Skip to content
BY 4.0 license Open Access Published by De Gruyter February 20, 2021

Weighted value distributions of the Riemann zeta function on the critical line

  • Alessandro Fazzari ORCID logo EMAIL logo
From the journal Forum Mathematicum

Abstract

We prove a central limit theorem for log|ζ(12+it)| with respect to the measure |ζ(m)(12+it)|2kdt (k,m), assuming RH and the asymptotic formula for twisted and shifted integral moments of zeta. Under the same hypotheses, we also study a shifted case, looking at the measure |ζ(12+it+iα)|2kdt, with α(-1,1). Finally, we prove unconditionally the analogue result in the random matrix theory context.

MSC 2010: 11M06

1 Introduction and statement of the main results

In 1946, Selberg [18] proved a central limit theorem for the real part of the logarithm of the Riemann zeta function on the critical line, showing that the distribution of log|ζ(12+it)| is approximately Gaussian with mean 0 and variance 12loglogT, i.e.

(1.1)1Tmeas{t[T,2T]:log|ζ(12+it)|12loglogTV}Ve-x22dx2π

for any fixed V, as T goes to infinity. The author [9] studied the distribution of log|ζ(12+it)| with respect to the weighted measure |ζ(12+it)|2dt and, assuming the Riemann hypothesis (RH), proved that it is asymptotically Gaussian with mean loglogT and variance 12loglogT. In this paper, we investigate the value distribution of log|ζ(12+it)| with respect to the measure

(1.2)|ζ(m)(12+it)|2kdt

for any fixed non-negative integers m,k. The motivation is due to the study of the large values of the Riemann zeta function, i.e. the uniformity in V in the central limit theorem (1.1). In [19], Soundararajan speculates that an upper bound like

1Tmeas{t[T,2T]:log|ζ(12+it)|12loglogTV}1Vexp(-V22)

holds in a large range for V. In particular, we expect (see [17, Conjecture 2] and [11]) that for any fixed k,

(1.3)1Tmeas{t[T,2T]:log|ζ(12+it)|kloglogT}k1loglogT1(logT)k2.

Expressing the characteristic function of positive reals as a Mellin transform, the left-hand side of (1.3) can be written as

1(logT)k212π-+1T(logT)k2T2Teiu(log|ζ(12+it)|-kloglogT)|ζ(12+it)|2k𝑑tdu2k+iu

and is then related to the distribution of log|ζ(12+it)| with respect to the weighted measure (1.2), with m=0.

In the case k=1,2, we can prove a central limit theorem for log|ζ(12+it)| with respect to the measure |ζ(m)(12+it)|2kdt, assuming RH only.

Corollary 1.1.

Assume the Riemann hypothesis, let m be a non-negative integer and let k=1 or k=2. As t varies in Tt2T, the distribution of log|ζ(12+it)| is asymptotically Gaussian with mean kloglogT and variance 12loglogT, with respect to the weighted measure |ζ(m)(12+it)|2kdt.

In the case k>2, since not even the moments of zeta are known, we cannot expect to prove unconditionally a central limit theorem. However, assuming the asymptotic formula [13, Conjecture 7.1] suggested by the recipe [6] for the twisted and shifted (2k)-th moments of the Riemann zeta function

T2T(ab)-itζ(12+α1+it)ζ(12+αk+it)ζ(12+β1-it)ζ(12+βk-it)𝑑t,

one can deal with the general case too. More precisely, we use the strategy of [4, Theorem 1.2] in order to re-write the conjecture of Hughes and Young and we assume the following statement.

Conjecture 1.2 (Hughes–Young).

Let T be a large parameter, let α1,,αk,β1,,βk(logT)-1, let Φj be the set of subsets of {α1,,αk} of cardinality j, for j=0,,k, and similarly let Ψj be the set of subsets of {β1,,βk} of cardinality j. If 𝒮Φj and 𝒯Ψj, then write 𝒮={αi1,,αij} and 𝒯={βl1,,βlj}, where i1<i2<<ij and l1<l2<<lj. Let (α𝒮;β𝒯) be the tuple obtained from (α1,,αk;β1,,βk) by replacing αlr with -βlr and replacing βir with -αir for 1rj. Consider

A(s)=aTθf(a)asandB(s)=bTθg(b)bs,

where f(a)aε and g(b)bε for any ε>0. We then conjecture that there exists a δ>0, depending on k, such that if θ<δ, then

T2TA(12+it)B(12+it)¯ζ(12+it+α1)ζ(12+it+αk)ζ(12-it+β1)ζ(12-it+βk)𝑑t

equals

a,bTθf(a)g(b)¯T2T0jk𝒮Φj𝒯ΨjZα𝒮;β𝒯,a,b(t)(t2π)-𝒮-𝒯dt+O(T1-η)

for some η>0, where we have written (t/2π)-𝒮-𝒯 for (t/2π)-x𝒮x-y𝒯y and

Zα;β,a,b(t):=am1mk=bn1nk1abm11/2+α1mk1/2+αkn11/2+β1nk1/2+βkV(m1mkn1nktk)

with

V(x):=12πi(1)G(s)s(2π)-ksx-s𝑑s,

where G(s) is an even entire function of rapid decay in any fixed strip |(s)|C satisfying G(0)=1.

With this assumption we can prove the main theorem.

Theorem 1.3.

Let k,mN and assume the Riemann hypothesis and Conjecture 1.2 for k. As t varies in Tt2T, the distribution of log|ζ(12+it)| is asymptotically Gaussian with mean kloglogT and variance 12loglogT, with respect to the weighted measure |ζ(m)(12+it)|2kdt.

In particular, since Conjecture 1.2 is known in the cases k=1 (see [2, 5] and, e.g., [3] for the easy modifications needed to account for the shifts) and k=2 (see [13, 4]), we notice that Corollary 1.1 trivially follows from Theorem 1.3.

We notice that Theorem 1.3 shows that the m-th derivative has no effect in the weighted distribution of log|ζ(12+it)|. This is consistent with the conjecture (see [12, Conjecture 6.1] and [8])

T2T|ζ(12+it)|2k-2h|ζ(12+it)|2h𝑑tc(h,k)T(logT)k2+2h,

which indicates that |ζ(12+it)|2h amplifies the contribution coming from the large values in the same way as |ζ(12+it)|2h would do (up to a normalization of (logT)2h). More generally, as far as moments are concerned, the m-th derivative of zeta should behave like zeta itself (see, e.g., [7]), up to a normalization of logmT, in accordance with Theorem 1.3. We also note that, while in Selberg’s classical case the mean is 0 because the contribution of the small values and that of the large values of zeta balance out, by tilting with |ζ(12+it)|2k, the mean of log|ζ(12+it)| moves to the right as k grows, and this reflects the fact that the measure |ζ(12+it)|2kdt gives more and more weight to the large values of the Riemann zeta function.

Moreover, we look at the shifted weighted measure |ζ(12+it+iα)|2kdt with a real number α such that |α|<1. As we will see, the distribution of log|ζ(12+it)| is quite sensitive to the parameter α. Indeed, in computing the integral

(1.4)T2Tlog|ζ(12+it)||ζ(12+it+iα)|2k𝑑t,

one expects the same magnitude as in the unshifted case if |α| is smaller than (logT)-1, which is the typical scale for the Riemann zeta function. On the other hand, if |α| is larger, then the two factors in the integral (1.4) start decorrelating, and thus the size of the integral decreases. This phenomenon is shown in the following result, in which we use the notation

μ~α={loglogT+O(1)if |α|logT1,-log|α|+O(1)if |α|logT>1,

for any α(-1,1).

Theorem 1.4.

Let kN and assume the Riemann hypothesis and Conjecture 1.2 for k. As t varies in Tt2T, for any fixed and real α such that |α|<1, the distribution of log|ζ(12+it)| is asymptotically Gaussian with mean kμ~α and variance 12loglogT, with respect to the measure |ζ(12+it+iα)|2kdt.

This theorem shows that the shift has no effect if it is smaller than (logT)-1. On the contrary, for larger values of the shift the mean gets smaller; for instance, if

α=(logT)δlogT

with δ(0,1), then the mean is (1-δ)kloglogT. In this shifted case too, if k2, then Theorem 1.4 holds assuming RH only.

Lastly, we show that in the random matrix theory setting, an analogous weighted central limit theorem can be proved unconditionally. We consider the characteristic polynomials

Z=Z(U,θ)=det(I-Ue-iθ)

of N×N unitary matrices U and we investigate their distribution of values with respect to the circular unitary ensemble (CUE). It has been conjectured that the limiting distribution of the non-trivial zeros of the Riemann zeta function, on the scale of their mean spacing, is the same as that of the eigenphases θn of matrices in the CUE in the limit as N (see, e.g., [14]). Then we consider a tilted version of the Haar measure and we have the following theorem.

Theorem 1.5.

As N, the value distribution of log|Z| is asymptotically Gaussian with mean klogN and variance 12logN with respect to the measure |Z|2kdHaar .

As usual, the correspondence with Theorem 1.3 holds if we identify the mean density of the eigenangles θn, N/2π, with the mean density of the Riemann zeros at a height T up the critical line 12πlogT2π, i.e. if

N=logT2π.

2 Proof of Theorems 1.3 and 1.4

To prove both the theorems, we introduce a set of shifts α1,,αk,β1,,βk and we denote for the sake of brevity

ζα,β(t):=ζ(12+α1+it)ζ(12+αk+it)ζ(12+β1-it)ζ(12+βk-it).

The general strategy of the proof is similar to the one of [9, Theorem 1], but here we avoid the detailed combinatorial analysis we performed in [9], by working with Euler products instead of Dirichlet series, inspired by [1, Proposition 5.1]. The first step is then approximating the logarithm of the Riemann zeta function with a suitable Dirichlet polynomial. Let us set

P~(t):=px1p1/2+it,

where x:=Tε, with ε:=(logloglogT)-1. Now we show that log|ζ(12+it)| has the same distribution as P~(t) with respect to the measure ζα,β(t)dt. This is achieved in the following proposition by bounding the second moment of the difference.

Proposition 2.1.

For a non-negative integer k, assume Conjecture 1.2 and RH. Let T be a large parameter and let α1,,αk, β1,,βkC such that |αi|,|βj|<1, |(αi)|,|(βj)|(logT)-1 and |αi-βj|(logT)-1 for all 1i,jk. Then we have

T2T|log|ζ(12+it)|-P~(t)|2ζα,β(t)𝑑tkT(logT)k2(logloglogT)5/2.

Proof.

From Tsang’s work [20, equation (5.15)], we know that

logζ(12+it)-P~(t)=S1+S2+S3+O(R)-L(t)

with

S1:=px(p-1/2-4/logx-p-1/2)p-it,
S2:=prxr2p-r(1/2+4/logx+it)r,
S3:=x<nx3Λ(n)lognn-1/2-4/logx-it,
R:=5logx(|nx3Λ(n)n1/2+4/logx+it|+logT),
L(t):=ρ1/21/2+4/logx(12+4logx-u)1u+it-ρ112+4logx-ρ𝑑u,

where the sum in the definition of L(t) is over all non-trivial zeros of ζ. Then we have to bound the second moments of the terms on the right-hand side with respect to the weighted measure ζα,β(t)dt by using Conjecture 1.2 (note that we are allowed to apply the conjecture since the shifts are small up to a change of variable, being |αi-βj|(logT)-1).

Let us start with

T2T|S1|2ζα,β(t)𝑑t|T2T0jk𝒮Φj𝒯Ψj(t2π)-𝒮-𝒯p,qx(p-4/logx-1)(q-4/logx-1)Zα𝒮,β𝒯,p,q(t)dt|
T2T0jk𝒮Φj𝒯Ψj|a,bx𝟙x(a)𝟙x(b)(a-4logx-1)(b-4logx-1)Zα𝒮,β𝒯,a,b(t)|dt,

where 𝟙x() is the indicator function of primes up to x. If we denote by Ω(n) the function which counts the number of prime factors of n with multiplicity, then we have

nx𝟙x(n)f(n)=n:p|npxz[zΩ(n)f(n)]z=0.

Hence we get

(2.1)T2T|S1|2ζα,β(t)dtT2T0jk𝒮Φj𝒯Ψj|zw[12πi(1)G(s)s(t2π)ksIα,βj,𝒮,𝒯(z,w;s)ds]z=0w=0|dt,

where

(2.2)Iα,βj,𝒮,𝒯(z,w;s):=a,b:p|abpxzΩ(a)wΩ(b)gx(a)gx(b)Z~α𝒮,β𝒯,a,b(s)

with gx being the multiplicative function defined by gx(pα)=p-4α/logx-1 and

(2.3)Z~α,β,a,b(s):=am1mk=bn1nk1abm11/2+α1+smk1/2+αk+sn11/2+β1+snk1/2+βk+s.

We now analyze the first term Iα,β0(z,w;s) and then we will see how to apply the method to deal with all others. Since the sum in the definition (2.2) is multiplicative, we have

Iα,β0(z,w;s)=am1mk=bn1nk:p|abpxzΩ(a)wΩ(b)gx(a)gx(b)abm112+α1+smk12+αk+sn112+β1+snk12+βk+s
=pxa+m1++mk=b+n1++nkzΩ(pa)wΩ(pb)gx(pa)gx(pb)pa2+b2+m1(12+α1+s)++mk(12+αk+s)+n1(12+β1+s)++nk(12+βk+s)
(2.4)p>xm1++mk=n1++nk1pm1(12+α1+s)++mk(12+αk+s)+m1(12+β1+s)++nk(12+βk+s)

By putting in evidence the first terms in the Euler products, this equals

Aα,β(z,w;s)px(1+zwgx(p)2p+zgx(p)p1+β1+s++wgx(p)p1+αk+s)
   px(1+1p1+α1+β1+2s++1p1+αk+βk+2s)p>x(1+1p1+α1+β1+2s++1p1+αk+βk+2s)
(2.5)=Aα,β*(z,w;s)i,j=1kζ(1+αi+βj+2s)exp(px{zwgx(p)2p+zgx(p)p1+β1+s++wgx(p)p1+αk+s}),

where Aα,β(z,w;s) and Aα,β*(z,w;s) are arithmetical factors (Euler products) converging absolutely in a half-plane (s)>-δ for some δ>0, uniformly for |z|,|w|1, such that their derivatives with respect to z and w at 0 also converge in the same half-plane. We now have extracted the polar part, and hence we are ready to shift the integral over s in (2.1) to the left of zero. To do so, it is convenient to prescribe the same conditions as in [4, remarks after Lemma 2.1], assuming that G(s) vanishes at s=-αi+βj2 for all i,j, so that the only pole we pick in the contour shift is at s=0. Moreover, we assume that the shifts are such that |αi+βj|(logT)-1 for every 1i,jk, so that

(2.6)i,j=1k|ζ(1+αi+βj)|k(logT)k2.

Hence by (2.4)–(2.6), we get

zw[12πi(1)G(s)s(t2π)ksIα,β0(z,w;s)ds]z=0w=0
k(logT)k2|zw[exp(px{zwgx(p)2p+zgx(p)p1+β1++wgx(p)p1+αk})]z=0w=0|
k(logT)k2(px|gx(p)|p)2
k(logT)k2(logloglogT)2,

where (see, e.g., [9, p. 5])

|(αi)|,|(β)|(logT)-1andpx|p-4/logx-1|plogloglogT.

This last bound can be obtained by using the Taylor approximation e-z=1+O(z) for z1, which yields

px|p-4/logx-1|p=px|p-4logp/logx-1|p1logxpxlogpp1

by Merten’s first theorem. All the other terms Iα,βj,𝒮,𝒯(z,w) can be treated exactly in the same way as Iα,β0(z,w) by assuming that |αi±βj|(logT)-1 for every i,j, since they only differ from the first case by permutations and changes of signs of the shifts. Therefore, we get

zw[12πi(1)G(s)s(t2π)ksIα,βj,𝒮,𝒯(z,w;s)ds]z=0w=0k(logT)k2(logloglogT)2

provided that

(2.7)|αi±βj|(logT)-1for every 1i,jk.

Plugging this into (2.1), we prove that

(2.8)T2T|S1|2ζα,β(t)𝑑tkT(logT)k2(logloglogT)2.

Moreover, since the left-hand side in (2.8) is holomorphic in terms of the shifts, although we have proved the above for αi,βj such that (2.7) holds, the maximum modulus principle can be applied to obtain the bound to the enlarged domain we need.

We will treat the other pieces similarly. As regards the second one, we have that

T2T|S2|2ζα,β(t)𝑑tT2T0jk𝒮Φj𝒯Ψj|p1r1,p2r2xr1,r22p1-4r1logxr1p2-4r2logxr2Zα𝒮,β𝒯,p1r1,p2r2(t)|dt.

As before, we analyze the first term only since all others are completely analogous. The term for j=0 is

12πi(1)G(s)s(t2π)ksp1r1m1mk=p2r2n1nk:p1r1,p2r2x,r1,r22p1-4r1logxp2-4r2logxr1r21p1r1p2r2m112+α1+snk12+βk+sds.

This time, because of the condition r1,r22, when we estimate the sum via the first terms of its Euler product, we just get that the above is

ki,j=1k|ζ(1+αi+βj)|

Applying the same machinery as before, this yields

T2T|S2|2ζα,β(t)𝑑tkT(logT)k2.

We use the same approach in order to bound the second moment of S3 as well, which is

kT2T|12πi(1)G(s)s(t2π)ksam1mk=bn1nkx<a,b<x3Λ(a)Λ(b)logalogba-4/logxb-4/logxabm112+α1+snk12+βk+s|𝑑t
kT(x<px31p)2i,j=1k|ζ(1+αi+βj)|
kT(logT)k2

since the sum x<px3p-1 is bounded.

We deal with R in the same way and we get that

T2T|R|2ζα,β(t)𝑑tkT(logT)k2(logloglogT)2,

where the extra factor with the triple log comes from the second term in the definition of R, while the first one can be treated analogously to S3.

Finally, we have to bound the second moment of L(t). To do so, in view of equations [9, (2.8) and (2.9)], it suffices to study

1(logx)2T2T(log+1ηtlogx)2(|nx3Λ(n)n-4/logxn1/2+it|+logT)2ζα,β(t)𝑑t,

where ηt:=minρ|t-γ| and log+t:=max(logt,0), with the aim of proving that this is

kT(logT)k2(logloglogT)5/2.

By applying the Cauchy–Schwarz inequality, the above is

(T2T(1logx|nx3Λ(n)n-4/logxn1/2+it|+1ε)4|ζ(12+β1-it)|2|ζ(12+βk-it)|2𝑑t)12
(2.9)(T2T(log+1ηtlogx)4|ζ(12+α1+it)|2|ζ(12+αk+it)|2𝑑t)12

and the first term can be treated as R above and it is

kT(logT)k2(logloglogT)4.

We now conclude the proof, bounding the second term in (2.9). If we denote

𝒯:=[T-1logx,2T+1logx],

then we have

T2T(log+1ηtlogx)4|ζ(12+α1+it)|2|ζ(12+αk+it)|2𝑑t
γ𝒯-1/logx1/logx(log+1|w|logx)4j=1k|ζ(12+αj+i(w+γ))|2dw
=γ𝒯-11(log+1|t|)4j=1k|ζ(12+αj+i(γ+tlogx))|2dtlogx
(2.10)1logx-11(log|t|)4j=1k(γ𝒯|ζ(12+iγ+αj+itlogx)|2k)1/kdt

by the Hölder inequality. The remaining sum can be bounded under RH in view of Kirila’s [15, Theorem 1.2] and Milinovich’s [16] works, which generalize the well-known result due to Gonek about the sum over the non-trivial zeros of zeta [10, Corollary 1]. Indeed, since the shifts αj+itlogx in the sum over zeros in (2.10) have modulus 1 and real part (logT)-1 in absolute value, then we have

γ𝒯|ζ(12+iγ+αj+itlogx)|2kkTlogT(logT)k2

for every j=1,,k. Putting this into (2.10), we get

(2.11)T2T(log+1ηtlogx)4|ζ(12+α1+it)|2|ζ(12+αk+it)|2𝑑tkTlogTlogx(logT)k2.

Plugging (2.11) into (2.9), we prove that

T2T|L(t)|2ζα,β(t)𝑑tkT(logT)k2(logloglogT)4T(logT)k2logloglogT
kT(logT)k2(logloglogT)5/2,

concluding the proof of the proposition. ∎

The second step is getting rid of the small primes, showing that their contribution does not affect the distribution asymptotically. This simple fact will simplify the third and last step of the proof, as we will see in the following. Let us define

P(t):=pX1p1/2+it,

where X:=(logT,x] (we recall that x=Tε and ε=(logloglogT)-1).

Proposition 2.2.

For a non-negative integer k, assume Conjecture 1.2 and RH. Let T be a large parameter and let α1,,αk, β1,,βkC such that |αi|,|βj|<1, |(αi)|,|(βj)|(logT)-1 and |αi-βj|(logT)-1 for all 1i,jk. Then we have

T2T|P~(t)-P(t)|2ζα,β(t)𝑑tkT(logT)k2(logloglogT)2.

Proof.

This can be proved with the same method used in Proposition 2.1. We recall that 𝟙logT() denotes the indicator function of primes up to logT. We start by studying

(2.12)12πi(1)G(s)s(t2π)ksam1mk=bn1nk𝟙logT(a)𝟙logT(b)abm112+α1+snk12+βk+sds

and, as usual, we estimate the sum with the first terms of its Euler product, extract the polar part, shift the integral over s to the left, getting that (2.12) is

ki,j=1k|ζ(1+αi+βj)||zw[exp(plogT{zwp+zp1+β1++wp1+αk})]z=0w=0|.

By the same argument as in the proof of Proposition 2.1 the above is

k(logT)k2(plogT1p)2(logT)k2(logloglogT)2,

and this concludes the proof. ∎

Finally, we investigate the distribution of the polynomial P(t), which, thanks to Propositions 2.1 and 2.2, has the same distribution as log|ζ(12+it)|. The most natural method to do so is studying the moments, and this is achieved in the following result.

Proposition 2.3.

For a non-negative integer k, assume Conjecture 1.2 and RH. Let T be a large parameter and let α1,,αk, β1,,βkC such that |αi|,|βj|<1, |(αi)|,|(βj)|(logT)-1 and |αi-βj|(logT)-1 for all 1i,jk. Set L:=pX1ploglogT and μR such that μloglogT. Then for every fixed integer n we have

T2T(P(t)-kμ)nζα,β(t)𝑑t
=T2T0jk𝒮Φj𝒯Ψj(t2π)-𝒮-𝒯Mα𝒮,β𝒯(0)zn[ez24-kzμexp(z2pXgp(𝒮,𝒯)p)]z=0dt+Ok,n(T(logT)k2-1+ε)

where

gp(𝒮,𝒯):=x1𝒮p-x1+x2𝒯p-x2+x3𝒮px3+x4𝒯px4

and

Mα,β(s):=m1mk=n1nk1m112+α1+smk12+αk+sn112+β1+snk12+βk+s,

so that Mα,β(0) is the first term of the moment of ζα,β, predicted by the recipe [6] (the sum which defines Mα,β(s) does not converge for s=0, so we appeal to [6, Theorem 2.4.1] for the analytic continuation).

Proof.

Expanding out the powers, since 2(z)=z+z¯, we get

(2.13)T2T(P(t)-kμ)nζα,β(t)𝑑t=j+h=n(nh)(-kμ)j2-hr+s=h(hr)T2TP(t)rP(t)¯sζα,β(t)𝑑t.

By using Conjecture 1.2 and ignoring the error term which is negligible in this context, the inner integral in (2.13) is

(2.14)T2T0jk𝒮Φj𝒯Ψj(t2π)-𝒮-𝒯a,b𝟙X*r(a)𝟙X*s(b)Zα𝒮,β𝒯,a,b(t)dt,

where 𝟙X*r(a) denotes the indicator function of primes in the interval X, self-convoluted r times. If we define temporarily the multiplicative function g given by g(pn)=1/n!, then the inner sum over a,b in (2.14) equals

(2.15)zrws[a,b:p|abpXzΩ(a)wΩ(b)g(a)g(b)Zα𝒮,β𝒯,a,b(t)]z=0w=0.

Then, plugging (2.14) and (2.15) into (2.13) and recollecting together the powers we expanded before, we get

(2.16)T2T12πi(1)G(s)s(t2π)ks0jk𝒮Φj𝒯Ψj(t2π)-𝒮-𝒯zn[e-kμzIα,βj,𝒮,𝒯(z;s)]z=0dsdt,

with

Iα,βj,𝒮,𝒯(z;s):=a,b:p|abpX(z2)Ω(a)+Ω(b)g(a)g(b)Z~α𝒮,β𝒯,a,b(s);

see (2.3) for the definition of Z~α,β,a,b(s). Now study the first term in (2.16), i.e. j=0, since all other terms can be understood from the first one with a slight modification. Then we look at

Iα,β0(z;s)=am1mk=bn1nkp|abpX(z/2)Ω(a)+Ω(b)g(a)g(b)abm112+α1+smk12+αk+sn112+β1+snk12+βk+s
=pXa+m1++mk=b+n1++nk(z/2)a+bg(pa)g(pb)pa2+b2+m1(12+α1+s)++nk(12+βk+s)pXm1++mk=n1++nk1pm1(12+α1+s)++nk(12+βk+s)
=exp(pX{(z/2)2p+z/2p1+α1+s++z/2p1+αk+s+z/2p1+β1+s+z/2p1+βk+s})FX,α,β(z;s)Mα,β(s),

where FX,α,β(z;s) is an arithmetical factor (Euler product) converging absolutely in a product of half-planes containing the origin, such that FX,α,β(z;0) is holomorphic at z=0, FX,α,β(0,0)=1 and all its derivatives at z=0 are small, i.e.

zc[FX,α,β(z,0)]z=0(logT)-1

for any positive integer c. Now we want to shift the integral over s in (2.16) to the left of zero, picking the contribution of the (only) pole at s=0. To do so, we appeal to the meromorphic continuation of the function Mα,β(s); see [6, Theorem 2.4.1]. Thus we can shift the path of integration to the vertical line (say) (s)=-110, where the integral is trivially bounded by T1-1/10+ε for any positive ε. Moreover, the contribution from the pole at s=0 gives

Tzn[e-kμzez24exp(z2pXp-α1++p-βkp)FX,α,β(z;0)]z=0Mα,β(0).

Thanks to the bounds for μ, and the derivatives of FX,α,β being FX,α,β(0;0)=1 and Mα,β(0)(logT)k2 (again this is due to an argument similar to the one in the proof of Proposition 2.1, when we assume extra conditions on the shifts and then we appeal to the maximum modulus principle), we get that the term for j=0 in (2.16) equals

Tzn[e-kμzez24exp(z2pXp-α1++p-βkp)]z=0Mα,β(0)+Ok,n(T(logT)k2-1+ε).

Analogously, the general term turns out to be

T2T(t2π)-𝒮-𝒯zn[ez24-kμzexp(z2pXgp(𝒮,𝒯)p)]z=0Mα𝒮,β𝒯(0)dt+Ok,n(T(logT)k2-1+ε),

and putting this into (2.14), i.e. summing over j,𝒮,𝒯, we get the thesis. ∎

Proof of Theorem 1.3.

This proof follows easily from the three propositions we have proved above. If we take μ= and all shifts small enough, i.e.

αi,βj(logT)-1for any i,j,

then the exponent on the right-hand side of the thesis in Proposition 2.3 becomes

z24+z2pXgp(𝒮,𝒯)p-zk=z24+O(zpX|p-α1-1|++|p-βk-1|p)
=z24+Ok(zpX|p-α1-1|p)
=z24+Ok(z)

and does not depend on 𝒮 and 𝒯 asymptotically, in fact. Hence we can bring that factor outside and reconstruct the moment of ζα,β, as follows:

T2T(P(t)-k)nζα,β(t)𝑑t
=zn[ez24+Ok(z)]z=0T2T0jk𝒮Φj𝒯Ψj(t2π)-𝒮-𝒯Mα𝒮,β𝒯(0)dt+Ok,n(T(logT)k2-1+ε)
=(zn[ez24]z=0+Ok,n((loglogT)n-12))T2Tζα,β(t)dt+Ok,n(T(logT)k2-1+ε).

The thesis follows by analyzing

zn[ez24]z=0=[m1+2m2=nn!m1!m2!(2z4)m1(12!2)m2]z=0.

If n is odd then the coefficient

zn[ez24]z=0

vanishes, while if n is even, then only the term for m1=0 and m2=n2 survives and gives

n!(n/2)!(4)n/2=(n-1)!!(2)n/2,

i.e.

1T2Tζα,β(t)𝑑tT2T(P(t)-k)nζα,β(t)𝑑t={(1+ok,n(1))(n-1)!!(2)n/2if n is even,ok,n((loglogT)n/2)if n is odd.

This then matched with the Gaussian coefficient proves that, in the limit T, P(t) has Gaussian distribution, with mean kloglogT and variance 12loglogT, and then so does log|ζ(12+it)|, in view of Propositions 2.1 and 2.2. Theorem 1.3 follows by taking the derivatives with respect to the shifts. ∎

Proof of Theorem 1.4.

To derive Theorem 1.4, in Proposition 2.3 we set

α1==αk=iαandβ1==βk=-iα,

with α,|α|<1, and we take μ as

μα:=pXcos(αlogp)p=μ~α+O(1)={loglogT+O(1)if |α|logT1,-log|α|+O(1)if |α|logT>1,

by partial summation. Then we get

T2T(P(t)-kμα)n|ζ(12+iα+it)|2kdt=(1+ok,n(1))zn[ez24]z=0T2T|ζ(12+iα+it)|2kdt

since the quantity

z2pXgp(𝒮,𝒯)p-zkμα

vanishes for all 𝒮,𝒯 for this choice of the shifts. The thesis follows as in the proof of Theorem 1.3. ∎

3 Proof of Theorem 1.5

In the usual notations we set in the introduction, let us define the moment generating function

(3.1)MN(s)=|Z|s=j=0(log|Z|)jj!sj,

where the mean has to be considered over the group U(N) with respect to the Haar measure, and the cumulants Qj=Qj(N) by

logMN(s)=j=1Qjj!sj.

In [14], among other things, Keating and Snaith studied the cumulants showing that

Qn={0if n=1,12logN+O(1)if n=2,O(1)if n3,

and deduced a central limit theorem proving that the limiting distribution of log|Z| is Gaussian with mean 0 and variance 12logN. Here, for any k, we study the distribution of the random variable log|Z| with respect to the tilted measure |Z|2kdHaar. Before starting with our analysis, we recall that the moments of |Z| are known also for a non-negative integer k (see [14, (6) and (16)]):

(3.2)MN(2k)=|Z|2k=exp(j=1(2k)jj!Qj)=j=1NΓ(j)Γ(j+2k)Γ(j+k)2Nk2G2(1+k)G(1+2k),

where k and G denotes the Barnes G-function. We set

2k:=j=1NΓ(j)Γ(j+2k)Γ(j+k)2.

Now we are ready to consider the first moment

|Z|2klog|Z|=ddx[|Z|2k+x]x=0=ddx[j=1NΓ(j)Γ(j+2k+x)Γ(j+k+x2)2]x=0

by (3.1) and (3.2). We compute the derivative by Leibniz’s rule, writing

j=1NΓ(j)Γ(j+2k+x)Γ(j+k+x2)2=exp(j=1N{logΓ(j)+logΓ(j+2k+x)-2logΓ(j+k+x2)}),

and we get

(3.3)|Z|2klog|Z|=j=1NΓ(j)Γ(j+2k)Γ(j+k)2j=1N{ΓΓ(j+2k)-ΓΓ(j+k)}.

Moreover, an application of Stirling’s formula yields

(3.4)ΓΓ(j+2k)-ΓΓ(j+k)=kj+Ok(1j2).

Hence, by (3.3) and (3.4), the weighted mean of the random variable log|Z| is

μ2k:=12k|Z|2klog|Z|=klogN+Ok(1).

Then we study the weighted n-th moment of the random variable log|Z|:

|Z|2k(log|Z|-μ2k)n=h+j=n(nh)(-μ2k)j|Z|2k(log|Z|)h
=h+j=n(nh)djdxj[e-xμ2k]x=0dhdxh[j=1NΓ(j)Γ(j+2k+x)Γ(j+k+x2)2]x=0
(3.5)=dndxn[exp(-xμ2k+j=1Nlog(Γ(j)Γ(j+2k+x)Γ(j+k+x2)2))]x=0.

If we set

fj(x)=fN,k,j(x):=logΓ(j)+logΓ(j+2k+x)-2logΓ(j+k+x2),

then we can carry on the computation in (3.5) by computing the derivative, getting

dndxn[exp(-xμ2k+j=1Nlog(Γ(j)Γ(j+2k+x)Γ(j+k+x2)2))]x=0
=m1,,mnn!m1!mn!ej=1Nfj(0)i=1n(1i!didxi[-xμ2k+j=1Nfj(x)]x=0)mi
(3.6)=2km1,,mnn!m1!mn!(-μ2k+j=1Nfj(0))m1(12j=1Nfj′′(0))m2i=3n(1i!j=1Nfj(i)(0))mi,

where the sums in (3.6) are over the n-tuple (m1,,mn) such that

m1+2m2++nmn=n.

Using Stirling’s approximation formula, one can easily estimate the derivatives of fj(x) and prove

j=1Nfj(0)=j=1N{kj+Ok(j-2)}=klogN+Ok(1),
j=1Nfj′′(0)=j=1N{1j+2k-1/2j+k+O(j-2)}=12logN+Ok(1),
(3.7)j=1Nfj(i)(0)=j=1NO(j-2)=O(1)for all i3.

Putting together (3.5), (3.6) and (3.7), one has

|Z|2k(log|Z|-μ2k)n=2km1+2m2++nmn=nn!m1!mn!(14logN+Ok(1))m2(Ok(1))m1+m3++mn.

Then, if n is even, the asymptotic is given by m2=n2 and mi=0 for i2, giving

|Z|2k(log|Z|-μ2k)nk,n2kn!(n/2)!(14logN)n/2=2k(n-1)!!(12logN)n/2,

while if n is odd, then the n-th moment is surely ok,n(2k(logN)n/2).


Communicated by Valentin Blomer


Acknowledgements

I am grateful to Sandro Bettin for his support and encouragement as well as for many useful suggestions and to Jon Keating for helpful conversations and for pointing me out interesting connections with the random matrix theory setting, which inspired, among other things, the last section of this paper. I also wish to thank the referee for a very careful reading of the paper and for indicating several inaccuracies.

References

[1] L.-P. Arguin, D. Belius, P. Bourgade, M. Radziwił ł and K. Soundararajan, Maximum of the Riemann zeta function on a short interval of the critical line, Comm. Pure Appl. Math. 72 (2019), no. 3, 500–535. 10.1002/cpa.21791Search in Google Scholar

[2] R. Balasubramanian, J. B. Conrey and D. R. Heath-Brown, Asymptotic mean square of the product of the Riemann zeta-function and a Dirichlet polynomial, J. Reine Angew. Math. 357 (1985), 161–181. 10.1515/crll.1985.357.161Search in Google Scholar

[3] S. Bettin, The second moment of the Riemann zeta function with unbounded shifts, Int. J. Number Theory 6 (2010), no. 8, 1933–1944. 10.1142/S1793042110003861Search in Google Scholar

[4] S. Bettin, H. M. Bui, X. Li and M. Radziwił ł, A quadratic divisor problem and moments of the Riemann zeta-function, J. Eur. Math. Soc. (JEMS) 22 (2020), no. 12, 3953–3980. 10.4171/JEMS/999Search in Google Scholar

[5] S. Bettin, V. Chandee and M. Radziwił ł, The mean square of the product of the Riemann zeta-function with Dirichlet polynomials, J. Reine Angew. Math. 729 (2017), 51–79. 10.1515/crelle-2014-0133Search in Google Scholar

[6] J. B. Conrey, D. W. Farmer, J. P. Keating, M. O. Rubinstein and N. C. Snaith, Integral moments of L-functions, Proc. Lond. Math. Soc. (3) 91 (2005), no. 1, 33–104. 10.1112/S0024611504015175Search in Google Scholar

[7] J. B. Conrey and A. Ghosh, Zeros of derivatives of the Riemann zeta-function near the critical line, Analytic Number Theory (Allerton Park 1989), Progr. Math. 85, Birkhäuser, Boston (1990), 95–110. 10.1007/978-1-4612-3464-7_8Search in Google Scholar

[8] J. B. Conrey, M. O. Rubinstein and N. C. Snaith, Moments of the derivative of characteristic polynomials with an application to the Riemann zeta function, Comm. Math. Phys. 267 (2006), no. 3, 611–629. 10.1007/s00220-006-0090-5Search in Google Scholar

[9] A. Fazzari, A weighted central limit theorem for log|ζ(1/2+it)|, Mathematika 67 (2021), no. 2, 324–341. 10.1112/mtk.12078Search in Google Scholar

[10] S. M. Gonek, Mean values of the Riemann zeta function and its derivatives, Invent. Math. 75 (1984), no. 1, 123–141. 10.1007/BF01403094Search in Google Scholar

[11] A. Harper, Sharp conditional bounds for moments of the Riemann zeta functions, preprint (2013), https://arxiv.org/abs/1305.4618v1. Search in Google Scholar

[12] C. P. Hughes, On the characteristic polynomial of a random unitary matrix and the Riemann zeta function, PhD thesis, University of Bristol, 2001. 10.1007/s002200100453Search in Google Scholar

[13] C. P. Hughes and M. P. Young, The twisted fourth moment of the Riemann zeta function, J. Reine Angew. Math. 641 (2010), 203–236. 10.1515/crelle.2010.034Search in Google Scholar

[14] J. P. Keating and N. C. Snaith, Random matrix theory and ζ(1/2+it), Comm. Math. Phys. 214 (2000), no. 1, 57–89. 10.1007/s002200000261Search in Google Scholar

[15] S. Kirila, An upper bound for discrete moments of the derivative of the Riemann zeta-function, Mathematika 66 (2020), no. 2, 475–497. 10.1112/mtk.12008Search in Google Scholar

[16] M. B. Milinovich, Upper bounds for moments of ζ(ρ), Bull. Lond. Math. Soc. 42 (2010), no. 1, 28–44. 10.1112/blms/bdp096Search in Google Scholar

[17] M. Radziwill, Large deviations in Selberg’s central limit theorem, preprint (2011), https://arxiv.org/abs/1108.5092. Search in Google Scholar

[18] A. Selberg, Contributions to the theory of the Riemann zeta-function, Arch. Math. Naturvid. 48 (1946), no. 5, 89–155. Search in Google Scholar

[19] K. Soundararajan, Moments of the Riemann zeta function, Ann. of Math. (2) 170 (2009), no. 2, 981–993. 10.4007/annals.2009.170.981Search in Google Scholar

[20] K.-M. Tsang, The Distribution of the Values of the Riemann Zeta-Function, ProQuest LLC, Ann Arbor, 1984; Thesis (Ph.D.)–Princeton University. Search in Google Scholar

Received: 2020-10-04
Revised: 2021-01-12
Published Online: 2021-02-20
Published in Print: 2021-05-01

© 2021 Walter de Gruyter GmbH, Berlin/Boston

This work is licensed under the Creative Commons Attribution 4.0 International License.

Downloaded on 27.4.2024 from https://www.degruyter.com/document/doi/10.1515/forum-2020-0284/html
Scroll to top button