This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Explicit bounds for the Riemann zeta-function on the 1-line

Ghaith A. Hiary, Nicol Leong, and Andrew Yang GH: Department of Mathematics, The Ohio State University, 231 West 18th Ave, Columbus, OH 43210, USA hiary.1@osu.edu NL: University of New South Wales (Canberra) at the Australian Defence Force Academy, ACT, Australia nicol.leong@unsw.edu.au AY: University of New South Wales (Canberra) at the Australian Defence Force Academy, ACT, Australia andrew.yang1@unsw.edu.au
Abstract.

Explicit estimates for the Riemann zeta-function on the 11-line are derived using various methods, in particular van der Corput lemmas of high order and a theorem of Borel and Carathéodory.

Key words and phrases:
Van der Corput estimate, exponential sums, Riemann zeta function.
2020 Mathematics Subject Classification:
Primary: 11M06, 11Y35, 11L07.

1. Introduction

Let ζ(s)\zeta(s) denote the Riemann zeta-function. An open problem in analytic number theory is to determine the growth rate of ζ(s)\zeta(s) on the line s=1+its=1+it as t+t\to+\infty. Assuming the Riemann hypothesis, Littlewood [Lit26, Lit28] showed that the value of |ζ(1+it)||\zeta(1+it)| is limited to the range

1loglogtζ(1+it)loglogt.\frac{1}{\log\log t}\ll\zeta(1+it)\ll\log\log t.

In comparison, the sharpest estimates requiring no assumption furnish a wider range of the form

(1) 1(logt)2/3(loglogt)1/3ζ(1+it)log2/3t.\frac{1}{(\log t)^{2/3}(\log\log t)^{1/3}}\ll\zeta(1+it)\ll\log^{2/3}t.

The upper bound in the unconditional estimate (1) was made explicit by Ford [For02]. He computed numerical constants cc and t0t_{0} such that111Trudgian [Tru14] later lowered Ford’s value of cc from 76.276.2 to 62.662.6.

|ζ(1+it)|clog2/3t,(tt0).|\zeta(1+it)|\leq c\log^{2/3}t,\qquad(t\geq t_{0}).

The constants cc and t0t_{0} are often too large for standard applications, including explicit bounds on S(t)S(t) [Tru14a, HSW22] and explicit zero-density estimates [KLN18]. One instead defaults to an asymptotically worse bound of the form ζ(1+it)logt\zeta(1+it)\ll\log t that nevertheless is sharper in a finite range of tt of interest.

This motivates our Theorem 1. We establish a new explicit bound on ζ(1+it)\zeta(1+it) that is asymptotically sharper than O(logt)O(\log t) but still good at small values of tt.

Theorem 1.

For t3t\geq 3, we have

|ζ(1+it)|1.731logtloglogt.|\zeta(1+it)|\leq 1.731\frac{\log t}{\log\log t}.

Additionally, using a different method, we prove explicit upper bounds of the same form, on both 1/|ζ(1+it)|1/|\zeta(1+it)| and |ζ(1+it)/ζ(1+it)|\left|\zeta^{\prime}(1+it)/\zeta(1+it)\right|.

Theorem 2.

For t3t\geq 3, we have

1|ζ(1+it)|431.7logtloglogt.\frac{1}{|\zeta(1+it)|}\leq 431.7\frac{\log t}{\log\log t}.

Furthermore, for t500t\geq 500, if

1σ1+9250loglogtlogt,1\leq\sigma\leq 1+\frac{9}{250}\frac{\log\log t}{\log t},

then

|ζζ(σ+it)|154.5logtloglogt.\left|\frac{\zeta^{\prime}}{\zeta}(\sigma+it)\right|\leq 154.5\frac{\log t}{\log\log t}.

Theorem 1 makes explicit a result due to Weyl [Wey21] (see also [Tit86, Section 5.16]), and is sharper than the prior explicit bounds due to Patel [Pat22] on |ζ(1+it)||\zeta(1+it)| for tt in the range exp(3382)texp(3.61108)\exp(3382)\leq t\leq\exp(3.61\cdot 10^{8}).222We also note that Theorem 1 is sharper for t23256t\geq 23256 than the bound |ζ(1+it)|34logt|\zeta(1+it)|\leq\tfrac{3}{4}\log t given in [Tru14].

As far as we know, the bounds in Theorem 2 are the first unconditional explicit bounds of their kind. The first assertion of Theorem 2 supersedes [Car+22, Proposition A.2] when texp(exp(10.07))t\geq\exp(\exp(10.07)), while the second assertion supersedes [Tru15, Table 2 with (W,R1)=(12,40.14)(W,R_{1})=(12,40.14)] for texp(exp(3.85))t\geq\exp(\exp(3.85)).

Under the assumption of the Riemann Hypothesis, sharper explicit estimates for ζ/ζ(1+it)\zeta^{\prime}/\zeta(1+it) and 1/ζ(1+it)1/\zeta(1+it) of the order loglogt\log\log t are known [Sim23], [CHS24, Theorem 5], [Chi23]. Much work has also been done on bounding more general LL-functions [Lum18], [PS22].

Lastly, it is reasonable to try to adapt the remarkable approaches in [Sou09, p. 984] and [LLS15, Lemma 2.5] to our context. In these articles, very precise bounds, conditional on the Riemann hypothesis, are obtained. One can arrange for these approaches to yield unconditional bounds, and the outcome is similarly constrained by the width of available explicit zeros-free regions. See [Li10], for example, where the method from [Sou09] is used to establish bounds on LL-functions for σ1\sigma\geq 1.

2. Discussion

One can derive an explicit bound on |ζ(1+it)||\zeta(1+it)| of the form logt+α1\log t+\alpha_{1} using a version of the Euler–Maclaurin summation. For example, using the version in [Sim20, Corollary 2], applied with t0=30πt_{0}=30\pi and c=1/4c=1/4, together with the elementary harmonic sum bound333Here, γ=0.577216\gamma=0.577216\ldots is the Euler constant. (see Lemma 8 below),

(2) n=1N1nlogN+γ+12N,\sum_{n=1}^{N}\frac{1}{n}\leq\log N+\gamma+\frac{1}{2N},

yields after a small numerical computation, |ζ(1+it)|logt0.45|\zeta(1+it)|\leq\log t-0.45 for t3t\geq 3.

Patel [Pat22] replaced the use of the Euler–Maclaurin summation by an explicit version of the Riemann–Siegel formula on the 11-line. Consequently, the size of the main term in the Euler–Maclaurin bound was cut by half. Patel thus obtained |ζ(1+it)|12logt+1.93|\zeta(1+it)|\leq\frac{1}{2}\log t+1.93 for t3t\geq 3, a substantial improvement at the expense of a larger constant term.

In general, we expect that an explicit van der Corput lemma of order kk, where k2k\geq 2, will produce a bound of the form

(3) |ζ(1+it)|1klogt+αk,(t3).|\zeta(1+it)|\leq\frac{1}{k}\log t+\alpha_{k},\qquad(t\geq 3).

One proceeds by approximating ζ(1+it)\zeta(1+it) by an Euler–Maclaurin sum ntn1it\sum_{n\ll t}n^{-1-it}, or better yet by a Riemann–Siegel sum ntn1it\sum_{n\ll\sqrt{t}}n^{-1-it}. In either case, roughly speaking, the term (1/k)logt(1/k)\log t arises from bounding the subsum with nt1/kn\leq t^{1/k} via the triangle inequality and the harmonic sum estimate (2), while the constant αk\alpha_{k} predominantly arises from bounding the subsum with n>t1/kn>t^{1/k} via van der Corput lemmas of order k\leq k.

However, the αk\alpha_{k} obtained using this approach appear to grow fast with kk. For example, in [Pat21], α5\alpha_{5} is about 2323-times larger than α2\alpha_{2}, which already suggests exponential growth. Despite this, there is still gain in letting kk grow with tt, provided the growth is slow enough. We let kk grow like loglogt\log\log t, which is permitted by a recent optimized van der Corput test, uniform in kk, derived in [Yan24]. Subsequently, we obtain a bound on |ζ(1+it)||\zeta(1+it)| of the form c0logt/loglogtc_{0}\log t/\log\log t with an explicit number c0c_{0}.

While we could let kk increase faster than loglogt\log\log t, leading to a smaller contribution of the (1/k)logt(1/k)\log t term, the αk\alpha_{k} thus obtained would likely take over as the main term, asymptotically surpassing logt/loglogt\log t/\log\log t in size. This limits the order of the van der Corput lemmas we employ.

In proving Theorem 1, we encounter an unexpected “gap interval”, not covered by any of the other explicit estimates, which requires a special treatment. Namely, the interval I=[exp(16),exp(662)]I=[\exp(16),\exp(662)]. To prove the inequality in Theorem 1 for all tIt\in I, we use the following proposition.

Proposition 3.

We have

|ζ(1+it)|{13logt+4.66,texp(16),833logt+12.45,texp(88).|\zeta(1+it)|\leq\begin{cases}\displaystyle\frac{1}{3}\log t+4.66,&t\geq\exp(16),\\ &\\ \displaystyle\frac{8}{33}\log t+12.45,&t\geq\exp(88).\end{cases}

To clarify the need for Proposition 3, consider that if one inputs the bounds in [Pat22], as below, one finds

maxtImin(logt,12logt+1.93,15logt+44.02)logt/loglogt2.539,\max_{t\in I}\frac{\displaystyle\min\left(\log t,\frac{1}{2}\log t+1.93,\frac{1}{5}\log t+44.02\right)}{\log t/\log\log t}\geq 2.539,

and the maximum of this ratio occurs around t1=exp(140.3)t_{1}=\exp(140.3). On the other hand, the savings we otherwise establish here may materialize only when tt is well past t1t_{1}. Therefore, without the improvement enabled by Proposition 3, we cannot do any better than the bound |ζ(1+it)|2.539logt/loglogt|\zeta(1+it)|\leq 2.539\log t/\log\log t over tIt\in I. Proposition 3, which is based on a hybrid of van der Corput lemmas of orders k=3k=3, 44 and 55, allows us to circumvent this barrier.

In retrospect, a gap interval like II should be expected. This is because the estimates in [Pat21] rely, essentially, on van der Corput lemmas of orders k=2k=2 and k=5k=5.444The bound obtained via the Riemann–Siegel formula in [Pat21] should correspond to a k=2k=2 van der Corput lemma. In particular, bounds that would follow from using van der Corput lemmas of orders k=3k=3 and k=4k=4 are not utilized, which presumably causes the gap.

3. Proof of Theorem 1

Let t0t_{0} be a positive number to be chosen later, subject to t0exp(990/7)t_{0}\geq\exp(990/7). We suppose tt0t\geq t_{0} throughout this section. (We will use another method for t<t0t<t_{0}.)

Theorem 9 supplies a Riemann–Siegel approximation of ζ(1+it)\zeta(1+it) in terms of two sums, plus a remainder term \mathcal{R}. We use the triangle inequality to bound the entire second sum in this theorem. This gives

(4) |ζ(1+it)||n=1n11n1+it|+g(t)2π+(t),n1=t/(2π).\begin{split}|\zeta(1+it)|&\leq\left|\sum_{n=1}^{n_{1}}\frac{1}{n^{1+it}}\right|+\frac{g(t)}{\sqrt{2\pi}}+\mathcal{R}(t),\qquad n_{1}=\lfloor\sqrt{t/(2\pi)}\rfloor.\end{split}

Since g(t)g(t) and (t)\mathcal{R}(t) are well-understood, the bulk of the work is in bounding the main sum over nn1n\leq n_{1}.

We introduce two sequences, {Xm}0mM\{X_{m}\}_{0\leq m\leq M} and {Yk}3kr+1\{Y_{k}\}_{3\leq k\leq r+1}. The former sequence XmX_{m} will be used to partition the main sum in (4) in a dyadic manner. The latter sequence YkY_{k} will be used to easily bound the partition points XmX_{m} from above and below solely in terms of tt (or, what is essentially the same, in terms of n1n_{1}).

To this end, let hh and h2h_{2} be real positive numbers, also to be chosen later, subject to 1<h21<h\leq 2 and h2<1/log2h_{2}<1/\log 2. We use h2h_{2} immediately in defining

R:=2r1,r:=h2loglogt.\begin{split}&R:=2^{r-1},\qquad r:=\lfloor h_{2}\log\log t\rfloor.\end{split}

Furthermore we will assume that h2h_{2} and t0t_{0} are chosen so that

(5) r0:=h2loglogt06.r_{0}:=\lfloor h_{2}\log\log t_{0}\rfloor\geq 6.

Since tt0t\geq t_{0} by supposition, rr0r\geq r_{0}. The integer rr will correspond to the highest order van der Corput lemma we use in bounding the tail of the main sum in (4).

Furthermore, let

K:=2k1,θk:=2K(k2)K+4,(k3).\begin{split}&K:=2^{k-1},\qquad\theta_{k}:=\frac{2K}{(k-2)K+4},\qquad(k\geq 3).\end{split}

In the definition of θk\theta_{k}, KK is determined by kk. Also, θk\theta_{k} decreases monotonically to zero with k3k\geq 3, and is bounded from above by θ3=1\theta_{3}=1.

We now define the previously mentioned sequences {Xm}\{X_{m}\} and {Yk}\{Y_{k}\}. Set

X0:=12πt1/2,Xm:=hmX0,(m1).\begin{split}&X_{0}:=\frac{1}{\sqrt{2\pi}}t^{1/2},\qquad X_{m}:=h^{-m}X_{0},\qquad(m\geq 1).\end{split}

So, X0=n1\lfloor X_{0}\rfloor=n_{1} and since h>1h>1, XmX_{m} decreases monotonically with mm. Also set

(6) Yk:=X0θk,(k3),Y_{k}:=X_{0}^{\theta_{k}},\qquad(k\geq 3),

so that YkY_{k} decreases monotonically with k3k\geq 3, starting at Y3=X0Y_{3}=X_{0}. Note that if kk is large, then θk2/k\theta_{k}\approx 2/k, and so Ykt1/kY_{k}\approx t^{1/k} for large kk. Lastly, define

M:=(1θr+1)logX0logh.M:=\left\lceil\frac{(1-\theta_{r+1})\log X_{0}}{\log h}\right\rceil.

The integer MM corresponds to the number of points we will use to partition the main sum. From the definition, we have

(7) (1θr+1)logX0loghM<(1θr+1)logX0logh+1.\frac{(1-\theta_{r+1})\log X_{0}}{\log h}\leq M<\frac{(1-\theta_{r+1})\log X_{0}}{\log h}+1.

Thus, our choice of MM ensures that

(8) XMX0θr+1=Yr+1<XM1hX0θr+1.X_{M}\leq X_{0}^{\theta_{r+1}}=Y_{r+1}<X_{M-1}\leq hX_{0}^{\theta_{r+1}}.

Additionally, we have

(9) XMYr+1<Yr<<Y4<Y3=X0.X_{M}\leq Y_{r+1}<Y_{r}<\cdots<Y_{4}<Y_{3}=X_{0}.

The assumption (5) is required for the subsequent analysis to be valid. This condition implies, among other things, that θr+1<1\theta_{r+1}<1 and, hence, M1M\geq 1. This in turn ensures that some quantities appearing later (e.g. M1M-1) are nonnegative. All contingent conditions we impose (including on r0r_{0}, hh, h2h_{2} and t0t_{0}) will be verified at the end, after choosing the values of our free parameters.

By the chains of inequalities (8) and (9), for all m[1,M1]m\in[1,M-1], Yr+1<Xm<Y3Y_{r+1}<X_{m}<Y_{3}. Therefore, since the sequence Yr+1,,Y3Y_{r+1},\ldots,Y_{3} partitions the interval [Yr+1,Y3][Y_{r+1},Y_{3}], for each integer m[1,M1]m\in[1,M-1], there is a unique integer k[3,r]k\in[3,r] such that

(10) Xm(Yk+1,Yk].X_{m}\in(Y_{k+1},Y_{k}].

Note, though, that multiple mm’s could correspond to the same kk.

With this in mind, let us denote

(11) Sm:=Xm+1<nXm11n1+it=Xm+1<nXm11n1+it,S_{m}:=\sum_{X_{m}+1<n\leq X_{m-1}}\frac{1}{n^{1+it}}=\sum_{\lfloor X_{m}\rfloor+1<n\leq\lfloor X_{m-1}\rfloor}\frac{1}{n^{1+it}},

and observe that

(12) |XM1<nX01n1+it|m=1M1|1(Xm+1)1+it+Sm|m=1M1(1Xm+|Sm|).\left|\sum_{X_{M-1}<n\leq X_{0}}\frac{1}{n^{1+it}}\right|\leq\sum_{m=1}^{M-1}\left|\frac{1}{(\lfloor X_{m}\rfloor+1)^{1+it}}+S_{m}\right|\leq\sum_{m=1}^{M-1}\left(\frac{1}{X_{m}}+|S_{m}|\right).

We calculate

(13) Xm1Xm+1<Xm1Xm=h.\frac{\lfloor X_{m-1}\rfloor}{\lfloor X_{m}\rfloor+1}<\frac{X_{m-1}}{X_{m}}=h.

Here, we used the elementary inequality x1<xxx-1<\lfloor x\rfloor\leq x, valid for any real number xx together with the definition Xm=hmX0X_{m}=h^{-m}X_{0}.

We bound SmS_{m} using Lemma 13 with a=Xm+1a=\lfloor X_{m}\rfloor+1 and b=Xm1hab=\lfloor X_{m-1}\rfloor\leq ha. For each integer m[1,M1]m\in[1,M-1] such that Xm(Yk+1,Yk]X_{m}\in(Y_{k+1},Y_{k}], and with K=2k1K=2^{k-1}, we have, for any choice of the free parameter η>0\eta>0,

(14) |Sm|Ck(η,h)(Xm+1)k/(2K2)t1/(2K2)+Dk(η,h)(Xm+1)k/(2K2)2/Kt1/(2K2).\begin{split}\left|S_{m}\right|&\leq C_{k}(\eta,h)(\lfloor X_{m}\rfloor+1)^{-k/(2K-2)}t^{1/(2K-2)}\\ &\qquad+D_{k}(\eta,h)(\lfloor X_{m}\rfloor+1)^{k/(2K-2)-2/K}t^{-1/(2K-2)}.\end{split}

To bound the RHS, observe that

(15) Yk+1<Xm+1Xm+1=Xm(1+1Xm)Yk(1+1Xm).Y_{k+1}<\lfloor X_{m}\rfloor+1\leq X_{m}+1=X_{m}\left(1+\frac{1}{X_{m}}\right)\leq Y_{k}\left(1+\frac{1}{X_{m}}\right).

In addition, XmXM1>X0θr+1X_{m}\geq X_{M-1}>X_{0}^{\theta_{r+1}}. Moreover, it follows by definition that

(16) θr+1=2r1+22r.\theta_{r+1}=\frac{2}{r-1+2^{2-r}}.

Since θr+1\theta_{r+1} decreases continuously with rr, and rh2loglogtr\leq h_{2}\log\log t, we have X0θr+1=exp(θr+1logX0)κ(h2,t)X_{0}^{\theta_{r+1}}=\exp(\theta_{r+1}\log X_{0})\geq\kappa(h_{2},t) where

(17) κ(h2,t):=exp(log(t/(2π))h2loglogt1+22h2loglogt).\kappa(h_{2},t):=\exp\left(\frac{\log(t/(2\pi))}{h_{2}\log\log t-1+2^{2-h_{2}\log\log t}}\right).

It is straightforward to verify that κ(h2,t)\kappa(h_{2},t) is increasing in tt for tt0t\geq t_{0}, so κ(h2,t)κ(h2,t0)\kappa(h_{2},t)\geq\kappa(h_{2},t_{0}). Inserting (15) and (17) into (14),

(18) |Sm|Ck(η,h)Yk+1k/(2K2)t1/(2K2)+D^k(η,h,h2,t0)Ykk/(2K2)2/Kt1/(2K2)\begin{split}\left|S_{m}\right|&\leq C_{k}(\eta,h)Y_{k+1}^{-k/(2K-2)}t^{1/(2K-2)}\\ &\qquad+\widehat{D}_{k}(\eta,h,h_{2},t_{0})Y_{k}^{k/(2K-2)-2/K}t^{-1/(2K-2)}\end{split}

for tt0t\geq t_{0}, where

(19) D^k(η,h,h2,t0):=Dk(η,h)(1+1κ(h2,t0))k/(2K2)2/K.\widehat{D}_{k}(\eta,h,h_{2},t_{0}):=D_{k}(\eta,h)\left(1+\frac{1}{\kappa(h_{2},t_{0})}\right)^{k/(2K-2)-2/K}.

This is permissible since the exponents of Xm+1\lfloor X_{m}\rfloor+1 in the first and second terms of (14) respectively satisfy

(20) k2K2<0,andk2K22K0,(k3).-\frac{k}{2K-2}<0,\qquad\text{and}\qquad\frac{k}{2K-2}-\frac{2}{K}\geq 0,\qquad(k\geq 3).

It is worth clarifying that the right-hand side of (18) depends on mm via the requirement (10).

Next, we express the bound on SmS_{m} in (18) solely in terms of tt. Starting with the first term, we have

(21) Yk+1k/(2K2)t1/(2K2)=(2π)θk+1k/(4K4)tρ(k),\displaystyle Y_{k+1}^{-k/(2K-2)}t^{1/(2K-2)}=(2\pi)^{\theta_{k+1}k/(4K-4)}t^{-\rho(k)},

where after a quick calculation,

(22) ρ(k):=K2K112(k1)K+4.\rho(k):=\frac{K-2}{K-1}\cdot\frac{1}{2(k-1)K+4}.

We estimate the second term in (18) similarly, yielding

(23) Ykk/(2K2)2/Kt1/(2K2)<(2π)θk(k/(4K4)1/K)tKK112(k2)K+4<(2π)θk(k/(4K4)1/K)tρ(k).\begin{split}Y_{k}^{k/(2K-2)-2/K}t^{-1/(2K-2)}&<(2\pi)^{-\theta_{k}(k/(4K-4)-1/K)}t^{-\frac{K}{K-1}\cdot\frac{1}{2(k-2)K+4}}\\ &<(2\pi)^{-\theta_{k}(k/(4K-4)-1/K)}t^{-\rho(k)}.\end{split}

The exponent of 2π2\pi in the estimate (21) may be bounded for k>5k>5 via θk+132/81\theta_{k+1}\leq 32/81. Additionally, recalling the second inequality of (20), we may bound the contribution of the 2π2\pi-factor in the estimate (23) simply by 11. This motivates us to define

(24) A(k):={(2π)θk+1k4(K1)Ck+(2π)θk(k4(K1)1K)D^k,3k5,(2π)8k81(K1)Ck+D^k,6k.A(k):=\begin{cases}(2\pi)^{\theta_{k+1}\frac{k}{4(K-1)}}C_{k}+(2\pi)^{-\theta_{k}(\frac{k}{4(K-1)}-\frac{1}{K})}\widehat{D}_{k},&3\leq k\leq 5,\\ (2\pi)^{\frac{8k}{81(K-1)}}C_{k}+\widehat{D}_{k},&6\leq k.\end{cases}

Note that A(k)A(k) additionally depends on η\eta, hh, h2h_{2} and t0t_{0}. Assembling (18), (21), (23) and (24), we obtain that if 1mM11\leq m\leq M-1 and kk is determined via the requirement (10), then

(25) |Sm|<A(k)tρ(k).|S_{m}|<A(k)\,t^{-\rho(k)}.

We now divide the argument into two cases: 3k53\leq k\leq 5 and k>5k>5. In the former case, let NkN_{k} denote the number of sums SmS_{m} corresponding to each integer k3k\geq 3. In other words, NkN_{k} is the number of mm such that Xm(Yk+1,Yk]X_{m}\in(Y_{k+1},Y_{k}]. We observe that Xm(Yk+1,Yk]X_{m}\in(Y_{k+1},Y_{k}] if and only if Mkm<Mk+1M_{k}\leq m<M_{k+1}, where

(26) Mk:=(1θk)logX0logh.M_{k}:=\frac{(1-\theta_{k})\log X_{0}}{\log h}.

Therefore, by the bound on |Sm||S_{m}| in (25),

(27) 1m<M6|Sm|<k=35NkA(k)tρ(k).\sum_{1\leq m<M_{6}}|S_{m}|<\sum_{k=3}^{5}N_{k}\,A(k)\,t^{-\rho(k)}.

By counting the maximum number of integers in a continuous interval, for k3k\geq 3,

(28) Nk(θkθk+1)logX0logh+1ϕklogt,N_{k}\leq\frac{(\theta_{k}-\theta_{k+1})\log X_{0}}{\log h}+1\leq\phi_{k}\log t,

where

(29) ϕk=ϕk(h,t0):=θkθk+12logh+1logt0.\phi_{k}=\phi_{k}(h,t_{0}):=\frac{\theta_{k}-\theta_{k+1}}{2\log h}+\frac{1}{\log t_{0}}.

Thus, in view of (27) we obtain

(30) 1m<M6|Sm|b0,\sum_{1\leq m<M_{6}}|S_{m}|\leq b_{0},

where by (28) and on explicitly computing the values of ρ(k)\rho(k) for 3k53\leq k\leq 5,

(31) b0=b0(η,h,h2,t0):=A(3)ϕ3logt0t01/30+A(4)ϕ4logt0t03/182+A(5)ϕ5logt0t07/990.\begin{split}b_{0}=b_{0}(\eta,h,h_{2},t_{0})&:=A(3)\phi_{3}\frac{\log t_{0}}{t_{0}^{1/30}}+A(4)\phi_{4}\frac{\log t_{0}}{t_{0}^{3/182}}+A(5)\phi_{5}\frac{\log t_{0}}{t_{0}^{7/990}}.\end{split}

Here, we used the supposition tt0exp(990/7)t\geq t_{0}\geq\exp(990/7) which ensures that logt/tρ(k)\log t/t^{\rho(k)} is decreasing for tt0t\geq t_{0}, with 3k53\leq k\leq 5.

Next, for k>5k>5, we use a tail argument that removes the dependence of AA on kk in the bound (25). Define

(32) A0=A0(η,h,h2,t0):=supk6A(k).A_{0}=A_{0}(\eta,h,h_{2},t_{0}):=\sup_{k\geq 6}A(k).

We will bound A0A_{0} explicitly later (Section 4), after determining η\eta, hh, h2h_{2} and t0t_{0}. Importantly, under our assumptions, it will transpire that A0A_{0} is finite.

In the meantime, we observe that ρ(k)\rho(k) is continuously decreasing in k3k\geq 3 (towards zero), as can be seen from computing the derivative of ρ(k)\rho(k) with respect to kk. So, by simply counting the number of terms, and noting that as mm ranges over [M6,M1][M_{6},M-1], krk\leq r, we obtain from (25),

(33) M6mM1|Sm||MM6|A0tρ(r)A1|b1θr+1|tρ(r)logt,\begin{split}\sum_{M_{6}\leq m\leq M-1}|S_{m}|\leq|M-M_{6}|A_{0}\,t^{-\rho(r)}\leq A_{1}\,\frac{|b_{1}-\theta_{r+1}|}{t^{\rho(r)}}\log t,\end{split}

where, from (7) and (26),

A1=A1(η,h,h2,t0):=12A0logh,b1=b1(h,t0):=θ6+2loghlogt0.A_{1}=A_{1}(\eta,h,h_{2},t_{0}):=\frac{1}{2}\cdot\frac{A_{0}}{\log h},\qquad b_{1}=b_{1}(h,t_{0}):=\theta_{6}+\frac{2\log h}{\log t_{0}}.

We want to further simplify the bound on the right-side of (33). Via (16) and a routine calculation,

(34) 2r1+22r0θr+1<b1,(rr06).\frac{2}{r-1+2^{2-r_{0}}}\leq\theta_{r+1}<b_{1},\qquad(r\geq r_{0}\geq 6).

Hence, utilizing the inequality rh2loglogtr\leq h_{2}\log\log t as well as log(u)u1\log(u)\leq u-1, valid for any real number u>0u>0, yields that if r6r\geq 6, then

(35) b1θr+1tρ(r)exp(b1θr+11ρ(r)logt)1A21logt,\begin{split}\frac{b_{1}-\theta_{r+1}}{t^{\rho(r)}}\leq\exp(b_{1}-\theta_{r+1}-1-\rho(r)\log t)\leq\frac{1}{A_{2}}\frac{1}{\log t},\end{split}

where (identifying xx with logt\log t below)

(36) A2=A2(h,h2,t0):=infxlogt01xexp(1b1+2h2logx1+22r0+xρ(h2logx)).A_{2}=A_{2}(h,h_{2},t_{0}):=\inf_{x\geq\log t_{0}}\frac{1}{x}\exp\left(1-b_{1}+\frac{2}{h_{2}\log x-1+2^{2-r_{0}}}+x\,\rho(h_{2}\log x)\right).

In indicating the dependencies of A2A_{2}, we used that r0r_{0} is determined by h2h_{2} and t0t_{0}, and b1b_{1} further depends on hh. Now, from the definition of ρ\rho, and using

2h2logx=xh2log2,2^{h_{2}\log x}=x^{h_{2}\log 2},

we obtain555By the lower bound on h2h_{2} in (5), we have xh2log264x^{h_{2}\log 2}\geq 64 for xlogt0x\geq\log t_{0}. So, ρ(h2logx)\rho(h_{2}\log x) is well-defined for xlogt0x\geq\log t_{0}.

(37) ρ(h2logx)=xh2log24xh2log221(h2logx1)xh2log2+4.\rho(h_{2}\log x)=\frac{x^{h_{2}\log 2}-4}{x^{h_{2}\log 2}-2}\cdot\frac{1}{(h_{2}\log x-1)x^{h_{2}\log 2}+4}.

Since666This step is where a slow enough growth on rr is utilized, to ensure that (logt)h2log2=o(logt)(\log t)^{h_{2}\log 2}=o(\log t) as tt\to\infty, so that the constant A2(h2,t0)A_{2}(h_{2},t_{0}) defined in (36) is positive. by supposition 0<h2<1/log20<h_{2}<1/\log 2, the quantity defining A2A_{2} tends to ++\infty as x+x\to+\infty. As this quantity is continuous and positive for xlogt0x\geq\log t_{0}, A2A_{2} is finite and positive for all admissible h2h_{2}. From (33) and (35) we conclude that

(38) M6mM1|Sm|A1A2.\sum_{M_{6}\leq m\leq M-1}|S_{m}|\leq\frac{A_{1}}{A_{2}}.

Lastly, to bound the remaining terms in (12), in view of (7), we have the rough estimate

m=1M11Xm=1X0m=1M1hm=1X0hMhh1<X01θr+11X0hh1<hh11X0θr+1.\sum_{m=1}^{M-1}\frac{1}{X_{m}}=\frac{1}{X_{0}}\sum_{m=1}^{M-1}h^{m}=\frac{1}{X_{0}}\frac{h^{M}-h}{h-1}<\frac{X_{0}^{1-\theta_{r+1}}-1}{X_{0}}\frac{h}{h-1}<\frac{h}{h-1}\frac{1}{X_{0}^{\theta_{r+1}}}.

Since rh2loglogtr\leq h_{2}\log\log t and θr+1>2/r\theta_{r+1}>2/r, we have, for tt0t\geq t_{0},

(39) m=1M11Xm<hh1exp(log(t0/(2π))h2loglogt0)=:E(h,h2,t0),\sum_{m=1}^{M-1}\frac{1}{X_{m}}<\frac{h}{h-1}\exp\left(-\frac{\log(t_{0}/(2\pi))}{h_{2}\log\log t_{0}}\right)=:E(h,h_{2},t_{0}),

say.

In summary, combining (12), (30), (38) and (39), the tail sum satisfies the inequality

(40) |XM1<nX01n1+it|A1(η,h,h2,t0)A2(h,h2,t0)+b0(η,h,h2,t0)+E(h,h2,t0).\begin{split}\left|\sum_{X_{M-1}<n\leq X_{0}}\frac{1}{n^{1+it}}\right|&\leq\frac{A_{1}(\eta,h,h_{2},t_{0})}{A_{2}(h,h_{2},t_{0})}+b_{0}(\eta,h,h_{2},t_{0})+E(h,h_{2},t_{0}).\end{split}

So, for such rr, the tail sum is bounded by a constant independent of tt.

It remains to bound the initial sum over 1nXM11\leq n\leq X_{M-1}. We use the triangle inequality followed by the harmonic sum bound (2), which yields

(41) |1nXM11n1+it|logXM1+γ+12(XM11).\begin{split}\left|\sum_{1\leq n\leq X_{M-1}}\frac{1}{n^{1+it}}\right|&\leq\log X_{M-1}+\gamma+\frac{1}{2(X_{M-1}-1)}.\end{split}

We use the inequality (8) to bound XM1X_{M-1} from above and below, so that the right-side of (41) satisfies

(42) θr+12logt2π+logh+γ+12(X0θr+11).\begin{split}\leq\frac{\theta_{r+1}}{2}\log\frac{t}{2\pi}+\log h+\gamma+\frac{1}{2(X_{0}^{\theta_{r+1}}-1)}.\end{split}

Since θr+1<2/(r1)\theta_{r+1}<2/(r-1) by (16), the first term in (42) is bounded by

(43) log(t/(2π))r1<logtlog2πh2loglogt21log2πlogt0h22loglogt0logtloglogt,\frac{\log(t/(2\pi))}{r-1}<\frac{\log t-\log 2\pi}{h_{2}\log\log t-2}\leq\frac{1-\frac{\log 2\pi}{\log t_{0}}}{h_{2}-\frac{2}{\log\log t_{0}}}\frac{\log t}{\log\log t},

where the last inequality is due to the fact that (1log2πx)/(h22logx)(1-\frac{\log 2\pi}{x})/(h_{2}-\frac{2}{\log x}) is decreasing for x990/7x\geq 990/7 if h2<1/log2h_{2}<1/\log 2.

In addition, we use (17) to bound the last term. Putting this together with (41), (42) and (43), we obtain

(44) |1nXM11n1+it|<A3logtloglogt,\left|\sum_{1\leq n\leq X_{M-1}}\frac{1}{n^{1+it}}\right|<A_{3}\frac{\log t}{\log\log t},

where

A3=A3(h,h2,t0):=1log2πlogt0h22loglogt0+(logh+γ+12(κ(h2,t0)1))loglogt0logt0.\begin{split}A_{3}=A_{3}(h,h_{2},t_{0})&:=\frac{1-\frac{\log 2\pi}{\log t_{0}}}{h_{2}-\frac{2}{\log\log t_{0}}}+\left(\log h+\gamma+\frac{1}{2(\kappa(h_{2},t_{0})-1)}\right)\frac{\log\log t_{0}}{\log t_{0}}.\end{split}

Lastly, the bound on ζ(1+it)\zeta(1+it) in (4) carries an extra term g(t)/2π+(t)g(t)/\sqrt{2\pi}+\mathcal{R}(t). Using the definitions in Theorem 9, this term is easily bounded by g(t0)/2π+(t0)g(t_{0})/\sqrt{2\pi}+\mathcal{R}(t_{0}). Therefore, combining this with (40) and (44), we see that

|ζ(1+it)||1nn11n1+it|+g(t0)2π+(t0)A4logtloglogt,|\zeta(1+it)|\leq\left|\sum_{1\leq n\leq n_{1}}\frac{1}{n^{1+it}}\right|+\frac{g(t_{0})}{\sqrt{2\pi}}+\mathcal{R}(t_{0})\leq A_{4}\frac{\log t}{\log\log t},

for tt0t\geq t_{0}, where

A4=A4(η,h,h2,t0):=A3+(A1A2+b0+E+g(t0)2π+(t0))loglogt0logt0.A_{4}=A_{4}(\eta,h,h_{2},t_{0}):=A_{3}+\left(\frac{A_{1}}{A_{2}}+b_{0}+E+\frac{g(t_{0})}{\sqrt{2\pi}}+\mathcal{R}(t_{0})\right)\frac{\log\log t_{0}}{\log t_{0}}.

We choose the following values for our free parameters, which are suggested by numerical experimentation.

(45) η=1.2364,h=1.0224,h2=0.85532,t0=exp(3381).\eta=1.2364,\qquad h=1.0224,\qquad h_{2}=0.85532,\qquad t_{0}=\exp(3381).

With these choices, our preconditions on t0t_{0} and h2h_{2} hold, and the conditions h>1h>1 and r06r_{0}\geq 6 also hold (in fact, r0=6r_{0}=6). Additionally, in Section 4, we show that

A0\displaystyle A_{0} =0.0976483443,\displaystyle=0.0976483443\ldots,
A2\displaystyle A_{2} =0.0618341521.\displaystyle=0.0618341521\ldots.

Therefore, we can explicitly compute

b0\displaystyle b_{0} 4.629108,\displaystyle\leq 4.629\cdot 10^{-8},
A1\displaystyle A_{1} 2.2039724975,\displaystyle\leq 2.2039724975\ldots,
A3\displaystyle A_{3} 1.6420608139,\displaystyle\leq 1.6420608139\ldots,
A4\displaystyle A_{4} 1.7301296329.\displaystyle\leq 1.7301296329\ldots.

In summary, the assertion of Theorem 1 holds when tt0=exp(3381)t\geq t_{0}=\exp(3381).

For smaller tt0t\leq t_{0} we use the bound in Patel [Pat22] and Proposition 3. Specifically, the bound in [Pat22, Theorem 1.1] yields

|ζ(1+it)|15logt+44.021.731logtloglogt,exp(662)texp(3381),|ζ(1+it)|12logt+1.931.731logtloglogt,3texp(16).\begin{split}&|\zeta(1+it)|\leq\frac{1}{5}\log t+44.02\leq 1.731\frac{\log t}{\log\log t},\qquad\exp(662)\leq t\leq\exp(3381),\\ &|\zeta(1+it)|\leq\frac{1}{2}\log t+1.93\leq 1.731\frac{\log t}{\log\log t},\qquad 3\leq t\leq\exp(16).\end{split}

This leaves the gap interval exp(16)texp(662)\exp(16)\leq t\leq\exp(662) for which we invoke Proposition 3. Since the gap interval falls in the range of tt where Proposition 3 is applicable, our target bound is verified there as well.

4. Bounding A0A_{0} and A2A_{2}

We derive an upper bound on A0(η,h,h2,t0)A_{0}(\eta,h,h_{2},t_{0}), where the parameter values are given in (45). For the convenience of the reader, we recall the definition A0:=supk6A(k)A_{0}:=\sup_{k\geq 6}A(k), where for k6k\geq 6,

A(k):=(2π)k/(81K81)Ck(η,h)+D^k(η,h,h2,t0).A(k):=(2\pi)^{k/(81K-81)}C_{k}(\eta,h)+\widehat{D}_{k}(\eta,h,h_{2},t_{0}).

With our choice of free parameter values in (45), we numerically compute the following list of A(k)A(k).

A(6)\displaystyle A(6) =0.0976483443,A(7)=0.0914388403,\displaystyle=0.0976483443\ldots,\qquad A(7)=0.0914388403\ldots,
A(8)\displaystyle A(8) =0.0882612019,A(9)=0.0865130096,\displaystyle=0.0882612019\ldots,\qquad A(9)=0.0865130096\ldots,
A(10)\displaystyle A(10) =0.0855306439.\displaystyle=0.0855306439\ldots.

The values of CkC_{k} and DkD_{k} required for the above computation are obtained using the formulas in Lemma 13. In turn, evaluating Ck(η,h)C_{k}(\eta,h) and Dk(η,h)D_{k}(\eta,h) relies on evaluating Ak(η,hk)A_{k}(\eta,h^{k}) and Bk(η)B_{k}(\eta). When k=3k=3, AkA_{k} and BkB_{k} are computed directly using their definitions in Lemma 11. When k4k\geq 4, the recursive formulas in Lemma 12 are used to compute them. Let us examine these recursive formulas.

First, the following quantity, which appears in the formula for Ak+1(η,ω)A_{k+1}(\eta,\omega) in (92), satisfies

(46) maxk10219/12(K1)(2K1)(4K3)=219/125112092035=:μ10.\max_{k\geq 10}\frac{2^{19/12}(K-1)}{\sqrt{(2K-1)(4K-3)}}=\frac{2^{19/12}\cdot 511}{\sqrt{2092035}}=:\mu_{10}.

Here, K=2k1K=2^{k-1} and the maximum occurs when k=10k=10. Another quanitity appearing in the formula for Ak+1(η,ω)A_{k+1}(\eta,\omega) is δk\delta_{k}, which is clearly seen to decrease with kk. Noting that h>1h>1, hence ω=hk+1>1\omega=h^{k+1}>1 for any k0k\geq 0, we deduce that

Ak+1(η,ω)δ10(1+μ10Ak(η,ω)1/2).A_{k+1}(\eta,\omega)\leq\delta_{10}\left(1+\mu_{10}A_{k}(\eta,\omega)^{1/2}\right).

So, we are led to consider the discrete map

xn+1=δ10(1+μ10xn1/2).x_{n+1}=\delta_{10}\left(1+\mu_{10}x_{n}^{1/2}\right).

By a routine calculation, this map has a single fixed point at

(47) x=(μ10δ102+μ102δ1024+δ10)2=2.7600429449,x^{*}=\left(\frac{\mu_{10}\delta_{10}}{2}+\sqrt{\frac{\mu_{10}^{2}\delta_{10}^{2}}{4}+\delta_{10}}\right)^{2}=2.7600429449\ldots,

which is a stable point. Since Ak+1(η,ω)xk+1A_{k+1}(\eta,\omega)\leq x_{k+1} for k10k\geq 10 and, by a direct numerical computation, A10(η,ω)xA_{10}(\eta,\omega)\leq x^{*}, it follows that

(48) Ak(η,hk)x,(k10).A_{k}(\eta,h^{k})\leq x^{*},\qquad(k\geq 10).

The analysis of Bk+1B_{k+1} for k10k\geq 10 proceeds similarly. The recursive formula for Bk+1B_{k+1} contains the following quantity satisfying

(49) maxk1023/2(K1)(2K3)(4K5)23/25113231767=:ν10.\max_{k\geq 10}\frac{2^{3/2}(K-1)}{\sqrt{(2K-3)(4K-5)}}\leq\frac{2^{3/2}\cdot 511}{3\sqrt{231767}}=:\nu_{10}.

So, Bk+1δ10ν10Bk1/2B_{k+1}\leq\delta_{10}\nu_{10}B_{k}^{1/2} for k10k\geq 10. Consider the discrete map yn+1=δ10ν10yn1/2y_{n+1}=\delta_{10}\nu_{10}y_{n}^{1/2}. This map has a single fixed point, namely,

y=(δ10ν10)2=1.0023463404,y^{*}=(\delta_{10}\nu_{10})^{2}=1.0023463404\ldots,

which is also stable. Since Bk+1yk+1B_{k+1}\leq y_{k+1} for k10k\geq 10 and, by a direct numerical computation, B10yB_{10}\leq y^{*}, it follows that

(50) Bk(η)y,(k10).B_{k}(\eta)\leq y^{*},\qquad(k\geq 10).

Furthermore, we numerically verify that for k10k\geq 10,

h2k/Kk/(2K2)(h1)((k1)!2π)12K2<0.023,h^{2k/K-k/(2K-2)}(h-1)\left(\frac{(k-1)!}{2\pi}\right)^{\frac{1}{2K-2}}<0.023,
hk/(2K2)(h1)12/K(2π(k1)!)12K2<0.023.h^{k/(2K-2)}(h-1)^{1-2/K}\left(\frac{2\pi}{(k-1)!}\right)^{\frac{1}{2K-2}}<0.023.

Combining this with (48) and (50), and recalling the definitions in Lemma 13 and (19), yields the inequalities Ck(η,h)<0.023xC_{k}(\eta,h)<0.023x^{*} and D^k(η,h)<0.024y\widehat{D}_{k}(\eta,h)<0.024y^{*} for k10k\geq 10. Consequently, A(k)<0.09A(k)<0.09 for k10k\geq 10.

Furthermore, in view of the values of A(k)A(k) listed at the beginning, we also see that A(k)A(6)=0.09764A(k)\leq A(6)=0.09764\ldots for 6k106\leq k\leq 10. We therefore conclude A0=A(6)A_{0}=A(6).

For the computation related to A2A_{2}, we shall use a derivative calculation to show that the function

(51) a2(u):=1b1+euρ(h2u)+2h2u1+22r0ua_{2}(u):=1-b_{1}+e^{u}\,\rho(h_{2}u)+\frac{2}{h_{2}u-1+2^{2-r_{0}}}-u

is increasing in uu0:=loglogt0u\geq u_{0}:=\log\log t_{0}. So, on identifying uu with logx\log x in the definition of A2A_{2} in (36), the value of A2A_{2} is exp(a2(u0))\exp(a_{2}(u_{0})).

To this end, we first calculate

ddueuρ(h2u)=euρ(h2u)(1+ddulogρ(h2u)),ddulogρ(h2u)=h2a~2(h2u),\frac{d}{du}e^{u}\,\rho(h_{2}u)=e^{u}\rho(h_{2}u)\left(1+\frac{d}{du}\log\rho(h_{2}u)\right),\qquad\frac{d}{du}\log\rho(h_{2}u)=-h_{2}\tilde{a}_{2}(h_{2}u),

where

a~2(v)=(v1)log2+12v(8(v1)log2+6)+4v(8(v2)log2+8)(v1+22v)(121v)(122v).\tilde{a}_{2}(v)=\frac{(v-1)\log 2+1-2^{-v}(8(v-1)\log 2+6)+4^{-v}(8(v-2)\log 2+8)}{(v-1+2^{2-v})(1-2^{1-v})(1-2^{2-v})}.

Note that vv is identified with h2uh_{2}u, so we need only consider vh2u0=:v0v\geq h_{2}u_{0}=:v_{0}. It is easy to see that

supvv0a~2(v)<supvv0(v1)log2+1(v1+22v0)(121v0)(122v0)log2+0.2,\sup_{v\geq v_{0}}\tilde{a}_{2}(v)<\sup_{v\geq v_{0}}\frac{(v-1)\log 2+1}{(v-1+2^{2-v_{0}})(1-2^{1-v_{0}})(1-2^{2-v_{0}})}\leq\log 2+0.2,

where the supremum occurs at v=v0v=v_{0}. Hence, the derivative of euρ(h2u)e^{u}\,\rho(h_{2}u) is at least euρ(h2u)(1(log2+0.2)h2)e^{u}\rho(h_{2}u)(1-(\log 2+0.2)h_{2}) for uu0u\geq u_{0}. Importantly, 1(log2+0.2)h21-(\log 2+0.2)h_{2} is positive. In comparison, the derivative of the remaining terms in the formula for a2(u)a_{2}(u) is

ddu(2h2u1+22r0u)=2h2(h2u1+22r0)211.04731,\frac{d}{du}\left(\frac{2}{h_{2}u-1+2^{2-r_{0}}}-u\right)=-\frac{2h_{2}}{(h_{2}u-1+2^{2-r_{0}})^{2}}-1\geq-1.04731\ldots,

since uu0u\geq u_{0} and r0=6r_{0}=6. Since

(52) euρ(h2u)(1(log2+0.2)h2)1.06112e^{u}\rho(h_{2}u)(1-(\log 2+0.2)h_{2})\geq 1.06112\ldots

for uu0u\geq u_{0}, the derivative of a2(u)a_{2}(u) is positive for uu0u\geq u_{0}, proving our claim.

5. Proof of Proposition 3

The calculation is similar to that in [Hia16] on the 1/21/2-line, so we shall be brief in some details. We divide the main sum in (4) into pieces of length t1/3\approx t^{1/3}, then apply the van der Corput Lemma 10 to each piece other than an initial segment, which is bounded by the triangle inequality.

Let t1t_{1} be a positive number to be chosen later, subject to t1exp(12)t_{1}\geq\exp(12). Suppose that tt1t\geq t_{1}, unless otherwise stated, and let n1n_{1} be as defined in (4). Let P=t1/3P=\lceil t^{1/3}\rceil, Q=n1/PQ=\lfloor n_{1}/P\rfloor, and let q0q_{0} a positive integer to be chosen later, subject to q0Qq_{0}\leq Q. Thus, we have

(53) |n=1n11n1+it|n=1q0P11n+q=q0Q1|n=qP(q+1)P11n1+it|+|QPn11n1+it|.\begin{split}\left|\sum_{n=1}^{n_{1}}\frac{1}{n^{1+it}}\right|&\leq\sum_{n=1}^{q_{0}P-1}\frac{1}{n}+\sum_{q=q_{0}}^{Q-1}\left|\sum_{n=qP}^{(q+1)P-1}\frac{1}{n^{1+it}}\right|+\left|\sum_{QP}^{n_{1}}\frac{1}{n^{1+it}}\right|.\end{split}

The harmonic sum bound (2) yields, subject to q0P>1q_{0}P>1,

(54) n=1q0P11nlog(q0P1)+γ+12(q0P1).\sum_{n=1}^{q_{0}P-1}\frac{1}{n}\leq\log(q_{0}P-1)+\gamma+\frac{1}{2(q_{0}P-1)}.

For q0qQq_{0}\leq q\leq Q, let N=qPN=qP and let NN^{\prime} be an integer such that NN<N+PN\leq N^{\prime}<N+P. Partial summation gives

(55) |n=NN1n1+it|\displaystyle\left|\sum_{n=N}^{N^{\prime}}\frac{1}{n^{1+it}}\right| 1Nmax1LP|SN(L)|,\displaystyle\leq\frac{1}{N}\max_{1\leq L\leq P}|S_{N}(L)|,

where

SN(L):=n=0L1exp(2πif(n))andf(x)=t2πlog(N+x).S_{N}(L):=\sum_{n=0}^{L-1}\exp(2\pi if(n))\qquad\text{and}\qquad f(x)=\frac{t}{2\pi}\log(N+x).

We estimate SN(L)S_{N}(L) using Lemma 10 with r=qr=q, K=PK=P and

K0=t11/3P.K_{0}=t_{1}^{1/3}\leq P.

This gives, for any positive η\eta,

1Nmax1LP|SN(L)|\displaystyle\frac{1}{N}\max_{1\leq L\leq P}|S_{N}(L)| 1qαW1/3+βW1/3P+αηP+ηβW2/3P2.\displaystyle\leq\frac{1}{q}\sqrt{\frac{\alpha}{W^{1/3}}+\frac{\beta W^{1/3}}{P}+\frac{\alpha\eta}{P}+\frac{\eta\beta W^{2/3}}{P^{2}}}.

where WW, λ\lambda, α=α(W,λ)\alpha=\alpha(W,\lambda), β=β(W)\beta=\beta(W) are defined in Lemma 10. We estimate the (nonnegative) terms under the square-root as follows.

αW1/3<απ1/3q,\displaystyle\frac{\alpha}{W^{1/3}}<\frac{\alpha}{\pi^{1/3}q}, βW1/3P\displaystyle\frac{\beta W^{1/3}}{P} βπ1/3(1+1/q0)qt1/3,\displaystyle\leq\frac{\beta\pi^{1/3}(1+1/q_{0})q}{t^{1/3}},
αηPαηt1/3,\displaystyle\frac{\alpha\eta}{P}\leq\frac{\alpha\eta}{t^{1/3}}, ηβW2/3P2\displaystyle\frac{\eta\beta W^{2/3}}{P^{2}} ηβπ2/3(1+1/q0)2q2t2/3.\displaystyle\leq\frac{\eta\beta\pi^{2/3}(1+1/q_{0})^{2}q^{2}}{t^{2/3}}.

Isolating the first term using the inequality x+yx+y\sqrt{x+y}\leq\sqrt{x}+\sqrt{y}, valid for nonnegative xx and yy, and using the inequality qQt1/6/(2π)1/2q\leq Q\leq t^{1/6}/(2\pi)^{1/2} to bound qq in the remaining terms, therefore gives

(56) 1Nmax1LP|SN(L)|α(W,λ)1/2π1/6q3/2+B(W,λ)qt1/12,\frac{1}{N}\max_{1\leq L\leq P}|S_{N}(L)|\leq\frac{\alpha(W,\lambda)^{1/2}}{\pi^{1/6}q^{3/2}}+\frac{B(W,\lambda)}{qt^{1/12}},

where, since tt1t\geq t_{1},

B(W,λ):=1+1/q021/2π1/6β(W)+ηt11/6α(W,λ)+η(1+1/q0)22π1/3t11/6β(W).B(W,\lambda):=\sqrt{\frac{1+1/q_{0}}{2^{1/2}\pi^{1/6}}\beta(W)+\frac{\eta}{t_{1}^{1/6}}\alpha(W,\lambda)+\frac{\eta(1+1/q_{0})^{2}}{2\pi^{1/3}t_{1}^{1/6}}\beta(W)}.

We also note that for all qq0q\geq q_{0} we have WW0W\geq W_{0} and λλ0\lambda\leq\lambda_{0}, where

W0:=π(q0+1)3,λ0:=(1+1q0)3.\begin{split}W_{0}:=\pi(q_{0}+1)^{3},\qquad&\lambda_{0}:=\left(1+\frac{1}{q_{0}}\right)^{3}.\end{split}

Therefore, for all qq under consideration, α(W)α(W0,λ0)\alpha(W)\leq\alpha(W_{0},\lambda_{0}) and β(W)β(W0)\beta(W)\leq\beta(W_{0}). It follows that for qq0q\geq q_{0},

(57) 1Nmax1LP|SN(L)|α01/2π1/6q3/2+B0qt1/12.\frac{1}{N}\max_{1\leq L\leq P}|S_{N}(L)|\leq\frac{\alpha_{0}^{1/2}}{\pi^{1/6}q^{3/2}}+\frac{B_{0}}{qt^{1/12}}.

where α0:=α(W0,λ0)\alpha_{0}:=\alpha(W_{0},\lambda_{0}) and B0:=B(W0,λ0)B_{0}:=B(W_{0},\lambda_{0}).

Put together, we have

(58) |n=q0Pn11n1+it|α01/2π1/6q=q0Q1q3/2+B0t1/12q=q0Q1q.\begin{split}\left|\sum_{n=q_{0}P}^{n_{1}}\frac{1}{n^{1+it}}\right|\leq&\frac{\alpha_{0}^{1/2}}{\pi^{1/6}}\sum_{q=q_{0}}^{Q}\frac{1}{q^{3/2}}+\frac{B_{0}}{t^{1/12}}\sum_{q=q_{0}}^{Q}\frac{1}{q}.\end{split}

Moreover, from the argument in [HPY24, (3.1)], and using the bound Qt1/6/(2π)1/2Q\leq t^{1/6}/(2\pi)^{1/2} once again,

(59) q=q0Q1q3/2q01/2Q+1/2dxx3/2<2q01/2,q=q0Q1qq01/2Q+1/2dxx<16logt+logτ1log(q01/2),\begin{split}&\sum_{q=q_{0}}^{Q}\frac{1}{q^{3/2}}\leq\int_{q_{0}-1/2}^{Q+1/2}\frac{\text{d}x}{x^{3/2}}<\frac{2}{\sqrt{q_{0}-1/2}},\\ \sum_{q=q_{0}}^{Q}\frac{1}{q}&\leq\int_{q_{0}-1/2}^{Q+1/2}\frac{\text{d}x}{x}<\frac{1}{6}\log t+\log\tau_{1}-\log\left(q_{0}-1/2\right),\end{split}

where

τ1:=1(2π)1/2+12t11/6.\tau_{1}:=\frac{1}{(2\pi)^{1/2}}+\frac{1}{2t_{1}^{1/6}}.

Therefore, observing that 16logtt1/12\frac{1}{6}\frac{\log t}{t^{1/12}} is decreasing for tt1exp(12)t\geq t_{1}\geq\exp(12), and combining this with the initial sum bound (54), we obtain

|n=1n11n1+it|\displaystyle\left|\sum_{n=1}^{n_{1}}\frac{1}{n^{1+it}}\right| <log(q0P1)+γ+12(q0P1)+α01/2π1/62q01/2\displaystyle<\log\left(q_{0}P-1\right)+\gamma+\frac{1}{2(q_{0}P-1)}+\frac{\alpha_{0}^{1/2}}{\pi^{1/6}}\frac{2}{\sqrt{q_{0}-1/2}}
+B0t11/12(16logt1+logτ1log(q01/2)).\displaystyle\quad+\frac{B_{0}}{t_{1}^{1/12}}\left(\frac{1}{6}\log t_{1}+\log\tau_{1}-\log\left(q_{0}-1/2\right)\right).

We now estimate

log(q0P1)log(q0t1/3+q01)13logt+log(q0+(q01)t11/3),\log\left(q_{0}P-1\right)\leq\log\left(q_{0}t^{1/3}+q_{0}-1\right)\leq\frac{1}{3}\log t+\log\left(q_{0}+(q_{0}-1)t_{1}^{-1/3}\right),

which follows using P<t1/3+1P<t^{1/3}+1. Furthermore,

12(q0P1)12(q0t11/31).\frac{1}{2(q_{0}P-1)}\leq\frac{1}{2\big{(}q_{0}t_{1}^{1/3}-1\big{)}}.

This yields

(60) |ζ(1+it)|13logt+B1(q0)+g(t1)2π+(t1),|\zeta(1+it)|\leq\frac{1}{3}\log t+B_{1}(q_{0})+\frac{g(t_{1})}{\sqrt{2\pi}}+\mathcal{R}(t_{1}),

where

B1(q0):=\displaystyle B_{1}(q_{0}):= γ+log(q0+(q01)t11/3)+12(q0t11/31)\displaystyle\,\gamma+\log\left(q_{0}+(q_{0}-1)t_{1}^{-1/3}\right)+\frac{1}{2\big{(}q_{0}t_{1}^{1/3}-1\big{)}}
+α01/2π1/62q01/2+B0t11/12(16logt1+logτ1log(q01/2)).\displaystyle+\frac{\alpha_{0}^{1/2}}{\pi^{1/6}}\frac{2}{\sqrt{q_{0}-1/2}}+\frac{B_{0}}{t_{1}^{1/12}}\left(\frac{1}{6}\log t_{1}+\log\tau_{1}-\log\left(q_{0}-1/2\right)\right).

Choosing t1=exp(16)t_{1}=\exp(16), η=1.267\eta=1.267 and q0=5q_{0}=5, we obtain

(61) |ζ(1+it)|13logt+4.66,texp(16).|\zeta(1+it)|\leq\frac{1}{3}\log t+4.66,\qquad t\geq\exp(16).

This implies Theorem 1 in the range exp(16)texp(88)\exp(16)\leq t\leq\exp(88).

Consider now the second part of Proposition 3. In this critical region, we make use of a few additional tools to sharpen our estimates. In particular, we employ a different method of bounding the second sum appearing in Theorem 9, leading to an estimate of size O(t1/12)O(t^{-1/12}) instead of O(1)O(1). Also, we employ the kk-th derivative tests used in the proof of Theorem 1, however this time we introduce additional scaling parameters to further optimise the switching points between derivative tests of different order.

To this end, let t2t_{2} be a real number to be chosen later, subject to t2exp(182/3)t_{2}\geq\exp(182/3). Assume tt2t\geq t_{2}. Recall from (58) and (59) we showed for q01q_{0}\geq 1, P=t1/3P=\lceil t^{1/3}\rceil and tt2t\geq t_{2} that the tail sum satisfies

(62) |n=q0Pn11n1+it|α01/2π1/62q01/2+B0t21/12(16logt2+logτ1log(q01/2)).\left|\sum_{n=q_{0}P}^{n_{1}}\frac{1}{n^{1+it}}\right|\leq\frac{\alpha_{0}^{1/2}}{\pi^{1/6}}\frac{2}{\sqrt{q_{0}-1/2}}+\frac{B_{0}}{t_{2}^{1/12}}\left(\frac{1}{6}\log t_{2}+\log\tau_{1}-\log\left(q_{0}-1/2\right)\right).

As for the initial sum over 1n<q0P1\leq n<q_{0}P, instead of bounding it using the triangle inequality throughout as in (54), we will refine our estimate by using kk-th order van der Corput lemmas with k=4k=4 and 55 in the subintervals (Z5,Z4](Z_{5},Z_{4}] and (Z6,Z5](Z_{6},Z_{5}] respectively, where

Z4:=q0P1,Z5:=μ5tθ5/2,Z6:=μ6tθ6/2,\displaystyle Z_{4}:=q_{0}P-1,\qquad Z_{5}:=\mu_{5}t^{\theta_{5}/2},\qquad Z_{6}:=\mu_{6}t^{\theta_{6}/2},

with θ5,θ6\theta_{5},\theta_{6} the same as in the proof of Theorem 1 and μ5,μ6\mu_{5},\mu_{6} real positive numbers to be chosen later, subject to

(63) 1<μ6t2θ6/2μ5t2θ5/2q0t21/31.1<\mu_{6}t_{2}^{\theta_{6}/2}\leq\mu_{5}t_{2}^{\theta_{5}/2}\leq q_{0}\lceil t_{2}^{1/3}\rceil-1.

We remark that Z5,Z6Z_{5},Z_{6} will serve a similar purpose as YkY_{k} in the proof of Theorem 1, whereas the new parameters μ5,μ6\mu_{5},\mu_{6} will allow us to fine-tune our estimates in this critical region.

Let h>1h^{\prime}>1 be a real number to be chosen later, and let Xm=(h)mZ4X_{m}^{\prime}=(h^{\prime})^{-m}Z_{4} for integer mm. Define

M:=log(Z4/Z6)logh.M^{\prime}:=\left\lceil\frac{\log(Z_{4}/Z_{6})}{\log h^{\prime}}\right\rceil.

The definition of MM^{\prime} guarantees XMZ6<XM1X^{\prime}_{M^{\prime}}\leq Z_{6}<X^{\prime}_{M^{\prime}-1}. Next, define the subsum

Sm:=Xm+1<nXm11n1+it,S_{m}^{\prime}:=\sum_{X_{m}^{\prime}+1<n\leq X_{m-1}^{\prime}}\frac{1}{n^{1+it}},

so that, similarly to (12),

|XM1<nZ41n1+it|1mM1(1Xm+|Sm|).\left|\sum_{X^{\prime}_{M^{\prime}-1}<n\leq Z_{4}}\frac{1}{n^{1+it}}\right|\leq\sum_{1\leq m\leq M^{\prime}-1}\left(\frac{1}{X_{m}^{\prime}}+|S_{m}^{\prime}|\right).

By a similar calculation as in (13), for each 1mM11\leq m\leq M^{\prime}-1,

Xm1Xm+1h.\frac{\lfloor X_{m-1}^{\prime}\rfloor}{\lfloor X_{m}^{\prime}\rfloor+1}\leq h^{\prime}.

Let k=4k=4 or 55. Following the argument of Theorem 1, if Zk+1<XmZkZ_{k+1}<X_{m}^{\prime}\leq Z_{k}, then

|Sm|\displaystyle\left|S_{m}^{\prime}\right| Ck(η,h)(Xm+1)k/(2K2)t1/(2K2)\displaystyle\leq C_{k}(\eta,h^{\prime})(\lfloor X_{m}^{\prime}\rfloor+1)^{-k/(2K-2)}t^{1/(2K-2)}
+Dk(η,h)(Xm+1)k/(2K2)2/Kt1/(2K2)\displaystyle\qquad\qquad+D_{k}(\eta,h^{\prime})(\lfloor X_{m}^{\prime}\rfloor+1)^{k/(2K-2)-2/K}t^{-1/(2K-2)}
Ck(η,h)Zk+1k/(2K2)t1/(2K2)+Dk(η,h)(Zk+1)k/(2K2)2/Kt1/(2K2).\displaystyle\leq C_{k}(\eta,h^{\prime})Z_{k+1}^{-k/(2K-2)}t^{1/(2K-2)}+D_{k}(\eta,h^{\prime})(Z_{k}+1)^{k/(2K-2)-2/K}t^{-1/(2K-2)}.

We write μ4:=q0+(q01)t21/3\mu_{4}:=q_{0}+(q_{0}-1)t_{2}^{-1/3} so that Z4μ4t1/3Z_{4}\leq\mu_{4}t^{1/3}. Since k/(2K2)2/K0k/(2K-2)-2/K\geq 0, for tt2t\geq t_{2},

|Sm|Ak(t):={μ52/7C4t3/182+(μ4+t21/3)1/28D4t5/84,k=4,μ61/6C5t7/990+(μ5+t24/13)1/24D5t4/195,k=5.|S_{m}^{\prime}|\leq A_{k}^{\prime}(t):=\begin{cases}\displaystyle\mu_{5}^{-2/7}C_{4}t^{-3/182}+(\mu_{4}+t_{2}^{-1/3})^{1/28}D_{4}t^{-5/84},&k=4,\\ \displaystyle\mu_{6}^{-1/6}C_{5}t^{-7/990}+(\mu_{5}+t_{2}^{-4/13})^{1/24}D_{5}t^{-4/195},&k=5.\end{cases}

Let NkN_{k}^{\prime} be the number of XmX_{m}^{\prime} satisfying Zk+1<XmZkZ_{k+1}<X_{m}^{\prime}\leq Z_{k}. Then

Nk\displaystyle N_{k}^{\prime} logZklogZk+1logh+1,thereforeNkBklogtlogh,\displaystyle\leq\frac{\log Z_{k}-\log Z_{k+1}}{\log h^{\prime}}+1,\qquad\text{therefore}\qquad N_{k}^{\prime}\leq B^{\prime}_{k}\frac{\log t}{\log h^{\prime}},

where

Bk:={139+max{0,logμ4logμ5+logh}logt2,k=4,28429+max{0,logμ5logμ6+logh}logt2,k=5.B^{\prime}_{k}:=\begin{cases}\displaystyle\frac{1}{39}+\frac{\max\{0,\log\mu_{4}-\log\mu_{5}+\log h^{\prime}\}}{\log t_{2}},&k=4,\\ \,&\\ \displaystyle\frac{28}{429}+\frac{\max\{0,\log\mu_{5}-\log\mu_{6}+\log h^{\prime}\}}{\log t_{2}},&k=5.\end{cases}

Put together, we have

Zk+1<XmZk|Sm|Ak(t)Bkloghlogt.\sum_{Z_{k+1}<X^{\prime}_{m}\leq Z_{k}}|S_{m}^{\prime}|\leq\frac{A^{\prime}_{k}(t)B^{\prime}_{k}}{\log h^{\prime}}\log t.

For tt2t\geq t_{2}, the function A4(t)logtA_{4}^{\prime}(t)\log t is decreasing, so A4(t)logtA4(t2)logt2A_{4}^{\prime}(t)\log t\leq A_{4}^{\prime}(t_{2})\log t_{2}. For k=5k=5, we use

t7/990logt9907eandt4/195logtt24/195logt2,(tt2).t^{-7/990}\log t\leq\frac{990}{7e}\qquad\text{and}\qquad t^{-4/195}\log t\leq t_{2}^{-4/195}\log t_{2},\qquad(t\geq t_{2}).

Therefore,

(64) 1mM1|Sm|1logh(A4(t2)B4logt2+9907eB5C5μ61/6+(μ5+t24/13)1/24B5D5t24/195logt2).\begin{split}\sum_{1\leq m\leq M^{\prime}-1}|S^{\prime}_{m}|&\leq\frac{1}{\log h^{\prime}}\bigg{(}A^{\prime}_{4}(t_{2})B^{\prime}_{4}\log t_{2}+\frac{990}{7e}\frac{B^{\prime}_{5}C_{5}}{\mu_{6}^{1/6}}\\ &\qquad\qquad+\left(\mu_{5}+t_{2}^{-4/13}\right)^{1/24}B^{\prime}_{5}D_{5}t_{2}^{-4/195}\log t_{2}\bigg{)}.\end{split}

Next, similarly to (39), we obtain

(65) 1mM11Xm<hh11Z6h(h1)μ6t2θ6/2,(tt2).\sum_{1\leq m\leq M^{\prime}-1}\frac{1}{X_{m}^{\prime}}<\frac{h^{\prime}}{h^{\prime}-1}\frac{1}{Z_{6}}\leq\frac{h^{\prime}}{(h^{\prime}-1)\mu_{6}t_{2}^{\theta_{6}/2}},\qquad(t\geq t_{2}).

In the remaining range 1nXM11\leq n\leq X^{\prime}_{M^{\prime}-1}, we use that hZ6XM1h^{\prime}Z_{6}\geq X^{\prime}_{M^{\prime}-1}. Combined with the triangle inequality and (2), this yields

(66) |1nXM11n1+it|1nhZ61nloghZ6+γ+12(hZ61)θ62logt+logμ6+logh+γ+12(hμ6t2θ6/21).\begin{split}\left|\sum_{1\leq n\leq\lfloor X_{M^{\prime}-1}\rfloor}\frac{1}{n^{1+it}}\right|&\leq\sum_{1\leq n\leq\lfloor h^{\prime}Z_{6}\rfloor}\frac{1}{n}\leq\log h^{\prime}Z_{6}+\gamma+\frac{1}{2(h^{\prime}Z_{6}-1)}\\ &\leq\frac{\theta_{6}}{2}\log t+\log\mu_{6}+\log h^{\prime}+\gamma+\frac{1}{2\big{(}h^{\prime}\mu_{6}t_{2}^{\theta_{6}/2}-1\big{)}}.\end{split}

Next, consider the second sum 1nn1nit\sum_{1\leq n\leq n_{1}}n^{it} in Theorem 9, which was just bounded by n1n_{1} in our previous calculations. Let qq, q0q_{0}, PP, QQ and NN be as defined in the proof of Proposition 3, and suppose as before that tt2exp(182/3)t\geq t_{2}\geq\exp(182/3). Using Pt1/3+1P\leq t^{1/3}+1, the estimate (57) gives for q0qQq_{0}\leq q\leq Q,

max1LP|SN(L)|(t1/3+1)(α01/2π1/6q1/2+B0t1/12).\max_{1\leq L\leq P}|S_{N}(L)|\leq\left(t^{1/3}+1\right)\left(\frac{\alpha_{0}^{1/2}}{\pi^{1/6}q^{1/2}}+\frac{B_{0}}{t^{1/12}}\right).

We combine this with the inequality

q=q0Q1q1/2q01Qdxx1/2<2Q1/2,\sum_{q=q_{0}}^{Q}\frac{1}{q^{1/2}}\leq\int_{q_{0}-1}^{Q}\frac{\text{d}x}{x^{1/2}}<2Q^{1/2},

as well as bound the sum over 1n<q0P1\leq n<q_{0}P by the number of terms in the sum, which gives

|n=1n1nit|\displaystyle\left|\sum_{n=1}^{n_{1}}n^{it}\right| <q0P1+(t1/3+1)(2α01/2Q1/2π1/6+B0Qt1/12).\displaystyle<q_{0}P-1+\left(t^{1/3}+1\right)\left(\frac{2\alpha_{0}^{1/2}Q^{1/2}}{\pi^{1/6}}+\frac{B_{0}Q}{t^{1/12}}\right).

Since Qt1/6/2πQ\leq t^{1/6}/\sqrt{2\pi}, we obtain for tt2t\geq t_{2},

(67) |n=1n1nit|<Ct5/12,\left|\sum_{n=1}^{n_{1}}n^{it}\right|<C\,t^{5/12},

where

C(q0,η,t2)\displaystyle C(q_{0},\eta,t_{2}) :=(1+t21/3)(q0t21/12+23/4α01/2π5/12+B02π).\displaystyle:=\left(1+t_{2}^{-1/3}\right)\left(\frac{q_{0}}{t_{2}^{1/12}}+\frac{2^{3/4}\alpha_{0}^{1/2}}{\pi^{5/12}}+\frac{B_{0}}{\sqrt{2\pi}}\right).

Therefore, after multiplying by g(t)t1/2g(t)t^{-1/2}, we obtain a sharper bound on the contribution of the second sum in Theorem 9. Namely, the bound Cg(t2)t1/12Cg(t_{2})\,t^{-1/12}, which replaces the original estimate g(t2)/2πg(t_{2})/\sqrt{2\pi}.

Lastly, combining (62), (64), (65), (66) and (67), and noting that θ6=16/33\theta_{6}=16/33, we have for tt2t\geq t_{2}

|ζ(1+it)|833logt+B3,|\zeta(1+it)|\leq\frac{8}{33}\log t+B_{3},

where

B3:=\displaystyle B_{3}:= 1logh(A4(t2)B4logt2+9907eB5C5μ61/6+(μ5+t24/13)1/24B5D5t24/195logt2)\displaystyle\,\frac{1}{\log h^{\prime}}\left(A^{\prime}_{4}(t_{2})B^{\prime}_{4}\log t_{2}+\frac{990}{7e}\frac{B^{\prime}_{5}C_{5}}{\mu_{6}^{1/6}}+(\mu_{5}+t_{2}^{-4/13})^{1/24}B^{\prime}_{5}D_{5}t_{2}^{-4/195}\log t_{2}\right)
+h(h1)μ6t2θ6/2+α0(W0,λ0)1/2π1/62q01/2\displaystyle\qquad+\frac{h^{\prime}}{(h^{\prime}-1)\mu_{6}t_{2}^{\theta_{6}/2}}+\frac{\alpha_{0}(W_{0},\lambda_{0})^{1/2}}{\pi^{1/6}}\frac{2}{\sqrt{q_{0}-1/2}}
+B0(W0,λ0)t21/12(16logt2+logτ1log(q01/2))\displaystyle\qquad+\frac{B_{0}(W_{0},\lambda_{0})}{t_{2}^{1/12}}\left(\frac{1}{6}\log t_{2}+\log\tau_{1}-\log\left(q_{0}-1/2\right)\right)
+logμ6+logh+γ+12(hμ6t2θ6/21)+C(q0,η,t2)g(t2)t21/12+(t2).\displaystyle\qquad+\log\mu_{6}+\log h^{\prime}+\gamma+\frac{1}{2\big{(}h^{\prime}\mu_{6}t_{2}^{\theta_{6}/2}-1\big{)}}+C(q_{0},\eta,t_{2})\frac{g(t_{2})}{t_{2}^{1/12}}+\mathcal{R}(t_{2}).

We choose

η\displaystyle\eta =1.3348,\displaystyle=1.3348,\qquad\qquad h=1.056,\displaystyle h^{\prime}=1.056,\qquad\qquad q0=46,\displaystyle q_{0}=46,
μ5\displaystyle\mu_{5} =48.575,\displaystyle=48.575,\qquad\qquad μ6=51.296,\displaystyle\mu_{6}=51.296,\qquad\qquad t2=exp(90).\displaystyle t_{2}=\exp(90).

Verifying that (63) holds, we calculate B312.45B_{3}\leq 12.45, i.e. that

|ζ(1+it)|833logt+12.45,texp(88).|\zeta(1+it)|\leq\frac{8}{33}\log t+12.45,\qquad t\geq\exp(88).

This implies Theorem 1 for exp(88)texp(662)\exp(88)\leq t\leq\exp(662).

6. Proof of Theorem 2

We begin by proving some bounds on ζ(s)\zeta(s). To help make our notation more intuitive, we will allow reuse of variable names from previous sections. The reader should keep in mind these variables mean different things in the confines of this section. Since this section is independent from Section 3, Section 4 and Section 5, this will hopefully cause no confusion.

Lemma 4.

Let C>0C>0 and t0eet_{0}\geq e^{e} be constants. Suppose tt0t\geq t_{0}. Define

σt:=1C(loglogt)2logt.\sigma_{t}:=1-\frac{C(\log\log t)^{2}}{\log t}.

Additionally, suppose that CC satisfies

(68) C(loglogte)2logte12,wherete:=ee2.\frac{C(\log\log t_{e})^{2}}{\log t_{e}}\leq\frac{1}{2},\qquad\text{where}\qquad t_{e}:=e^{e^{2}}.

For each tt0t\geq t_{0}, if σtσ2\sigma_{t}\leq\sigma\leq 2, then

(69) |ζ(σ+it)|AlogBtwhereA=76.2,B=23+71.2C3/2e2.|\zeta(\sigma+it)|\leq A\log^{B}t\qquad\text{where}\qquad A=76.2,\qquad B=\frac{2}{3}+\frac{71.2\,C^{3/2}}{e^{2}}.
Proof.

From Ford [For02], if 1/2σ11/2\leq\sigma\leq 1, then

(70) |ζ(σ+it)|76.2t4.45(1σ)3/2log2/3t.|\zeta(\sigma+it)|\leq 76.2\,t^{4.45(1-\sigma)^{3/2}}\log^{2/3}t.

As σt1/2\sigma_{t}\geq 1/2, a consequence of the assumption (68) and considering that (loglogt)2/logt(\log\log t)^{2}/\log t reaches its maximum at t=tet=t_{e}, we may use (70) for any σ[σt,1]\sigma\in[\sigma_{t},1].

In view of this, since (1σ)3/2(1-\sigma)^{3/2} is monotonically decreasing with σ\sigma, if σtσ1\sigma_{t}\leq\sigma\leq 1, then (70) gives

(71) |ζ(σ+it)|76.2exp(4.45(1σt)3/2logt+23loglogt).\displaystyle|\zeta(\sigma+it)|\leq 76.2\exp\left(4.45(1-\sigma_{t})^{3/2}\log t+\frac{2}{3}\log\log t\right).

Or, written differently,

(72) |ζ(σ+it)|Aexp(B~loglogt),|\zeta(\sigma+it)|\leq A\exp\left(\tilde{B}\log\log t\right),

where AA is defined as in (69) and

B~:=23+4.45C3/2(loglogt)2logt.\tilde{B}:=\frac{2}{3}+4.45\,C^{3/2}\,\frac{(\log\log t)^{2}}{\sqrt{\log t}}.

As (loglogt)2/logt(\log\log t)^{2}/\sqrt{\log t} reaches its maximum at t=ee4t=e^{e^{4}}, attaining the value 16/e216/e^{2}, we obtain after a simple computation that B~B\tilde{B}\leq B, with BB defined as in (69). The desired result therefore follows for σtσ1\sigma_{t}\leq\sigma\leq 1 on replacing B~\tilde{B} with BB in (72).

To show that the result holds for 1σ21\leq\sigma\leq 2 as well, we use the Phragmén–Lindelöf Principle on the holomorphic function

f(s)=(s1)ζ(s).f(s)=(s-1)\zeta(s).

To this end, on the 11-line we have

|f(1+it)|=|t||ζ(1+it)|62.6|2+it|log2/3|2+it|.|f(1+it)|=|t|\,|\zeta(1+it)|\leq 62.6\,|2+it|\log^{2/3}|2+it|.

This inequality is verified numerically for |t|<3|t|<3 and is a consequence of the bound |ζ(1+it)|62.6log2/3t|\zeta(1+it)|\leq 62.6\log^{2/3}t for |t|3|t|\geq 3, given in [Tru14]. Also, on the 22-line, plainly,

|f(2+it)||1+it|ζ(2)<62.6|3+it|log2/3|3+it|,|f(2+it)|\leq|1+it|\,\zeta(2)<62.6|3+it|\log^{2/3}|3+it|,

for all real tt. By [Tru14a, Lemma 3], we thus obtain777We apply [Tru14a, Lemma 3] with the following parameter values (in the notation in the cited paper): a=1a=1, b=2b=2, Q=1Q=1, α1=β1=1\alpha_{1}=\beta_{1}=1, α2=β2=2/3\alpha_{2}=\beta_{2}=2/3 and A=B=62.6A=B=62.6. Note that the AA and BB in [Tru14a, Lemma 3] are different from the AA and BB in the statement of our lemma.

(73) |f(s)|62.6|1+s|log2/3|1+s|,(1σ2).|f(s)|\leq 62.6|1+s|\log^{2/3}|1+s|,\qquad(1\leq\sigma\leq 2).

Moreover, for 1σ21\leq\sigma\leq 2 we have the easy estimates,

|1+s||s1|9+t2t1+9t02,log|1+s|log9+t2logt+log1+9t02.\begin{split}&\frac{|1+s|}{|s-1|}\leq\frac{\sqrt{9+t^{2}}}{t}\leq\sqrt{1+9t_{0}^{-2}},\\ &\log|1+s|\leq\log\sqrt{9+t^{2}}\leq\log t+\log\sqrt{1+9t_{0}^{-2}}.\end{split}

So that, for 1σ21\leq\sigma\leq 2 and tt0t\geq t_{0}, the inequality (73) implies that

|ζ(σ+it)|\displaystyle|\zeta(\sigma+it)| 62.61+9t02(1+log1+9t02logt0)log2/3t<AlogBt,\displaystyle\leq 62.6\sqrt{1+9t_{0}^{-2}}\left(1+\frac{\log\sqrt{1+9t_{0}^{-2}}}{\log t_{0}}\right)\log^{2/3}t<A\log^{B}t,

where AA and BB are defined as in (69), as desired. ∎

Remark.

Following Titchmarsh’s argument [Tit86, Theorem 5.17], we could make use of [Yan24, Theorem 1.1]. For simplicity, however, we used the Richert-type bound (70) due to Ford [For02] instead.

Lemma 5.

Let t0eet_{0}\geq e^{e} and d1>0d_{1}>0 be constants. Suppose tt0t\geq t_{0}. We have

|ζ(1+d1loglogtlogt+it)|C(d1,t0)logtloglogt,\left|\zeta\left(1+d_{1}\frac{\log\log t}{\log t}+it\right)\right|\leq C^{\prime}(d_{1},t_{0})\frac{\log t}{\log\log t},

where γ\gamma is Euler’s constant, and

(74) C(d1,t0)=1d1exp(γd1loglogt0logt0).C^{\prime}(d_{1},t_{0})=\frac{1}{d_{1}}\exp\left(\gamma d_{1}\frac{\log\log t_{0}}{\log t_{0}}\right).
Proof.

We use the following uniform bound due to Ramaré [Ram16, Lemma 5.4]:

|ζ(σ+it)|ζ(σ)eγ(σ1)σ1,(σ>1).|\zeta(\sigma+it)|\leq\zeta(\sigma)\leq\frac{e^{\gamma(\sigma-1)}}{\sigma-1},\qquad(\sigma>1).

Substituting σ=1+d1loglogt/logt\sigma=1+d_{1}\log\log t/\log t therefore gives for tt0t\geq t_{0},

ζ(σ)eγd1loglogt/logtd1loglogt/logt1d1exp(γd1loglogt0logt0)logtloglogt,\zeta(\sigma)\leq\frac{e^{\gamma d_{1}\log\log t/\log t}}{d_{1}\log\log t/\log t}\leq\frac{1}{d_{1}}\exp\left(\frac{\gamma d_{1}\log\log t_{0}}{\log t_{0}}\right)\frac{\log t}{\log\log t},

as claimed. In the last inequality, we used that loglogt/logt\log\log t/\log t reaches its maximum at t=eet=e^{e} and is monotonically decreasing after that. ∎

6.1. Bounding |ζ(1+it)/ζ(1+it)||\zeta^{\prime}(1+it)/\zeta(1+it)|

We follow the argument of [Tru15], which gives an explicit version of results in [Tit86, §3]. The method relies on Theorem 14.

We will therefore construct concentric disks, centred just to the right of the line s=1+its=1+it, and extend their radius slightly to the left and into the critical strip. Ultimately, we want to apply Lemmas 15 and 16 with f(s)=ζ(s)f(s)=\zeta(s), which will give us the desired results.

The process is as follows. Let t0eet_{0}\geq e^{e} and suppose tt0t^{\prime}\geq t_{0}. Let dd be a real positive constant, which will be chosen later. Denote the center of the concentric disks to be constructed by s=σ+its^{\prime}=\sigma^{\prime}+it^{\prime}, where

(75) σ=1+δt,withδt:=dloglogtlogt.\sigma^{\prime}=1+\delta_{t^{\prime}},\qquad\text{with}\qquad\delta_{t^{\prime}}:=\frac{d\log\log t^{\prime}}{\log t^{\prime}}.

Notice that δt\delta_{t^{\prime}} is decreasing with tt^{\prime} and is at most d/ed/e. Let r>0r>0 denote the radius of our largest concentric disk. In the sequel, rr will be determined as function of ss^{\prime}.

In addition, suppose σ+r1+ϵ\sigma^{\prime}+r\leq 1+\epsilon where ϵ1\epsilon\leq 1 is another positive constant to be chosen later. This implies, for instance, that rϵ1r\leq\epsilon\leq 1. In order to enforce this inequality on σ\sigma^{\prime} and rr, it is enough to demand that

(76) de+rϵ.\frac{d}{e}+r\leq\epsilon.

Let s=σ+its=\sigma+it be a complex number. Aiming to apply Lemma 15 in the disk |ss|r|s-s^{\prime}|\leq r, we first seek a valid A1A_{1} in that disk. Similarly, for Lemma 16, we seek a valid A2A_{2} at the disk center s=ss=s^{\prime}. In applying Lemma 16, we moreover need to ensure that the non-vanishing condition on f(s)=ζ(s)f(s)=\zeta(s) is fulfilled, which we do by way of a zero-free region of zeta.

Let us first determine a valid A1A_{1}, with the aid of Lemma 4. To this end, let C>0C>0 be a parameter restricted by the inequality in (68), so that (in the notation of Lemma 4) σt1/2\sigma_{t}\geq 1/2 for tet\geq e. Note that σt\sigma_{t} is monotonically decreasing for ettee\leq t\leq t_{e} and monotonically increasing thereafter. So, the maximum of σt\sigma_{t} over the interval trtt+rt^{\prime}-r\leq t\leq t^{\prime}+r occurs at one of the end-points. For example, if trtet^{\prime}-r\geq t_{e}, then the maximum occurs at t=t+rt=t^{\prime}+r. And if t+rtet^{\prime}+r\leq t_{e}, then the maximum occurs at t=trt=t^{\prime}-r. With this in mind, we calculate

(77) max|tt|rσt=max{σtr,σt+r}<1C0(loglogt)2logt,\max_{|t-t^{\prime}|\leq r}\sigma_{t}=\max\{\sigma_{t^{\prime}-r},\sigma_{t^{\prime}+r}\}<1-\frac{C_{0}\,(\log\log t^{\prime})^{2}}{\log t^{\prime}},

where, since rϵr\leq\epsilon,

C0:=Cmin{(loglog(tϵ)loglogt)2,logtlog(t+ϵ)}.C_{0}:=C\,\min\left\{\left(\frac{\log\log(t^{\prime}-\epsilon)}{\log\log t^{\prime}}\right)^{2},\frac{\log t^{\prime}}{\log(t^{\prime}+\epsilon)}\right\}.

Clearly, C0=C0(t)C_{0}=C_{0}(t^{\prime}) increases monotonically with tt^{\prime} (towards CC). We therefore set

(78) C1:=C0(t0)andσ1,t:=1C1(loglogt)2logt.C_{1}:=C_{0}(t_{0})\qquad\text{and}\qquad\sigma_{1,t^{\prime}}:=1-\frac{C_{1}\,(\log\log t^{\prime})^{2}}{\log t^{\prime}}.

So that, by (77), no matter tt^{\prime}, rr and ϵ\epsilon, so long as they are subject to our conditions,

σ1,tmax|tt|rσt.\sigma_{1,t^{\prime}}\geq\max_{|t-t^{\prime}|\leq r}\sigma_{t}.

Consequently, Lemma 4 gives that for each t[tr,t+r]t\in[t^{\prime}-r,t^{\prime}+r] and any σ[σ1,t,2]\sigma\in[\sigma_{1,t^{\prime}},2], and on recalling rϵr\leq\epsilon,

(79) |ζ(s)|AlogB(t+r)A3logBt,A3=A(1+log(1+ϵ/t)logt)B.|\zeta(s)|\leq A\log^{B}(t^{\prime}+r)\leq A_{3}\log^{B}t^{\prime},\qquad A_{3}=A\left(1+\frac{\log(1+\epsilon/t^{\prime})}{\log t^{\prime}}\right)^{B}.

We would like the inequality (79) to hold throughout the disk |ss|r|s-s^{\prime}|\leq r. This will be certainly the case if this disk lies entirely in the rectangle specified by σ[σ1,t,2]\sigma\in[\sigma_{1,t^{\prime}},2] and t[tr,t+r]t\in[t^{\prime}-r,t^{\prime}+r]. This follows, in turn, on requiring σrσ1,t\sigma^{\prime}-r\geq\sigma_{1,t^{\prime}} or, equivalently, that

rC1(loglogt)2logt+δt.r\leq\frac{C_{1}(\log\log t^{\prime})^{2}}{\log t^{\prime}}+\delta_{t^{\prime}}.

Therefore, subject to the constraint (76), the following choice of rr works:

(80) r=rt:=(C1+dloglogt)(loglogt)2logt.r=r_{t^{\prime}}:=\left(C_{1}+\frac{d}{\log\log t^{\prime}}\right)\frac{(\log\log t^{\prime})^{2}}{\log t^{\prime}}.

To ensure the constraint (76) is met for all tt0t^{\prime}\geq t_{0}, it suffices that

(81) de+(C1+d)4e2ϵ1.\frac{d}{e}+\left(C_{1}+d\right)\frac{4}{e^{2}}\leq\epsilon\leq 1.

Here, we used that (loglogt)2/logt(\log\log t^{\prime})^{2}/\log t^{\prime} reaches its maximum of 4/e24/e^{2} at t=tet^{\prime}=t_{e}.

Overall, combining (79) and (80) with the trivial bound |ζ(s)|ζ(2σ)/ζ(σ)|\zeta(s^{\prime})|\geq\zeta(2\sigma^{\prime})/\zeta(\sigma^{\prime}), easily seen on considering the Euler product of zeta, we obtain that throughout the disk |ss|r|s-s^{\prime}|\leq r,

|ζ(s)ζ(s)|\displaystyle\left|\frac{\zeta(s)}{\zeta(s^{\prime})}\right|\leq A3ζ(σ)ζ(2σ)logBt=A3X(σ)σ1logBt,\displaystyle\,A_{3}\,\frac{\zeta(\sigma^{\prime})}{\zeta(2\sigma^{\prime})}\log^{B}t^{\prime}=A_{3}\,\frac{X(\sigma^{\prime})}{\sigma^{\prime}-1}\log^{B}t^{\prime},

where

X(σ)=ζ(σ)(σ1)ζ(2σ).X(\sigma^{\prime})=\frac{\zeta(\sigma^{\prime})(\sigma^{\prime}-1)}{\zeta(2\sigma^{\prime})}.

Since σ=σ(t)\sigma^{\prime}=\sigma^{\prime}(t^{\prime}) decreases to 11 with tt0t^{\prime}\geq t_{0}, X(σ)X(\sigma^{\prime}) decreases to 1/ζ(2)1/\zeta(2) with tt^{\prime}. Thus, recalling from (75) that σ=1+δt\sigma^{\prime}=1+\delta_{t^{\prime}} and observing A3A_{3} is also decreasing in tt0t^{\prime}\geq t_{0}, we obtain that throughout the disk |ss|r|s-s^{\prime}|\leq r,

(82) |ζ(s)ζ(s)|AmaxXmaxdlogB+1tloglogt,\left|\frac{\zeta(s)}{\zeta(s^{\prime})}\right|\leq\frac{A_{\max}\,X_{\max}}{d}\frac{\log^{B+1}t^{\prime}}{\log\log t^{\prime}},

where Amax=A3(t0)A_{\max}=A_{3}(t_{0}) and Xmax=X(σ(t0))X_{\max}=X(\sigma^{\prime}(t_{0})).

In summary, when applying Lemma 15 in the disk |ss|r|s-s^{\prime}|\leq r, with rr as in (80) and subject to (81), we may take

logA1=(B+1)loglogtlogloglogt+log(AmaxXmaxd).\log A_{1}=(B+1)\log\log t^{\prime}-\log\log\log t^{\prime}+\log\left(\frac{A_{\max}\,X_{\max}}{d}\right).

Next, in preparation for applying Lemma 16, we want to ensure that the intersection of the disk |ss|r|s-s^{\prime}|\leq r and the right half-plane σσαr\sigma\geq\sigma^{\prime}-\alpha r lies entirely in a zero-free region for zeta. By [Yan24, Corollary 1.2], we have the following zero-free region, valid for t3t\geq 3.

ζ(σ+it)0forσ>1c0loglogtlogt,wherec0=121.233.\zeta(\sigma+it)\neq 0\quad\text{for}\quad\sigma>1-\frac{c_{0}\log\log t}{\log t},\quad\text{where}\quad c_{0}=\frac{1}{21.233}.

Since 1c0loglogt/logt1-c_{0}\log\log t/\log t is monotonically increasing for tt0t\geq t_{0} and since tt0t^{\prime}\geq t_{0} and rϵr\leq\epsilon, it suffices for our purposes to require that

σαr1c0loglog(t+ϵ)log(t+ϵ).\sigma^{\prime}-\alpha r\geq 1-\frac{c_{0}\log\log(t^{\prime}+\epsilon)}{\log(t^{\prime}+\epsilon)}.

In other words, it suffices that

α1r(δt+c0loglog(t+ϵ)log(t+ϵ)).\alpha\leq\frac{1}{r}\left(\delta_{t^{\prime}}+\frac{c_{0}\log\log(t^{\prime}+\epsilon)}{\log(t^{\prime}+\epsilon)}\right).

Therefore, subject to α<1/2\alpha<1/2 as demanded by Lemma 16, the following choice of α\alpha works.

(83) α=1r(d+c1)loglogtlogt=d+c1d+C1loglogt,\alpha=\frac{1}{r}\cdot\frac{(d+c_{1})\log\log t^{\prime}}{\log t^{\prime}}=\frac{d+c_{1}}{d+C_{1}\log\log t^{\prime}},

where (using monotonicity to deduce the inequality below)

(84) c1:=c0logt0log(t0+ϵ),so thatc1c0.c_{1}:=\frac{c_{0}\log t_{0}}{\log(t_{0}+\epsilon)},\qquad\text{so that}\qquad c_{1}\leq c_{0}.

In particular, the constraint α<1/2\alpha<1/2 is fulfilled if

(85) d+c1d+C1loglogt0<12.\frac{d+c_{1}}{d+C_{1}\log\log t_{0}}<\frac{1}{2}.

Now, to determine A2A_{2}, we utilise [Del87], which asserts that if σ>1\sigma>1, then

|ζ(s)ζ(s)|ζ(σ)ζ(σ)<1σ1.\left|\frac{\zeta^{\prime}(s)}{\zeta(s)}\right|\leq-\frac{\zeta^{\prime}(\sigma)}{\zeta(\sigma)}<\frac{1}{\sigma-1}.

Taking s=ss=s^{\prime} in this inequality leads to

|ζ(s)ζ(s)|<1δt=logtdloglogt.\left|\frac{\zeta^{\prime}(s^{\prime})}{\zeta(s^{\prime})}\right|<\frac{1}{\delta_{t^{\prime}}}=\frac{\log t^{\prime}}{d\log\log t^{\prime}}.

So, we may take A2A_{2} in Lemma 16 to be

A2=logtdloglogt.A_{2}=\frac{\log t^{\prime}}{d\log\log t^{\prime}}.

Lastly, in applying Lemma 16, we need to be able to reach the 11-line from our position at ss^{\prime}. So, we set

β=δtαr=dc1+d.\beta=\frac{\delta_{t^{\prime}}}{\alpha r}=\frac{d}{c_{1}+d}.

Clearly, β<1\beta<1, as required by Lemma 16. The conclusion of the lemma therefore holds throughout the disk |ss|αβr=δt|s-s^{\prime}|\leq\alpha\beta r=\delta_{t^{\prime}}. This includes, in particular, the horizontal line segment s=σ+its=\sigma+it^{\prime}, with 1σ1+2δt1\leq\sigma\leq 1+2\delta_{t^{\prime}}.

Putting it all together, and making the substitution ttt^{\prime}\to t, Lemma 16 hence furnishes the bound

(86) |ζ(s)ζ(s)|8β(1β)(12α)2logA1r+1+β1βA2,\left|\frac{\zeta^{\prime}(s)}{\zeta(s)}\right|\leq\frac{8\beta}{(1-\beta)(1-2\alpha)^{2}}\,\frac{\log A_{1}}{r}+\frac{1+\beta}{1-\beta}\,A_{2},

valid in the region

1σ1+2dloglogtlogt,tt0.1\leq\sigma\leq 1+\frac{2d\log\log t}{\log t},\qquad t\geq t_{0}.

We now input the values we determined for A1A_{1}, A2A_{2}, rr, α\alpha and β\beta into (86). We also note

1+β1β=1+2dc1,so that1+β1βA2=(1d+2c1)logtloglogt,\frac{1+\beta}{1-\beta}=1+\frac{2d}{c_{1}},\qquad\text{so that}\qquad\frac{1+\beta}{1-\beta}\,A_{2}=\left(\frac{1}{d}+\frac{2}{c_{1}}\right)\frac{\log t}{\log\log t},

as well as

8β1β=8dc1,1(12α)2=(C1loglogt+dC1loglogtd2c1)2.\frac{8\beta}{1-\beta}=\frac{8d}{c_{1}},\qquad\frac{1}{(1-2\alpha)^{2}}=\left(\frac{C_{1}\log\log t+d}{C_{1}\log\log t-d-2c_{1}}\right)^{2}.

Hence, on observing that 1/(12α)21/(1-2\alpha)^{2} is decreasing in tt, we obtain the following lemma.

Lemma 6.

Let t0eet_{0}\geq e^{e} and suppose that tt0t\geq t_{0}. Let CC, dd and ϵ\epsilon be any positive constants satisfying the constraints (68), (81) and (85). If

1σ1+2dloglogtlogt,1\leq\sigma\leq 1+\frac{2d\log\log t}{\log t},

then

|ζ(σ+it)ζ(σ+it)|Q1(C,d,ϵ,t0)logtloglogt,\left|\frac{\zeta^{\prime}(\sigma+it)}{\zeta(\sigma+it)}\right|\leq Q_{1}(C,d,\epsilon,t_{0})\frac{\log t}{\log\log t},

where

Q1(C,d,ϵ,t0):=λ1(B+1+1loglogt0log(AmaxXmaxd))+λ2.\displaystyle Q_{1}(C,d,\epsilon,t_{0}):=\lambda_{1}\left(B+1+\frac{1}{\log\log t_{0}}\log\Bigg{(}\frac{A_{\max}X_{\max}}{d}\Bigg{)}\right)+\lambda_{2}.

Here,

λ1:=1C18dc1(C1loglogt0+dC1loglogt0d2c1)2,λ2:=1d+2c1.\begin{split}\lambda_{1}:=\frac{1}{C_{1}}\cdot\frac{8d}{c_{1}}\cdot\left(\frac{C_{1}\log\log t_{0}+d}{C_{1}\log\log t_{0}-d-2c_{1}}\right)^{2},\qquad\lambda_{2}:=\frac{1}{d}+\frac{2}{c_{1}}.\end{split}

The number BB is defined as in Lemma 4 and depends on CC. The numbers AmaxA_{\max} and XmaxX_{\max} are defined as in (82) and depend on ϵ\epsilon and t0t_{0}. The number C1C_{1} is defined in (78), and depends on CC, ϵ\epsilon and t0t_{0}. The number c1c_{1} is defined in (84) and depends on ϵ\epsilon, t0t_{0} and c0=1/21.233c_{0}=1/21.233.

6.2. Bounding |1/ζ(1+it)||1/\zeta(1+it)|

Moving from a bound on the logarithmic derivative of ζ\zeta to one on the reciprocal of ζ\zeta is done in the usual way (see [Tit86, Theorem 3.11]), with some improvements using the trigonometric polynomial (see [Car+22, Proposition A.2]).

Lemma 7.

Let t0eet_{0}\geq e^{e} and d>0d>0 be constants. Suppose that for each tt0t\geq t_{0},

if1σ1+2dloglogtlogt,then|ζ(σ+it)ζ(σ+it)|R1logtloglogt.\text{if}\quad 1\leq\sigma\leq 1+\frac{2d\log\log t}{\log t},\quad\text{then}\quad\left|\frac{\zeta^{\prime}(\sigma+it)}{\zeta(\sigma+it)}\right|\leq R_{1}\frac{\log t}{\log\log t}.

Then for any tt0t\geq t_{0} and any real number d1d_{1} such that 0<d12d0<d_{1}\leq 2d, we have

|1ζ(1+it)|Q2(d1,t0)logtloglogt,\left|\frac{1}{\zeta(1+it)}\right|\leq Q_{2}(d_{1},t_{0})\frac{\log t}{\log\log t},

where

Q2(d1,t0)=exp(R1d1)(loglogt0logt0+1d1)3/4(C(d1,2t0)(log2logt0+1))1/4.Q_{2}(d_{1},t_{0})=\exp(R_{1}d_{1})\left(\frac{\log\log t_{0}}{\log t_{0}}+\frac{1}{d_{1}}\right)^{3/4}\left(C^{\prime}(d_{1},2t_{0})\left(\frac{\log 2}{\log t_{0}}+1\right)\right)^{1/4}.

The number CC^{\prime} is defined as in (74).

Proof.

Suppose tt0t\geq t_{0}. Let d1d_{1} be a real positive parameter such that d12dd_{1}\leq 2d. Let δ1=d1loglogt/logt\delta_{1}=d_{1}\log\log t/\log t. It is easy to see that

log|1ζ(1+it)|\displaystyle\log\left|\frac{1}{\zeta(1+it)}\right| =Relogζ(1+it)\displaystyle=-\textup{Re}\log\zeta(1+it)
=Relogζ(1+δ1+it)+11+δ1Re(ζζ(σ+it))dσ.\displaystyle=-\textup{Re}\log\zeta\left(1+\delta_{1}+it\right)+\int_{1}^{1+\delta_{1}}\textup{Re}\left(\frac{\zeta^{\prime}}{\zeta}(\sigma+it)\right)\text{d}\sigma.

So, using our suppositions, we have

(87) log|1ζ(1+it)|log|ζ(1+δ1+it)|+R1d1.\log\left|\frac{1}{\zeta(1+it)}\right|\leq-\log\left|\zeta\left(1+\delta_{1}+it\right)\right|+R_{1}d_{1}.

On the other hand, for σ>1\sigma>1, and using the classical nonnegativity argument involving the trigonometric polynomial 3+4cosθ+cos2θ3+4\cos\theta+\cos 2\theta [Tit86, Section 3.3], gives that

ζ3(σ)|ζ4(σ+it)ζ(σ+2it)|1.\zeta^{3}(\sigma)|\zeta^{4}(\sigma+it)\zeta(\sigma+2it)|\geq 1.

Combining this with the inequality ζ(σ)σ/(σ1)\zeta(\sigma)\leq\sigma/(\sigma-1) and Lemma 5, we therefore obtain, on taking σ=1+δ1\sigma=1+\delta_{1},

(88) |1ζ(1+δ1+it)||ζ(1+δ1)|3/4|ζ(1+δ1+2it)|1/4(1+δ1δ1)3/4(C(d1,2t0)log2tloglog2t)1/4.\begin{split}\left|\frac{1}{\zeta(1+\delta_{1}+it)}\right|&\leq|\zeta(1+\delta_{1})|^{3/4}|\zeta(1+\delta_{1}+2it)|^{1/4}\\ &\leq\left(\frac{1+\delta_{1}}{\delta_{1}}\right)^{3/4}\left(C^{\prime}(d_{1},2t_{0})\frac{\log 2t}{\log\log 2t}\right)^{1/4}.\end{split}

We now observe

(89) 1+δ1δ1(loglogt0logt0+1d1)logtloglogt,log2tloglog2t(log2logt0+1)logtloglogt.\begin{split}\frac{1+\delta_{1}}{\delta_{1}}&\leq\left(\frac{\log\log t_{0}}{\log t_{0}}+\frac{1}{d_{1}}\right)\frac{\log t}{\log\log t},\\ \frac{\log 2t}{\log\log 2t}&\leq\left(\frac{\log 2}{\log t_{0}}+1\right)\frac{\log t}{\log\log t}.\end{split}

So that, combining (89) and (88), and noting by (87) we have

|1ζ(1+it)|\displaystyle\left|\frac{1}{\zeta(1+it)}\right| exp(R1d1)|ζ(1+δ1+it)|,\displaystyle\leq\frac{\exp(R_{1}d_{1})}{|\zeta(1+\delta_{1}+it)|},

the lemma follows. ∎

6.3. Numeric calculations

We define the following upper bounds:

Q1(C,d,ϵ,t0)R1 and Q2(d1,t0)R2.Q_{1}(C,d,\epsilon,t_{0})\leq R_{1}\qquad\text{ and }\qquad Q_{2}(d_{1},t_{0})\leq R_{2}.

The numeric calculations for R1R_{1} and R2R_{2} proceed as follows. First fix t0eet_{0}\geq e^{e} and a positive ϵ1\epsilon\leq 1. For the purposes of obtaining an upper bound R1R_{1}, we then optimise Q1(C,d,ϵ,t0)Q_{1}(C,d,\epsilon,t_{0}) in Lemma 6 over CC and dd, all the while ensuring CC, dd, and ϵ\epsilon are subject to our constraints (68), (81) and (85)

Thereafter, with these values of dd and R1R_{1}, the relationship d12dd_{1}\leq 2d allows us to obtain R2R_{2} by optimising Q2(d1,t0)Q_{2}(d_{1},t_{0}) in Lemma 7 over d1d_{1}.

Our choices of

t0=500,ϵ=0.517,\displaystyle t_{0}=500,\qquad\epsilon=0.517,
C=e28,d=0.018,d1=0.0065,\displaystyle C=\frac{e^{2}}{8},\qquad d=0.018,\qquad d_{1}=0.0065,

give us

r\displaystyle r 0.502,\displaystyle\leq 0.502,\qquad α0.0381,\displaystyle\alpha\leq 0.0381,\qquad β=0.274,\displaystyle\beta=0.274\ldots,

and the upper bounds

R1=154.5,R2=431.7.R_{1}=154.5,\qquad R_{2}=431.7.

In summary, we have for t500t\geq 500,

(90) 1|ζ(1+it)|431.7logtloglogt.\frac{1}{|\zeta(1+it)|}\leq 431.7\frac{\log t}{\log\log t}.

Furthermore, in the region

1σ1+9250loglogtlogt,1\leq\sigma\leq 1+\frac{9}{250}\frac{\log\log t}{\log t},

we have for t500t\geq 500,

(91) |ζζ(σ+it)|154.5logtloglogt.\left|\frac{\zeta^{\prime}}{\zeta}(\sigma+it)\right|\leq 154.5\frac{\log t}{\log\log t}.

Finally, to cover the range from 3t5003\leq t\leq 500 in (90), we refer to the proof of Proposition A.2A.2 in [Car+22], where via interval arithmetic, they computed that for 2t5002\leq t\leq 500,

1|ζ(1+it)|2.079logt.\frac{1}{|\zeta(1+it)|}\leq 2.079\log t.

Comparing this with the estimate in (90), we see after a simple verification that the bound

1|ζ(1+it)|431.7logtloglogt\frac{1}{|\zeta(1+it)|}\leq 431.7\frac{\log t}{\log\log t}

holds for all t3t\geq 3.

This concludes our proof of Theorem 2. These bounds can certainly be improved. Some possible avenues of attack would be to treat the sum over zeros in the proof of Lemma 16 more carefully, or to use a larger zero-free region.

Remark.

It is possible with our methods to directly get a bound like (91) for t<500t<500, especially if one were not concerned with the size of R2R_{2}. For instance, choosing t0=16t_{0}=16, ϵ=0.508\epsilon=0.508, C=e2/8C=e^{2}/8, d=0.0147d=0.0147 and d1=0.0056d_{1}=0.0056, would return for t16t\geq 16 that if

1σ1+1475000loglogtlogt,1\leq\sigma\leq 1+\frac{147}{5000}\frac{\log\log t}{\log t},

then

|ζζ(σ+it)|177.5logtloglogt.\left|\frac{\zeta^{\prime}}{\zeta}(\sigma+it)\right|\leq 177.5\frac{\log t}{\log\log t}.

While this would cause a modest jump in R1R_{1}, the size of R2R_{2} would increase significantly to R2=511.1R_{2}=511.1. This is because R2R_{2} increases exponentially with R1R_{1}, as evident from Lemma 7. The dominant factor there is CC^{\prime}, and any improvement in that quantity would help greatly.

7. Background results

7.1. Background results related to Sections 3-5

Lemma 8.

[Fra21, Lemma 2.1] Let γ\gamma denote Euler’s constant. For N1N\geq 1,

n=1N1n<logN+γ+12N112N2+164N4.\sum_{n=1}^{N}\frac{1}{n}<\log N+\gamma+\frac{1}{2N}-\frac{1}{12N^{2}}+\frac{1}{64N^{4}}.
Theorem 9.

[Pat22, Theorem 3.9] For t>0t>0 and n1=t/(2π)n_{1}=\lfloor\sqrt{t/(2\pi)}\rfloor we have

|ζ(1+it)||n=1n11n1+it|+g(t)t1/2|n=1n11nit|+|\zeta(1+it)|\leq\left|\sum_{n=1}^{n_{1}}\frac{1}{n^{1+it}}\right|+\frac{g(t)}{t^{1/2}}\left|\sum_{n=1}^{n_{1}}\frac{1}{n^{-it}}\right|+\mathcal{R}

where

g(t)=2πexp(53t2+π6t),g(t)=\sqrt{2\pi}\exp\left(\frac{5}{3t^{2}}+\frac{\pi}{6t}\right),

and

=(t):=1t1/2(π2+g(t)2)+1t(9π2+g(t)π(32log2))+1t3/2(968π3/2+g(t)242π700).\begin{split}\mathcal{R}=\mathcal{R}(t):=&\frac{1}{t^{1/2}}\left(\sqrt{\frac{\pi}{2}}+\frac{g(t)}{2}\right)+\frac{1}{t}\left(9\sqrt{\frac{\pi}{2}}+\frac{g(t)}{\sqrt{\pi(3-2\log 2)}}\right)\\ &+\frac{1}{t^{3/2}}\left(\frac{968\pi^{3/2}+g(t)242\pi}{700}\right).\end{split}

Next we review some explicit estimates of exponential sums from the literature. The following lemma is a specialised third-derivative test for a phase function commonly encountered when estimating the Riemann zeta-function. It is due to [HPY24, Lemma 2.6] which builds on [Hia16].

Lemma 10 (Explicit third derivative test).

[HPY24, Lemma 2.6] Let rr and KK be positive integers, and let tt and K0K_{0} be positive numbers. Suppose KK0>1K\geq K_{0}>1. Let

f(x):=t2πlog(rK+x),W:=π(r+1)3K3t,λ:=(1+r)3r3,\displaystyle f(x):=\frac{t}{2\pi}\log\left(rK+x\right),\qquad W:=\frac{\pi(r+1)^{3}K^{3}}{t},\qquad\lambda:=\frac{(1+r)^{3}}{r^{3}},

so that 1/W|f′′′(x)|λ/W1/W\leq|f^{\prime\prime\prime}(x)|\leq\lambda/W for 0xK0\leq x\leq K. Then, for each positive integer LKL\leq K and any η>0\eta>0,

|n=0L1e2πif(n)|2(KW1/3+η)(αK+βW2/3)\left|\sum_{n=0}^{L-1}e^{2\pi if(n)}\right|^{2}\leq\left(\frac{K}{W^{1/3}}+\eta\right)(\alpha K+\beta W^{2/3})

where

α\displaystyle\alpha =α(W,λ):=1η+ημ3W1/3+μ3W2/3+32μ15πη+W1/3,\displaystyle=\alpha(W,\lambda):=\frac{1}{\eta}+\frac{\eta\mu}{3W^{1/3}}+\frac{\mu}{3W^{2/3}}+\frac{32\mu}{15\sqrt{\pi}}\sqrt{\eta+W^{-1/3}},
β\displaystyle\beta =β(W):=323πη+(24π)1W1/3,\displaystyle=\beta(W):=\frac{32}{3\sqrt{\pi\eta}}+\left(2-\frac{4}{\pi}\right)\frac{1}{W^{1/3}},
μ\displaystyle\mu =μ(λ):=12λ2/3(1+1(1K01)λ1/3).\displaystyle=\mu(\lambda):=\frac{1}{2}\lambda^{2/3}\left(1+\frac{1}{(1-K_{0}^{-1})\lambda^{1/3}}\right).
Remark.

If W1W\leq 1, say, then α\alpha will be greater than 11, and so αL2L2\alpha L^{2}\geq L^{2}. Hence, for small WW, it is better to use the triangle inequality than Lemma 10.

Remark.

We mention that there is a typo in the statement of [HPY24, Lemma 2.6] where NN in the statement of the lemma should be 0. This lemma is only applied with N=0N=0 in [HPY24] in any case.888In passing, let us also point out that on [HPY24, 211], the minimum in the definition of yjy_{j} should be against LmL-m rather than Lm1L-m-1. This does not affect Equation 5.4 on p. 211 since, as stated in [HPY24, Lemma 2.5], the bound on |g′′(x)||g^{\prime\prime}(x)| holds up to KmK-m, not only Km1K-m-1. None of the results in [HPY24] are affected.

The next two lemmas respectively represent another third derivative test and a kkth derivative test for k4k\geq 4, due to [Yan24]. The reason we state this additional third derivative test is because A3A_{3} and B3B_{3} comprise the boundary cases for the recursive formulas in Lemma 12.

Lemma 11 (Another explicit third derivative test).

[Yan24, Lemma 2.4] Let aa and NN be integers such that N1N\geq 1. Suppose f(x)f(x) has three continuous derivatives and f′′′(x)f^{\prime\prime\prime}(x) is monotonic over the interval (a,a+N](a,a+N]. Suppose further that there are numbers λ3>0\lambda_{3}>0 and ω>1\omega>1 such that λ3|f′′′(x)|ωλ3\lambda_{3}\leq|f^{\prime\prime\prime}(x)|\leq\omega\lambda_{3} for all x(a,a+N]x\in(a,a+N]. Then for any η>0\eta>0,

|n=a+1a+Ne2πif(n)|A3ω1/2Nλ31/6+B3N1/2λ31/6\left|\sum_{n=a+1}^{a+N}e^{2\pi if(n)}\right|\leq A_{3}\omega^{1/2}N\lambda_{3}^{1/6}+B_{3}N^{1/2}\lambda_{3}^{-1/6}

where

A3=A3(η,ω):=1ηω+3215πη+λ01/3+13(η+λ01/3)λ01/3δ3,\begin{split}&A_{3}=A_{3}(\eta,\omega):=\sqrt{\frac{1}{\eta\omega}+\frac{32}{15\sqrt{\pi}}\sqrt{\eta+\lambda_{0}^{1/3}}+\frac{1}{3}\left(\eta+\lambda_{0}^{1/3}\right)\lambda_{0}^{1/3}}\delta_{3},\end{split}
B3=B3(η):=323π1/4η1/4δ3,\begin{split}&B_{3}=B_{3}(\eta):=\frac{\sqrt{32}}{\sqrt{3}\pi^{1/4}\eta^{1/4}}\delta_{3},\end{split}

and λ0\lambda_{0} and δ3\delta_{3} are defined by

λ0=λ0(η,ω):=(1η+32η1/2ω15π)3,\begin{split}&\lambda_{0}=\lambda_{0}(\eta,\omega):=\left(\frac{1}{\eta}+\frac{32\eta^{1/2}\omega}{15\sqrt{\pi}}\right)^{-3},\end{split}
δ3=δ3(η):=12+121+38π1/2η3/2.\begin{split}&\delta_{3}=\delta_{3}(\eta):=\sqrt{\frac{1}{2}+\frac{1}{2}\sqrt{1+\frac{3}{8}\pi^{1/2}\eta^{3/2}}}.\end{split}
Lemma 12 (General explicit kk-th derivative test).

[Yan24, Lemma 2.5] Let aa, NN and kk be integers such that N1N\geq 1 and k4k\geq 4. Suppose f(x)f(x) is equipped with kk continuous derivatives and f(k)(x)f^{(k)}(x)is monotonic over the interval (a,a+N](a,a+N]. Suppose further that there are numbers λk>0\lambda_{k}>0 and ω>1\omega>1 such that λk|f(k)(x)|ωλk\lambda_{k}\leq|f^{(k)}(x)|\leq\omega\lambda_{k} for all x(a,a+N]x\in(a,a+N]. Then for any η>0\eta>0,

|n=a+1a+Ne2πif(n)|Akω2/KNλk1/(2K2)+BkN12/Kλk1/(2K2),\left|\sum_{n=a+1}^{a+N}e^{2\pi if(n)}\right|\leq A_{k}\omega^{2/K}N\lambda_{k}^{1/(2K-2)}+B_{k}N^{1-2/K}\lambda_{k}^{-1/(2K-2)},

where K=2k1K=2^{k-1}, AkA_{k} and BkB_{k} are defined recursively via the formulas

(92) Ak+1(η,ω):=δk(ω1/K+219/12(K1)(2K1)(4K3)Ak(η,ω)1/2),A_{k+1}(\eta,\omega):=\delta_{k}\left(\omega^{-1/K}+\frac{2^{19/12}(K-1)}{\sqrt{(2K-1)(4K-3)}}A_{k}(\eta,\omega)^{1/2}\right),
(93) Bk+1(η):=δk23/2(K1)(2K3)(4K5)Bk(η)1/2,B_{k+1}(\eta):=\delta_{k}\frac{2^{3/2}(K-1)}{\sqrt{(2K-3)(4K-5)}}B_{k}(\eta)^{1/2},

with A3A_{3} and B3B_{3} as in Lemma 11, and δk\delta_{k} defined by

(94) δk=δk(η):=1+2233712/K(9π1024η)1/K.\delta_{k}=\delta_{k}(\eta):=\sqrt{1+\frac{2}{2337^{1-2/K}}\left(\frac{9\pi}{1024}\eta\right)^{1/K}}.
Lemma 13.

Let aa, bb and kk be integers such that a>0a>0, b>ab>a and k3k\geq 3. Let t>0t>0. Let h>1h>1 be a number such that bhab\leq ha. Then for any η>0\eta>0,

|a<nb1n1+it|Ckak/(2K2)t1/(2K2)+Dkak/(2K2)2/Kt1/(2K2),\left|\sum_{a<n\leq b}\frac{1}{n^{1+it}}\right|\leq C_{k}a^{-k/(2K-2)}t^{1/(2K-2)}+D_{k}a^{k/(2K-2)-2/K}t^{-1/(2K-2)},

where

Ck=Ck(η,h):=Ak(η,hk)h2k/Kk/(2K2)(h1)((k1)!2π)12K2,Dk=Dk(η,h):=Bk(η)hk/(2K2)(h1)12/K(2π(k1)!)12K2.\begin{split}C_{k}=C_{k}(\eta,h)&:=A_{k}(\eta,h^{k})h^{2k/K-k/(2K-2)}(h-1)\left(\frac{(k-1)!}{2\pi}\right)^{\frac{1}{2K-2}},\\ D_{k}=D_{k}(\eta,h)&:=B_{k}(\eta)h^{k/(2K-2)}(h-1)^{1-2/K}\left(\frac{2\pi}{(k-1)!}\right)^{\frac{1}{2K-2}}.\end{split}

and Ak(η,hk)A_{k}(\eta,h^{k}) and Bk(η)B_{k}(\eta) defined in accordance with Lemma 12.

Proof.

This is a special case of Lemma 12 and Lemma 11 with the phase function

f(x)=t2πlogx,and hencef(k)(x)=(1)kt2π(k1)!xk.f(x)=-\frac{t}{2\pi}\log x,\qquad\text{and hence}\qquad f^{(k)}(x)=(-1)^{k}\frac{t}{2\pi}\frac{(k-1)!}{x^{k}}.

Since a<bhaa<b\leq ha by assumption, λk|f(k)(x)|ωλk\lambda_{k}\leq|f^{(k)}(x)|\leq\omega\lambda_{k} for a<xba<x\leq b, where

λk=t2π(k1)!hkak,ω=hk.\lambda_{k}=\frac{t}{2\pi}\frac{(k-1)!}{h^{k}a^{k}},\qquad\omega=h^{k}.

Suppose that k4k\geq 4. Then, we may apply Lemma 12 with these values of λk\lambda_{k} and ω\omega, and with N=baN=b-a. This furnishes a bound that increases with NN. Noting that N(h1)aN\leq(h-1)a, we thus obtain

|a<nb1nit|\displaystyle\left|\sum_{a<n\leq b}\frac{1}{n^{it}}\right| Ak(η,hk)(hk)2/K(h1)a(t2π(k1)!hkak)1/(2K2)\displaystyle\leq A_{k}(\eta,h^{k})(h^{k})^{2/K}(h-1)a\left(\frac{t}{2\pi}\frac{(k-1)!}{h^{k}a^{k}}\right)^{1/(2K-2)}
+Bk(η)((h1)a)12/K(t2π(k1)!hkak)1/(2K2)\displaystyle\qquad\qquad+B_{k}(\eta)((h-1)a)^{1-2/K}\left(\frac{t}{2\pi}\frac{(k-1)!}{h^{k}a^{k}}\right)^{-1/(2K-2)}
=Cka1k/(2K2)t1/(2K2)+Dka12/K+k/(2K2)t1/(2K2).\displaystyle=C_{k}a^{1-k/(2K-2)}t^{1/(2K-2)}+D_{k}a^{1-2/K+k/(2K-2)}t^{-1/(2K-2)}.

The desired result hence follows from partial summation in this case. If k=3k=3, then the result follows from Lemma 11 using a similar calculation. ∎

7.2. Background results related to Section 6

Here we give some preliminary lemmas for our proofs of estimates on |ζ(1+it)/ζ(1+it)||\zeta^{\prime}(1+it)/\zeta(1+it)| and |1/ζ(1+it)||1/\zeta(1+it)|. We reiterate that Section 6 reuses variable names from previous sections. These variables mean different things in the confines of this section as this section is independent from Sections 3, 4, and 5.

The first two lemmas, which form the basis of our method, are built upon a theorem of Borel and Carathéodory. This theorem enables us to deduce an upper bound for the modulus of a function and its derivatives on a circle, from bounds for its real part on a larger concentric circle.

Theorem 14 (Borel–Carathéodory).

Let s0s_{0} be a complex number. Let RR be a positive number, possibly depending on s0s_{0}. Suppose that the function f(s)f(s) is analytic in a region containing the disk |ss0|R|s-s_{0}|\leq R. Let MM denote the maximum of Ref(s)\textup{Re}\,f(s) on the boundary |ss0|=R|s-s_{0}|=R. Then, for any r(0,R)r\in(0,R) and any ss such that |ss0|r|s-s_{0}|\leq r,

|f(s)|2rRrM+R+rRr|f(s0)|.|f(s)|\leq\frac{2r}{R-r}M+\frac{R+r}{R-r}|f(s_{0})|.

If in addition f(s0)=0f(s_{0})=0, then for any r(0,R)r\in(0,R) and any ss such that |ss0|r|s-s_{0}|\leq r,

|f(s)|2R(Rr)2M.|f^{\prime}(s)|\leq\frac{2R}{(R-r)^{2}}M.
Proof.

Proofs for estimating both |f(s)||f(s)| and |f(s)||f^{\prime}(s)| are available in [Tit39, §5.5]. However, if we include the restriction f(s0)=0f(s_{0})=0, we can refer to [MV07, Lemma 6.2], which provides alternative proofs, resulting in a sharper estimate for |f(s)||f^{\prime}(s)|. ∎

Lemma 15.

Let s0s_{0} be a complex number. Let rr and α\alpha be positive numbers (possibly depending on s0s_{0}) such that α<1/2\alpha<1/2. Suppose that the function f(s)f(s) is analytic in a region containing the disk |ss0|r|s-s_{0}|\leq r. Suppose further there is a number A1A_{1} independent of ss such that,

|f(s)f(s0)|A1,|ss0|r.\left|\frac{f(s)}{f(s_{0})}\right|\leq A_{1},\qquad|s-s_{0}|\leq r.

Then, for any ss in the disk |ss0|αr|s-s_{0}|\leq\alpha r we have

|f(s)f(s)ρ1sρ|4logA1r(12α)2,\left|\frac{f^{\prime}(s)}{f(s)}-\sum_{\rho}\frac{1}{s-\rho}\right|\leq\frac{4\log A_{1}}{r(1-2\alpha)^{2}},

where ρ\rho runs through the zeros of f(s)f(s) in the disk |ss0|12r|s-s_{0}|\leq\tfrac{1}{2}r, counted with multiplicity.

Proof.

We follow the proof in [Tit86, Section 3]. Let gg be the function

g(s)=f(s)ρ1sρ,g(s)=f(s)\prod_{\rho}\frac{1}{s-\rho},

where ρ\rho in the (finite) product runs through the zeros of f(s)f(s), counted with multiplicity, that satisfy |ρs0|12r|\rho-s_{0}|\leq\tfrac{1}{2}r. Since by construction the poles and zeros cancel, gg is analytic in |ss0|r|s-s_{0}|\leq r, and does not vanish in the disk |ss0|12r|s-s_{0}|\leq\tfrac{1}{2}r. Therefore, the function

h(s)=logg(s)g(s0),h(s)=\log\frac{g(s)}{g(s_{0})},

where the logarithm branch is determined by h(s0)=0h(s_{0})=0, is analytic in some region containing the disk |ss0|12r|s-s_{0}|\leq\tfrac{1}{2}r.

Now, on the circle |ss0|=r|s-s_{0}|=r, we have |sρ|12r|s0ρ||s-\rho|\geq\tfrac{1}{2}r\geq|s_{0}-\rho|. So, on this circle,

|g(s)g(s0)|=|f(s)f(s0)ρ(s0ρsρ)||f(s)f(s0)|A1.\left|\frac{g(s)}{g(s_{0})}\right|=\left|\frac{f(s)}{f(s_{0})}\prod_{\rho}\left(\frac{s_{0}-\rho}{s-\rho}\right)\right|\leq\left|\frac{f(s)}{f(s_{0})}\right|\leq A_{1}.

The inequality also holds in the interior of the circle, by the maximum modulus principle. Hence, Reh(s)logA1\textup{Re}\,h(s)\leq\log A_{1} throughout the disk |ss0|r|s-s_{0}|\leq r.

Given this, we apply Theorem 14 to h(s)h(s) with the concentric circles centered at s0s_{0} of radii αr\alpha r and 12r\frac{1}{2}r. This yields throughout the disk |ss0|αr|s-s_{0}|\leq\alpha r that

(95) |h(s)|4logA1r(12α)2.|h^{\prime}(s)|\leq\frac{4\log A_{1}}{r(1-2\alpha)^{2}}.

On the other hand, since g(s0)g(s_{0}) is constant, we have

|h(s)|=|g(s)g(s)|=|f(s)f(s)ρ1sρ|.|h^{\prime}(s)|=\left|\frac{g^{\prime}(s)}{g(s)}\right|=\left|\frac{f^{\prime}(s)}{f(s)}-\sum_{\rho}\frac{1}{s-\rho}\right|.

Combining this formula with the inequality (95) completes the proof. ∎

Lemma 16.

Let ss and s0s_{0} be complex numbers with real parts σ\sigma and σ0\sigma_{0}, respectively. Let rr, α\alpha, β\beta, A1A_{1} and A2A_{2} be positive numbers, possibly depending on s0s_{0}, such that α<1/2\alpha<1/2 and β<1\beta<1. Suppose that the function f(s)f(s) satisfies the conditions of Lemma 15 with rr, α\alpha and A1A_{1}, and that

|f(s0)f(s0)|A2.\left|\frac{f^{\prime}(s_{0})}{f(s_{0})}\right|\leq A_{2}.

Suppose, in addition, that f(s)0f(s)\neq 0 for any ss in both the disk |ss0|r|s-s_{0}|\leq r and the right half-plane σσ0αr\sigma\geq\sigma_{0}-\alpha r. Then, for any ss in the disk |ss0|αβr|s-s_{0}|\leq\alpha\beta r,

|f(s)f(s)|8βlogA1r(1β)(12α)2+1+β1βA2.\left|\frac{f^{\prime}(s)}{f(s)}\right|\leq\frac{8\beta\log A_{1}}{r(1-\beta)(1-2\alpha)^{2}}+\frac{1+\beta}{1-\beta}A_{2}.
Proof.

Using the inequality |z|Rez|z|-|z|\leq-\textup{Re}\,z\leq|z|, valid for any complex number zz, together with Lemma 15 in |ss0|αr|s-s_{0}|\leq\alpha r, we have throughout this disk,

(96) Ref(s)f(s)\displaystyle-\textup{Re}\,\frac{f^{\prime}(s)}{f(s)} 4logA1r(12α)2ρRe1sρ,\displaystyle\leq\frac{4\log A_{1}}{r(1-2\alpha)^{2}}-\sum_{\rho}\textup{Re}\,\frac{1}{s-\rho},

where ρ\rho runs through the zeros of ff (with multiplicity) satisfying |ρs0|12r|\rho-s_{0}|\leq\tfrac{1}{2}r. By the assumption on the nonvanishing of ff, each ρ\rho in this sum satisfies Reρ<σ0αr\textup{Re}\,\rho<\sigma_{0}-\alpha r. On the other hand, all ss in the disk |ss0|αr|s-s_{0}|\leq\alpha r clearly satisfy σσ0αr\sigma\geq\sigma_{0}-\alpha r. Hence,

ρRe1sρ=ρσReρ|sρ|2>0,\sum_{\rho}\textup{Re}\,\frac{1}{s-\rho}=\sum_{\rho}\frac{\sigma-\textup{Re}\,\rho}{|s-\rho|^{2}}>0,

throughout the disk |ss0|αr|s-s_{0}|\leq\alpha r. We may thus drop the sum in (96) and still obtain a valid upper bound, call it M1M_{1}.

Now, f(s)/f(s)-f^{\prime}(s)/f(s) is analytic in the disk |ss0|αr|s-s_{0}|\leq\alpha r since by assumption ff does not vanish there. We can therefore apply Theorem 14 to f(s)/f(s)-f^{\prime}(s)/f(s) using the two concentric circles |ss0|αβr|s-s_{0}|\leq\alpha\beta r and |ss0|αr|s-s_{0}|\leq\alpha r, and the upper bound M1MM_{1}\geq M on the circle |ss0|αr|s-s_{0}|\leq\alpha r, which yields the result. ∎

Acknowledgements

We would like to thank Fatima Majeed for pointing out errors in an early version of this manuscript and giving us some helpful suggestions.

References

  • [Car+22] Emanuel Carneiro, Andrés Chirre, Harald Andrés Helfgott and Julián Mejía-Cordero “Optimality for the two-parameter quadratic sieve” In Acta Arithmetica 203 Instytut Matematyczny Polskiej Akademii Nauk, 2022, pp. 195–226
  • [Chi23] Andrés Chirre “Bounding zeta on the 11-line under the partial Riemann hypothesis” In Bulletin of the Australian Mathematical Society Cambridge University Press, 2023, pp. 1–8
  • [CHS24] Andrés Chirre, Markus Valås Hagen and Aleksander Simonič “Conditional estimates for the logarithmic derivative of Dirichlet L-functions” In Indagationes Mathematicae 35.1, 2024, pp. 14–27
  • [Del87] Hubert Delange “Une remarque sur la dérivée logarithmique de la fonction zêta de Riemann” In Colloquium Mathematicum 53.2, 1987, pp. 333–335 Polska Akademia Nauk. Instytut Matematyczny PAN
  • [For02] Kevin Ford “Vinogradov’s Integral and Bounds for the Riemann Zeta Function” In Proceedings of the London Mathematical Society 85.3, 2002, pp. 565–633
  • [Fra21] F. J. Francis “An investigation into explicit versions of Burgess’ bound” In J. Number Theory 228 Elsevier, 2021, pp. 87–107
  • [HSW22] E. Hasanalizade, Q. Shen and P-J. Wong “Counting zeros of the Riemann zeta function” In J. Number Theory 235 Elsevier, 2022, pp. 219–241
  • [HPY24] G. A. Hiary, D. Patel and A. Yang “An improved explicit estimate for ζ(1/2+it)\zeta(1/2+it) In J. Number Theory 256 Elsevier, 2024, pp. 195–217
  • [Hia16] Ghaith A. Hiary “An explicit van der Corput estimate for ζ(1/2+it)\zeta(1/2+it) In Indagationes Mathematicae 27.2, 2016, pp. 524–533
  • [KLN18] Habiba Kadiri, Allysa Lumley and Nathan Ng “Explicit zero density for the Riemann zeta function” In Journal of Mathematical Analysis and Applications 465.1, 2018, pp. 22–46
  • [LLS15] Youness Lamzouri, Xiannan Li and Kannan Soundararajan “Conditional bounds for the least quadratic non-residue and related problems” In Mathematics of Computation 84.295, 2015, pp. 2391–2412 DOI: 10.1090/S0025-5718-2015-02925-1
  • [Li10] Xiannan Li “Upper bounds on LL-functions at the edge of the critical strip” In Int. Math. Res. Not. IMRN, 2010, pp. 727–755 DOI: 10.1093/imrn/rnp148
  • [Lit26] J. E. Littlewood “On the Riemann Zeta-Function” In Proceedings of the London Mathematical Society s2-24.1, 1926, pp. 175–201 DOI: https://doi.org/10.1112/plms/s2-24.1.175
  • [Lit28] J. E. Littlewood “Mathematical Notes (5): On the Function 1/ζ(1+ti)1/\zeta(1+ti) In Proceedings of the London Mathematical Society s2-27.1, 1928, pp. 349–357
  • [Lum18] Allysa Lumley “Explicit bounds for L{L}-functions on the edge of the critical strip” In Journal of Number Theory 188 Elsevier, 2018, pp. 186–209
  • [MV07] H. L. Montgomery and R. C. Vaughan “Multiplicative number theory I: Classical theory” Cambridge university press, 2007
  • [PS22] Neea Palojärvi and Aleksander Simonič “Conditional estimates for L{L}-functions in the Selberg class” In arXiv preprint arXiv:2211.01121, 2022
  • [Pat21] Dhir Patel “Explicit sub-Weyl bound for the Riemann Zeta function”, 2021
  • [Pat22] Dhir Patel “An explicit upper bound for |ζ(1+it)||\zeta(1+it)| In Indagationes Mathematicae 33.5, 2022, pp. 1012–1032
  • [Ram16] O. Ramaré “An explicit density estimate for Dirichlet LL-series” In Mathematics of Computation 85.297, 2016, pp. 325–356
  • [Sim20] Aleksander Simonič “Explicit zero density estimate for the Riemann zeta-function near the critical line” In Journal of Mathematical Analysis and Applications 491.1, 124303, 41pp., 2020
  • [Sim23] Aleksander Simonič “Estimates for L-functions in the Critical Strip Under GRH with Effective Applications” In Mediterranean Journal of Mathematics 20.2, 2023, pp. 87
  • [Sou09] Kannan Soundararajan “Moments of the Riemann zeta function” In Ann. of Math. (2) 170.2, 2009, pp. 981–993 DOI: 10.4007/annals.2009.170.981
  • [Tit39] E. C. Titchmarsh “The Theory of Functions” Oxford university press, 1939
  • [Tit86] E. C. Titchmarsh “The Theory of the Riemann Zeta-function” Oxford: Oxford Science Publications, 1986
  • [Tru14] T. Trudgian “A new upper bound for |ζ(1+it)||\zeta(1+it)| In Bulletin of the Australian Mathematical Society 89.2, 2014
  • [Tru15] Tim Trudgian “Explicit bounds on the logarithmic derivative and the reciprocal of the Riemann zeta-function” In Functiones et Approximatio Commentarii Mathematici 52.2 Adam Mickiewicz University, Faculty of MathematicsComputer Science, 2015, pp. 253–261
  • [Tru14a] Timothy Trudgian “An improved upper bound for the argument of the Riemann zeta-function on the critical line II” In Journal of Number Theory 134, 2014, pp. 280–292
  • [Wey21] Hermann Weyl “Zur Abschätzung von ζ(1+ti)\zeta(1+ti) In Mathematische Zeitschrift 10.1-2, 1921, pp. 88–101 DOI: 10.1007/BF02102307
  • [Yan24] A. Yang “Explicit bounds on ζ(s)\zeta(s) in the critical strip and a zero-free region” In Journal of Mathematical Analysis and Applications 534.2 Elsevier, 2024, pp. 128124