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

Explicit bounds on ζ(s)\zeta(s) in the critical strip and a zero-free region

Andrew Yang
Abstract.

We derive explicit upper bounds for the Riemann zeta-function ζ(σ+it)\zeta(\sigma+it) on the lines σ=1k/(2k2)\sigma=1-k/(2^{k}-2) for integer k4k\geq 4. This is used to show that the zeta-function has no zeroes in the region

σ>1loglog|t|21.233log|t|,|t|3.\sigma>1-\frac{\log\log|t|}{21.233\log|t|},\qquad|t|\geq 3.

This is the largest known zero-free region for exp(171)texp(5.3105)\exp(171)\leq t\leq\exp(5.3\cdot 10^{5}). Our results rely on an explicit version of the van der Corput AnBA^{n}B process for bounding exponential sums.

Key words and phrases:
Riemann zeta-function, exponential sums, van der Corput method, zero-free region.
2010 Mathematics Subject Classification:
Primary: 11M06, 11M26; Secondary: 11Y35

1. Introduction

Bounding the size of the Riemann zeta-function ζ(s)\zeta(s) within the critical strip 0<s<10<\Re s<1 is a central goal in analytic number theory, with results having broad implications for the zeroes of ζ(s)\zeta(s) and thus the distribution of prime numbers [For02a, KLN18, Tru14a]. By far the most common result of this type are estimates of ζ(σ+it)\zeta(\sigma+it) as tt\to\infty and σ\sigma is fixed to either 1/21/2 or 11, i.e. bounds along the critical line and 1-line respectively. Currently, the best known unconditional results are ζ(1/2+it)εt13/84+ε\zeta(1/2+it)\ll_{\varepsilon}t^{13/84+\varepsilon} for any ε>0\varepsilon>0, due to Bourgain [Bou16], and ζ(1+it)log2/3t\zeta(1+it)\ll\log^{2/3}t, due to Vinogradov [Vin58]. Explicit results are also known, with the current best bound (for large tt) being |ζ(1/2+it)|307.098t27/164|\zeta(1/2+it)|\leq 307.098t^{27/164} (t3t\geq 3), proved by Patel [Pat21] and |ζ(1+it)|62.6log2/3t|\zeta(1+it)|\leq 62.6\log^{2/3}t (t3t\geq 3) due to Trudgian [Tru14].

The main focus of this paper is to prove explicit bounds for ζ(σ+it)\zeta(\sigma+it) for special values of σ(1/2,1)\sigma\in(1/2,1). For many applications of interest, such as zero-free regions and zero-density estimates, we require bounds holding uniformly in the strip 1/2σ11/2\leq\sigma\leq 1. The Vinogradov–Korobov zero-free region, for instance, relies on a Ford–Richert type result |ζ(σ+it)|AtB(1σ)3/2log2/3t|\zeta(\sigma+it)|\leq At^{B(1-\sigma)^{3/2}}\log^{2/3}t for all 1/2σ11/2\leq\sigma\leq 1 [Che00, Ric67, For02a]. Via a convexity argument (see e.g. [Tit86, §7.8]), we may use estimates of ζ(1/2+it)\zeta(1/2+it) and ζ(1+it)\zeta(1+it) to obtain bounds on ζ(σ+it)\zeta(\sigma+it) for any 1/2σ11/2\leq\sigma\leq 1. However, through van der Corput’s method of bounding exponential sums, we can directly derive asymptotically sharper bounds for specific values of σ\sigma. Titchmarsh [Tit86, Thm. 5.13] shows for instance that if σ=1k/(2k2)\sigma=1-k/(2^{k}-2) for some integer k4k\geq 4, then

ζ(σ+it)t1/(2k2)logt.\zeta(\sigma+it)\ll t^{1/(2^{k}-2)}\log t. (1.1)

This is sharper than what is achievable via convexity arguments for all k4k\geq 4.

Having an explicit version of (1.1) will allow us to improve many existing explicit results about ζ(s)\zeta(s). The main obstacle to making (1.1) explicit is the difficulty with bounding the implied constants of the kkth derivative test, obtained through van der Corput’s Ak2BA^{k-2}B process. In this work we refine existing approaches to explicit exponential sum theory, to obtain an explicit kkth derivative test with constants holding uniformly for all k3k\geq 3. This allows us to show the following theorem.

Theorem 1.1.

Let k4k\geq 4 be an integer and σk:=1k/(2k2)\sigma_{k}:=1-k/(2^{k}-2). Then

|ζ(σk+it)|1.546t1/(2k2)logt,t3.\left|\zeta\left(\sigma_{k}+it\right)\right|\leq 1.546t^{1/(2^{k}-2)}\log t,\qquad t\geq 3. (1.2)

For example, substituting k=4k=4 gives |ζ(5/7+it)|1.546t1/14logt|\zeta(5/7+it)|\leq 1.546t^{1/14}\log t. By comparison, the sharpest bound that can currently be obtained using bounds on ζ(1/2+it)\zeta(1/2+it), ζ(1+it)\zeta(1+it) and the convexity principle is ζ(5/7+it)εt13/147+ε\zeta(5/7+it)\ll_{\varepsilon}t^{13/147+\varepsilon}, where 13/147>1/1413/147>1/14. Theorem 1.1 is sharpest for small to moderately sized kk and tt. In particular, as kk\to\infty, (1.2) reduces to |ζ(1+it)|1.546logt|\zeta(1+it)|\leq 1.546\log t, which is weaker than other known bounds on the 1-line [Bac16, Pat22]. If we are only interested in large kk and tt, then Theorem 1.1 can be sharpened (see remarks in §5).

We briefly highlight some immediate applications of our results. The explicit kkth derivative test is an ingredient in deriving an explicit version of Littlewood’s bound ζ(1+it)logt/loglogt\zeta(1+it)\ll\log t/\log\log t [Tit86, Thm. 5.16]. Similarly, the kkth derivative test can be used to make explicit the bounds 1/ζ(1+it),ζ/ζ(1+it)logt/loglogt1/\zeta(1+it),\zeta^{\prime}/\zeta(1+it)\ll\log t/\log\log t, which are useful for bounding Mertens’ function M(x)M(x) [Tru15, LL22]. Additionally, Theorem 1.1 can be used to improve explicit bounds on S(t)S(t), the argument of the zeta-function along the critical line (see e.g. [Tru14a, HSW21]). In this work, we use Theorem 1.1 to prove an explicit version of Littlewood’s [Lit22] zero-free region of the form 1σloglogt/logt1-\sigma\ll\log\log t/\log t.

1.1. Littlewood’s zero-free region

Zero-free regions for ζ(s)\zeta(s) are widely studied partly due to their implications for prime distributions; some recent results include [Ste70, RS75, Kon77, Che00, For02a, Kad05, JK14, MT14, MTY24]. The current best explicit zero-free region for small tt is due to Mossinghoff, Trudgian and Yang [MTY24], who proved that there are no zeroes of ζ(σ+it)\zeta(\sigma+it) in the region

σ>115.558691log|t|,|t|2.\sigma>1-\frac{1}{5.558691\log|t|},\qquad|t|\geq 2. (1.3)

For intermediate tt, the following zero-free region is currently the sharpest known

σ>10.049620.0196/(J(|t|)+1.15)J(|t|)+0.685+0.155loglog|t|,|t|3,\sigma>1-\frac{0.04962-0.0196/(J(|t|)+1.15)}{J(|t|)+0.685+0.155\log\log|t|},\qquad|t|\geq 3, (1.4)

where J(t):=16logt+loglogt+log0.618J(t):=\frac{1}{6}\log t+\log\log t+\log 0.618. This is formed by substituting [HPY24, Thm. 1.1] into [For02a, Thm. 3] and noting that J(t)<14logt+1.8521J(t)<\frac{1}{4}\log t+1.8521 for t3t\geq 3.

For large tt, the following Vinogradov–Korobov zero-free region due to [MTY24], building on the method of Ford [For02a, For22], is currently the sharpest known

σ>1155.241(log|t|)2/3(loglog|t|)1/3,|t|3.\sigma>1-\frac{1}{55.241(\log|t|)^{2/3}(\log\log|t|)^{1/3}},\qquad|t|\geq 3. (1.5)

In this work we use Theorem 1.1 and the explicit kkth derivative test to prove the following zero-free region.

Corollary 1.2.

There are no zeroes of ζ(σ+it)\zeta(\sigma+it) in the region

σ>1loglog|t|21.233log|t|,|t|3.\sigma>1-\frac{\log\log|t|}{21.233\log|t|},\qquad|t|\geq 3. (1.6)

This represents the largest known zero-free region in the range exp(170.3)texp(532 141)\exp(170.3)\leq t\leq\exp(532\,141). To summarise the current state of knowledge for other ranges of tt: for t31012t\leq 3\cdot 10^{12}, all zeroes are known to lie on the critical line, due to the computational verification performed in [PT21]. For 31012<texp(46.2)3\cdot 10^{12}<t\leq\exp(46.2), (1.3) is the sharpest known zero-free region; for exp(46.3)texp(170.2)\exp(46.3)\leq t\leq\exp(170.2), (1.4) is sharpest; for texp(532 142)t\geq\exp(532\,142), (1.5) is the sharpest.

1.2. Approach

The main tool used to establish Theorem 1.1 are upper bounds on sums of the form

Sf(a,N):=|a<na+Ne(f(n))|S_{f}(a,N):=\left|\sum_{a<n\leq a+N}e(f(n))\right| (1.7)

where e(x):=exp(2πix)e(x):=\exp(2\pi ix) and ff is a sufficiently smooth function. In Titchmarsh [Tit86, Ch. V], it was shown that if f(x)f(x) has k3k\geq 3 continuous derivatives satisfying 0<λkf(k)(x)hλk0<\lambda_{k}\leq f^{(k)}(x)\leq h\lambda_{k}, then

Sf(a,N)A1h1/2k2Nλk1/(2k2)+A2N11/2k2λk1/(2k2)S_{f}(a,N)\leq A_{1}h^{1/2^{k-2}}N\lambda_{k}^{1/(2^{k}-2)}+A_{2}N^{1-1/2^{k-2}}\lambda_{k}^{-1/(2^{k}-2)} (1.8)

for some unspecified absolute constants A1A_{1} and A2A_{2}. This is also known as a kkth derivative test, and the method of derivation was to use van der Corput’s Ak2BA^{k-2}B process, where a single application of Poisson summation is followed by k2k-2 applications of the Weyl–van der Corput inequality. To our knowledge, to date the constants A1A_{1} and A2A_{2} in (1.8) have not been explicitly computed. However, Granville and Ramaré [GR96, Prop. 8.2] have proved an explicit bound of the form

Sf(a,N)k,hN(λk2k1/(2k2)+N1logk1(λk1))1/2k1S_{f}(a,N)\ll_{k,h}N\left(\lambda_{k}^{2^{k-1}/(2^{k}-2)}+N^{-1}\log^{k-1}(\lambda_{k}^{-1})\right)^{1/2^{k-1}} (1.9)

for k2k\geq 2. To facilitate a comparison, in our eventual application the value of λk\lambda_{k} is such that (1.8) and (1.9) respectively reduce to bounds of the form

SfN11/2k2andSfN11/2k2log(k1)/2k1N.S_{f}\ll N^{1-1/2^{k-2}}\qquad\text{and}\qquad S_{f}\ll N^{1-1/2^{k-2}}\log^{(k-1)/2^{k-1}}N.

Therefore, (1.8) produces a log-power saving (see also remarks after Lemma 2.5). More is known for small values of kk; see for instance [Bor12, Thm. 6.9] for k=2k=2 and Patel [Pat22] for k=3,4,5k=3,4,5. In §2 we derive a result of the form (1.8) and provide a comparison of our respective results.

The main challenge to proving (1.8) with reasonable constants, is that A1A_{1} and A2A_{2} tend to grow rapidly when applying the bound on an ill-suited summation interval, either because the resulting sum is too short, or because the phase function f(x)f(x) cannot be controlled properly on that domain. Meanwhile, an Ak2BA^{k-2}B process involves k2k-2 successive applications of Weyl-differencing, which for large kk limits our ability to isolate and properly address such pathological intervals. In our approach, we make progress through the repeated use of the trivial bound with each application of Weyl-differencing to avoid applying the kkth derivative test in intervals for which it is poorly suited.

1.3. Structure of this paper

In §2 we review the van der Corput method and construct explicit kkth-derivative tests corresponding to the Ak2B(0,1)A^{k-2}B(0,1) exponent pair. The results of this section are agnostic to the choice of phase function f(x)f(x). In §3 we specialise to a specific phase function to bound ζ(s)\zeta(s) on certain vertical lines inside the critical strip. Finally, in §4 we use the results of the previous two sections to prove Corollary 1.2.

2. An explicit kkth derivative test

The primary tool we use to bound ζ(s)\zeta(s) in the critical strip is an upper bound on the exponential sum Sf(a,N)S_{f}(a,N) (defined in (1.7)), where ff is a smooth function possessing at least k1k\geq 1 continuous derivatives. Four established methods of bounding Sf(a,N)S_{f}(a,N) are the Weyl–Hardy–Littlewood method, the van der Corput method, Vinogradov’s method and the Bombieri–Iwaniec method (for an exposition, see Titchmarsh [Tit86, Ch. V] and [BI86]). Here, we review relevant aspects of the van der Corput method in an explicit context. First, we have the trivial bound

Sf(a,N)N+1S_{f}(a,N)\leq N+1 (2.1)

arising from applying the triangle inequality and counting the maximum number of integers in (a,a+N](a,a+N]. If NN is an integer, we can improve this to Sf(a,N)NS_{f}(a,N)\leq N. The explicit Kuzmin–Landau lemma improves on the trivial bound if f(x)f(x) is sufficiently well-behaved. Let x\|x\| denote the distance to the nearest integer to xx. Suppose that f(x)f(x) is a real-valued function with a monotonic and continuous derivative on [a,a+N][a,a+N], satisfying f(x)λ1>0\|f^{\prime}(x)\|\geq\lambda_{1}>0. Then

Sf(a,N)2πλ1.S_{f}(a,N)\leq\frac{2}{\pi\lambda_{1}}. (2.2)

Proofs of this result can be found in [Lan28], [Hia16], [Pat22] and [HPY24]. See also [Cor21], [Kuz27], [HP49], [Tit86, p. 91], [GK91, p. 7] and the survey in [Rey20]. Note that the bound in (2.2) does not depend on the length of the summation interval (a,a+N](a,a+N]. In [KT50] and [HPY24], the following generalisation was proved. If ff^{\prime} is monotonic and continuous on (a,a+N](a,a+N], and l+λ1f(x)l+1μ1l+\lambda_{1}\leq f^{\prime}(x)\leq l+1-\mu_{1} for some integer ll and λ1,μ1>0\lambda_{1},\mu_{1}>0, then

Sf(a,N)λ11+μ11π.S_{f}(a,N)\leq\frac{\lambda_{1}^{-1}+\mu_{1}^{-1}}{\pi}. (2.3)

In practice, the conditions imposed on f(x)f^{\prime}(x) are rarely satisfied on the entire summation interval (a,a+N](a,a+N]. Instead, we divide the interval of summation into multiple subintervals and apply (2.3) within some of the intervals, and the trivial bound (2.1) in the remaining intervals. By appropriately choosing the locations of the subdivisions, we arrive at an explicit version of an inequality due to van der Corput (see also [Bor12, Thm. 6.9] for a similar explicit result).

Lemma 2.1 (Second-derivative test).

Suppose f(x)f(x) is real-valued and twice continuous differentiable on [a,a+N][a,a+N] for some integers a,Na,N, with f′′(x)f^{\prime\prime}(x) monotonic and satisfying

λ2|f′′(x)|hλ2,x[a,a+N],\lambda_{2}\leq|f^{\prime\prime}(x)|\leq h\lambda_{2},\qquad x\in[a,a+N],

for some λ2>0\lambda_{2}>0 and h>1h>1. Then,

Sf(a,N)4πNhλ21/2+Nhλ2+4πλ21/2.S_{f}(a,N)\leq\frac{4}{\sqrt{\pi}}Nh\lambda_{2}^{1/2}+Nh\lambda_{2}+\frac{4}{\sqrt{\pi}}\lambda_{2}^{-1/2}.
Proof.

We follow closely the argument in [HPY24, Lem. 2.5]. There are two notable differences. First, we implement a suggested refinement mentioned in the concluding remarks of [HPY24] to eliminate the constant term of 24π12-4\pi^{-1}. This slightly sharpens the bound and simplifies the arguments that follow. Second, we generalise the argument to a broader class of functions. In doing so we incur a penalty of h1/3h^{1/3} in the first two terms. We opt for this generalisation because in our eventual application (the kkth derivative test) it becomes increasingly difficult to leverage the benefits of specialising f(x)f(x) as kk grows.

Consider first the case if λ2>π/16\lambda_{2}>\pi/16. Then, using the trivial bound, since aa and NN are integers, and using h>1h>1,

Sf(a,N)N<4πNhλ21/2S_{f}(a,N)\leq N<\frac{4}{\sqrt{\pi}}Nh\lambda_{2}^{1/2} (2.4)

so the desired result is true for all λ2>π/16\lambda_{2}>\pi/16.111We can expand the range of λ2\lambda_{2} under consideration via the following argument, as remarked by Timothy S. Trudgian. For all λ2>1+4(24+π)/π=0.1439\lambda_{2}>1+4(2-\sqrt{4+\pi})/\pi=0.1439\ldots, we have (4/π)Nλ21/2+Nλ2>N(4/\sqrt{\pi})N\lambda_{2}^{1/2}+N\lambda_{2}>N so the desired result once again follows from the trivial bound, as h>1h>1. This allows us to assume a sharper upper bound on λ2\lambda_{2} in the subsequent argument. In the remainder of the proof we will assume that λ2π/16\lambda_{2}\leq\pi/16.

The conditions imposed on f(x)f(x) imply that either f′′(x)>0f^{\prime\prime}(x)>0 on [a,a+N][a,a+N] or f′′(x)<0f^{\prime\prime}(x)<0. Without loss of generality, assume that f′′(x)>0f^{\prime\prime}(x)>0, since we may replace ff with f-f without changing the value of Sf(a,N)S_{f}(a,N). Due to the continuity of f′′f^{\prime\prime}, there exists some ξ[a,a+N]\xi\in[a,a+N] for which

f(a+N)f(a)=Nf′′(ξ)Nhλ2.f^{\prime}(a+N)-f^{\prime}(a)=Nf^{\prime\prime}(\xi)\leq Nh\lambda_{2}. (2.5)

Meanwhile, define

C0:=f(a),Ck:=f(a+N),C_{0}:=\lfloor f^{\prime}(a)\rfloor,\qquad C_{k}:=\lfloor f^{\prime}(a+N)\rfloor, (2.6)

and let {f(a)}=ε1\{f^{\prime}(a)\}=\varepsilon_{1} and {f(b)}=ε2\{f^{\prime}(b)\}=\varepsilon_{2}, where {x}\{x\} denotes the fractional part of xx. Let 0<Δ<1/20<\Delta<1/2 be a parameter to be chosen later, and let

Cj\displaystyle C_{j} :=Cj1+1,\displaystyle:=C_{j-1}+1, 1jk1,\displaystyle 1\leq j\leq k-1,
xj\displaystyle x_{j} :=max{(f)1(CjΔ),a},\displaystyle:=\max\{(f^{\prime})^{-1}(C_{j}-\Delta),a\}, 1jk,\displaystyle 1\leq j\leq k, (2.7)
yj\displaystyle y_{j} :=min{(f)1(Cj+Δ),a+N},\displaystyle:=\min\{(f^{\prime})^{-1}(C_{j}+\Delta),a+N\}, 0jk.\displaystyle 0\leq j\leq k.

By (2.5), we have

k=CkC0=f(a+N)ε2(f(a)ε1)Nhλ2+ε1ε2,k=C_{k}-C_{0}=f^{\prime}(a+N)-\varepsilon_{2}-(f^{\prime}(a)-\varepsilon_{1})\leq Nh\lambda_{2}+\varepsilon_{1}-\varepsilon_{2}, (2.8)

Furthermore, since both ff^{\prime} and its inverse function (f)1(f^{\prime})^{-1} are increasing, for 1jk1\leq j\leq k we have

yjxj\displaystyle y_{j}-x_{j} =min{(f)1(Cj+Δ),a+N}max{(f)1(CjΔ),a}\displaystyle=\min\{(f^{\prime})^{-1}(C_{j}+\Delta),a+N\}-\max\{(f^{\prime})^{-1}(C_{j}-\Delta),a\}
2Δ|((f)1)(ξj)|,\displaystyle\leq 2\Delta|((f^{\prime})^{-1})^{\prime}(\xi_{j})|, (2.9)

for some ξj\xi_{j} satisfying

max{CjΔ,f(a)}ξjmin{Cj+Δ,f(a+N)}.\max\{C_{j}-\Delta,f^{\prime}(a)\}\leq\xi_{j}\leq\min\{C_{j}+\Delta,f^{\prime}(a+N)\}. (2.10)

Therefore, νj:=(f)1(ξj)[a,a+N]\nu_{j}:=(f^{\prime})^{-1}(\xi_{j})\in[a,a+N] and hence

yjxj2Δ|((f)1)(ξj)|=2Δ|f′′(νj)|2Δλ21.y_{j}-x_{j}\leq 2\Delta|((f^{\prime})^{-1})^{\prime}(\xi_{j})|=\frac{2\Delta}{|f^{\prime\prime}(\nu_{j})|}\leq 2\Delta\lambda_{2}^{-1}. (2.11)

Next, because Δ<1/2\Delta<1/2 by assumption, and since (f)1(f^{\prime})^{-1} is increasing, we have ax1<y1<x2<y2<<xk<yka+Na\leq x_{1}<y_{1}<x_{2}<y_{2}<\cdots<x_{k}<y_{k}\leq a+N. First, by the trivial bound and (2.11), we have for 1jk1\leq j\leq k,

Sf(xj,yjxj)yjxj+12Δλ21+1.S_{f}(x_{j},y_{j}-x_{j})\leq y_{j}-x_{j}+1\leq 2\Delta\lambda_{2}^{-1}+1. (2.12)

Next, in intervals of the form [yj,xj+1)[y_{j},x_{j+1}) for 1jk11\leq j\leq k-1, we have, by construction, f(x)Δ\|f^{\prime}(x)\|\geq\Delta. By Lemma 2.2, for 1jk11\leq j\leq k-1,

Sf(yj,xj+1yj)2πΔS_{f}(y_{j},x_{j+1}-y_{j})\leq\frac{2}{\pi\Delta} (2.13)

It remains to consider the boundary sums Sf(a,x1a)S_{f}(a,x_{1}-a) and Sf(yk,a+Nyk)S_{f}(y_{k},a+N-y_{k}). Here we use the same argument as [HPY24]. First, consider Sf(a,x1a)S_{f}(a,x_{1}-a). There are three cases.

Case 1: 0ε1<Δ0\leq\varepsilon_{1}<\Delta.

Then, y0=(f)1(f(a)+Δ)>(f)1(f(a))=ay_{0}=(f^{\prime})^{-1}(\lfloor f^{\prime}(a)\rfloor+\Delta)>(f^{\prime})^{-1}(f^{\prime}(a))=a, as (f)1(f^{\prime})^{-1} is increasing. We divide (a,x1](a,x_{1}] into (a,y0](a,y_{0}] and (y0,x1](y_{0},x_{1}]. In (a,y0](a,y_{0}] we use the trivial bound in a similar fashion as (2.11). In (y0,x1](y_{0},x_{1}] we have f(n)Δ\|f^{\prime}(n)\|\geq\Delta, so we use the Kuzmin–Landau lemma (2.2) to cover this subinterval. Together, we have

Sf(a,x1a)Sf(a,y1a)+Sf(y1,x1y1)(Δε1)λ21+1+2πΔ.S_{f}(a,x_{1}-a)\leq S_{f}(a,y_{1}-a)+S_{f}(y_{1},x_{1}-y_{1})\leq(\Delta-\varepsilon_{1})\lambda_{2}^{-1}+1+\frac{2}{\pi\Delta}. (2.14)

Case 2: Δε11Δ\Delta\leq\varepsilon_{1}\leq 1-\Delta.

We have y0ax1y_{0}\leq a\leq x_{1} and C0+ε1f(n)C0+1ΔC_{0}+\varepsilon_{1}\leq f^{\prime}(n)\leq C_{0}+1-\Delta for all n(a,x1]n\in(a,x_{1}]. By (2.3), we have

Sf(a,x1a)1π(1ε1+1Δ).S_{f}(a,x_{1}-a)\leq\frac{1}{\pi}\left(\frac{1}{\varepsilon_{1}}+\frac{1}{\Delta}\right). (2.15)

Case 3: 1Δ<ε111-\Delta<\varepsilon_{1}\leq 1.

Then f(a)=f(a)+ε1>C0+1Δ=C1Δf^{\prime}(a)=\lfloor f^{\prime}(a)\rfloor+\varepsilon_{1}>C_{0}+1-\Delta=C_{1}-\Delta. This implies (f)1(C1Δ)<a(f^{\prime})^{-1}(C_{1}-\Delta)<a. Hence x1=ax_{1}=a and thus Sf(a,x1a)=0S_{f}(a,x_{1}-a)=0 in this case.

Combining the three cases, we conclude that

Sf(a,x1a)1πΔ+HΔ(ε1),S_{f}(a,x_{1}-a)\leq\frac{1}{\pi\Delta}+H_{\Delta}(\varepsilon_{1}), (2.16)

where

HΔ(ε):={(Δε)λ21+1+1πΔif ε[0,Δ)1πεif ε[Δ,1Δ]1πΔif ε(1Δ,1]H_{\Delta}(\varepsilon):=\begin{cases}\displaystyle(\Delta-\varepsilon)\lambda_{2}^{-1}+1+\frac{1}{\pi\Delta}&\text{if }\varepsilon\in[0,\Delta)\\ \displaystyle\frac{1}{\pi\varepsilon}&\text{if }\varepsilon\in[\Delta,1-\Delta]\\ \displaystyle-\frac{1}{\pi\Delta}&\text{if }\varepsilon\in(1-\Delta,1]\end{cases} (2.17)

Via a similar argument, we have

Sf(yk,a+Nyk)1πΔ+HΔ(1ε2).S_{f}(y_{k},a+N-y_{k})\leq\frac{1}{\pi\Delta}+H_{\Delta}(1-\varepsilon_{2}). (2.18)

Combining (2.12), (2.13) and (2.8), we obtain that Sf(a,N)S_{f}(a,N) is majorised by

Sf(a,x1a)+j=1kSf(xj,yjxj)+j=1k1Sf(yj,xj+1yj)+Sf(yk,a+Nyk)1πΔ+HΔ(ε1)+k(2Δλ21+1)+(k1)2πΔ+1πΔ+HΔ(1ε2)2(Nhλ2+ε1ε2)(1πΔ+Δλ21+12)+HΔ(ε1)+HΔ(1ε2).\begin{split}&S_{f}(a,x_{1}-a)+\sum_{j=1}^{k}S_{f}(x_{j},y_{j}-x_{j})+\sum_{j=1}^{k-1}S_{f}(y_{j},x_{j+1}-y_{j})+S_{f}(y_{k},a+N-y_{k})\\ &\qquad\qquad\leq\frac{1}{\pi\Delta}+H_{\Delta}(\varepsilon_{1})+k\left(2\Delta\lambda_{2}^{-1}+1\right)+(k-1)\frac{2}{\pi\Delta}+\frac{1}{\pi\Delta}+H_{\Delta}(1-\varepsilon_{2})\\ &\qquad\qquad\leq 2\left(Nh\lambda_{2}+\varepsilon_{1}-\varepsilon_{2}\right)\left(\frac{1}{\pi\Delta}+\Delta\lambda_{2}^{-1}+\frac{1}{2}\right)+H_{\Delta}(\varepsilon_{1})+H_{\Delta}(1-\varepsilon_{2}).\end{split} (2.19)

We choose Δ=Δ0:=λ2/π\Delta=\Delta_{0}:=\sqrt{\lambda_{2}/\pi} to minimise the second factor, noting that the upper bound on λ2\lambda_{2} guarantees our choice satisfies the previous assumption that Δ<1/2\Delta<1/2. By (2.19), we have

Sf(a,N)4πNhλ21/2+Nhλ2+J(ε1)+J(1ε2)S_{f}(a,N)\leq\frac{4}{\sqrt{\pi}}Nh\lambda_{2}^{1/2}+Nh\lambda_{2}+J(\varepsilon_{1})+J(1-\varepsilon_{2}) (2.20)

where

J(ε)=HΔ0(ε)+(4πΔ0+1)(ε12).J(\varepsilon)=H_{\Delta_{0}}(\varepsilon)+\left(\frac{4}{\pi\Delta_{0}}+1\right)\left(\varepsilon-\frac{1}{2}\right). (2.21)

To complete the proof, it suffices to show that

J(ε)2πΔ0,0ε<1.J(\varepsilon)\leq\frac{2}{\pi\Delta_{0}},\qquad 0\leq\varepsilon<1. (2.22)

We once again consider three cases.

Case 1: ε[0,Δ0)\varepsilon\in\left[0,\Delta_{0}\right).

From (2.17) and (2.21), and using 0ε<Δ00\leq\varepsilon<\Delta_{0},

J(ε)<Δ0λ2+1+1πΔ0+(4πΔ0+1)(Δ012)=Δ0+4π+12<2πΔ0,\begin{split}J(\varepsilon)&<\frac{\Delta_{0}}{\lambda_{2}}+1+\frac{1}{\pi\Delta_{0}}+\left(\frac{4}{\pi\Delta_{0}}+1\right)\left(\Delta_{0}-\frac{1}{2}\right)=\Delta_{0}+\frac{4}{\pi}+\frac{1}{2}<\frac{2}{\pi\Delta_{0}},\end{split} (2.23)

where the last inequality holds since 0<Δ01/40<\Delta_{0}\leq 1/4.

Case 2: ε[Δ0,1Δ0]\varepsilon\in\left[\Delta_{0},1-\Delta_{0}\right].

Then,

J(ε)=1πε+(4πΔ0+1)ε1πΔ012J(\varepsilon)=\frac{1}{\pi\varepsilon}+\left(\frac{4}{\pi\Delta_{0}}+1\right)\varepsilon-\frac{1}{\pi\Delta_{0}}-\frac{1}{2} (2.24)

is a convex function, so J(ε)max{J(Δ0),J(1Δ0)}J(\varepsilon)\leq\max\left\{J(\Delta_{0}),J(1-\Delta_{0})\right\}. However, for any 0<Δ01/40<\Delta_{0}\leq 1/4 we have

J(Δ0)=4π12+Δ0<2πΔ0J(\Delta_{0})=\frac{4}{\pi}-\frac{1}{2}+\Delta_{0}<\frac{2}{\pi\Delta_{0}} (2.25)

and

J(1Δ0)=1π(1Δ0)+(1Δ0)+2πΔ04π12<2πΔ0.\begin{split}J(1-\Delta_{0})&=\frac{1}{\pi(1-\Delta_{0})}+(1-\Delta_{0})+\frac{2}{\pi\Delta_{0}}-\frac{4}{\pi}-\frac{1}{2}<\frac{2}{\pi\Delta_{0}}.\end{split} (2.26)

Case 3: ε(1Δ0,1)\varepsilon\in(1-\Delta_{0},1).

From ε<1\varepsilon<1 and Δ0<1/4\Delta_{0}<1/4,

J(ε)<1πΔ0+12(4πΔ0+1)<2πΔ0,J(\varepsilon)<-\frac{1}{\pi\Delta_{0}}+\frac{1}{2}\left(\frac{4}{\pi\Delta_{0}}+1\right)<\frac{2}{\pi\Delta_{0}}, (2.27)

as required. ∎

Remark.

A qualitatively similar result is historically obtained via Poisson summation (known as process BB, see e.g. [Tit86, Ch. V]). In our treatment we bypass Poisson summation altogether to obtain more favourable constants, while still achieving the goal of shortening the lengths of the exponential sums under consideration. In van der Corput notation, Lemma 2.1 corresponds to the B(0,1)B(0,1) exponent pair.

In our application it is convenient to have the second-derivative test to be of the same form as all higher derivative tests, which motivates the following lemma. This may be compared to [Bor12, Thm. 6.9], which has A2=4/πA_{2}=4/\sqrt{\pi} and B2=8/πB_{2}=8/\sqrt{\pi}.

Lemma 2.2.

Let f(x)f(x), aa, NN, hh and λ2\lambda_{2} satisfy the same conditions as Lemma 2.1. Then

Sf(a,N)A2Nhλ21/2+B2λ21/2,S_{f}(a,N)\leq A_{2}Nh\lambda_{2}^{1/2}+B_{2}\lambda_{2}^{-1/2},

where

A2:=2+4+ππ,B2:=4π.A_{2}:=\frac{2+\sqrt{4+\pi}}{\sqrt{\pi}},\qquad B_{2}:=\frac{4}{\sqrt{\pi}}. (2.28)
Proof.

Let λ0:=1+4(24+π)/π=0.1439\lambda_{0}:=1+4(2-\sqrt{4+\pi})/\pi=0.1439\ldots be the unique solution to

(4π+λ01/2)λ01/2=1.\left(\frac{4}{\sqrt{\pi}}+\lambda_{0}^{1/2}\right)\lambda_{0}^{1/2}=1. (2.29)

If λ2λ0\lambda_{2}\leq\lambda_{0}, then by Lemma 2.1,

Sf(a,N)4πNhλ21/2+Nhλ2+4πλ21/2(4π+λ01/2)Nhλ21/2+4πλ21/2.S_{f}(a,N)\leq\frac{4}{\sqrt{\pi}}Nh\lambda_{2}^{1/2}+Nh\lambda_{2}+\frac{4}{\sqrt{\pi}}\lambda_{2}^{-1/2}\leq\left(\frac{4}{\sqrt{\pi}}+\lambda_{0}^{1/2}\right)Nh\lambda_{2}^{1/2}+\frac{4}{\sqrt{\pi}}\lambda_{2}^{-1/2}.

On the other hand if λ2>λ0\lambda_{2}>\lambda_{0}, then by (2.29) and the trivial bound,

Sf(a,N)N<(4π+λ01/2)Nhλ01/2<(4π+λ01/2)Nhλ21/2+4πλ21/2,S_{f}(a,N)\leq N<\left(\frac{4}{\sqrt{\pi}}+\lambda_{0}^{1/2}\right)Nh\lambda_{0}^{1/2}<\left(\frac{4}{\sqrt{\pi}}+\lambda_{0}^{1/2}\right)Nh\lambda_{2}^{1/2}+\frac{4}{\sqrt{\pi}}\lambda_{2}^{-1/2},

hence the result follows in either case. ∎

To obtain higher-derivative tests, we use an explicit AA process, which makes use of the Weyl-differencing operation. Here, SfS_{f} is expressed in terms of SgS_{g} with g(x)=f(x+r)f(x)g(x)=f(x+r)-f(x) for some integer r>0r>0. Intuitively, if f(x)f(x) is well-approximated by a degree dd polynomial on (a,b](a,b], with d>0d>0, then we can expect that g(x)g(x) is well-approximated by a degree d1d-1 polynomial on (a,b](a,b]. We can achieve sharper bounds on SgS_{g} since the lower order of g(x)g(x) means it likely satisfies the conditions of (2.3) over longer intervals, hence increasing the savings produced by the bound.

Lemma 2.3 (Explicit AA process).

Let f(x)f(x) be real-valued and defined on (a,a+N](a,a+N], for some integers a,Na,N. For all integers q>0q>0, we have

(Sf(a,N))2(N1+q)(Nq+2qr=1q1(1rq)Sgr(a,Nr))(S_{f}(a,N))^{2}\leq\left(N-1+q\right)\left(\frac{N}{q}+\frac{2}{q}\sum_{r=1}^{q-1}\left(1-\frac{r}{q}\right)S_{g_{r}}(a,N-r)\right)

where gr(x):=f(x+r)f(x)g_{r}(x):=f(x+r)-f(x).

Proof.

See [CG04, Lem. 5] and [PT15, Lem. 2]. The first factor originally appears as N+1+qN+1+q in [CG04] (upon making the substitution aN+1a\mapsto N+1, NL1N\mapsto L-1). As remarked in [PT15], it is possible to reduce this factor to N+qN+q via a more careful bound. Here, we further decrease the factor by 1 by assuming that aa and NN are integers, as sums over n(a,a+N]n\in(a,a+N] are equivalent to sums over n[a+1,a+N]n\in[a+1,a+N]. ∎

By using Lemma 2.1 to estimate Sgr(a,Nr)S_{g_{r}}(a,N-r), we obtain the following estimate, which is historically obtained by applying the Poisson summation formula then applying the AA process. The next lemma corresponds to an explicit AB(0,1)AB(0,1) process.

Lemma 2.4 (Explicit third derivative test).

Let f(x)f(x) have three continuous derivatives and suppose f′′′f^{\prime\prime\prime} is monotonic and satisfies λ3|f′′′(x)|hλ3\lambda_{3}\leq|f^{\prime\prime\prime}(x)|\leq h\lambda_{3} for all x(a,a+N]x\in(a,a+N], where λ3>0\lambda_{3}>0, h>1h>1 and aa, NN are integers. Then, for any η3>0\eta_{3}>0, we have

Sf(a,N)A3(η3,h)h1/2Nλ31/6+B3(η3)N1/2λ31/6S_{f}(a,N)\leq A_{3}(\eta_{3},h)h^{1/2}N\lambda_{3}^{1/6}+B_{3}(\eta_{3})N^{1/2}\lambda_{3}^{-1/6}

where

A3:=1η3h+3215πη3+λ01/3+13(η3+λ01/3)λ01/3δ3,B3:=323π1/4η31/4δ3,A_{3}:=\sqrt{\frac{1}{\eta_{3}h}+\frac{32}{15\sqrt{\pi}}\sqrt{\eta_{3}+\lambda_{0}^{1/3}}+\frac{1}{3}\left(\eta_{3}+\lambda_{0}^{1/3}\right)\lambda_{0}^{1/3}}\delta_{3},\quad B_{3}:=\frac{\sqrt{32}}{\sqrt{3}\pi^{1/4}\eta_{3}^{1/4}}\delta_{3},
λ0:=(1η3+32η31/2h15π)3,δ3:=12+121+38π1/2η33/2.\lambda_{0}:=\left(\frac{1}{\eta_{3}}+\frac{32\eta_{3}^{1/2}h}{15\sqrt{\pi}}\right)^{-3},\qquad\delta_{3}:=\sqrt{\frac{1}{2}+\frac{1}{2}\sqrt{1+\frac{3}{8}\pi^{1/2}\eta_{3}^{3/2}}}.
Proof.

We will first consider the case if λ3λ0\lambda_{3}\geq\lambda_{0}. Using the trivial bound, since aa and NN are integers, we have

Sf(a,N)N=1η3+32η31/2h15πNλ01/6\displaystyle S_{f}(a,N)\leq N=\sqrt{\frac{1}{\eta_{3}}+\frac{32\eta_{3}^{1/2}h}{15\sqrt{\pi}}}N\lambda_{0}^{1/6} 1η3h+32η31/215πh1/2Nλ31/6\displaystyle\leq\sqrt{\frac{1}{\eta_{3}h}+\frac{32\eta_{3}^{1/2}}{15\sqrt{\pi}}}h^{1/2}N\lambda_{3}^{1/6}
<A3h1/2Nλ31/6,\displaystyle<A_{3}h^{1/2}N\lambda_{3}^{1/6}, (2.30)

and hence the desired result follows if λ3λ0\lambda_{3}\geq\lambda_{0}. In the last inequality we have used δ3>1\delta_{3}>1 and λ0>0\lambda_{0}>0.

Next, consider the case of λ3λ1\lambda_{3}\leq\lambda_{1}, where

λ1:=(η3κN)3,κ:=121+38π1/2η33/212.\lambda_{1}:=\left(\frac{\eta_{3}}{\kappa N}\right)^{3},\qquad\kappa:=\frac{1}{2}\sqrt{1+\frac{3}{8}\pi^{1/2}\eta_{3}^{3/2}}-\frac{1}{2}. (2.31)

Note that κ\kappa satisfies

32κ(1+κ)3π1/2η33/2=1andδ3=1+κ.\frac{32\kappa(1+\kappa)}{3\pi^{1/2}\eta_{3}^{3/2}}=1\qquad\text{and}\qquad\delta_{3}=\sqrt{1+\kappa}. (2.32)

We once again apply the trivial bound to obtain

Sf(a,N)N\displaystyle S_{f}(a,N)\leq N =(1+κ)1/2323π1/4η33/4Nκ1/2\displaystyle=(1+\kappa)^{1/2}\cdot\frac{\sqrt{32}}{\sqrt{3}\pi^{1/4}\eta_{3}^{3/4}}N\kappa^{1/2}
=δ3323π1/4η31/4N1/2λ11/6B3N1/2λ31/6,\displaystyle=\delta_{3}\frac{\sqrt{32}}{\sqrt{3}\pi^{1/4}\eta_{3}^{1/4}}N^{1/2}\lambda_{1}^{-1/6}\leq B_{3}N^{1/2}\lambda_{3}^{-1/6}, (2.33)

hence the desired result follows in this case too.

Suppose now that λ1<λ3<λ0\lambda_{1}<\lambda_{3}<\lambda_{0} (we may assume that λ1<λ0\lambda_{1}<\lambda_{0}, since otherwise the proof is complete). Let gr(x)=f(x+r)f(x)g_{r}(x)=f(x+r)-f(x) as in Lemma 2.3. By the mean value theorem, since f′′′f^{\prime\prime\prime} is continuous we have gr′′(x)=f′′(x+r)f′′(x)=rf′′′(ξ)g_{r}^{\prime\prime}(x)=f^{\prime\prime}(x+r)-f^{\prime\prime}(x)=rf^{\prime\prime\prime}(\xi) for some ξ[x,x+r][a,a+N]\xi\in[x,x+r]\subseteq[a,a+N]. Therefore, by the lemma’s assumption we have

rλ3|gr′′(x)|hrλ3,x[a,a+Nr]r\lambda_{3}\leq|g_{r}^{\prime\prime}(x)|\leq h\cdot r\lambda_{3},\qquad x\in[a,a+N-r] (2.34)

and hence we may apply Lemma 2.1 with f=grf=g_{r} and λ2=rλ3\lambda_{2}=r\lambda_{3}, since both aa and a+Nra+N-r are integers. This gives

Sgr(a,br)4π(Nr)h(rλ3)1/2+(Nr)h(rλ3)+4π(rλ3)1/2.S_{g_{r}}(a,b-r)\leq\frac{4}{\sqrt{\pi}}(N-r)h(r\lambda_{3})^{1/2}+(N-r)h(r\lambda_{3})+\frac{4}{\sqrt{\pi}}(r\lambda_{3})^{-1/2}. (2.35)

We apply the inequality

r=1q(1rq)rsq1+s(1+s)(2+s),1<s1,\sum_{r=1}^{q}\left(1-\frac{r}{q}\right)r^{s}\leq\frac{q^{1+s}}{(1+s)(2+s)},\qquad-1<s\leq 1, (2.36)

(see e.g. [Pat22, (32)] for 1<s<1-1<s<1 and [Hia16, (45)] for s=1s=1) to obtain, after using Nr<NN-r<N,

2qr=1q1(1rq)Sgr(a,Nr)3215πhNq1/2λ31/2+13hNqλ3+323πq1/2λ31/2.\frac{2}{q}\sum_{r=1}^{q-1}\left(1-\frac{r}{q}\right)S_{g_{r}}(a,N-r)\leq\frac{32}{15\sqrt{\pi}}hNq^{1/2}\lambda_{3}^{1/2}+\frac{1}{3}hNq\lambda_{3}+\frac{32}{3\sqrt{\pi}}q^{-1/2}\lambda_{3}^{-1/2}. (2.37)

We choose q=η3λ31/3q=\lceil\eta_{3}\lambda_{3}^{-1/3}\rceil for some η3>0\eta_{3}>0, so that η3λ31/3qη3λ31/3+1\eta_{3}\lambda_{3}^{-1/3}\leq q\leq\eta_{3}\lambda_{3}^{-1/3}+1 and the RHS is

32hN15πη3λ31/3+1λ31/2+hN3(η3λ31/3+1)λ3+323π(η3λ31/3)1/2λ31/2\displaystyle\leq\frac{32hN}{15\sqrt{\pi}}\sqrt{\eta_{3}\lambda_{3}^{-1/3}+1}\lambda_{3}^{1/2}+\frac{hN}{3}(\eta_{3}\lambda_{3}^{-1/3}+1)\lambda_{3}+\frac{32}{3\sqrt{\pi}}\left(\eta_{3}\lambda_{3}^{-1/3}\right)^{-1/2}\lambda_{3}^{-1/2}
=hNλ31/3(3215πη3+λ31/3+13(η3+λ31/3)λ31/3)+323πη3λ31/3.\displaystyle=hN\lambda_{3}^{1/3}\left(\frac{32}{15\sqrt{\pi}}\sqrt{\eta_{3}+\lambda_{3}^{1/3}}+\frac{1}{3}\left(\eta_{3}+\lambda_{3}^{1/3}\right)\lambda_{3}^{1/3}\right)+\frac{32}{3\sqrt{\pi\eta_{3}}}\lambda_{3}^{-1/3}. (2.38)

Substituting this estimate into Lemma 2.3,

(Sf(a,N))2\displaystyle(S_{f}(a,N))^{2} (N+q1)(Nq+2qr=1q1(1rq)Sgr(a,Nr))\displaystyle\leq\left(N+q-1\right)\left(\frac{N}{q}+\frac{2}{q}\sum_{r=1}^{q-1}\left(1-\frac{r}{q}\right)S_{g_{r}}(a,N-r)\right) (2.39)
(N+η3λ31/3)(Nη3λ31/3+323πη3λ31/3\displaystyle\leq\left(N+\eta_{3}\lambda_{3}^{-1/3}\right)\Bigg{(}\frac{N}{\eta_{3}\lambda_{3}^{-1/3}}+\frac{32}{3\sqrt{\pi\eta_{3}}}\lambda_{3}^{-1/3}
+hNλ31/3(3215πη3+λ31/3+13(η3+λ31/3)λ31/3))\displaystyle\qquad\qquad+hN\lambda_{3}^{1/3}\left(\frac{32}{15\sqrt{\pi}}\sqrt{\eta_{3}+\lambda_{3}^{1/3}}+\frac{1}{3}\left(\eta_{3}+\lambda_{3}^{1/3}\right)\lambda_{3}^{1/3}\right)\Bigg{)} (2.40)
<(1+η3λ31/3N)((A3δ3)2hN2λ31/3+(B3δ3)2Nλ31/3),\displaystyle<\left(1+\frac{\eta_{3}\lambda_{3}^{-1/3}}{N}\right)\left(\left(\frac{A_{3}}{\delta_{3}}\right)^{2}hN^{2}\lambda_{3}^{1/3}+\left(\frac{B_{3}}{\delta_{3}}\right)^{2}N\lambda_{3}^{-1/3}\right), (2.41)

where the last inequality follows from the assumption that λ3<λ0\lambda_{3}<\lambda_{0}. Applying the assumption λ3>λ1\lambda_{3}>\lambda_{1} to the first factor, then taking square roots,

Sf(a,N)\displaystyle S_{f}(a,N) (1+η3λ11/3N)((A3δ3)2hN2λ31/3+(B3δ3)2Nλ31/3)\displaystyle\leq\sqrt{\left(1+\frac{\eta_{3}\lambda_{1}^{-1/3}}{N}\right)\left(\left(\frac{A_{3}}{\delta_{3}}\right)^{2}hN^{2}\lambda_{3}^{1/3}+\left(\frac{B_{3}}{\delta_{3}}\right)^{2}N\lambda_{3}^{-1/3}\right)} (2.42)
1+κA3δ3h1/2Nλ31/6+1+κB3δ3N1/2λ31/6,\displaystyle\leq\sqrt{1+\kappa}\frac{A_{3}}{\delta_{3}}h^{1/2}N\lambda_{3}^{1/6}+\sqrt{1+\kappa}\frac{B_{3}}{\delta_{3}}N^{1/2}\lambda_{3}^{-1/6}, (2.43)

since x+yx+y\sqrt{x+y}\leq\sqrt{x}+\sqrt{y} for all x,y0x,y\geq 0. However δ3=1+κ\delta_{3}=\sqrt{1+\kappa} so the result follows. ∎

If we now use Lemma 2.4 instead of Lemma 2.1 to bound Sgr(a,Nr)S_{g_{r}}(a,N-r) in Lemma 2.3, then we obtain the fourth-derivative test, via the A2B(0,1)A^{2}B(0,1) process. Performing this substitution recursively as necessary, it is possible to derive an explicit Ak2B(0,1)A^{k-2}B(0,1) process, for any k2k\geq 2. The following lemma is our main result.

Lemma 2.5 (Explicit kkth derivative test).

Let a,Na,N be integers with N>0N>0. Let f(x)f(x) be equipped with k4k\geq 4 continuous derivatives, with f(k)f^{(k)} monotonic, and suppose that 0<λk|f(k)(x)|hλk0<\lambda_{k}\leq|f^{(k)}(x)|\leq h\lambda_{k} for all x(a,a+N]x\in(a,a+N] and some h>1h>1. Then

Sf(a,N)Akh2/KNλk1/(2K2)+BkN12/Kλk1/(2K2)S_{f}(a,N)\leq A_{k}h^{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}, A3(η3,h)A_{3}(\eta_{3},h) and B3(η3)B_{3}(\eta_{3}) are defined in Lemma 2.4, and Ak,BkA_{k},B_{k} for k4k\geq 4 are defined recursively via

Aj+1(η3,h):=δj(h1/J+219/12(J1)(2J1)(4J3)Aj(η3,h)1/2),A_{j+1}(\eta_{3},h):=\delta_{j}\left(h^{-1/J}+\frac{2^{19/12}(J-1)}{\sqrt{(2J-1)(4J-3)}}A_{j}(\eta_{3},h)^{1/2}\right), (2.44)
Bj+1(η3):=δj23/2(J1)(2J3)(4J5)Bj(η3)1/2,B_{j+1}(\eta_{3}):=\delta_{j}\frac{2^{3/2}(J-1)}{\sqrt{(2J-3)(4J-5)}}B_{j}(\eta_{3})^{1/2}, (2.45)
δj:=1+2233712/J(9π1024η3)1/J,\delta_{j}:=\sqrt{1+\frac{2}{2337^{1-2/J}}\left(\frac{9\pi}{1024}\eta_{3}\right)^{1/J}}, (2.46)

where J:=2j1J:=2^{j-1}.

Proof.

The central argument remains the same as that of Lemma 2.4, however to achieve a bound that holds uniformly for all k4k\geq 4 without excessive tedium, we forsake sharpness in several key inequalities. One motivation for the separate treatment of the k=3k=3 case is to establish good starting constants A3A_{3} and B3B_{3} which feed into all higher derivative tests. An immediate avenue for further refinement is therefore to extend the argument of Lemma 2.4 to higher kk before switching to the general argument.

As before, we begin by considering a few edge cases. First, suppose N2336N\leq 2336. Since δk>1\delta_{k}>1, for all k3k\geq 3 we have

Ak+1Bk+1δk2237/12(K1)2(2K1)(2K3)(4K3)(4K5)Ak1/2Bk1/2>21/12Ak1/2Bk1/2A_{k+1}B_{k+1}\geq\delta_{k}^{2}\frac{2^{37/12}(K-1)^{2}}{\sqrt{(2K-1)(2K-3)(4K-3)(4K-5)}}A_{k}^{1/2}B_{k}^{1/2}>2^{1/12}A_{k}^{1/2}B_{k}^{1/2}

which implies that

AkBk>21/62/(3K)(A3B3)4/K.A_{k}B_{k}>2^{1/6-2/(3K)}(A_{3}B_{3})^{4/K}. (2.47)

Therefore, from the arithmetic-geometric means inequality,

Akh2/KNλk1/(2K2)+BkN12/Kλk1/(2K2)2(AkBk)1/2h1/KN11/K>213/121/(3K)(A3B3)2/Kh1/KN11/K.\begin{split}&A_{k}h^{2/K}N\lambda_{k}^{1/(2K-2)}+B_{k}N^{1-2/K}\lambda_{k}^{-1/(2K-2)}\geq 2(A_{k}B_{k})^{1/2}h^{1/K}N^{1-1/K}\\ &\qquad\qquad\qquad>2^{13/12-1/(3K)}(A_{3}B_{3})^{2/K}h^{1/K}N^{1-1/K}.\end{split} (2.48)

The last expression is no smaller than NN if 213/121/(3K)(A3B3)2/Kh1/KN1/K2^{13/12-1/(3K)}(A_{3}B_{3})^{2/K}h^{1/K}\geq N^{1/K}, which is true since

A3B3(32η31/215π)1/2323π1/4η31/4=3235π,A_{3}B_{3}\geq\left(\frac{32\eta_{3}^{1/2}}{15\sqrt{\pi}}\right)^{1/2}\frac{\sqrt{32}}{\sqrt{3}\pi^{1/4}\eta_{3}^{1/4}}=\frac{32}{3\sqrt{5\pi}}, (2.49)

and, since h>1h>1, k4k\geq 4,

213K/121/3(A3B3)2h>225/3(3235π)2>2336N.2^{13K/12-1/3}(A_{3}B_{3})^{2}h>2^{25/3}\cdot\left(\frac{32}{3\sqrt{5\pi}}\right)^{2}>2336\geq N. (2.50)

Therefore, the desired result follows from the trivial bound.

Next, suppose N2337N\geq 2337 and λkλ0(k)\lambda_{k}\leq\lambda_{0}(k), where

λ0(k):=(9π1024η3)2+2/KN4+4/K.\lambda_{0}(k):=\left(\frac{9\pi}{1024}\eta_{3}\right)^{-2+2/K}N^{-4+4/K}. (2.51)

Since δk>1\delta_{k}>1, we have Bk>Bk11/2B_{k}>B_{k-1}^{1/2} for all k4k\geq 4, hence Bk>B34/KB_{k}>B_{3}^{4/K}. Thus

BkN12/Kλk1/(2K2)B34/KN12/Kλ01/(2K2)=(323π1/4η31/4)4/K[(9π1024η3)2+2/KN4+4/K]1/(2K2)N12/K=N.\begin{split}&B_{k}N^{1-2/K}\lambda_{k}^{-1/(2K-2)}\geq B_{3}^{4/K}N^{1-2/K}\lambda_{0}^{-1/(2K-2)}\\ &\qquad\qquad=\left(\frac{\sqrt{32}}{\sqrt{3}\pi^{1/4}\eta_{3}^{1/4}}\right)^{4/K}\left[\left(\frac{9\pi}{1024}\eta_{3}\right)^{-2+2/K}N^{-4+4/K}\right]^{-1/(2K-2)}N^{1-2/K}\\ &\qquad\qquad=N.\end{split} (2.52)

Hence the desired result once again follows from the trivial bound.

We now proceed to the main argument, assuming that N2337N\geq 2337 and λkλ0(k)\lambda_{k}\geq\lambda_{0}(k) for all k4k\geq 4. By Lemma 2.4, the desired result holds for k=3k=3. Assume for an induction that the lemma holds for some j3j\geq 3. For convenience denote J:=2j1J:=2^{j-1}.

Suppose that f(x)f(x) has j+1j+1 continuous derivatives on [a,b][a,b], such that λj+1|f(j+1)(x)|hλj+1\lambda_{j+1}\leq|f^{(j+1)}(x)|\leq h\lambda_{j+1} for some λj+1>0\lambda_{j+1}>0. Let gr(x):=f(x+r)f(x)g_{r}(x):=f(x+r)-f(x) so that, via the mean value theorem,

gr(j)(x)=f(j)(x+r)f(j)(x)=rf(j+1)(ξ),g_{r}^{(j)}(x)=f^{(j)}(x+r)-f^{(j)}(x)=rf^{(j+1)}(\xi), (2.53)

for all r1r\geq 1, x(a,a+Nr]x\in(a,a+N-r] and some ξ[x,x+r]\xi\in[x,x+r]. From the conditions on f(j+1)(x)f^{(j+1)}(x), we have

rλj+1|gr(j)(x)|rhλj+1.r\lambda_{j+1}\leq|g_{r}^{(j)}(x)|\leq rh\lambda_{j+1}. (2.54)

By the inductive assumption, we have

Sgr(a,Nr)Ajh2/JN(rλj+1)1/(2J2)+BjN12/J(rλj+1)1/(2J2),S_{g_{r}}(a,N-r)\leq A_{j}h^{2/J}N(r\lambda_{j+1})^{1/(2J-2)}+B_{j}N^{1-2/J}(r\lambda_{j+1})^{-1/(2J-2)}, (2.55)

for some constants AjA_{j}, BjB_{j} not depending on rr or λj+1\lambda_{j+1}. Note we have used the inequality Nr<NN-r<N for simplicity. We once again apply (2.36) to obtain

r=1q1(1rq)r1/(2J2)αjq1+1/(2J2),r=1q1(1rq)r1/(2J2)βjq11/(2J2)\sum_{r=1}^{q-1}\left(1-\frac{r}{q}\right)r^{1/(2J-2)}\leq\alpha_{j}q^{1+1/(2J-2)},\quad\sum_{r=1}^{q-1}\left(1-\frac{r}{q}\right)r^{-1/(2J-2)}\leq\beta_{j}q^{1-1/(2J-2)}

where

αj:=4(J1)2(2J1)(4J3),βj:=4(J1)2(2J3)(4J5).\alpha_{j}:=\frac{4(J-1)^{2}}{(2J-1)(4J-3)},\qquad\beta_{j}:=\frac{4(J-1)^{2}}{(2J-3)(4J-5)}. (2.56)

Substituting into (2.55), we obtain

2qr=1q1(1rq)Sgr(a,Nr)2αjAjh2/JNλj+11/(2J2)q1/(2J2)+2βjBjN12/Jλj+11/(2J2)q1/(2J2).\begin{split}\frac{2}{q}\sum_{r=1}^{q-1}\left(1-\frac{r}{q}\right)S_{g_{r}}(a,N-r)&\leq 2\alpha_{j}A_{j}h^{2/J}N\lambda_{j+1}^{1/(2J-2)}q^{1/(2J-2)}\\ &+2\beta_{j}B_{j}N^{1-2/J}\lambda_{j+1}^{-1/(2J-2)}q^{-1/(2J-2)}.\end{split} (2.57)

Using Lemma 2.3 and the inequality x+yx+y\sqrt{x+y}\leq\sqrt{x}+\sqrt{y},

Sf(a,N)\displaystyle S_{f}(a,N) DjNq1/2+DjN1/2[(2αjAj)1/2h2/JN(qλj+1)1/(2J2)\displaystyle\leq\frac{D_{j}N}{q^{1/2}}+D_{j}N^{1/2}\Bigg{[}(2\alpha_{j}A_{j})^{1/2}\sqrt{h^{2/J}N(q\lambda_{j+1})^{1/(2J-2)}}
+(2βjBj)1/2N12/J(qλj+1)1/(2J2)]\displaystyle\qquad\qquad+(2\beta_{j}B_{j})^{1/2}\sqrt{N^{1-2/J}(q\lambda_{j+1})^{-1/(2J-2)}}\Bigg{]} (2.58)
=DjNq1/2+Dj(2αjAj)1/2h1/JN(qλj+1)1/(4J4)\displaystyle=\frac{D_{j}N}{q^{1/2}}+D_{j}(2\alpha_{j}A_{j})^{1/2}h^{1/J}N(q\lambda_{j+1})^{1/(4J-4)}
+Dj(2βjBj)1/2N11/J(qλj+1)1/(4J4)\displaystyle\qquad\qquad+D_{j}(2\beta_{j}B_{j})^{1/2}N^{1-1/J}(q\lambda_{j+1})^{-1/(4J-4)} (2.59)

where

Dj:=1+qN.D_{j}:=\sqrt{1+\frac{q}{N}}. (2.60)

We choose q=λj+11/(2J1)+1q=\left\lfloor\lambda_{j+1}^{-1/(2J-1)}\right\rfloor+1, so that we have (wastefully)

λj+11/(2J1)<q2λj+11/(2J1).\lambda_{j+1}^{-1/(2J-1)}<q\leq 2\lambda_{j+1}^{-1/(2J-1)}. (2.61)

This choice of qq gives, via the assumption λj+1λ0(j+1)\lambda_{j+1}\geq\lambda_{0}(j+1) and N2337N\geq 2337,

Dj1+2N((9π1024η3)2+1/JN4+2/J)1/(2J1)δj.D_{j}\leq\sqrt{1+\frac{2}{N}\left(\left(\frac{9\pi}{1024}\eta_{3}\right)^{-2+1/J}N^{-4+2/J}\right)^{-1/(2J-1)}}\leq\delta_{j}. (2.62)

Additionally, since j3j\geq 3, we also have

14J4112,14J4(112J1)=14J2.\frac{1}{4J-4}\leq\frac{1}{12},\qquad\frac{1}{4J-4}\left(1-\frac{1}{2J-1}\right)=\frac{1}{4J-2}. (2.63)

This gives the following estimates.

(qλj+1)1/(4J4)21/(4J4)λj+1[11/(2J1)]/(4J4)<21/12λj+11/(4J2),(q\lambda_{j+1})^{1/(4J-4)}\leq 2^{1/(4J-4)}\lambda_{j+1}^{\left[1-1/(2J-1)\right]/(4J-4)}<2^{1/12}\lambda_{j+1}^{1/(4J-2)}, (2.64)
(qλj+1)1/(4J4)<λj+11/(4J2),(q\lambda_{j+1})^{-1/(4J-4)}<\lambda_{j+1}^{-1/(4J-2)}, (2.65)
q1/2<λj+11/(4J2),q^{-1/2}<\lambda_{j+1}^{1/(4J-2)}, (2.66)
(qλj+1)1/(2J2)λj+1(1/(2J1)1)/(2J2)=λj+11/(2J1).(q\lambda_{j+1})^{-1/(2J-2)}\leq\lambda_{j+1}^{(1/(2J-1)-1)/(2J-2)}=\lambda_{j+1}^{-1/(2J-1)}. (2.67)

Combining these with (2.62) and (2.59), we have

Sf(a,N)\displaystyle S_{f}(a,N) δjNλj+11/(4J2)+27/12δj(αjAj)1/2h1/JNλj+11/(4J2)\displaystyle\leq\delta_{j}N\lambda_{j+1}^{1/(4J-2)}+2^{7/12}\delta_{j}(\alpha_{j}A_{j})^{1/2}h^{1/J}N\lambda_{j+1}^{1/(4J-2)}
+δj(2βjBj)1/2N11/Jλj+11/(4J2)\displaystyle\qquad\qquad+\delta_{j}(2\beta_{j}B_{j})^{1/2}N^{1-1/J}\lambda_{j+1}^{-1/(4J-2)} (2.68)
=δj[h1/J+27/12(αjAj)1/2]h1/JNλj+11/(4J2)\displaystyle=\delta_{j}\left[h^{-1/J}+2^{7/12}(\alpha_{j}A_{j})^{1/2}\right]h^{1/J}N\lambda_{j+1}^{1/(4J-2)}
+δj(2βjBj)1/2N11/Jλj+11/(4J2)\displaystyle\qquad\qquad+\delta_{j}(2\beta_{j}B_{j})^{1/2}N^{1-1/J}\lambda_{j+1}^{-1/(4J-2)} (2.69)
=Aj+1h1/JNλj+11/(4J2)+Bj+1N11/Jλj+11/(4J2),\displaystyle=A_{j+1}h^{1/J}N\lambda_{j+1}^{1/(4J-2)}+B_{j+1}N^{1-1/J}\lambda_{j+1}^{-1/(4J-2)}, (2.70)

hence the induction is complete. ∎

Remark.

In our eventual application we have λktNk\lambda_{k}\asymp tN^{-k} where tK/((k+1)K2K+1)NtK/(kK2K+2)t^{K/((k+1)K-2K+1)}\ll N\ll t^{K/(kK-2K+2)}, so that in particular,

N2+2/KλkN1+1/K.N^{-2+2/K}\ll\lambda_{k}\ll N^{-1+1/K}.

Therefore, Lemma 2.5 implies (for fixed hh) that Sf(a,N)N11/(2K)=N11/2k1S_{f}(a,N)\ll N^{1-1/(2K)}=N^{1-1/2^{k-1}}. Meanwhile, logλklogN\log\lambda_{k}\asymp\log N so (1.9) reduces to Sf(a,N)N11/2k1log(k1)/2k1NS_{f}(a,N)\ll N^{1-1/2^{k-1}}\log^{(k-1)/2^{k-1}}N. In particular, in our application we take k=2k=2 when t2/3Ntt^{2/3}\ll N\ll t. Therefore, using (1.9) in place of Lemma 2.5 in the argument of Theorem 1.1 produces a bound of strength ζ(σk+it)t1/(2K2)log3/2t\zeta(\sigma_{k}+it)\ll t^{1/(2K-2)}\log^{3/2}t.

For many applications we are interested in uniform bounds holding for all kk0k\geq k_{0}. To this end we provide the following completely explicit result.

Lemma 2.6.

Let k10k\geq 10, a,N>0a,N>0 be integers. Suppose f(x)f(x) is any function having kk continuous derivatives with f(k)f^{(k)} monotonic, and λk|f(k)(x)|hλk\lambda_{k}\leq|f^{(k)}(x)|\leq h\lambda_{k} for all x(a,a+N]x\in(a,a+N] and some λk>0\lambda_{k}>0, 1<h31<h\leq 3. Then

Sf(a,N)2.762h2/KNλk1/(2K2)+1.02N12/Kλk1/(2K2).S_{f}(a,N)\leq 2.762\,h^{2/K}N\lambda_{k}^{1/(2K-2)}+1.02N^{1-2/K}\lambda_{k}^{-1/(2K-2)}.
Proof.

From Lemma 2.4 and (2.44), we observe that Ak(η3,h)A_{k}(\eta_{3},h) is a decreasing function of hh, for all k3k\geq 3. It thus suffices to bound AkA_{k} with h=3h=3. We choose η3=4.7399\eta_{3}=4.7399 in Lemma 2.4 and recursively compute AkA_{k}, BkB_{k} for 4k104\leq k\leq 10, using Lemma 2.5, to ultimately obtain

A10(4.7399,3)2.744,B10(4.7399)1.020.A_{10}(4.7399,3)\leq 2.744,\qquad B_{10}(4.7399)\leq 1.020. (2.71)

For k10k\geq 10, we note that δk\delta_{k} is decreasing in kk, so δkδ10\delta_{k}\leq\delta_{10}. Also, for any K>1K>1 we have

219/12(K1)(2K1)(4K3)=219/12(2+1K1)(4+1K1)<21/12.\frac{2^{19/12}(K-1)}{\sqrt{(2K-1)(4K-3)}}=\frac{2^{19/12}}{\sqrt{\left(2+\frac{1}{K-1}\right)\left(4+\frac{1}{K-1}\right)}}<2^{1/12}. (2.72)

By Lemma 2.5,

Ak+1δ10(1+21/12Ak1/2),k10.A_{k+1}\leq\delta_{10}\left(1+2^{1/12}A_{k}^{1/2}\right),\qquad k\geq 10. (2.73)

The discrete map xn+1=δ10(1+21/12xn1/2)x_{n+1}=\delta_{10}\left(1+2^{1/12}x_{n}^{1/2}\right) has a single stable fixed point

x=(211/12δ10+211/6δ102+δ10)22.762.x^{*}=\left(2^{-11/12}\delta_{10}+\sqrt{2^{-11/6}\delta_{10}^{2}+\delta_{10}}\right)^{2}\leq 2.762. (2.74)

Since Ak+1xk+1A_{k+1}\leq x_{k+1} and A10xA_{10}\leq x^{*}, we have Ak2.762A_{k}\leq 2.762 for all k10k\geq 10. As for BkB_{k}, for all k10k\geq 10,

δk23/2(K1)(2K3)(4K5)1.002.\delta_{k}\frac{2^{3/2}(K-1)}{\sqrt{(2K-3)(4K-5)}}\leq 1.002. (2.75)

The map yn+1=1.002yn1/2y_{n+1}=1.002y_{n}^{1/2} has a single stable fixed point y=1.0022<1.02y^{*}=1.002^{2}<1.02, so it follows from (2.71) that Bk1.02B_{k}\leq 1.02 for all k10k\geq 10.

3. Bounds on ζ(s)\zeta(s) in the critical strip

In this section we use the explicit kkth derivative test to bound ζ(σ+it)\zeta(\sigma+it) for certain values of σ\sigma, specifically those corresponding to

σk:=1k2k2\sigma_{k}:=1-\frac{k}{2^{k}-2} (3.1)

for integers k4k\geq 4. Such values of σ\sigma lie in the interval (1/2,1)(1/2,1), so that we are bounding ζ(s)\zeta(s) along vertical lines residing between the half-line and the 1-line. Such bounds can be used to develop explicit zero-free regions through the method of Ford [For02a], which rely primarily on sharp bounds on ζ(s)\zeta(s) slightly to the left of σ=1\sigma=1. Using the kkth derivative test, we can establish asymptotically sharper bounds on ζ(s)\zeta(s) than what is possible by considering bounds on the half-line and convexity principle alone [Tit86].

Throughout this section, we specialise the phase function f(x)f(x) encountered in lemmas 2.1, 2.4 and 2.5 to

f(x):=t2πlogx,f(x):=-\frac{t}{2\pi}\log x, (3.2)

so that

Sf(a,N)=|a<na+Nnit|.S_{f}(a,N)=\left|\sum_{a<n\leq a+N}n^{-it}\right|. (3.3)

Our proof of Theorem 1.1 is divided into two sections. First, in §3.1 we prove the theorem for tTkt\leq T_{k} where

Tk:=exp(2.6134(2k11)+2.8876kk3),k4.T_{k}:=\exp\left(\frac{2.6134(2^{k-1}-1)+2.8876k}{k-3}\right),\qquad k\geq 4. (3.4)

The main tools in this range are the Phragmén–Lindelöf Principle combined with the following two bounds on σ=1/2\sigma=1/2 and σ=1\sigma=1 respectively:

|ζ(1/2+it)|\displaystyle|\zeta(1/2+it)| 0.618t1/6logt,t3,\displaystyle\leq 0.618t^{1/6}\log t,\qquad t\geq 3, (3.5)
|ζ(1+it)|\displaystyle|\zeta(1+it)| logt,t3.\displaystyle\leq\log t,\qquad t\geq 3. (3.6)

The first bound is due to [HPY24], and the second is due to [Bac16]. We note that sharper bounds on the 1-line are known for large tt [Tru14, Pat22], however our argument requires a bound holding for all t3t\geq 3. Second, in §3.2, we prove Theorem 1.1 for t>Tkt>T_{k} using Euler–Maclaurin summation and Lemma 2.4 and 2.5 to make explicit the argument of [Tit86, Thm. 5.13].

As in the previous section, throughout let K:=2k1K:=2^{k-1}. We will write x\lfloor x\rfloor and x\lceil x\rceil to mean the largest integer no greater than xx, and the smallest integer no smaller than xx, respectively. Unless otherwise specified, s=σ+its=\sigma+it with σ(0,1]\sigma\in(0,1] and t>0t>0.

3.1. Proof for 3tTk3\leq t\leq T_{k}

In this range, we use the following version of the Phragmén–Lindelöf Principle, due to Trudgian [Tru14a].

Lemma 3.1 (Phragmén–Lindelöf Principle).

Let a,b,Qa,b,Q be real numbers satisfying a<ba<b and a+Q>1a+Q>1. Suppose f(s)f(s) is a holomorphic function in asba\leq\Re s\leq b such that |f(s)|<Cexp(ek|t|)|f(s)|<C\exp(e^{k|t|}) for some C>0C>0 and k<π/(ba)k<\pi/(b-a). Suppose further that

|f(s)|{A|Q+s|α1logα2|Q+s|for s=aB|Q+s|β1logβ2|Q+s|for s=b|f(s)|\leq\begin{cases}A|Q+s|^{\alpha_{1}}\log^{\alpha_{2}}|Q+s|&\text{for }\Re s=a\\ B|Q+s|^{\beta_{1}}\log^{\beta_{2}}|Q+s|&\text{for }\Re s=b\end{cases}

for some A,α1,α2,B,β1,β2>0A,\alpha_{1},\alpha_{2},B,\beta_{1},\beta_{2}>0. Then, for all a<s<ba<\Re s<b,

|f(s)|(A|Q+s|α1logα2|Q+s|)bsba(B|Q+s|β1logβ2|Q+s|)saba.|f(s)|\leq\left(A|Q+s|^{\alpha_{1}}\log^{\alpha_{2}}|Q+s|\right)^{\frac{b-\Re s}{b-a}}\left(B|Q+s|^{\beta_{1}}\log^{\beta_{2}}|Q+s|\right)^{\frac{\Re s-a}{b-a}}.
Proof.

See [Tru14a, Lem. 3]. ∎

The motivation for using such a convexity argument for small tt is the bounds (3.5) and (3.6) are comparatively sharp for small tt. We choose the holomorphic function f(s):=(s1)ζ(s)f(s):=(s-1)\zeta(s).

First, we verify numerically that if s=1/2+its=1/2+it, then

sup|t|3|(s1)ζ(s)|<0.618|1.31+s|7/6log|1.31+s|\sup_{|t|\leq 3}|(s-1)\zeta(s)|<0.618|1.31+s|^{7/6}\log|1.31+s| (3.7)

and if s=1+its=1+it, then

sup|t|3|(s1)ζ(s)|<|1.31+s|log|1.31+s|.\sup_{|t|\leq 3}|(s-1)\zeta(s)|<|1.31+s|\log|1.31+s|. (3.8)

Therefore, by combining with (3.5) and (3.6) and using Lemma 3.1,

|ζ(s)|1|s1|(0.618|Q0+s|7/6log|Q0+s|)22σ(|Q0+s|log|Q0+s|)2σ1=0.61822σ|s1||Q0+s|(4σ)/3log|Q0+s|,1/2s1,\begin{split}|\zeta(s)|&\leq\frac{1}{|s-1|}\left(0.618|Q_{0}+s|^{7/6}\log|Q_{0}+s|\right)^{2-2\sigma}\left(|Q_{0}+s|\log|Q_{0}+s|\right)^{2\sigma-1}\\ &=\frac{0.618^{2-2\sigma}}{|s-1|}|Q_{0}+s|^{(4-\sigma)/3}\log|Q_{0}+s|,\qquad 1/2\leq\Re s\leq 1,\end{split} (3.9)

where Q0=1.31Q_{0}=1.31. If s=σk+its=\sigma_{k}+it with 1/2σk11/2\leq\sigma_{k}\leq 1, we have |Q0+s|=(Q0+σk)2+t22.312+t2|Q_{0}+s|=\sqrt{(Q_{0}+\sigma_{k})^{2}+t^{2}}\leq\sqrt{2.31^{2}+t^{2}}. If furthermore tt0t\geq t_{0}, then

log|Q0+s|log(t2.312t02+1)logtlogt(1+log(2.312t02+1)2logt0)logt\log|Q_{0}+s|\leq\frac{\log\left(t\sqrt{\frac{2.31^{2}}{t_{0}^{2}}+1}\right)}{\log t}\log t\leq\left(1+\frac{\log\left(\frac{2.31^{2}}{t_{0}^{2}}+1\right)}{2\log t_{0}}\right)\log t

and for k4k\geq 4, σk5/7\sigma_{k}\geq 5/7, hence for tt0t\geq t_{0}

|Q0+s|(4σk)/3(2.312t02+1)23/42t(4σk)/3.|Q_{0}+s|^{(4-\sigma_{k})/3}\leq\left(\frac{2.31^{2}}{t_{0}^{2}}+1\right)^{23/42}t^{(4-\sigma_{k})/3}. (3.10)

Hence

|ζ(σk+it)|0.61822σkA(t0)t(1σk)/3logt,tt0,|\zeta(\sigma_{k}+it)|\leq 0.618^{2-2\sigma_{k}}A(t_{0})t^{(1-\sigma_{k})/3}\log t,\qquad t\geq t_{0}, (3.11)

where

A(t0):=(2.312t02+1)23/42(1+log(2.312t02+1)2logt0).A(t_{0}):=\left(\frac{2.31^{2}}{t_{0}^{2}}+1\right)^{23/42}\left(1+\frac{\log\left(\frac{2.31^{2}}{t_{0}^{2}}+1\right)}{2\log t_{0}}\right). (3.12)

The RHS of (3.11) is majorized by 1.546t1/(2K2)logt1.546t^{1/(2K-2)}\log t if

0.618k/(K1)A(t0)1.546t(3k)/(6K6),0.618^{k/(K-1)}A(t_{0})\leq 1.546t^{(3-k)/(6K-6)}, (3.13)

i.e. if

tt1(k)=(A(t0)1.546)(6K6)/(3k)0.6186k/(3k).t\leq t_{1}(k)=\left(\frac{A(t_{0})}{1.546}\right)^{(6K-6)/(3-k)}0.618^{6k/(3-k)}. (3.14)

Taking t0=3t_{0}=3, we have A(t0)1.4747A(t_{0})\leq 1.4747, in which case t1(k)exp(8.7)t_{1}(k)\geq\exp(8.7) for all k4k\geq 4. Therefore, (3.13) is satisfied for all 3texp(8.7)3\leq t\leq\exp(8.7). Similarly, taking t0=exp(8.7)t_{0}=\exp(8.7), we have A(t0)1.0001A(t_{0})\leq 1.0001 and

t1(k)exp((6K6)log1.5461.00016klog0.618k3)Tk,k4.t_{1}(k)\geq\exp\left(\frac{(6K-6)\log\frac{1.546}{1.0001}-6k\log 0.618}{k-3}\right)\geq T_{k},\qquad k\geq 4. (3.15)

It follows that

|ζ(σk+it)|1.546t1/(2K2)logt,3tTk,k4,|\zeta(\sigma_{k}+it)|\leq 1.546t^{1/(2K-2)}\log t,\qquad 3\leq t\leq T_{k},\;k\geq 4, (3.16)

as required.

3.2. Proof for tTkt\geq T_{k}

This subsection contains our main argument. We begin by bounding the difference between ζ(s)\zeta(s) and its partial sum using Euler–Maclaurin summation. Recent explicit bounds on ζ(1/2+it)\zeta(1/2+it) have instead used the Riemann–Siegel formula [PT15, Hia16, HPY24], and the Gabcke [Gab79] remainder bound. This produces better constants on the critical line since it allows us to consider an exponential sum of length O(t1/2)O(t^{1/2}) instead of O(t)O(t). Off the critical line, applying the Riemann–Siegel formula requires an explicit bound on the remainder term holding for 1/2σ11/2\leq\sigma\leq 1. This can be achieved by appealing to the results of [Rey11], and has been done, for instance, on the 1-line in [Pat22]. For simplicity, however, we instead use Euler–Maclaurin summation, where explicit bounds on the remainder term have already been computed.

To bound the remainder term arising from the Euler–Maclaurin summation, we use the following result due to Simonič [Sim20, Cor. 2], which builds on results in [Kad13, Thm. 1.2] and [Che00, Prop. 1].

Lemma 3.2 ([Sim20] Corollary 2).

Let s=σ+its=\sigma+it where 1/2σ11/2\leq\sigma\leq 1 and tt0>0t\geq t_{0}>0. If h>(2π)1h>(2\pi)^{-1}, then

|ζ(s)1nhtns|1(ht0)σ(h+12+31+t02(112hcot12h)).\left|\zeta(s)-\sum_{1\leq n\leq ht}n^{-s}\right|\leq\frac{1}{(ht_{0})^{\sigma}}\left(h+\frac{1}{2}+3\sqrt{1+t_{0}^{-2}}\left(1-\frac{1}{2h}\cot\frac{1}{2h}\right)\right). (3.17)

Suppose now that j,k,rj,k,r are positive integers with k4k\geq 4 and

J:=2j1,K:=2k1,R:=2r1.J:=2^{j-1},\qquad K:=2^{k-1},\qquad R:=2^{r-1}. (3.18)

Throughout, we will write s=σk+its=\sigma_{k}+it where σk=1k/(2K2)\sigma_{k}=1-k/(2K-2) and tTkt\geq T_{k}. Furthermore let

θr:=RrR2R+2,\theta_{r}:=\frac{R}{rR-2R+2}, (3.19)

for all 2rk2\leq r\leq k. Roughly speaking, our approach is to use Lemma 3.2 then use partial summation and Lemma 2.5 to bound

0<nX2nσkit.\sum_{0<n\leq X_{2}}n^{-\sigma_{k}-it}. (3.20)

We divide the sum (3.20) into three subsums — in (0,X0](0,X_{0}], we use the trivial bound, and in (X0,X1](X_{0},X_{1}] and (X1,X2](X_{1},X_{2}] we apply Lemma 2.6 with different choices of kk. Here, we ultimately make the choice

X1:=X1(k,h0)=h0tθk,X2:=X2(h0,h2)=h0h2tX_{1}:=X_{1}(k,h_{0})=\left\lfloor h_{0}t^{\theta_{k}}\right\rfloor,\qquad X_{2}:=X_{2}(h_{0},h_{2})=\lfloor h_{0}h_{2}t\rfloor (3.21)

for some scaling parameters h0,h2>0h_{0},h_{2}>0, to be chosen later.

Lemma 3.3.

Let s=σk+its=\sigma_{k}+it, where tt0>0t\geq t_{0}>0 for some integer k4k\geq 4. Let r2r\geq 2 and R=2r1R=2^{r-1} and let η3>0\eta_{3}>0 be any real number. Furthermore, suppose aa, bb are integers satisfying 0<a<bha0<a<b\leq ha for some h>1h>1. Then

|a<nbns|Cr(η3,h)ak/(2K2)r/(2R2)t1/(2R2)+Dr(η3,h)ak/(2K2)2/R+r/(2R2)t1/(2R2)\begin{split}\left|\sum_{a<n\leq b}n^{-s}\right|&\leq C_{r}(\eta_{3},h)a^{k/(2K-2)-r/(2R-2)}t^{1/(2R-2)}\\ &\qquad\qquad+D_{r}(\eta_{3},h)a^{k/(2K-2)-2/R+r/(2R-2)}t^{-1/(2R-2)}\end{split} (3.22)

where

Cr(η3,h):=Ar(η3,hr)h2r/Rr/(2R2)(h1)((r1)!2π)12R2,Dr(η3,h):=Br(η3)hr/(2R2)(h1)12/R(2π(r1)!)12R2.\begin{split}C_{r}(\eta_{3},h)&:=A_{r}(\eta_{3},h^{r})h^{2r/R-r/(2R-2)}(h-1)\left(\frac{(r-1)!}{2\pi}\right)^{\frac{1}{2R-2}},\\ D_{r}(\eta_{3},h)&:=B_{r}(\eta_{3})h^{r/(2R-2)}(h-1)^{1-2/R}\left(\frac{2\pi}{(r-1)!}\right)^{\frac{1}{2R-2}}.\end{split} (3.23)
Proof.

Observe that with our choice of f(x)f(x) in (3.2),

f(r)(x)=(1)r(r1)!t2πxr.f^{(r)}(x)=\frac{(-1)^{r}(r-1)!t}{2\pi x^{r}}. (3.24)

Hence for all x(a,b]x\in(a,b], where aa and bb are integers satisfying a<bhaa<b\leq ha for some h>1h>1, we have

(r1)!2πt(ha)r|f(r)(x)|(r1)!2πtar.\frac{(r-1)!}{2\pi}\frac{t}{(ha)^{r}}\leq|f^{(r)}(x)|\leq\frac{(r-1)!}{2\pi}\frac{t}{a^{r}}. (3.25)

Therefore, we may apply Lemma 2.5 with

λr=(r1)!2πt(ha)r,h=hr,N=ba(h1)a.\lambda_{r}=\frac{(r-1)!}{2\pi}\frac{t}{(ha)^{r}},\qquad h=h^{r},\qquad N=b-a\leq(h-1)a. (3.26)

This gives

a<nbnit\displaystyle\sum_{a<n\leq b}n^{-it} Ar(η3,hr)h2r/R(h1)a((r1)!t2π(ha)r)1/(2R2)\displaystyle\leq A_{r}(\eta_{3},h^{r})h^{2r/R}(h-1)a\left(\frac{(r-1)!t}{2\pi(ha)^{r}}\right)^{1/(2R-2)} (3.27)
+Br(η3)((h1)a)12/R((r1)!t2π(ha)r)1/(2R2)\displaystyle\qquad\qquad+B_{r}(\eta_{3})((h-1)a)^{1-2/R}\left(\frac{(r-1)!t}{2\pi(ha)^{r}}\right)^{-1/(2R-2)}
=Cra1r/(2R2)t1/(2R2)+Dra12/R+r/(2R2)t1/(2R2).\displaystyle=C_{r}a^{1-r/(2R-2)}t^{1/(2R-2)}+D_{r}a^{1-2/R+r/(2R-2)}t^{-1/(2R-2)}. (3.28)

By partial summation, for any ff and σ>0\sigma>0 we have

|a<nbnσe(f(n))|\displaystyle\left|\sum_{a<n\leq b}n^{-\sigma}e(f(n))\right| =|bσa<nbe(f(n))+abσtσ1a<nte(f(n))dt|\displaystyle=\left|b^{-\sigma}\sum_{a<n\leq b}e(f(n))+\int_{a}^{b}\sigma t^{-\sigma-1}\sum_{a<n\leq t}e(f(n))\text{d}t\right| (3.29)
aσmaxa<Lb|a<nLe(f(n))|.\displaystyle\leq a^{-\sigma}\max_{a<L\leq b}\left|\sum_{a<n\leq L}e(f(n))\right|. (3.30)

The result follows from combining (3.28) and (3.30). ∎

3.2.1. The small region [1,X1][1,X_{1}]

Lemma 3.4.

Let s=σk+its=\sigma_{k}+it with tt0t\geq t_{0}. Then, for any h0>0h_{0}>0, h1>1h_{1}>1, η3>0\eta_{3}>0 and (kK2K+2)1ϕk1(kK-2K+2)^{-1}\leq\phi\leq k^{-1},

|1nX1ns|αk(h0,h1,η3,ϕ,t0)t1/(2K2)logt.\left|\sum_{1\leq n\leq X_{1}}n^{-s}\right|\leq\alpha_{k}(h_{0},h_{1},\eta_{3},\phi,t_{0})\;t^{1/(2K-2)}\log t.

where X1=X1(k,h0)X_{1}=X_{1}(k,h_{0}) is defined in (3.21), and

αk(h0,h1,η3,ϕ,t):=2K2kt(1ϕk)/(2K2)logt+12K2kt1/(2K2)logt+(θkϕlogh1+max{0,log(h0h1)logh1}logt)(Ck(η3,h1)+Dk(η3,h1)h12/Kk/(K1))+1logth11k/(2K2)h11k/(2K2)1h0k/(2K2)1t1/(kK2K+2)ϕ.\begin{split}&\alpha_{k}(h_{0},h_{1},\eta_{3},\phi,t):=\frac{2K-2}{kt^{(1-\phi k)/(2K-2)}\log t}+\frac{1-\frac{2K-2}{k}}{t^{1/(2K-2)}\log t}\\ &\qquad+\left(\frac{\theta_{k}-\phi}{\log h_{1}}+\frac{\max\left\{0,\frac{\log(h_{0}h_{1})}{\log h_{1}}\right\}}{\log t}\right)\left(C_{k}(\eta_{3},h_{1})+D_{k}(\eta_{3},h_{1})h_{1}^{2/K-k/(K-1)}\right)\\ &\qquad+\frac{1}{\log t}\frac{h_{1}^{1-k/(2K-2)}}{h_{1}^{1-k/(2K-2)}-1}h_{0}^{k/(2K-2)-1}t^{1/(kK-2K+2)-\phi}.\end{split} (3.31)

with θk\theta_{k} is as defined in (3.19).

Proof.

Let ϕ\phi be a parameter to be chosen later, satisfying

1kK2K+2ϕ1k.\frac{1}{kK-2K+2}\leq\phi\leq\frac{1}{k}. (3.32)

Define

M:=(θkϕ)logt+logh0logh1,M:=\left\lceil\frac{(\theta_{k}-\phi)\log t+\log h_{0}}{\log h_{1}}\right\rceil, (3.33)
X1:=h0tθk,X0:=h1MX1.X_{1}:=\left\lfloor h_{0}t^{\theta_{k}}\right\rfloor,\qquad X_{0}:=\lfloor h_{1}^{-M}X_{1}\rfloor. (3.34)

First, observe that

X0h0h1((θkϕ)logt+logh0)/logh1tθk=tϕ.X_{0}\leq h_{0}h_{1}^{-\left(\left(\theta_{k}-\phi\right)\log t+\log h_{0}\right)/\log h_{1}}t^{\theta_{k}}=t^{\phi}. (3.35)

Hence, from the trivial bound we have

|1nX0ns|1nX0nσk1+1X0uσkdu=1+X01σ11σ1+2K2k(tϕk/(2K2)1)\begin{split}\left|\sum_{1\leq n\leq X_{0}}n^{-s}\right|&\leq\sum_{1\leq n\leq X_{0}}n^{-\sigma_{k}}\leq 1+\int_{1}^{X_{0}}u^{-\sigma_{k}}\text{d}u\\ &=1+\frac{X_{0}^{1-\sigma}-1}{1-\sigma}\leq 1+\frac{2K-2}{k}(t^{\phi k/(2K-2)}-1)\end{split} (3.36)

since 1σ=k/(2K2)1-\sigma=k/(2K-2). If kk is so large that X0<1X_{0}<1, then the sum on the LHS is empty while the RHS is positive, so the inequality holds regardless. Furthermore, as ϕk1\phi k\leq 1, we have, for tt0t\geq t_{0},

|1nX0ns|(2K2kt0(1ϕk)/(2K2)logt0+12K2kt01/(2K2)logt0)t1/(2K2)logt.\left|\sum_{1\leq n\leq X_{0}}n^{-s}\right|\leq\left(\frac{2K-2}{kt_{0}^{(1-\phi k)/(2K-2)}\log t_{0}}+\frac{1-\frac{2K-2}{k}}{t_{0}^{1/(2K-2)}\log t_{0}}\right)t^{1/(2K-2)}\log t. (3.37)

Next, consider the sum over the interval (X1,X0](X_{1},X_{0}]. We divide the interval into MM pieces of the form (Ym+1,Ym](Y_{m+1},Y_{m}], where

Ym:=h1mX1,0mM.Y_{m}:=\left\lfloor h_{1}^{-m}X_{1}\right\rfloor,\qquad 0\leq m\leq M. (3.38)

Note that Y0=X1Y_{0}=X_{1} and YMX0Y_{M}\leq X_{0}, so the entire interval (X0,X1](X_{0},X_{1}] is covered. We divide the sum over (Ym,Ym1](Y_{m},Y_{m-1}] into

Ym<nYm1ns=Ym<nh1Ymns+h1Ym<nYm1ns=S1+S2,\sum_{Y_{m}<n\leq Y_{m-1}}n^{-s}=\sum_{Y_{m}<n\leq\lfloor h_{1}Y_{m}\rfloor}n^{-s}+\sum_{\lfloor h_{1}Y_{m}\rfloor<n\leq Y_{m-1}}n^{-s}=S_{1}+S_{2}, (3.39)

say. Recalling that σ=1k/(2K2)\sigma=1-k/(2K-2), we take a=Yma=Y_{m}, r=kr=k and h=h1h=h_{1} in (3.28) and combining with (3.30), we obtain

S1\displaystyle S_{1} Ck(η3,h1)Ymk/(2K2)k/(2K2)t1/(2K2)\displaystyle\leq C_{k}(\eta_{3},h_{1})Y_{m}^{k/(2K-2)-k/(2K-2)}t^{1/(2K-2)}
+Dk(η3,h1)Ymk/(2K2)2/K+k/(2K2)t1/(2K2)\displaystyle\qquad\qquad+D_{k}(\eta_{3},h_{1})Y_{m}^{k/(2K-2)-2/K+k/(2K-2)}t^{-1/(2K-2)} (3.40)
Ckt1/(2K2)+Dk(h1mh0tθk)k/(K1)2/Kt1/(2K2)\displaystyle\leq C_{k}t^{1/(2K-2)}+D_{k}\left(h_{1}^{-m}h_{0}t^{\theta_{k}}\right)^{k/(K-1)-2/K}t^{-1/(2K-2)} (3.41)
=(Ck+Dk(h1mh0)k/(K1)2/K)t1/(2K2)\displaystyle=\left(C_{k}+D_{k}(h_{1}^{-m}h_{0})^{k/(K-1)-2/K}\right)t^{1/(2K-2)} (3.42)

where, passing from (3.40) to (3.41) we used

Ymh1mX1h1mh0tθk,Y_{m}\leq h_{1}^{-m}X_{1}\leq h_{1}^{-m}h_{0}t^{\theta_{k}}, (3.43)

and passing from (3.41) to (3.42) we used

θk(kK12K)12K2=12K2,\theta_{k}\left(\frac{k}{K-1}-\frac{2}{K}\right)-\frac{1}{2K-2}=\frac{1}{2K-2}, (3.44)

for all k4k\geq 4.

To bound S2S_{2} of (3.39), we note that

Ym1h1Ym<h1(m1)X1h1(h1mX11)+1h1+1.Y_{m-1}-\lfloor h_{1}Y_{m}\rfloor<h_{1}^{-(m-1)}X_{1}-h_{1}(h_{1}^{-m}X_{1}-1)+1\leq h_{1}+1. (3.45)

However Ym1h1YmY_{m-1}-\lfloor h_{1}Y_{m}\rfloor is an integer so Ym1h1Ymh1Y_{m-1}-\lfloor h_{1}Y_{m}\rfloor\leq\lceil h_{1}\rceil, i.e. there are at most h1\lceil h_{1}\rceil terms in S2S_{2}. Therefore, via the trivial bound, and noting that k/(2K2)1<0k/(2K-2)-1<0,

S2h1(h1Ym+1)k/(2K2)1h1(h1(m1)h0tθk)k/(2K2)1\begin{split}S_{2}\leq\lceil h_{1}\rceil(\lfloor h_{1}Y_{m}\rfloor+1)^{k/(2K-2)-1}\leq\lceil h_{1}\rceil(h_{1}^{-(m-1)}h_{0}t^{\theta_{k}})^{k/(2K-2)-1}\end{split} (3.46)

Therefore, writing δ1=h1k/(K1)+2/K\delta_{1}=h_{1}^{-k/(K-1)+2/K} and δ2=h11k/(2K2)\delta_{2}=h_{1}^{1-k/(2K-2)},

|X0<nX1ns|m=1M|Ym<nYm1ns|=(CkM+Dkm=1Mδ1m)t1/(2K2)+S3\left|\sum_{X_{0}<n\leq X_{1}}n^{-s}\right|\leq\sum_{m=1}^{M}\left|\sum_{Y_{m}<n\leq Y_{m-1}}n^{-s}\right|=\left(C_{k}M+D_{k}\sum_{m=1}^{M}\delta_{1}^{m}\right)t^{1/(2K-2)}+S_{3} (3.47)

where

S3=h0k/(2K2)1tθk(k/(2K2)1)m=1Mδ2m1.S_{3}=h_{0}^{k/(2K-2)-1}t^{\theta_{k}(k/(2K-2)-1)}\sum_{m=1}^{M}\delta_{2}^{m-1}. (3.48)

Since δ2>1\delta_{2}>1 we have

m=1Mδ2m1=δ2M1δ21h11k/(2K2)tθkϕ1h11k/(2K2)1=δ2δ21(tθkϕ1).\sum_{m=1}^{M}\delta_{2}^{m-1}=\frac{\delta_{2}^{M}-1}{\delta_{2}-1}\leq\frac{h_{1}^{1-k/(2K-2)}t^{\theta_{k}-\phi}-1}{h_{1}^{1-k/(2K-2)}-1}=\frac{\delta_{2}}{\delta_{2}-1}(t^{\theta_{k}-\phi}-1). (3.49)

Meanwhile, from the lower bound on ϕ\phi, we have

θk(k2K21)+θkϕ12K2=1kK2K+2ϕ0\begin{split}&\theta_{k}\left(\frac{k}{2K-2}-1\right)+\theta_{k}-\phi-\frac{1}{2K-2}=\frac{1}{kK-2K+2}-\phi\leq 0\end{split} (3.50)

so that, in light of (3.48), (3.49) and (3.50), we have, for tt0t\geq t_{0},

S3E1(t0)t1/2K2logt,S_{3}\leq E_{1}(t_{0})t^{1/2K-2}\log t, (3.51)
E1(t):=1logtδ2δ21h0k/(2K2)1t1/(kK2K+2)ϕ.E_{1}(t):=\frac{1}{\log t}\frac{\delta_{2}}{\delta_{2}-1}h_{0}^{k/(2K-2)-1}t^{1/(kK-2K+2)-\phi}. (3.52)

Next, since δ1<1\delta_{1}<1, we have trivially

m=1Mδ1m<δ1M\displaystyle\sum_{m=1}^{M}\delta_{1}^{m}<\delta_{1}M Mh12/Kk/(K1).\displaystyle\leq Mh_{1}^{2/K-k/(K-1)}. (3.53)

Meanwhile, if tt0t\geq t_{0} and h1>1h_{1}>1 then

M(θkϕ)logtlogh1+logh0logh1+1(θkϕlogh1+1logt0max{0,logh0logh1+1})logtM\leq\frac{(\theta_{k}-\phi)\log t}{\log h_{1}}+\frac{\log h_{0}}{\log h_{1}}+1\leq\left(\frac{\theta_{k}-\phi}{\log h_{1}}+\frac{1}{\log t_{0}}\max\left\{0,\frac{\log h_{0}}{\log h_{1}}+1\right\}\right)\log t

Therefore, combining the previous estimates, we have for tt0t\geq t_{0},

|X0<nX1ns|\displaystyle\left|\sum_{X_{0}<n\leq X_{1}}n^{-s}\right| (MCklogt+MDkh12/Kk/(K1)+E1(t0))t1/(2K2)logt\displaystyle\leq\left(\frac{MC_{k}}{\log t}+MD_{k}h_{1}^{2/K-k/(K-1)}+E_{1}(t_{0})\right)t^{1/(2K-2)}\log t (3.54)
E2(h0,h1,η3,ϕ,t0)t1/(2K2)logt\displaystyle\leq E_{2}(h_{0},h_{1},\eta_{3},\phi,t_{0})\;t^{1/(2K-2)}\log t (3.55)

where

E2(h0,h1,η3,ϕ,t)=1logth11k/(2K2)h11k/(2K2)1h0k/(2K2)1t1/(kK2K+2)ϕ\displaystyle E_{2}(h_{0},h_{1},\eta_{3},\phi,t)=\frac{1}{\log t}\frac{h_{1}^{1-k/(2K-2)}}{h_{1}^{1-k/(2K-2)}-1}h_{0}^{k/(2K-2)-1}t^{1/(kK-2K+2)-\phi}
+(θkϕlogh1+max{0,log(h0h1)logh1}logt)(Ck(η3,h1)+Dk(η3,h1)h12/Kk/(K1))\displaystyle\qquad+\left(\frac{\theta_{k}-\phi}{\log h_{1}}+\frac{\max\left\{0,\frac{\log(h_{0}h_{1})}{\log h_{1}}\right\}}{\log t}\right)\left(C_{k}(\eta_{3},h_{1})+D_{k}(\eta_{3},h_{1})h_{1}^{2/K-k/(K-1)}\right)

is decreasing in tt. Combining with (3.37), the result follows. ∎

3.2.2. The large region (X1,X2](X_{1},X_{2}]

In this region we use Lemma 2.6 with smaller choices of kk.

Lemma 3.5.

Let k4k\geq 4 and s=σk+its=\sigma_{k}+it with tt0>0t\geq t_{0}>0. Then, for any h0,η3>0h_{0},\eta_{3}>0, h2(0,e)h_{2}\in(0,e) and h3>1h_{3}>1, we have

|X1<nX2ns|βk(t0)t1/(2K2)logt\left|\sum_{X_{1}<n\leq X_{2}}n^{-s}\right|\leq\beta_{k}(t_{0})t^{1/(2K-2)}\log t (3.56)

where X1(k,h0)X_{1}(k,h_{0}) and X2(h0,h2)X_{2}(h_{0},h_{2}) are defined in (3.21), and

βk(t)=βk(h0,h2,h3,η3,t):=r=2k1Fk(r,t),\beta_{k}(t)=\beta_{k}(h_{0},h_{2},h_{3},\eta_{3},t):=\sum_{r=2}^{k-1}F_{k}(r,t), (3.57)
Fk(r,t):=(θrθr+1logh3logh2(k2)logt+1logt)t(θrk1)/(2K2)×[(Cr(η3,h3)Hk/(2K2)r/(2R2)+Dr(η3,h3)Hk/(2K2)2/R+r/(2R2))tθr/R+h3(h3H)k/(2K2)1tθr]\begin{split}&F_{k}(r,t):=\left(\frac{\theta_{r}-\theta_{r+1}}{\log h_{3}}-\frac{\log h_{2}}{(k-2)\log t}+\frac{1}{\log t}\right)t^{(\theta_{r}k-1)/(2K-2)}\times\\ &\quad\bigg{[}\left(C_{r}(\eta_{3},h_{3})H^{k/(2K-2)-r/(2R-2)}+D_{r}(\eta_{3},h_{3})H^{k/(2K-2)-2/R+r/(2R-2)}\right)t^{-\theta_{r}/R}\\ &\qquad\qquad+\lceil h_{3}\rceil(h_{3}H)^{k/(2K-2)-1}t^{-\theta_{r}}\bigg{]}\end{split} (3.58)

and

H:=H(r,k)=h0h2(kr)/(k2)h3.H:=H(r,k)=\frac{h_{0}h_{2}^{(k-r)/(k-2)}}{h_{3}}. (3.59)
Proof.

We begin by defining

Zr:=h0h2(kr)/(k2)tθr,2rk.Z_{r}:=\lfloor h_{0}h_{2}^{(k-r)/(k-2)}t^{\theta_{r}}\rfloor,\qquad 2\leq r\leq k. (3.60)

where θk\theta_{k} is defined in (3.19). Noting that Z2=X2Z_{2}=X_{2} and Zk=X1Z_{k}=X_{1}, we have

|X1<nX2ns|r=2k1|Zr+1<nZrns|.\left|\sum_{X_{1}<n\leq X_{2}}n^{-s}\right|\leq\sum_{r=2}^{k-1}\left|\sum_{Z_{r+1}<n\leq Z_{r}}n^{-s}\right|. (3.61)

We further divide each interval (Zr+1,Zr](Z_{r+1},Z_{r}] into intervals of the form (Wm,Wm1](W_{m},W_{m-1}], m=1,2,,Mrm=1,2,\ldots,M_{r} where

Wm:=max{Zr+1,h3mZr}W_{m}:=\max\{Z_{r+1},\lfloor h_{3}^{-m}Z_{r}\rfloor\} (3.62)

and MrM_{r} is the smallest integer for which WMr=Zr+1W_{M_{r}}=Z_{r+1}. Note in particular that since h3>1h_{3}>1,

h31logh3((θrθr+1)logtlogh2k2)Zrh0h2(kr1)/(k2)t(θrθr+1)tθr=Zr+1,\left\lfloor h_{3}^{-\left\lceil\frac{1}{\log h_{3}}\left((\theta_{r}-\theta_{r+1})\log t-\frac{\log h_{2}}{k-2}\right)\right\rceil}Z_{r}\right\rfloor\leq\lfloor h_{0}h_{2}^{(k-r-1)/(k-2)}t^{-(\theta_{r}-\theta_{r+1})}t^{\theta_{r}}\rfloor=Z_{r+1}, (3.63)

so that

Mr1logh3((θrθr+1)logtlogh2k2).M_{r}\leq\left\lceil\frac{1}{\log h_{3}}\left((\theta_{r}-\theta_{r+1})\log t-\frac{\log h_{2}}{k-2}\right)\right\rceil. (3.64)

Consider now the sum over (Wm,Wm1](W_{m},W_{m-1}]. We divide the sum as before, with

Wm<nWm1ns=Wm<nh3Wmns+h3Wm<nWm1ns.\sum_{W_{m}<n\leq W_{m-1}}n^{-s}=\sum_{W_{m}<n\leq\lfloor h_{3}W_{m}\rfloor}n^{-s}+\sum_{\lfloor h_{3}W_{m}\rfloor<n\leq W_{m-1}}n^{-s}. (3.65)

Since h3Wmh3Wm\lfloor h_{3}W_{m}\rfloor\leq h_{3}W_{m}, we may apply (3.28) to the first sum with a=Wma=W_{m} and h=h3h=h_{3} to obtain, for r2r\geq 2 and 1mMr1\leq m\leq M_{r},

|Wm<nh3Wmns|Cr(η3,h3)Wmκ3t1/(2R2)+Dr(η3,h3)Wmκ4t1/(2R2)\begin{split}\left|\sum_{W_{m}<n\leq h_{3}\lfloor W_{m}\rfloor}n^{-s}\right|&\leq C_{r}(\eta_{3},h_{3})W_{m}^{\kappa_{3}}t^{1/(2R-2)}+D_{r}(\eta_{3},h_{3})W_{m}^{\kappa_{4}}t^{-1/(2R-2)}\end{split} (3.66)

where

κ3:=κ(r,k)=k2K2r2R2,κ4:=κ4(r,k)=k2K22R+r2R2.\kappa_{3}:=\kappa(r,k)=\frac{k}{2K-2}-\frac{r}{2R-2},\qquad\kappa_{4}:=\kappa_{4}(r,k)=\frac{k}{2K-2}-\frac{2}{R}+\frac{r}{2R-2}. (3.67)

For k>r2k>r\geq 2, we have the identities

θrκ3+12R2=θr(k2K2rθr12R2)=θr(k2K21R),\theta_{r}\kappa_{3}+\frac{1}{2R-2}=\theta_{r}\left(\frac{k}{2K-2}-\frac{r-\theta_{r}^{-1}}{2R-2}\right)=\theta_{r}\left(\frac{k}{2K-2}-\frac{1}{R}\right), (3.68)
θrκ412R2=θr(k2K21R).\theta_{r}\kappa_{4}-\frac{1}{2R-2}=\theta_{r}\left(\frac{k}{2K-2}-\frac{1}{R}\right). (3.69)

Therefore, noting that WmW_{m} is decreasing in mm and κ3>0\kappa_{3}>0 for all r<kr<k, we have, via (3.68),

Wmκ3t1/(2R2)W1κ3t1/(2R2)(h31h0h2(kr)/(k2)tθr)κ3t1/(2R2)=Hκ3tθr(k/(2K2)1/R)\begin{split}W_{m}^{\kappa_{3}}t^{1/(2R-2)}\leq W_{1}^{\kappa_{3}}t^{1/(2R-2)}&\leq\left(h_{3}^{-1}h_{0}h_{2}^{(k-r)/(k-2)}t^{\theta_{r}}\right)^{\kappa_{3}}t^{1/(2R-2)}\\ &=H^{\kappa_{3}}t^{\theta_{r}(k/(2K-2)-1/R)}\end{split} (3.70)

where HH is defined in (3.59). Similarly, via (3.69) and κ4>0\kappa_{4}>0 we have

Wmκ4t1/(2R2)Hκ4tθr(k/(2K2)1/R).W_{m}^{\kappa_{4}}t^{-1/(2R-2)}\leq H^{\kappa_{4}}t^{\theta_{r}(k/(2K-2)-1/R)}.

For the second sum in (3.65), we note as before that

Wm1h1Wm<h3(m1)Zrh3(h3mZr1)+1h3+1.W_{m-1}-\lfloor h_{1}W_{m}\rfloor<h_{3}^{-(m-1)}Z_{r}-h_{3}(h_{3}^{-m}Z_{r}-1)+1\leq h_{3}+1.

and hence, since Wm1h1WmW_{m-1}-\lfloor h_{1}W_{m}\rfloor is an integer, we in fact have h1Wmh3\lfloor h_{1}W_{m}\rfloor\leq\lceil h_{3}\rceil, which is the maximum number of terms in the second sum of (3.65)\eqref{s3_second_sum_div}. Therefore, using the fact that nσn^{-\sigma} is decreasing and h3Wm<h3Wm+1h_{3}W_{m}<\lfloor h_{3}W_{m}\rfloor+1,

|h3Wm<nWm1ns|<h3(h3Wm)k/(2K2)1h3(h3(m1)h0h2(kr)/(k2)tθr)k/(2K2)1.\begin{split}\left|\sum_{\lfloor h_{3}W_{m}\rfloor<n\leq W_{m-1}}n^{-s}\right|&<\lceil h_{3}\rceil(h_{3}W_{m})^{k/(2K-2)-1}\\ &\leq\lceil h_{3}\rceil\left(h_{3}^{-(m-1)}h_{0}h_{2}^{(k-r)/(k-2)}t^{\theta_{r}}\right)^{k/(2K-2)-1}.\end{split} (3.71)

Combining (3.66), (3.70) and (3.71) gives

|Zr+1<nZrns|\displaystyle\left|\sum_{Z_{r+1}<n\leq Z_{r}}n^{-s}\right| m=1Mr|Wm<nWm1ns|\displaystyle\leq\sum_{m=1}^{M_{r}}\left|\sum_{W_{m}<n\leq W_{m-1}}n^{-s}\right| (3.72)
Mr[Cr(η3,h3)Hκ3+Dr(η3,h3)Hκ4]tθr(k/(2K2)1/R)\displaystyle\leq M_{r}\left[C_{r}(\eta_{3},h_{3})H^{\kappa_{3}}+D_{r}(\eta_{3},h_{3})H^{\kappa_{4}}\right]t^{\theta_{r}(k/(2K-2)-1/R)}
+Mrh3(h3H)k/(2K2)1tθr(k/(2K2)1).\displaystyle\qquad\qquad+M_{r}\lceil h_{3}\rceil(h_{3}H)^{k/(2K-2)-1}t^{\theta_{r}(k/(2K-2)-1)}. (3.73)

Using

Mr(θrθr+1)logtlogh3logh2k2+1M_{r}\leq\frac{(\theta_{r}-\theta_{r+1})\log t}{\log h_{3}}-\frac{\log h_{2}}{k-2}+1 (3.74)

and substituting into (3.61), we obtain

|X1<nX2ns|βk(t)t1/(2K2)logt\left|\sum_{X_{1}<n\leq X_{2}}n^{-s}\right|\leq\beta_{k}(t)t^{1/(2K-2)}\log t (3.75)

where βk(t)\beta_{k}(t) is defined in (3.57).

Next, we will show that if kk is fixed, then Fk(r,t)F_{k}(r,t), and hence βk(t)\beta_{k}(t), is a decreasing function of tt, so that βk(t)βk(t0)\beta_{k}(t)\leq\beta_{k}(t_{0}) for tt0t\geq t_{0}. Upon noting that for 2xx+12^{x}\geq x+1 for x1x\geq 1, we have

(kr)R(2kr1)R=KR(k-r)R\leq(2^{k-r}-1)R=K-R (3.76)

and hence

θr(k2K21R)12K2=12K2(kRrR2R+21)1rR2R+2=(kr)R+2(RK)(2K2)(rR2R+2)0.\begin{split}\theta_{r}\left(\frac{k}{2K-2}-\frac{1}{R}\right)-\frac{1}{2K-2}&=\frac{1}{2K-2}\left(\frac{kR}{rR-2R+2}-1\right)-\frac{1}{rR-2R+2}\\ &=\frac{(k-r)R+2(R-K)}{(2K-2)(rR-2R+2)}\leq 0.\end{split} (3.77)

Additionally,

θr(k2K21)12K2<θr(k2K21R)12K20.\theta_{r}\left(\frac{k}{2K-2}-1\right)-\frac{1}{2K-2}<\theta_{r}\left(\frac{k}{2K-2}-\frac{1}{R}\right)-\frac{1}{2K-2}\leq 0. (3.78)

Finally, from the assumption that h2<eh_{2}<e and k4k\geq 4,

logh2k2+1>0-\frac{\log h_{2}}{k-2}+1>0 (3.79)

and so Fk(r,t)F_{k}(r,t) is decreasing, so βk(t)βk(t0)\beta_{k}(t)\leq\beta_{k}(t_{0}) for tt0t\geq t_{0}. ∎

3.2.3. Putting it all together

For each row (k,η3,h0,h1,h2,h3,αk,βk,γk)(k,\eta_{3},h_{0},h_{1},h_{2},h_{3},\alpha_{k},\beta_{k},\gamma_{k}) of Table 1, we substitute the appropriate variables into Lemma 3.4 and 3.5 and take t0=Tkt_{0}=T_{k}. This gives

|1nX2ns|(αk+βk)t1/(2K2)logt,tTk.\left|\sum_{1\leq n\leq X_{2}}n^{-s}\right|\leq(\alpha_{k}+\beta_{k})t^{1/(2K-2)}\log t,\qquad t\geq T_{k}. (3.80)

Subsequently, taking h=h0h2>1/(2π)h=h_{0}h_{2}>1/(2\pi) and σ=σk[5/7,1)\sigma=\sigma_{k}\in[5/7,1) in Lemma 3.2,

|ζ(σk+it)|(αk+βk)t1/(2K2)logt+G(h0h2,σk)γkt1/(2K2)logt,\begin{split}|\zeta(\sigma_{k}+it)|&\leq(\alpha_{k}+\beta_{k})t^{1/(2K-2)}\log t+G(h_{0}h_{2},\sigma_{k})\leq\gamma_{k}t^{1/(2K-2)}\log t,\end{split} (3.81)

for tt01010t\geq t_{0}\geq 10^{10}, where

G(h,σ):=1(ht0)σ(h+12+31+t02(112hcot12h)).G(h,\sigma):=\frac{1}{(ht_{0})^{\sigma}}\left(h+\frac{1}{2}+3\sqrt{1+t_{0}^{-2}}\left(1-\frac{1}{2h}\cot\frac{1}{2h}\right)\right). (3.82)

The last inequality of (3.81) follows from

αk+βk+G(h0h2,σ)t1/(2K2)logtαk+βk+G(h0h2,σ)t01/(2K2)logt0γk\alpha_{k}+\beta_{k}+\frac{G(h_{0}h_{2},\sigma)}{t^{1/(2K-2)}\log t}\leq\alpha_{k}+\beta_{k}+\frac{G(h_{0}h_{2},\sigma)}{t_{0}^{1/(2K-2)}\log t_{0}}\leq\gamma_{k} (3.83)

for all tt0t\geq t_{0}. This proves Theorem 1.1 for 4k94\leq k\leq 9 and tTkt\geq T_{k}.

Table 1. Values appearing in §3.2.3
kk η3\eta_{3} h0h_{0} h1h_{1} h2h_{2} h3h_{3} αk\alpha_{k} βk\beta_{k} γk\gamma_{k}
4 1.22626 0.03640 1.30262 4.37500 1.30021 1.1796 0.3655 1.546
5 1.43074 0.10750 1.17205 17.2191 1.28297 0.7253 0.6401 1.366
6 1.79198 0.40548 1.08095 25.8377 1.19628 0.4944 0.6267 1.122
7 1.95195 0.97083 1.02940 6.87426 1.09787 0.3634 0.5350 0.899
8 1.94390 0.98846 1.01101 5.00587 1.05355 0.2824 0.4405 0.723
9 1.85285 0.99604 1.00392 3.80684 1.02923 0.2285 0.3652 0.594

Suppose now that k10k\geq 10. We take h0=h1=h3=eh_{0}=h_{1}=h_{3}=e, h2=1h_{2}=1 and η3=4.7399\eta_{3}=4.7399 (as in Lemma 2.6). This choice of parameters gives us H=1H=1 and, by Lemma 2.6,

Ak(η3,h1)=Ak(η3,h3)2.762,Bk(η3,h1)=Bk(η3,h3)1.02.A_{k}(\eta_{3},h_{1})=A_{k}(\eta_{3},h_{3})\leq 2.762,\qquad B_{k}(\eta_{3},h_{1})=B_{k}(\eta_{3},h_{3})\leq 1.02. (3.84)

We will show that Fk(r,t)F_{k}(r,t) is decreasing in kk for k10k\geq 10, for fixed r<kr<k and t>1t>1. Let

f(k)=θrk12K2f(k)=\frac{\theta_{r}k-1}{2K-2} (3.85)

so that, since θr\theta_{r} is decreasing in rr,

f(k)f(k+1)=θr(k2K2k+14K2)+14K212K2θk1(kKK+2)K2(K1)(2K1)\begin{split}f(k)-f(k+1)&=\theta_{r}\left(\frac{k}{2K-2}-\frac{k+1}{4K-2}\right)+\frac{1}{4K-2}-\frac{1}{2K-2}\\ &\geq\frac{\theta_{k-1}(kK-K+2)-K}{2(K-1)(2K-1)}\end{split} (3.86)

However,

θk1=K(k1)K2K+4>KkKK+2\theta_{k-1}=\frac{K}{(k-1)K-2K+4}>\frac{K}{kK-K+2} (3.87)

and thus f(k)f(k+1)>0f(k)-f(k+1)>0, i.e. ff is decreasing. Therefore, the factor t(θrk1)/(2K2)t^{(\theta_{r}k-1)/(2K-2)} appearing in (3.58) is decreasing in kk for fixed t>1t>1. Since H=1H=1, the only other way Fk(r,t)F_{k}(r,t) depends on kk is through the factor

(h3H)k/(2K2)1(h_{3}H)^{k/(2K-2)-1} (3.88)

which is decreasing in kk since h3H=h3>1h_{3}H=h_{3}>1. Hence, Fk(r,t)F_{k}(r,t) is decreasing in kk. Therefore, for all kk04k\geq k_{0}\geq 4,

βk(t)=r=2k1Fk(r,t)r=2k01Fk0(r,t)+r=k0k1Fk(r,t)=βk0(t)+r=k0k1Fk(r,t),\beta_{k}(t)=\sum_{r=2}^{k-1}F_{k}(r,t)\leq\sum_{r=2}^{k_{0}-1}F_{k_{0}}(r,t)+\sum_{r=k_{0}}^{k-1}F_{k}(r,t)=\beta_{k_{0}}(t)+\sum_{r=k_{0}}^{k-1}F_{k}(r,t), (3.89)

with the understanding that if k=k0k=k_{0} then the last sum is 0. Now, for all r<kr<k, we have

kr+2<22kr=2KRk-r+2<2\cdot 2^{k-r}=\frac{2K}{R} (3.90)

hence

k2K2R<r2+2R=1θr,k-\frac{2K-2}{R}<r-2+\frac{2}{R}=\frac{1}{\theta_{r}}, (3.91)

from which it follows that

θrk12K2θrR<0.\frac{\theta_{r}k-1}{2K-2}-\frac{\theta_{r}}{R}<0. (3.92)

This implies that for t1t\geq 1,

t(θrk1)/(2K2)<1,tθr(k/(2K2)1/R)1/(2K2)<1t^{(\theta_{r}k-1)/(2K-2)}<1,\qquad t^{\theta_{r}(k/(2K-2)-1/R)-1/(2K-2)}<1 (3.93)

Substituting (3.93) into (3.58), and using h2=H=1h_{2}=H=1, we obtain

Fk(r,t)(θrθr+1logh3+1logt)(h3h0k/(2K2)1+Cr(η3,h3)+Dr(η3,h3)).F_{k}(r,t)\leq\left(\frac{\theta_{r}-\theta_{r+1}}{\log h_{3}}+\frac{1}{\log t}\right)\left(\lceil h_{3}\rceil h_{0}^{k/(2K-2)-1}+C_{r}(\eta_{3},h_{3})+D_{r}(\eta_{3},h_{3})\right). (3.94)

Next, for k10k\geq 10, and given the choices of h0h_{0}, h3h_{3},

h3h0k/(2K2)11.115.\lceil h_{3}\rceil h_{0}^{k/(2K-2)-1}\leq 1.115. (3.95)

Meanwhile, for tTkt\geq T_{k},

r=10k1(θrθr+1logh3+1logt)<θ10+k10logtθ10+(k10)(k3)3.2078(K1)+2.8876k0.12494.\sum_{r=10}^{k-1}\left(\frac{\theta_{r}-\theta_{r+1}}{\log h_{3}}+\frac{1}{\log t}\right)<\theta_{10}+\frac{k-10}{\log t}\leq\theta_{10}+\frac{(k-10)(k-3)}{3.2078(K-1)+2.8876k}\leq 0.12494. (3.96)

If r10r\geq 10, then via Lemma 2.6, since h3<3h_{3}<3, we have Ar2.762A_{r}\leq 2.762, Br1.02B_{r}\leq 1.02 and

h32r/Rr/(2R2)(h31)(Γ(r)2π)12R21.0179,h_{3}^{2r/R-r/(2R-2)}(h_{3}-1)\left(\frac{\Gamma(r)}{2\pi}\right)^{\frac{1}{2R-2}}\leq 1.0179, (3.97)
h3r/(2R2)(h31)12/R(2πΓ(r))12R2<1,h_{3}^{r/(2R-2)}(h_{3}-1)^{1-2/R}\left(\frac{2\pi}{\Gamma(r)}\right)^{\frac{1}{2R-2}}<1, (3.98)

so that

Cr(h3)2.804,Dr(h3)1.02,r10.C_{r}(h_{3})\leq 2.804,\qquad D_{r}(h_{3})\leq 1.02,\qquad r\geq 10. (3.99)

Combining (3.94), (3.95), (3.96), and (3.99), we have

r=15k1Fr,k(t)<0.12494(1.115+2.804+1.02)0.6171.\sum_{r=15}^{k-1}F_{r,k}(t)<0.12494\left(1.115+2.804+1.02\right)\leq 0.6171. (3.100)

Hence, using k0=10k_{0}=10 in (3.89),

βk(h0,h2,h3,η3,Tk)β10(h0,h2,h3,η3,T10)+r=10k1Fr,k(Tk)0.6064+0.6171=1.2235.\begin{split}\beta_{k}(h_{0},h_{2},h_{3},\eta_{3},T_{k})&\leq\beta_{10}(h_{0},h_{2},h_{3},\eta_{3},T_{10})+\sum_{r=10}^{k-1}F_{r,k}(T_{k})\\ &\leq 0.6064+0.6171=1.2235.\end{split} (3.101)

Finally, substituting η3=1,h0=e,h1=e,ϕ=k1,t0=TkT10\eta_{3}=1,h_{0}=e,h_{1}=e,\phi=k^{-1},t_{0}=T_{k}\geq T_{10} into Lemma 3.4,

αk=2K2klogt+12K2kt1/(2K2)logt+(θk1k+2logt)(Ck(η3,h1)+Dk(η3,h1)h12/Kk/(K1))+1logth11k/(2K2)h11k/(2K2)1h0k/(2K2)1t1/(kK2K+2)1/k.\begin{split}\alpha_{k}&=\frac{2K-2}{k\log t}+\frac{1-\frac{2K-2}{k}}{t^{1/(2K-2)}\log t}\\ &\qquad+\left(\theta_{k}-\frac{1}{k}+\frac{2}{\log t}\right)\left(C_{k}(\eta_{3},h_{1})+D_{k}(\eta_{3},h_{1})h_{1}^{2/K-k/(K-1)}\right)\\ &\qquad+\frac{1}{\log t}\frac{h_{1}^{1-k/(2K-2)}}{h_{1}^{1-k/(2K-2)}-1}h_{0}^{k/(2K-2)-1}t^{1/(kK-2K+2)-1/k}.\end{split} (3.102)

First, we use 1exx1-e^{-x}\leq x to obtain

2K2klogt(1t1/(2K2))2K2klogtlogt2K2110,\frac{2K-2}{k\log t}\left(1-t^{-1/(2K-2)}\right)\leq\frac{2K-2}{k\log t}\frac{\log t}{2K-2}\leq\frac{1}{10}, (3.103)

hence the first two terms of (3.102) are majorized by

110+1Tk1/(2K2)logTk0.105.\frac{1}{10}+\frac{1}{T_{k}^{1/(2K-2)}\log T_{k}}\leq 0.105. (3.104)

Next,

θk1k+2logt2k(k2)+2logTk0.036,\theta_{k}-\frac{1}{k}+\frac{2}{\log t}\leq\frac{2}{k(k-2)}+\frac{2}{\log T_{k}}\leq 0.036, (3.105)

and, as before, Ck(η3,h1)2.804C_{k}(\eta_{3},h_{1})\leq 2.804 and Dk(η3,h1)1.02D_{k}(\eta_{3},h_{1})\leq 1.02 since we have chosen h1=h3h_{1}=h_{3} and k10k\geq 10. Combined with the trivial bound h12/Kk/(K1)<1h_{1}^{2/K-k/(K-1)}<1, the third term of (3.102) is no greater than 0.1380.138.

Finally, to bound the last term we combine the (wasteful) estimates

h11k/(2K2)2.691,h0k/(2K2)1<1,1kK2K+21k<0,h_{1}^{1-k/(2K-2)}\geq 2.691,\qquad h_{0}^{k/(2K-2)-1}<1,\qquad\frac{1}{kK-2K+2}-\frac{1}{k}<0, (3.106)

each valid for k10k\geq 10. Combined with t>1t>1 and the decreasingness of x/(x1)x/(x-1) for x>1x>1, the last term of (3.102) is bounded by

1logTk2.6912.69110.009.\frac{1}{\log T_{k}}\frac{2.691}{2.691-1}\leq 0.009. (3.107)

Hence, combining (3.104), (3.105) and (3.107), we finally have

αk(h0,h1,η3,ϕ,Tk)0.105+0.138+0.009=0.252.\alpha_{k}(h_{0},h_{1},\eta_{3},\phi,T_{k})\leq 0.105+0.138+0.009=0.252. (3.108)

Combining with (3.101) gives

|1netns|1.476t1/(2K2)logt,tTk,k10.\left|\sum_{1\leq n\leq\lfloor et\rfloor}n^{-s}\right|\leq 1.476t^{1/(2K-2)}\log t,\qquad t\geq T_{k},\;k\geq 10. (3.109)

Meanwhile, applying σkσ10\sigma_{k}\geq\sigma_{10} and tTkT10t\geq T_{k}\geq T_{10} to (3.82) gives the crude bound G(h0h2,σk)0.001G(h_{0}h_{2},\sigma_{k})\leq 0.001. Thus

|ζ(σk+it)|1.476t1/(2K2)logt+0.001<1.546t1/(2K2)logt|\zeta(\sigma_{k}+it)|\leq 1.476t^{1/(2K-2)}\log t+0.001<1.546t^{1/(2K-2)}\log t (3.110)

for all k10k\geq 10 and tTkt\geq T_{k}, as required.

4. Proof of Corollary 1.2

Existing methods of deriving explicit zero-free regions of ζ(s)\zeta(s) fall into two broad categories — the “global” approach of Hadamard, and the “local” method of Landau. The first method relies on explicit estimates of ζ/ζ(s)\Re\zeta^{\prime}/\zeta(s) to the right of the line σ=1\sigma=1, which in turn rely on estimates of Γ/Γ(s)\Re\Gamma^{\prime}/\Gamma(s) [Kad05]. Currently, the sharpest known classical zero-free region is derived using a variant of this method [MTY24].

The second method uses upper bounds on |ζ(s)||\zeta(s)| slightly to the left of σ=1\sigma=1, i.e. within the critical strip. The strength of the resulting zero-free region depends on the sharpness of this bound. The explicit Vinogradov–Korobov zero-free region uses a Ford–Richert type bound

ζ(σ+it)εtB(1σ)3/2+ε,\zeta(\sigma+it)\ll_{\varepsilon}t^{B(1-\sigma)^{3/2}+\varepsilon}, (4.1)

uniformly for 1/2σ11/2\leq\sigma\leq 1, for some constant B>0B>0 and any ε>0\varepsilon>0. On the other hand, the classical zero-free region can be obtained via a bound of the form

ζ(σ+it)εtθ+ε\zeta(\sigma+it)\ll_{\varepsilon}t^{\theta+\varepsilon} (4.2)

for some fixed σ,θ>0\sigma,\theta>0 and any ε>0\varepsilon>0 [For02a]. In this work, we use Theorem 1.1 to obtain a bound of the form

ζ(σ(t)+it)εtθ(t)+ε\zeta(\sigma(t)+it)\ll_{\varepsilon}t^{\theta(t)+\varepsilon} (4.3)

for some functions σ(t)1\sigma(t)\to 1 and θ(t)0\theta(t)\to 0 as tt\to\infty, to prove Corollary 1.2.

Both methods also make critical use of a non-negative trigonometric polynomial P(x)P(x), of the form

P(x):=j=0Dbjcos(jx)P(x):=\sum_{j=0}^{D}b_{j}\cos(jx) (4.4)

such that P(x)0P(x)\geq 0, bj0b_{j}\geq 0 and b0<b1b_{0}<b_{1}. In this work, we use the polynomial of degree D=46D=46 presented in [MTY24, Table 1], which was found via a large-scale computational search. The coefficients bjb_{j} of this polynomial satisfy

b0=1,b1=1.74708744081848,b:=j=1Dbj=3.57440943022073.b_{0}=1,\qquad b_{1}=1.74708744081848,\qquad b:=\sum_{j=1}^{D}b_{j}=3.57440943022073. (4.5)

4.1. Notation

Throughout, let k4k\geq 4 be an integer, and let

ηk:=k2k2.\eta_{k}:=\frac{k}{2^{k}-2}. (4.6)

The variables bj,bb_{j},b and DD will be reserved for the trigonometric polynomial (4.5). Non-trivial zeroes of ζ(s)\zeta(s) are denoted by ρ=β+it\rho=\beta+it and ρ=β+it\rho^{\prime}=\beta^{\prime}+it^{\prime}. We also find it convenient to write

L1:=L1(t)=log(Dt+1),L2:=L2(t)=loglog(Dt+1).L_{1}:=L_{1}(t)=\log(Dt+1),\qquad L_{2}:=L_{2}(t)=\log\log(Dt+1). (4.7)

The variables A=76.2A=76.2 and B=4.45B=4.45 will be used to denote the constants appearing in the Ford–Richert bound

|ζ(σ+it)|AtB(1σ)3/2log2/3t,12σ1,t3.|\zeta(\sigma+it)|\leq At^{B(1-\sigma)^{3/2}}\log^{2/3}t,\qquad\frac{1}{2}\leq\sigma\leq 1,t\geq 3. (4.8)

This inequality is instrumental in the proof of Ford’s [For02a] Vinogradov–Korobov zero-free region (1.5). Throughout, we write

N(t,η):=#{ρ:|1+itρ|η}N(t,\eta):=\#\left\{\rho:|1+it-\rho|\leq\eta\right\} (4.9)

where the zeroes are counted with multiplicity.

Our argument is roughly the same as [For02a] — an upper bound on ζ(s)\zeta(s) within the critical strip is used to construct an inequality involving the real and imaginary parts of a hypothetical zero, then a contradiction is obtained if the real part of the zero is too large. We divide the argument into two sections, and this is reflected in the intermediary lemmas below. For small tt, we use Theorem 1.1 and [HPY24, Thm. 1.1], combined with the Phragmén–Lindelöf Principle to bound ζ(s)\zeta(s) uniformly in 1/2σ5/71/2\leq\sigma\leq 5/7. For large tt, we use Theorem 1.1 directly by setting k=k(t)k=k(t), tending to infinity with tt. The two-part argument allows us to customise our tools for a specific region, significantly improving the constant in Corollary 1.2. Throughout, we will reuse Ford’s results where possible, making changes only when necessary or if a substantial improvement can be obtained.

We begin by bounding an integral involving ζ(s)\zeta(s), taken on a vertical line within the critical strip. Ultimately, this bound is combined with an upper bound on ζ(s)\zeta(s) to produce the “main” term in the zero-free region constant. The following lemma is largely the same as [For02a, Lem. 3.4], the main difference being that we only require a bound on ζ(σ+it)\zeta(\sigma+it) to hold for large tt instead of for all t3t\geq 3. This technical change allows us to avoid applying the Phragmén–Lindelöf Principle (Lemma 3.1) at very small values of tt where it is poorly suited.

Lemma 4.1.

Let 0<a1/20<a\leq 1/2, t0100t_{0}\geq 100, t3t0t\geq 3t_{0} and 1/2σ1t11/2\leq\sigma\leq 1-t^{-1}, and suppose

|ζ(σ+iy)|XtYlogt,t0|y|t,|\zeta(\sigma+iy)|\leq Xt^{Y}\log t,\qquad t_{0}\leq|y|\leq t,

for some X,Y>0X,Y>0. Then

12log|ζ(σ+it+iau)|cosh2udulogX+Ylogt+loglogt.\frac{1}{2}\int_{-\infty}^{\infty}\frac{\log|\zeta(\sigma+it+iau)|}{\cosh^{2}u}\text{d}u\leq\log X+Y\log t+\log\log t.
Proof.

We follow essentially the same argument as [For02a, Lem. 3.4] by splitting the integral into four parts. Let

log|ζ(σ+it+iau)|cosh2udu=2t/a+2t/a(t+t0)/a+(t+t0)/a(tt0)/a+(tt0)/a=I1+I2+I3+I4.\begin{split}\int_{-\infty}^{\infty}\frac{\log|\zeta(\sigma+it+iau)|}{\cosh^{2}u}\text{d}u&=\int_{-\infty}^{-2t/a}+\int_{-2t/a}^{-(t+t_{0})/a}+\int_{-(t+t_{0})/a}^{-(t-t_{0})/a}+\int_{-(t-t_{0})/a}^{\infty}\\ &=I_{1}+I_{2}+I_{3}+I_{4}.\end{split} (4.10)

To bound I1I_{1}, if u2t/au\leq-2t/a then t+au(,t]t+au\in(-\infty,-t] and by the lemma’s assumptions, we have

log|ζ(σ+i(t+au))|logX+Ylog(tau)+loglog(tau)log(XtYlogt)+(Y+1logt)(au2tt)\begin{split}\log|\zeta(\sigma+i(t+au))|&\leq\log X+Y\log(-t-au)+\log\log(-t-au)\\ &\leq\log(Xt^{Y}\log t)+\left(Y+\frac{1}{\log t}\right)\left(\frac{-au-2t}{t}\right)\end{split} (4.11)

where in the second inequality we have used log(x+1)x\log(x+1)\leq x with x=aut20x=-\frac{au}{t}-2\geq 0. Therefore,

I1[log(XtYlogt)+(Y+1logt)(au2tt)]2t/aducosh2u.I_{1}\leq\left[\log(Xt^{Y}\log t)+\left(Y+\frac{1}{\log t}\right)\left(\frac{-au-2t}{t}\right)\right]\int_{-\infty}^{-2t/a}\frac{\text{d}u}{\cosh^{2}u}. (4.12)

Next, if 2t/au(t+t0)/a-2t/a\leq u\leq-(t+t_{0})/a, then t+au[t,t0]t+au\in[-t,-t_{0}], so by the lemma’s assumption, we have |ζ(σ+i(t+au))|XtYlogt|\zeta(\sigma+i(t+au))|\leq Xt^{Y}\log t and

I2log(XtYlogt)2t/a(t+t0)/aducosh2uI_{2}\leq\log(Xt^{Y}\log t)\int_{-2t/a}^{-(t+t_{0})/a}\frac{\text{d}u}{\cosh^{2}u} (4.13)

If (t+t0)/au(tt0)/a-(t+t_{0})/a\leq u\leq-(t-t_{0})/a, then t+au[t0,t0]t+au\in[-t_{0},t_{0}] and we use

ζ(s)=1s1+12+s1xx+1/2xs+1dx.\zeta(s)=\frac{1}{s-1}+\frac{1}{2}+s\int_{1}^{\infty}\frac{\lfloor x\rfloor-x+1/2}{x^{s+1}}\text{d}x.

Letting s=σ+i(t+au)s=\sigma+i(t+au), we have |s1||(s1)|t1|s-1|\geq|\Re(s-1)|\geq t^{-1}. Also, |s|=σ2+(t+au)2<12+t02|s|=\sqrt{\sigma^{2}+(t+au)^{2}}<\sqrt{1^{2}+t_{0}^{2}}. Using σ1/2\sigma\geq 1/2 and t01t_{0}\geq 1, we have

|ζ(s)|1|s1|+12+|s|21dxxσ+1t+12+12σ1+t02<t+t0+1,|\zeta(s)|\leq\frac{1}{|s-1|}+\frac{1}{2}+\frac{|s|}{2}\int_{1}^{\infty}\frac{\text{d}x}{x^{\sigma+1}}\leq t+\frac{1}{2}+\frac{1}{2\sigma}\sqrt{1+t_{0}^{2}}<t+t_{0}+1,

so that

I3log(t+t0+1)(t+t0)/a(tt0)/aducosh2u.I_{3}\leq\log(t+t_{0}+1)\int_{-(t+t_{0})/a}^{-(t-t_{0})/a}\frac{\text{d}u}{\cosh^{2}u}. (4.14)

Finally, if u(t0t)/au\geq(t_{0}-t)/a, then t+au[t0,)t+au\in[t_{0},\infty). By the lemma’s assumption, and applying log(x+1)x\log(x+1)\leq x and log(x+1)x12x2+13x3\log(x+1)\leq x-\frac{1}{2}x^{2}+\frac{1}{3}x^{3} with x=au/t>1x=au/t>-1, gives

log|ζ(σ+i(t+au))|\displaystyle\log|\zeta(\sigma+i(t+au))| logX+Ylog(t+au)+loglog(t+au)\displaystyle\leq\log X+Y\log(t+au)+\log\log(t+au)
log(XtYlogt)+(Y+1logt)(aut(au)22t2+(au)33t3)\displaystyle\leq\log(Xt^{Y}\log t)+\left(Y+\frac{1}{\log t}\right)\left(\frac{au}{t}-\frac{(au)^{2}}{2t^{2}}+\frac{(au)^{3}}{3t^{3}}\right)

so that

I4[log(XtYlogt)+(Y+1logt)(aut(au)22t2+(au)33t3)](t0t)/aducosh2uI_{4}\leq\left[\log(Xt^{Y}\log t)+\left(Y+\frac{1}{\log t}\right)\left(\frac{au}{t}-\frac{(au)^{2}}{2t^{2}}+\frac{(au)^{3}}{3t^{3}}\right)\right]\int_{(t_{0}-t)/a}^{\infty}\frac{\text{d}u}{\cosh^{2}u} (4.15)

Combining (4.12), (4.13), (4.14) and (4.15), we have

log|ζ(σ+i(t+au))|cosh2udu<2log(XtYlogt)+E,\int_{-\infty}^{\infty}\frac{\log|\zeta(\sigma+i(t+au))|}{\cosh^{2}u}\text{d}u<2\log(Xt^{Y}\log t)+E,
E:=log(t+t0+1)(t+t0)/a(tt0)/aducosh2u+(Y+1logt)((t0t)/aaut(au)22t2+(au)33t3cosh2udu2t/aau+2ttcosh2udu)\begin{split}E&:=\log(t+t_{0}+1)\int_{-(t+t_{0})/a}^{-(t-t_{0})/a}\frac{\text{d}u}{\cosh^{2}u}\\ &\qquad+\left(Y+\frac{1}{\log t}\right)\left(\int_{(t_{0}-t)/a}^{\infty}\frac{\frac{au}{t}-\frac{(au)^{2}}{2t^{2}}+\frac{(au)^{3}}{3t^{3}}}{\cosh^{2}u}\text{d}u-\int_{-\infty}^{-2t/a}\frac{\frac{au+2t}{t}}{\cosh^{2}u}\text{d}u\right)\end{split} (4.16)

However, since (t+t0)/a<(tt0)/a<0-(t+t_{0})/a<-(t-t_{0})/a<0,

(t+t0)/a(tt0)/aducosh2u2t0acosh2(tt0a).\int_{-(t+t_{0})/a}^{-(t-t_{0})/a}\frac{\text{d}u}{\cosh^{2}u}\leq\frac{2t_{0}}{a\cosh^{2}\left(\frac{t-t_{0}}{a}\right)}.

Using a1/2a\leq 1/2, t3t0t\geq 3t_{0}, t0100t_{0}\geq 100 and cosh2u14e2|u|\cosh^{2}u\geq\frac{1}{4}e^{2|u|}, the first term on the RHS of (4.16) is majorised by

t0log(t+t0+1)acosh2(tt0a)4t0log(t+t0+1)ae2(tt0)/aet/a8t0log(4t0+1)e2t0<et/a.\frac{t_{0}\log(t+t_{0}+1)}{a\cosh^{2}(\frac{t-t_{0}}{a})}\leq\frac{4t_{0}\log(t+t_{0}+1)}{ae^{2(t-t_{0})/a}}\leq e^{-t/a}\frac{8t_{0}\log(4t_{0}+1)}{e^{2t_{0}}}<e^{-t/a}. (4.17)

Next, letting v=au2tv=-au-2t and using cosh2u14e2|u|\cosh^{2}u\geq\frac{1}{4}e^{2|u|},

2t/aaut2cosh2udu2t/aaut2e2udu=4ate4t/a0ve2vdv=ate4t/a.\int_{-\infty}^{-2t/a}\frac{-\frac{au}{t}-2}{\cosh^{2}u}\text{d}u\leq\int_{-\infty}^{-2t/a}\frac{-\frac{au}{t}-2}{e^{-2u}}\text{d}u=\frac{4a}{t}e^{-4t/a}\int_{0}^{\infty}ve^{-2v}\text{d}v=\frac{a}{t}e^{-4t/a}. (4.18)

Finally, if x=autx=\frac{au}{t} and utt0au\geq\frac{t-t_{0}}{a} then x1t0t23x\geq 1-\frac{t_{0}}{t}\geq\frac{2}{3} and that x+x22+x33103x3x+\frac{x^{2}}{2}+\frac{x^{3}}{3}\leq\frac{10}{3}x^{3}. Hence

(t0t)/aaut(au)22t2+(au)33t3cosh2uduaut(au)22t2+(au)33t3cosh2udu+(tt0)/aaut+(au)22t2+(au)33t314e2uduπ212a2t2+40a33t3(tt0)/au3e2udu<π212a2t2+10100a3t3,\begin{split}&\int_{(t_{0}-t)/a}^{\infty}\frac{\frac{au}{t}-\frac{(au)^{2}}{2t^{2}}+\frac{(au)^{3}}{3t^{3}}}{\cosh^{2}u}\text{d}u\\ &\qquad\qquad\leq\int_{-\infty}^{\infty}\frac{\frac{au}{t}-\frac{(au)^{2}}{2t^{2}}+\frac{(au)^{3}}{3t^{3}}}{\cosh^{2}u}\text{d}u+\int_{(t-t_{0})/a}^{\infty}\frac{\frac{au}{t}+\frac{(au)^{2}}{2t^{2}}+\frac{(au)^{3}}{3t^{3}}}{\frac{1}{4}e^{2u}}\text{d}u\\ &\qquad\qquad\leq-\frac{\pi^{2}}{12}\frac{a^{2}}{t^{2}}+\frac{40a^{3}}{3t^{3}}\int_{(t-t_{0})/a}^{\infty}u^{3}e^{-2u}\text{d}u<-\frac{\pi^{2}}{12}\frac{a^{2}}{t^{2}}+10^{-100}\frac{a^{3}}{t^{3}},\end{split} (4.19)

where the last inequality follows from (tt0)/a2t0/a400(t-t_{0})/a\geq 2t_{0}/a\geq 400 and evaluating the integral explicitly. Combining (4.17), (4.18) and (4.19), we have

Eet/a+(Y+1logt)(π212a2t2+ate4t/a+10100a3t3)<et/a1logta22t2<0.E\leq e^{-t/a}+\left(Y+\frac{1}{\log t}\right)\left(-\frac{\pi^{2}}{12}\frac{a^{2}}{t^{2}}+\frac{a}{t}e^{-4t/a}+10^{-100}\frac{a^{3}}{t^{3}}\right)<e^{-t/a}-\frac{1}{\log t}\frac{a^{2}}{2t^{2}}<0. (4.20)

We use the above lemma in two forms - the first (Lemma 4.1 below) is used for large tt and the second (Lemma 4.2 below) is used for small values of tt.

Lemma 4.2.

Let k4k\geq 4 be an integer, 0<a1/20<a\leq 1/2 and tmax{31012,2k}t\geq\max\{3\cdot 10^{12},2^{k}\}. Then

12log|ζ(1ηk+it+iau)|cosh2udulogt2k2+loglogt+log1.546.\frac{1}{2}\int_{-\infty}^{\infty}\frac{\log|\zeta(1-\eta_{k}+it+iau)|}{\cosh^{2}u}\text{d}u\leq\frac{\log t}{2^{k}-2}+\log\log t+\log 1.546.
Proof.

Follows by taking t0=1012t_{0}=10^{12}, σ=1ηk\sigma=1-\eta_{k}, X=1.546X=1.546 and Y=1/(2k2)Y=1/(2^{k}-2) in Lemma 4.1, and using Theorem 1.1. The condition that σ1t1\sigma\leq 1-t^{-1} is satisfied since for k4k\geq 4, ηk2kt1\eta_{k}\geq 2^{-k}\geq t^{-1}. ∎

Lemma 4.3.

For all 2/7η1/22/7\leq\eta\leq 1/2, 0<a1/20<a\leq 1/2 and t31012t\geq 3\cdot 10^{12}, we have

12log|ζ(1η+it+iau)|cosh2udu8η118logt+loglogt+1.6594.279η.\frac{1}{2}\int_{-\infty}^{\infty}\frac{\log|\zeta(1-\eta+it+iau)|}{\cosh^{2}u}\text{d}u\leq\frac{8\eta-1}{18}\log t+\log\log t+1.659-4.279\eta.
Proof.

Let s=σ+its=\sigma+it. We begin with the estimates

|(s1)ζ(s)|0.618|Q+s|7/6log|Q+s|,s=1/2,|(s-1)\zeta(s)|\leq 0.618|Q+s|^{7/6}\log|Q+s|,\qquad\Re s=1/2,
|(s1)ζ(s)|1.546|Q+s|15/14log|Q+s|,s=5/7|(s-1)\zeta(s)|\leq 1.546|Q+s|^{15/14}\log|Q+s|,\qquad\Re s=5/7

with Q=1.31Q=1.31. These bounds are verified numerically for |t|3|t|\leq 3, then combined with [HPY24, Thm. 1.1] and the case k=4k=4 in Theorem 1.1, respectively. Applying Lemma 3.1, for all 1/2σ5/71/2\leq\sigma\leq 5/7,

|ζ(s)|0.618(1014σ)/31.546(14σ7)/3|Q+s|(258σ)/18log|Q+s||s1|.|\zeta(s)|\leq 0.618^{(10-14\sigma)/3}1.546^{(14\sigma-7)/3}\frac{|Q+s|^{(25-8\sigma)/18}\log|Q+s|}{|s-1|}. (4.21)

Next, we combine the estimates (for 1/2σ5/71/2\leq\sigma\leq 5/7 and tt0t\geq t_{0})

|Q+s|(258σ)/18|s1|<((Q+σ)2+t2)(258σ)/36t((Q+57t0)2+1)7/12t(78σ)/18,\frac{|Q+s|^{(25-8\sigma)/18}}{|s-1|}<\frac{\left((Q+\sigma)^{2}+t^{2}\right)^{(25-8\sigma)/36}}{t}\leq\left(\left(\frac{Q+\frac{5}{7}}{t_{0}}\right)^{2}+1\right)^{7/12}t^{(7-8\sigma)/18},
log|Q+s|log(Q+σ)2t02+1+logt(1+1logt0log(Q+57)2t02+1)logt\log|Q+s|\leq\log\sqrt{\frac{(Q+\sigma)^{2}}{t_{0}^{2}}+1}+\log t\leq\left(1+\frac{1}{\log t_{0}}\log\sqrt{\frac{(Q+\frac{5}{7})^{2}}{t_{0}^{2}}+1}\right)\log t

to obtain

|ζ(s)|C1C2σt(78σ)/18logt,tt0,12σ57|\zeta(s)|\leq C_{1}\,C_{2}^{\sigma}\,t^{(7-8\sigma)/18}\log t,\qquad t\geq t_{0},\frac{1}{2}\leq\sigma\leq\frac{5}{7} (4.22)

where

C1:=0.61810/31.5467/3((Q+57t0)2+1)7/12(1+12logt0log((Q+57)2t02+1)),C_{1}:=\frac{0.618^{10/3}}{1.546^{7/3}}\left(\left(\frac{Q+\frac{5}{7}}{t_{0}}\right)^{2}+1\right)^{7/12}\left(1+\frac{1}{2\log t_{0}}\log\left(\frac{(Q+\frac{5}{7})^{2}}{t_{0}^{2}}+1\right)\right), (4.23)
C2:=(1.5460.618)14/3.C_{2}:=\left(\frac{1.546}{0.618}\right)^{14/3}. (4.24)

Substituting t0=1012t_{0}=10^{12}, Q=1.31Q=1.31, σ=1η\sigma=1-\eta gives

log|ζ(1η+it)|1.6594.279η+8η118logt+loglogt,tt0.\log|\zeta(1-\eta+it)|\leq 1.659-4.279\eta+\frac{8\eta-1}{18}\log t+\log\log t,\qquad t\geq t_{0}.

The result follows from applying Lemma 4.1. ∎

Next, we slightly modify a lemma due to Ford [For02a], which bounds N(t,η)N(t,\eta), the number of zeroes ρ\rho satisfying |1+itρ|η|1+it-\rho|\leq\eta (counted with multiplicity). The main change is to use sharper bounds on ζ(s)\zeta(s) to the right of the 1-line, and to change the constants so that the lemma continues to hold for 1/4η2/71/4\leq\eta\leq 2/7. Note that a result like Lemma 4.1 can be used to produce a sharper bound for small values of tt, however we do not pursue this here for sake of brevity.

Lemma 4.4.

For 0<η2/70<\eta\leq 2/7 and t100t\geq 100, we have

N(t,η)5.9975η3/2logt+6.12+23loglogtlogη1.879.N(t,\eta)\leq 5.9975\eta^{3/2}\log t+6.12+\frac{\frac{2}{3}\log\log t-\log\eta}{1.879}. (4.25)
Proof.

We follow the argument of [For02a, Lem. 4.2]. In place of [For02a, (4.1)] we use a result appearing in [BR02] that

ζ(σ)eγ(σ1)σ1,σ>1,\zeta(\sigma)\leq\frac{e^{\gamma(\sigma-1)}}{\sigma-1},\qquad\sigma>1, (4.26)

where γ=0.5772\gamma=0.5772\ldots is the Euler-Mascheroni constant, to obtain

ζ(1+3.1421η)e3.1421γη3.1421η.\zeta(1+3.1421\eta)\leq\frac{e^{3.1421\gamma\eta}}{3.1421\eta}. (4.27)

Note that we are using η\eta in place of RR in [For02a, Lem. 4.2]. The constant 3.1421 is chosen so as to minimise the first term on the RHS of (4.25). Then, using [For02a, Lem. 4.1] with s=1+0.6421ηs=1+0.6421\eta,

10.6421η\displaystyle-\frac{1}{0.6421\eta} |ζζ(1+0.6421η+it)|ζζ(1+0.6421η+it)\displaystyle\leq-\left|\frac{\zeta^{\prime}}{\zeta}(1+0.6421\eta+it)\right|\leq-\Re\frac{\zeta^{\prime}}{\zeta}(1+0.6421\eta+it)
0.3758N(t,η)η+15η(23loglogt+(1.8579η)3/2Blogt\displaystyle\leq-\frac{0.3758N(t,\eta)}{\eta}+\frac{1}{5\eta}\Big{(}\frac{2}{3}\log\log t+(1.8579\eta)^{3/2}B\log t
+logA+3.1421γηlog3.1421η)\displaystyle\qquad\qquad+\log A+3.1421\gamma\eta-\log 3.1421\eta\Big{)}

However, for η2/7\eta\leq 2/7 we have 3.1421γηlog3.1421η0.62673.1421\gamma\eta-\log 3.1421\eta\leq-0.6267, so the result follows from substituting A=76.2A=76.2 and B=4.45B=4.45 originating from (4.8).222Here we applied (4.8) with σ=11.8579η\sigma=1-1.8579\eta where 0<η2/70<\eta\leq 2/7. Therefore, we need to extend the range of validity of this bound to σ0.4691\sigma\geq 0.4691\ldots. This is permissible by examining the proof of [For02, Lem. 7.1] (in fact, (4.8) is sharpest near σ=1\sigma=1).

Next, we use the above lemma to bound a sum over zeroes away from 1+it1+it, reproduced below. This result is the same as [For02a, Lem 4.3], except we use a sharper estimate of S(t)S(t), due to Trudgian [Tru14a], and extend the range of permissible values of η\eta up to 2/72/7. We note that for large tt, the estimate of S(t)S(t) due to [HSW21] is sharper, however our results are most sensitive to sharpness of the estimate for “small” tt. Using Theorem 1.1, it is possible to further refine the arguments of [Tru14a] and [HSW21] to improve bounds on S(t)S(t), and thus improve Lemma 4.5. We leave such considerations for possible future work (see remarks in §5).

Lemma 4.5.

Let t31012t\geq 3\cdot 10^{12}. If 0<η2/70<\eta\leq 2/7, then

|1+itρ|η1|1+itρ|2(23.99η40.385)logt+(0.3548η2+1.2031)loglogt40.236+5.86η2logη1.879η2N(t,η)η2.\begin{split}\sum_{|1+it-\rho|\geq\eta}\frac{1}{|1+it-\rho|^{2}}&\leq\left(\frac{23.99}{\sqrt{\eta}}-40.385\right)\log t+\left(\frac{0.3548}{\eta^{2}}+1.2031\right)\log\log t\\ &\qquad-40.236+\frac{5.86}{\eta^{2}}-\frac{\log\eta}{1.879\eta^{2}}-\frac{N(t,\eta)}{\eta^{2}}.\end{split}
Proof.

We follow the approach taken by Ford [For02a, Lem. 4.3]. We divide the zeroes satisfying |1+itρ|η|1+it-\rho|\geq\eta into the following sets

Z1:={ρ:|ρt|δ},Z2:={ρ:ρZ1,|1+itρ|η0,|itρ|η0},Z3:={ρ:ρZ1,ρZ2,|1+itρ|η}\begin{split}Z_{1}&:=\{\rho:|\Im\rho-t|\geq\delta\},\\ Z_{2}&:=\{\rho:\rho\notin Z_{1},|1+it-\rho|\geq\eta_{0},|it-\rho|\geq\eta_{0}\},\\ Z_{3}&:=\{\rho:\rho\notin Z_{1},\rho\notin Z_{2},|1+it-\rho|\geq\eta\}\end{split} (4.28)

for some δη0>0\delta\geq\eta_{0}>0 to be chosen later. We will also assume that δ1\delta\leq 1, this will be verified after choosing δ\delta. Let N(t)N(t) denote the number of zeroes ρ\rho of ζ(s)\zeta(s), with multiplicity, satisfying 0<ρ<t0<\Im\rho<t. We use the following result, obtained by using [PT15, Cor. 1] in place of [Tru14a, Thm. 1] in the proof of [Tru14a, Cor. 1]: 333The proof of [PT15, Cor. 1] uses [PT15, Thm. 1], which in turn uses an erroneous lemma [CG04, Lem. 3] (see e.g. [Pat21, Pat22, For22, HPY24] for a discussion). However, [PT15, Thm. 1] has since been proved independently in [HPY24, Thm. 1.1], so we may use it here.

|N(t)t2πlogt2πe78|0.11logt+0.29loglogt+2.29+0.2t0,tt0e.\left|N(t)-\frac{t}{2\pi}\log\frac{t}{2\pi e}-\frac{7}{8}\right|\leq 0.11\log t+0.29\log\log t+2.29+\frac{0.2}{t_{0}},\qquad t\geq t_{0}\geq e. (4.29)

Therefore, for t14t\geq 14

N(t)=t2πlogt2πe+78+Q(t),N(t)=\frac{t}{2\pi}\log\frac{t}{2\pi e}+\frac{7}{8}+Q(t), (4.30)

for some Q(t)Q(t) satisfying

|Q(t)|0.11logt+0.29loglogt+2.305.|Q(t)|\leq 0.11\log t+0.29\log\log t+2.305. (4.31)

Let

Sj:=ρZj1|1+itρ|2,j=1,2,3.S_{j}:=\sum_{\rho\in Z_{j}}\frac{1}{|1+it-\rho|^{2}},\qquad j=1,2,3. (4.32)

Then, since there are no zeroes with |ρ|14|\Im\rho|\leq 14, and using |1+itρ||tρ||1+it-\rho|\geq|t-\Im\rho|,

S1t+δdN(u)(ut)2+14tδdN(u)(ut)2+14dN(u)(u+t)2=I1+I2+I3,S_{1}\leq\int_{t+\delta}^{\infty}\frac{\text{d}N(u)}{(u-t)^{2}}+\int_{14}^{t-\delta}\frac{\text{d}N(u)}{(u-t)^{2}}+\int_{14}^{\infty}\frac{\text{d}N(u)}{(u+t)^{2}}=I_{1}+I_{2}+I_{3},

where, since dN(u)=12πlogu2π+dQ(u)\text{d}N(u)=\frac{1}{2\pi}\log\frac{u}{2\pi}+\text{d}Q(u)

I1=12π[(δ1+t1)log(t+δ)logδtlog2πδ]+δdQ(t+x)x2\displaystyle I_{1}=\frac{1}{2\pi}\left[(\delta^{-1}+t^{-1})\log(t+\delta)-\frac{\log\delta}{t}-\frac{\log 2\pi}{\delta}\right]+\int_{\delta}^{\infty}\frac{\text{d}Q(t+x)}{x^{2}}

and

δdQ(t+x)x2=[Q(t+x)x2]δ+2δQ(t+x)x3dx.\displaystyle\int_{\delta}^{\infty}\frac{\text{d}Q(t+x)}{x^{2}}=\left[\frac{Q(t+x)}{x^{2}}\right]_{\delta}^{\infty}+2\int_{\delta}^{\infty}\frac{Q(t+x)}{x^{3}}\text{d}x.

Since log(1+u)u\log(1+u)\leq u if u0u\geq 0, we have that log(t+x)logt=log(1+xt)xt\log(t+x)-\log t=\log(1+\frac{x}{t})\leq\frac{x}{t} and

loglog(t+x)loglogt=log(log(t+x)logt)=log(1+xtlogt)<xtlogt.\log\log(t+x)-\log\log t=\log\left(\frac{\log(t+x)}{\log t}\right)=\log\left(1+\frac{x}{t\log t}\right)<\frac{x}{t\log t}.

Therefore, for δ1\delta\leq 1 and t10000t\geq 10000,

|Q(t+δ)|\displaystyle|Q(t+\delta)| 0.11log(t+δ)+0.29loglog(t+δ)+2.305\displaystyle\leq 0.11\log(t+\delta)+0.29\log\log(t+\delta)+2.305
0.11(logt+δt)+0.29(loglogt+δtlogt)+2.305\displaystyle\leq 0.11\left(\log t+\frac{\delta}{t}\right)+0.29\left(\log\log t+\frac{\delta}{t\log t}\right)+2.305
0.11logt+0.29loglogt+2.306.\displaystyle\leq 0.11\log t+0.29\log\log t+2.306.

Furthermore,

δ|Q(t+x)|x3dx\displaystyle\int_{\delta}^{\infty}\frac{|Q(t+x)|}{x^{3}}\text{d}x δ(0.11logt+0.29loglogt+2.305x3+(0.11t+0.29tlogt)δx2)dx\displaystyle\leq\int_{\delta}^{\infty}\left(\frac{0.11\log t+0.29\log\log t+2.305}{x^{3}}+\left(\frac{0.11}{t}+\frac{0.29}{t\log t}\right)\frac{\delta}{x^{2}}\right)\text{d}x
12δ2(0.11logt+0.29loglogt+2.306).\displaystyle\leq\frac{1}{2}\delta^{-2}(0.11\log t+0.29\log\log t+2.306).

Finally, if δ1\delta\leq 1 then logδ0\log\delta\leq 0 and for tt0t\geq t_{0}, we have

12π[(δ1+t1)log(t+δ)logδtlog2πδ]δ1+t012π(logt+δt0)logδt0log2π2πδ.\frac{1}{2\pi}\left[\left(\delta^{-1}+t^{-1}\right)\log(t+\delta)-\frac{\log\delta}{t}-\frac{\log 2\pi}{\delta}\right]\leq\frac{\delta^{-1}+t_{0}^{-1}}{2\pi}\left(\log t+\frac{\delta}{t_{0}}\right)-\frac{\log\delta}{t_{0}}-\frac{\log 2\pi}{2\pi\delta}.

Therefore

I1A1logt+B1loglogt+C1,I_{1}\leq A_{1}^{\prime}\log t+B_{1}^{\prime}\log\log t+C_{1}^{\prime}, (4.33)
A1=0.22δ2+δ1+t012π,B1=0.58δ2,C1=4.612δ2+1+δ/t02πt0logδt0log2π2πδ.A_{1}^{\prime}=0.22\delta^{-2}+\frac{\delta^{-1}+t_{0}^{-1}}{2\pi},\quad B_{1}^{\prime}=0.58\delta^{-2},\quad C_{1}^{\prime}=4.612\delta^{-2}+\frac{1+\delta/t_{0}}{2\pi t_{0}}-\frac{\log\delta}{t_{0}}-\frac{\log 2\pi}{2\pi\delta}.

Next,

I2\displaystyle I_{2} =12π14tδ1(ut)2logu2πdu+14tδdQ(u)(ut)2\displaystyle=\frac{1}{2\pi}\int_{14}^{t-\delta}\frac{1}{(u-t)^{2}}\log\frac{u}{2\pi}\text{d}u+\int_{14}^{t-\delta}\frac{\text{d}Q(u)}{(u-t)^{2}}
12πlogt2π14tδdu(ut)2+[Q(u)(ut)2]14tδ+214tδQ(u)(ut)3du\displaystyle\leq\frac{1}{2\pi}\log\frac{t}{2\pi}\int_{14}^{t-\delta}\frac{\text{d}u}{(u-t)^{2}}+\left[\frac{Q(u)}{(u-t)^{2}}\right]_{14}^{t-\delta}+2\int_{14}^{t-\delta}\frac{Q(u)}{(u-t)^{3}}\text{d}u
<δ12πlogt2π+δ2|Q(tδ)|+2|Q(tδ)|14tδdu|ut|3(since Q(14)>0)\displaystyle<\frac{\delta^{-1}}{2\pi}\log\frac{t}{2\pi}+\delta^{-2}|Q(t-\delta)|+2|Q(t-\delta)|\int_{14}^{t-\delta}\frac{\text{d}u}{|u-t|^{3}}\qquad\text{(since $Q(14)>0$)}
<δ12πlogt2π+δ2(0.22logt+0.58loglogt+4.61)\displaystyle<\frac{\delta^{-1}}{2\pi}\log\frac{t}{2\pi}+\delta^{-2}(0.22\log t+0.58\log\log t+4.61)
=A1′′logt+B1′′loglogt+C1′′,\displaystyle=A_{1}^{\prime\prime}\log t+B_{1}^{\prime\prime}\log\log t+C_{1}^{\prime\prime}, (4.34)

where

A1′′=δ12π+0.22δ2,B1′′=0.58δ2,C1′′=4.61δ2log2π2πδ.A_{1}^{\prime\prime}=\frac{\delta^{-1}}{2\pi}+0.22\delta^{-2},\qquad B_{1}^{\prime\prime}=0.58\delta^{-2},\qquad C_{1}^{\prime\prime}=4.61\delta^{-2}-\frac{\log 2\pi}{2\pi\delta}.

Next,

I312π14log(u+t2π)(u+t)2du+214|Q(u)|(u+t)3du0.00014.I_{3}\leq\frac{1}{2\pi}\int_{14}^{\infty}\frac{\log\left(\frac{u+t}{2\pi}\right)}{(u+t)^{2}}\text{d}u+2\int_{14}^{\infty}\frac{|Q(u)|}{(u+t)^{3}}\text{d}u\leq 0.00014. (4.35)

Therefore, combining (4.33), (4.34) and (4.35),

S1A1logt+B1loglogt+C1S_{1}\leq A_{1}\log t+B_{1}\log\log t+C_{1} (4.36)

where

A1\displaystyle A_{1} :=0.44δ2+2δ1+t012π,B1=1.16δ2,\displaystyle:=0.44\delta^{-2}+\frac{2\delta^{-1}+t_{0}^{-1}}{2\pi},\qquad B_{1}=1.16\delta^{-2},
C1=9.222δ2+1+δ/t02πt0logδt0log2ππδ+0.00014.C_{1}=9.222\delta^{-2}+\frac{1+\delta/t_{0}}{2\pi t_{0}}-\frac{\log\delta}{t_{0}}-\frac{\log 2\pi}{\pi\delta}+0.00014.

Next, if ν0:=(η02+(1η0)2)/2\nu_{0}:=(\eta_{0}^{-2}+(1-\eta_{0})^{-2})/2, then

S2max{4|Z2|,ν0|Z2|}=ν0|Z2|.S_{2}\leq\max\left\{4|Z_{2}|,\nu_{0}|Z_{2}|\right\}=\nu_{0}|Z_{2}|. (4.37)

We also have

|Z2|+|Z3|\displaystyle|Z_{2}|+|Z_{3}| =N(t+δ)N(tδ)N(t,η)\displaystyle=N(t+\delta)-N(t-\delta)-N(t,\eta)
=t+δ2πlogt+δ2πtδ2πlogtδ2πδπ+Q(t+δ)Q(tδ)N(t,η)\displaystyle=\frac{t+\delta}{2\pi}\log\frac{t+\delta}{2\pi}-\frac{t-\delta}{2\pi}\log\frac{t-\delta}{2\pi}-\frac{\delta}{\pi}+Q(t+\delta)-Q(t-\delta)-N(t,\eta)

If f(t):=(t/2π)log(t/2π)f(t):=(t/2\pi)\log(t/2\pi), then, since f(t)f^{\prime}(t) is increasing,

t+δ2πlogt+δ2πtδ2πlogtδ2π=f(t+δ)f(tδ)2δf(t+δ)\displaystyle\frac{t+\delta}{2\pi}\log\frac{t+\delta}{2\pi}-\frac{t-\delta}{2\pi}\log\frac{t-\delta}{2\pi}=f(t+\delta)-f(t-\delta)\leq 2\delta f^{\prime}(t+\delta)
=δπ(log(t+δ)log2π+1)<δπlogt+δπ(δt0log2π+1)\displaystyle\qquad\qquad\qquad\qquad=\frac{\delta}{\pi}\left(\log(t+\delta)-\log 2\pi+1\right)<\frac{\delta}{\pi}\log t+\frac{\delta}{\pi}\left(\frac{\delta}{t_{0}}-\log 2\pi+1\right)

and, since logt\log t and loglogt\log\log t are both concave, we have log(tδ)+log(t+δ)2logt\log(t-\delta)+\log(t+\delta)\leq 2\log t and loglog(tδ)+loglog(t+δ)2loglogt\log\log(t-\delta)+\log\log(t+\delta)\leq 2\log\log t by Jensen’s inequality. Hence

Q(t+δ)Q(tδ)\displaystyle Q(t+\delta)-Q(t-\delta) 0.22logt+0.58loglogt+4.61\displaystyle\leq 0.22\log t+0.58\log\log t+4.61

so that

|Z2|+|Z3|Alogt+Bloglogt+CN(t,η),|Z_{2}|+|Z_{3}|\leq A^{\prime}\log t+B^{\prime}\log\log t+C^{\prime}-N(t,\eta), (4.38)

where

A=δπ+0.22,B=0.58C=δπ(δt0log2π)+4.61.A^{\prime}=\frac{\delta}{\pi}+0.22,\qquad B^{\prime}=0.58\qquad C^{\prime}=\frac{\delta}{\pi}\left(\frac{\delta}{t_{0}}-\log 2\pi\right)+4.61.

Next, since N3+N(t,η)=2N(t,η0)N_{3}+N(t,\eta)=2N(t,\eta_{0}),

S3N(t,η0)(1η0)2+ηη0dN(t,u)u2=1(1η0)2N(t,η0)+[N(t,u)u2]ηη0+2ηη0N(t,u)u3du=2ν0N(t,η0)N(t,η)η2+2ηη0N(t,u)u3du=ν0|Z3|+(ν01η2)N(t,η)+2ηη0N(t,u)u3du\begin{split}S_{3}&\leq\frac{N(t,\eta_{0})}{(1-\eta_{0})^{2}}+\int_{\eta}^{\eta_{0}}\frac{\text{d}N(t,u)}{u^{2}}=\frac{1}{(1-\eta_{0})^{2}}N(t,\eta_{0})+\left[\frac{N(t,u)}{u^{2}}\right]_{\eta}^{\eta_{0}}+2\int_{\eta}^{\eta_{0}}\frac{N(t,u)}{u^{3}}\text{d}u\\ &=2\nu_{0}N(t,\eta_{0})-\frac{N(t,\eta)}{\eta^{2}}+2\int_{\eta}^{\eta_{0}}\frac{N(t,u)}{u^{3}}\text{d}u\\ &=\nu_{0}|Z_{3}|+\left(\nu_{0}-\frac{1}{\eta^{2}}\right)N(t,\eta)+2\int_{\eta}^{\eta_{0}}\frac{N(t,u)}{u^{3}}\text{d}u\end{split} (4.39)

and

2ηη0N(t,u)u3du\displaystyle 2\int_{\eta}^{\eta_{0}}\frac{N(t,u)}{u^{3}}\text{d}u 2ηη05.9975u3/2logt+6.12+23loglogtlogu1.879u3du\displaystyle\leq 2\int_{\eta}^{\eta_{0}}\frac{5.9975u^{3/2}\log t+6.12+\frac{\frac{2}{3}\log\log t-\log u}{1.879}}{u^{3}}\text{d}u
23.99(η1/2η01/2)logt+(η2η02)(5.86+0.3548loglogt)\displaystyle\leq 23.99\left(\eta^{-1/2}-\eta_{0}^{-1/2}\right)\log t+\left(\eta^{-2}-\eta_{0}^{-2}\right)(5.86+0.3548\log\log t)
+11.879(logη0η02logηη2).\displaystyle\qquad\qquad+\frac{1}{1.879}\left(\frac{\log\eta_{0}}{\eta_{0}^{2}}-\frac{\log\eta}{\eta^{2}}\right).

Therefore, combining (4.37), (4.38) and (4.39),

S2+S3ν0(Alogt+Bloglogt+C)+23.99(η1/2η01/2)logt+(η2η02)(5.86+0.3548loglogt)+11.879(logη0η02logηη2)N(t,η)η2.\begin{split}S_{2}+S_{3}&\leq\nu_{0}\left(A^{\prime}\log t+B^{\prime}\log\log t+C^{\prime}\right)+23.99\left(\eta^{-1/2}-\eta_{0}^{-1/2}\right)\log t\\ &+\left(\eta^{-2}-\eta_{0}^{-2}\right)(5.86+0.3548\log\log t)+\frac{1}{1.879}\left(\frac{\log\eta_{0}}{\eta_{0}^{2}}-\frac{\log\eta}{\eta^{2}}\right)-\frac{N(t,\eta)}{\eta^{2}}.\end{split} (4.40)

Hence finally, combining (4.36) and (4.40),

|1+itρ|η1|1+itρ|2A′′logt+B′′loglogt+C′′N(t,η)η2,\sum_{|1+it-\rho|\geq\eta}\frac{1}{|1+it-\rho|^{2}}\leq A^{\prime\prime}\log t+B^{\prime\prime}\log\log t+C^{\prime\prime}-\frac{N(t,\eta)}{\eta^{2}},

where

A′′=0.44δ2+2δ1+t012π+ν0(δπ+0.22)+23.99(η1/2η01/2),A^{\prime\prime}=0.44\delta^{-2}+\frac{2\delta^{-1}+t_{0}^{-1}}{2\pi}+\nu_{0}\left(\frac{\delta}{\pi}+0.22\right)+23.99\left(\eta^{-1/2}-\eta_{0}^{-1/2}\right),
B′′=1.16δ2+0.58ν0+0.3548(η2η02),B^{\prime\prime}=1.16\delta^{-2}+0.58\nu_{0}+0.3548\left(\eta^{-2}-\eta_{0}^{-2}\right),
C′′\displaystyle C^{\prime\prime} =9.222δ2+1+δ/t02πt0logδt0log2ππδ+ν0(δπ(δt0log2π)+4.61)\displaystyle=9.222\delta^{-2}+\frac{1+\delta/t_{0}}{2\pi t_{0}}-\frac{\log\delta}{t_{0}}-\frac{\log 2\pi}{\pi\delta}+\nu_{0}\left(\frac{\delta}{\pi}\left(\frac{\delta}{t_{0}}-\log 2\pi\right)+4.61\right)
+5.86(η2η02)+11.879(logη0η02logηη2)+0.00014.\displaystyle\qquad\qquad+5.86\left(\eta^{-2}-\eta_{0}^{-2}\right)+\frac{1}{1.879}\left(\frac{\log\eta_{0}}{\eta_{0}^{2}}-\frac{\log\eta}{\eta^{2}}\right)+0.00014.

We choose t0=31012t_{0}=3\cdot 10^{12}, η0=2/7\eta_{0}=2/7, and δ=0.90114\delta=0.90114 to minimise the bound when η=2/7\eta=2/7, t=exp(1000)t=\exp(1000), which are close to the values most relevant to our application. The desired result follows. ∎

Remark.

Lemma 4.5 is ultimately used to bound a secondary term that is asymptotically smaller, as tt\to\infty, than the “main” term, which is bounded using Lemma 4.1. However, the values of tt we encounter are small enough that Lemma 4.5 noticeably influences the resulting constant of Corollary 1.2. For this reason we derive a second version of the result (Lemma 4.6 below) that is sharper for “large” values of η\eta.

Lemma 4.6.

If t31012t\geq 3\cdot 10^{12} and 2/7η1/22/7\leq\eta\leq 1/2, then

|1+itρ|η1|1+itρ|2(0.5576+0.6079ν)logt+(0.7813+0.58ν)loglogt+5.732+3.898νN(t,η)η2,\begin{split}\sum_{|1+it-\rho|\geq\eta}\frac{1}{|1+it-\rho|^{2}}&\leq(0.5576+0.6079\nu)\log t+(0.7813+0.58\nu)\log\log t\\ &\qquad+5.732+3.898\nu-\frac{N(t,\eta)}{\eta^{2}},\end{split}

where ν:=12(η2+(1η)2)\nu:=\frac{1}{2}(\eta^{-2}+(1-\eta)^{-2}).

Proof.

For some δ(1,2)\delta\in(1,2) to be chosen later (independent of the δ\delta parameter appearing in Lemma 4.5), define

Z1:={ρ:|ρt|δ},Z2:={ρ:ρZ1,|1+itρ|η,|itρ|η},Z3:={ρ:ρZ1,|itρ|<η}\begin{split}Z_{1}^{\prime}&:=\{\rho:|\Im\rho-t|\geq\delta\},\\ Z_{2}^{\prime}&:=\{\rho:\rho\notin Z_{1}^{\prime},|1+it-\rho|\geq\eta,|it-\rho|\geq\eta\},\\ Z_{3}^{\prime}&:=\{\rho:\rho\notin Z_{1}^{\prime},|it-\rho|<\eta\}\end{split} (4.41)

and let SjS_{j}^{\prime}, j=1,2,3j=1,2,3 be defined analogously to (4.32). This is equivalent to setting η0=η\eta_{0}=\eta in (4.28). We use a similar method as before to bound S1S_{1}^{\prime}, however, since δ\delta is now greater than 1, we drop the logδ/t0-\log\delta/t_{0} term in the bound of S1S_{1} to ensure the monotonicity argument remains valid. Explicitly, we have

S1A1logt+B1loglogt+C1S_{1}^{\prime}\leq A_{1}^{\prime}\log t+B_{1}^{\prime}\log\log t+C_{1}^{\prime} (4.42)

where

A1\displaystyle A_{1}^{\prime} :=0.44δ2+2δ1+t012π,B1=1.16δ2,\displaystyle:=0.44\delta^{-2}+\frac{2\delta^{-1}+t_{0}^{-1}}{2\pi},\qquad B_{1}^{\prime}=1.16\delta^{-2},
C1=9.222δ2+1+δ/t02πt0log2ππδ+0.00014.C_{1}^{\prime}=9.222\delta^{-2}+\frac{1+\delta/t_{0}}{2\pi t_{0}}-\frac{\log 2\pi}{\pi\delta}+0.00014.

Next, similarly to before,

S2max{4|Z2|,ν|Z2|}=ν|Z2|S_{2}^{\prime}\leq\max\left\{4|Z_{2}^{\prime}|,\nu|Z_{2}^{\prime}|\right\}=\nu|Z_{2}^{\prime}| (4.43)

and

|Z2|=N(t+δ)N(tδ)2N(t,η)A2logt+B2loglogt+C22N(t,η),\begin{split}|Z_{2}^{\prime}|&=N(t+\delta)-N(t-\delta)-2N(t,\eta)\\ &\leq A_{2}^{\prime}\log t+B_{2}^{\prime}\log\log t+C_{2}^{\prime}-2N(t,\eta),\end{split} (4.44)

where

A2=δπ+0.22,B2=0.58C2=δπ(δt0log2π)+4.61.A_{2}^{\prime}=\frac{\delta}{\pi}+0.22,\qquad B_{2}^{\prime}=0.58\qquad C_{2}^{\prime}=\frac{\delta}{\pi}\left(\frac{\delta}{t_{0}}-\log 2\pi\right)+4.61.

Furthermore, since each zero in Z3Z_{3} contributes at most (1η)2(1-\eta)^{-2} each to the sum S3S_{3}, and there are N(t,η)N(t,\eta) of them,

S3N(t,η)(1η)2.S_{3}^{\prime}\leq\frac{N(t,\eta)}{(1-\eta)^{2}}. (4.45)

Combining (4.42), (4.43), (4.44) and (4.45),

|1+itρ|η1|1+itρ|2A′′logt+B′′loglogt+C′′N(t,η)η2\begin{split}\sum_{|1+it-\rho|\geq\eta}\frac{1}{|1+it-\rho|^{2}}\leq A^{\prime\prime}\log t+B^{\prime\prime}\log\log t+C^{\prime\prime}-\frac{N(t,\eta)}{\eta^{2}}\end{split} (4.46)

where

A′′=0.44δ2+2δ1+t012π+ν(δπ+0.22),B′′=1.16δ2+0.58ν,A^{\prime\prime}=0.44\delta^{-2}+\frac{2\delta^{-1}+t_{0}^{-1}}{2\pi}+\nu\left(\frac{\delta}{\pi}+0.22\right),\qquad B^{\prime\prime}=1.16\delta^{-2}+0.58\nu,
C′′=9.222δ2+1+δ/t02πt0logδt0log2ππδ+0.00014+ν(δπ(δt0log2π)+4.61).C^{\prime\prime}=9.222\delta^{-2}+\frac{1+\delta/t_{0}}{2\pi t_{0}}-\frac{\log\delta}{t_{0}}-\frac{\log 2\pi}{\pi\delta}+0.00014+\nu\left(\frac{\delta}{\pi}\left(\frac{\delta}{t_{0}}-\log 2\pi\right)+4.61\right).

Choosing δ=1.2185\delta=1.2185 and t0=31012t_{0}=3\cdot 10^{12} gives the desired result. ∎

The following lemma, due to [For02a, Lem. 4.6] using the methods of [HB92, Lem. 5.1], provides an upper bound on a Dirichlet series “mollified” using a smoothing function ff.

Lemma 4.7 ([For02a] Lemma 4.6).

Let 0η1/20\leq\eta\leq 1/2, f(u)0f(u)\geq 0 be a real function with a continuous derivative, and a Laplace transform F(z):=0f(y)ezuduF(z):=\int_{0}^{\infty}f(y)e^{-zu}\text{d}u that is absolutely convergent for z>0\Re z>0. Let F0(z):=F(z)z1f(0)F_{0}(z):=F(z)-z^{-1}f(0) and suppose

|F0(z)|δ|z|2,z0,|z|η,|F_{0}(z)|\leq\frac{\delta}{|z|^{2}},\qquad\Re z\geq 0,\;|z|\geq\eta,

for some constant δ>0\delta>0. Then, if s=1+its=1+it with t1000t\geq 1000,

n1Λ(n)f(logn)ns|sρ|η{F(sρ)+f(0)(π2ηcot(π(sρ)2η)1sρ)}\displaystyle\Re\sum_{n\geq 1}\frac{\Lambda(n)f(\log n)}{n^{s}}\leq-\sum_{|s-\rho|\leq\eta}\Re\left\{F(s-\rho)+f(0)\left(\frac{\pi}{2\eta}\cot\left(\frac{\pi(s-\rho)}{2\eta}\right)-\frac{1}{s-\rho}\right)\right\}
+f(0)4ηlog|ζ(sη+2ηuiπ)|log|ζ(s+η+2ηuiπ)|cosh2udu\displaystyle\qquad+\frac{f(0)}{4\eta}\int_{-\infty}^{\infty}\frac{\log|\zeta(s-\eta+\frac{2\eta ui}{\pi})|-\log|\zeta(s+\eta+\frac{2\eta ui}{\pi})|}{\cosh^{2}u}\text{d}u
+δ{1.8+logt3+|sρ|η1|sρ|2},\displaystyle\qquad+\delta\left\{1.8+\frac{\log t}{3}+\sum_{|s-\rho|\geq\eta}\frac{1}{|s-\rho|^{2}}\right\},

and

n1Λ(n)f(logn)nF(0)+1.8δ.\displaystyle\Re\sum_{n\geq 1}\frac{\Lambda(n)f(\log n)}{n}\leq F(0)+1.8\delta.
Proof.

Follows by combining [For02a] Lemma 4.5 and Lemma 4.6. ∎

Remark.

As noted in [MTY24], in the above lemma it is possible to take tt0t\geq t_{0} with t0t_{0} much larger than 1000, however this has no noticeable effect in the final result.

The choice of the smoothing function ff in the above lemma has a direct impact on the size of the resulting zero-free region. Here, we choose ff the same way as Ford [For02a], which is in turn based on [HB92]. Such functions are derived via a calculus-of-variations argument and are provably optimal under certain assumptions. Jang and Kwon [JK14] presented an alternative smoothing function that produced better results for the classical zero-free region (see also [MTY24]), however this smoothing function did not lead to significant improvements in Corollary 1.2.

Given some λ>0\lambda>0 to be fixed later, define

f(u):=λeλug(λu),u0,f(u):=\lambda e^{\lambda u}g(\lambda u),\qquad u\geq 0, (4.47)

where, successively,

g(u):=(hh)(u)=0h(ut)h(t)dt,u0,g(u):=(h*h)(u)=\int_{0}^{\infty}h(u-t)h(t)\text{d}t,\qquad u\geq 0, (4.48)
h(u):={(cos(utanθ)cosθ)sec2θif |u|θcotθ0otherwise,h(u):=\begin{cases}(\cos(u\tan\theta)-\cos\theta)\sec^{2}\theta&\text{if }|u|\leq\theta\cot\theta\\ 0&\text{otherwise}\end{cases}, (4.49)

and θ=θ(b0,b1)\theta=\theta(b_{0},b_{1}) is defined as the unique solution to

sin2θ=b1b0(1θcotθ),0<θ<π2,\sin^{2}\theta=\frac{b_{1}}{b_{0}}(1-\theta\cot\theta),\qquad 0<\theta<\frac{\pi}{2}, (4.50)

where b0b_{0} and b1b_{1} are coefficients of the non-negative trigonometric polynomial (4.5). Via a direct computation, we have

θ=θ(b0,b1)=1.132693699,\theta=\theta(b_{0},b_{1})=1.132693699\ldots, (4.51)
f(0)=λg(0),f(0)=\lambda g(0),\\ (4.52)
g(0)=sec2θ(θtanθ+3θcotθ3)=5.610921922.\begin{split}g(0)&=\sec^{2}\theta(\theta\tan\theta+3\theta\cot\theta-3)=5.610921922\ldots.\end{split} (4.53)

We furthermore require information about the Laplace transforms of f(u)f(u) and g(u)g(u). Let

F(z):=0ezuf(u)du,G(z):=0ezug(u)duF(z):=\int_{0}^{\infty}e^{-zu}f(u)\text{d}u,\qquad G(z):=\int_{0}^{\infty}e^{-zu}g(u)\text{d}u (4.54)

denote the Laplace transforms of f(u)f(u) and g(u)g(u) respectively, and for convenience write

F0(z):=F(z)f(0)z,G0(z):=G(z)g(0)z.F_{0}(z):=F(z)-\frac{f(0)}{z},\qquad G_{0}(z):=G(z)-\frac{g(0)}{z}. (4.55)

The function G0(z)G_{0}(z) has an explicit formula (given in [For02a, (7.3)] 444Note that in [For02a], g(u)g(u), G(z)G(z) and G0(z)G_{0}(z) appear as w(u)w(u), W(z)W(z) and W0(z)W_{0}(z) respectively. We have renamed these functions to prevent confusion with W0(x)W_{0}(x), the product log function, which appears later.) in terms of θ\theta which can be used to bound |F0(z)||F_{0}(z)| via the relation

F0(z)=G(zλ1)λg(0)z=G0(zλ1)+λf(0)z(zλ).F_{0}(z)=G\left(\frac{z}{\lambda}-1\right)-\frac{\lambda g(0)}{z}=G_{0}\left(\frac{z}{\lambda}-1\right)+\frac{\lambda f(0)}{z(z-\lambda)}. (4.56)

Here we have used (4.47) and properties of the Laplace transform. Repeating the steps in [For02a, Lem. 7.1], we obtain, for all z1\Re z\geq-1 and |z|R3|z|\geq R\geq 3,

|F0(z)|cλf(0)|z|2,|F_{0}(z)|\leq\frac{c\lambda f(0)}{|z|^{2}}, (4.57)

where

c:=c(R)=H(R)(R+1)2R3w(0)+1+1Rc:=c(R)=\frac{H(R)(R+1)^{2}}{R^{3}w(0)}+1+\frac{1}{R} (4.58)

and

H(R):=c0(1tan2θ/R2)2(c2(R+1)R3(e2θ/tanθ+1)+c1R2+c3),H(R):=\frac{c_{0}}{(1-\tan^{2}\theta/R^{2})^{2}}\left(\frac{c_{2}(R+1)}{R^{3}}(e^{2\theta/\tan\theta}+1)+\frac{c_{1}}{R^{2}}+c_{3}\right), (4.59)

where, as in [For02a], and using (4.51),

c0:=1sinθcos3θ=14.464,c1:=(θsinθcosθ)tan4θ=15.541,c2:=tan3θsin2θ=7.9763,c3:=(θsinθcosθ)tan2θ=3.4108.\begin{split}c_{0}&:=\frac{1}{\sin\theta\cos^{3}\theta}=14.464\ldots,\\ c_{1}&:=(\theta-\sin\theta\cos\theta)\tan^{4}\theta=15.541\ldots,\\ c_{2}&:=\tan^{3}\theta\sin^{2}\theta=7.9763\ldots,\\ c_{3}&:=(\theta-\sin\theta\cos\theta)\tan^{2}\theta=3.4108\ldots.\end{split} (4.60)

We also have, via a direct computation using (4.51) and (4.54),

G(0)=13csc2θ(3(4θ25)+θ(154θ2)cotθ)θcscθsecθ0.659108.G^{\prime}(0)=\frac{1}{3}\csc^{2}\theta(3(4\theta^{2}-5)+\theta(15-4\theta^{2})\cot\theta)-\theta\csc\theta\sec\theta\geq-0.659108. (4.61)

In the next lemma, we combine all of our above results with Lemma 4.7 to form an explicit inequality involving the real and imaginary parts of a zero ρ=β+it\rho=\beta+it.

Lemma 4.8.

Let k4k\geq 4 be an integer and R3R\geq 3. Suppose β+it\beta+it is a zero satisfying t31012t\geq 3\cdot 10^{12} and 1βηk/21-\beta\leq\eta_{k}/2. Furthermore, suppose that there are no zeroes in the rectangle

1λ<s1,t1sDt+11-\lambda<\Re s\leq 1,\qquad t-1\leq\Im s\leq Dt+1

where λ\lambda is a constant satisfying 0<λmin{1β,ηk/(R+1)}0<\lambda\leq\min\left\{1-\beta,\eta_{k}/(R+1)\right\}. Then

1λ(0.179960.20523(1βλ1))0.087π2b1b01βηk2\displaystyle\frac{1}{\lambda}\left(0.17996-0.20523\left(\frac{1-\beta}{\lambda}-1\right)\right)\leq 0.087\pi^{2}\frac{b_{1}}{b_{0}}\frac{1-\beta}{\eta_{k}^{2}}
+12ηk{bb0(L13.3772k2+L2+log1.546)+logζ(1+ηk)}\displaystyle\qquad\qquad+\frac{1}{2\eta_{k}}\left\{\frac{b}{b_{0}}\left(\frac{L_{1}-3.377}{2^{k}-2}+L_{2}+\log 1.546\right)+\log\zeta(1+\eta_{k})\right\}
+c(R)λ[bb0{(23.99ηk40.051)(L13.377)+1.2031L238.58\displaystyle\qquad\qquad+c(R)\lambda\Bigg{[}\frac{b}{b_{0}}\,\bigg{\{}\left(\frac{23.99}{\sqrt{\eta_{k}}}-40.051\right)(L_{1}-3.377)+1.2031L_{2}-38.58
+2.0373L20.5322logηkηk2}+1.8],\displaystyle\qquad\qquad\qquad\qquad+\frac{2.0373L_{2}-0.5322\log\eta_{k}}{\eta_{k}^{2}}\bigg{\}}+1.8\Bigg{]},

with c(R)c(R) defined in (4.58).

Proof.

We largely follow the approach of [For02a, Lem. 7.1] (see also [MTY24, Lem. 4.7]). The main difference is instead of using the Ford–Richert bound (4.8), we use Theorem 1.1. We also use Lemma 4.1 instead of [For02a, Lem. 3.4] and Lemma 4.5 instead of [For02a, Lem. 4.3]. Furthermore, since we eventually apply this lemma with smaller values of tt, more care is taken with bounding secondary error terms which can be significant for small tt. Throughout, we retain the definitions of θ\theta, ff, gg, FF, GG, F0F_{0}, G0G_{0} and H(R)H(R) encountered previously.

If |z|ηk|z|\geq\eta_{k}, then |z|(R+1)λ|z|\geq(R+1)\lambda via the lemma’s assumptions. Hence

|F0(z)|c(R)λf(0)|z|2,z0,|z|ηk.|F_{0}(z)|\leq\frac{c(R)\lambda f(0)}{|z|^{2}},\qquad\Re z\geq 0,|z|\geq\eta_{k}. (4.62)

Therefore the conditions of Lemma 4.7 are satisfied with δ=c(R)λf(0)\delta=c(R)\lambda f(0) and η=ηk\eta=\eta_{k}. Since t1000t\geq 1000, we may apply Lemma 4.7 with s=sj:=1+ijts=s_{j}:=1+ijt with j=0,1,,Dj=0,1,\ldots,D. Summing the resulting expressions, and using the non-negativity of the trigonometric polynomial (4.5), we obtain

0\displaystyle 0 n=1Λ(n)nf(logn)[j=0Dbjcos(jtlogn)]\displaystyle\leq\sum_{n=1}^{\infty}\frac{\Lambda(n)}{n}f(\log n)\left[\sum_{j=0}^{D}b_{j}\cos(jt\log n)\right]
j=1Dbj|sjρ|ηk{F(sjρ)+f(0)(π2ηkcot(π2ηk(sjρ))1sjρ)}\displaystyle\leq-\Re\sum_{j=1}^{D}b_{j}\sum_{|s_{j}-\rho|\leq\eta_{k}}\left\{F(s_{j}-\rho)+f(0)\left(\frac{\pi}{2\eta_{k}}\cot\left(\frac{\pi}{2\eta_{k}}(s_{j}-\rho)\right)-\frac{1}{s_{j}-\rho}\right)\right\}
+f(0)4ηkj=1Dbjlog|ζ(sjηk+2ηkuiπ)|log|ζ(sj+ηk+2ηkuiπ)|cosh2udu\displaystyle\quad+\frac{f(0)}{4\eta_{k}}\sum_{j=1}^{D}b_{j}\int_{-\infty}^{\infty}\frac{\log|\zeta(s_{j}-\eta_{k}+\frac{2\eta_{k}ui}{\pi})|-\log|\zeta(s_{j}+\eta_{k}+\frac{2\eta_{k}ui}{\pi})|}{\cosh^{2}u}\text{d}u
+b0F(0)+c(R)λf(0){b(1.8+logt3)+1.8b0+j=1Dbj|sjρ|ηk1|sjρ|2}.\displaystyle\quad+b_{0}F(0)+c(R)\lambda f(0)\left\{b\left(1.8+\frac{\log t}{3}\right)+1.8b_{0}+\sum_{j=1}^{D}b_{j}\sum_{|s_{j}-\rho|\geq\eta_{k}}\frac{1}{|s_{j}-\rho|^{2}}\right\}. (4.63)

We now seek an upper bound for the right-hand side of (4.63). First, by [For02a, Lem. 5.1], we have

121cosh2uj=1Dbjlog|ζ(1+ηk+ijt+i2ηkuπ)|dub0logζ(1+ηk).-\frac{1}{2}\int_{-\infty}^{\infty}\frac{1}{\cosh^{2}u}\sum_{j=1}^{D}b_{j}\log\left|\zeta\left(1+\eta_{k}+ijt+i\frac{2\eta_{k}u}{\pi}\right)\right|\text{d}u\leq b_{0}\log\zeta(1+\eta_{k}). (4.64)

Second, note that

1bj=1Dbjlog(jt)1bj=1Dbj(L1logDj)L13.377,\frac{1}{b}\sum_{j=1}^{D}b_{j}\log(jt)\leq\frac{1}{b}\sum_{j=1}^{D}b_{j}\left(L_{1}-\log\frac{D}{j}\right)\leq L_{1}-3.377, (4.65)

so that, by Lemma 4.1,

12j=1Dbjlog|ζ(1ηk+ijt+2ηkui/π)|cosh2udub(L13.3772k2+L2+log1.546).\frac{1}{2}\sum_{j=1}^{D}b_{j}\int_{-\infty}^{\infty}\frac{\log|\zeta(1-\eta_{k}+ijt+2\eta_{k}ui/\pi)|}{\cosh^{2}u}\text{d}u\leq b\left(\frac{L_{1}-3.377}{2^{k}-2}+L_{2}+\log 1.546\right). (4.66)

Third, by Lemma 4.5, and using 5.861.6825L25.86\leq 1.6825L_{2} for t31012t\geq 3\cdot 10^{12},

j=1Dbj|sjρ|ηk1|sjρ|2j=1Dbj{(23.99ηk40.385)log(jt)+1.2031loglog(jt)40.236+0.3548loglog(jt)+5.860.5322logηkηk2N(t,ηk)ηk2}<b{(23.99ηk40.385)(L13.377)+1.2031L240.236+2.0373L20.5322logηkηk2}1ηk2j=1DbjN(jt,ηk).\begin{split}&\sum_{j=1}^{D}b_{j}\sum_{|s_{j}-\rho|\geq\eta_{k}}\frac{1}{|s_{j}-\rho|^{2}}\\ &\qquad\qquad\leq\sum_{j=1}^{D}b_{j}\bigg{\{}\left(\frac{23.99}{\sqrt{\eta_{k}}}-40.385\right)\log(jt)+1.2031\log\log(jt)-40.236\\ &\qquad\qquad\qquad\qquad+\frac{0.3548\log\log(jt)+5.86-0.5322\log\eta_{k}}{\eta_{k}^{2}}-\frac{N(t,\eta_{k})}{\eta_{k}^{2}}\bigg{\}}\\ &\qquad\qquad<b\Bigg{\{}\left(\frac{23.99}{\sqrt{\eta_{k}}}-40.385\right)(L_{1}-3.377)+1.2031L_{2}-40.236\\ &\qquad\qquad\qquad\qquad+\frac{2.0373L_{2}-0.5322\log\eta_{k}}{\eta_{k}^{2}}\Bigg{\}}-\frac{1}{\eta_{k}^{2}}\sum_{j=1}^{D}b_{j}N(jt,\eta_{k}).\end{split} (4.67)

Fourth, from [For02a, (7.7)], if λ1ρ\lambda\leq 1-\Re\rho, ληk/(R+1)\lambda\leq\eta_{k}/(R+1) and |1+ijtρ|ηk|1+ijt-\rho|\leq\eta_{k}, then

{F(1+ijtρ)+f(0)(π2ηkcot(π2ηk(1+ijtρ))11+ijtρ)}cπ2λf(0)\begin{split}&-\Re\left\{F(1+ijt-\rho)+f(0)\left(\frac{\pi}{2\eta_{k}}\cot\left(\frac{\pi}{2\eta_{k}}(1+ijt-\rho)\right)-\frac{1}{1+ijt-\rho}\right)\right\}\\ &\qquad\qquad\leq c^{\prime}\pi^{2}\lambda f(0)\end{split} (4.68)

where

c:=4π2(1+(R+1)2H(R)w(0)R3)=4π2(c1R).c^{\prime}:=\frac{4}{\pi^{2}}\left(1+\frac{(R+1)^{2}H(R)}{w(0)R^{3}}\right)=\frac{4}{\pi^{2}}\left(c-\frac{1}{R}\right). (4.69)

We use (4.68) for j1j\neq 1. If j=1j=1, then 1+ijtρ=1β1+ijt-\rho=1-\beta. Furthermore,

cotxx10.348x,0<xπ4\cot x-x^{-1}\geq-0.348x,\qquad 0<x\leq\frac{\pi}{4} (4.70)

and 1βηk/21-\beta\leq\eta_{k}/2 by assumption, so π2η(1β)π4\frac{\pi}{2\eta}(1-\beta)\leq\frac{\pi}{4} and

{F(1+itρ)+f(0)(π2ηkcot(π2ηk(1+itρ))11+itρ)}F(1β)0.348π2f(0)(1β).\begin{split}&-\Re\left\{F(1+it-\rho)+f(0)\left(\frac{\pi}{2\eta_{k}}\cot\left(\frac{\pi}{2\eta_{k}}(1+it-\rho)\right)-\frac{1}{1+it-\rho}\right)\right\}\\ &\qquad\qquad\qquad\qquad\leq F(1-\beta)-0.348\pi^{2}f(0)(1-\beta).\end{split} (4.71)

Combining (4.68) and (LABEL:classic_bound4), and noting that cπ2λf(0)b1N(t,ηk)0c^{\prime}\pi^{2}\lambda f(0)b_{1}N(t,\eta_{k})\geq 0,

j=0Dbj|1+ijtρ|ηk{F(1+ijtρ)\displaystyle-\Re\sum_{j=0}^{D}b_{j}\sum_{|1+ijt-\rho|\leq\eta_{k}}\bigg{\{}F(1+ijt-\rho)
+f(0)(π2ηkcot(π2ηk(1+ijtρ))11+ijtρ)}\displaystyle\qquad\qquad\qquad\qquad+f(0)\left(\frac{\pi}{2\eta_{k}}\cot\left(\frac{\pi}{2\eta_{k}}(1+ijt-\rho)\right)-\frac{1}{1+ijt-\rho}\right)\bigg{\}} (4.72)
b1[F(1β)0.348π2f(0)(1β)]+cπ2λf(0)j=0DbjN(jt,ηk).\displaystyle\qquad\qquad\leq-b_{1}\left[F(1-\beta)-0.348\pi^{2}f(0)(1-\beta)\right]+c^{\prime}\pi^{2}\lambda f(0)\sum_{j=0}^{D}b_{j}N(jt,\eta_{k}).

Finally we also have logt<L1logD\log t<L_{1}-\log D, so that

b(1.8+logt3)b(L13.3773+1.65).b\left(1.8+\frac{\log t}{3}\right)\leq b\left(\frac{L_{1}-3.377}{3}+1.65\right). (4.73)

Substituting (4.64), (4.66), (4.67), (4.72) and (4.73) into (4.63), we obtain

0\displaystyle 0 b1[F(1β)0.348π24ηk2f(0)(1β)]+(cπ2λf(0)cλf(0)ηk2)j=1DbjN(jt,η)\displaystyle\leq-b_{1}\left[F(1-\beta)-0.348\frac{\pi^{2}}{4\eta_{k}^{2}}f(0)(1-\beta)\right]+\left(c^{\prime}\pi^{2}\lambda f(0)-\frac{c\lambda f(0)}{\eta_{k}^{2}}\right)\sum_{j=1}^{D}b_{j}N(jt,\eta)
+f(0)[b(L13.3772k2+L2+log1.546)+b0ζ(1+ηk)]+b0F(0)\displaystyle\qquad+f(0)\left[b\left(\frac{L_{1}-3.377}{2^{k}-2}+L_{2}+\log 1.546\right)+b_{0}\zeta(1+\eta_{k})\right]+b_{0}F(0)
+cλf(0)[b{(23.99ηk40.051)(L13.377)+1.2031L238.58\displaystyle\qquad+c\lambda f(0)\Bigg{[}b\;\bigg{\{}\left(\frac{23.99}{\sqrt{\eta_{k}}}-40.051\right)(L_{1}-3.377)+1.2031L_{2}-38.58
+2.0373L20.5322logηkηk2}+1.8b0].\displaystyle\qquad\qquad\qquad+\frac{2.0373L_{2}-0.5322\log\eta_{k}}{\eta_{k}^{2}}\bigg{\}}+1.8b_{0}\Bigg{]}. (4.74)

However, for all ηk1/2\eta_{k}\leq 1/2,

cπ2λf(0)cλf(0)ηk2[4(c1R)4c]λf(0)<0c^{\prime}\pi^{2}\lambda f(0)-\frac{c\lambda f(0)}{\eta_{k}^{2}}\leq\left[4\left(c-\frac{1}{R}\right)-4c\right]\lambda f(0)<0 (4.75)

so we may drop the sum in (4.74). Furthermore,

b0F(0)b1F(1β)\displaystyle b_{0}F(0)-b_{1}F(1-\beta) =b1(G(0)G(1βλ1))(b1G(0)b0G(1))\displaystyle=b_{1}\left(G(0)-G\left(\frac{1-\beta}{\lambda}-1\right)\right)-\left(b_{1}G(0)-b_{0}G(-1)\right)
=b1(G(0)G(1βλ1))b0f(0)cos2θλ.\displaystyle=b_{1}\left(G(0)-G\left(\frac{1-\beta}{\lambda}-1\right)\right)-\frac{b_{0}f(0)\cos^{2}\theta}{\lambda}. (4.76)

Note that G(z)G^{\prime}(z) is continuous and decreasing for z>0z>0, so by the mean-value theorem, and using (4.61) followed by (4.52) and (4.53),

G(0)G(1βλ1)(11βλ)G(0)0.659108(1βλ1)0.11747b0f(0)λ(1βλ1).\begin{split}G(0)-G\left(\frac{1-\beta}{\lambda}-1\right)\leq\left(1-\frac{1-\beta}{\lambda}\right)G^{\prime}(0)\leq 0.659108\left(\frac{1-\beta}{\lambda}-1\right)\\ \leq\frac{0.11747b_{0}f(0)}{\lambda}\left(\frac{1-\beta}{\lambda}-1\right).\end{split} (4.77)

Combining with (4.76), (4.75) and (4.77) with (4.74), we have

b0f(0)cos2θλ0.11747b1b0f(0)λ(1βλ1)b1F(1β)b0F(0)\displaystyle\frac{b_{0}f(0)\cos^{2}\theta}{\lambda}-\frac{0.11747b_{1}b_{0}f(0)}{\lambda}\left(\frac{1-\beta}{\lambda}-1\right)\leq b_{1}F(1-\beta)-b_{0}F(0)
f(0)2ηk[b(L13.3772k2+L2+log1.546)+b0logζ(1+ηk)]\displaystyle\qquad\leq\frac{f(0)}{2\eta_{k}}\bigg{[}b\left(\frac{L_{1}-3.377}{2^{k}-2}+L_{2}+\log 1.546\right)+b_{0}\log\zeta(1+\eta_{k})\bigg{]} (4.78)
+c(R)λf(0)[b{(23.99ηk40.051)(L13.377)+1.2031L238.58\displaystyle\qquad\qquad+c(R)\lambda f(0)\Bigg{[}b\;\bigg{\{}\left(\frac{23.99}{\sqrt{\eta_{k}}}-40.051\right)(L_{1}-3.377)+1.2031L_{2}-38.58
+2.0373L20.5322logηkηk2}+1.8b0]+0.348b1π24ηk2f(0)(1β).\displaystyle\qquad\qquad\qquad+\frac{2.0373L_{2}-0.5322\log\eta_{k}}{\eta_{k}^{2}}\bigg{\}}+1.8b_{0}\Bigg{]}+0.348b_{1}\frac{\pi^{2}}{4\eta_{k}^{2}}f(0)(1-\beta).

Dividing through by b0f(0)b_{0}f(0), the result follows. ∎

We now proceed to the main proof of Corollary 1.2, which is similar to the proof of [For02a, Thm. 2]. Throughout, we take t0:=exp(170.2)t_{0}:=\exp(170.2) and t1:=exp(967.6)t_{1}:=\exp(967.6). If 3tH:=5.451083\leq t\leq H:=5.45\cdot 10^{8}, then Corollary 1.2 follows from the partial verification of the Riemann Hypothesis up to height HH, performed independently in e.g. [LRW86, Wed03, Gou04, Pla17, PT21]. Note that in [PT21], the Riemann Hypothesis is actually verified up to height 310123\cdot 10^{12}, however our results do not require this, and the lower value of HH has been independently verified by multiple authors. If Ht<t0H\leq t<t_{0}, then Corollary 1.2 follows from (1.4). Assume now that tt0t\geq t_{0}. Throughout, let

Z(β,t):=(1β)logtloglogt,M:=infζ(β+it)=0,tt0Z(β,t),Z(\beta,t):=(1-\beta)\frac{\log t}{\log\log t},\qquad M:=\inf_{\begin{subarray}{c}\zeta(\beta+it)=0,\\ t\geq t_{0}\end{subarray}}Z(\beta,t), (4.79)
α:=0.13913,M1:=0.0470978.\alpha:=0.13913,\qquad M_{1}:=0.0470978. (4.80)

For integer k4k\geq 4 let

Tk:=(2k2k)1/αandηk:=k2k2T_{k}:=\left(2^{k2^{k}}\right)^{1/\alpha}\qquad\text{and}\qquad\eta_{k}:=\frac{k}{2^{k}-2} (4.81)

so that

k=W0(αlogTk)log2,k=\frac{W_{0}(\alpha\log T_{k})}{\log 2}, (4.82)

where W0(x)W_{0}(x) is the principal branch of the Lambert WW function, which satisfies W0(x)eW0(x)=xW_{0}(x)e^{W_{0}(x)}=x for any x>0x>0.

Note that MM1M\geq M_{1} implies Corollary 1.2. Assume for a contradiction that M<M1M<M_{1}. Then, there exists a zero β+it\beta+it for which Z(β,t)Z(\beta,t) is arbitrarily close to MM. In particular, let β+it\beta+it be a zero such that tHt\geq H and satisfying

MZ(β,t)M(1+δ),δ:=min{10100logt0,M1M2M}.M\leq Z(\beta,t)\leq M(1+\delta),\qquad\delta:=\min\left\{\frac{10^{-100}}{\log t_{0}},\frac{M_{1}-M}{2M}\right\}. (4.83)

In particular, we have Z(β,t)M1Z(\beta,t)\leq M_{1}. If we let

λ:=ML2L1,\lambda:=\frac{ML_{2}}{L_{1}}, (4.84)

then there are no zeroes of ζ(s)\zeta(s) in the rectangle

1λ<s1,t1sDt+1.1-\lambda<\Re s\leq 1,\qquad t-1\leq\Im s\leq Dt+1. (4.85)

This is because if a zero β+it\beta^{\prime}+it^{\prime} exists in that rectangle, then as loglogx/logx\log\log x/\log x is decreasing for x>eex>e^{e}, and ee<t<Dt+1e^{e}<t^{\prime}<Dt^{\prime}+1,

1β<λMloglogtlogt1-\beta^{\prime}<\lambda\leq\frac{M\log\log t^{\prime}}{\log t^{\prime}} (4.86)

so that Z(β,t)<MZ(\beta^{\prime},t^{\prime})<M. If tt0t^{\prime}\geq t_{0}, then this is in contradiction to the definition of MM. On the other hand if t01t<t0t_{0}-1\leq t^{\prime}<t_{0}, then by (1.4) and a short numerical computation,

1β0.049620.0196/(J(t01)+1.15)J(t01)+0.685+0.155loglog(t01)<M1loglogt0logt0<M1loglogtlogt.1-\beta^{\prime}\leq\frac{0.04962-0.0196/(J(t_{0}-1)+1.15)}{J(t_{0}-1)+0.685+0.155\log\log(t_{0}-1)}<\frac{M_{1}\log\log t_{0}}{\log t_{0}}<\frac{M_{1}\log\log t^{\prime}}{\log t^{\prime}}.

Thus, Z(β,t)M1>MZ(\beta^{\prime},t^{\prime})\geq M_{1}>M, which is a contradiction. Therefore, (4.85) is zero-free, and any zero β+it\beta+it must satisfy

λ1β.\lambda\leq 1-\beta. (4.87)

We now divide our argument into two parts. For t0t<t1t_{0}\leq t<t_{1}, our argument relies on a uniform bound on ζ(s)\zeta(s) for 1/2s5/71/2\leq\Re s\leq 5/7. Here, we use Lemma 4.3 and Lemma 4.6. These tools allow us to derive a sharp estimate of the zero-free region over a finite range of tt close to t0t_{0}, however are insufficient to obtain an asymptotic estimate of the required strength. Therefore, for tt1t\geq t_{1} we switch to bounds on ζ(σk+it)\zeta(\sigma_{k}+it), where we take kk\to\infty as tt\to\infty, in the form of Lemma 4.8.

First, consider the case tt1t\geq t_{1}. Since T5=exp(797.1)<t1T_{5}=\exp(797.1\ldots)<t_{1}, there is always an integer k5k\geq 5 such that t[Tk,Tk+1)t\in[T_{k},T_{k+1}). Then, since (2x2)/x(2^{x}-2)/x is increasing for all x>0x>0, and W0(αlogt)/log2k>0W_{0}(\alpha\log t)/\log 2\geq k>0, we have, for tt1t\geq t_{1},

1ηk=2k2kexp(W0(αlogt))2W0(αlogt)/log2=A0(αlogt)logt(loglogt)2\frac{1}{\eta_{k}}=\frac{2^{k}-2}{k}\leq\frac{\exp(W_{0}(\alpha\log t))-2}{W_{0}(\alpha\log t)/\log 2}=A_{0}(\alpha\log t)\frac{\log t}{(\log\log t)^{2}} (4.88)

where, since (loglogt)2=(log(αlogt)logα)2(\log\log t)^{2}=(\log(\alpha\log t)-\log\alpha)^{2},

A0(x):=αlog2(logxlogα)2W0(x)2(12W0(x)x).A_{0}(x):=\frac{\alpha\log 2(\log x-\log\alpha)^{2}}{W_{0}(x)^{2}}\left(1-\frac{2W_{0}(x)}{x}\right). (4.89)

Note that

A0(x)=2αlog2log(x/α)x2W0(x)2(W0(x)+1)A1(x)A_{0}^{\prime}(x)=\frac{2\alpha\log 2\log(x/\alpha)}{x^{2}W_{0}(x)^{2}(W_{0}(x)+1)}A_{1}(x) (4.90)

where

A1(x):=(W0(x)2+2W0(x)x)logxα2W0(x)2(2x)W0(x)+x.A_{1}(x):=(W_{0}(x)^{2}+2W_{0}(x)-x)\log\frac{x}{\alpha}-2W_{0}(x)^{2}-(2-x)W_{0}(x)+x. (4.91)

Since W0(x)=logxloglogx+o(1)W_{0}(x)=\log x-\log\log x+o(1), we have A1(x)=xloglogx+O(x)A_{1}(x)=-x\log\log x+O(x), i.e. A1(x)<0A_{1}(x)<0 for sufficiently large xx. We in fact find computationally that A1(x)<0A_{1}(x)<0 for all x>x=15.832x>x^{*}=15.832\ldots. It follows from (4.90) that if x0:=max{x,αlogt1}x_{0}:=\max\{x^{*},\alpha\log t_{1}\}, then

A0(αlogt)A0(x0)0.3297,tt1.A_{0}(\alpha\log t)\leq A_{0}(x_{0})\leq 0.3297,\qquad t\geq t_{1}. (4.92)

This also implies that

1ηk<C1logtloglogt,C1:=A0(x0)loglogt00.047958.\begin{split}\frac{1}{\eta_{k}}<C_{1}\frac{\log t}{\log\log t},\qquad C_{1}:=\frac{A_{0}(x_{0})}{\log\log t_{0}}\leq 0.047958.\end{split} (4.93)

Then

λMloglogtlogt0.047958M1ηkηk442.729,\lambda\leq\frac{M\log\log t}{\log t}\leq 0.047958M_{1}\eta_{k}\leq\frac{\eta_{k}}{442.729}, (4.94)

i.e. ληk/(R+1)\lambda\leq\eta_{k}/(R+1) with R=441.729R=441.729. This allows us to compute, via (4.58) and (4.59),

c(R)1.02268.c(R)\leq 1.02268. (4.95)

Collecting (4.87) and (4.94), all conditions of Lemma 4.8 are met with η=ηk\eta=\eta_{k}, R=441.729R=441.729. We thus have, for all zeroes β+it\beta+it such that Tkt<Tk+1T_{k}\leq t<T_{k+1},

1λ(0.179960.20523(1βλ1))0.087π2b1b01βηk2+12ηk{bb0[L13.3772k2+L2+log1.546]+logζ(1+ηk)}+c(R)λ[bb0{(23.99ηk40.051)(L13.377)+1.2031L238.58+2.0373L20.5322logηkηk2}+1.8].\begin{split}&\frac{1}{\lambda}\left(0.17996-0.20523\left(\frac{1-\beta}{\lambda}-1\right)\right)\leq 0.087\pi^{2}\frac{b_{1}}{b_{0}}\frac{1-\beta}{\eta_{k}^{2}}\\ &\qquad+\frac{1}{2\eta_{k}}\left\{\frac{b}{b_{0}}\left[\frac{L_{1}-3.377}{2^{k}-2}+L_{2}+\log 1.546\right]+\log\zeta(1+\eta_{k})\right\}\\ &\qquad+c(R)\lambda\Bigg{[}\frac{b}{b_{0}}\,\bigg{\{}\left(\frac{23.99}{\sqrt{\eta_{k}}}-40.051\right)(L_{1}-3.377)+1.2031L_{2}-38.58\\ &\qquad\qquad\qquad\qquad+\frac{2.0373L_{2}-0.5322\log\eta_{k}}{\eta_{k}^{2}}\bigg{\}}+1.8\Bigg{]}.\end{split} (4.96)

We now proceed to bound each term appearing on the right hand side. First,

W0(αlogt)W0(αlogTk)<W0(αlogTk+1)W0(αlogTk)=k+1k.\frac{W_{0}(\alpha\log t)}{W_{0}(\alpha\log T_{k})}<\frac{W_{0}(\alpha\log T_{k+1})}{W_{0}(\alpha\log T_{k})}=\frac{k+1}{k}. (4.97)

Furthermore, logx/W0(αx)\log x/W_{0}(\alpha x) is decreasing for x>x1:=max{e1+αe,logt1}x>x_{1}:=\max\{e^{1+\alpha e},\log t_{1}\}, since

ddxlogxW0(αx)=W0(αx)logx+1xW0(αx)(W0(αx)+1).\frac{\text{d}}{\text{d}x}\frac{\log x}{W_{0}(\alpha x)}=\frac{W_{0}(\alpha x)-\log x+1}{xW_{0}(\alpha x)(W_{0}(\alpha x)+1)}. (4.98)

Taking x=logtx=\log t and combining with (4.97) and (4.82) gives

1ηkL13.3772k2=(L13.377)log2W0(αlogTk)<(L13.377)(1+k1)log2W0(αlogt)=(1+k1)log2(L13.377logt)loglogtW0(αlogt)logtloglogtC2logtloglogt,\begin{split}\frac{1}{\eta_{k}}\frac{L_{1}-3.377}{2^{k}-2}&=\frac{(L_{1}-3.377)\log 2}{W_{0}(\alpha\log T_{k})}<\frac{(L_{1}-3.377)(1+k^{-1})\log 2}{W_{0}(\alpha\log t)}\\ &=(1+k^{-1})\log 2\left(\frac{L_{1}-3.377}{\log t}\right)\frac{\log\log t}{W_{0}(\alpha\log t)}\frac{\log t}{\log\log t}\leq C_{2}\frac{\log t}{\log\log t},\end{split} (4.99)

where, since L13.377logt+log(D+t11)3.377L_{1}-3.377\leq\log t+\log(D+t_{1}^{-1})-3.377 and log(D+t11)3.377>0\log(D+t_{1}^{-1})-3.377>0,

C2:=65log2(1+log(D+t11)3.377logt1)logx1W0(αx1)1.58176.C_{2}:=\frac{6}{5}\log 2\left(1+\frac{\log(D+t_{1}^{-1})-3.377}{\log t_{1}}\right)\frac{\log x_{1}}{W_{0}(\alpha x_{1})}\leq 1.58176. (4.100)

Next, using (4.88), we have

L2ηk<(A0(x0)L2loglogt)logtloglogtC3logtloglogt,\begin{split}\frac{L_{2}}{\eta_{k}}&<\left(\frac{A_{0}(x_{0})L_{2}}{\log\log t}\right)\frac{\log t}{\log\log t}\leq C_{3}\frac{\log t}{\log\log t},\end{split} (4.101)

where

C3:=A0(x0)(1+log(1+log(D+t11)/logt1)loglogt1)0.32989.C_{3}:=A_{0}(x_{0})\left(1+\frac{\log\left(1+\log(D+t_{1}^{-1})/\log t_{1}\right)}{\log\log t_{1}}\right)\leq 0.32989. (4.102)

Next, since x1logx-x^{-1}\log x is decreasing for x<1x<1, by (4.88) we have

logηkηk\displaystyle-\frac{\log\eta_{k}}{\eta_{k}} log[(loglogt)2/(A0(x0)logt)](loglogt)2/(A0(x0)logt)\displaystyle\leq-\frac{\log[(\log\log t)^{2}/(A_{0}(x_{0})\log t)]}{(\log\log t)^{2}/(A_{0}(x_{0})\log t)}
=A0(x0)[logA0(x0)+loglogt2logloglogtloglogt]logtloglogt<C4logtloglogt,\displaystyle=A_{0}(x_{0})\left[\frac{\log A_{0}(x_{0})+\log\log t-2\log\log\log t}{\log\log t}\right]\frac{\log t}{\log\log t}<C_{4}\frac{\log t}{\log\log t}, (4.103)

where

C4:=A0(x0)(logA0(x0)loglogt1+1)0.27649.C_{4}:=A_{0}(x_{0})\left(\frac{\log A_{0}(x_{0})}{\log\log t_{1}}+1\right)\leq 0.27649. (4.104)

Finally, using (4.26) with σ=1+ηk\sigma=1+\eta_{k}, we have

logζ(1+ηk)γηklogηk.\log\zeta(1+\eta_{k})\leq\gamma\eta_{k}-\log\eta_{k}. (4.105)

Combining (4.93), (4.99), (4.101), (4.103) and (4.105),

12ηk{bb0(L13.3772k2+L2+log1.546)+logζ(1+ηk)}C5logtloglogt\frac{1}{2\eta_{k}}\left\{\frac{b}{b_{0}}\left(\frac{L_{1}-3.377}{2^{k}-2}+L_{2}+\log 1.546\right)+\log\zeta(1+\eta_{k})\right\}\leq C_{5}\frac{\log t}{\log\log t} (4.106)

where, from (4.5) we have b=3.57440943022073b=3.57440943022073, b0=1b_{0}=1, and

C5:=[C1log1.5462+C22+C32]bb0+C42+γ2loglogt0logt03.59415.C_{5}:=\left[\frac{C_{1}\log 1.546}{2}+\frac{C_{2}}{2}+\frac{C_{3}}{2}\right]\frac{b}{b_{0}}+\frac{C_{4}}{2}+\frac{\gamma}{2}\frac{\log\log t_{0}}{\log t_{0}}\leq 3.59415. (4.107)

This constitutes the “main” term. Next, from (4.88) and using (4.83),

0.087π2b1b01βηk21.5002Z(β,t)loglogtlogt(A0(x0)logt(loglogt)2)2C6logtloglogt\begin{split}\frac{0.087\pi^{2}b_{1}}{b_{0}}\frac{1-\beta}{\eta_{k}^{2}}&\leq 1.5002\frac{Z(\beta,t)\log\log t}{\log t}\cdot\left(\frac{A_{0}(x_{0})\log t}{(\log\log t)^{2}}\right)^{2}\leq C_{6}\frac{\log t}{\log\log t}\end{split} (4.108)

where

C6:=1.5002A0(x0)2M1(loglogt1)20.00017.C_{6}:=\frac{1.5002A_{0}(x_{0})^{2}M_{1}}{(\log\log t_{1})^{2}}\leq 0.00017. (4.109)

The remaining term of (4.96) is majorised by

c(R)M{bb0{(23.99ηk40.051)L22L1+(1.2031L238.58)L22L12+2.0373L20.5322logηkηk2L22L12}+1.8L22L12}L1L2.\begin{split}&c(R)M\bigg{\{}\frac{b}{b_{0}}\bigg{\{}\left(\frac{23.99}{\sqrt{\eta_{k}}}-40.051\right)\frac{L_{2}^{2}}{L_{1}}+(1.2031L_{2}-38.58)\frac{L_{2}^{2}}{L_{1}^{2}}\\ &\qquad\qquad\qquad+\frac{2.0373L_{2}-0.5322\log\eta_{k}}{\eta_{k}^{2}}\frac{L_{2}^{2}}{L_{1}^{2}}\bigg{\}}+1.8\frac{L_{2}^{2}}{L_{1}^{2}}\bigg{\}}\frac{L_{1}}{L_{2}}.\end{split} (4.110)

We have, by (4.88), for tt1t\geq t_{1},

1ηkL22L1A0(x0)1/2log1/2tloglogtL22L1<A0(x0)1/2L2L11/2\frac{1}{\sqrt{\eta_{k}}}\frac{L_{2}^{2}}{L_{1}}\leq A_{0}(x_{0})^{1/2}\frac{\log^{1/2}t}{\log\log t}\frac{L_{2}^{2}}{L_{1}}<A_{0}(x_{0})^{1/2}\frac{L_{2}}{L_{1}^{1/2}} (4.111)

so that

(23.99ηk40.051)L22L1+(1.2031L238.58)L22L12U(L1),\begin{split}\left(\frac{23.99}{\sqrt{\eta_{k}}}-40.051\right)\frac{L_{2}^{2}}{L_{1}}+(1.2031L_{2}-38.58)\frac{L_{2}^{2}}{L_{1}^{2}}\leq U(L_{1}),\end{split} (4.112)

where, using (4.92),

U(x):=13.775logxx1/240.051log2xx+(1.2031logx38.58)log2xx2.U(x):=13.775\frac{\log x}{x^{1/2}}-40.051\frac{\log^{2}x}{x}+(1.2031\log x-38.58)\frac{\log^{2}x}{x^{2}}. (4.113)

By solving for stationary points explicitly, we find that for xL1(t1)x\geq L_{1}(t_{1}),

U(x)C7:=1.18399.U(x)\leq C_{7}:=1.18399. (4.114)

Meanwhile, we also have the estimates

1ηk2L23L12A0(x0)2log2t(loglogt)4L23L12C8,\frac{1}{\eta_{k}^{2}}\frac{L_{2}^{3}}{L_{1}^{2}}\leq A_{0}(x_{0})^{2}\frac{\log^{2}t}{(\log\log t)^{4}}\frac{L_{2}^{3}}{L_{1}^{2}}\leq C_{8}, (4.115)

where, since L1>logtL_{1}>\log t and L2<loglogt+log(N+t11)logt1L_{2}<\log\log t+\frac{\log(N+t_{1}^{-1})}{\log t_{1}},

C8:=A0(x0)2(1+log(N+t11)/logt1loglogt1)31loglogt10.01584,C_{8}:=A_{0}(x_{0})^{2}\left(1+\frac{\log(N+t_{1}^{-1})/\log t_{1}}{\log\log t_{1}}\right)^{3}\frac{1}{\log\log t_{1}}\leq 0.01584, (4.116)

and, since x2logx-x^{-2}\log x is increasing and x2log2xx^{-2}\log^{2}x is decreasing for xex\geq e,

logηkηk2L22L12log[(loglogt)2/(A0(x0)logt)](loglogt)4/(A0(x0)logt)2(loglogt)2log2tC8L1L2-\frac{\log\eta_{k}}{\eta_{k}^{2}}\frac{L_{2}^{2}}{L_{1}^{2}}\leq-\frac{\log[(\log\log t)^{2}/(A_{0}(x_{0})\log t)]}{(\log\log t)^{4}/(A_{0}(x_{0})\log t)^{2}}\frac{(\log\log t)^{2}}{\log^{2}t}\leq C_{8}\frac{L_{1}}{L_{2}} (4.117)

where

C9:=A0(x0)2loglogt1(1+logA0(x0)loglogt1)L2(t1)L1(t1)0.0001.C_{9}:=\frac{A_{0}(x_{0})^{2}}{\log\log t_{1}}\left(1+\frac{\log A_{0}(x_{0})}{\log\log t_{1}}\right)\frac{L_{2}(t_{1})}{L_{1}(t_{1})}\leq 0.0001. (4.118)

Next, we have L22/L12C10:=0.00006L_{2}^{2}/L_{1}^{2}\leq C_{10}:=0.00006 for tt1t\geq t_{1}. Combining this with the bound for U(x)U(x) and (4.115), we have that (4.110) is C11L1/L2\leq C_{11}L_{1}/L_{2}, where

C11:=c(R)M1(bb0(C7+2.0373C8+0.5322C9)+1.8C10)0.20942.C_{11}:=c(R)M_{1}\left(\frac{b}{b_{0}}\left(C_{7}+2.0373C_{8}+0.5322C_{9}\right)+1.8C_{10}\right)\leq 0.20942. (4.119)

Lastly, applying (4.83) we have

1βλ1=Z(β,t)loglogtlogtL1ML21<(1+δ)L1logt1C12logt1,\begin{split}\frac{1-\beta}{\lambda}-1&=\frac{Z(\beta,t)\log\log t}{\log t}\frac{L_{1}}{ML_{2}}-1<(1+\delta)\frac{L_{1}}{\log t}-1\leq\frac{C_{12}}{\log t_{1}},\end{split} (4.120)

where, since tt1t\geq t_{1},

C12:=10100(1+log(D+t11)logt1)+log(D+t11)3.82865.C_{12}:=10^{-100}\left(1+\frac{\log(D+t_{1}^{-1})}{\log t_{1}}\right)+\log(D+t_{1}^{-1})\leq 3.82865. (4.121)

Combining (4.96), (4.107), (4.108), (4.119), (4.120) and (4.121), and from the definition of M1M_{1},

M=λL1L20.179960.7857/logt1C5+C6+C110.04709785>M1,M=\lambda\frac{L_{1}}{L_{2}}\geq\frac{0.17996-0.7857/\log t_{1}}{C_{5}+C_{6}+C_{11}}\geq 0.04709785>M_{1}, (4.122)

so the desired contradiction is reached if tt1t\geq t_{1}.

For t0t<t1t_{0}\leq t<t_{1}, we use a similar argument, except with Lemma 4.3 and Lemma 4.6. Here, we choose

η:=η(t)=(8EL2(t))1,E:=30.95461.\eta:=\eta(t)=\left(8-\frac{E}{L_{2}(t)}\right)^{-1},\qquad E:=30.95461. (4.123)

These parameters guarantee that, for all t0tt1t_{0}\leq t\leq t_{1}, we have 2/7η(t)1/22/7\leq\eta(t)\leq 1/2. Observe that for tt0t\geq t_{0}, we have

L2loglogt1+log(D+t01)logt0loglogt01.0044\frac{L_{2}}{\log\log t}\leq 1+\frac{\log(D+t_{0}^{-1})}{\log t_{0}\log\log t_{0}}\leq 1.0044

so that

(8EL2)loglogtlogt8loglogtE/1.0044logt0.06039.\left(8-\frac{E}{L_{2}}\right)\frac{\log\log t}{\log t}\leq\frac{8\log\log t-E/1.0044}{\log t}\leq 0.06039. (4.124)

Here, the last inequality follows from solving explicitly the maximum of (8logxE/1.0044)/x(8\log x-E/1.0044)/x for xlogt0x\geq\log t_{0}. Therefore, we have

1η0.06039logtloglogtλ0.06039M1ηη351.588,\frac{1}{\eta}\leq 0.06039\frac{\log t}{\log\log t}\qquad\implies\qquad\lambda\leq 0.06039M_{1}\eta\leq\frac{\eta}{351.588},

so that λη/(R+1)\lambda\leq\eta/(R^{\prime}+1) with R=350.588R^{\prime}=350.588. This implies that c(R)1.0288c(R^{\prime})\leq 1.0288.

Since 2/7η(t)1/22/7\leq\eta(t)\leq 1/2 for all t0tt1t_{0}\leq t\leq t_{1}, we may apply Lemma 4.3 and 4.6. We use these in place of Lemma 4.2 and 4.5 respectively, to obtain, via the same argument as Lemma 4.8,

1λ(0.179960.20523(1βλ1))0.087π2b1b01βη2+12η{bb0(8η118(L13.377)+L2+1.6594.279η)+logζ(1+η)}+c(R)λ[bb0{(0.891+0.6079ν)(L13.377)+(0.7813+0.58ν)L2+7.329+3.898ν}+1.8],\begin{split}&\frac{1}{\lambda}\left(0.17996-0.20523\left(\frac{1-\beta}{\lambda}-1\right)\right)\leq 0.087\pi^{2}\frac{b_{1}}{b_{0}}\frac{1-\beta}{\eta^{2}}\\ &\qquad+\frac{1}{2\eta}\left\{\frac{b}{b_{0}}\left(\frac{8\eta-1}{18}(L_{1}-3.377)+L_{2}+1.659-4.279\eta\right)+\log\zeta(1+\eta)\right\}\\ &\qquad+c(R^{\prime})\lambda\Bigg{[}\frac{b}{b_{0}}\bigg{\{}(0.891+0.6079\nu)(L_{1}-3.377)+(0.7813+0.58\nu)L_{2}\\ &\qquad\qquad\qquad+7.329+3.898\nu\bigg{\}}+1.8\Bigg{]},\end{split} (4.125)

where, as before, ν:=12(η2+(1η)2)\nu:=\frac{1}{2}(\eta^{-2}+(1-\eta)^{-2}). First, analogously to before, we have

0.087π2b1b01βη21.5002Z(β,t)loglogtlogt(0.06039logtloglogt)20.00015L1L2.\begin{split}\frac{0.087\pi^{2}b_{1}}{b_{0}}\frac{1-\beta}{\eta^{2}}&\leq 1.5002\frac{Z(\beta,t)\log\log t}{\log t}\cdot\left(\frac{0.06039\log t}{\log\log t}\right)^{2}\leq\frac{0.00015L_{1}}{L_{2}}.\end{split} (4.126)

Next, since η2/7\eta\geq 2/7 and by (4.26),

logζ(1+η)2ηγ2logη2η2.481.\frac{\log\zeta(1+\eta)}{2\eta}\leq\frac{\gamma}{2}-\frac{\log\eta}{2\eta}\leq 2.481. (4.127)

Hence

12η{bb0(1.6594.279η+8η118(L13.377)+L2)+logζ(1+η)}\displaystyle\frac{1}{2\eta}\left\{\frac{b}{b_{0}}\left(1.659-4.279\eta+\frac{8\eta-1}{18}(L_{1}-3.377)+L_{2}\right)+\log\zeta(1+\eta)\right\}
b2b0{1.847(8EL2)5.779+E18L1L2+L2(8EL2)}+2.481\displaystyle\qquad\qquad\leq\frac{b}{2b_{0}}\left\{1.847\left(8-\frac{E}{L_{2}}\right)-5.779+\frac{E}{18}\frac{L_{1}}{L_{2}}+L_{2}\left(8-\frac{E}{L_{2}}\right)\right\}+2.481
(3.07346+14.298L22L136.761L2L1102.18L1)L1L2\displaystyle\qquad\qquad\leq\left(3.07346+14.298\frac{L_{2}^{2}}{L_{1}}-36.761\frac{L_{2}}{L_{1}}-\frac{102.18}{L_{1}}\right)\frac{L_{1}}{L_{2}}
3.59852L1L2,\displaystyle\qquad\qquad\leq 3.59852\frac{L_{1}}{L_{2}}, (4.128)

where the last inequality follows from

14.298log2x36.761logx102.18x0.52506,xL1(t0).\frac{14.298\log^{2}x-36.761\log x-102.18}{x}\leq 0.52506,\qquad x\geq L_{1}(t_{0}). (4.129)

Next, note that since η1/2\eta\leq 1/2,

ν=12(1η2+1(1η)2)12((8EL2)2+4)=348EL2+E22L22\nu=\frac{1}{2}\left(\frac{1}{\eta^{2}}+\frac{1}{(1-\eta)^{2}}\right)\leq\frac{1}{2}\left(\left(8-\frac{E}{L_{2}}\right)^{2}+4\right)=34-\frac{8E}{L_{2}}+\frac{E^{2}}{2L_{2}^{2}}

so that

νL2(34L228EL2+E22L1)L1L2B2L1L2\nu L_{2}\leq\left(\frac{34L_{2}^{2}-8EL_{2}+\frac{E^{2}}{2}}{L_{1}}\right)\frac{L_{1}}{L_{2}}\leq B_{2}\frac{L_{1}}{L_{2}} (4.130)

where

B2:=maxxL1(t0)34log2x8Elogx+E2/2x0.61184.B_{2}:=\max_{x\geq L_{1}(t_{0})}\frac{34\log^{2}x-8E\log x+E^{2}/2}{x}\leq 0.61184. (4.131)

Furthermore,

c(R)λ(bb0{(0.891+0.6079ν)(L13.377)+(0.7813+0.58ν)L2\displaystyle c(R^{\prime})\lambda\bigg{(}\frac{b}{b_{0}}\bigg{\{}(0.891+0.6079\nu)(L_{1}-3.377)+(0.7813+0.58\nu)L_{2}
+7.329+3.898ν}+1.8)B3(t)L1L2,\displaystyle\qquad\qquad\qquad\qquad+7.329+3.898\nu\bigg{\}}+1.8\bigg{)}\leq B_{3}(t)\frac{L_{1}}{L_{2}}, (4.132)

where, since (7.3293.3770.891)b/b0+1.8<18(7.329-3.377\cdot 0.891)b/b_{0}+1.8<18 and 3.8983.3770.6079<23.898-3.377\cdot 0.6079<2, we may take

B3(t):=c(R)M1(bb0{0.891L22L1+0.6079B2+0.7813L23L12+0.58B2L2L1+2B2L1}+18L22L12).\begin{split}B_{3}(t)&:=c(R^{\prime})M_{1}\bigg{(}\frac{b}{b_{0}}\bigg{\{}0.891\frac{L_{2}^{2}}{L_{1}}+0.6079B_{2}+0.7813\frac{L_{2}^{3}}{L_{1}^{2}}\\ &\qquad\qquad+0.58B_{2}\frac{L_{2}}{L_{1}}+\frac{2B_{2}}{L_{1}}\bigg{\}}+18\frac{L_{2}^{2}}{L_{1}^{2}}\bigg{)}.\end{split} (4.133)

Since B3(t)B_{3}(t) is decreasing in tt for tt0t\geq t_{0}, we have B3(t)B3(t0)0.09245B_{3}(t)\leq B_{3}(t_{0})\leq 0.09245. Therefore, collecting (4.125), (4.126), (4.128) and (4.132), and from the definition of M1M_{1},

M=λL1L20.179960.7857/logt00.00015+3.59852+0.092450.0475>M1,M=\lambda\frac{L_{1}}{L_{2}}\geq\frac{0.17996-0.7857/\log t_{0}}{0.00015+3.59852+0.09245}\geq 0.0475>M_{1}, (4.134)

hence we have also obtained the desired contradiction for t0t<t1t_{0}\leq t<t_{1}.

Therefore, we have shown that if tt0t\geq t_{0},

(1β)logtloglogt=Z(β,t)MM1=0.04709781βloglogt21.233logt,(1-\beta)\frac{\log t}{\log\log t}=Z(\beta,t)\geq M\geq M_{1}=0.0470978\implies 1-\beta\leq\frac{\log\log t}{21.233\log t}, (4.135)

so that Corollary 1.2 is proved for tt0t\geq t_{0}.

5. Conclusion and future work

In this article we proved an explicit version of Littlewood’s zero-free region by deriving an explicit kkth derivative test using van der Corput’s Ak2BA^{k-2}B process. Our method represents a significant departure from recent approaches to proving classical zero-free regions. In this section we identify some potential refinements, which we have forgone for sake of brevity, but which may serve as basis for future research.

One avenue to improve Theorem 1.1 is to reduce the constants AkA_{k} and BkB_{k} appearing in the explicit kkth derivative test (Lemma 2.5). By specialising the argument to a specific value of kk, we can use sharper variants of the inequalities used in the general argument of Lemma 2.5. Alternatively, we can also achieve savings by specialising the choice of the phase function f(x)f(x) earlier on (this was demonstrated in [HPY24] for the ABAB process). Both approaches inevitably involve long and tedious arithmetic, however the iterative nature of the kkth derivative test means it may be well suited for an automated or computer-assisted proof program.

Theorem 1.1 is presented as a bound holding uniformly for all k4k\geq 4 and t3t\geq 3. However, the proof of Corollary 1.2 only requires a bound holding for ttkt\geq t_{k}, where tkt_{k}\to\infty as kk\to\infty. Therefore, one way to improve Corollary 1.2 is to replace Theorem 1.1 with a bound of the form |ζ(σk+it)|a(k)t1/(2k2)logt|\zeta(\sigma_{k}+it)|\leq a(k)t^{1/(2^{k}-2)}\log t (ttkt\geq t_{k}) with a(k)a(k) a decreasing function of kk. In particular, if a(k)0a(k)\to 0 sufficiently quickly, then we can derive an asymptotically better bound while leaving the rest of the argument largely unchanged. However, we anticipate limited benefits of applying this method to solely improve the constant of Corollary 1.2, since the bottleneck appears to be in small values of kk, and thus tt.

Lastly, we also note that the proof of Corollary 1.2 relies on multiple results that can be sharpened using Theorem 1.1. For instance, Lemma 4.5 and 4.6 primarily depend on an estimate of S(t)S(t), which in turn depend on upper bounds on ζ(s)\zeta(s) in the critical strip. This is readily obtained via Theorem 1.1 and the Phragmén–Lindelöf Principle. Another example is Lemma 4.4, where the Ford–Richert bound (4.8) is used, which for small tt can be improved by using a similar bound derived from Theorem 1.1.

Acknowledgements

I would like to thank my supervisor Timothy S. Trudgian for suggesting the initial idea for this project, and for various helpful ideas throughout the writing of this paper. Thanks also to the anonymous referee for useful feedback and suggestions.

References

  • [Bac16] R. J. Backlund “Über die Nullstellen der Riemannschen Zetafunktion” In Acta Mathematica 41, 1916, pp. 345–375
  • [BR02] G. Bastien and M. Rogalski “Convexité, complête monotonie et inégalités sur les fonctions zêta et gamma, sur les fonctions des opérateurs de Baskakov et sur des fonctions arithmétiques” In Canadian Journal of Mathematics 54.5 Cambridge University Press, 2002, pp. 916–944
  • [BI86] E. Bombieri and H. Iwaniec “On the order of ζ(12+it)\zeta(\frac{1}{2}+it) In Annali della Scuola Normale Superiore di Pisa - Classe di Scienze Ser. 4, 13.3 Scuola normale superiore, 1986, pp. 449–472
  • [Bor12] Olivier Bordelles “Arithmetic tales”, Universitext New York: Springer, 2012
  • [Bou16] J. Bourgain “Decoupling, exponential sums and the Riemann zeta function” In Journal of the American Mathematical Society 30.1, 2016, pp. 205–224
  • [Che00] Y. F. Cheng “An explicit zero-free region for the Riemann zeta-function” In The Rocky Mountain Journal of Mathematics 30.1 Rocky Mountain Mathematics Consortium, 2000, pp. 135–148
  • [CG04] Yuanyou F. Cheng and Sidney W. Graham “Explicit Estimates for the Riemann Zeta Function” In Rocky Mountain Journal of Mathematics 34.4, 2004, pp. 1261–1280
  • [Cor21] J.G. van der Corput “Zahlentheoretische Abschätzungen” In Mathematische Annalen 84, 1921, pp. 53–79
  • [For02] K. Ford “Vinogradov’s Integral and Bounds for the Riemann Zeta Function” In Proceedings of the London Mathematical Society 85.3, 2002, pp. 565–633
  • [For02a] K. Ford “Zero-free regions for the Riemann zeta function” In Number Theory for the Millennium, II (Urbana, IL, 2000) A K Peters, Natick, MA, 2002, pp. 25–56
  • [For22] K. Ford “Zero-free regions for the Riemann zeta function” Preprint available at arXiv:1910.08205, 2022
  • [Gab79] Wolfgang Gabcke “Neue Herleitung und explizite Restabschätzung der Riemann-Siegel-Formel”, 1979
  • [Gou04] X. Gourdon “The 101310^{13} first zeros of the Riemann Zeta function, and zeros computation at very large height” Available at http://numbers.computation.free.fr/Constants/Miscellaneous/zetazeros1e13-1e24.pdf, 2004
  • [GK91] S. W. Graham and Grigori Kolesnik “Van der Corput’s Method of Exponential Sums”, London Mathematical Society Lecture Note Series Cambridge University Press, 1991
  • [GR96] Andrew Granville and Olivier Ramaré “Explicit bounds on exponential sums and the scarcity of squarefree binomial coefficients” In Mathematika 43.1, 1996, pp. 73–107
  • [HSW21] Elchin Hasanalizade, Quanli Shen and Peng-Jie Wong “Counting zeros of the Riemann zeta function” In Journal of Number Theory 235, 2021, pp. 219–241
  • [HB92] D. R. Heath-Brown “Zero-Free Regions for Dirichlet L-Functions, and the Least Prime in an Arithmetic Progression” In Proceedings of the London Mathematical Society s3-64.2, 1992, pp. 265–338
  • [HP49] Fritz Herzog and George Piranian “Sets of convergence of Taylor series I” In Duke Mathematical Journal 16.3 Duke University Press, 1949, pp. 529 –534
  • [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
  • [HPY24] Ghaith A. Hiary, Dhir Patel and Andrew Yang “An improved explicit estimate for ζ(1/2+it)\zeta(1/2+it) In Journal of Number Theory 256, 2024, pp. 195–217
  • [JK14] Woo-Jin Jang and Soun-Hi Kwon “A note on Kadiri’s explicit zero free region for Riemann zeta function” In Journal of the Korean Mathematical Society 51.6, 2014, pp. 1291–1304
  • [Kad05] Habiba Kadiri “Une région explicite sans zéros pour la fonction ζ\zeta de Riemann” In Acta Arithmetica 117.4, 2005, pp. 303–339
  • [Kad13] Habiba Kadiri “A zero density result for the Riemann zeta function” In Acta Arithmetica 160.2, 2013, pp. 185–200
  • [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
  • [KT50] J. Karamata and M. Tomic “Sur une inégalité de Kusmin-Landau relative aux sommes trigonométriques et son application à la somme de Gauss.” In Publications de l’Institut Mathématique 3, 1950, pp. 207–218
  • [Kon77] V. P. Kondrat’ev “Some extremal properties of positive trigonometric polynomials” In Mathematical notes of the Academy of Sciences of the USSR 22.3, 1977, pp. 696–698
  • [Kuz27] R. Kuzmin “Sur quelques inégalités trigonométriques” In Soc. Phys.-Math. Léningrade 1, 1927, pp. 233–239
  • [Lan28] E. Landau “Üeber eine trigonometrische Summe” In Nachrichten von der Gesellschaft der Wissenschaften zu Göttingen, 1928, pp. 21–24
  • [LL22] Ethan S. Lee and Nicol Leong “Explicit bounds on the summatory function of the Möbius function using the Perron formula” Preprint available at arXiv:2208.06141, 2022
  • [Lit22] J. E. Littlewood “Researches in the theory of the Riemann ζ\zeta-function” In Proc. London Math. Soc. 20.2, 1922, pp. 22–27
  • [LRW86] J. Lune, H. J. J. Riele and D. T. Winter “On the zeros of the Riemann zeta function in the critical strip. IV” In Mathematics of Computation 46.174, 1986, pp. 667–681
  • [MT14] Michael J. Mossinghoff and Tim S. Trudgian “Nonnegative trigonometric polynomials and a zero-free region for the Riemann zeta-function” In Journal of Number Theory 157, 2014, pp. 329–349
  • [MTY24] Mike J. Mossinghoff, Timothy S. Trudgian and Andrew Yang “Explicit zero-free regions for the Riemann zeta-function” In Research in Number Theory 10.11, 2024
  • [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
  • [PT15] D. J. Platt and T. S. Trudgian “An improved explicit bound on |ζ(1/2+it)||\zeta(1/2+it)| In Journal of Number Theory 147, 2015, pp. 842–851
  • [PT21] D. J. Platt and T. S. Trudgian “The Riemann hypothesis is true up to 310123\cdot 10^{12} In Bull. Lond. Math. Soc. 53.3, 2021, pp. 792–797
  • [Pla17] David J. Platt “Isolating some non-trivial zeros of zeta” In Mathematics of Computation 86.307, 2017, pp. 2449–2467
  • [Rey11] J. Arias Reyna “High precision computation of Riemann’s zeta function by the Riemann-Siegel formula, I” In Mathematics of Computation 80.274, 2011, pp. 995–995
  • [Rey20] J. Arias Reyna “On Kuzmin-Landau Lemma” Preprint available at arXiv:2002.05982, 2020
  • [Ric67] H. E. Richert “Zur Abschätzung der Riemannschen Zetafunktion in der Nähe der Vertikalen σ=1\sigma=1 In Mathematische Annalen 169.1, 1967, pp. 97–101
  • [RS75] J. Barkley Rosser and Lowell Schoenfeld “Sharper Bounds for the Chebyshev Functions θ(x)\theta(x) and ψ(x)\psi(x) In Mathematics of Computation 29.129 American Mathematical Society, 1975, pp. 243–269
  • [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
  • [Ste70] S. B. Stechkin “Zeros of the Riemann zeta-function” In Mathematical notes of the Academy of Sciences of the USSR 8.4, 1970, pp. 706–711
  • [Tit86] E. C. Titchmarsh “The Theory of the Riemann Zeta-function” Oxford: Oxford Science Publications, 1986
  • [Tru14] T. S. Trudgian “A new upper bound for |ζ(1+it)||\zeta(1+it)| In Bulletin of the Australian Mathematical Society 89.2, 2014, pp. 259–264
  • [Tru15] Tim S. Trudgian “Explicit bounds on the logarithmic derivative and the reciprocal of the Riemann zeta-function” In Functiones et Approximatio Commentarii Mathematici 52.2, 2015, pp. 253–261
  • [Tru14a] Timothy S. 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
  • [Vin58] I. M. Vinogradov “Eine neue Absch ätzung der Funktion ζ(1+it)\zeta(1+it) In Izv. Akad. Nauk SSSR, Ser. Mat. 22, 1958, pp. 161–164
  • [Wed03] S. Wedeniwski “ZetaGrid — Computational verification of the Riemann Hypothesis” Banff, Alberta, Canada: Conference in number theory in honour of Professor H. C. Williams, 2003