Explicit bounds on in the critical strip and a zero-free region
Abstract.
We derive explicit upper bounds for the Riemann zeta-function on the lines for integer . This is used to show that the zeta-function has no zeroes in the region
This is the largest known zero-free region for . Our results rely on an explicit version of the van der Corput 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: 11Y351. Introduction
Bounding the size of the Riemann zeta-function within the critical strip is a central goal in analytic number theory, with results having broad implications for the zeroes of and thus the distribution of prime numbers [For02a, KLN18, Tru14a]. By far the most common result of this type are estimates of as and is fixed to either or , i.e. bounds along the critical line and 1-line respectively. Currently, the best known unconditional results are for any , due to Bourgain [Bou16], and , due to Vinogradov [Vin58]. Explicit results are also known, with the current best bound (for large ) being (), proved by Patel [Pat21] and () due to Trudgian [Tru14].
The main focus of this paper is to prove explicit bounds for for special values of . For many applications of interest, such as zero-free regions and zero-density estimates, we require bounds holding uniformly in the strip . The Vinogradov–Korobov zero-free region, for instance, relies on a Ford–Richert type result for all [Che00, Ric67, For02a]. Via a convexity argument (see e.g. [Tit86, §7.8]), we may use estimates of and to obtain bounds on for any . However, through van der Corput’s method of bounding exponential sums, we can directly derive asymptotically sharper bounds for specific values of . Titchmarsh [Tit86, Thm. 5.13] shows for instance that if for some integer , then
(1.1) |
This is sharper than what is achievable via convexity arguments for all .
Having an explicit version of (1.1) will allow us to improve many existing explicit results about . The main obstacle to making (1.1) explicit is the difficulty with bounding the implied constants of the th derivative test, obtained through van der Corput’s process. In this work we refine existing approaches to explicit exponential sum theory, to obtain an explicit th derivative test with constants holding uniformly for all . This allows us to show the following theorem.
Theorem 1.1.
Let be an integer and . Then
(1.2) |
For example, substituting gives . By comparison, the sharpest bound that can currently be obtained using bounds on , and the convexity principle is , where . Theorem 1.1 is sharpest for small to moderately sized and . In particular, as , (1.2) reduces to , which is weaker than other known bounds on the 1-line [Bac16, Pat22]. If we are only interested in large and , then Theorem 1.1 can be sharpened (see remarks in §5).
We briefly highlight some immediate applications of our results. The explicit th derivative test is an ingredient in deriving an explicit version of Littlewood’s bound [Tit86, Thm. 5.16]. Similarly, the th derivative test can be used to make explicit the bounds , which are useful for bounding Mertens’ function [Tru15, LL22]. Additionally, Theorem 1.1 can be used to improve explicit bounds on , 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 .
1.1. Littlewood’s zero-free region
Zero-free regions for 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 is due to Mossinghoff, Trudgian and Yang [MTY24], who proved that there are no zeroes of in the region
(1.3) |
For intermediate , the following zero-free region is currently the sharpest known
(1.4) |
where . This is formed by substituting [HPY24, Thm. 1.1] into [For02a, Thm. 3] and noting that for .
For large , the following Vinogradov–Korobov zero-free region due to [MTY24], building on the method of Ford [For02a, For22], is currently the sharpest known
(1.5) |
In this work we use Theorem 1.1 and the explicit th derivative test to prove the following zero-free region.
Corollary 1.2.
There are no zeroes of in the region
(1.6) |
This represents the largest known zero-free region in the range . To summarise the current state of knowledge for other ranges of : for , all zeroes are known to lie on the critical line, due to the computational verification performed in [PT21]. For , (1.3) is the sharpest known zero-free region; for , (1.4) is sharpest; for , (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
(1.7) |
where and is a sufficiently smooth function. In Titchmarsh [Tit86, Ch. V], it was shown that if has continuous derivatives satisfying , then
(1.8) |
for some unspecified absolute constants and . This is also known as a th derivative test, and the method of derivation was to use van der Corput’s process, where a single application of Poisson summation is followed by applications of the Weyl–van der Corput inequality. To our knowledge, to date the constants and in (1.8) have not been explicitly computed. However, Granville and Ramaré [GR96, Prop. 8.2] have proved an explicit bound of the form
(1.9) |
for . To facilitate a comparison, in our eventual application the value of is such that (1.8) and (1.9) respectively reduce to bounds of the form
Therefore, (1.8) produces a log-power saving (see also remarks after Lemma 2.5). More is known for small values of ; see for instance [Bor12, Thm. 6.9] for and Patel [Pat22] for . 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 and 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 cannot be controlled properly on that domain. Meanwhile, an process involves successive applications of Weyl-differencing, which for large 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 th 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 th-derivative tests corresponding to the exponent pair. The results of this section are agnostic to the choice of phase function . In §3 we specialise to a specific phase function to bound 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 th derivative test
The primary tool we use to bound in the critical strip is an upper bound on the exponential sum (defined in (1.7)), where is a smooth function possessing at least continuous derivatives. Four established methods of bounding 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
(2.1) |
arising from applying the triangle inequality and counting the maximum number of integers in . If is an integer, we can improve this to . The explicit Kuzmin–Landau lemma improves on the trivial bound if is sufficiently well-behaved. Let denote the distance to the nearest integer to . Suppose that is a real-valued function with a monotonic and continuous derivative on , satisfying . Then
(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 . In [KT50] and [HPY24], the following generalisation was proved. If is monotonic and continuous on , and for some integer and , then
(2.3) |
In practice, the conditions imposed on are rarely satisfied on the entire summation interval . 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 is real-valued and twice continuous differentiable on for some integers , with monotonic and satisfying
for some and . Then,
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 . 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 in the first two terms. We opt for this generalisation because in our eventual application (the th derivative test) it becomes increasingly difficult to leverage the benefits of specialising as grows.
Consider first the case if . Then, using the trivial bound, since and are integers, and using ,
(2.4) |
so the desired result is true for all .111We can expand the range of under consideration via the following argument, as remarked by Timothy S. Trudgian. For all , we have so the desired result once again follows from the trivial bound, as . This allows us to assume a sharper upper bound on in the subsequent argument. In the remainder of the proof we will assume that .
The conditions imposed on imply that either on or . Without loss of generality, assume that , since we may replace with without changing the value of . Due to the continuity of , there exists some for which
(2.5) |
Meanwhile, define
(2.6) |
and let and , where denotes the fractional part of . Let be a parameter to be chosen later, and let
(2.7) | |||||
By (2.5), we have
(2.8) |
Furthermore, since both and its inverse function are increasing, for we have
(2.9) |
for some satisfying
(2.10) |
Therefore, and hence
(2.11) |
Next, because by assumption, and since is increasing, we have . First, by the trivial bound and (2.11), we have for ,
(2.12) |
Next, in intervals of the form for , we have, by construction, . By Lemma 2.2, for ,
(2.13) |
It remains to consider the boundary sums and . Here we use the same argument as [HPY24]. First, consider . There are three cases.
Case 1: .
Case 2: .
We have and for all . By (2.3), we have
(2.15) |
Case 3: .
Then . This implies . Hence and thus in this case.
Combining the three cases, we conclude that
(2.16) |
where
(2.17) |
Via a similar argument, we have
(2.18) |
Combining (2.12), (2.13) and (2.8), we obtain that is majorised by
(2.19) |
We choose to minimise the second factor, noting that the upper bound on guarantees our choice satisfies the previous assumption that . By (2.19), we have
(2.20) |
where
(2.21) |
To complete the proof, it suffices to show that
(2.22) |
We once again consider three cases.
Case 1: .
Case 2: .
Then,
(2.24) |
is a convex function, so . However, for any we have
(2.25) |
and
(2.26) |
Case 3: .
From and ,
(2.27) |
as required. ∎
Remark.
A qualitatively similar result is historically obtained via Poisson summation (known as process , 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 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 and .
Lemma 2.2.
Proof.
To obtain higher-derivative tests, we use an explicit process, which makes use of the Weyl-differencing operation. Here, is expressed in terms of with for some integer . Intuitively, if is well-approximated by a degree polynomial on , with , then we can expect that is well-approximated by a degree polynomial on . We can achieve sharper bounds on since the lower order of means it likely satisfies the conditions of (2.3) over longer intervals, hence increasing the savings produced by the bound.
Lemma 2.3 (Explicit process).
Let be real-valued and defined on , for some integers . For all integers , we have
where .
Proof.
See [CG04, Lem. 5] and [PT15, Lem. 2]. The first factor originally appears as in [CG04] (upon making the substitution , ). As remarked in [PT15], it is possible to reduce this factor to via a more careful bound. Here, we further decrease the factor by 1 by assuming that and are integers, as sums over are equivalent to sums over . ∎
By using Lemma 2.1 to estimate , we obtain the following estimate, which is historically obtained by applying the Poisson summation formula then applying the process. The next lemma corresponds to an explicit process.
Lemma 2.4 (Explicit third derivative test).
Let have three continuous derivatives and suppose is monotonic and satisfies for all , where , and , are integers. Then, for any , we have
where
Proof.
We will first consider the case if . Using the trivial bound, since and are integers, we have
(2.30) |
and hence the desired result follows if . In the last inequality we have used and .
Next, consider the case of , where
(2.31) |
Note that satisfies
(2.32) |
We once again apply the trivial bound to obtain
(2.33) |
hence the desired result follows in this case too.
Suppose now that (we may assume that , since otherwise the proof is complete). Let as in Lemma 2.3. By the mean value theorem, since is continuous we have for some . Therefore, by the lemma’s assumption we have
(2.34) |
and hence we may apply Lemma 2.1 with and , since both and are integers. This gives
(2.35) |
We apply the inequality
(2.36) |
(see e.g. [Pat22, (32)] for and [Hia16, (45)] for ) to obtain, after using ,
(2.37) |
We choose for some , so that and the RHS is
(2.38) |
Substituting this estimate into Lemma 2.3,
(2.39) | ||||
(2.40) | ||||
(2.41) |
where the last inequality follows from the assumption that . Applying the assumption to the first factor, then taking square roots,
(2.42) | ||||
(2.43) |
since for all . However so the result follows. ∎
If we now use Lemma 2.4 instead of Lemma 2.1 to bound in Lemma 2.3, then we obtain the fourth-derivative test, via the process. Performing this substitution recursively as necessary, it is possible to derive an explicit process, for any . The following lemma is our main result.
Lemma 2.5 (Explicit th derivative test).
Let be integers with . Let be equipped with continuous derivatives, with monotonic, and suppose that for all and some . Then
where , and are defined in Lemma 2.4, and for are defined recursively via
(2.44) |
(2.45) |
(2.46) |
where .
Proof.
The central argument remains the same as that of Lemma 2.4, however to achieve a bound that holds uniformly for all without excessive tedium, we forsake sharpness in several key inequalities. One motivation for the separate treatment of the case is to establish good starting constants and 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 before switching to the general argument.
As before, we begin by considering a few edge cases. First, suppose . Since , for all we have
which implies that
(2.47) |
Therefore, from the arithmetic-geometric means inequality,
(2.48) |
The last expression is no smaller than if , which is true since
(2.49) |
and, since , ,
(2.50) |
Therefore, the desired result follows from the trivial bound.
Next, suppose and , where
(2.51) |
Since , we have for all , hence . Thus
(2.52) |
Hence the desired result once again follows from the trivial bound.
We now proceed to the main argument, assuming that and for all . By Lemma 2.4, the desired result holds for . Assume for an induction that the lemma holds for some . For convenience denote .
Suppose that has continuous derivatives on , such that for some . Let so that, via the mean value theorem,
(2.53) |
for all , and some . From the conditions on , we have
(2.54) |
By the inductive assumption, we have
(2.55) |
for some constants , not depending on or . Note we have used the inequality for simplicity. We once again apply (2.36) to obtain
where
(2.56) |
Substituting into (2.55), we obtain
(2.57) |
Using Lemma 2.3 and the inequality ,
(2.58) | ||||
(2.59) |
where
(2.60) |
We choose , so that we have (wastefully)
(2.61) |
This choice of gives, via the assumption and ,
(2.62) |
Additionally, since , we also have
(2.63) |
This gives the following estimates.
(2.64) |
(2.65) |
(2.66) |
(2.67) |
Combining these with (2.62) and (2.59), we have
(2.68) | ||||
(2.69) | ||||
(2.70) |
hence the induction is complete. ∎
Remark.
In our eventual application we have where , so that in particular,
Therefore, Lemma 2.5 implies (for fixed ) that . Meanwhile, so (1.9) reduces to . In particular, in our application we take when . Therefore, using (1.9) in place of Lemma 2.5 in the argument of Theorem 1.1 produces a bound of strength .
For many applications we are interested in uniform bounds holding for all . To this end we provide the following completely explicit result.
Lemma 2.6.
Let , be integers. Suppose is any function having continuous derivatives with monotonic, and for all and some , . Then
Proof.
From Lemma 2.4 and (2.44), we observe that is a decreasing function of , for all . It thus suffices to bound with . We choose in Lemma 2.4 and recursively compute , for , using Lemma 2.5, to ultimately obtain
(2.71) |
For , we note that is decreasing in , so . Also, for any we have
(2.72) |
By Lemma 2.5,
(2.73) |
The discrete map has a single stable fixed point
(2.74) |
Since and , we have for all . As for , for all ,
(2.75) |
The map has a single stable fixed point , so it follows from (2.71) that for all .
∎
3. Bounds on in the critical strip
In this section we use the explicit th derivative test to bound for certain values of , specifically those corresponding to
(3.1) |
for integers . Such values of lie in the interval , so that we are bounding 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 slightly to the left of . Using the th derivative test, we can establish asymptotically sharper bounds on than what is possible by considering bounds on the half-line and convexity principle alone [Tit86].
Throughout this section, we specialise the phase function encountered in lemmas 2.1, 2.4 and 2.5 to
(3.2) |
so that
(3.3) |
Our proof of Theorem 1.1 is divided into two sections. First, in §3.1 we prove the theorem for where
(3.4) |
The main tools in this range are the Phragmén–Lindelöf Principle combined with the following two bounds on and respectively:
(3.5) | ||||
(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 [Tru14, Pat22], however our argument requires a bound holding for all . Second, in §3.2, we prove Theorem 1.1 for 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 . We will write and to mean the largest integer no greater than , and the smallest integer no smaller than , respectively. Unless otherwise specified, with and .
3.1. Proof for
In this range, we use the following version of the Phragmén–Lindelöf Principle, due to Trudgian [Tru14a].
Lemma 3.1 (Phragmén–Lindelöf Principle).
Let be real numbers satisfying and . Suppose is a holomorphic function in such that for some and . Suppose further that
for some . Then, for all ,
Proof.
See [Tru14a, Lem. 3]. ∎
The motivation for using such a convexity argument for small is the bounds (3.5) and (3.6) are comparatively sharp for small . We choose the holomorphic function .
First, we verify numerically that if , then
(3.7) |
and if , then
(3.8) |
Therefore, by combining with (3.5) and (3.6) and using Lemma 3.1,
(3.9) |
where . If with , we have . If furthermore , then
and for , , hence for
(3.10) |
Hence
(3.11) |
where
(3.12) |
The RHS of (3.11) is majorized by if
(3.13) |
i.e. if
(3.14) |
Taking , we have , in which case for all . Therefore, (3.13) is satisfied for all . Similarly, taking , we have and
(3.15) |
It follows that
(3.16) |
as required.
3.2. Proof for
This subsection contains our main argument. We begin by bounding the difference between and its partial sum using Euler–Maclaurin summation. Recent explicit bounds on 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 instead of . Off the critical line, applying the Riemann–Siegel formula requires an explicit bound on the remainder term holding for . 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 where and . If , then
(3.17) |
Suppose now that are positive integers with and
(3.18) |
Throughout, we will write where and . Furthermore let
(3.19) |
for all . Roughly speaking, our approach is to use Lemma 3.2 then use partial summation and Lemma 2.5 to bound
(3.20) |
We divide the sum (3.20) into three subsums — in , we use the trivial bound, and in and we apply Lemma 2.6 with different choices of . Here, we ultimately make the choice
(3.21) |
for some scaling parameters , to be chosen later.
Lemma 3.3.
Let , where for some integer . Let and and let be any real number. Furthermore, suppose , are integers satisfying for some . Then
(3.22) |
where
(3.23) |
Proof.
3.2.1. The small region
Lemma 3.4.
Proof.
Let be a parameter to be chosen later, satisfying
(3.32) |
Define
(3.33) |
(3.34) |
First, observe that
(3.35) |
Hence, from the trivial bound we have
(3.36) |
since . If is so large that , then the sum on the LHS is empty while the RHS is positive, so the inequality holds regardless. Furthermore, as , we have, for ,
(3.37) |
Next, consider the sum over the interval . We divide the interval into pieces of the form , where
(3.38) |
Note that and , so the entire interval is covered. We divide the sum over into
(3.39) |
say. Recalling that , we take , and in (3.28) and combining with (3.30), we obtain
(3.40) | ||||
(3.41) | ||||
(3.42) |
where, passing from (3.40) to (3.41) we used
(3.43) |
and passing from (3.41) to (3.42) we used
(3.44) |
for all .
To bound of (3.39), we note that
(3.45) |
However is an integer so , i.e. there are at most terms in . Therefore, via the trivial bound, and noting that ,
(3.46) |
Therefore, writing and ,
(3.47) |
where
(3.48) |
Since we have
(3.49) |
Meanwhile, from the lower bound on , we have
(3.50) |
so that, in light of (3.48), (3.49) and (3.50), we have, for ,
(3.51) |
(3.52) |
Next, since , we have trivially
(3.53) |
Meanwhile, if and then
Therefore, combining the previous estimates, we have for ,
(3.54) | ||||
(3.55) |
where
is decreasing in . Combining with (3.37), the result follows. ∎
3.2.2. The large region
In this region we use Lemma 2.6 with smaller choices of .
Lemma 3.5.
Let and with . Then, for any , and , we have
(3.56) |
where and are defined in (3.21), and
(3.57) |
(3.58) |
and
(3.59) |
Proof.
We begin by defining
(3.60) |
where is defined in (3.19). Noting that and , we have
(3.61) |
We further divide each interval into intervals of the form , where
(3.62) |
and is the smallest integer for which . Note in particular that since ,
(3.63) |
so that
(3.64) |
Consider now the sum over . We divide the sum as before, with
(3.65) |
Since , we may apply (3.28) to the first sum with and to obtain, for and ,
(3.66) |
where
(3.67) |
For , we have the identities
(3.68) |
(3.69) |
Therefore, noting that is decreasing in and for all , we have, via (3.68),
(3.70) |
where is defined in (3.59). Similarly, via (3.69) and we have
For the second sum in (3.65), we note as before that
and hence, since is an integer, we in fact have , which is the maximum number of terms in the second sum of . Therefore, using the fact that is decreasing and ,
(3.71) |
Combining (3.66), (3.70) and (3.71) gives
(3.72) | ||||
(3.73) |
Using
(3.74) |
and substituting into (3.61), we obtain
(3.75) |
where is defined in (3.57).
Next, we will show that if is fixed, then , and hence , is a decreasing function of , so that for . Upon noting that for for , we have
(3.76) |
and hence
(3.77) |
Additionally,
(3.78) |
Finally, from the assumption that and ,
(3.79) |
and so is decreasing, so for . ∎
3.2.3. Putting it all together
For each row of Table 1, we substitute the appropriate variables into Lemma 3.4 and 3.5 and take . This gives
(3.80) |
Subsequently, taking and in Lemma 3.2,
(3.81) |
for , where
(3.82) |
The last inequality of (3.81) follows from
(3.83) |
for all . This proves Theorem 1.1 for and .
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 . We take , and (as in Lemma 2.6). This choice of parameters gives us and, by Lemma 2.6,
(3.84) |
We will show that is decreasing in for , for fixed and . Let
(3.85) |
so that, since is decreasing in ,
(3.86) |
However,
(3.87) |
and thus , i.e. is decreasing. Therefore, the factor appearing in (3.58) is decreasing in for fixed . Since , the only other way depends on is through the factor
(3.88) |
which is decreasing in since . Hence, is decreasing in . Therefore, for all ,
(3.89) |
with the understanding that if then the last sum is 0. Now, for all , we have
(3.90) |
hence
(3.91) |
from which it follows that
(3.92) |
This implies that for ,
(3.93) |
Substituting (3.93) into (3.58), and using , we obtain
(3.94) |
Next, for , and given the choices of , ,
(3.95) |
Meanwhile, for ,
(3.96) |
If , then via Lemma 2.6, since , we have , and
(3.97) |
(3.98) |
so that
(3.99) |
Combining (3.94), (3.95), (3.96), and (3.99), we have
(3.100) |
Hence, using in (3.89),
(3.101) |
Finally, substituting into Lemma 3.4,
(3.102) |
First, we use to obtain
(3.103) |
hence the first two terms of (3.102) are majorized by
(3.104) |
Next,
(3.105) |
and, as before, and since we have chosen and . Combined with the trivial bound , the third term of (3.102) is no greater than .
Finally, to bound the last term we combine the (wasteful) estimates
(3.106) |
each valid for . Combined with and the decreasingness of for , the last term of (3.102) is bounded by
(3.107) |
Hence, combining (3.104), (3.105) and (3.107), we finally have
(3.108) |
Combining with (3.101) gives
(3.109) |
Meanwhile, applying and to (3.82) gives the crude bound . Thus
(3.110) |
for all and , as required.
4. Proof of Corollary 1.2
Existing methods of deriving explicit zero-free regions of fall into two broad categories — the “global” approach of Hadamard, and the “local” method of Landau. The first method relies on explicit estimates of to the right of the line , which in turn rely on estimates of [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 slightly to the left of , 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
(4.1) |
uniformly for , for some constant and any . On the other hand, the classical zero-free region can be obtained via a bound of the form
(4.2) |
for some fixed and any [For02a]. In this work, we use Theorem 1.1 to obtain a bound of the form
(4.3) |
for some functions and as , to prove Corollary 1.2.
Both methods also make critical use of a non-negative trigonometric polynomial , of the form
(4.4) |
such that , and . In this work, we use the polynomial of degree presented in [MTY24, Table 1], which was found via a large-scale computational search. The coefficients of this polynomial satisfy
(4.5) |
4.1. Notation
Throughout, let be an integer, and let
(4.6) |
The variables and will be reserved for the trigonometric polynomial (4.5). Non-trivial zeroes of are denoted by and . We also find it convenient to write
(4.7) |
The variables and will be used to denote the constants appearing in the Ford–Richert bound
(4.8) |
This inequality is instrumental in the proof of Ford’s [For02a] Vinogradov–Korobov zero-free region (1.5). Throughout, we write
(4.9) |
where the zeroes are counted with multiplicity.
Our argument is roughly the same as [For02a] — an upper bound on 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 , we use Theorem 1.1 and [HPY24, Thm. 1.1], combined with the Phragmén–Lindelöf Principle to bound uniformly in . For large , we use Theorem 1.1 directly by setting , tending to infinity with . 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 , taken on a vertical line within the critical strip. Ultimately, this bound is combined with an upper bound on 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 to hold for large instead of for all . This technical change allows us to avoid applying the Phragmén–Lindelöf Principle (Lemma 3.1) at very small values of where it is poorly suited.
Lemma 4.1.
Let , , and , and suppose
for some . Then
Proof.
We follow essentially the same argument as [For02a, Lem. 3.4] by splitting the integral into four parts. Let
(4.10) |
To bound , if then and by the lemma’s assumptions, we have
(4.11) |
where in the second inequality we have used with . Therefore,
(4.12) |
Next, if , then , so by the lemma’s assumption, we have and
(4.13) |
If , then and we use
Letting , we have . Also, . Using and , we have
so that
(4.14) |
Finally, if , then . By the lemma’s assumption, and applying and with , gives
so that
(4.15) |
Combining (4.12), (4.13), (4.14) and (4.15), we have
(4.16) |
However, since ,
Using , , and , the first term on the RHS of (4.16) is majorised by
(4.17) |
Next, letting and using ,
(4.18) |
Finally, if and then and that . Hence
(4.19) |
where the last inequality follows from and evaluating the integral explicitly. Combining (4.17), (4.18) and (4.19), we have
(4.20) |
∎
We use the above lemma in two forms - the first (Lemma 4.1 below) is used for large and the second (Lemma 4.2 below) is used for small values of .
Lemma 4.2.
Let be an integer, and . Then
Proof.
Lemma 4.3.
For all , and , we have
Proof.
Let . We begin with the estimates
with . These bounds are verified numerically for , then combined with [HPY24, Thm. 1.1] and the case in Theorem 1.1, respectively. Applying Lemma 3.1, for all ,
(4.21) |
Next, we combine the estimates (for and )
to obtain
(4.22) |
where
(4.23) |
(4.24) |
Substituting , , gives
The result follows from applying Lemma 4.1. ∎
Next, we slightly modify a lemma due to Ford [For02a], which bounds , the number of zeroes satisfying (counted with multiplicity). The main change is to use sharper bounds on to the right of the 1-line, and to change the constants so that the lemma continues to hold for . Note that a result like Lemma 4.1 can be used to produce a sharper bound for small values of , however we do not pursue this here for sake of brevity.
Lemma 4.4.
For and , we have
(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
(4.26) |
where is the Euler-Mascheroni constant, to obtain
(4.27) |
Note that we are using in place of 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 ,
However, for we have , so the result follows from substituting and originating from (4.8).222Here we applied (4.8) with where . Therefore, we need to extend the range of validity of this bound to . This is permissible by examining the proof of [For02, Lem. 7.1] (in fact, (4.8) is sharpest near ). ∎
Next, we use the above lemma to bound a sum over zeroes away from , reproduced below. This result is the same as [For02a, Lem 4.3], except we use a sharper estimate of , due to Trudgian [Tru14a], and extend the range of permissible values of up to . We note that for large , the estimate of due to [HSW21] is sharper, however our results are most sensitive to sharpness of the estimate for “small” . Using Theorem 1.1, it is possible to further refine the arguments of [Tru14a] and [HSW21] to improve bounds on , and thus improve Lemma 4.5. We leave such considerations for possible future work (see remarks in §5).
Lemma 4.5.
Let . If , then
Proof.
We follow the approach taken by Ford [For02a, Lem. 4.3]. We divide the zeroes satisfying into the following sets
(4.28) |
for some to be chosen later. We will also assume that , this will be verified after choosing . Let denote the number of zeroes of , with multiplicity, satisfying . 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.
(4.29) |
Therefore, for
(4.30) |
for some satisfying
(4.31) |
Let
(4.32) |
Then, since there are no zeroes with , and using ,
where, since
and
Since if , we have that and
Therefore, for and ,
Furthermore,
Finally, if then and for , we have
Therefore
(4.33) |
Next,
(4.34) |
where
Next,
(4.35) |
Therefore, combining (4.33), (4.34) and (4.35),
(4.36) |
where
Next, if , then
(4.37) |
We also have
If , then, since is increasing,
and, since and are both concave, we have and by Jensen’s inequality. Hence
so that
(4.38) |
where
Next, since ,
(4.39) |
and
Therefore, combining (4.37), (4.38) and (4.39),
(4.40) |
Hence finally, combining (4.36) and (4.40),
where
We choose , , and to minimise the bound when , , 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 , than the “main” term, which is bounded using Lemma 4.1. However, the values of 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 .
Lemma 4.6.
If and , then
where .
Proof.
For some to be chosen later (independent of the parameter appearing in Lemma 4.5), define
(4.41) |
and let , be defined analogously to (4.32). This is equivalent to setting in (4.28). We use a similar method as before to bound , however, since is now greater than 1, we drop the term in the bound of to ensure the monotonicity argument remains valid. Explicitly, we have
(4.42) |
where
Next, similarly to before,
(4.43) |
and
(4.44) |
where
Furthermore, since each zero in contributes at most each to the sum , and there are of them,
(4.45) |
Combining (4.42), (4.43), (4.44) and (4.45),
(4.46) |
where
Choosing and 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 .
Lemma 4.7 ([For02a] Lemma 4.6).
Let , be a real function with a continuous derivative, and a Laplace transform that is absolutely convergent for . Let and suppose
for some constant . Then, if with ,
and
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 with much larger than 1000, however this has no noticeable effect in the final result.
The choice of the smoothing function in the above lemma has a direct impact on the size of the resulting zero-free region. Here, we choose 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 to be fixed later, define
(4.47) |
where, successively,
(4.48) |
(4.49) |
and is defined as the unique solution to
(4.50) |
where and are coefficients of the non-negative trigonometric polynomial (4.5). Via a direct computation, we have
(4.51) |
(4.52) |
(4.53) |
We furthermore require information about the Laplace transforms of and . Let
(4.54) |
denote the Laplace transforms of and respectively, and for convenience write
(4.55) |
The function has an explicit formula (given in [For02a, (7.3)] 444Note that in [For02a], , and appear as , and respectively. We have renamed these functions to prevent confusion with , the product log function, which appears later.) in terms of which can be used to bound via the relation
(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 and ,
(4.57) |
where
(4.58) |
and
(4.59) |
where, as in [For02a], and using (4.51),
(4.60) |
We also have, via a direct computation using (4.51) and (4.54),
(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 .
Lemma 4.8.
Let be an integer and . Suppose is a zero satisfying and . Furthermore, suppose that there are no zeroes in the rectangle
where is a constant satisfying . Then
with 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 , more care is taken with bounding secondary error terms which can be significant for small . Throughout, we retain the definitions of , , , , , , and encountered previously.
If , then via the lemma’s assumptions. Hence
(4.62) |
Therefore the conditions of Lemma 4.7 are satisfied with and . Since , we may apply Lemma 4.7 with with . Summing the resulting expressions, and using the non-negativity of the trigonometric polynomial (4.5), we obtain
(4.63) |
We now seek an upper bound for the right-hand side of (4.63). First, by [For02a, Lem. 5.1], we have
(4.64) |
Second, note that
(4.65) |
so that, by Lemma 4.1,
(4.66) |
Third, by Lemma 4.5, and using for ,
(4.67) |
Fourth, from [For02a, (7.7)], if , and , then
(4.68) |
where
(4.69) |
We use (4.68) for . If , then . Furthermore,
(4.70) |
and by assumption, so and
(4.71) |
Combining (4.68) and (LABEL:classic_bound4), and noting that ,
(4.72) | |||
Finally we also have , so that
(4.73) |
Substituting (4.64), (4.66), (4.67), (4.72) and (4.73) into (4.63), we obtain
(4.74) |
However, for all ,
(4.75) |
so we may drop the sum in (4.74). Furthermore,
(4.76) |
Note that is continuous and decreasing for , so by the mean-value theorem, and using (4.61) followed by (4.52) and (4.53),
(4.77) |
Combining with (4.76), (4.75) and (4.77) with (4.74), we have
(4.78) | |||
Dividing through by , 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 and . If , then Corollary 1.2 follows from the partial verification of the Riemann Hypothesis up to height , performed independently in e.g. [LRW86, Wed03, Gou04, Pla17, PT21]. Note that in [PT21], the Riemann Hypothesis is actually verified up to height , however our results do not require this, and the lower value of has been independently verified by multiple authors. If , then Corollary 1.2 follows from (1.4). Assume now that . Throughout, let
(4.79) |
(4.80) |
For integer let
(4.81) |
so that
(4.82) |
where is the principal branch of the Lambert function, which satisfies for any .
Note that implies Corollary 1.2. Assume for a contradiction that . Then, there exists a zero for which is arbitrarily close to . In particular, let be a zero such that and satisfying
(4.83) |
In particular, we have . If we let
(4.84) |
then there are no zeroes of in the rectangle
(4.85) |
This is because if a zero exists in that rectangle, then as is decreasing for , and ,
(4.86) |
so that . If , then this is in contradiction to the definition of . On the other hand if , then by (1.4) and a short numerical computation,
Thus, , which is a contradiction. Therefore, (4.85) is zero-free, and any zero must satisfy
(4.87) |
We now divide our argument into two parts. For , our argument relies on a uniform bound on for . 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 close to , however are insufficient to obtain an asymptotic estimate of the required strength. Therefore, for we switch to bounds on , where we take as , in the form of Lemma 4.8.
First, consider the case . Since , there is always an integer such that . Then, since is increasing for all , and , we have, for ,
(4.88) |
where, since ,
(4.89) |
Note that
(4.90) |
where
(4.91) |
Since , we have , i.e. for sufficiently large . We in fact find computationally that for all . It follows from (4.90) that if , then
(4.92) |
This also implies that
(4.93) |
Then
(4.94) |
i.e. with . This allows us to compute, via (4.58) and (4.59),
(4.95) |
Collecting (4.87) and (4.94), all conditions of Lemma 4.8 are met with , . We thus have, for all zeroes such that ,
(4.96) |
We now proceed to bound each term appearing on the right hand side. First,
(4.97) |
Furthermore, is decreasing for , since
(4.98) |
Taking and combining with (4.97) and (4.82) gives
(4.99) |
where, since and ,
(4.100) |
Next, using (4.88), we have
(4.101) |
where
(4.102) |
Next, since is decreasing for , by (4.88) we have
(4.103) |
where
(4.104) |
Finally, using (4.26) with , we have
(4.105) |
Combining (4.93), (4.99), (4.101), (4.103) and (4.105),
(4.106) |
where, from (4.5) we have , , and
(4.107) |
This constitutes the “main” term. Next, from (4.88) and using (4.83),
(4.108) |
where
(4.109) |
The remaining term of (4.96) is majorised by
(4.110) |
We have, by (4.88), for ,
(4.111) |
so that
(4.112) |
where, using (4.92),
(4.113) |
By solving for stationary points explicitly, we find that for ,
(4.114) |
Meanwhile, we also have the estimates
(4.115) |
where, since and ,
(4.116) |
and, since is increasing and is decreasing for ,
(4.117) |
where
(4.118) |
Next, we have for . Combining this with the bound for and (4.115), we have that (4.110) is , where
(4.119) |
Lastly, applying (4.83) we have
(4.120) |
where, since ,
(4.121) |
Combining (4.96), (4.107), (4.108), (4.119), (4.120) and (4.121), and from the definition of ,
(4.122) |
so the desired contradiction is reached if .
For , we use a similar argument, except with Lemma 4.3 and Lemma 4.6. Here, we choose
(4.123) |
These parameters guarantee that, for all , we have . Observe that for , we have
so that
(4.124) |
Here, the last inequality follows from solving explicitly the maximum of for . Therefore, we have
so that with . This implies that .
Since for all , 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,
(4.125) |
where, as before, . First, analogously to before, we have
(4.126) |
Next, since and by (4.26),
(4.127) |
Hence
(4.128) |
where the last inequality follows from
(4.129) |
Next, note that since ,
so that
(4.130) |
where
(4.131) |
Furthermore,
(4.132) |
where, since and , we may take
(4.133) |
Since is decreasing in for , we have . Therefore, collecting (4.125), (4.126), (4.128) and (4.132), and from the definition of ,
(4.134) |
hence we have also obtained the desired contradiction for .
5. Conclusion and future work
In this article we proved an explicit version of Littlewood’s zero-free region by deriving an explicit th derivative test using van der Corput’s 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 and appearing in the explicit th derivative test (Lemma 2.5). By specialising the argument to a specific value of , 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 earlier on (this was demonstrated in [HPY24] for the process). Both approaches inevitably involve long and tedious arithmetic, however the iterative nature of the th 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 and . However, the proof of Corollary 1.2 only requires a bound holding for , where as . Therefore, one way to improve Corollary 1.2 is to replace Theorem 1.1 with a bound of the form () with a decreasing function of . In particular, if 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 , and thus .
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 , which in turn depend on upper bounds on 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 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 ” 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 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 ” In Indagationes Mathematicae 27.2, 2016, pp. 524–533
- [HPY24] Ghaith A. Hiary, Dhir Patel and Andrew Yang “An improved explicit estimate for ” 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 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 -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 ” In Indagationes Mathematicae 33.5, 2022, pp. 1012–1032
- [PT15] D. J. Platt and T. S. Trudgian “An improved explicit bound on ” 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 ” 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 ” In Mathematische Annalen 169.1, 1967, pp. 97–101
- [RS75] J. Barkley Rosser and Lowell Schoenfeld “Sharper Bounds for the Chebyshev Functions and ” 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 ” 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 ” 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