Explicit bounds for the Riemann zeta function
and a new zero free region
Abstract.
We prove that for and . As a consequence, we improve the explicit zero-free region for , showing that has no zeros in the region for and asymptotically in the region for sufficiently large.
Key words and phrases:
Riemann zeta function, explicit bounds, zero-free region, Vinogradov integral2020 Mathematics Subject Classification:
Primary 11M06, 11N05, 11L15; Secondary 11D72, 11M35.1. Introduction
Let be the Riemann zeta function, with a complex variable. It is known that all the non trivial zeros of have real part . Detecting zero free regions for inside the critical strip is an open problem that has always caught much interest in analytic number theory. Great effort has been put in trying to find both asymptotically and explicit regions inside the critical strip where there are no zeros of . The classical zero-free region is of the form , where is a positive constant. The best known result of this form is due to Mossinghoff, Trudgian and Yang [MTY22] with for every (see [Ste70, RS75, Kon77, Kad05, JK14, MT14] for previous results). Littlewood zero-free region [Lit22] is instead of the form of , where is a positive constant. It has been made first explicit by Yang [Yan23], who found for .
Asymptotically larger zero-free regions for the Riemann zeta function , known as Korobov–Vinogradov zero-free regions, are of the form
(1.1) |
where is a positive constant and are due to the method of Korobov [Kor58] and Vinogradov [Vin58], in which the main tool is an upper bound on when is near to the line . Upper bounds of this form were first made explicit by Richert [Ric67], who used Korobov-Vinogradov method to prove that
(1.2) |
with and a certain absolute constant. Although smaller values for were already found (see [Kul99]), the first completely explicit bound of the form (1.2) is due to Cheng [Che99], with and . This estimate was further improved in 2002 by Ford [For02a] to and for every and .
Explicit bounds of the form (1.2) play a fundamental role in detecting explicit Korobov–Vinogradov zero-free regions for . There are several results regarding the value of the constant in (1.1) and the current best known estimate is due to Mossinghoff, Trudgian and Yang [MTY22], with for every (see [Che00, For02, For22] for previous results).
In this paper, we will improve the values for both the constants in (1.2) and, as a consequence, we find an improved Korobov-Vinogradov zero-free region for . More precisely, denoting with the Hurwitz zeta function defined for every and , we will prove the following result for both and a generic Hurwitz function .
Theorem 1.1.
The following estimate holds for every and :
(1.3) | ||||
with and .
The bound on found in Theorem 1.1 might be useful to bound Dirichlet -functions due to the relation , where is a Dirichlet character modulo .
Although the new values and found in Theorem 1.1 are modest improvements on those found by Ford in [For02a] ( and respectively), their importance relies on the fact that just a small improvement for both and can lead to improvements in several other results in analytic number theory. In particular, they have many applications in finding improved estimates for Korobov–Vinogrado zero-free region and in estimating the error term in the prime number theorem, both in an effective and ineffective way.
Furthermore, improvements on and only of the size found in Theorem 1.1 were already expected. In [For02a] (Section 8, point ), Ford already predicted that an optimization of the argument involving the Vinogradov integral would have lower by less than , which is consistent with our new value for found in Theorem 1.1.
As already mentioned, an immediate consequence of Theorem 1.1 is a new explicit zero free region for the Riemann zeta function.
Theorem 1.2.
There are no zeros of for and
Theorem 1.1 has also an influence on asymptotically Korobov-Vinogradov zero-free regions of the form
(1.4) |
where is a positive constant for sufficiently large. More precisely, we get , improving on the current best value of due to Mossinghoff, Trudgian and Yang [MTY22] (see [For02] for previous results).
Theorem 1.3.
For sufficiently large there are no zeros of with
(1.5) |
As in [MTY22], the proofs of Theorem 1.2 and Theorem 1.3 rely on some non-negative trigonometric polynomials. We recall that, for every , a -th degree non-negative trigonometric polynomial is defined as
(1.6) |
where are constants such that and for all real .
Finally, one can use Theorem 1.3 to improve on the error term in the prime number theorem. As per Ford [For02a], we get the following estimate for the error term in the prime number theorem
with
where is the constant in (1.4). Using the new value for found in Theorem 1.3 we obtain , which is a slightly improvement on found by Mossinghoff, Trudgian and Yang [MTY22].
As in [For02a], Theorem 1.1 follows from a uniform upper bound on the sum
(1.7) |
where is a positive integer and . This upper bound is of the form
with and The strength of Ford’s argument in [For02a] relies on some explicit estimates for both the Vinogradov integral and a quantity that counts the number of solutions of incomplete Diophantine systems that, combined with estimates for the exponential sum , give a completely explicit uniform upper bound on . Explicit bounds for the Vinogradov integral were further improved by Preobrazhenskiĭ [Pre11] in 2011 and Steiner in 2019 [Ste19] (see [Hua49, Ste70, Tyr87, ACK04] for previous results).
We recall that the Vinogradov integral is defined as
(1.8) |
where and , or equivalently, is defined as the number of solutions of the simultaneous equations
Regarding incomplete systems, we denote with the number of solutions of the system
where is a suitable set. For our purpose, we will use the definition given by Ford in [For02a] where is the set of integers composed only of prime factors in . Hence, is defined as
(1.9) |
where
and .
In his paper [For02a], Ford estimated working separately and with different techniques on three different ranges of . More precisely, he considered the ranges , and , where the critical case happens to be around . However, it is possible to shift the critical case to , with a consequent improved upper bound on and so improved values of and . Indeed, in his paper [For02a] (p. 590), Ford stated a variant of his argument to find even better estimates for the Vinogradov integral when is small. These results, combined with Preobrazhenskiĭ’s argument for in [Pre11] and some explicit bounds for the Vinogradov integral when due to Tyrina [Tyr87], will give the following explicit bounds for .
Theorem 1.4.
Let and be integers with and
with . Then
where, denoting with the unique integer such that ,
(1.10) | ||||
Furthermore, if , there is an integer such that for ,
with
Using these new bounds for the Vinogradov integral we will prove the following result.
Theorem 1.5.
Suppose is a positive integer, and set . Then
In proving both Theorem 1.4 and Theorem 1.5, we tried to make near-optimal choices for all parameters that are used in the argument. Hence, it seems unlikely that a substantial improvement in the constant in Theorem 1.1 can be obtained via better choices of parameters alone.
An immediate corollary of Theorem 1.5 involving Dirichlet characters is the following one.
Corollary 1.6.
Suppose is a Dirichlet character modulo , where and . Then
The proof is the same as that of Corollary 2A in [For02a].
2. Background
We recall some preliminary results we will use later in the proof of the theorems.
First of all, given a fixed , suppose and is a positive integer. The -tuple of polynomials is said to be of type if is identically zero for , and for some integer , when has degree with leading coefficient .
Then, let be the number of solutions of
with and . Also, let be the number of solutions of
with , and .
Lemma 2.1 ([For02a] Lemma 3.2’).
Suppose and are integers with
Let and be real numbers with
Suppose is a positive integer and is a system of polynomials of type with . Denote by the set of the smallest primes greater than , and suppose . Then there are a system of polynomials of type and a prime such that
Lemma 2.2 ([For02a] Lemma 3.3’).
Suppose that is a prime and is a system of polynomials of type . Then there is a system of polynomials of type with such that
Lemma 2.3.
Suppose and
Let be an integer satisfying
(2.1) |
Define
and suppose and are chosen so that for every . Suppose
If , then
where
We will use the definition of given in Lemma 2.3 for the range . However, when becomes large, the influence that the new funcitons have on estimating the Vinogradov’s integral becomes negligible, as already pointed out by Ford in [For02a] (p. 590). Hence, for , we use the following result used by Ford in his work.
Lemma 2.4 ([For02a] Lemma 3.4).
Suppose and
Let be an integer satisfying the same relations as in (2.1). Then, define
and suppose and are chosen so that for every . Suppose
If , then
where
For a given and , we let be the value of coming from Lemma 2.3 or Lemma 2.4, where we take maximal satisfying (2.1).
Lemma 2.5 ([For02a] Lemma 3.5).
As already mentioned before, some explicit bounds for the Vinogradov’s integral were found by Tyrina [Tyr87], and they give evident improvements when . In her argument, Tyrina defines recursively sequences , and , with and . More precisely, for we take to be the integer nearest to the number
and we define recursively
where
Then, we consider the function :
This function increases monotonically and assumes the values from to infinity as increases from to . The function which is inverse to is also monotonically increasing. Finally, let denote the integer , where
Tyrina proved the following result.
Theorem 2.6 ([Tyr87] Thm. 1).
The mean value satisfies the estimate
where . Here
-
(1)
if , then ;
-
(2)
if , then
-
(3)
if , then ;
-
(4)
if , then .
New explicit bounds for were found by Ford in 2002 [For02a].
Theorem 2.7 ([For02a] Thm. 3).
Let and be integers with and . Then
where
Further, if , there is an integer such that for ,
with
(2.2) |
In 2011, Preobrazhenskiĭ [Pre11] improved the value of in Theorem 2.7 when to . Some further explicit bounds for the Vinogradov’s integral were found also by Steiner [Ste19] using Wooley’s efficient congruencing method [Woo12, Woo13, Woo17]. He proved that for every , , and , we have , where is roughly . However, although the exponent of is the optimal one, the constant is far too large compared to that one found by Ford in Theorem 2.7, which is of order . In order to have a reasonable estimate for , one should find a constant that is at most of order , since otherwise, after having taken the th root (this passage is required by the argument in Section 4, where the constant appearing in the upper bound for the Vinogradov integral will be elevated to the power of , with both and of order ), the constant cannot be controlled. Hence, as already Steiner pointed out in his paper ([Ste19], p. 359), the estimate he found for cannot be applied to Ford’s argument.
Theorem 2.8 ([For02a] Thm.4).
Let , , and for some . Assume that
Then
where
Finally, we recall a few results found in [For02a] involving both defined in (1.7) and some bounds for both and that we will use to prove Theorem 1.5 and Theorem 1.1 respectively.
Lemma 2.9 ([For02a] Lemma 5.1).
Suppose and are integers , and and are integers satisfying . Let be a positive integer, and be real numbers with . Let be a nonempty subset of the positive integers . Then
where
Lemma 2.10 ([For02a], Lemma 7.1).
Suppose and . If either or , then
Lemma 2.11 ([For02a], Lemma 7.2).
If and , then
Lemma 2.12 ([For02a], Lemma 7.3).
Suppose that for positive constants and , where . Let . Then, for and , we have
3. Proof of Theorem 1.4
Let . Due to the size of , we are in the first or third cases of Tyrina’s Theorem 2.6. However, since the third case gives an estimate for those between two consecutive for which we have an estimate in the first case, we can restrict ourselves to the first case. By Theorem 2.6, we know that
where is the inverse function of , with
A direct computation gives the following explicit expression for :
where is the principal branch of the Lambert product-log function.
Furthermore, the function is monotonically increasing in the interval . Therefore, to lower-bound , it suffices to upper-bound . We want to find such that . This means that for every .
However,
(3.1) |
It follows that, given
we have
(3.2) |
since is decreasing in , for every fixed .
In order to prove the first part of Theorem 1.4, we follow Ford’s method in [For02a] and we find an estimate for . By (3.2), we know that, given fixed, we have for every . Hence, in particular, if we denote , we have . As a result, for the case , we follow Ford’s method in Lemma 3.6 of [For02a] but instead of starting from , we start from .
Lemma 3.1.
For every , and
we have
with
and
Proof.
We take
in Lemma 2.5, where are defined as in Lemma 2.4. Then, we define for every . Now, we fix and we define , , , and .
If , then the upper bound for follows trivially, since
Hence, from now on, we just consider .
By (3.2), we know that, given fixed, we have for every . Hence, in particular, if we denote , we have . At this point, we follow Ford’s method in Lemma 3.6 of [For02a] but instead of starting from , we start from .
Let
Using the definition of , we have
Hence,
and so the hypotheses of Lemma 2.5 are satisfied.
As per Ford [For02a], we have
(3.3) |
Since , denoting with
we have
(3.4) | ||||
as
It follows that
As per Ford [For02a],
(3.5) | ||||
because
As a function of the real variable ,
has a positive second derivative and a minimum at
It follows that in the interval
the maximum occurs in
Hence,
(3.6) | ||||
and
We want to estimate the quantity
using Taylor expansion for . We observe that
Furthermore, for all we have
Now,
and
It follows that
(3.7) |
Now, we repeat the process for the term of second order. We have
As before, using the Taylor expansion in ( for ), we have
(3.8) |
However, in order to have an upper bound for
a direct computation shows that we need to use the slightly weaker bound
(3.9) |
where we have instead of . Using the upper bounds found in (3.7) and (3.9) we obtain
(3.10) |
From (3.5) and (3.10), it follows that
where in the last passage we used the fact that .
Now we define
and let
Since is increasing on , and hence in , we have
As per Ford [For02a], given
we have
Hence, using an iterative argument we get
Since we are working with , that is , and we found the inequality
(3.11) |
we have
It follows that
Therefore
Now,
Furthermore
and, since , and for , if we write , where , we have
Also, using the fact that , we have
Furthermore, for , we have
It follows that
and so
Now we shift our attention on the constant. As per Ford [For02a], to bound the constants , we choose in Lemma 2.5, so that
for every . Then,
(3.12) |
since it was proved by Ford in [For02a] for the wider range .
Now, let . By (3.12) and Lemma 2.5,
and, for ,
Iterating this last inequality gives, for ,
This concludes the proof of the lemma. ∎
Remark.
In Lemma 3.1, contrary to Lemma 3.6 in [For02a], we are not allow to consider as a lower bound for the range of for which the lemma holds. Indeed, the upper bound for , that is , is a decreasing function in . As a result, this upper bound reaches the minimum when, for fixed, . Substituting in the above expression, one gets , which is less than for . Hence, in order to have an estimate for and which is uniform for every and , we took as lower bound for , in order to ensure that for every and there exists always at least a value of in the range given in the hypotheses of Lemma 3.1. Also, the choice as lower bound for is admissible, as in Lemma 3.1 the starting point is which is less than for .
Now, we will use Lemma 3.1 to prove the first part of the theorem.
Given , every admissible such that , is of the form , where , and
Using Hölder’s inequality, we get
where and
By Lemma 3.1, we have
where
Further,
Hence,
This completes the first part of Theorem 1.4.
We now turn to the second part of the theorem.
For , we follow Ford’s argument for proving Theorem 2.7 in [For02a]. Running Program 1 listed in Section 9 with defined as in Lemma 2.3, we get the desired bounds for in the ranges
For we use Program 2 in Section 9, with defined as in Lemma 2.3. Program 2 differs from Program 1 in the definition of the variable , as, due to (3.2), we know that, for , for every . Hence, in order to initialize , we start from instead of starting from .
For , we follow Preobrazhenskiĭ’s argument [Pre11]. Taking in Lemma 3.1, we observe that for every ,
Hence, since from now on we will work with , we will restrict the range of in Lemma 3.1 to
(3.13) |
Choosing
(3.14) |
we have, by Lemma 3.1, that
where we used the fact that
Hence, for every fixed, there exists such that
where
and ( the value of follows from a uniform upper bound for the constant
found in the first part of Theorem 1.4 which holds for every , and in the range (3.13)). This estimate holds since, using our choice (3.14) for , we have and
Remark.
The starting point with is nearly the optimal choice. Indeed, taking with would imply a greater value for . A direct computation shows that, although the constant in the estimate of would decrease, the quantity would increase, leading to an overall estimate for which is worse than that one found by Ford in Lemma 3.6 of [For02a] when .
Furthermore, the lower bound for in (3.1) is optimal, as the inequality in (3.1) is no longer valid for . Finally, also the value in (3.1) is nearly optimal, as this inequality is satisfied only with and .
4. Proof of Theorem 1.5
We will consider the cases and separately.
4.1. Case
We combine Ford’s method [For02a] with the new estimates for the Vinogradov’s integral found in Theorem 1.4 to obtain an improved upper bound for in Lemma 2.9.
As in [For02a], we make the following assumptions:
Also, we define
and
(4.1) |
Following the proof of Theorem 2 in [For02a] for large, using the notation of Lemma 2.9, we have
(4.2) |
where
(4.3) | ||||
At this point, we shall take the near-optimal values
(4.4) |
where is taken from Theorem 1.4.
Furthermore, we choose and we want to estimate when , since otherwise for Theorem 1.5 follows trivially:
Finally, we consider
(4.5) |
and the following bounds for the quantity :
First of all, by (4.4) and Theorem 1.4 we have
(4.6) |
where and comes from Theorem 1.4.
Since , from (4.4) and (4.5) we have . Let and , where . By the assumption on , (4.5), the hypotheses of Theorem 2.8 are satisfied if we take and . Hence, by Theorem 2.8,
(4.7) |
where
Since by the hypotheses of Theorem 2.8 and the range (4.5) for we have
following exactly the argument used by Ford in the proof of Theorem 2 in [For02a] for large with instead of we get
where
By (4.4), we have
(4.8) |
and
It follows that
(4.9) |
Lemma 4.1.
For and we have the following estimate
Proof.
From Lemma 2.9, (4.2), (4.4), (4.6), (4.7) and (4.8) we have
(4.10) | ||||
Now, we choose
Furthermore, taking , we make the same assumptions as in Lemma 5.3 of [For02a]
where satisfies (4.5). Then, we bound the exponent of in each interval where each of the quantities and defined in (4.4) is constant. We also take constant values of and in , so that are also fixed.
As in [For02a], we define for the quantities
so that
It follows that
At this point we use Program 3 in Section 9 to find the best values for and , under the condition , so that, for ,
where and . Running Program 3 we can notice that in each interval we have and . The conclusion follows. ∎
Lemma 4.2.
For and we have the following estimate
Proof.
From Lemma 2.9, (4.2), (4.4), (4.6), (4.7) and (4.8) we have
(4.11) | ||||
Now, we assume
(4.12) |
where
(4.13) |
From (4.1), (4.12) and (4.13), the relation (4.5) holds and, furthermore, we have
From (4.4), (4.12) and (4.13) we have
Following exactly the proof of Lemma 5.2 in [For02a], with instead of , with the adjusted value of instead of , we still have
It remains to estimate . We have
Following Ford’s argument in [For02a], we have
where
Using Ford’s estimate for the expression inside the brackets in the definition of , we have
It follows that
In the range
is increasing, hence the maximum occurring at , with
Using from Theorem 1.4, it follows that
∎
Remark.
One might have considered the ranges and , with . However, a direct computation shows that even for really large values of , of the order of , and hence, by definition of , for values of much greater than , the improvement is negligible. Indeed, one can see that a new choice for would not influence the estimate for in the range we found in Lemma 4.1, while for the case , if , and hence , would be sufficiently large, the value of in Theorem 1.4 would be for sufficiently large, which is slightly smaller that we found for . However, a direct computation shows that this really small improvment on for sufficiently large would not lead to an improvement in Lemma 4.2, when and sufficently large.
4.2. Case
5. Proof of Theorem 1.1
First of all, since and , we restrict our attention to lying in the upper half-plane. Then, we consider separately the cases or or and . The main contribution will come from the case and .
5.1. Case
5.2. Case
Following the proof of Lemma 2.10 in [For02a], one has
for and .
If , from [For02a] we know that
If , from the argument of Lemma 2.10 in [For02a], one gets
Hence, we also have
Remark.
One can notice that the estimate found for and in this range is much better than that one found for and . Indeed, the constant we found in this current case is just , which is less than the final value .
5.3. Case
Remark.
The choice of comes from the proof of Lemma 7.1 in [For02a], where for .
With one should get . However, the influence of this second choice on the constant is negligible and it does not lead to any further improvements on .
6. Proof of Theorem 1.2
For we use the best known explicit Littlewood zero-free region due to Yang in [Yan23]. For we follow the argument used to prove Theorem 1.1 in [MTY22] which relies on a non-negative trigonometric polynomial defined by (1.6) with degree 40 having
Using the new values and found in Theorem 1.1 and making the following new assumptions
7. Proof of Theorem 1.3
8. Some possible improvements on
In this section we give some suggestions for some possible improvements on the constant in Theorem 1.1. We will provide some quite detailed proofs of some useful lemmas to help the reader follow the argument more easily.
Instead of the definition (1.7) for , one could consider the following sum
(8.1) |
with . For , we can find sharper estimates.
Theorem 8.1.
For and we have:
If one could find a bound of the form also for , then the constant in Theorem 1.1 might be reduced to around . We briefly outline the proof of Theorem 8.1.
We start with a preliminary lemma that is a more general version of Lemma 2.9.
Lemma 8.2.
Suppose and are integers , and and are integers satisfying . Let be a positive integer, and be real numbers with . Let be a nonempty subset of the positive integers . Then
where
and is defined in (8.1).
Proof.
We define . For , and , we have
For we have
Also for real and . Thus, for some ,
where and .
At this point, from Ford’s argument in Lemma 5.1 of [For02a], we have
The conclusion follows. ∎
At this point, from Ford’s argument in section of [For02a] we have the following bounds:
and
where
and are the same as for Theorem 1.5. If we use these bounds in Lemma 8.2 we get
(8.2) |
Now, we consider , since otherwise we have trivially
Lemma 8.3.
For and we have the following estimate:
Proof.
From Lemma 4.1 we know that
Using this result in (8.2) we get
where the last passage comes from the fact that, for , we have
Indeed, for and , we have
if
which holds, since with our choice . ∎
Lemma 8.4.
For and we have the following estimate:
Proof.
Theorem 8.1 follows immediately from Lemma 8.3 and Lemma 8.4.
At this point, if one could find an estimate of the form also for , then we can use a more general version of Lemma 2.12.
Lemma 8.5.
If and , then
Lemma 8.6.
Suppose that for positive constants and and , where . Let . Then, for and , we have
Proof.
Let
By Lemma 8.5, . Put . By partial summation,
where
As a function of is increasing on and decreasing on , where . Thus
where .
To bound the last integral, we make use of the inequality
where the maximum occurs near . Therefore
∎
However, Ford’s method in [For02a] for (or as in our paper), cannot be modified to extract the factor . Indeed, following the proof of Lemma 6.3 in [For02a], one should estimates the following quantity
where
However, since an estimate on the whole sum of over is required, instead of one for the maximum of over the interval , it is not possible to extract a factor at this step. A clever argument which would overcome this problem could lead to a suitable estimate also for the case , with a consequent improved value for .
9. Code listings
Program 1 for Theorem 1.4
Program 2 for Theorem 1.4
Program 3 for Lemma 4.1
Acknowledgements
I would like to thank my supervisor Timothy S. Trudgian for his support and helpful suggestions throughout the writing of this article.
References
- [ACK04] Gennady I Arkhipov, Vladimir N Chubarikov and Anatoly A Karatsuba In Trigonometric Sums in Number Theory and Analysis Berlin–New-York: Walter de Gruyter, 2004
- [Che99] Y. F. Cheng “An explicit upper bound for the Riemann zeta-function near the line ” In The Rocky Mountain Journal of Mathematics 29.1 Rocky Mountain Mathematics Consortium, 1999, pp. 115–140
- [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
- [For02] 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
- [For02a] Kevin Ford “Vinogradov’s integral and bounds for the Riemann zeta function” In Proceedings of the London Mathematical Society 85.3, 2002, pp. 565–633
- [Hua49] Loo-Keng Hua “An improvement of Vinogradov’s mean-value theorem and several applications” In The Quarterly Journal of Mathematics os-20.1, 1949, pp. 48–61
- [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
- [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
- [Kor58] N. M. Korobov “Estimates of trigonometric sums and their applications” In Uspehi Mat. Nauk, 13, 1958, pp. 185–192
- [Kul99] Mieczysław Kulas “Refinement of an estimate for the Hurwitz zeta function in a neighbourhood of the line ” In Acta Arithmetica 89.4, 1999, pp. 301–309
- [Lit22] J. E. Littlewood “Researches in the theory of the Riemann -function” In Proc. London Math. Soc. 20.2, 1922, pp. 22–27
- [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
- [MTY22] Michael J. Mossinghoff, Timothy S. Trudgian and Andrew Yang “Explicit zero-free regions for the Riemann zeta-function” arXiv:2212.06867, 2022
- [Pre11] S. N. Preobrazhenskiĭ “New estimate in Vinogradov’s mean-value theorem” In Mathematical Notes 89.1-2, 2011, pp. 277–290
- [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
- [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
- [Ste19] Raphael S. Steiner “Effective Vinogradov’s mean value theorem via efficient boxing” In Journal of Number Theory 204, 2019, pp. 354–404
- [Tyr87] O. V. Tyrina “A new estimate for a trigonometric integral of I. M. Vinogradov” In Izv. Akad. Nauk SSSR Ser. Mat. 51, 1987, pp. 363–378
- [Vin58] I. M. Vinogradov “Eine neue Abschätzung der Funktion ” In Izv. Akad. Nauk SSSR, Ser. Mat. 22, 1958, pp. 161–164
- [Woo12] Trevor D Wooley “Vinogradov’s mean value theorem via efficient congruencing” In Annals of Mathematics, 2012, pp. 1575–1627
- [Woo13] Trevor D. Wooley “Vinogradov’s mean value theorem via efficient congruencing, II” In Duke Mathematical Journal 162.4, 2013, pp. 673 –730
- [Woo17] Trevor D. Wooley “Nested efficient congruencing and relatives of Vinogradov’s mean value theorem” In Proceedings of the London Mathematical Society 118, 2017
- [Yan23] Andrew Yang “Explicit bounds on in the critical strip and a zero-free region” arXiv:2301.03165, 2023