Explicit bounds for the Riemann zeta-function on the 1-line
Abstract.
Explicit estimates for the Riemann zeta-function on the -line are derived using various methods, in particular van der Corput lemmas of high order and a theorem of Borel and Carathéodory.
Key words and phrases:
Van der Corput estimate, exponential sums, Riemann zeta function.2020 Mathematics Subject Classification:
Primary: 11M06, 11Y35, 11L07.1. Introduction
Let denote the Riemann zeta-function. An open problem in analytic number theory is to determine the growth rate of on the line as . Assuming the Riemann hypothesis, Littlewood [Lit26, Lit28] showed that the value of is limited to the range
In comparison, the sharpest estimates requiring no assumption furnish a wider range of the form
(1) |
The upper bound in the unconditional estimate (1) was made explicit by Ford [For02]. He computed numerical constants and such that111Trudgian [Tru14] later lowered Ford’s value of from to .
The constants and are often too large for standard applications, including explicit bounds on [Tru14a, HSW22] and explicit zero-density estimates [KLN18]. One instead defaults to an asymptotically worse bound of the form that nevertheless is sharper in a finite range of of interest.
This motivates our Theorem 1. We establish a new explicit bound on that is asymptotically sharper than but still good at small values of .
Theorem 1.
For , we have
Additionally, using a different method, we prove explicit upper bounds of the same form, on both and .
Theorem 2.
For , we have
Furthermore, for , if
then
Theorem 1 makes explicit a result due to Weyl [Wey21] (see also [Tit86, Section 5.16]), and is sharper than the prior explicit bounds due to Patel [Pat22] on for in the range .222We also note that Theorem 1 is sharper for than the bound given in [Tru14].
As far as we know, the bounds in Theorem 2 are the first unconditional explicit bounds of their kind. The first assertion of Theorem 2 supersedes [Car+22, Proposition A.2] when , while the second assertion supersedes [Tru15, Table 2 with ] for .
Under the assumption of the Riemann Hypothesis, sharper explicit estimates for and of the order are known [Sim23], [CHS24, Theorem 5], [Chi23]. Much work has also been done on bounding more general -functions [Lum18], [PS22].
Lastly, it is reasonable to try to adapt the remarkable approaches in [Sou09, p. 984] and [LLS15, Lemma 2.5] to our context. In these articles, very precise bounds, conditional on the Riemann hypothesis, are obtained. One can arrange for these approaches to yield unconditional bounds, and the outcome is similarly constrained by the width of available explicit zeros-free regions. See [Li10], for example, where the method from [Sou09] is used to establish bounds on -functions for .
2. Discussion
One can derive an explicit bound on of the form using a version of the Euler–Maclaurin summation. For example, using the version in [Sim20, Corollary 2], applied with and , together with the elementary harmonic sum bound333Here, is the Euler constant. (see Lemma 8 below),
(2) |
yields after a small numerical computation, for .
Patel [Pat22] replaced the use of the Euler–Maclaurin summation by an explicit version of the Riemann–Siegel formula on the -line. Consequently, the size of the main term in the Euler–Maclaurin bound was cut by half. Patel thus obtained for , a substantial improvement at the expense of a larger constant term.
In general, we expect that an explicit van der Corput lemma of order , where , will produce a bound of the form
(3) |
One proceeds by approximating by an Euler–Maclaurin sum , or better yet by a Riemann–Siegel sum . In either case, roughly speaking, the term arises from bounding the subsum with via the triangle inequality and the harmonic sum estimate (2), while the constant predominantly arises from bounding the subsum with via van der Corput lemmas of order .
However, the obtained using this approach appear to grow fast with . For example, in [Pat21], is about -times larger than , which already suggests exponential growth. Despite this, there is still gain in letting grow with , provided the growth is slow enough. We let grow like , which is permitted by a recent optimized van der Corput test, uniform in , derived in [Yan24]. Subsequently, we obtain a bound on of the form with an explicit number .
While we could let increase faster than , leading to a smaller contribution of the term, the thus obtained would likely take over as the main term, asymptotically surpassing in size. This limits the order of the van der Corput lemmas we employ.
In proving Theorem 1, we encounter an unexpected “gap interval”, not covered by any of the other explicit estimates, which requires a special treatment. Namely, the interval . To prove the inequality in Theorem 1 for all , we use the following proposition.
Proposition 3.
We have
To clarify the need for Proposition 3, consider that if one inputs the bounds in [Pat22], as below, one finds
and the maximum of this ratio occurs around . On the other hand, the savings we otherwise establish here may materialize only when is well past . Therefore, without the improvement enabled by Proposition 3, we cannot do any better than the bound over . Proposition 3, which is based on a hybrid of van der Corput lemmas of orders , and , allows us to circumvent this barrier.
In retrospect, a gap interval like should be expected. This is because the estimates in [Pat21] rely, essentially, on van der Corput lemmas of orders and .444The bound obtained via the Riemann–Siegel formula in [Pat21] should correspond to a van der Corput lemma. In particular, bounds that would follow from using van der Corput lemmas of orders and are not utilized, which presumably causes the gap.
3. Proof of Theorem 1
Let be a positive number to be chosen later, subject to . We suppose throughout this section. (We will use another method for .)
Theorem 9 supplies a Riemann–Siegel approximation of in terms of two sums, plus a remainder term . We use the triangle inequality to bound the entire second sum in this theorem. This gives
(4) |
Since and are well-understood, the bulk of the work is in bounding the main sum over .
We introduce two sequences, and . The former sequence will be used to partition the main sum in (4) in a dyadic manner. The latter sequence will be used to easily bound the partition points from above and below solely in terms of (or, what is essentially the same, in terms of ).
To this end, let and be real positive numbers, also to be chosen later, subject to and . We use immediately in defining
Furthermore we will assume that and are chosen so that
(5) |
Since by supposition, . The integer will correspond to the highest order van der Corput lemma we use in bounding the tail of the main sum in (4).
Furthermore, let
In the definition of , is determined by . Also, decreases monotonically to zero with , and is bounded from above by .
We now define the previously mentioned sequences and . Set
So, and since , decreases monotonically with . Also set
(6) |
so that decreases monotonically with , starting at . Note that if is large, then , and so for large . Lastly, define
The integer corresponds to the number of points we will use to partition the main sum. From the definition, we have
(7) |
Thus, our choice of ensures that
(8) |
Additionally, we have
(9) |
The assumption (5) is required for the subsequent analysis to be valid. This condition implies, among other things, that and, hence, . This in turn ensures that some quantities appearing later (e.g. ) are nonnegative. All contingent conditions we impose (including on , , and ) will be verified at the end, after choosing the values of our free parameters.
By the chains of inequalities (8) and (9), for all , . Therefore, since the sequence partitions the interval , for each integer , there is a unique integer such that
(10) |
Note, though, that multiple ’s could correspond to the same .
With this in mind, let us denote
(11) |
and observe that
(12) |
We calculate
(13) |
Here, we used the elementary inequality , valid for any real number together with the definition .
We bound using Lemma 13 with and . For each integer such that , and with , we have, for any choice of the free parameter ,
(14) |
To bound the RHS, observe that
(15) |
In addition, . Moreover, it follows by definition that
(16) |
Since decreases continuously with , and , we have where
(17) |
It is straightforward to verify that is increasing in for , so . Inserting (15) and (17) into (14),
(18) |
for , where
(19) |
This is permissible since the exponents of in the first and second terms of (14) respectively satisfy
(20) |
It is worth clarifying that the right-hand side of (18) depends on via the requirement (10).
Next, we express the bound on in (18) solely in terms of . Starting with the first term, we have
(21) |
where after a quick calculation,
(22) |
We estimate the second term in (18) similarly, yielding
(23) |
The exponent of in the estimate (21) may be bounded for via . Additionally, recalling the second inequality of (20), we may bound the contribution of the -factor in the estimate (23) simply by . This motivates us to define
(24) |
Note that additionally depends on , , and . Assembling (18), (21), (23) and (24), we obtain that if and is determined via the requirement (10), then
(25) |
We now divide the argument into two cases: and . In the former case, let denote the number of sums corresponding to each integer . In other words, is the number of such that . We observe that if and only if , where
(26) |
Therefore, by the bound on in (25),
(27) |
By counting the maximum number of integers in a continuous interval, for ,
(28) |
where
(29) |
Thus, in view of (27) we obtain
(30) |
where by (28) and on explicitly computing the values of for ,
(31) |
Here, we used the supposition which ensures that is decreasing for , with .
Next, for , we use a tail argument that removes the dependence of on in the bound (25). Define
(32) |
We will bound explicitly later (Section 4), after determining , , and . Importantly, under our assumptions, it will transpire that is finite.
In the meantime, we observe that is continuously decreasing in (towards zero), as can be seen from computing the derivative of with respect to . So, by simply counting the number of terms, and noting that as ranges over , , we obtain from (25),
(33) |
We want to further simplify the bound on the right-side of (33). Via (16) and a routine calculation,
(34) |
Hence, utilizing the inequality as well as , valid for any real number , yields that if , then
(35) |
where (identifying with below)
(36) |
In indicating the dependencies of , we used that is determined by and , and further depends on . Now, from the definition of , and using
we obtain555By the lower bound on in (5), we have for . So, is well-defined for .
(37) |
Since666This step is where a slow enough growth on is utilized, to ensure that as , so that the constant defined in (36) is positive. by supposition , the quantity defining tends to as . As this quantity is continuous and positive for , is finite and positive for all admissible . From (33) and (35) we conclude that
(38) |
Lastly, to bound the remaining terms in (12), in view of (7), we have the rough estimate
Since and , we have, for ,
(39) |
say.
In summary, combining (12), (30), (38) and (39), the tail sum satisfies the inequality
(40) |
So, for such , the tail sum is bounded by a constant independent of .
It remains to bound the initial sum over . We use the triangle inequality followed by the harmonic sum bound (2), which yields
(41) |
We use the inequality (8) to bound from above and below, so that the right-side of (41) satisfies
(42) |
Since by (16), the first term in (42) is bounded by
(43) |
where the last inequality is due to the fact that is decreasing for if .
In addition, we use (17) to bound the last term. Putting this together with (41), (42) and (43), we obtain
(44) |
where
Lastly, the bound on in (4) carries an extra term . Using the definitions in Theorem 9, this term is easily bounded by . Therefore, combining this with (40) and (44), we see that
for , where
We choose the following values for our free parameters, which are suggested by numerical experimentation.
(45) |
With these choices, our preconditions on and hold, and the conditions and also hold (in fact, ). Additionally, in Section 4, we show that
Therefore, we can explicitly compute
In summary, the assertion of Theorem 1 holds when .
4. Bounding and
We derive an upper bound on , where the parameter values are given in (45). For the convenience of the reader, we recall the definition , where for ,
With our choice of free parameter values in (45), we numerically compute the following list of .
The values of and required for the above computation are obtained using the formulas in Lemma 13. In turn, evaluating and relies on evaluating and . When , and are computed directly using their definitions in Lemma 11. When , the recursive formulas in Lemma 12 are used to compute them. Let us examine these recursive formulas.
First, the following quantity, which appears in the formula for in (92), satisfies
(46) |
Here, and the maximum occurs when . Another quanitity appearing in the formula for is , which is clearly seen to decrease with . Noting that , hence for any , we deduce that
So, we are led to consider the discrete map
By a routine calculation, this map has a single fixed point at
(47) |
which is a stable point. Since for and, by a direct numerical computation, , it follows that
(48) |
The analysis of for proceeds similarly. The recursive formula for contains the following quantity satisfying
(49) |
So, for . Consider the discrete map . This map has a single fixed point, namely,
which is also stable. Since for and, by a direct numerical computation, , it follows that
(50) |
Furthermore, we numerically verify that for ,
Combining this with (48) and (50), and recalling the definitions in Lemma 13 and (19), yields the inequalities and for . Consequently, for .
Furthermore, in view of the values of listed at the beginning, we also see that for . We therefore conclude .
For the computation related to , we shall use a derivative calculation to show that the function
(51) |
is increasing in . So, on identifying with in the definition of in (36), the value of is .
To this end, we first calculate
where
Note that is identified with , so we need only consider . It is easy to see that
where the supremum occurs at . Hence, the derivative of is at least for . Importantly, is positive. In comparison, the derivative of the remaining terms in the formula for is
since and . Since
(52) |
for , the derivative of is positive for , proving our claim.
5. Proof of Proposition 3
The calculation is similar to that in [Hia16] on the -line, so we shall be brief in some details. We divide the main sum in (4) into pieces of length , then apply the van der Corput Lemma 10 to each piece other than an initial segment, which is bounded by the triangle inequality.
Let be a positive number to be chosen later, subject to . Suppose that , unless otherwise stated, and let be as defined in (4). Let , , and let a positive integer to be chosen later, subject to . Thus, we have
(53) |
The harmonic sum bound (2) yields, subject to ,
(54) |
For , let and let be an integer such that . Partial summation gives
(55) |
where
We estimate using Lemma 10 with , and
This gives, for any positive ,
where , , , are defined in Lemma 10. We estimate the (nonnegative) terms under the square-root as follows.
Isolating the first term using the inequality , valid for nonnegative and , and using the inequality to bound in the remaining terms, therefore gives
(56) |
where, since ,
We also note that for all we have and , where
Therefore, for all under consideration, and . It follows that for ,
(57) |
where and .
Put together, we have
(58) |
Moreover, from the argument in [HPY24, (3.1)], and using the bound once again,
(59) |
where
Therefore, observing that is decreasing for , and combining this with the initial sum bound (54), we obtain
We now estimate
which follows using . Furthermore,
This yields
(60) |
where
Choosing , and , we obtain
(61) |
This implies Theorem 1 in the range .
Consider now the second part of Proposition 3. In this critical region, we make use of a few additional tools to sharpen our estimates. In particular, we employ a different method of bounding the second sum appearing in Theorem 9, leading to an estimate of size instead of . Also, we employ the -th derivative tests used in the proof of Theorem 1, however this time we introduce additional scaling parameters to further optimise the switching points between derivative tests of different order.
To this end, let be a real number to be chosen later, subject to . Assume . Recall from (58) and (59) we showed for , and that the tail sum satisfies
(62) |
As for the initial sum over , instead of bounding it using the triangle inequality throughout as in (54), we will refine our estimate by using -th order van der Corput lemmas with and in the subintervals and respectively, where
with the same as in the proof of Theorem 1 and real positive numbers to be chosen later, subject to
(63) |
We remark that will serve a similar purpose as in the proof of Theorem 1, whereas the new parameters will allow us to fine-tune our estimates in this critical region.
Let be a real number to be chosen later, and let for integer . Define
The definition of guarantees . Next, define the subsum
so that, similarly to (12),
By a similar calculation as in (13), for each ,
Let or . Following the argument of Theorem 1, if , then
We write so that . Since , for ,
Let be the number of satisfying . Then
where
Put together, we have
For , the function is decreasing, so . For , we use
Therefore,
(64) |
Next, similarly to (39), we obtain
(65) |
In the remaining range , we use that . Combined with the triangle inequality and (2), this yields
(66) |
Next, consider the second sum in Theorem 9, which was just bounded by in our previous calculations. Let , , , and be as defined in the proof of Proposition 3, and suppose as before that . Using , the estimate (57) gives for ,
We combine this with the inequality
as well as bound the sum over by the number of terms in the sum, which gives
Since , we obtain for ,
(67) |
where
Therefore, after multiplying by , we obtain a sharper bound on the contribution of the second sum in Theorem 9. Namely, the bound , which replaces the original estimate .
6. Proof of Theorem 2
We begin by proving some bounds on . To help make our notation more intuitive, we will allow reuse of variable names from previous sections. The reader should keep in mind these variables mean different things in the confines of this section. Since this section is independent from Section 3, Section 4 and Section 5, this will hopefully cause no confusion.
Lemma 4.
Let and be constants. Suppose . Define
Additionally, suppose that satisfies
(68) |
For each , if , then
(69) |
Proof.
From Ford [For02], if , then
(70) |
As , a consequence of the assumption (68) and considering that reaches its maximum at , we may use (70) for any .
In view of this, since is monotonically decreasing with , if , then (70) gives
(71) |
Or, written differently,
(72) |
where is defined as in (69) and
As reaches its maximum at , attaining the value , we obtain after a simple computation that , with defined as in (69). The desired result therefore follows for on replacing with in (72).
To show that the result holds for as well, we use the Phragmén–Lindelöf Principle on the holomorphic function
To this end, on the -line we have
This inequality is verified numerically for and is a consequence of the bound for , given in [Tru14]. Also, on the -line, plainly,
for all real . By [Tru14a, Lemma 3], we thus obtain777We apply [Tru14a, Lemma 3] with the following parameter values (in the notation in the cited paper): , , , , and . Note that the and in [Tru14a, Lemma 3] are different from the and in the statement of our lemma.
(73) |
Moreover, for we have the easy estimates,
So that, for and , the inequality (73) implies that
where and are defined as in (69), as desired. ∎
Remark.
Lemma 5.
Let and be constants. Suppose . We have
where is Euler’s constant, and
(74) |
Proof.
We use the following uniform bound due to Ramaré [Ram16, Lemma 5.4]:
Substituting therefore gives for ,
as claimed. In the last inequality, we used that reaches its maximum at and is monotonically decreasing after that. ∎
6.1. Bounding
We follow the argument of [Tru15], which gives an explicit version of results in [Tit86, §3]. The method relies on Theorem 14.
We will therefore construct concentric disks, centred just to the right of the line , and extend their radius slightly to the left and into the critical strip. Ultimately, we want to apply Lemmas 15 and 16 with , which will give us the desired results.
The process is as follows. Let and suppose . Let be a real positive constant, which will be chosen later. Denote the center of the concentric disks to be constructed by , where
(75) |
Notice that is decreasing with and is at most . Let denote the radius of our largest concentric disk. In the sequel, will be determined as function of .
In addition, suppose where is another positive constant to be chosen later. This implies, for instance, that . In order to enforce this inequality on and , it is enough to demand that
(76) |
Let be a complex number. Aiming to apply Lemma 15 in the disk , we first seek a valid in that disk. Similarly, for Lemma 16, we seek a valid at the disk center . In applying Lemma 16, we moreover need to ensure that the non-vanishing condition on is fulfilled, which we do by way of a zero-free region of zeta.
Let us first determine a valid , with the aid of Lemma 4. To this end, let be a parameter restricted by the inequality in (68), so that (in the notation of Lemma 4) for . Note that is monotonically decreasing for and monotonically increasing thereafter. So, the maximum of over the interval occurs at one of the end-points. For example, if , then the maximum occurs at . And if , then the maximum occurs at . With this in mind, we calculate
(77) |
where, since ,
Clearly, increases monotonically with (towards ). We therefore set
(78) |
So that, by (77), no matter , and , so long as they are subject to our conditions,
Consequently, Lemma 4 gives that for each and any , and on recalling ,
(79) |
We would like the inequality (79) to hold throughout the disk . This will be certainly the case if this disk lies entirely in the rectangle specified by and . This follows, in turn, on requiring or, equivalently, that
Therefore, subject to the constraint (76), the following choice of works:
(80) |
To ensure the constraint (76) is met for all , it suffices that
(81) |
Here, we used that reaches its maximum of at .
Overall, combining (79) and (80) with the trivial bound , easily seen on considering the Euler product of zeta, we obtain that throughout the disk ,
where
Since decreases to with , decreases to with . Thus, recalling from (75) that and observing is also decreasing in , we obtain that throughout the disk ,
(82) |
where and .
Next, in preparation for applying Lemma 16, we want to ensure that the intersection of the disk and the right half-plane lies entirely in a zero-free region for zeta. By [Yan24, Corollary 1.2], we have the following zero-free region, valid for .
Since is monotonically increasing for and since and , it suffices for our purposes to require that
In other words, it suffices that
Therefore, subject to as demanded by Lemma 16, the following choice of works.
(83) |
where (using monotonicity to deduce the inequality below)
(84) |
In particular, the constraint is fulfilled if
(85) |
Now, to determine , we utilise [Del87], which asserts that if , then
Taking in this inequality leads to
So, we may take in Lemma 16 to be
Lastly, in applying Lemma 16, we need to be able to reach the -line from our position at . So, we set
Clearly, , as required by Lemma 16. The conclusion of the lemma therefore holds throughout the disk . This includes, in particular, the horizontal line segment , with .
Putting it all together, and making the substitution , Lemma 16 hence furnishes the bound
(86) |
valid in the region
We now input the values we determined for , , , and into (86). We also note
as well as
Hence, on observing that is decreasing in , we obtain the following lemma.
Lemma 6.
Let and suppose that . Let , and be any positive constants satisfying the constraints (68), (81) and (85). If
then
where
Here,
The number is defined as in Lemma 4 and depends on . The numbers and are defined as in (82) and depend on and . The number is defined in (78), and depends on , and . The number is defined in (84) and depends on , and .
6.2. Bounding
Moving from a bound on the logarithmic derivative of to one on the reciprocal of is done in the usual way (see [Tit86, Theorem 3.11]), with some improvements using the trigonometric polynomial (see [Car+22, Proposition A.2]).
Lemma 7.
Let and be constants. Suppose that for each ,
Then for any and any real number such that , we have
where
The number is defined as in (74).
Proof.
Suppose . Let be a real positive parameter such that . Let . It is easy to see that
So, using our suppositions, we have
(87) |
On the other hand, for , and using the classical nonnegativity argument involving the trigonometric polynomial [Tit86, Section 3.3], gives that
Combining this with the inequality and Lemma 5, we therefore obtain, on taking ,
(88) |
We now observe
(89) |
So that, combining (89) and (88), and noting by (87) we have
the lemma follows. ∎
6.3. Numeric calculations
We define the following upper bounds:
The numeric calculations for and proceed as follows. First fix and a positive . For the purposes of obtaining an upper bound , we then optimise in Lemma 6 over and , all the while ensuring , , and are subject to our constraints (68), (81) and (85)
Thereafter, with these values of and , the relationship allows us to obtain by optimising in Lemma 7 over .
Our choices of
give us
and the upper bounds
In summary, we have for ,
(90) |
Furthermore, in the region
we have for ,
(91) |
Finally, to cover the range from in (90), we refer to the proof of Proposition in [Car+22], where via interval arithmetic, they computed that for ,
Comparing this with the estimate in (90), we see after a simple verification that the bound
holds for all .
This concludes our proof of Theorem 2. These bounds can certainly be improved. Some possible avenues of attack would be to treat the sum over zeros in the proof of Lemma 16 more carefully, or to use a larger zero-free region.
Remark.
It is possible with our methods to directly get a bound like (91) for , especially if one were not concerned with the size of . For instance, choosing , , , and , would return for that if
then
While this would cause a modest jump in , the size of would increase significantly to . This is because increases exponentially with , as evident from Lemma 7. The dominant factor there is , and any improvement in that quantity would help greatly.
7. Background results
7.1. Background results related to Sections 3-5
Lemma 8.
[Fra21, Lemma 2.1] Let denote Euler’s constant. For ,
Theorem 9.
Next we review some explicit estimates of exponential sums from the literature. The following lemma is a specialised third-derivative test for a phase function commonly encountered when estimating the Riemann zeta-function. It is due to [HPY24, Lemma 2.6] which builds on [Hia16].
Lemma 10 (Explicit third derivative test).
[HPY24, Lemma 2.6] Let and be positive integers, and let and be positive numbers. Suppose . Let
so that for . Then, for each positive integer and any ,
where
Remark.
If , say, then will be greater than , and so . Hence, for small , it is better to use the triangle inequality than Lemma 10.
Remark.
We mention that there is a typo in the statement of [HPY24, Lemma 2.6] where in the statement of the lemma should be . This lemma is only applied with in [HPY24] in any case.888In passing, let us also point out that on [HPY24, 211], the minimum in the definition of should be against rather than . This does not affect Equation 5.4 on p. 211 since, as stated in [HPY24, Lemma 2.5], the bound on holds up to , not only . None of the results in [HPY24] are affected.
The next two lemmas respectively represent another third derivative test and a th derivative test for , due to [Yan24]. The reason we state this additional third derivative test is because and comprise the boundary cases for the recursive formulas in Lemma 12.
Lemma 11 (Another explicit third derivative test).
[Yan24, Lemma 2.4] Let and be integers such that . Suppose has three continuous derivatives and is monotonic over the interval . Suppose further that there are numbers and such that for all . Then for any ,
where
and and are defined by
Lemma 12 (General explicit -th derivative test).
[Yan24, Lemma 2.5] Let , and be integers such that and . Suppose is equipped with continuous derivatives and is monotonic over the interval . Suppose further that there are numbers and such that for all . Then for any ,
where , and are defined recursively via the formulas
(92) |
(93) |
with and as in Lemma 11, and defined by
(94) |
Lemma 13.
Let , and be integers such that , and . Let . Let be a number such that . Then for any ,
where
and and defined in accordance with Lemma 12.
Proof.
This is a special case of Lemma 12 and Lemma 11 with the phase function
Since by assumption, for , where
Suppose that . Then, we may apply Lemma 12 with these values of and , and with . This furnishes a bound that increases with . Noting that , we thus obtain
The desired result hence follows from partial summation in this case. If , then the result follows from Lemma 11 using a similar calculation. ∎
7.2. Background results related to Section 6
Here we give some preliminary lemmas for our proofs of estimates on and . We reiterate that Section 6 reuses variable names from previous sections. These variables mean different things in the confines of this section as this section is independent from Sections 3, 4, and 5.
The first two lemmas, which form the basis of our method, are built upon a theorem of Borel and Carathéodory. This theorem enables us to deduce an upper bound for the modulus of a function and its derivatives on a circle, from bounds for its real part on a larger concentric circle.
Theorem 14 (Borel–Carathéodory).
Let be a complex number. Let be a positive number, possibly depending on . Suppose that the function is analytic in a region containing the disk . Let denote the maximum of on the boundary . Then, for any and any such that ,
If in addition , then for any and any such that ,
Proof.
Lemma 15.
Let be a complex number. Let and be positive numbers (possibly depending on ) such that . Suppose that the function is analytic in a region containing the disk . Suppose further there is a number independent of such that,
Then, for any in the disk we have
where runs through the zeros of in the disk , counted with multiplicity.
Proof.
We follow the proof in [Tit86, Section 3]. Let be the function
where in the (finite) product runs through the zeros of , counted with multiplicity, that satisfy . Since by construction the poles and zeros cancel, is analytic in , and does not vanish in the disk . Therefore, the function
where the logarithm branch is determined by , is analytic in some region containing the disk .
Now, on the circle , we have . So, on this circle,
The inequality also holds in the interior of the circle, by the maximum modulus principle. Hence, throughout the disk .
Lemma 16.
Let and be complex numbers with real parts and , respectively. Let , , , and be positive numbers, possibly depending on , such that and . Suppose that the function satisfies the conditions of Lemma 15 with , and , and that
Suppose, in addition, that for any in both the disk and the right half-plane . Then, for any in the disk ,
Proof.
Using the inequality , valid for any complex number , together with Lemma 15 in , we have throughout this disk,
(96) |
where runs through the zeros of (with multiplicity) satisfying . By the assumption on the nonvanishing of , each in this sum satisfies . On the other hand, all in the disk clearly satisfy . Hence,
throughout the disk . We may thus drop the sum in (96) and still obtain a valid upper bound, call it .
Now, is analytic in the disk since by assumption does not vanish there. We can therefore apply Theorem 14 to using the two concentric circles and , and the upper bound on the circle , which yields the result. ∎
Acknowledgements
We would like to thank Fatima Majeed for pointing out errors in an early version of this manuscript and giving us some helpful suggestions.
References
- [Car+22] Emanuel Carneiro, Andrés Chirre, Harald Andrés Helfgott and Julián Mejía-Cordero “Optimality for the two-parameter quadratic sieve” In Acta Arithmetica 203 Instytut Matematyczny Polskiej Akademii Nauk, 2022, pp. 195–226
- [Chi23] Andrés Chirre “Bounding zeta on the -line under the partial Riemann hypothesis” In Bulletin of the Australian Mathematical Society Cambridge University Press, 2023, pp. 1–8
- [CHS24] Andrés Chirre, Markus Valås Hagen and Aleksander Simonič “Conditional estimates for the logarithmic derivative of Dirichlet L-functions” In Indagationes Mathematicae 35.1, 2024, pp. 14–27
- [Del87] Hubert Delange “Une remarque sur la dérivée logarithmique de la fonction zêta de Riemann” In Colloquium Mathematicum 53.2, 1987, pp. 333–335 Polska Akademia Nauk. Instytut Matematyczny PAN
- [For02] Kevin Ford “Vinogradov’s Integral and Bounds for the Riemann Zeta Function” In Proceedings of the London Mathematical Society 85.3, 2002, pp. 565–633
- [Fra21] F. J. Francis “An investigation into explicit versions of Burgess’ bound” In J. Number Theory 228 Elsevier, 2021, pp. 87–107
- [HSW22] E. Hasanalizade, Q. Shen and P-J. Wong “Counting zeros of the Riemann zeta function” In J. Number Theory 235 Elsevier, 2022, pp. 219–241
- [HPY24] G. A. Hiary, D. Patel and A. Yang “An improved explicit estimate for ” In J. Number Theory 256 Elsevier, 2024, pp. 195–217
- [Hia16] Ghaith A. Hiary “An explicit van der Corput estimate for ” In Indagationes Mathematicae 27.2, 2016, pp. 524–533
- [KLN18] Habiba Kadiri, Allysa Lumley and Nathan Ng “Explicit zero density for the Riemann zeta function” In Journal of Mathematical Analysis and Applications 465.1, 2018, pp. 22–46
- [LLS15] Youness Lamzouri, Xiannan Li and Kannan Soundararajan “Conditional bounds for the least quadratic non-residue and related problems” In Mathematics of Computation 84.295, 2015, pp. 2391–2412 DOI: 10.1090/S0025-5718-2015-02925-1
- [Li10] Xiannan Li “Upper bounds on -functions at the edge of the critical strip” In Int. Math. Res. Not. IMRN, 2010, pp. 727–755 DOI: 10.1093/imrn/rnp148
- [Lit26] J. E. Littlewood “On the Riemann Zeta-Function” In Proceedings of the London Mathematical Society s2-24.1, 1926, pp. 175–201 DOI: https://doi.org/10.1112/plms/s2-24.1.175
- [Lit28] J. E. Littlewood “Mathematical Notes (5): On the Function ” In Proceedings of the London Mathematical Society s2-27.1, 1928, pp. 349–357
- [Lum18] Allysa Lumley “Explicit bounds for -functions on the edge of the critical strip” In Journal of Number Theory 188 Elsevier, 2018, pp. 186–209
- [MV07] H. L. Montgomery and R. C. Vaughan “Multiplicative number theory I: Classical theory” Cambridge university press, 2007
- [PS22] Neea Palojärvi and Aleksander Simonič “Conditional estimates for -functions in the Selberg class” In arXiv preprint arXiv:2211.01121, 2022
- [Pat21] Dhir Patel “Explicit sub-Weyl bound for the Riemann Zeta function”, 2021
- [Pat22] Dhir Patel “An explicit upper bound for ” In Indagationes Mathematicae 33.5, 2022, pp. 1012–1032
- [Ram16] O. Ramaré “An explicit density estimate for Dirichlet -series” In Mathematics of Computation 85.297, 2016, pp. 325–356
- [Sim20] Aleksander Simonič “Explicit zero density estimate for the Riemann zeta-function near the critical line” In Journal of Mathematical Analysis and Applications 491.1, 124303, 41pp., 2020
- [Sim23] Aleksander Simonič “Estimates for L-functions in the Critical Strip Under GRH with Effective Applications” In Mediterranean Journal of Mathematics 20.2, 2023, pp. 87
- [Sou09] Kannan Soundararajan “Moments of the Riemann zeta function” In Ann. of Math. (2) 170.2, 2009, pp. 981–993 DOI: 10.4007/annals.2009.170.981
- [Tit39] E. C. Titchmarsh “The Theory of Functions” Oxford university press, 1939
- [Tit86] E. C. Titchmarsh “The Theory of the Riemann Zeta-function” Oxford: Oxford Science Publications, 1986
- [Tru14] T. Trudgian “A new upper bound for ” In Bulletin of the Australian Mathematical Society 89.2, 2014
- [Tru15] Tim Trudgian “Explicit bounds on the logarithmic derivative and the reciprocal of the Riemann zeta-function” In Functiones et Approximatio Commentarii Mathematici 52.2 Adam Mickiewicz University, Faculty of MathematicsComputer Science, 2015, pp. 253–261
- [Tru14a] Timothy Trudgian “An improved upper bound for the argument of the Riemann zeta-function on the critical line II” In Journal of Number Theory 134, 2014, pp. 280–292
- [Wey21] Hermann Weyl “Zur Abschätzung von ” In Mathematische Zeitschrift 10.1-2, 1921, pp. 88–101 DOI: 10.1007/BF02102307
- [Yan24] A. Yang “Explicit bounds on in the critical strip and a zero-free region” In Journal of Mathematical Analysis and Applications 534.2 Elsevier, 2024, pp. 128124