Abstract
It is proved that if T is sufficiently large, then uniformly for all positive integers , we have
where γ is the Euler constant. We also establish lower bounds for maximum of when and are fixed.
1. INTRODUCTION
This paper establishes the following new results for extreme values of derivatives of the Riemann zeta function (in this paper, we use the short‐hand notations, , and ).
Theorem 1
If T is sufficiently large, then uniformly for all positive integers , we have
Remark 1
In our Theorem 1, ℓ does not have to be fixed. In particular, if , then for sufficiently large T, we have
This value is even larger than the conditional upper bound of extreme value of the Riemann zeta function on the ‐line in the same interval . Recall that Littlewood [20] proved that the Riemann hypothesis (RH) implies the existence of a constant C such that for large T we have . Chandee and Soundararajan [10] proved that on RH, one can take any constant .
Theorem 2
Let and be fixed.
- (A)
Let c be a positive number less than . If T is sufficiently large, then
- (B)
Let be given and κ be a positive number less than . Then for sufficiently large T, we havewhere is an absolute positive constant.
Remark 2
By Soundararajan's original resonance method [22], we can also establish lower bounds for the maximum of derivatives of the zeta function on the ‐line on the shorter interval . In this case we obtain , losing a factor compared to the above result on the longer interval .
The research for extreme values of the Riemann zeta function has a long history. In 1910, Bohr and Landau first established the result (see [23, Theorem 8.5]). In 1924, Littlewood (see [23, Theorem 8.9(A)]) was able to find an explicit constant in the Ω‐result of Bohr and Landau, by proving that . Littlewood's result was improved by Levinson [18] in 1972, and by Granville–Soundararajan [14] in 2005. The currently best‐known lower bound is established by Aistleitner–Mahatab–Munsch [3] in 2017, who proved that , for some constant C.
On the other hand, when assuming the RH, Littlewood proved that , for sufficiently large t (see [23, Theorem 14.9]). Furthermore, Littlewood conjectured that . In [14], Granville–Soundararajan made the stronger conjecture: , for some constant C 1 which can be effectively computed.
Compared to the research on extreme values of the Riemann zeta function, much less is known about the extreme values of its derivatives. In [17], Kalmynin obtained Ω‐results for the Riemann zeta function and its derivatives in some regions inside the critical strip near the line = 1. He also mentioned that his methods do not provide any non‐trivial results about the domains of the form with . Note that Kalmynin did not obtain Ω‐results for when and are given.
It is still uncertain whether the methods of [3, 14, 18, 23] are able to establish the result in our Theorem 1, since those methods basically rely on the fact that the k‐divisors function is multiplicative and/or the fact that the Riemann zeta function has an Euler product: , . Note that the function is not multiplicative and the derivative does not have an Euler product.
We also emphasize that the key points in Theorem 1 are the range and the constant in front of . In fact, one can use the method of Bohr–Landau to prove a much weaker result, that is, when is fixed. See Section 7 for such a short proof.
We will use Soundararajan's original resonance method [22] to prove Theorem 1. The new ingredient for the proof is the following Proposition 1.
Proposition 1
If T is sufficiently large, then uniformly for all positive numbers ℓ, we have
where the supremum is taken over all functions satisfying that the denominator is not equal to zero, when the parameter T is given.
The following Proposition 2 will not be used to prove our theorems. However, it is closely related to Proposition 1 and can be viewed as a “log‐type” greatest common divisor (GCD) sum, so we list it here for independent interest.
Proposition 2
Let and let be a positive number less than . For sufficiently large N, we have
where the supremum is taken over all subsets with size N.
Remark 3
Actually we can also use Proposition 2 and Hilberdink's version of the resonance method [15] to prove a similar result to the one in Theorem 1. But the constant in front of will be much worse.
Soundararajan introduced his resonance method in [22] and proved that
which improved earlier results of Montgomery and Balasubramanian–Ramachandra. Montgomery [21] proved it under RH and with the constant 1/20 instead of in Soundararajan's result. Balasubramanian–Ramachandra [4] proved the result unconditionally but also with a smaller constant compared to Soundararajan's result.
By constructing large GCD sums, Aistleitner [1] used a modified version of Soundararajan's resonance method to establish lower bounds for maximum of when is fixed. His results improved early results of Voronin [24] and Hilberdink [15] via resonance methods. He proved that
for large T, and one can take . The same result had been proved by Montgomery in [21] with a smaller value for . In [9], Bondarenko and Seip improved the value in Aistleitner's result.
By constructing large GCD sums, using a convolution formula for ζ in the resonance method, Bondarenko and Seip [7, 8] proved the following surprising result:
After optimizing the GCD sums, de la Bretèche and Tenenbaum [11] improved the factor from to in the above result.
Following the work of Bondarenko–Seip and de la Bretèche–Tenenbaum, we use their modified versions of resonance methods to prove Theorem 2. The new ingredient is our convolution formula for . Throughout the paper, define the function as follows:
(1) |
Throughout the paper, also define the sequence as , and for . Then we have the following identity and the Dirichlet series converge absolutely:
(2) |
The reason why we add the part is that we want to make for all . Since when , the factor has very small influence on the log‐type GCD sums compared to the case , we will simply use the fact that and then come to the situation of optimizing GCD sums.
Let be given and let be a finite set. The GCDs sums of are defined as follows:
where denotes the GCD of m and n and denotes the least common multiple of m and n.
The case was studied by Gál [13], who proved that
(3) |
The asymptotically sharp constant in (3) is . This fact was proved by Lewko and Radziwiłł in [19].
Bondarenko and Seip [6, 7] proved the following result for GCD sums when
Later, based on constructions of [6, 7], de la Bretèche and Tenenbaum [11] optimized the result of Bondarenko–Seip and obtained the following:
(4) |
Aistleitner, Berkes, and Seip [2] proved the following essentially optimal result for GCD sums when , where and are positive constants only depending on σ:
(5) |
Moreover, in [2, p. 1526], they also gave an example (following ideas of [13]) for the lower bound when . Let and let be the set of all square‐free integers composed of the first r primes. Then
(6) |
for some positive constant . For simplicity, in our proof we will use this construction. For more constructions, see Bondarenko–Seip [9, pp. 131–136] and Z. Dong‐B.Wei [12, Theorem 1.2].
2. LEMMAS FOR THE RIEMANN ZETA FUNCTION
Lemma 1
Let be fixed. If T is sufficiently large, then uniformly for , , and all positive integers ℓ, we have
(7) where the implied constant in big only depends on σ0.
It follows from Hardy–Littlewood's classical approximation formula (see [23, Theorem 4.11]) for and Cauchy's integral formula for derivatives.
Lemma 2
Let and be fixed. Then uniformly for all and ,
(8) where the implied constant depends on ℓ and ε only.
It follows from classical convex estimates for and Cauchy's integral formula.
In the following, we will derive a “double version” convolution formula, similar to Lemma of 5.3 of de la Bretèche and Tenenbaum [11]. The proof is same as the proof of “single version” convolution formulas in Lemma 1 of Bondarenko and Seip [8].
Define the Fourier transform of K as
Lemma 3
Let and be fixed. Write . Assume that is a holomorphic function in the strip , satisfying the growth condition
(9) If , then
(10) where
(11) and
(12)
Define . is a meromorphic function in the vertical strip , with two poles, namely, at and . Let Y be large and consider straight line integrals for . Set .
Note that , where is an entire function. The residue theorem gives that
(13) By (2) and (9) and applying Cauchy's theorem term by term,
Clearly,
By the trivial estimate , estimates for (Lemma 2) and (9), we obtain
Take , then , as . Similarly for J 3.
The following results are due to Hadamard, Landau, and Schnee (also see [16]).
Lemma 4
(Hadamard, Landau, and Schnee) Let and be fixed. Suppose , then
Remark 4
In this paper, if it is not stated, the limit notations “∼” and “o()” are under the condition when the corresponding variable tends to infinity. Namely, as , , or .
In particular, when and are fixed, one has
(14) |
For , Ingham [16, p. 294, Theorem A″] has proved the following result on second moments of .
Lemma 5
(Ingham) Let be fixed. Then
3. PROOF OF PROPOSITION 1
We will use the construction of Bondarenko and Seip in [9].
Let . Given a positive number y and a positive integer b, define
We will choose a number x and an integer b later to make . Let be the set of divisors of and be the set of divisors of . Let be the complement of in . Note that both and are divisor‐closed which means and . Define the function to be the characteristic function of , then
As showed in [9],
Also in [9, p. 129, lines 3–9], it is proved that
(15) Next, we split the sum into the following two parts:
We will prove the following identity:
(16) To see this, let m be the largest integer such that and let w be the largest integer such that ( denotes the nth prime). Then we have
Note that , then we immediately get (16). Now (15) together with (16) give that
(17) where we omit the term inside the second big term since . Thus we obtain
By the definition of , if , then . So we have
Now we set and . By the prime number theorem, when T is sufficiently large. Take the choices of and into the above inequality, then we are done.
4. PROOF OF THEOREM 1
Set and let . Define the moments as follows:
As in [22], denotes a smooth function, compactly supported in [1, 2], with for all y, and for . Partial integration gives that for any positive integer ν.
Also in [22], Soundararajan proved that
(18) Since Φ is compactly supported in [1, 2], we deduce that
Since , for the off‐diagonal terms we have , by the rapid decay of (see [22, p. 471]). Thus the contribution of the off‐diagonal terms to the above summands can be bounded by
Again, by , we obtain
(19) By Lemma 1, we have the following approximation formula and the implied constant in the big term is absolute:
In the integral of , the big term above contributes at most
Combining this with (19), we have
Finally, the above formula together with (18) gives that
Now let . By Stirling's formula, if T is sufficiently large, then for all positive integers , we have . Other big terms can be easily bounded. Together with Proposition 1, we finish the proof of Theorem 1.
5. PROOF OF THEOREM 2
5.1. Constructing the resonator
Given a set of positive integers and a parameter T, we will construct a resonator , following ideas from [1], [7], and [11]. Define
Let be the set of integers j such that and let be the minimum of for . We then set
and
for every in . Then the resonator is defined as follows:
(20) |
By Cauchy's inequality, one has the following trivial estimates [11]:
As in [7] , set . Its Fourier transform satisfies .
Replacing T by in [8, Lemma 5], gives that
(21) |
5.2. The proof
Let . Choose and set . Fix such that .
As in [8], choose
which has Fourier transform
(22) Define
Following [8] and [11], we will show that the integral on and gives the main term for . We will frequently use the following trivial estimates (Lemma 2):
(23) A simple computation gives
Note that
Thus
where the last step follows from Lemmas 4 and 5 and the Cauchy–Schwarz inequality. Trivially, by and ,
The fast decay of Φ and (23) give that
Using (21) and (23), one can compute
Combining the above estimates, one gets
Note that and give . Again, by (21)
(24) Next, let
(25) and set
By the convolution formula (10), one obtains
We will bound as follows:
(26) By Cauchy's integral for derivatives and the explicit expression for K, we have the following estimates for all :
(27) where the implied constants depend on ε and ℓ only.
And trivially, for all , one has
(28) where the implied constants depend only on ℓ.
Note that there are finitely many non‐negative integer pairs satisfying , so
(29) Proceed similarly for , so we get (26).
Next, in order to relate to the GCD sums, we would like to use Fourier transform on the whole real line. So set
By (22), if . Clearly, . So one can get
We obtain . Thus we have
(30) We compute the integral by expanding the product of the resonator and the infinite series of , and then integrate term by term, as in [7, p. 1699]. Using the fact for every k and if , one gets
Next, proceed as in [11, p. 127–128] (following ideas from [7]),
(31) Combining (30) with (31), we have
(32) Next, we will consider the two cases and separately.
Case 1: .
In this case, let be the set in (4) with . Recall that , so
(33) Also, in [11, p. 128], de la Bretèche and Tenenbaum showed that for this set ,
(34) So the second term on the right‐hand side of (32) is o(1). And clearly, the big terms in (32) can be ignored. Thus
(35) Case 2: .
In this case, let be the set in (6) with . Again, , so
(36) Similarly as (31), we have
And by (4), we can get the following estimates:
Hence,
(37) Make κ slightly larger in the beginning then one can get (B).
6. PROOF OF PROPOSITION 2
The idea of the proof is basically the same as in the proof of Proposition 1. The new ingredient is Gál's identity. In this section, in order to avoid confusion about notations, we use the notation for the ordered pair of m and n.
Let , where denotes the nth prime. Define to be the set of divisors of , then . By Gál's identity [13],
Let , then by the prime number theorem. Let b be the integer satisfying that
then , as . Choose a set such that and .
Following Lewko–Radziwiłł in [19], we use Gál's identity for the GCD sum and then split the product into two parts:
(38) By Mertens' theorem, the first product is asymptotically equal to as . The second product converges as to
Next, let and define the sets as follows:
, where m and n have prime factorizations as .
, where m and n have prime factorizations as .
Then define to be the union of the above two sets and to be the complement of in :
Now we split the GCD sum into two parts:
(39) By symmetry, we have
(40) By the definition of and Gál's identity, we have
Again, we have , the first product converges to , and the third product is asymptotically equal to as .
For the second product, it can be bounded as
And by Mertens' theorem and the prime number theorem, we have
As a result, we obtain that
Hence by (38), (39), and (40), we get
By the construction of , if , then
Thus
By our choice of , we are done.
7. A SHORT PROOF FOR A WEAKER RESULT
One can use the method of Bohr–Landau (see [23, Theorem 8.5]) to prove the weaker result that , when is fixed.
Write . When ,
For given positive integers N and q, by Dirichlet's theorem, there exists , such that for all integers . Hence
Take to get
(41) One can compute that
(42) and for large N that
(43) Now fix a positive constant A (only depending on ℓ) such that and let . Combining with (41) gives that
(44) Next, define
Suppose that . So . Clearly, . Then we get a contradiction with (44) by the Phragmén–Lindelöf principle (for instance, see[23, p. 189]).
8. DISCUSSIONS, OPEN PROBLEMS, AND CONJECTURES
Let and , define the following normalized log‐type GCD sums as:
Problem 1
Given σ and ℓ, optimize .
Remark 5
We are particularly interested in the case . Given ℓ, what is the optimal constant such that ? (See [25] for both unconditional and conditional upper bounds). When , is it true that for some positive constant ? (These bounds are inspired by the work of Bondarenko–Hilberdink–Seip in [5], where the authors studied GCD sums for ). It is not difficult to obtain the upper bounds that for some positive constant , by [5, Theorem 1] and arguments in the proof of Proposition 4 of [25].
We are also interested in extreme values of in the left half strip. It is unlike the situation of the zeta function, where the values on the left half strip can be easily determined by the right half strip via the functional equation. Thus it is worth to study when , even for this reason.
Problem 2
Study extreme values of , when and are given.
We can use Theorem A of Ingham [16] to prove the following claim, from which we obtain the lower bounds (45) on maximum of . But we expect something slightly better.
Claim 1
Let and be fixed. Then
In Theorem A of Ingham [16], let and , then
where
and
Immediately, we obtain the following corollary.
Corollary 1
Let , and be fixed. Then for large T,
(45)
Note that the lower bound in Theorem 1 increases when ℓ increases. So it is natural to have the following conjecture.
Conjecture 1
If T is sufficiently large, then uniformly for all positive integers ⩽ , such that , we have
When ℓ is fixed, we have the following conjecture, inspired by the conjecture of Granville–Soundararajan .
Conjecture 2
Let be given. Then there exists a polynomial of total degree such that
In particular, there exists a positive constant such that
Remark 6
Does exist? In particular, do we have ?
Remark 7
When assuming the RH, one can get for sufficiently large (see [25]).
When and are given, we think that the maximum of derivatives of zeta function and maximum of zeta function only differs by multiplying some small factors. More precisely, we have the conjecture.
Conjecture 3
Let and be fixed, then there exists constants and which depend on σ and ℓ, such that for sufficiently large T, we have
where the implied constants depend at most on σ and ℓ. Moreover, when , then we can take and , where and are constants depending at most on σ.
When we try to give a different proof of Theorem 1 via Levinson's approach [18], we meet with the following problem. In particular, if the following problem has a positive solution, then a new proof for our Theorem 1 can be given.
Problem 3
Let be given. Find and some positive constant such that if k is sufficiently large, then we have
and
where is defined as
Remark 8
The arithmetic function is not multiplicative, which makes the problem difficult.
Problem 4
Study extreme values of derivatives of L‐functions.
Problem 5
In our Theorem 1, we require . What is the largest possible range for ℓ, that the result of Theorem 1 can still be valid. For instance, what can we say about the extreme values if , or ?
Problem 6
Can one find some range for ℓ, such that the results in Theorem 2 can still hold?
Remark 9
The main terms always satisfy since we have for all n and ℓ. It is not clear about the moments of derivatives of the zeta function if ℓ can depend on T. For instance, if we let , then what can we say about the second moments as ,
When ℓ depends on T, it also seems difficult to bound the contributions of .
Moreover, we have the following general problem, which asks how large or how small the extreme values of can be if ℓ can be taken arbitrary large with respect to the length T of the interval .
Problem 7
Given , decide which one of following four properties can be true.
(A) Given any function , there always exists some function such that if , then for sufficiently large T, we have
(B) Given any function , there always exists some function such that if , then for sufficiently large T, we have
(C) There exists some function , such that for all function , if , then for sufficiently large T, we have
(D) There exists some function , such that for all function , if , then for sufficiently large T, we have
JOURNAL INFORMATION
Mathematika is owned by University College London and published by the London Mathematical Society. All surplus income from the publication of Mathematika is returned to mathematicians and mathematics research via the Society's research grants, conference grants, prizes, initiatives for early career researchers and the promotion of mathematics.
ACKNOWLEDGEMENTS
I am grateful to Christoph Aistleitner for his guidance and many helpful discussions. I thank Marc Munsch for a valuable suggestion. I thank Zikang Dong and Daniel El‐Baz for their comments on early version of the paper. The work was supported by the Austrian Science Fund (FWF), project W1230.
REFERENCES
- 1. Aistleitner C., Lower bounds for the maximum of the Riemann zeta function along vertical lines, Math. Ann. 365 (2016), 473–496. [Google Scholar]
- 2. Aistleitner C., Berkes I., and Seip K., GCD sums from Poisson integrals and systems of dilated functions, J. Eur. Math. Soc. 17 (2015), 1517–1546. [Google Scholar]
- 3. Aistleitner C., Mahatab K., and Munsch M., Extreme values of the Riemann zeta function on the 1‐line, Int. Math. Res. Not. IMRN 22 (2019), 6924–6932. [Google Scholar]
- 4. Balasubramanian R. and Ramachandra K., On the frequency of Titchmarsh's phenomenon for . III, Proc. Indian Acad. Sci. Sect. A 86 (1977), 341–351. [Google Scholar]
- 5. Bondarenko A., Hilberdink T., and Seip K., Gál‐type GCD sums beyond the critical line, J. Number Theory 166 (2016), 93–104. [Google Scholar]
- 6. Bondarenko A. and Seip K., GCD sums and complete sets of square‐free numbers, Bull. London Math. Soc. 47 (2015), 29–41. [Google Scholar]
- 7. Bondarenko A. and Seip K., Large greatest common divisor sums and extreme values of the Riemann zeta function, Duke Math. J. 166 (2017), 1685–1701. [Google Scholar]
- 8. Bondarenko A. and Seip K., Extreme values of the Riemann zeta function and its argument, Math. Ann. 372 (2018), 999–1015. [Google Scholar]
- 9. Bondarenko A. and Seip K., Note on the resonance method for the Riemann zeta function, Oper. Theory Adv. Appl. 261 (2018), 121–139. [Google Scholar]
- 10. Chandee V. and Soundararajan K., Bounding on the Riemann hypothesis, Bull. Lond. Math. Soc. 43 (2011), 243–250. [Google Scholar]
- 11. de la Bretèche R. and Tenenbaum G., Sommes de Gál et applications, Proc. London Math. Soc. (3) 119 (2019), 104–134. [Google Scholar]
- 12. Dong Z. and Wei B., On large values of , arXiv:2110.04278.
- 13. Gál I. S., A theorem concerning Diophantine approximations, Nieuw Arch. Wiskunde 23 (1949), 13–38. [Google Scholar]
- 14. Granville A. and Soundararajan K., Extreme values of , The Riemann Zeta function and related themes: papers in honour of Professor K. Ramachandra, Ramanujan Mathematical Society Lecture Notes Series, vol. 2, Ramanujan Mathematical Society, Mysore, 2006, pp. 65–80. [Google Scholar]
- 15. Hilberdink T., An arithmetical mapping and applications to Ω‐results for the Riemann zeta function, Acta Arith. 139 (2009), 341–367. [Google Scholar]
- 16. Ingham A. E., Mean‐value theorems in the theory of the Riemann zeta‐function, Proc. Lond. Math. Soc. 27 (1926), 273–300. [Google Scholar]
- 17. Kalmynin A., Omega‐theorems for the Riemann zeta function and its derivatives near the line Re s = 1 , Acta Arith. 186 (2018), 201–217. [Google Scholar]
- 18. Levinson N., Ω‐theorems for the Riemann zeta‐function, Acta Arith. 20 (1972), 317–330. [Google Scholar]
- 19. Lewko M. and Radziwiłł M., Refinements of Gál's theorem and applications, Adv. Math. 305 (2017), 280–297. [Google Scholar]
- 20. Littlewood J. E., On the zeros of the Riemann zeta‐function, Proc. Camb. Philos. Soc. 22 (1924), 295–318. [Google Scholar]
- 21. Montgomery H. L., Extreme values of the Riemann zeta function, Comment. Math. Helv. 52 (1977), 511–518. [Google Scholar]
- 22. Soundararajan K., Extreme values of zeta and L‐functions, Math. Ann. 342 (2008), 467–486. [Google Scholar]
- 23. Titchmarsh E. C., The theory of the Riemann zeta‐function, 2nd ed., Oxford University Press, New York, 1986. [Google Scholar]
- 24. Voronin S. M., Lower bounds in Riemann zeta‐function theory, Izv. Akad. Nauk SSSR Ser. Mat. 52 (1988), 882–892. [Google Scholar]
- 25. Yang D., A note on log‐type GCD sums and derivatives of the Riemann zeta function , arXiv:2201.12968.