Real zeros of quadratic Hecke -functions
Abstract.
We study real zeros of a family of quadratic Hecke -functions in the Gaussian field to show that more than twenty percent of the members have no zeros on the interval .
Mathematics Subject Classification (2010): 11M06, 11M20, 11M41, 11R42
Keywords: Hecke -functions, quadratic Hecke characters, mollifier, real zeros
1. Introduction
The non-vanishing issue of central values of -functions has been studied intensively in the literature due to its deep arithmetic implications. For the classical case, it is expected that the corresponding never vanishes at for any Dirichlet character . This statement first appeared as a conjecture of S. Chowla [Chow], who concerned with the special case of primitive real characters.
Partial resolutions to Chowla’s conjecture was first given by M. Jutila [Jutila], who evaluated the first two moments of the family of quadratic Dirichlet -functions to show that infinitely many such -functions do not vanish at the central value. By evaluating mollified moments instead, K. Soundararajan [sound1] improved the result of Jutila to show that unconditionally, at least of the members of the quadratic family have non-vanishing central values. This percentage can be further improved to be if one assumes the generalized Riemann hypothesis (GRH). In fact, this follows from a result of A. E. Özluk and C. Snyder [O&S] on the one-level density of low-lying zeros of the family of quadratic Dirichlet -functions, replacing the test function used in [O&S] by an optimal one given in a paper of H. Iwaniec, W. Luo and P. Sarnak [ILS, Appendix A] .
The statements of GRH and Chowla’s conjecture imply that for every primitive quadratic Dirichlet character and all . However, it was not even known previously whether every such Dirichlet -function of sufficiently large conductor has a non-trivial real zero or not until J. B. Conrey and K. Soundararajan [C&S] proved in 2002 that for at least of the odd square-free integers we have 0 for , where is the Kronecker symbol.
Motivated by the above result of Conrey and Soundararajan, we investigate in this paper the real zeros of a family of quadratic Hecke -functions in the Gaussian number field . To state our result, we first introduce some notations. Throughout the paper, we denote and for the ring of integers of . For any , we write for its norm. We denote for the -function associated to a Hecke character of and the Dedekind zeta function of , respectively.
We say a Hecke character is primitive modulo if it does not factor through for any divisor of such that . We also say that is of trivial infinite type if its component at infinite places of is trivial. For any , we reserve in this paper the symbol for the quadratic residue symbol defined in Section 2.1. It is shown in [Gao2, Section 2.1] that defines a primitive quadratic Hecke character modulo of trivial infinite type for any odd, square-free . Here, we say that is odd if and is square-free if the ideal is not divisible by the square of any prime ideal. The family of -functions that we consider in this paper is given by
Notice that every -function in the above family satisfies a functional equation given by (2.4) which implies that is entire on the whole complex plane. As has a simple pole at , this implies that . Thus, if one assumes GRH for the above family of -functions as well as the non-vanishing of these -functions at the central point , then one can deduce that for all . Consequently, we have for all by continuity and the fact that . In this paper, we show that unconditionally, at least a positive percent of these -functions do not vanish on . By further keeping in mind that the number of odd, square-free such that is asymptotically (see [G&Zhao4, Section 3.1]), our result is given as follows.
Theorem 1.1.
For at least of the members of -functions in , we have for . More precisely, the number of -functions in such that and for all exceeds for all large .
Our proof of Theorem 1.1 is mainly based on the approach of Conrey and Soundararajan in their above mentioned work [C&S] concerning the non-vanishing of quadratic Dirichlet -functions on the real line. We also note here that similar to what is pointed out in [C&S], the approach in our paper implies that for all large , the number of odd, square-free with such that has a zero in is for any fixed and . Another outcome of our approach is that there are many -functions having no non-trivial zeros in a thin rectangle containing the real line. More precisely, there exists a constant such that for at least of the members of with , the corresponding has no zeros in the rectangle for all large .
2. Preliminaries
The proof of our result requires many tools as well as auxiliary results, which we gather them here first.
2.1. Quadratic residue symbol and Gauss sum
It is well-known that has class number one and that every ideal in co-prime to has a unique generator which is . Such a generator is called primary. Further notice that is the only prime ideal in that lies above the ideal . We now fix a generator for each ideal in by taking to be of the form with and primary. We denote the set of these generators by throughout the paper. For , we denote for their greatest common divisor such that .
We denote for a prime element in , which means that the ideal generated by is a prime ideal. We denote the group of units in by and the discriminant of by . Thus, we have and . We write for a perfect square in , by which we mean that for some .
The quadratic residue symbol is defined for any odd prime such that for any , when and with when . The definition is then extended to multiplicatively for any odd . Here, we define for .
As mentioned in Section 1, we denote for the quadratic residue symbol . In this paper, we regard as a principal character modulo , so that for all . Note that this implies that . For other values of , we shall regard as a Hecke character of trivial infinite type modulo as this is justified in [G&Zhao2019, Section 2.1]. In particular, we notice that when and .
We define the quadratic Gauss sum for with being odd, by
where we define for any complex number ,
We denote for the analogue on of the usual Möbius function on and for the number of elements in the reduced residue class of . We recall the following explicitly evaluations of given in [G&Zhao4, Lemma 2.2].
Lemma 2.2.
-
(i)
We have
-
(ii)
Let be a primary prime in . Suppose is the largest power of dividing . (If then set .) Then for ,
2.3. The approximate functional equation
For any primitive quadratic Hecke character of of trivial infinite type, a well-known result of E. Hecke asserts that has an analytic continuation to the entire complex plane with a simple pole at only when is principal. Moreover, satisfies the following functional equation (see [iwakow, Theorem 3.8])
(2.1) |
where is the conductor of , and
(2.2) |
In particular, we have the following functional equation for :
(2.3) |
When for any odd, square-free , we combine [iwakow, Theorem 3.8] and [Gao2, Lemma 2.2] to see that . Thus, the functional equation (2.1) in this case becomes
(2.4) |
We now set
(2.5) |
It follows from (2.2) and (2.4) that we have the following functional equation
(2.6) |
Throughout the paper, we fix a positive real number and let and be two complex numbers satisfying . We further denote , so that we have as well. We define for real numbers ,
(2.7) |
where
The following Lemma establishes some analytical properties concerning .
Lemma 2.4.
The function is a smooth complex-valued function for . For near , we have
(2.8) |
For large and any integer , we have
(2.9) |
Proof.
We move the line of integration in (2.7) to to deduce readily (2.8). On the other hand, for any , we have
This implies that is smooth. To prove (2.9), we may suppose that . Using the facts that for and , we see that
where the last estimation above follows from Stirling’s formula (see [iwakow, (5.113)]). The last assertion of the lemma now follows by taking above. ∎
For any complex number and any primary , we define
Note that is easily seen to be an even function of . We also define for any odd, square-free ,
Now, we present an approximate functional equation for .
Lemma 2.5.
With the notations above for and . We have for any odd, square-free ,
Proof.
We begin with the following integral for some ,
We evaluate the above integral by first writing as a Dirichlet series to see that the above expression equals upon integrating term by term. On the other hand, we move the line of integration to to encounter poles at , . The residues contribute by the functional equation (2.6). Using (2.6) again, we see that the remaining integral on the line equals to via a change of variable . This leads to the desired result. ∎
2.6. Quadratic large sieves
We include in this section two large sieve results concerning quadratic Hecke characters. They are generalizations in of the well-known large sieve results due to D. R. Heath-Brown [DRHB] on quadratic Dirichlet characters. The first lemma can be obtained by applying a large sieve result of K. Onodera [Onodera] on quadratic residue symbols in in the proof of [DRHB, Corollary 2] and [sound1, Lemma 2.4].
Lemma 2.7.
Let be positive integers, and let be arbitrary complex numbers. Let denote the set of for square-free satisfying . Then for any ,
Let be a positive integer, and for each satisfying , we write with square-free and . Suppose the sequence satisfies , then
Similarly, combining the above result of Onodera with the derivation of [DRHB, Theorem 2] and [sound1, Lemma 2.5], we have the following result.
Lemma 2.8.
Let be as in Lemma 2.7. For any complex number with , we have
2.9. Poisson summation
We recall the following two dimensional Poisson summation formula, which follows from [G&Zhao4, Lemma 2.7, Corollary 2.8].
Lemma 2.10.
Let be primary and be the quadratic residue symbol modulo . For any smooth function of compact support, we have for ,
where
(2.10) |
When applying the above lemma in the proof of our result, we are led to consider the behaviors of a particular function. In the rest of this section, we include a result on this. We begin by defining
(2.11) |
where is a smooth function compacted in and is given in (2.7).
The function that we are interested is then defined for and by
(2.12) |
Thus, is the Mellin transform of the function . Here we recall that the Mellin transform of a function is given by
For further reference, we note that if we further assume that is support in , then integration by parts implies that for , we have
(2.13) |
where we define for integers ,
Now, we have the following result concerning analytical properties of .
Lemma 2.11.
Proof.
Notice that for any smooth function , we can evaluate the function defined in (2.10) in polar coordinates as
Using this and the definition of in (2.7), we deduce that, for ,
We apply the relation (see [Gao1, Section 2.4])
to see that
Substituting this into the above expression for , we readily derive (2.14) by noticing that we can now take . This also implies that is an entire function of for .
It remains to establish (2.15). For this, we set and we may assume that here. By apply Stirling’s formula given in [iwakow, (5.112)], we see that
(2.16) |
2.12. Analytical behaviors of certain functions
Besides the function considered in the previous section, we also need to know analytical behaviors of a few other functions that are needed in the paper. We include several results in this section. First, we note that the following result can be established similar to [sound1, Lemma 5.3].
Lemma 2.13.
Let be primary and for each , we write uniquely by
(2.17) |
with square-free and . For , we have
Here is defined by
Moreover, the function is holomorphic for and satisfies the bound that uniformly for ,
Our next two lemmas provide bounds for certain dyadic sums involving and .
Lemma 2.14.
Let be two integers and let be defined in (2.17). For any sequence of complex numbers satisfying and when , we have for ,
Proof.
We write for any as with . We define for the appearing in this product,
(2.18) |
It follows from the definition of in Lemma 2.13 and part (ii) of Lemma 2.2 that we may write with , and square-free, for otherwise we have . Further, we see that when ,
Applying the above and the Cauchy-Schwarz inequality, we deduce that
where
We use the bound for given in Lemma 2.13 in the above expression to see that
(2.19) |
Note that as ,
Using this in (2.19), we obtain via another application of the Cauchy-Schwarz inequality that
We relabel by and note that for all and , we have
This implies that
(2.20) |
Lemma 2.15.
Let be two integers and let . For and any sequence of complex numbers satisfying , the expression
(2.22) |
is bounded by
and also by
Proof.
We apply Lemma 2.11, Lemma 2.13 to bound respectively and to see that the expression in (2.22) is
Upon writing , we readily deduce the first bound of the lemma from the above estimation.
To derive the second bound, we set to recast the integral in (2.14) as
This implies that
Notice that (2.16) is still valid with and so that it implies that . We apply this estimation and the Cauchy-Schwarz inequality to deduce that
By inserting the above bound into (2.22) and applying Lemma 2.14, we obtain the second bound of the lemma. ∎
In the remaining of the section, we include two more results concerning various functions studied in this paper.
Lemma 2.16.
Let be a polynomial satisfying . Let be a multiplicative function satisfying for some fixed . Let and be two bounded complex numbers such that and are for an absolute positive constant and a large real number . Then we have for ,
(2.23) |
where is a holomorphic function in defined by
Moreover, for any odd with , we have
(2.24) |
where is an absolute constant and .
Proof.
First note that we can establish (2.23) by considering Euler products. To prove (2.24), we may assume that . We then apply the Taylor expansion to write the sum in (2.24) as
We note that the inner sum above can be regarded as a Riesz type means so that we can apply the treatment given in [MVa1, Sect 5.1] to further write the above sums as
(2.25) |
To evaluate the integral above, we set and apply the zero free region for (see [MVa1, Section 8.4]) to choose a positive constant such that has no zeros in the region . We then notice that we have uniformly for . Moreover, similar to the bound given for the Riemann zeta function in [MVa1, Theorem 6.7], we have the following estimation for which asserts that for ,
(2.26) |
We now truncate the integral in (2.25) to the line segment to with . By applying the above estimations for and , we see that the error introduced by doing so is
(2.27) |
We then shift the remaining integral to the line segment . Again by the above estimations for and , we see that the two horizontal integrals are
(2.28) |
For the vertical integral, we note that as , we have
We can thus divide the vertical integral into two parts, one over the segment and one over the rest. We apply (2.26) and the bound to see that the vertical integral is
As is a polynomial, we know that only for finitely many so that we may assume that is bounded. It follows that for an appropriate positive constant , the above is
(2.29) |
Furthermore, we note that we encounter a multiple pole at in the above process. Thus, by combining (2.27), (2.28) and (2.29), we conclude that the expression given in (2.25) is
(2.30) |
We now calculate the residue above using the Taylor expansion of around and discarding the powers of that are since they make no contributions. This way, we see that the residue equals
where the first equality above following by setting while noting that . Applying this in (2.30) allows us to deduce (2.24) and this completes the proof of the lemma. ∎
For our next result, we define for odd primary primes ,
and extend the above definitions multiplicatively to functions on primary odd, square-free algebraic numbers .
Moreover, we write any odd as , with being square-free and . We define an absolutely convergent function for complex numbers in the region and such that and for primes ,
(2.31) |
Lastly, we denote for a closed contour (oriented counter-clockwise) containing the points with perimeter length being . Also, for we have , for some absolute constant and such that . We then have the following result.
Lemma 2.17.
With notations as above, we have for on the contour and ,
(2.32) |
Also, for any smooth function on and , we have
(2.33) |
Proof.
Applying the definition of given in Lemma 2.16 allows us to write the expression given in (2.32) as
where
Note here that is a multiplicative function such that . This allows us to write so that is holomorphic in .
Now we apply Perron’s formula as given in [MVa1, Theorem 5.2, Corollary 5.3] to obtain that for , ,
where
It is then easy to see that
(2.34) |
We now shift the contour of integration to to see that we have
(2.35) |
Moreover, by applying the following subconvexity bound for on the critical line given in [HB1988],
we deduce that
(2.36) |
By combining (2.34), (2.35) and (2.36), together with the observation that , , we then conclude that
If we set for brevity that , then a little calculation implies that and . This implies (2.32), which in turn implies (2.33) by partial summation and this completes the proof of the lemma. ∎
3. Plan of the proof
We define for any large number and any smooth function supported in ,
so that is a weighted count on the number of odd, square-free algebraic integers such that and that has a non-trivial real zero. The proof of Theorem 1.1 is based on an estimation of . To achieve this, we apply the following argument principle given in [C&S, Lemma 2.1], which is originally due to A. Selberg [Selberg46].
Lemma 3.1.
Let be a holomorphic function that does not vanish in the region . Let be the rectangular box with vertices , where and are two real numbers such that . Then
To proceed further, we define for any sequence of complex numbers and any smooth function supported in ,
For a real parameter to be determined later, we write
with
We then deduce that , where
Let be a small number and let be two parameters such that and . Also, let be a polynomial satisfying and , . We define a sequence such that when and , we have
(3.1) |
For other values of , we define . The definition above then implies that for all . We use the to define for any odd , the following mollifier function
(3.2) |
Further, we define for any complex number ,
We now apply Lemma 3.1 to the function with , , and and to be chosen later, such that has no zeros in . Arguing as in [C&S, Section 2], we deduce that
(3.3) |
where
(3.4) |
It remains to estimate the quantities on the right-hand side of (3.3). We first note the following result concerning , whose proof is given in Section 4.1.
Proposition 3.2.
Let be a smooth function supported on such that and . For large , and , we have
Moreover, the function has no zeros in , and we have
Next, notice that an evaluation on the terms in (3.3) involving with and requires one to study . As Proposition 3.2 enables us to treat when is slightly away from , we may focus on the case when is near . For this purpose, we shall evaluate more generally the following expression:
(3.5) |
where is given in (2.5), is a smooth function supported on and satisfy the conditions given in Section 2.3. Later, we shall set and in (3.5) to retrieve the expression .
In order to evaluate (3.5), we apply the approximate functional equation for to see that we may recast the expression in (3.5) as
In Section 4.2, we obtain the following estimation for .
Proposition 3.3.
With the above notations, we have for ,
The treatment on is more involved. Our proof in fact requires us to evaluate more generally for any odd primary . To state our result, we introduce a few notations. We define for any two complex numbers and ,
(3.6) |
Our evaluation on is given in the following result.
Proposition 3.4.
With the above notations and writing any odd primary as , with being primary, square-free and , we have
(3.7) |
Here is a remainder term bounded on average by
(3.8) |
With Propositions 3.2-3.4 available, we are able to obtain the following result concerning for being small.
Proposition 3.5.
Let be a non-negative smooth function on satisfying and . Let be a complex number such that , . We take so that , and . Then with the mollifier function being given in (3.2), we have for ,
4. Proofs of Propositions 3.2 and 3.3
4.1. Proof of Proposition 3.2
In this section we estimate by proving Proposition 3.2. We note first that similar to the bounds given in [C&S, (4.1), (4.2)], it follows from Lemma 2.7 that for any odd such that , we have
(4.1) |
Now, we write and we observe that it follows from [G&Zhao4, Section 3.1] and partial summation that we have
(4.2) |
Using the above, we see that
(4.3) |
To estimate the error terms above, for any real number , consider the integral
By moving the line of integration above to , we see that the pole at contributes . This implies that
We apply the Cauchy-Schwarz inequality to see that
It follows from this and the rapid decay of when that we have
Applying the Cauchy-Schwarz inequality again, we see that
Applying Lemma 2.8 and (4.1) in the above estimation, we see that for ,
(4.4) |
With one more application of the Cauchy-Schwarz inequality, we deduce from the above that
(4.5) |
Now, we consider the contribution from . For , we write
By (3.1) and (3.2), we see that for all and that for all , where we denote for the analogue on of the usual divisor function on . Moreover, as is supported on square-free numbers, we have . It follows that we have for all primary square values . Now, we have
(4.6) |
Using the result that
we see that the second term on the right side of (4.6) contributes . We then move the line of integration in the first term on the right side of (4.6)to to see that
(4.7) |
It follows from this and the Cauchy-Schwarz inequality that we have
We split the sum over above into dyadic blocks and apply Lemma 2.7 to see that
Combining this with (4.4), we deduce that
(4.8) |
Next, we bound by applying (4.7) to see that
(4.9) |
We now apply the Mellin transform to see that for any ,
Moving the line of integration above to , we see that we encounter a pole at if and only if and the corresponding contribution of the residue is . It follows that
where is if and otherwise. As for all perfect squares with norm , we deduce that
(4.10) |
Note that Lemma 2.8 implies that
Applying this together with (2.13) by setting and there, we obtain from (4.10) that
4.2. Proof of Proposition 3.3
In this section, we estimate by proving Proposition 3.3. We first notice that and that unless where is square-free and . It follows that
(4.12) |
where means that the sum is over odd and square-free . Now, applying twice the Cauchy-Schwarz inequality, we see that the sum over above is
(4.13) |
Note that for any , we have
(4.14) |
We write the sum above as
(4.15) |
with
Observe that the left side of (4.15) is analytic for all since is non-principal. Thus, we we may move the line of integration in (4.14) to without encountering any pole. Using the estimations that , , and the rapid decay of when , we apply the Cauchy-Schwarz inequality to see that
It follows from this and Lemma 2.8 that we have
We apply this with (4.1) to see that the expression in (4.13) is bounded by
Inserting the above in (4.12) and keeping in mind that , we see that the assertion of Proposition 3.3 follows.
5. Proof of Proposition 3.4
5.1. A first decomposition
By definition, we have
We apply the Poisson summation formula given in Lemma 2.10 to treat the last sum above to obtain that
(5.2) |
The above allows us to deduce from (5.1) that
(5.3) |
where arises from the term in (5.2) and includes the remaining non-zero terms in (5.2). Hence
(5.4) |
We show in what follows that contributes to a main term and also contributes to secondary main terms.
5.2. The principal term
Note that if and otherwise. Also, we have
Using the above observations, we see that
As with primary and square-free, we see that is equivalent to for some primary . Thus
Note that we have for any ,
where we deduce the last equality above by noticing that
We then conclude that for any ,
(5.5) |
where
(5.6) |
Now, by comparing Euler factors, we see that for ,
where and are as defined in (3.6) and (2.31). Using this in (5.6), we see that
(5.7) |
Taking above implies that . We now move the line of integration in (5.7) to to encounter simple poles at , in the process. We note the following convexity bounds for from [iwakow, Exercise 3, p. 100]:
(5.8) |
It follows that on the new line. Moreover, we have
Thus, we deduce that the integral on the new line is
It follows that
Applying this in (5.5), we obtain that
(5.9) |
5.3. The secondary main terms
We apply Lemma 2.13 in (5.10) and move the line of integration to . In this process, we encounter poles only when , with simple poles at . The residues of these poles contribute secondary main terms. We therefore write where
(5.11) |
and (after replacing by )
(5.12) |
Note that it follows from Lemma 2.2 and Lemma 2.13 that we have . Note also that we have
(5.13) |
where the last equality above follows from the observation that we have by Lemma 2.2 and Lemma 2.13.
We now define for and any complex number with ,
It follows from Lemma 2.13 that the above series converges absolutely when . We further apply (5.13) to recast as
We now apply Lemma 2.2 to evaluate defined in Lemma 2.13 to see that
(5.14) |
where with
It then follows from this and (5.14) that is analytic in the domain .
We now apply Lemma 2.11 and the observation that is an even function of to deduce that, for ,
(5.15) |
The analytical properties of discussed above allow us to move the line of integration in (LABEL:Hint) to without encountering any pole. We then deduce from this and (5.12) that
(5.16) |
We now extend the sum over in (5.16) to infinity. To estimate the error introduced, we let be the circle centred at , with radius , and oriented counter-clockwise. Notice that, for any complex number with , the function is analytic for inside . Thus, we deduce from Cauchy’s theorem that
(5.17) |
We observe that for on and that . Also, by Stirling’s formula (see [iwakow, (5.112)]), we have
Moreover, applying the convexity bounds for given in (5.8), we deduce from (5.14) that . These estimates allow us to bound the quantities in (5.17) by
It follows that
Applying the above in (5.16), we see that the error introduced by extending the sum over to infinity is
We then conclude that we have
(5.18) |
where
A direct calculation using the expression for given in (5.14) yields
Using this together with the functional equation for given in (2.3), and the identity , we obtain the following identity
(5.19) |
As it is easy to see that the left side above is invariant under , we deduce that .
We now evaluate the integral in (5.18) for by moving the line of integration to to encounter simple poles at in the process. It follows that
where the last integral above follows from the previous one upon a change of variable while noticing that . Thus we deduce that
Applying the above in (5.18), we obtain that
(5.20) |
5.4. Estimation of
In this section, we estimate given in (5.11) on average by deriving the bound given in (3.8). We denote if , and otherwise. We deduce from (5.11) that
where
Recall the definition of in (2.17). We apply the Cauchy-Schwarz inequality to see that for any integer ,
Using the Cauchy-Schwarz inequality again together with Lemma 2.8, we see that
This implies that
(5.21) |
5.5. Conclusion
Combining (5.3), (5.9), (5.20) and our result for the average size of given in the previous section, we see that in order to prove Proposition 3.4, it remains to simply the expression given in (5.9) and (5.20). We now employ the identity given in (5.19) to see that the contribution from the poles at to (5.9) and (5.20) cancel precisely each other, while the contribution from the poles at in both these expressions gives rise to the main term on the right side of (3.7), This completes the proof of Proposition 3.4.
6. Proof of Theorem 1.1
We now proceed to complete our proof of Theorem 1.1. We begin by evaluating when is near .
6.1. Proof of Proposition 3.5
We further set to rewrite the above as
(6.2) | ||||
(6.3) |
where the last estimation above follows from Proposition 3.3.
Note that we have
(6.4) | ||||
(6.5) |
We now apply Proposition 3.4 to evaluate . Using the estimations that , , , and , we deduce that various remainder terms in Proposition 3.4 contribute
(6.6) |
By taking to be the closed contour described in the paragraph above Lemma 2.17, we can evaluate the contribution of to (6.7) by
(6.8) |
Using the fact that is supported on square-free elements in , we now write , with , , being primary, square-free and . This implies that , , and so that the sum over in (6.8) can be written as
(6.9) |
We then apply the relation
to recast the expressions given in (6.9) as
Next, we use the Möbius function to detect the condition in the above expression to see that it equals to
(6.10) |
We further group terms according to to see that (6.10) becomes
(6.11) |
We now apply Lemma 2.16 to evaluate the sum over given in (6.11) by first considering the case . We set , , and we denote by in Lemma 2.16 to see that by applying it with , and then with , , the sum of these two applications yields
(6.12) |
Here one checks that the sum of the terms of the above two applications of Lemma 2.16 cancel each other. Thus the main term above comes from the sum of the terms only. Note also that the main term above is due to the choice of the contour .
Now, replacing by and by in (6.12), we see that a similar expression for the sum over in (6.11) holds. It follows that the case in (6.11) equals
We apply the last expression above in (6.8) to see that, based on our choice of and , the contribution of the error term above is
(6.13) |
On the other hand, the main term coming from the contribution of the terms to (6.8) equals to
(6.14) |
Note that as vanishes at , the integrand in (6.14) has only a simple pole at inside . It then follows from Cauchy’s theorem that the expression in (6.14) equals
(6.15) |
We apply Lemma 2.17 to evaluate (6.15) and then combine the result together with the error term given in (6.13) to see that the contribution to from the part of restricting to the case equals
(6.16) |
We now evaluate (6.11) for the case by applying Lemma 2.16 with , , , as before, and with , this time. We deduce then that for any odd ,
(6.17) |
To evaluate the residue above, we write the Taylor expansion of as to see that , , , and that for . It follows from this that the residue term in (6.17) equals
(6.18) |
Again, replacing by and by in (6.18) yields an expression for the sum over in (6.11) when . We can thus apply these expressions to evaluate (6.11) for the case . By doing so, we see that the contribution of the remainder terms is
(6.19) |
On the other hand, by writing for , we obtain that the main term is
We apply Lemma 2.17 to compute the expression above. By a suitable change of variables, we see that it equals
(6.20) |
where
We now apply (6.19) and (6.20) to evaluate the expression in (6.8) when . Based on our choices for and , we see that the error terms in (6.19) and (6.20) contribute
(6.21) |
Moreover, we see that the main term in (6.8) equals
(6.22) |
Applying the relations
we see that the expression in (6.22) equals
(6.23) |
Now, using integration by parts together with the observations that , and , , we obtain after a little calculation that
The above expressions allow us to recast (6.23) as
It follows from this and (6.21) that the contribution of to from the terms equals
Combining the above with (6.16), we conclude that the contribution to equals
(6.24) |
We obtain the contribution to (6.7) in a similar way to see that it equals
(6.25) |
6.2. Completion of the proof
We are now able to complete our proof of Theorem 1.1. For this, we take our parameters exactly as those used by Conrey and Soundararajan in their proof of [C&S, Theorem 1]. Precisely, we take in (3.2) and with . We further take and . Let be given as in Proposition 3.2 and let to be a smooth function supported in satisfying for all , for , and . We apply Proposition 3.2 in (3.3) to deduce that
(6.26) |
where and are given in (3.4).
Now, Proposition 3.5 implies that for real numbers satisfying and , we have
where
The above expression implies that , from which we deduce that
(6.27) |
Moreover, based on Proposition 3.2 and our choice for , we can extend the integral in the definition of in (3.4) to infinity with an negligible error. This way, we obtain that
(6.28) |
We conclude from (6.26)-(6.28) that
As shown in [C&S, p. 10], the above further implies that . By summing over , , , we obtain the proof of Theorem 1.1.
Acknowledgments. P. G. is supported in part by NSFC grant 11871082.