The fourth moment of central values of quadratic Hecke -functions in the Gaussian field
Abstract.
We obtain an asymptotic formula for the fourth moment of central values of a family of quadratic Hecke -functions in the Gaussian field under the generalized Riemann hypothesis (GRH). We also establish lower bounds unconditionally and upper bounds under GRH for higher moments of the same family.
Mathematics Subject Classification (2010): 11M06, 11M41, 11M50
Keywords: central values, Hecke -functions, quadratic Hecke characters, moments
1. Introduction
Moments of central values of families of -functions have been intensively studied in the literature in order to understand important arithmetic information they carry. Although much progress has been made towards establishing asymptotic formulas for the first few moments for various families of -functions, little is known for the higher moments. In connection with random matrix theory, conjectures on the order of magnitude for the moments were made by J. P. Keating and N. C. Snaith in [Keating-Snaith02]. A simple and powerful method towards establishing lower bounds of these conjectured results was developed by Z. Rudnick and K. Soundararajan in [R&Sound] and [R&Sound1].
For the family of quadratic Dirichlet -functions, the above method of Rudnick and Soundararajan allows them to show in [R&Sound1] that for every even natural number ,
(1.1) |
where is the Kronecker symbol and denotes the set of fundamental discriminants.
In the other direction, Soundararajan [Sound01] proved that, assuming the generalized Riemann hypothesis (GRH), for all real ,
(1.2) |
The above result was further sharpened by A. J. Harper in [Harper] to remove the power.
Besides the lower and upper bounds for the moments given in (1.1) and (1.2), asymptotic formulas are known for integers in the form
(1.3) |
where is an explicit linear polynomials of degree and is the error term. Here we note that for odd, square-free , the character is primitive modulo satisfying . Note also that we choose to present the asymptotic formulas for a family which is preferred by most recent studies due to technical reasons instead of the family appearing in (1.1) and (1.2).
Evaluation of the first two moments () in (1.3) was initiated by M. Jutila in [Jutila]. The error terms in Jutila’s results were subsequently improved in [ViTa, DoHo, Young1] for the first moment and in [sound1, Sono] for the second moment. For smoothed first moment, a result of D. Goldfeld and J. Hoffstein in [DoHo] implies that one can take . An error term of the same size was later obtained by M. P. Young in [Young1] using a recursive approach. The same approach was then adapted by K. Sono [Sono] to show that one can take for smoothed second moment. The sizes of these error terms then make the expressions given in (1.3) in agreement with a conjecture made by J. B. Conrey, D. W. Farmer, J. P. Keating, M. O. Rubinstein and N. C. Snaith in [CFKRS] on asymptotic behaviours of these moments of the family of quadratic Dirichlet -functions.
The third moment given in (1.3) was originally obtained by Soundararajan in [sound1]. The error term in Soundararajan’s result was later improved by Young in [Young2] to be for the smoothed version. Under GRH, Q. Shen [Shen] established an asymptotic formula for the fourth moment given in (1.3) with some savings on the powers of in the error term. The result of Shen is achieved by combining lower and upper bounds for the fourth moment, with a lower bound obtained unconditionally using the method of Rudnick and Soundararajan in [R&Sound, R&Sound1] (note that such a lower bound is stated in [R&Sound1] without proof). An upper bound for the fourth moment is obtained under GRH by making use of upper bounds for the shifted moments of quadratic Dirichlet -functions. This approach originated from earlier work of Soundararajan [Sound01] as well as Soundararajan and Young [S&Y], especially the treatment in [S&Y] on the second moment of quadratic twists of modular -functions.
The result of Soundararajan and Young in [S&Y] supplies another example on moments of quadratic families of -functions. In [Gao1], the author considered the analogue case of evaluating moments of central values of a family of quadratic Hecke -functions in the Gaussian field to obtain the smoothed first three moments of such family with power saving error terms.
To describe the family studied in [Gao1], we let be the Gaussian field throughout the paper and be the ring of integers of . We denote for a Hecke character of and we recall that a Hecke character of is said to be of trivial infinite type if its component at infinite places of is trivial. We write for the -function associated to and for the Dedekind zeta function of . In the rest of the paper, we reserve the expression for the the quadratic residue symbol defined in Section 2.1. We also denote for the norm of any .
We then consider the family of -functions:
(1.4) |
Here we say that any is odd if and is square-free if the ideal is not divisible by the square of any prime ideal. Note also that (see Section 2.1 below) the symbol defines a primitive quadratic Hecke character modulo of trivial infinite type when is odd and square-free.
The first three moments of the above family is studied in [Gao1]. In this paper, our main purpose is to establish the fourth moment of the same family under GRH. In this process, we shall need to make use of results on upper bounds for shifted moments of the family under GRH. Therefore, we begin by considering lower and upper bounds for higher moments of this family. Unconditionally, we have the following result concerning the lower bounds.
Theorem 1.1.
For every even natural number , we have
(1.5) |
Here the “” on the sum over means that the sum is restricted to square-free elements in .
The proof of Theorem 1.1 is given in Section 3 and it follows the arguments in the proof of [R&Sound1, Theorem 2] for the case of Dirichlet -functions. Also similar to the remarks given below [R&Sound1, Theorem 2], the proof of Theorem 1.1 can be applied to give lower bounds for the moments for all rational numbers , provided that we replace by .
Before we state a corresponding result on the upper bounds, we need to introduce some notations. Let such that and , we define
(1.9) |
We further define for ,
(1.10) | ||||
(1.11) |
Then we have the following result on upper bounds for shifted moments of the family given in (1.4) under GRH.
Theorem 1.2.
Assume GRH for and for all odd, square-free . Let be large and let with , and . Then for any positive real number and any , we have
The proof of Theorem 1.2 is given in Section 4 and our proof follows closely the approaches in [Sound01, S&Y, Shen]. We note that similar results to Theorem 1.2 were obtained for the moments of the Riemann zeta function by V. Chandee [Chandee] and for the moments of all Dirichlet -functions modulo by M. Munsch [Munsch].
We now give two consequences of Theorem 1.2. First, by setting in Theorem 1.2, we deduce immediately the following upper bounds for moments of the central values of the family given in (1.4).
Corollary 1.3.
Assume GRH for and for all odd, square-free . For any positive real number and any , we have for large ,
The second consequence concerns upper bounds for shifted fourth moment of the family given in (1.4), which is what we need in a refined study in Section 5 on the fourth moment under GRH.
Corollary 1.4.
Assume GRH for and for all odd, square-free . Let be large and let with and . Then
We shall omit the proof of the above corollary as it is analogous to the proof [Shen, Theorem 2.4]. With the aide of Corollary 1.4, we are able to obtain a more precise expression for the fourth moment. To present our result, we define constants for any real number by
(1.12) |
Here and in what follows, we denote for a prime number in , by which we mean that the ideal generated by is a prime ideal. We also note that the expression indicates that is a primary element in defined in Section 2.1.
Now, we state our result on a conditional asymptotic evaluation on the fourth moment.
Theorem 1.5.
Assume GRH for and for all odd, square-free . Then for any , we have
Here the “” on the sum over means that the sum is restricted to square-free elements in and is defined as in (1.12).
Without assuming GRH, we also have the following lower bound for the fourth moment.
Theorem 1.6.
Unconditionally, we have
Here the “” on the sum over means that the sum is restricted to square-free elements in and is defined as in (1.12).
Theorems 1.5 and 1.6 are similar to those of Shen given in [Shen, Theorems 1.1-1.2] on the fourth moment of the family of quadratic Dirichlet -functions as well as the result of Soundararajan and Young given in [S&Y, Theorem 1.1] on the second moment of the family of quadratic twists of modular -functions. The proof for Theorems 1.5 and 1.6 given in Section 5, also proceed along the same lines of the proofs of [Shen, Theorems 1.1-1.2] and [S&Y, Theorem 1.1].
2. Preliminaries
As a preparation, we first include some auxiliary results needed in the proofs of our theorems.
2.1. Quadratic residue symbol and quadratic Gauss sum
Recall that and it is well-known that have class number one. We denote and for the group of units in and the discriminant of , respectively.
Every ideal in co-prime to has a unique generator congruent to modulo which is called primary. It follows from Lemma 6 on [I&R, p. 121] that an element with is primary if and only if or .
For , we denote the symbol for the quadratic residue symbol modulo in . For a prime with , the quadratic symbol is defined for , by , with . When , we define . Then the quadratic symbol is extended to any composite with multiplicatively. We further define for .
The following quadratic reciprocity law (see [G&Zhao4, (2.1)]) holds for two co-prime primary elements :
(2.1) |
Moreover, we deduce from Lemma 8.2.1 and Theorem 8.2.4 in [BEW] that the following supplementary laws hold for primary with :
(2.2) |
For any complex number , we define
For any , we define the quadratic Gauss sum associated to any quadric Hecke character modulo of trivial infinite type and the quadratic Gauss sum associated to the quadratic residue symbol for any by
(2.3) |
When , we shall denote for and for . Recall from [G&Zhao3, (2.2)] that for primary , we have
(2.4) |
A Hecke character is said to be primitive modulo if it does not factor through for any divisor of such that . Recall from Section 1 that we denote for the the quadratic residue symbol and we define to be when . In Section 2.1 of [G&Zhao4], it is shown that the symbol defines a primitive quadratic Hecke character modulo of trivial infinite type for any odd and square-free . When replacing by , we see that the symbol also defines a primitive quadratic Hecke character modulo of trivial infinite type for any odd and square-free . Our next lemma evaluates exactly.
Lemma 2.2.
For any odd, square-free , we have
(2.5) |
Proof.
It suffices to prove (2.5) with replaced by , where or and is primary and square-free. It follows from the Chinese remainder theorem that varies over the residue class modulo when and vary over the residue class modulo and , respectively. We then deduce that
As is a Hecke character of trivial infinite type modulo , we deduce that
(2.6) |
On the other hand, we denote to be the unique element in such that is primary for any . It follows from the quadratic reciprocity law (2.1) that
(2.7) |
In order to evaluate the last sum in (2.8), we note that it suffices to take to vary over the reduced residue class modulo . One representation of such class consists of the following elements (note that consists of the reduced residue class modulo and consists of the residue class modulo ):
We further write with (recall that or ) and check by direct calculations using (2.2) to see that (2.5) is valid with replaced by , where or and is primary and square-free. This completes the proof of the lemma. ∎
Let denote the number of elements in the reduced residue class of , we recall from [G&Zhao4, Lemma 2.2] the following explicitly evaluations of for primary .
Lemma 2.3.
-
(i)
We have
-
(ii)
Let be a primary prime in . Suppose is the largest power of dividing . (If then set .) Then for ,
2.4. The approximate functional equation
Let be a primitive quadratic Hecke character modulo of trivial infinite type of . Let
(2.9) |
A well-known result of E. Hecke shows that has an analytic continuation to the whole complex plane and satisfies the functional equation (see [iwakow, Theorem 3.8])
(2.10) |
where .
When we take for any odd, square-free , it follows from [iwakow, Theorem 3.8] that we have , so that Lemma 2.2 implies that in this case the functional equation becomes
For and rational integer , we let denote the analogue on of the usual function on . Thus equals the coefficient of in the Dirichlet series expansion of the -th power of . We shall also write for . In particular, when is primary, we have and
We denote also for rational integer and any real number ,
(2.11) |
We shall also write for respectively in the rest of the paper.
By setting in [Gao1, Lemma 2.6], we obtain the following approximate functional equation for . Note here that the derivation of [Gao1, Lemma 2.6] assumes a rapid decay of but the proof also carries over to the case due to the rapid decay of .
Lemma 2.5 (Approximate functional equation).
For any odd, square-free , we have for ,
(2.12) |
2.6. Poisson summation
In this section we gather some Poisson summation formulas over . We first recall that the Mellin transform for any function is defined to be
We now state a formula for smoothed character sums over all elements in .
Lemma 2.7.
Let and let be a Hecke character of trivial infinite type. For any smooth function of compact support, we have for ,
(2.13) |
The above is also valid when we replace by and by . Here are defined in (2.3) and
(2.14) |
Moreover, the function is real-valued for all and when , we have for ,
(2.15) |
Proof.
This lemma is essentially [G&Zhao4, Lemma 2.7] except for the last assertion. To establish it, we evaluate (2.14) using polar coordinates to see that
(2.16) |
The first equality above shows that for all .
We make some changes of variables (first , then ) to see that
(2.17) |
where the last line above follows from the relation (see [Gao1, Section 2.4]) that
By a further change of variable in the last integral of (2.17), we can recast as
(2.18) |
where we can retake as well and this completes the proof. ∎
We remark that when is a primitive Hecke character, we have
It follows from this and the expression given in (2.15) that the formula given (2.13) is equivalent to a version of the Poisson summation formula over number fields by L. Goldmakher and B. Louvel in [G&L, Lemma 3.2] for the case of the Gaussian field.
In the proof of Theorems 1.5 and 1.6, we need to consider a smoothed character sum over odd algebraic integers in . For this, we quote the following Poisson summation formula from [G&Zhao4, Corollary 2.8], which is a consequence of Lemma 2.7 above.
Lemma 2.8.
Let be primary and be the quadratic residue symbol . For any smooth function of compact support, we have for ,
2.9. Analytical behaviors of certain Dirichlet series
In this section, we discuss the analytical behaviors of several Dirichlet series that are needed in the proof of Theorems 1.5 and 1.6. We first define for ,
(2.19) |
where we define for primary by
(2.20) |
Our first result concerns the analytical behaviors of . The proof is similar to [Shen, Lemma 4.1], so we omit it here.
Lemma 2.10.
For , we have
where
and where for primary ,
Moreover, is analytic and uniformly bounded in the region .
Let be defined as in (2.3). We now fix a generator for every prime ideal by taking to be primary if and for the ideal (noting that is the only prime ideal in that lies above the integral ideal ). We also fix as the generator for the ring itself and extend the choice of the generator for any ideal of multiplicatively. We denote the set of such generators by . For any , we shall hence denote to be the unique pair of elements in such that with being square-free and . In this way, we define
Our next lemma gives the analytic properties of , we omit the proof here since it is similar to [Shen, Lemma 5.2].
Lemma 2.11.
For any , let with square-free and . Then for , we have
where
Here for ,
and for ,
Moreover, is analytic in the region and for , we have
(2.21) |
where the implied constant is absolute.
We define for a prime , ,
With the above notations, we note the following result concerning a Dirichlet series related to .
Lemma 2.12.
For , we have
(2.22) |
where
Here for , and for , .
In addition, we have
-
(1)
is analytic and uniformly bounded for .
-
(2)
for with the implied constant being absolute.
Proof.
We let and we divide the sum over in (2.22) into two sums, according to or not, to see that
Note that when , by Lemma 2.3. It follows that we have so that
We then deduce from this and the definition of given in Lemma 2.11 that
where the last equality above follows from the observation that is multiplicative with respect to and that we have when is primary from Lemma 2.3. The assertions of Lemma 2.12 now follows by arguing similarly to the proof of [Shen, Lemma 5.3]. ∎
Lastly, we note the following result which can be established similar to the proof of [Shen, Lemma 5.4].
Lemma 2.13.
For , we have
where
Moreover, is analytic and uniformly bounded for .
2.14. A mean value estimate for quadratic Hecke -functions
In [DRHB, Theorem 1], D. R. Heath-Brown established a powerful quadratic large sieve result for Dirichlet characters. Such result was extended by K. Onodera in [Onodera] to quadratic residue symbols in the Gaussian field. Applying Onodera’s result in a similar fashion as in the derivation of [DRHB, Theorem 2] by Heath-Brown to obtain a mean value estimation for the fourth moment of the family of primitive quadratic Dirichlet -functions, we have the following upper bound for the fourth moment of quadratic Hecke -functions unconditionally.
Lemma 2.15.
Suppose is a complex number with . Then
3. Proof of Theorem 1.1
We recall a result of Landau [Landau] implies that for an algebraic number field of degree and any primitive ideal character of with conductor , we have for ,
(3.1) |
where denotes the norm of and respectively, denotes the discriminant of and runs over integral ideas of .
We deduce from (3.1) by partial summation that for odd, square-free , the series
is convergent and equals to . In particular, this implies that . It follows from this and the observation that is an even natural number that we may further restrict the sum over in (1.5) to satisfy .
We set and apply Hölder’s inequality to see that
(3.2) |
where
and
In the remaining of the proof, it thus suffices to bound and . We bound first by noting that
where is any non-negative smooth function that is supported on for some fixed small such that for .
We now expand to see that
We consider the inner sum above by setting . Using Möbius function to express the condition that is square-free, we see that
(3.3) |
Note that for smoothed sums involving any non-principal Hecke character modulo of trivial infinite type, we have (see [G&Zhao2019, (1.4)]) that for
(3.4) |
Now, we write with and being primary and apply the quadratic reciprocity law (2.1) to see that
(3.5) |
Note that if is not a square then the symbol can be regarded as a non-principal Hecke character of trivial infinite type. By decomposing the sum over in (3.3) into sums over and apply (3.4) to the sum over , we deduce that the sum over in (3.3) is . On the other hand, the sum over is trivially . We then conclude that if is not a square,
(3.6) |
If is a perfect square, then by applying the following result for the Gauss circle problem (see [Huxley1]) with ,
(3.7) |
together with a routine argument, we see that
(3.8) |
where is defined in (2.20).
We note that
(3.10) |
Similar to [Selberg, Theorem 2], we have that for a positive constant ,
(3.11) |
We evaluate next. By applying the approximate functional equation (2.12) for , we see that
(3.13) |
Here we note that (see [sound1, Lemma 2.1]) is real-valued and smooth on and the -th derivative of satisfies
(3.14) |
If is not a square, then using (3.4), (3.14) and partial summation we see that
(3.15) |
If is an square, then using the right-hand side expression in (3.8) together with (3.14) and partial summation implies that the sum over in (3.13) is
(3.16) |
Applying (3.15) and (3.16) in (3.13), we see that the error terms in (3.15) and (3.16) contribute to (3.13) an amount .
To estimate contribution of the main term in (3.16) to (3.13), we write as where and are primary and is square-free. Then must be of the form where is primary. It follows that the contribution of the main term in (3.16) to (3.13) is
Note that , and by a standard argument, we see that the sum over above is
4. Proof of Theorem 1.2
4.1. A few lemmas
We first include a few lemmas needed in the proof of Theorem 1.2. We denote for the von Mangoldt function on . Thus equals the coefficient of in the Dirichlet series expansion of . Our first lemma provides an upper bound of in terms of a sum involving prime powers.
Lemma 4.2.
Let be a non-principal primitive quadratic Hecke character modulo of trivial infinite type. Assume GRH for and . Let be a large number and . Let denote the unique positive real number satisfying . For all we have uniformly for and that
(4.1) |
where the sum means that the sum is over integral ideals of .
Proof.
We denote for and we interpret as when . Thus we may suppose in the rest of the proof. Recall the associated function defined as in (2.9). Here is analytical in the entire complex plane since is non-principal. As has simple poles at the non-positive rational integers (see [Da, §10]), we see from the expression of from (2.9) that has simple zeros at , these are called the trivial zeros of . Since we are assuming GRH, we know that the non-trivial zeros of are precisely the zeros of .
Let run over the non-trivial zeros of . We then deduce from [HIEK, Theorem 5.6] and the observation that is analytic at since is non-principal and primitive that
(4.2) |
where are constants.
Taking the logarithmic derivative on both sides of (4.2) and making use of (2.9), we obtain that
This implies that
(4.3) |
On the other hand, combining the functional equation (2.10) with (4.2), we see that
Taking logarithmic derivative on both sides of the above expression, we obtain that
Here the second equality above follows by noting that the terms containing and cancel as both are zeros from the functional equation (2.10). We note here that
is convergent since if is a zero, so is and we have
and we know that is convergent by [iwakow, Lemma 5.5].
We then deduce that
Combining the above with (4.3), we see that
(4.4) |
where the third line above follows from by (6) of [Da, §10] and where we define
Integrating the last expression given for in (4.4) from to , we obtain by setting that
(4.5) |
where the last inequality above follows from the observation that for all satisfying .
Next, we deduce upon integrating term by term using the Dirichlet series expansion of that
where is a large number. Now moving the line of integration in the above expression to the left and calculating residues, we see also that
Comparing the above two expressions, we deduce that unconditionally, for any , we have
(4.6) |
Our next lemma treats essentially the sum over prime squares in (4.1).
Lemma 4.3.
Proof.
We first note that under GRH for , the prime ideal theorem has the following form (see [iwakow, Theorem 5.15])
(4.10) |
It follows from this and that we have
We then deduce from this that it suffices to establish (4.9) with the left-hand side expression in (4.9) being replaced by , where
It is easily seen that when using (4.10) and partial summation. The same procedure also implies that
(4.11) |
It follows that when , we have
where the integral is along the line segment connecting the origin and the point on the complex plane.
In the remaining case when , we note that by (4.10) and partial summation,
Now we assume that and and we integrate from to to see that
We break the integration into two parts, one horizontal integration along the -axis from to , and the other vertical integration from to . The horizontal integration is
If we write , then the vertical integration can be evaluated by breaking the integral over for and . We obtain this way that the vertical integration is
It follows that we have when and . In particular, this applies to the case when , thus completes the proof. ∎
Lastly, we present a mean value estimation which will be applied to estimate the sum over primes in (4.1) in our proof of Theorem 1.2.
Lemma 4.4.
Let and be real numbers. For fixed , let be a natural number with . Then for any complex numbers , we have
Proof.
Let be any non-negative smooth function that is supported on for some fixed small such that for . We have that
We further expand out the square in the last sum above and treat the sum over by applying (3.5) first and then using the smoothed version of Pólya-Vinogradov inequality (3.4) for number fields when the product of the primes involved is not a perfect square to see that we have
where we write for a square of an element in .
We take small enough so that . Then an argument similar to that in the proof of [S&Y, Lemma 6.3] leads to the assertion of the lemma. ∎
4.5. Completion of the proof
With lemmas 4.2-4.4 now available, we proceed to establish an upper bound for the frequency of large values of .
Proposition 4.6.
Assume GRH for and for all odd, square-free . Let be large and let with , and . Let denote the number of odd, square-free such that and
Then for , we have
for , we have
for , we have
Proof.
Let be the constant defined in Lemma 4.2 and apply this Lemma with and to for , we see that for ,
We then deduce that
(4.12) |
The terms with in the the above sum contribute . Using the fact , we deduce from Lemma 4.3 that
where is defined as in (1.10).
Applying the above estimation in (4.12), we obtain that
(4.13) |
By taking in (4.13) and bounding the sum over in (4.13) trivially (with the help of (4.10)), we see that for . Thus, we can assume .
In what follows, we shall denote for defined in (1.11) and we note that . We now set with
We further denote , for the real part of the sum in (4.13) truncated to , for the real part of the sum in (4.13) over . We then deduce that
It follows from this that if , then we have either
Now, we define
We then take to see that by Lemma 4.4, we have
where the last estimation above follows from (4.11) and Stirling’s formula (see [iwakow, (5.112)]), which implies that
We then deduce that
(4.14) |
Next, we estimate . For any , we obtain using Lemma 4.4 that
(4.15) |
where
By arguing as in the proof of Lemma 4.3, we see that
Combining with (4.15), this implies that
We now take when and otherwise to see that in either case, we have for large,
A little calculation then shows that
We then deduce from the above and (4.14) that
It is then easy to check that this leads to the assertion of the proposition. ∎
5. Proof of Theorems 1.5 and 1.6
5.1. Initial treatment
Let be a smooth Schwartz class function which is compactly supported on satisfying for all . We apply the approximate functional equation (2.12) to see that
where
(5.1) |
For two parameters satisfying , we define
(5.2) |
We let
(5.3) |
Then applying (5.1) to (5.2) and using the Möbius inversion to remove the square-free condition in (5.2), we obtain that
Now we separate the terms with and with for some to be chosen later, writing , respectively. We bound first in the following result.
Lemma 5.2.
Unconditionally, we have . Under GRH, we have .
Proof.
We first write with square-free and primary. We then let and apply the definition of in (5.3) to see that
(5.4) |
where we define to be the function by removing the Euler factors from at prime ideals dividing .
Applying the estimation
we can bound by moving the lines of the integrations in (5.4) to to see that
(5.5) |
By Corollary 1.4, we see that for ,
(5.6) |
Note that Lemma 2.15 implies that
(5.7) |
Next, we treat by applying the Poisson summation formula given in Lemma 2.8 to recast it as
(5.9) |
where
Now we write , where corresponds to the term with . By applying (2.11) and (2.18), we see that when ,
(5.10) |
where
and where the last expression in (5.10) follows from moving the lines of the first triple integral in (5.10) to , and a change the variables .
Substituting the last expression in (5.10) to (5.9), we see by using our notation for given in Section 2.9 that
(5.11) |
We observe that if we move the lines of integrations over in the last expression in (5.11) to the left, then we encounter poles at only when . For this reason, we further write , where
(5.12) |
and
(5.13) |
5.3. Computing : the term
Note that by Lemma 2.3 we have if , and otherwise. Thus we get
(5.14) |
We note that
Due to the rapid decay of , we can estimate the error term above as
We now denote with being primary together with a change of variables: to see that for the new variables , we have and the condition that further implies that both are squares now so that we have
where the last line follows by applying estimations similar to that given in (5.8).
Further note that we have
It follows from (5.15) and Lemma 2.10 that
(5.16) |
where
Applying Lemma 2.10 again, we see that is analytic for .
We first move the lines of the integrations in (5.16) to by noting that we encounter no poles. We then move the line of the integration over to to see that we encounter two poles of order at and in the process. It follows that
(5.17) |
It follows by moving the line of the integration over from to that we have
Similarly, we have that
5.4. Computing : the term
As is analytic in the region by (1) of Lemma 2.12 and is analytic at , we move the lines of the integration above to without encountering any poles to see that
(5.19) |
We extend the sum over in (5.19) to include all primary elements in , introducing an error term
(5.20) |
To facilitate our estimation of the triple integral in the above expression and other similar integrals in what follows, we gather here a few bounds on that hold uniformly in specified regions. On write , we have
(5.21) |
The first and third estimation above can be established similar to the proofs of [MVa1, Corollary 1.17] and [MVa1, Lemma 6.7], respectively. The second estimation above is the convexity bound for (see [iwakow, Exercise 3, p. 100]).
Also, by applying (3.7), we see that for the -th derivative of with , we have for ,
We deduce readily from the above that for and , we have
(5.22) |
We further note that integrating by parts implies that for and any integer ,
(5.23) |
where
We move the lines of the integrations in (5.20) over to without encountering any poles. Then by (2) of Lemma 2.12 and the estimations given in (5.21) and (5.23), we see that on the new lines of integrations, the expression in (5.20) is
where the last estimation above follows from (5.23) with and the bound .
We conclude from the above discussions and (5.19) that
(5.24) |
We now move the lines of the integrations in (5.25) to without encountering any poles, as is analytic and uniformly bounded in the region by Lemma 2.13. Next, we move the line of the integration over to to encounter a pole of order at and a pole of order at to see that triple integral in (5.25) is
(5.26) |
where the error term above follows from using estimations given in (5.21) and (5.23) and where we denote for the residues of the integrand in (5.25) at and , respectively.
It is easy to see that is analytic for by Lemma 2.13 and that
It follows from this that by moving the line of the double integral in (5.26) involving from to and applying (5.21), (5.22), we have
(5.27) |
Note that
As one checks that the expression in the parenthesis above is analytic for , we can move the line of the integral in (5.27) involving to . In this process, we encounter a pole at so that we have
(5.28) |
where
To evaluate , we use the fact that when and the functional equation (2.10) for :
to obtain that . On the other hand, a direct calculation shows that . We then deduce that
Similarly, by moving the line of the integration over from to in the double integral of in (5.26) and applying estimations given in (5.21)-(5.23), we see that
(5.29) |
We now summarize our result on in the following lemma, by combining the estimations from (5.25)-(5.29).
Lemma 5.5.
We have
5.6. Computing : the term
In this section, we estimate . We first deduce from (5.13) that
(5.30) |
Let be a parameter to be chosen later. We denote for the right-hand side expression above truncated to and for the right-hand side expression above over . For , we shift the the lines of the integrations to . For , we shift the the lines of the integrations to . Thus we obtain via (2.21) that
(5.31) |
where denotes the sum over all square-free elements in .
By the Cauchy-Schwarz inequality, we have that
(5.32) |
Similar to our proof of Theorem 1.2 (and hence Corollary 1.3), we can show that under GRH, for ,
Applying the above estimation in (5.32), we deduce that for ,
(5.33) |
Unconditionally, we have similar to Lemma 2.15 that
It follows from this and partial summation that we have
(5.34) |
Applying the estimation given (5.33) to (5.31) for and using (5.34) otherwise, we obtain by noting the exponential decay of that
(5.35) |
Similarly, we have
(5.36) |
We further observe that if instead we apply (5.34) to (5.31) for all , then we obtain corresponding estimations for by replacing with in both (5.35) and (5.36). On setting and keeping in mind that our choices of and will be at most powers of (see Section 5.8), we immediately derive from (5.30), (5.35) and (5.36) the following result.
Lemma 5.7.
Unconditionally, we have
Under GRH, we have
5.8. Proof of Theorem 1.5
We now complete the proof for Theorem 1.5 in this section. By setting and , we deduce from (5.2), (5.18), Lemma 5.2, Lemma 5.5 and Lemma 5.7 that under GRH,
(5.37) |
Here we note that, similar to [sound1, Lemma 2.1], we can show that the function appearing in the definition of given in (5.1) is a real-valued function, so that we have .
We define to see that
We move the line of the integral in the above expression for to by realizing that is entire there. Further applying the bound that for , we see that
It follows that
We estimation the sum over above using Corollary 1.4 when and Lemma 2.15 otherwise. We then deduce by the exponential decay of in that
(5.38) |
Combining (5.37) and (5.38), we see that
(5.39) |
where the last expression above follows from an application of the Cauchy-Schwarz inequality to estimate the sum involving the product of and .
We now take to be supported on satisfying for and for all rational integer . We then deduce that , and that . Thus (5.39) implies that
We then deduce by taking that
(5.40) |
Similarly, we can choose in (5.39) such that for , for all , and for all . Taking , we can deduce that
(5.41) |
Combining (5.40) and (5.41), we obtain that
The assertion of Theorem 1.5 now follows by summing the above over , , and then resetting to be .
5.9. Proof of Theorem 1.6
In this section, we complete the proof for Theorem 1.6. We first apply the Cauchy-Schwartz inequality to see that
(5.42) |
where
Here we recall that is defined as in (5.1).
We can evaluate similar to our evaluation of in Section 5.8, except that we now set . By setting again, we deduce from (5.18), Lemma 5.2, Lemma 5.5 and Lemma 5.7 that unconditionally,
with the implied constant in being absolute.
To evaluate , we recast it as
where
A similar argument to our evaluation of above implies that, by taking , we have unconditionally,
with the implied constant in being absolute.
We now take and take so that for , for all , and for all . Applying our estimations for and in (5.42), we deduce that
The assertion of Theorem 1.6 now follows by summing the above over , , and then resetting to be .
Acknowledgments. P. G. is supported in part by NSFC grant 11871082.