Sifting for small split primes
of an imaginary quadratic field
in a given ideal class
Abstract.
Let , be a prime, and let be an ideal class in the field . In this article, we give a new proof that , the smallest norm of a split prime , satisfies for some absolute constant . Our proof is sieve theoretic. In particular, this allows us to avoid the use of log-free zero-density estimates (for class group -functions) and the repulsion properties of exceptional zeros, two crucial inputs to previous proofs of this result.
Keywords: sieve, primes, binary quadratic forms, Linnik theorem.
1. Introduction
For integers and , let denote the least prime . In 1944, Linnik [29] showed that
(1) |
where both and the implied constant are absolute. Since then, there have been many improvements on this result. Building on the work of Heath-Brown [19], Xylouris [37] showed that unconditionally one can take in (1), which is the current record. This comes quite close to the bound
which is what follows assuming that the Riemann hypothesis holds for the Dirichlet -functions .
Such results are difficult to establish unconditionally, and have traditionally (following Linnik) depended on deep results on the zeros of these -functions, namely a log-free zero-density estimate and a quantitative version of the Deuring-Heilbronn phenomenon (exceptional zero repulsion effect).
Linnik’s theorem has been generalized in the setting of the Chebotarev density theorem: given a Galois extension of number fields with Galois group , each prime of (unramified in ) can be associated to a conjugacy class by the Artin symbol . The analogue of Linnik’s theorem is a bound (in terms of the various number field parameters involved) on the least norm of a prime with prescribed Artin symbol. There have been many works in this direction, both conditional (see [28] and [2]) and unconditional (see [8], [27], [36], [26], [39], and [35]). These unconditional results proceed by establishing analogues of Linnik’s log-free zero-density estimate and quantitative Deuring-Heilbronn phenomenon for the Hecke -functions.
In a different direction, there has been a growing interest in finding new proofs of Linnik’s theorem that avoid using input about the zeros of -functions—see for instance [7], [16], [25], [32], [11], [12], [31], [30], and Chapter 24 in [10]. Often these works combine sieve theoretic techniques with “pretentious methods” and/or with techniques coming from additive combinatorics. While such proofs are not usually (as of yet) as numerically strong as those that use the zeros of -functions, they are very interesting from a conceptual point of view. Similar to the “elementary proof” of the prime number theorem, such proofs show how challenging results in arithmetic can be achieved without (or with minimal) use of results on the zeros of -functions.
In this article, we are interested in a particular analogue of Linnik’s theorem for primes in imaginary quadratic fields, which we prove without using zero-density theorems or exceptional zero repulsion results. Let be a prime, , so that is a negative fundamental discriminant. (We work with prime for simplicity, though we expect that our methods could be adapted without major modifications to work for general fundamental discriminants .) For an integral ideal , we denote by its (absolute) norm. The class group of is a finite abelian group of order , the class number.
In analogy to Dirichlet’s theorem on primes in arithmetic progression, one can show using class group characters that for any specified ideal class , there are infinitely many split primes (i.e., unramified primes with , a rational prime) in the class . Therefore, in analogy to Linnik’s theorem, we are interested in bounding
the least norm of a split prime in the class . Our main result is
Theorem 1.1:
There is an absolute constant such that
and the implied constant is absolute.
This result is a special case of the results for the Chebotarev theorem, so it has been established several times before in those works listed above. In particular, Thorner and Zaman [35] showed
for all negative fundamental discriminants . However, our work is novel in that it is the first time such a result has been established without the use of zero-density theorems and quantitative Deuring-Heilbronn results. We also handle new sieve-theoretic challenges (in comparison with works on Linnik’s theorem) in the sifting dimension aspect; see Section 3 for details.
Remarks:
Ditchen [6] has shown that except for a density zero subset of negative fundamental discriminants , one has . For this result, they establish a large sieve inequality for class group characters on average over discriminants, as well as an analogue of the Bombieri-Vinogradov theorem for primes in ideal classes.
Given the correspondence between imaginary quadratic fields and binary quadratic forms (see [5], for instance), for the principal class , the quantity is the also the least prime of the form when . In our case with , is the least prime of the form . The distribution of such primes has been studied by Fouvry and Iwaniec [9] in connection with low-lying zeros of dihedral -functions.
2. Statement of results
Theorem 1.1 is the result of combining Theorems 2.2 and 2.3 below; we now establish our notations and state these theorems.
Given an ideal class in the class group of , we put
Given a character of the class group, we define
(2) |
the sum being taken over integral ideals of norm . Then by the orthogonality of the class group characters we have
(3) |
For the trivial character we have
(4) |
which is the number of ideals in of norm , and
is the Kronecker symbol. Indeed, is a primitive real Dirichlet character of conductor . By the Dirichlet class number formula, we have
Our main object of study is the sequence
(5) |
a smooth function with , denoting the Fourier transform,
and . We assume that is supported in the segment
where is a small concrete number whose value can be determined in the course of our arguments (though the exact value is not important to us). Here, is a large parameter going to infinity, and our goal is to estimate
By the prime number theorem, we have
where denotes the Mellin transform of ,
One expects prime ideals to equidistribute among the ideal classes in even when the discriminant is comparable in size (in the logarithmic scale) to the norm . Thus we expect the asymptotic formula
to hold for for some absolute constant . Indeed, the Riemann hypothesis for the class group -functions implies that the above asymptotic formula holds with . Here we establish the bound
(6) |
uniformly for for some absolute . (In actuality, we prove a slightly weaker lower bound in the case of an exceptional character; see the precise statement in Theorem 2.2 and the remarks that follow.) In particular, it follows from this lower bound that for all prime , we have
(7) |
In other words, every class contains a prime ideal with .
To establish the bound (6) unconditionally, we split our argument into two cases depending on the non/existence of real zeros of the Dirichlet -function . We will use assumptions about such zeros in several places in the work (and also for other -functions), so to clarify this, we make the following
Definition 2.1:
Let be an -function of conductor (see Appendix A for definitions). For a real number , we say that “Hypothesis ” holds for if every zero of with satisfies
(8) |
Now we state our main two theorems, which together prove Theorem 1.1.
Theorem 2.2:
There exists an absolute constant such that if the Dirichlet -function has a real zero that satisfies
then we have
(9) |
Theorem 2.3:
Let be the constant from the theorem above, and suppose that Hypothesis holds for . Then we have
(10) |
Remarks:
In both theorems above, the implied constants are absolute and effective, though we make no attempt at computing them here. See [13], where they compute an explicit admissible value for the exponent in (7).
We have , so the bound (10) implies (6). On the other hand, the lower bound in (9) is somewhat smaller than the true order of magnitude. This is because in the proof we have used the crude bound that does not involve cancellation in sums over . We would recover the correct order of magnitude if we showed that is well-approximated by the product . This is of course expected to be the case, and we prove it in Appendix A under the complementary assumption that is not exceptional. In any case, the bound (9) still shows that there are many primes with .
3. Outline of the arguments
We follow the general approach of Friedlander and Iwaniec’s proof of Linnik’s theorem in Chapter 24 of [10]—see also their recent related articles [11] and [12]. Our work differs from theirs in a few key aspects, which we now explain.
The goal is to give a positive lower bound for the sum , where for them is the characteristic function of the arithmetic progression , , and for us is defined by (5). They begin with an application of Buchstab’s identity,
(11) |
where denotes the sum of over having no prime factor less than , and denotes the subsequence of over multiples of . Here we take with taken to be as large as necessary.
First they treat the case where there is an exceptional character . In this case, they apply (11) to the “twisted” sequence , where
They show using the Fundamental Lemma of Sieve Theory that the two terms on the right-hand side of (11) are (asymptotically as becomes large) of the same size, up to a factor of
present in the second term. Assuming that is exceptional, is very small, and a positive lower bound for follows. Friedlander and Iwaniec also give in [11] an alternative approach via Selberg’s sieve that works on similar principles and gives comparable results.
We follow their approach in [10] to prove our Theorem 2.2, which is under the assumption of an exceptional character. This is taken up in Section 9. We require little modification of their arguments, since their method does not require very specific properties of the sequence beyond some basic sieve assumptions that also apply in our case. In fact, it is even simpler for us, since we have no need to twist our sequence by the weights above—such a factor naturally appears in this particular sequence already; see Proposition 7.1 for a precise statement.
For the non-exceptional case, Friedlander and Iwaniec work with a combinatorial sieve identity that leads to
(12) |
where
is the lower bound coming from the beta-sieve (so that are the lower-bound beta-sieve weights, and are the congruence sums for the sequence —see Section 5.1 for details), and
where the variables run over specific segments .
Their sifting problem is linear (i.e. of sieve dimension ; again, see 5.1 for details), which means that one can show that is negligible (relatively very small) for close to . The upshot is that they show that
up to some comparably negligible contributions. This reduces the problem of counting primes to finding a lower bound for , which counts products of prime quintuplets in arithmetic progression. Indeed, the common parity here (products of 1 and 5 primes, respectively) is an artifact of the sieve process.
By contrast, in this work we cannot so readily work with (12). The sifting density function for our sequence is given on primes by
and the presence of the character causes fluctuations that hinder one from easily claiming the one-dimensionality of the sieve problem. One possible approach would be to use the fact that we are working in the non-exceptional case (i.e., assuming Hypothesis for , say) to effectively bound the sum and hence control the sieve dimension.
However, here we choose to proceed differently: by the trivial bound , we can work with a -dimensional sieve. We can no longer show that is negligible for so close to (only for ; see Section 5.1), and so we employ a different combinatorial identity than in [12]. This identity comes from applying a second iteration of the Buchstab formula to each term in (11), which gives
(13) |
Rather than work with an inequality, we evaluate (nearly asymptotically) each of the three terms on the right-hand side of (13) and show that the result is positive. The first two terms are readily handled via the Fundamental Lemma, so we reduce the problem to analyzing the third term, which is
(14) |
This sum is our analogue of —note that it is supported on integers which are products of three (almost-) primes, the same parity as in .
Friedlander and Iwaniec handle via a multiplicative analogue of a additive ternary problem treated by the classical circle method. They use Dirichlet characters to decouple the prime variables . After removing the contribution of the principal character (the “major arc”), they use two of the prime variables and the orthogonality of the characters to recover the cost of opening the sum with the characters. The remaining three prime variables are used to obtain a nontrivial cancellation in the character sums over primes—importantly, they do not have need for any zero density bounds or repulsion properties of the exceptional zeros.
We handle the sum (14) in a similar manner, here using the class group characters instead of Dirichlet characters to decouple our variables. Just as above, two variables ( and ) and the orthogonality of the class group characters are used to recover the cost in using these characters. This involves a type of large sieve inequality for these characters (over integers free from small prime factors) that we develop in Section 8. For nontrivial cancellation in a character sum over the final prime variable , we apply the explicit formula and use a zero-free region for the class group -functions. It is a technical reason that we do not use three prime variables for this as they do in [12]. While their sequence is localized dyadically, , ours is supported in a longer segment . This means that we work with a longer sum over the prime variable , which effectively localizes the dual sum over zeros in the explicit formula to essentially be supported on zeros within the classical zero-free region. It is in this way that we do not make use of any zero density estimates or repulsion effects of exceptional zeros.
4. Acknowledgments
This work was completed as part of the author’s PhD thesis. He is deeply grateful to his advisor, Henryk Iwaniec, who provided constant support and guidance throughout this project, and who provided very insightful and helpful feedback during the writing of this article.
5. Preliminaries
5.1. The beta-sieve
For a nonnegative sequence of real numbers we define
for , where . The congruence sums for are
which we will evaluate in the form
(15) |
where is a multiplicative function with for prime , is a smooth approximation to , and is a remainder term that is small (on average over ) in comparison to . The range of the modulus for which (15) holds is called the level of distribution of the sequence .
By the inclusion-exclusion principle, one expects that
(16) |
where
To establish the estimates (16), we use a sequence of sieve weights , which are real numbers supported on squarefree integers satisfying
and we call the level of the sieve. We assume that they satisfy
(17) |
For sieve weights , we put ; that is,
To achieve lower- and upper-bounds as in (16), we use two sets of weights and , called lower- and upper-bound sieve weights. We put
(18) |
and we require that
which implies that
Finally, we say that our sifting problem has dimension at most if
(19) |
for every , for some constant .
While many choices of sieve weights would suffice for our purposes (any that furnish a strong-enough “fundamental lemma” result), for concreteness we will from here on work with a specific construction of sieve weights known as the beta-sieve. These weights were first constructed by Iwaniec [21] and also appear in unpublished work of Rosser. They are of combinatorial type, and they satisfy all of the general properties discussed above, including (17)—see Chapter 11 in [10] for a comprehensive treatment.
The main result we require about the beta-sieve weights is
Proposition 5.1 (see Theorem 11.13 in [10]):
Let be the upper- and lower-bound beta-sieve weights of level . Let be a sequence of nonnegative reals, let be defined by (15), and assume that satisfies (19) with .
Let and put . Define by (18), and put
Then we have
for , and
for , where is a specific absolute constant that depends only on the sifting dimension , and are the continuous solutions to the following system of differential-difference equations,
and and are specific absolute constants that depend only on the sifting dimension . As , we have
We will only apply Proposition 5.1 in the case . In this case we have ; in fact, if . On the other hand, it is nontrivial to compute the values of and for ; see Chapter §11.19 in [10] for a discussion of this and a number of useful inequalities. In particular, they provide a table of numerical values of and for specific values of that were computed by Sara Blight in 2009 and confirmed by Alastair J. Irving in 2014 (who also corrected one value in the table). When , we have
Remark:
Since for the beta-sieve weights, are bounded by
We will always bound the sieve remainder terms absolutely in this work; we have no need to extract additional cancellation from among these terms.
5.2. Class group -functions
Given a character of the class group , we define the associated -function by
the first sum being taken over all nonzero integral ideals of . These functions are entire except in the case that the character is the trivial one, ; in this case, the above -function is the Dedekind zeta function associated to the field ,
where is the Dirichlet -function associated to the Kronecker symbol . Note that is primitive, since is a fundamental discriminant.
The functions
are modular forms of weight 1 for the group with nebentypus . When , they are Hecke eigencuspforms, and in fact they are newforms because the character is primitive. Thus it follows that the coefficients satisfy the Hecke relations
(20) |
A convenient reference for these facts is [22]; see in particular §6.6. The class group -functions are self-dual in the sense that
since for all we have
even though the character need not be real. The completed -functions
satisfy the functional equation (with root number )
(21) |
5.3. The explicit formula and zeros of
By logarithmic differentiation of the Euler products
we get
where (by a slight abuse of notation)
Thus is supported on powers of prime ideals , and is supported on powers of (rational) primes . Note that for a rational prime we have
Using standard arguments (see Theorem 5.11 in [23], for instance), we have
Lemma 5.2:
Let be a smooth, compactly supported function on with Mellin transform . Then we have
(22) |
where the sum over is taken over all nontrivial zeros of .
We will apply this formula with test functions that have the form
(23) |
where is a compactly supported function on . In this case we have
Proposition 5.3:
Let be a smooth function supported on . Putting , suppose that . Then we have
(24) |
where is the class number, is defined as in (23), and runs over all nontrivial zeros of . The implied constant depends only on .
Proof.
Since for , we see that for all , and so the third and fourth terms on the right-hand side of (22) vanish. The rest of the proof follows in a standard way: integrate by parts to estimate the integral, and estimate trivially (using the assumption ) the contribution of prime powers to the left-hand side of (22). ∎
Finally, we record the following result that we will use in Subsection 12.4 to estimate a sum over the zeros of .
Proposition 5.4:
Suppose that the Dirichlet -function satisfies Hypothesis with . Then each of the class group -functions satisfy Hypothesis as well.
Proof.
Zaman [40] has shown that for sufficiently large, the product of -functions has at most one zero in the region
(25) |
and that if such a zero exists, then both it and the associated class group character are real. By the genus theory, the only real class group character is the trivial character because is prime, and in this case the Kronecker factorization theorem gives us
Assuming now that satisfies Hypothesis , we deduce from the above that each does as well (assuming that ). ∎
Remarks:
An explicit value for as in (25) is not necessary for our result in this article. Without an explicit value of the constant, the above result is due to Fogels [8]. Additionally, in [40] they give many other explicit results for more general Hecke -functions. See also [24] and [1].
The above proof is precisely the moment where we make use of the fact that is prime. To work with general fundamental discriminants , one may adjust the hypothesis of Proposition 5.4 to read “Suppose that for every divisor , the Dirichlet -function satisfies Hypothesis with ,” and the conclusion of the Proposition would still be true. In this case, one would have to correspondingly prove a version of Theorem 2.2 with a different hypothesis (i.e., “There exists a constant and a divisor such that if has a real zero …”), which we have chosen not to do here for the sake of a cleaner exposition.
6. The congruence sums
In this section, we consider the sequence defined in (5) and evaluate the associated congruence sums,
Expressing in terms of by (3), this is accomplished via
Proposition 6.1:
Let be a class group character. Let be a smooth function supported on , and let for some . Then
where is the multiplicative function given by
(26) |
and if , and otherwise.
Remark:
The evaluation of the congruence sums follows directly from the proposition above using (3). In the sieve terminology, this shows that the sequence has level of distribution .
To prove Proposition 6.1, we use Lemma 6.2, a summation formula for the harmonics (see for instance (5.16) in [23]). We omit its proof, which is standard—it follows essentially from the functional equation (21).
Lemma 6.2:
Let be a class group character. Then for a smooth, compactly supported function on with Mellin transform , we have
where if the trivial character, if , and
(27) |
with .
Next, we establish a version of the above formula in the logarithmic scale.
Lemma 6.3:
Let be a smooth function with support contained in . Then for any , we have
for any , where the implied constant depends only on and .
Proof.
We apply Lemma 6.2 with the choice
It is straightforward to verify that
and by partial integration times we derive
where , and is at our disposal. By Stirling’s formula, we have
Now we estimate the function given by (27): we move the line of integration to and take to get
Now , so as long as , we get
for any given by taking sufficiently large. ∎
Note that the main term in the above lemma does not depend on , which is a convenient feature in the forthcoming transformations. Using the lemma above, we prove Proposition 6.1.
Proof of Proposition 6.1.
We write and use the Hecke relations (20), and then we put to get
(28) | ||||
Next we split the -sum according to whether or , with to be chosen later. For the latter range where , we estimate the -sum trivially using , which gives us
whose contribution to (28) is
(29) |
In the other range where , we apply Proposition 6.3, which is applicable as long as , which we will arrange for with our later choice of . This gives
Plugging the above into (28) essentially gives the expression for in (26), except that the -sum here is restricted to . The range is easily extended to all up to the same error term in (29). Thus in total we have now shown
Finally, choosing gives the result. ∎
7. Sums of twisted by sieve weights
In this section, let , and let be the beta-sieve weights (upper- or lower-bound) of level ; put .
Proposition 7.1:
Proof.
We have
Applying Proposition 6.1 (with ) for the -sum on the right-hand side then shows that when , the above is bounded by (recall that )
(31) |
after using the bound . For , we observe that
is the sifted sum for the sequence given by
By Proposition 6.1, the congruence sums for this sequence are
where
and is given by (26). We have
(32) |
so we can take in (19) for some . Applying Proposition 5.1 gives
Just as in (31), the remainder terms are each bounded by
which is covered by , and the proof is complete. ∎
Here we state a corollary of the above proposition that is ready-to-use for the applications below. From here on we take , , and .
Corollary 7.2:
Let , and let be a smooth function supported on . Let be the beta-sieve weights of level . Suppose that
Then for sufficiently large we have
where and are given by (30). Furthermore, if Hypothesis holds for , then we can replace in the above with
8. A large sieve-type inequality for over almost-primes
In this section we give a type of large sieve inequality for the harmonics where runs over integers with no small prime factors. General large sieve inequalities for Hecke characters in number fields were given by Huxley [20] and Schaal [33]; results of the same analytic strength but with explicit constants were later established by Schumer [34].
However, these results are not sufficient for our particular application here. The above results (after specializing to our case) imply a bound
(33) |
for any complex numbers satisfying , as long as . In our case, the variable has no small prime factors, say . Rather than applying the bound (33) by restriction of the coefficients , we apply a sieve that allows us to gain two factors of , which is crucial for our application here. Our precise result is
Proposition 8.1:
Put and , and suppose that Hypothesis holds for . Then as long as and , we have
(34) |
where are any complex numbers satisfying .
Proof.
Let . Let be a smooth function supported on such that for , and let be the upper-bound beta-sieve weights of level .
Expanding the square and rearranging, the left-hand side of (34) is
Using the definition (2) of , the above sum over is equal to
where means and are in the same ideal class. In particular, the above sum is real and nonnegative. Therefore, taking absolute values, we have
The right-hand side above is majorized by
and so we have shown that
(35) |
Next we apply Corollary 7.2 to evaluate the -sum on the right-hand side of the above, which gives
as long as . Plugging this into (35), we have shown
We have , and the term is superfluous since we assume . This completes the proof. ∎
9. The exceptional case: proof of Theorem 2.2
Proof of Theorem 2.2.
We begin by applying the Buchstab formula for the sequence given by (5). This gives
(36) |
where , with to be chosen later. Let be the lower-bound beta-sieve weights of level , and put . The orthogonality (3) of the class group characters implies
Now we apply Corollary 7.2 to get
where and are given by (30), as long as
(37) |
Now we consider the second term in (36). For each , we use the upper bound
We have
(38) |
the error term appearing from the terms with . Now we let be the upper-bound beta-sieve weights supported on , putting . Expanding in terms of the characters , we have
We remove the condition up to the same error term in (38). Note that is supported on for any since . So Corollary 7.2 gives
as long as
(39) |
The second error term above is subsumed by the first after summing over , and hence we get
where we have put
One can show (see for instance §24.2 of [10]) that
where is any real zero of . Assuming that has a real zero , we get
Putting everything together, we have shown that
We choose values for the parameters, each in terms of . We take
With these choices, we see that both (37) and (39) hold for sufficiently large, and that . Finally, by (32) we have
and hence
which completes the proof. ∎
10. The non-exceptional case: proof of Theorem 2.3
In this section we prove Theorem 2.3 under the assumption of the three propositions below, whose proofs we give in the final section. Throughout this entire section and the following two, we assume that Hypothesis holds for .
We begin by applying the Buchstab formula
where , with to be chosen later. (Note that this is independent from the one in the previous section, and may be chosen to have a different numerical value.) We then apply a second iteration of the Buchstab formula to each term , which gives
(40) |
where we have put
and
We evaluate each term in the following three propositions.
Proposition 10.1:
With and , suppose that
(41) |
Then we have
(42) |
where the implied constant is absolute.
Proposition 10.2:
With and , suppose that
(43) |
Then we have
(44) |
where the implied constant is absolute.
Proposition 10.3:
Take and ; suppose and . Let be such that
(45) |
Then we have
(46) |
where is an explicit quantity defined in Section 12, and the implied constant is absolute. Importantly, does not depend on the ideal class in the definition of the sequence .
Assuming these propositions for now, we prove Theorem 2.3.
Proof of Theorem 2.3.
The sequence defined by
depends on the ideal class , so for the moment we write to emphasize this dependence. We now define the sequence by
the second inequality following from the orthogonality (3) of the class group characters. Applying (40) to and and taking the difference, we get
For each we have . Since each of the right-hand sides of (42), (44), and (46) does not depend on the class , it follows that each of those equations holds with replaced by . Therefore we get
Now we choose our parameters, each in terms of . (NB: the and here are chosen independently from those in Section 9, and here is fixed.) We take
(47) |
With these choices, one verifies that the conditions (41), (43), and (45) are verified for sufficiently large, and that we have
On the other hand, from (4) we have
From the prime number theorem we have
and from (76) (which assumes Hypothesis holds for ) we have
Therefore for our choices of parameters (47), we get
From (40) this implies for sufficiently large, which completes the proof. ∎
11. Proofs of Propositions 10.1 and 10.2
11.1. Evaluating
11.2. Evaluating
Proof of Proposition 10.2.
We evaluate the negative of ,
To evaluate the terms
we first attach sieve weights, putting
where , and are upper- and lower-bound beta-sieve weights of level . Next, using (3) (note that , since ), we have
where we have put
To use the multiplicativity of , we first remove the terms where . Such terms contribute to the above sum at most
and hence we get
Now for the -sum, we apply Corollary 7.2 with , where . Note that is supported on , which accounts for the condition (43). Following the same lines as in the proof of Proposition 10.1, we get
Next, we sum over . The contribution of the second -term above is at most
which is negligible. For the main term, we evaluate
which follows from Mertens’ theorem and Corollary A.2 (using that Hypothesis holds for ). Putting everything together completes the proof. ∎
12. Proof of Proposition 10.3
In this section we prove Proposition 10.3, where we evaluate the sum
(48) |
Proof of Proposition 10.3.
We define the quantity
where are defined in (56). From the definitions of in Section 12.2, it is apparent that does not depend on the ideal class in the definition of the sequence . Combining the results of Lemmas 12.1, 12.2, 12.3, and 12.6 below, we see that (46) holds, subject to the conditions (45). ∎
12.1. First arrangements
We separate from (48) the terms where either or , putting
where
gives the main contribution, and we will show that
give lesser contributions to . First we estimate : using , we have
Write with or . In either case, the -sum above is bounded by
(49) |
and summing over shows that the same bound holds for .
For , we have
The support of forces , hence , and the same for . We now open using the class group characters, getting
Now we apply Proposition 8.1: we choose the coefficients appropriately so that the summation
is supported on prime . Then as long as
the bound (34) gives
Thus we have now established
Lemma 12.1:
As long as and , we have
Remarks:
Here we have used our large sieve-type inequality (Proposition 8.1) to not lose the factor (or any logarithmic factors) after expanding via the class group characters .
Lemma 12.1 indicates that the sum does in fact contribute to a positive proportion of the sum . However, this proportion is (so to speak) of a lower order of magnitude (proportional to ) than the full sum (proportional to ), which is due to the short range of the variables .
12.2. A smooth decoupling
For the remaining terms from , we have , hence
and so . We make a smooth partition of the variable into segments that are geometric in the logarithmic scale. First, we partition the range
into segments with , where are given by
so is at our disposal, and it determines . Now we make a smooth partition from these points in the following way:
-
•
we now let the index run over half-integers ;
-
•
for each such , let be a smooth bump function supported on , such that for we have
We put
and from the above properties we get
(50) | ||||
We now use this smooth partition of unity to decouple the variables . From (50) we have
The conditions
(51) |
entangle and with , so we adjust them to decouple these variables. The variable lies in the restricted range by the support of , and so (by positivity) we replace the three conditions (51) respectively by
in the lower bound for , and with
in the upper bound for . After these adjustments we have
(52) | ||||
(53) |
Now we open using characters: by (5) and (3) we have
which we put into (52) and (53). By our adjustments above, we have arranged that are automatically pairwise coprime, so the multiplicativity of now gives
and
Finally, we wish to remove the conditions now that we have made use of them to decouple . To do so, we add back the missing terms, which are bounded by the same error term in (49) from before. Putting
we have shown the following
Lemma 12.2:
We have
(54) | ||||
(55) |
12.3. The contribution of the principal character
Now we extract from (54) and (55) the contribution from the principal character , which constitutes the main part of . Accordingly we put
(56) |
and we will show that the difference
is comparably small.
Lemma 12.3:
Let , and suppose that , , and
(57) |
Then we have
Remark:
The variable plays no essential theoretical role, but it must be present for technical reasons. On a mechanical level, it is a parameter that can be taken larger to ensure that Corollary 7.2 is applicable even when the range of the involved summation includes (relatively) very small integers.
To prove this lemma, we will use the following couple of estimates.
Lemma 12.4:
Let , and suppose that satisfies Hypothesis . Then we have
Proof.
We have , so the result follows directly from Mertens’ theorem and Corollary A.2. ∎
Lemma 12.5:
Let . Let , and suppose that
(58) |
Then for any , we have
(59) |
Proof.
Proof of Lemma 12.3.
In the following, we will apply Lemma 12.5 in situations where the variable always satisfies , i.e., . Therefore we assume throughout that is chosen so that the condition (57) holds. This ensures that (58) holds any time we apply Lemma 12.5.
First we handle . For these terms we have , or
as well as , or
(60) |
Using and the above bounds, we get
(61) |
Assuming that , , and , we replace (by positivity) the inequalities (60) and (61) by the simpler ones
Now applying Lemmas 12.4 and 12.5, we get (using )
(62) |
Next we analyze . We have , so for these terms we have
The same lower bounds in (60) and (61) hold, and hence gives
We apply Lemmas 12.4 and 12.5 again, and we find that satisfies the same bound as ,
(63) |
The differences are more complicated, but using positivity we can majorize them by two simpler sums,
where
and
and the set is given by the difference of sets
For , the condition implies
Applying Lemmas 12.4 and 12.5 then gives
(64) |
For , we observe that if and , then must be prime, since (assuming ). Otherwise, the elements of are , where , , and . In other words,
Therefore, using , we get
Using the condition , we apply Lemmas 12.4 and 12.5 to get
which we combine with (64) and sum over to get
Now we make a choice for the parameter : we take
(65) |
where denotes the integer part. This determines via the relation
and hence our choice of implies
(66) |
Therefore we have
Similarly, from (62) and (63) and the bound (66), we see that both and satisfy the same bound. This gives the result. ∎
12.4. The contribution of the other characters
In this section we estimate the contributions of the nonprincipal characters to the lower and upper bounds (54) and (55).
Lemma 12.6:
Proof.
We treat both at the same time because our arguments do not depend on the specific ranges of the variables and , only that the inequalities
hold for every . First, we evaluate the sum over using the explicit formula. Specifically, we apply Proposition 5.3 with the choice
which gives us
where
and the sum runs over the nontrivial zeros of . (We don’t write the conditions for explicitly, since they differ for , but of course we keep them in our minds as necessary.) Note that there is no polar contribution as in (24) since here we treat the nonprincipal characters . We decouple the variables from the above integral by applying the Fourier inversion
which gives us
where
Summing over and taking the absolute value, we get
where we have put
(67) |
and
(68) |
To estimate , we integrate by parts three times (after borrowing/returning a factor ) to get
For we have
for every , with , and thus we get
which holds for all . Using this to estimate the above integral, we derive
We use this bound to estimate (67). The number of zeros of with and is , so the above bound implies
By Proposition 5.4, the remaining zeros of fall in the range of Hypothesis , so they satisfy
and hence . Applying Lemma A.3 now gives
(69) |
This bound covers the one above for , and it is uniform in , so we deduce that is also bounded by the right-hand side of (69).
For , we apply the Cauchy inequality to the -sum in (68) to get
(70) | ||||
Here we have (by positivity) added back the principal character to the -sums. We apply Proposition 8.1: with chosen so that (57) holds, we have
where we choose the coefficients to agree with those of in (70) when , and we choose them to be 0 otherwise. This gives
For the integration over , we have
The derivatives are supported on and satisfy
and hence we derive
Therefore
Finally we sum over , which gives
From (65) we have , which completes the proof. ∎
Appendix A Approximating -functions by finite Euler products
Let be an entire -function of degree , where we think of as some interesting arithmetic object to which is attached. It is natural to try to approximate by a partial Euler product. This question and similar ones have been addressed by many works in the literature—see for instance [18], [17], [15], [3], [4], and [14], where such approximations are both developed and used for interesting arithmetic applications.
The main result in this section, Proposition A.1, would follow from results in the cited works above (for , at least, from a slight modification of the results in [3]) after using an appropriate zero-density estimate for . However, we wish to give here a self-contained proof of the result that does not rely on any zero-density estimates.
We assume that is given by a Dirichlet series and Euler product,
if , where . Further we assume that is entire and that it satisfies a functional equation of conductor , where
is its gamma factor. This means (see Chapter 5 of [23])
where denotes the root number, .
Our goal is to approximate by the finite product
(71) |
when comparable to in the logarithmic scale. Indeed, assuming the Riemann hypothesis for shows that
as . By comparison, our result will be unconditional. We do not require any zero-density estimate for , only that it have a zero-free region of “classical” type; that is, we assume that Hypothesis holds for .
Proposition A.1:
Suppose is entire and that it satisfies Hypothesis . Let for some . Then
(72) |
where
(73) |
The implied constant above depends on the degree and parameters .
In this article, we use the above result only for the quadratic Dirichlet -function . In this case the result reads: assuming Hypothesis holds for , then for we have
(74) |
Before giving the proof of Proposition A.1, we give a corollary that we need at other points in the work.
Corollary A.2:
Assume Hypothesis holds for , and that where . Let . Then for sufficiently large we have
(75) |
If is a smooth function supported on , then we have
(76) |
Proof of Corollary A.2.
To prove Proposition A.1, we begin with
Lemma A.3:
Let denote a zero of . Then for all ,
where the sum is taken over nontrivial zeros of , , and the implied constant depends only on the gamma factor parameters .
Proof of Lemma A.3.
For we have
with supported on prime powers,
Hence and
(77) |
On the other hand the Hadamard product yields
Note the contribution of the trivial zeros to the above sum is . For the constant , we have
Therefore for with we have
(78) |
where the summation now runs over all nontrivial zeros of . With and we have
(79) |
Thus now we choose (where ) and combine (77), (78), and (79) to prove the lemma. ∎
Proof of Proposition A.1.
We have
where is defined in (71), and
We put
Note that this sum converges by virtue of Hypothesis —it does not converge absolutely. We then check that
and therefore we have
where the implied constant depends on . To estimate , we put
where is a smooth function satisfying if , if , and if , with . Then we have
where again the implied constant depends on . Thus now we have
(80) |
To estimate , we apply the explicit formula for ,
(81) |
where is the Mellin transform of
We have
(82) |
by partial integration three times. We can choose so that for ,
and for we have
Using these bounds in (82), we get
(83) |
We also have the bound
(84) |
which is derived in a similar way, using the identity
and the bound
Recall that the number of zeros with and is for all . Using this with (84), we see that the contribution of with to the explicit formula (81) is bounded by
(85) |
The remaining zeros satisfy (8), so for these zeros we have
(86) |
Noting that
we apply Lemma A.3 to see that the contribution of these zeros to (81) is
(87) |
by (83) and (86). The bound above is larger than the one in (85), and hence is also bounded by the right-hand side of (87). Combining this bound with (80) then shows
Finally we choose , which gives (73). ∎
References
- [1] J.-H. Ahn and S.-H. Kwon “Some explicit zero-free regions for Hecke -functions” In J. Number Theory 145 Elsevier, 2014, pp. 433–473
- [2] E. Bach and J. Sorenson “Explicit bounds for primes in residue classes” In Math. Comp. 65.216, 1996, pp. 1717–1735
- [3] H.. Bui and J.. Keating “On the mean values of Dirichlet -functions” In Proc. London Math. Soc. 95.2 Oxford University Press, 2007, pp. 273–298
- [4] H.M. Bui and J.P. Keating “On the mean values of -functions in orthogonal and symplectic families” In Proc. London Math. Soc. 96.2 Oxford University Press, 2008, pp. 335–366
- [5] D.. Cox “Primes of the Form : Fermat, Class Field Theory, and Complex Multiplication” Amer. Math. Soc., 2022
- [6] J. Ditchen “Primes of the shape : the distribution on average and prime number races”, 2013
- [7] P…. Elliot “The least prime primitive root and Linnik’s theorem” In Number Theory for the Millenium I, 2002, pp. 393–418
- [8] E. Fogels “On the distribution of prime ideals” In Acta Arith. 7, 1962, pp. 255–269
- [9] É. Fouvry and H. Iwaniec “Low-lying zeros of dihedral -functions” In Duke Math. J. 116, 2013, pp. 189–217
- [10] J.. Friedlander and H. Iwaniec “Opera de Cribro” Amer. Math. Soc., 2010
- [11] J.. Friedlander and H. Iwaniec “Selberg’s sieve of irregular density” In Acta Arith. 209, 2023, pp. 385–396
- [12] J.. Friedlander and H. Iwaniec “Sifting for small primes from an arithmetic progression” In Sci. China Math. 66.12, 2023, pp. 2715–2730
- [13] L.. Gaudet “On the least prime represented by a positive-definite binary quadratic form”, 2023
- [14] S.. Gonek “Finite Euler products and the Riemann hypothesis” In Trans. Amer. Math. Soc. 364.4, 2012, pp. 2157–2191
- [15] S.. Gonek, C.. Hughes and J.. Keating “A hybrid Euler-Hadamard product for the Riemann zeta function” In Duke Math. J. 136.3, 2007
- [16] A. Granville, A. Harper and K. Soundararajan “A new proof of Hálasz’s theorem, and its consequences” In Compositio Math. 155.1 London Mathematical Society, 2019, pp. 126–163
- [17] A. Granville and K. Soundararajan “Extreme values of ” In The Riemann zeta function and related themes: papers in honour of Professor K. Ramachandra, Ramanujan Math. Soc. Lect. Notes Ser., vol. 2 Ramanujan Math. Soc., Mysore, 2006, pp. 65–80
- [18] A. Granville and K. Soundararajan “The distribution of values of ” In Geom. Func. Anal. 5.13, 2003, pp. 992–1028
- [19] D.. Heath-Brown “Zero-free regions for Dirichlet -functions, and the least prime in an arithmetic progression” In Proc. London Math. Soc. 64, 1992, pp. 265–338
- [20] M.. Huxley “The large sieve inequality for algebraic number fields” In Mathematika 15.2 London Mathematical Society, 1968, pp. 178–187
- [21] H. Iwaniec “Rosser’s sieve” In Acta Arith. 36.2, 1980, pp. 171–202
- [22] H. Iwaniec “Topics in Classical Automorphic Forms” Amer. Math. Soc., 1997
- [23] H. Iwaniec and E. Kowalski “Analytic Number Theory” Amer. Math. Soc., 2004
- [24] H. Kadiri “Explicit zero-free regions for Dedekind zeta functions” In Int. J. Number Theory 8.1 World Scientific, 2012, pp. 125–147
- [25] D. Koukoulopoulos “The Distribution of Prime Numbers” Amer. Math. Soc., 2019
- [26] E. Kowalski and P. Michel “Zeros of families of automorphic -functions close to 1” In Pacific J. Math. 207.2, 2002, pp. 411–431
- [27] J.. Lagarias, H.. Montgomery and A.. Odlyzko “A bound for the least prime ideal in the Chebotarev density theorem” In Invent. Math. 54.3, 1979, pp. 271–296
- [28] J.. Lagarias and A.. Odlyzko “Effective versions of the Chebotarev density theorem” In Algebraic number fields: -functions and Galois properties (Proc. Sympos., Univ. Durham, Durham, 1975) 7, 1977, pp. 409–464
- [29] Y.. Linnik “On the least prime in an arithmetic progression. I. The basic theorem” In Rec. Math. [Mat. Sbornik] N.S. 15.2, 1944, pp. 139–178
- [30] K. Matomäki, J. Merikoski and J. Teräväinen “Primes in arithmetic progressions and short intervals without -functions” In arXiv:2401.17570 [math.NT], 2024
- [31] K. Matomäki and J. Teräväinen “Products of primes in arithmetic progressions” In J. reine angew. Math. 2024.808, 2024, pp. 193–240
- [32] S. Sachpazis “A pretentious proof of Linnik’s estimate for primes in arithmetic progressions” In Mathematika 69.3, 2023, pp. 879–902
- [33] W. Schaal “On the large sieve method in algebraic number fields” In J. Number Theory 2.3 Elsevier, 1970, pp. 249–270
- [34] P.. Schumer “On the large sieve inequality in an algebraic number field” In Mathematika 33.1 London Mathematical Society, 1986, pp. 31–54
- [35] J. Thorner and A. Zaman “An explicit bound for the least prime ideal in the Chebotarev density theorem” In Algebra & Number Theory 11.5, 2017, pp. 1135–1197
- [36] A. Weiss “The least prime ideal” In J. reine angew. Math. 338, 1983, pp. 56–94
- [37] T. Xylouris “On Linnik’s constant” In Acta Arith. 1.150, 2011, pp. 65–91
- [38] A. Zaman “Analytic estimates for the Chebotarev Density Theorem and their applications”, 2017
- [39] A. Zaman “Bounding the least prime ideal in the Chebotarev density theorem” In Funct. Approx. Comment. Math. 57.1, 2017, pp. 115–142
- [40] A. Zaman “Explicit estimates for the zeros of Hecke -functions” In J. Number Theory 162 Elsevier, 2016, pp. 312–375