Sparse Variance for primes in arithmetic progression

advertisement
SPARSE VARIANCE FOR PRIMES IN
ARITHMETIC PROGRESSION
J. BRÜDERN AND T. D. WOOLEY∗
Abstract. An analogue of the Montgomery-Hooley asymptotic formula is established
for the variance of the number of primes in arithmetic progressions, in which the moduli
are restricted to the values of a polynomial.
1. Introduction
The distribution of primes is a fundamental problem in the theory of numbers. Our
concern here is with the error terms that arise in the prime number theorem for arithmetic
progressions. Let πœ™ denote Euler’s totient, and define the relevant quantity 𝐸(π‘₯; π‘˜, 𝑙) for
real π‘₯ ≥ 2, and coprime natural numbers π‘˜ and 𝑙, through the equation
∑
π‘₯
+ 𝐸(π‘₯; π‘˜, 𝑙).
log 𝑝 =
πœ™(π‘˜)
𝑝≤π‘₯
𝑝≡𝑙 mod π‘˜
Montgomery [8] obtained an asymptotic formula for the variance
𝑉 (π‘₯, 𝑦) =
π‘˜
∑ ∑
π‘˜≤𝑦
∣𝐸(π‘₯; π‘˜, 𝑙)∣2
𝑙=1
(𝑙,π‘˜)=1
that was promptly refined by Hooley [5], and now takes the shape
𝑉 (π‘₯, 𝑦) = π‘₯𝑦 log 𝑦 + 𝑐π‘₯𝑦 + 𝑂(π‘₯1/2 𝑦 3/2 + π‘₯2 (log π‘₯)−𝐴 ).
(1)
Here, the coefficient 𝑐 is a certain real constant, the exponent 𝐴 is a preassigned positive
real number, and the relation (1) holds uniformly for 1 ≤ 𝑦 ≤ π‘₯. In this range, the
asymptotic formula (1) implies the upper bound
𝑉 (π‘₯, 𝑦) β‰ͺ π‘₯𝑦 log π‘₯ + π‘₯2 (log π‘₯)−𝐴
(2)
which is known as the Barban-Davenport-Halberstam theorem. If one singles out an
individual progression 𝑙 modulo π‘˜ from the mean 𝑉 (π‘₯, 𝑦), then one recovers from the
bound (2) the Siegel-Walfisz theorem, the latter asserting that the estimate
𝐸(π‘₯; π‘˜, 𝑙) β‰ͺ π‘₯(log π‘₯)−𝐴/2
1991 Mathematics Subject Classification. 11N13, 11P55.
Key words and phrases. Primes in progressions, Barban-Davenport-Halberstam theorem.
∗
The authors thank Shandong University for providing a creative atmosphere during the period while
this paper was conceived, and the Hausdorff Research Institute for very good working conditions while
the paper was completed. The second author is supported in part by a Royal Society Wolfson Research
Merit Award.
1
2
J. BRÜDERN AND T. D. WOOLEY
holds uniformly in π‘˜ and 𝑙. Although it seems difficult to improve this last bound, the
asymptotic formula for 𝑉 (π‘₯, 𝑦) shows that in mean square, the average size of 𝐸(π‘₯; π‘˜, 𝑙)
is about the square-root of (π‘₯ log π‘₯)/π‘˜.
One might ask whether conclusions similar to those described above are possible for
thinner averages, for example with the moduli π‘˜ restricted to the values of a polynomial.
This was studied recently by Mikawa and Peneva [7], who obtained a result analogous
to the Barban-Davenport-Halberstam theorem. More precisely, let 𝑓 ∈ β„š[𝑑] denote an
integer-valued polynomial of degree 𝑑 ≥ 2, with positive leading coefficient and derivative
𝑓 ′ . Then, there exist real numbers 𝑦1 such that the inequalities 𝑓 (𝑦) ≥ 2 and 𝑓 ′ (𝑦) ≥ 1
hold for all 𝑦 ≥ 𝑦1 . Denote the smallest such 𝑦1 with 𝑦1 ≥ 1 by 𝑦0 (𝑓 ). We may now define
𝑉𝑓 (π‘₯, 𝑦) =
∑
𝑦0 (𝑓 )<π‘˜≤𝑦
′
𝑓 (π‘˜)
𝑓 (π‘˜)
∑
∣𝐸(π‘₯; 𝑓 (π‘˜), 𝑙)∣2 .
(3)
𝑙=1
(𝑙,𝑓 (π‘˜))=1
In this notation, the aforementioned estimate of Mikawa and Peneva asserts that whenever
𝐴 ≥ 1 is a real number, then there is a 𝐡 ≥ 1 such that, uniformly for 𝑓 (𝑦) ≤ π‘₯(log π‘₯)−𝐡 ,
one has
𝑉𝑓 (π‘₯, 𝑦) β‰ͺ π‘₯2 (log π‘₯)−𝐴 .
(4)
′
We note here that Theorem 2 of [7] relates to a variant of 𝑉𝑓 (π‘₯, 𝑦), with the weight 𝑓 (π‘˜)
in (3) replaced by π‘˜ −1 πœ™(𝑓 (π‘˜)). However, as is easily checked, the bound (4) above is
equivalent to the conclusion in [7]. Our purpose in this paper is to establish a version of
the Montgomery-Hooley asymptotic formula for 𝑉𝑓 (π‘₯, 𝑦).
Theorem. Let 𝑓 be an integer-valued polynomial of degree 𝑑 ≥ 2, with positive leading
coefficient. Then, whenever 𝑓 (𝑦) ≤ π‘₯, one has
(
)
𝑉𝑓 (π‘₯, 𝑦) = π‘₯𝑓 (𝑦) log 𝑓 (𝑦) + 𝐢(𝑓 )π‘₯𝑓 (𝑦) + 𝑂 π‘₯1/2 𝑓 (𝑦)3/2 + π‘₯2 (log π‘₯)−𝐴 ,
wherein 𝐢(𝑓 ) denotes a certain real number depending only on 𝑓 .
A new approach1 to the Montgomery-Hooley theorem was developed by Goldston and
Vaughan [3] and Vaughan [10]. In contrast with previous work that depended on the
additive theory of primes, the evaluation of 𝑉 (π‘₯, 𝑦) is now linked to the problem of finding
an asymptotic formula for the number of solutions of the equation 𝑝1 − 𝑝2 = β„Žπ‘˜ in primes
𝑝1 , 𝑝2 not exceeding π‘₯, and natural numbers β„Ž and π‘˜ with π‘˜ ≤ 𝑦. This may be viewed
as a ternary additive problem that is well within the competence of the circle method.
Moreover, the primes enter the scene only through an application of the Siegel-Walfisz
theorem. Vinogradov’s estimate for exponential sums over primes is not required since
the minor arc analysis can be performed primarily with the product β„Žπ‘˜. Following these
patterns, we are led to consider the equation 𝑝1 − 𝑝2 = β„Žπ‘“ (π‘˜), and again, a treatment by
the circle method is possible. A detailed discussion of this equation is to be found in §§3
and 4, following a preparatory analysis in the next section. Reflecting a similar problem
in the work of Vaughan [10], the treatment of the major arcs invokes certain auxiliary
averages of multiplicative functions that are supplied in §§5 and 6, thereby completing
the proof of the theorem.
There is an extensive literature relating to improvements of the error term in the asymptotic formula (1) that are subject to the truth of the Riemann hypothesis for Dirichlet
1For
historical comments on the genesis of this method, the curious reader is directed to p. 118 of [3].
SPARSE VARIANCE FOR PRIMES IN ARITHMETIC PROGRESSION
3
𝐿-functions (see, in particular, Hooley [5], [6], Friedlander and Goldston [2], Goldston and
Vaughan [3] and Vaughan [11]). Likewise, in the conclusion of our theorem above, the
term π‘₯2 (log π‘₯)−𝐴 may be replaced conditionally by π‘₯2−𝛿(𝑑) , where 𝛿(𝑑) is a positive real
number the size of which is determined by the maximal savings available for Weyl sums
involving a degree 𝑑 polynomial. We spare the reader any details.
Our methods are rather flexible. One may replace the primes with other sequences of
interest, and functions can be substituted for the polynomial 𝑓 with growth more rapid
than any polynomial. For example, by working along the lines of Brüdern and Perelli
[1], it should be possible to deduce an asymptotic formula for 𝑉𝑓 (π‘₯, 𝑦) when 𝑓 (π‘˜) =
[ exp((log π‘˜)𝛾 )], and 𝛾 is any positive number smaller than 3/2.
2. Two preparatory steps
Our initial advance on the analysis of the variance 𝑉𝑓 (π‘₯, 𝑦) involves the isolation of a
term accessible to the circle method, and this is the objective of this section. Fix the
value of 𝐴 ≥ 1 for the rest of the paper, and choose a permissible value of 𝐡 = 𝐡(𝐴) so
that the asymptotic relation (4) holds uniformly for 𝑓 (𝑦) ≤ π‘₯(log π‘₯)−𝐡 . An inspection of
[7] reveals that one may take 𝐡 = 24𝑑 𝐴. When π‘₯ is sufficiently large, there is a unique
solution 𝑦 to the equation 𝑓 (𝑦) = π‘₯(log π‘₯)−𝐡 that we denote by 𝑦1 = 𝑦1 (π‘₯). As a truncated
analogue of the sum 𝑉𝑓 (π‘₯, 𝑦) defined in (3), we define
𝑉𝑓′ (π‘₯, 𝑦)
=
∑
′
𝑓 (π‘˜)
𝑦1 <π‘˜≤𝑦
𝑓 (π‘˜)
∑
∣𝐸(π‘₯; 𝑓 (π‘˜), 𝑙)∣2 .
(5)
𝑙=1
(𝑙,𝑓 (π‘˜))=1
Then in view of the definition (3) and the estimate (4), it is easily verified that
𝑉𝑓 (π‘₯, 𝑦) = 𝑉𝑓′ (π‘₯, 𝑦) + 𝑂(π‘₯2 (log π‘₯)−𝐴 ).
(6)
We have removed small values of π‘˜ from 𝑉𝑓 (π‘₯, 𝑦) mainly for technical convenience. To
do so, we had to invoke the work of Mikawa and Peneva [7], but it would be possible,
and perhaps more coherent, to cover small π‘˜ by our method. However, a critical step for
small π‘˜ would depend on the ideas of section 2.2 in [7], and that we would have need to
duplicate in any case. In the interest of brevity, we therefore prefer to work temporarily
with 𝑉𝑓′ (π‘₯, 𝑦).
In the next step, we open the square in (5). Let
∑
πœ—(π‘₯; π‘˜, 𝑙) =
log 𝑝
𝑝≤π‘₯
𝑝≡𝑙 mod π‘˜
and
Φ𝑓 (𝑧, 𝑦) =
𝑓 ′ (π‘˜)
.
πœ™(𝑓
(π‘˜))
𝑧<π‘˜≤𝑦
∑
(7)
Then it follows from (5) that
𝑉𝑓′ (π‘₯, 𝑦) = 𝑆1 − 2π‘₯𝑆2 + π‘₯2 Φ𝑓 (𝑦1 , 𝑦),
(8)
4
J. BRÜDERN AND T. D. WOOLEY
where
𝑆1 =
∑
𝑓 (π‘˜)
𝑦1 <π‘˜≤𝑦
𝑆2 =
𝑓 (π‘˜)
∑
′
𝑙=1
(𝑙,𝑓 (π‘˜))=1
𝑓 ′ (π‘˜)
πœ™(𝑓 (π‘˜))
<π‘˜≤𝑦
𝑓 (π‘˜)
∑
∑
𝑦1
πœ—(π‘₯; 𝑓 (π‘˜), 𝑙)2 ,
πœ—(π‘₯; 𝑓 (π‘˜), 𝑙).
𝑙=1
(𝑙,𝑓 (π‘˜))=1
By the Prime Number Theorem, one finds that
𝑓 (π‘˜)
∑
∑
πœ—(π‘₯; 𝑓 (π‘˜), 𝑙) =
log 𝑝 = π‘₯ + 𝑂(π‘₯(log π‘₯)−3𝐴 ).
𝑝≤π‘₯
π‘βˆ€π‘“ (π‘˜)
𝑙=1
(𝑙,𝑓 (π‘˜))=1
Consequently, using only straightforward estimates, one discovers that whenever 𝑓 (𝑦) ≤ π‘₯,
one has
𝑆2 = π‘₯Φ𝑓 (𝑦1 , 𝑦) + 𝑂(π‘₯(log π‘₯)2−3𝐴 ).
(9)
Similarly, one may derive the relation
𝑓 (π‘˜)
∑
πœ—(π‘₯; 𝑓 (π‘˜), 𝑙)2 =
∑ ∑
(log 𝑝1 )(log 𝑝2 ).
𝑝1 ≤π‘₯ 𝑝2 ≤π‘₯
𝑝1 βˆ€π‘“ (π‘˜) 𝑝2 βˆ€π‘“ (π‘˜)
𝑝1 ≡𝑝2 mod 𝑓 (π‘˜)
𝑙=1
(𝑙,𝑓 (π‘˜))=1
Here we separate the diagonal terms with 𝑝1 = 𝑝2 from the off-diagonal ones. Note
that the conditions 𝑝1 βˆ£π‘“ (π‘˜) and 𝑝1 ≡ 𝑝2 mod 𝑓 (π‘˜) imply that 𝑝1 = 𝑝2 . We may therefore
ignore the constraint 𝑝𝑗 ∀ 𝑓 (π‘˜) (𝑗 = 1, 2) when considering the off-diagonal terms. In this
way, by symmetry, we derive the formula
𝑓 (π‘˜)
∑
πœ—(π‘₯; 𝑓 (π‘˜), 𝑙)2 =
∑
(log 𝑝1 )(log 𝑝2 ).
𝑝1 <𝑝2 ≤π‘₯
𝑝1 ≡𝑝2 mod 𝑓 (π‘˜)
𝑝≤π‘₯
π‘βˆ€π‘“ (π‘˜)
𝑙=1
(𝑙,𝑓 (π‘˜))=1
∑∑
(log 𝑝)2 + 2
We next reintroduce the terms with π‘βˆ£π‘“ (π‘˜), and sum over π‘˜. Then, much as in the analysis
leading to the relation (9), we find that
∑
∑
𝑆1 = 2𝑆0 +
𝑓 ′ (π‘˜)
(log 𝑝)2 + 𝑂(π‘₯(log π‘₯)2 ),
(10)
𝑝≤π‘₯
𝑦1 <π‘˜≤𝑦
where
𝑆0 =
∑
𝑦1 <π‘˜≤𝑦
𝑓 ′ (π‘˜)
∑∑
(log 𝑝1 )(log 𝑝2 ).
(11)
𝑝1 <𝑝2 ≤π‘₯
𝑝1 ≡𝑝2 mod 𝑓 (π‘˜)
On applying partial summation within (10), and substituting the outcome together
with (8) and (9) into (6), we obtain our final formula
∑
𝑉𝑓 (π‘₯, 𝑦) = 2𝑆0 − π‘₯2 Φ𝑓 (𝑦1 , 𝑦) + 𝑓 (𝑦)
(log 𝑝)2 + 𝑂(π‘₯2 (log π‘₯)−𝐴 ).
(12)
𝑝≤π‘₯
SPARSE VARIANCE FOR PRIMES IN ARITHMETIC PROGRESSION
5
3. The circle method
In this section we apply the circle method to evaluate the sum 𝑆0 defined in equation
(11). With this objective in view, we rewrite the congruence condition 𝑝1 ≡ 𝑝2 mod 𝑓 (π‘˜)
in (11) in the form 𝑝2 − 𝑝1 = β„Žπ‘“ (π‘˜), with β„Ž ∈ β„€. In addition, we observe that the
supplementary condition β„Ž ≥ 1 is equivalent to the summation condition 𝑝1 < 𝑝2 imposed
in the second summation of (11). Consequently, on writing
∑
∑
𝑇𝑓 (𝛼) =
𝑓 ′ (π‘˜)
𝑒(π›Όβ„Žπ‘“ (π‘˜)),
(13)
𝑦1 <π‘˜≤𝑦
π‘ˆ (𝛼) =
∑
β„Ž≤π‘₯/𝑓 (π‘˜)
(log 𝑝)𝑒(𝛼𝑝),
(14)
𝑝≤π‘₯
we may use orthogonality to infer from (11) that
∫ 1
𝑇𝑓 (𝛼)βˆ£π‘ˆ (𝛼)∣2 𝑑𝛼.
𝑆0 =
0
We analyse this integral by splitting the range of integration into major and minor arcs.
The dissection depends on the absolute constant 𝐷 ≥ 1 that arises in Lemma 1 below.
Let 𝐢 = 4𝑑3 𝐷𝐡, and write 𝑄 = (log π‘₯)𝐢 . Let 𝔐 denote the union of the pairwise disjoint
intervals βˆ£π›Ό−π‘Ž/π‘žβˆ£ ≤ 𝑄/π‘₯ with 1 ≤ π‘Ž ≤ π‘ž ≤ 𝑄 and (π‘Ž, π‘ž) = 1. Let π”ͺ = [𝑄/π‘₯, 1+𝑄/π‘₯]βˆ–π”.
The contribution from the minor arcs is controlled with the aid of the bound
sup βˆ£π‘‡π‘“ (𝛼)∣ β‰ͺ π‘₯(log π‘₯)−2𝐴
(15)
𝛼∈π”ͺ
that we establish in a moment. Equipped with (15) and the elementary estimate
∫ 1
∑
βˆ£π‘ˆ (𝛼)∣2 𝑑𝛼 =
(log 𝑝)2 β‰ͺ π‘₯ log π‘₯,
0
𝑝≤π‘₯
we infer that
𝑆0 = 𝑆𝔐 + 𝑂(π‘₯2 (log π‘₯)1−2𝐴 ),
(16)
where
∫
𝑆𝔐 =
𝑇𝑓 (𝛼)βˆ£π‘ˆ (𝛼)∣2 𝑑𝛼.
(17)
𝔐
The major arc contribution 𝑆𝔐 will be discussed in §4. In the remainder of this section,
we briefly sketch a proof of (15). This bound of Weyl’s type is certainly familiar territory
for workers in the field, but a suitable reference does not seem to be available. The
argument below is streamlined so as to confirm (15) with little effort, but when one seeks
for the strongest possible bounds, one can do rather better, in particular when 𝑑 is small.
This is of no relevance for us, but an analysis of 𝑉𝑓 (π‘₯, 𝑦) under GRH is sensitive to this
comment.
Let 𝑦0 ≤ 𝑧0 ≤ 𝑧, and consider the weighted Weyl sum
∑
𝑆(𝛽; 𝑧0 , 𝑧) =
𝑓 ′ (𝑛)𝑒(𝛽𝑓 (𝑛)).
𝑧0 <𝑛≤𝑧
6
J. BRÜDERN AND T. D. WOOLEY
Lemma 1. There is an absolute constant 𝐷 ≥ 1 with the property that whenever 𝛽 ∈ ℝ,
and 𝑏 and π‘Ÿ are coprime natural numbers with βˆ£π‘Ÿπ›½ − π‘βˆ£ ≤ 1/π‘Ÿ, one has
)1/(𝐷𝑑3 )
(
.
𝑆(𝛽; 𝑧0 , 𝑧) β‰ͺ 𝑧 𝑑 (log 𝑧) (π‘Ÿ + 𝑧 𝑑 βˆ£π‘Ÿπ›½ − π‘βˆ£)−1 + 𝑧 −1 + (π‘Ÿ + 𝑧 𝑑 βˆ£π‘Ÿπ›½ − π‘βˆ£)𝑧 −𝑑
Proof. When π‘˜ is a natural number, let 𝐽𝑠,π‘˜ (𝑃 ) denote the number of integral solutions
of the simultaneous equations
𝑠
∑
(π‘₯𝑗𝑖 − 𝑦𝑖𝑗 ) = 0
(1 ≤ 𝑗 ≤ π‘˜),
𝑖=1
with 1 ≤ π‘₯𝑖 , 𝑦𝑖 ≤ 𝑃 (1 ≤ 𝑖 ≤ 𝑠). Then, by Theorem 7.4 of Vaughan [9], for example,
there exists an absolute constant 𝐷 ≥ 1 such that whenever 𝑠 ≥ 21 π·π‘˜ 3 , one has 𝐽𝑠,π‘˜ (𝑃 ) β‰ͺ
𝑃 2𝑠−π‘˜(π‘˜+1)/2 . Applying Theorem 5.2 of Vaughan [9] in conjunction with partial summation,
we deduce that
3
𝑆(𝛽; 𝑧0 , 𝑧) β‰ͺ 𝑧 𝑑 (log 𝑧)(π‘Ÿ−1 + 𝑧 −1 + π‘Ÿπ‘§ −𝑑 )1/(𝐷𝑑 ) .
The conclusion of the lemma now follows from a familiar transference principle (see, for
example, Exercise 2 of Chapter 2 of Vaughan [9]).
β–‘
We are now ready to deduce (15). Let 𝑦(β„Ž) be the upper bound for π‘˜ determined by
the simultaneous conditions π‘˜ ≤ 𝑦 and 𝑓 (π‘˜) ≤ π‘₯/β„Ž. Then 𝑓 (𝑦(β„Ž)) = min{𝑓 (𝑦), π‘₯/β„Ž}.
Reversing the order of summation in (13) yields
∑
𝑇𝑓 (𝛼) =
𝑆(π›Όβ„Ž; 𝑦1 , 𝑦(β„Ž)).
(18)
β„Ž≤(log π‘₯)𝐡
Suppose that 𝛼 ∈ π”ͺ. For any integer β„Ž with 1 ≤ β„Ž ≤ (log π‘₯)𝐡 , we apply Dirichlet’s
theorem to find coprime integers π‘Ÿ = π‘Ÿ(β„Ž) and 𝑏 = 𝑏(β„Ž) with
1 ≤ π‘Ÿ ≤ 𝑦(β„Ž)𝑑−2/3
and βˆ£π›Όβ„Žπ‘Ÿ − π‘βˆ£ ≤ 𝑦(β„Ž)−𝑑+2/3 .
The validity of the simultaneous conditions β„Žπ‘Ÿ ≤ 𝑄 and βˆ£π›Ό − 𝑏/(β„Žπ‘Ÿ)∣ ≤ 𝑄/π‘₯ would imply
that 𝛼 ∈ 𝔐, whence for each β„Ž, at least one of these inequalities fails. If β„Žπ‘Ÿ > 𝑄, then
π‘Ÿ > (log π‘₯)𝐢−𝐡 , and an application of Lemma 1 with 𝛽 = π›Όβ„Ž yields
3
3
𝑆(π›Όβ„Ž; 𝑦1 , 𝑦(β„Ž)) β‰ͺ π‘₯(log π‘₯)(π‘Ÿ−1 + 𝑦(β„Ž)−2/3 )1/(𝐷𝑑 ) β‰ͺ π‘₯(log π‘₯)1+(𝐡−𝐢)/(𝐷𝑑 ) .
If β„Žπ‘Ÿ ≤ 𝑄 but βˆ£π›Ό − 𝑏/(β„Žπ‘Ÿ)∣ > 𝑄/π‘₯, meanwhile, then in like manner Lemma 1 gives
3
3
𝑆(π›Όβ„Ž; 𝑦1 , 𝑦(β„Ž)) β‰ͺ π‘₯(log π‘₯)𝑄−1/(𝐷𝑑 ) β‰ͺ π‘₯(log π‘₯)1−𝐢/(𝐷𝑑 ) .
Summing over natural numbers β„Ž with β„Ž ≤ (log π‘₯)𝐡 , we deduce (15) from (18).
4. The major arcs
We move on to analyse the major arc contribution 𝑆𝔐 to 𝑆0 , and subsequently incorporate the conclusions of the previous section in order to obtain an asymptotic formula
for 𝑆0 . For 𝛼 ∈ 𝔐, one may write 𝛼 = 𝛽 + π‘Ž/π‘ž for some 𝛽 ∈ ℝ with βˆ£π›½βˆ£ ≤ 𝑄/π‘₯, and
coprime integers π‘Ž and π‘ž with 1 ≤ π‘Ž ≤ π‘ž ≤ 𝑄. Then, by (14) and Lemma 3.1 of Vaughan
[9], one has
πœ‡(π‘ž)
𝐽(𝛽) + 𝑂(π‘₯(log π‘₯)−5𝐢 ),
π‘ˆ (𝛼) =
πœ™(π‘ž)
SPARSE VARIANCE FOR PRIMES IN ARITHMETIC PROGRESSION
7
where
𝐽(𝛽) =
∑
𝑒(𝛽𝑛).
𝑛≤π‘₯
The trivial upper bound 𝑇𝑓 (𝛼) β‰ͺ π‘₯ log π‘₯ (which is available from (13)) and (17) now
suffice to confirm that
∫ 𝑄/π‘₯
π‘ž
∑ πœ‡(π‘ž)2 ∑
𝑆𝔐 =
𝑇𝑓 (𝛽 + π‘Ž/π‘ž)∣𝐽(𝛽)∣2 𝑑𝛽 + 𝑂(π‘₯2 (log π‘₯)−𝐢 ).
2
πœ™(π‘ž)
−𝑄/π‘₯
π‘Ž=1
π‘ž≤𝑄
(π‘Ž,π‘ž)=1
We would like to extend the sum over all natural numbers π‘ž here. It is convenient
to do this in two steps. In the first of these steps, we extend the range for π‘ž above to
π‘ž ≤ π‘₯1/(2𝑑) . For this, we note that the intervals βˆ£π›Ό − π‘Ž/π‘žβˆ£ ≤ 𝑄/π‘₯ with 1 ≤ π‘Ž ≤ π‘ž ≤ π‘₯1/(2𝑑)
and (π‘Ž, π‘ž) = 1 are pairwise disjoint. When π‘Ž and π‘ž are coprime natural numbers with
1 ≤ π‘Ž ≤ π‘ž and 𝑄 < π‘ž ≤ π‘₯1/(2𝑑) , and βˆ£π›½βˆ£ ≤ 𝑄/π‘₯, one has 𝛽 + π‘Ž/π‘ž ∈ π”ͺ. Therefore, on
recalling (15), for any such 𝛽, π‘Ž and π‘ž, one has
βˆ£π‘‡π‘“ (𝛽 + π‘Ž/π‘ž)∣ ≤ sup βˆ£π‘‡π‘“ (𝛼)∣ β‰ͺ π‘₯(log π‘₯)−2𝐴 .
𝛼∈π”ͺ
Combining this estimate with the elementary inequality
∫ 1/2
∫ 𝑄/π‘₯
2
∣𝐽(𝛽)∣2 𝑑𝛽 = [π‘₯],
∣𝐽(𝛽)∣ 𝑑𝛽 ≤
−1/2
−𝑄/π‘₯
we deduce that
∑
𝑆𝔐 =
π‘ž≤π‘₯1/(2𝑑)
∫ 𝑄/π‘₯
π‘ž
πœ‡(π‘ž)2 ∑
𝑇𝑓 (𝛽 + π‘Ž/π‘ž)∣𝐽(𝛽)∣2 𝑑𝛽 + 𝑂(π‘₯2 (log π‘₯)1−2𝐴 ).
2
πœ™(π‘ž) π‘Ž=1 −𝑄/π‘₯
(19)
(π‘Ž,π‘ž)=1
Before extending the range for π‘ž even further, we complete the integral in (19) to
βˆ£π›½βˆ£ ≤ 21 . In this range for 𝛽, one has 𝐽(𝛽) β‰ͺ βˆ£π›½βˆ£−1 , so that another application of the
trivial bound 𝑇𝑓 (𝛼) β‰ͺ π‘₯ log π‘₯ in combination with straightforward estimates yields the
upper bound
∫
π‘ž
∑ πœ‡(π‘ž)2 ∑
βˆ£π‘‡π‘“ (𝛽 + π‘Ž/π‘ž)𝐽(𝛽)2 ∣ 𝑑𝛽
2
πœ™(π‘ž) π‘Ž=1
1/(2𝑑)
π‘ž≤π‘₯
1
(π‘Ž,π‘ž)=1 𝑄/π‘₯≤βˆ£π›½βˆ£≤ 2
∑ 1 ∫ 1/2
𝛽 −2 𝑑𝛽 β‰ͺ π‘₯2 (log π‘₯)2−𝐢 .
β‰ͺ π‘₯(log π‘₯)
πœ™(π‘ž) 𝑄/π‘₯
π‘ž≤π‘₯
Hence, we may indeed extend the integration in (19) to [− 21 , 21 ] with the introduction of
acceptable errors. However, by orthogonality and (13), one has
∫ 1/2
∑
∑
∑ ∑ ( π‘Žβ„Žπ‘“ (π‘˜) )
2
′
𝑇𝑓 (𝛽 + π‘Ž/π‘ž)∣𝐽(𝛽)∣ 𝑑𝛽 =
𝑓 (π‘˜)
𝑒
π‘ž
−1/2
𝑛
≤π‘₯
𝑛
≤π‘₯
𝑦 <π‘˜≤𝑦
β„Ž≤π‘₯/𝑓 (π‘˜)
1
1
2
𝑛2 −𝑛1 =β„Žπ‘“ (π‘˜)
=
∑
𝑦1 <π‘˜≤𝑦
𝑓 ′ (π‘˜)
∑
β„Ž≤π‘₯/𝑓 (π‘˜)
( π‘Žβ„Žπ‘“ (π‘˜) )
𝑒
([π‘₯] − β„Žπ‘“ (π‘˜)).
π‘ž
8
J. BRÜDERN AND T. D. WOOLEY
In the last sum, we may replace [π‘₯] by π‘₯, introducing an error bounded by 𝑂(π‘₯ log π‘₯) into
this identity. We summarize and inject these results into (19) to infer that
𝑆𝔐 = 𝑀0 + 𝑂(π‘₯2 (log π‘₯)−𝐴 ),
where
𝑀0 =
∑
π‘ž≤π‘₯1/(2𝑑)
(20)
∑
πœ‡(π‘ž)2 ∑ ′
𝑓
(π‘˜)
π‘π‘ž (β„Žπ‘“ (π‘˜))(π‘₯ − β„Žπ‘“ (π‘˜)),
πœ™(π‘ž)2 𝑦 <π‘˜≤𝑦
1
in which we have written
π‘π‘ž (𝑛) =
β„Ž≤π‘₯/𝑓 (π‘˜)
π‘ž
∑
π‘Ž=1
(π‘Ž,π‘ž)=1
( π‘Žπ‘› )
𝑒
π‘ž
for Ramanujan’s sum.
Recall that 𝑦(β„Ž) was defined through the equation 𝑓 (𝑦(β„Ž)) = min{𝑓 (𝑦), π‘₯/β„Ž}. In the
definition of 𝑀0 , we exchange the order of summation of β„Ž and π‘˜ and then sort π‘˜ according
to residue classes modulo π‘ž. Then
π‘ž
∑ πœ‡(π‘ž)2 ∑ ∑
∑
𝑀0 =
𝑐
(β„Žπ‘“
(π‘Ž))
𝑓 ′ (π‘˜)(π‘₯ − β„Žπ‘“ (π‘˜)).
(21)
π‘ž
2
πœ™(π‘ž)
1/(2𝑑)
π‘Ž=1
β„Ž≤π‘₯/𝑓 (𝑦1 )
π‘ž≤π‘₯
𝑦1 <π‘˜≤𝑦(β„Ž)
π‘˜≡π‘Ž mod π‘ž
For small π‘ž, Euler’s summation formula readily yields an asymptotic formula for the
innermost sum, with a main term independent of π‘Ž. It is then possible to sum π‘π‘ž (β„Žπ‘“ (π‘Ž))
over π‘Ž. This process will largely disentangle the sum in (21). Two lemmata make this
analysis precise.
Lemma 2. With π‘₯, 𝑦 and β„Ž in the ranges as above, and uniformly in π‘ž, one has
∫
∑
1 𝑓 (𝑦(β„Ž))
′
(π‘₯ − β„Žπ‘‘) 𝑑𝑑 + 𝑂(π‘₯𝑦 𝑑−1 + 𝑦 2𝑑−1 β„Ž).
𝑓 (π‘˜)(π‘₯ − β„Žπ‘“ (π‘˜)) =
π‘ž 𝑓 (𝑦1 )
𝑦1 <π‘˜≤𝑦(β„Ž)
π‘˜≡π‘Ž mod π‘ž
Proof. Observe first that by Euler’s summation formula, if 𝑔 : [𝑋, π‘Œ ] → ℝ is continuously
differentiable, then
∫
∑
1 π‘Œ
𝑔(π‘˜) =
𝑔(𝑑) 𝑑𝑑 + 𝐸,
(22)
π‘ž
𝑋
𝑋<π‘˜≤π‘Œ
π‘˜≡π‘Ž mod π‘ž
where the implicitly defined real number 𝐸 satisfies the bound
∫ π‘Œ
∣𝐸∣ ≤
βˆ£π‘” ′ (𝑑)∣ 𝑑𝑑 + βˆ£π‘”(𝑋)∣ + βˆ£π‘”(π‘Œ )∣.
𝑋
For the application at hand, we take 𝑔(𝑑) = 𝑓 ′ (𝑑)(π‘₯ − β„Žπ‘“ (𝑑)), and put 𝑋 = 𝑦1 and
π‘Œ = 𝑦(β„Ž). Note, in particular, that π‘₯ − β„Žπ‘“ (𝑑) = 𝑂(π‘₯) for all relevant choices of 𝑑, and
hence 𝑔(𝑑) β‰ͺ 𝑦 𝑑−1 π‘₯ for 𝑋 ≤ 𝑑 ≤ π‘Œ . Also, one has 𝑔 ′ (𝑑) = 𝑓 ′′ (𝑑)(π‘₯ − β„Žπ‘“ (𝑑)) − β„Žπ‘“ ′ (𝑑)2 . Then
it follows that
𝐸 β‰ͺ π‘₯𝑦 𝑑−1 + 𝑦 2𝑑−1 β„Ž.
(23)
Finally, by substitution one obtains
∫ 𝑦(β„Ž)
∫ 𝑓 (𝑦(β„Ž))
′
𝑓 (𝑑)(π‘₯ − β„Žπ‘“ (𝑑)) 𝑑𝑑 =
(π‘₯ − β„Žπœ ) π‘‘πœ.
(24)
𝑦1
𝑓 (𝑦1 )
SPARSE VARIANCE FOR PRIMES IN ARITHMETIC PROGRESSION
The conclusion of the lemma follows by collecting together (22), (23) and (24).
9
β–‘
In advance of the statement of the next lemma, for each natural number π‘š, we define
𝜌(π‘š) to be the number of solutions of the congruence 𝑓 (π‘Ž) ≡ 0 mod π‘š.
Lemma 3. For any β„Ž ∈ β„•, the sum
π‘ž
1∑
π‘π‘ž (β„Žπ‘“ (π‘Ž))
π‘€β„Ž (π‘ž) =
π‘ž π‘Ž=1
is a multiplicative function of π‘ž. When 𝑝 is a prime, one has
{
𝑝 − 1,
when π‘βˆ£β„Ž,
π‘€β„Ž (𝑝) =
𝜌(𝑝) − 1, when 𝑝 ∀ β„Ž.
Proof. Since π‘π‘ž (𝑛) is multiplicative as a function of π‘ž (see, for example, Theorem 272 of
Hardy and Wright [4]), it suffices to apply the Chinese Remainder Theorem to the sum
over π‘Ž in order to confirm the multiplicativity of π‘€β„Ž (π‘ž). Next, when 𝑝 is a prime, we apply
the definition of π‘π‘ž (𝑛) and orthogonality to obtain the relation
𝑝
𝑝−1
1 ∑ ∑ ( π‘™β„Žπ‘“ (π‘Ž) )
𝑒
π‘€β„Ž (𝑝) =
= card{1 ≤ π‘Ž ≤ 𝑝 : β„Žπ‘“ (π‘Ž) ≡ 0 mod 𝑝} − 1,
𝑝 𝑙=1 π‘Ž=1
𝑝
and the lemma follows at once.
β–‘
We inject the asymptotic formula supplied by Lemma 2 into (21). This introduces an
error term
∑ πœ‡(π‘ž)2 ∑
π‘žπœ™(π‘ž)(π‘₯𝑦 𝑑−1 + 𝑦 2𝑑−1 β„Ž) β‰ͺ π‘₯2−1/(2𝑑)+πœ€ ,
2
πœ™(π‘ž)
1/(2𝑑)
β„Ž≤π‘₯/𝑓 (𝑦1 )
π‘ž≤π‘₯
so that we now have
∑
𝑀0 =
π‘ž≤π‘₯1/(2𝑑)
πœ‡(π‘ž)2
πœ™(π‘ž)2
∑
∫
𝑓 (𝑦(β„Ž))
π‘€β„Ž (π‘ž)
(π‘₯ − β„Žπ‘‘) 𝑑𝑑 + 𝑂(π‘₯2−1/(2𝑑)+πœ€ ).
(25)
𝑓 (𝑦1 )
β„Ž≤π‘₯/𝑓 (𝑦1 )
We next exchange the order of summation and complete the sum over π‘ž. By Lemma 3,
the multiplicative function π‘€β„Ž (π‘ž)πœ‡(π‘ž)2 is non-negative and bounded by 𝑂((π‘ž, β„Ž)π‘ž πœ€ ). This
implies that
∑ πœ‡(π‘ž)2
π‘€β„Ž (π‘ž) = π‘Š (β„Ž) + 𝑂(β„Žπœ€ π‘₯−1/(2𝑑) ),
2
πœ™(π‘ž)
1/(2𝑑)
π‘ž≤π‘₯
with
π‘Š (β„Ž) =
∞
∑
πœ‡(π‘ž)2
π‘ž=1
(
)
β„Ž ∏
𝜌(𝑝) − 1
1+
π‘€β„Ž (π‘ž) =
.
πœ™(π‘ž)2
πœ™(β„Ž)
(𝑝 − 1)2
(26)
π‘βˆ€β„Ž
A trivial estimate for the integral in (25) shows it to be 𝑂(𝑓 (𝑦)π‘₯), and this suffices here
to conclude that
∫ 𝑓 (𝑦(β„Ž))
∑
𝑀0 =
π‘Š (β„Ž)
(π‘₯ − β„Žπ‘‘) 𝑑𝑑 + 𝑂(π‘₯2−1/(2𝑑)+πœ€ ).
β„Ž≤π‘₯/𝑓 (𝑦1 )
𝑓 (𝑦1 )
10
J. BRÜDERN AND T. D. WOOLEY
The derivative of 𝑑(π‘₯ − 12 β„Žπ‘‘) with respect to 𝑑 is π‘₯ − β„Žπ‘‘, and we have 𝑓 (𝑦(β„Ž)) = 𝑓 (𝑦) when
β„Ž ≤ π‘₯/𝑓 (𝑦), and 𝑓 (𝑦(β„Ž)) = π‘₯/β„Ž when β„Ž > π‘₯/𝑓 (𝑦). Hence, if we compute the integral and
split the sum over β„Ž at π‘₯/𝑓 (𝑦) when appropriate, we readily find that
(
)
(27)
𝑀0 = 12 𝑓 (𝑦1 )2 Θ𝑓 (π‘₯/𝑓 (𝑦1 )) − 𝑓 (𝑦)2 Θ𝑓 (π‘₯/𝑓 (𝑦)) + 𝑂(π‘₯2−1/(2𝑑)+πœ€ )
where
Θ𝑓 (𝐻) =
∑ π‘Š (β„Ž)
(𝐻 − β„Ž)2 .
β„Ž
β„Ž≤𝐻
(28)
We summarize the analysis based on the application of the circle method by combining
(16), (20) and (27), thereby obtaining the formula
2𝑆0 = 𝑓 (𝑦1 )2 Θ𝑓 (π‘₯/𝑓 (𝑦1 )) − 𝑓 (𝑦)2 Θ𝑓 (π‘₯/𝑓 (𝑦)) + 𝑂(π‘₯2 (log π‘₯)−𝐴 ),
(29)
and thus we enter the opening phase of the endgame.
5. An exercise in multiplicative number theory
In view of the relations (29) and (12), the evaluation of 𝑉𝑓 (π‘₯, 𝑦) is reduced to the
deduction of appropriate asymptotic formulae for the functions Φ𝑓 and Θ𝑓 , defined in (7)
and (28), respectively. This is accomplished largely by a series of exercises in multiplicative
number theory. We begin with an analysis of the Dirichlet series associated with the
function π‘Š (β„Ž)/β„Ž, restricting attention for the time being to the situation wherein 𝜌(2) ≥
1. An inspection of (26) shows that the series
𝐷(𝑠) =
∞
∑
π‘Š (𝑛)𝑛−1−𝑠
𝑛=1
converges absolutely in the half-plane Re (𝑠) > 0, and that π‘Š (β„Ž)/π‘Š (1) is multiplicative.
Hence, for Re (𝑠) > 0, one has
∏
𝐷(𝑠) = π‘Š (1)
𝐷𝑝 (𝑠)
(30)
𝑝
with
(
𝜌(𝑝) − 1
𝐷𝑝 (𝑠) = 1 + 1 +
(𝑝 − 1)2
)−1 ∑
∞
πœ™(𝑝𝑙 )−1 𝑝−𝑙𝑠 .
(31)
𝑙=1
Notice here that since 𝜌(𝑝) ≥ 0, the Euler factor 𝐷𝑝 (𝑠) is defined for all values of 𝑝, with
the exception of 𝑝 = 2 in the special case 𝜌(2) = 0. It is for this reason that we isolate
the latter situation below.2
The Euler product (30) yields an analytic continuation of 𝐷(𝑠) to a meromorphic
function on the half-plane Re (𝑠) > −2. In order to see this, it will be convenient to write
πœ‰ = 𝑝−1−𝑠 , and define the real number πœ† = πœ†(𝑝) by means of the identity
)−1
(
πœ†(𝑝)
𝜌(𝑝) − 1
=1− 2 .
(32)
1+
2
(𝑝 − 1)
𝑝
2We
are grateful to the referee for prompting our separate treatment of this case.
SPARSE VARIANCE FOR PRIMES IN ARITHMETIC PROGRESSION
11
It is apparent from the relation defining πœ†(𝑝) that 0 ≤ πœ† β‰ͺ 1, wherein the implicit
constant depends only on 𝑓 . We may rearrange the Euler factors into the shape
(
)
𝑝
πœ†
πœ‰
𝐷𝑝 (𝑠) = 1 +
1− 2
,
𝑝−1
𝑝 1−πœ‰
and consequently,
(
)
πœ†
πœ‰
1−
,
(1 − πœ‰)𝐷𝑝 (𝑠) = 1 +
𝑝−1
𝑝
(
)
( )
( )2
(
)
πœ‰
πœ‰ 1−πœ†
πœ‰
𝑝
πœ†
1−
(1 − πœ‰)𝐷𝑝 (𝑠) = 1 +
−
1−
,
𝑝
𝑝 𝑝−1
𝑝 𝑝−1
𝑝
)(
)
(
(1 − πœ‰/𝑝)(1 − πœ‰)
1−πœ†
πœ‰/𝑝
.
𝐷𝑝 (𝑠) = 1 +
1 − πœ‰ 2 /𝑝2
1 + πœ‰/𝑝
𝑝−1
On recalling (30), and rewriting πœ‰ as 𝑝−1−𝑠 again, the previous formulae transform into
the relations
𝜁(𝑠 + 1)𝜁(𝑠 + 2)
𝐷(𝑠)
= 𝜁(𝑠 + 1)𝐸1 (𝑠) = 𝜁(𝑠 + 1)𝜁(𝑠 + 2)𝐸2 (𝑠) =
𝐸3 (𝑠),
(33)
π‘Š (1)
𝜁(2𝑠 + 4)
where we have written
∏(
𝐸1 (𝑠) =
1+
𝑝
𝑝−𝑠
𝑝(𝑝 − 1)
(
πœ†(𝑝)
1−
𝑝
))
,
𝑝−𝑠
𝑝−2𝑠
(1
−
πœ†(𝑝))
−
𝑝2 (𝑝 − 1)
𝑝3 (𝑝 − 1)
𝑝
)
∏(
1 − πœ†(𝑝)
.
𝐸3 (𝑠) =
1+
(𝑝 − 1)(1 + 𝑝𝑠+2 )
𝑝
∏(
𝐸2 (𝑠) =
1+
(34)
(
πœ†(𝑝)
1−
𝑝
))
,
(35)
(36)
Here 𝜁 denotes Riemann’s zeta function, and the product 𝐸𝑗 converges absolutely and
locally uniformly in the half-plane Re (𝑠) > −(𝑗 + 1)/2 for 𝑗 = 1, 2, 3. In particular,
the relations recorded in (33) provide the analytic continuation of 𝐷(𝑠) to the half-plane
Re (𝑠) > −2. In the half-plane Re (𝑠) ≥ −3/2, meanwhile, the only singularities of 𝐷(𝑠)
are simple poles at 𝑠 = 0 and 𝑠 = −1.
Returning now to (28), we deduce from one of Perron’s formulae that
∫
1 2+𝑖∞ 𝐷(𝑠)𝐻 𝑠+2
Θ𝑓 (𝐻) =
𝑑𝑠.
(37)
πœ‹π‘– 2−𝑖∞ 𝑠(𝑠 + 1)(𝑠 + 2)
Observe that from (36), one finds that the function 𝐸3 (𝑠) is holomorphic and uniformly
bounded in Re (𝑠) ≥ −3/2. Invoking only standard estimates for the zeta factors in (33),
one can move the line of integration on the right hand side of (37) to Re (𝑠) = −3/2.
The reader may care to inspect the analysis on p. 141 of Goldston and Vaughan [3] for
the necessary ideas, though in the present circumstances one should not use the Riemann
hypothesis. One may also compare the discussion on p. 791 of Vaughan [10]. If 𝑅(𝑧) is
the residue of the integrand in (37) at 𝑠 = 𝑧, then it follows that
1
Θ (𝐻)
2 𝑓
= 𝑅(0) + 𝑅(−1) + 𝑂(𝐻 1/2 ).
(38)
12
J. BRÜDERN AND T. D. WOOLEY
We compute 𝑅(0) explicitly by using (33) and (34) along with the familiar expansion
1
𝜁(1 + 𝑠)
𝛾
= 2 + + 𝑂(1).
𝑠
𝑠
𝑠
A short calculation then reveals that
where
𝑅(0) = 21 𝑐(𝑓 )𝐻 2 log 𝐻 + Γ0 𝐻 2 ,
(39)
)
∏(
1 − πœ†(𝑝)𝑝−1
𝑐(𝑓 ) = π‘Š (1)
1+
𝑝(𝑝 − 1)
𝑝
(40)
and Γ0 = Γ0 (𝑓 ) is a certain real number, the precise value of which is not relevant to
our subsequent discussion. Another short calculation leads from (26) and (32) to the
alternative formulae
)(
)
∏(
𝜌(𝑝) − 1
1 − πœ†(𝑝)𝑝−1
𝑐(𝑓 ) =
1+
1+
2
(𝑝
−
1)
𝑝(𝑝 − 1)
𝑝
) ∑
∞
∏(
πœ‡(𝑛)2 𝜌(𝑛)
𝜌(𝑝)
=
.
(41)
=
1+
𝑝(𝑝 − 1)
π‘›πœ™(𝑛)
𝑛=1
𝑝
In the latter form, this constant will play an important role later.
A similar technique may be applied to evaluate 𝑅(−1). In this instance we use (33)
and (35) and then reach the transitional formula
𝑅(−1) = −π‘Š (1)𝐸2 (−1)𝜁(0)𝐻 log 𝐻 + Γ−1 𝐻,
(42)
where Γ−1 = Γ−1 (𝑓 ) is another real number, the precise value of which is again of little
importance in what follows. We remark, however, that Γ−1 (𝑓 ) could be made just as
explicit as 𝑐(𝑓 ). Observe next that 𝜁(0) = −1/2, and so on applying (26), (32) and (35),
we find that the Euler factors of π‘Š (1) and 𝐸2 (−1) combine to give the formula
)(
)
∏(
𝜌(𝑝) − 1
1 − πœ†(𝑝) 1 − πœ†(𝑝)𝑝−1
π‘Š (1)𝐸2 (−1) =
1+
−
1+
(𝑝 − 1)2
𝑝(𝑝 − 1)
𝑝(𝑝 − 1)
𝑝
(
)
(
)
∏
𝜌(𝑝) − 1
πœ†(𝑝)
=
1+
1− 2
= 1.
2
(𝑝 − 1)
𝑝
𝑝
We may therefore conclude that
𝑅(−1) = 21 𝐻 log 𝐻 + Γ−1 𝐻.
(43)
On substituting (39) and (43) into (38), we obtain the expansion
Θ𝑓 (𝐻) = 𝑐(𝑓 )𝐻 2 log 𝐻 + 2Γ0 𝐻 2 + 𝐻 log 𝐻 + 2Γ−1 𝐻 + 𝑂(𝐻 1/2 ),
which may be incorporated into (29), leading to the asymptotic formula
(
)
(
)
𝑓 (𝑦)
π‘₯
2
−π‘₯𝑓 (𝑦) log
− 2Γ−1 π‘₯𝑓 (𝑦)
2𝑆0 = 𝑐(𝑓 )π‘₯ log
𝑓 (𝑦1 )
𝑓 (𝑦)
(
)
+ 𝑂 π‘₯1/2 𝑓 (𝑦)3/2 + π‘₯2 (log π‘₯)−𝐴 .
This completes the evaluation of the expression 𝑆0 defined in equation (11).
(44)
SPARSE VARIANCE FOR PRIMES IN ARITHMETIC PROGRESSION
13
We turn our attention now to the situation with 𝜌(2) = 0. Here, as a consequence
of (26), one sees that π‘Š (β„Ž) = 0 whenever β„Ž is odd, and moreover that for all natural
numbers π‘š one has
)−1
(
π‘š ∏
π‘Š (2π‘š)
𝜌(𝑝) − 1
=
1+
.
π‘Š (2)
πœ™(2π‘š)
(𝑝 − 1)2
π‘βˆ£π‘š
𝑝>2
Define
{
2,
when 𝑝 = 2,
𝜌0 (𝑝) =
𝜌(𝑝), when 𝑝 > 2.
In the remainder of this section, we adopt the convention that the decoration of a symbol
with a subscript 0 is to be interpreted as indicating that the definition of that symbol is
to be modified by replacing 𝜌(𝑝) by 𝜌0 (𝑝) throughout. We now find that π‘Š (2) = π‘Š0 (1)
and
(
)−1 (
)
π‘Š (2π‘š) ∏
𝜌0 (𝑝) − 1
𝑝
=
1+
.
π‘Š0 (1)
(𝑝 − 1)2
𝑝−1
π‘βˆ£π‘š
In particular, it follows that π‘Š (2π‘š)/π‘Š0 (1) is a multiplicative function of π‘š. With the
Dirichlet series 𝐷(𝑠) defined as before, we again find that 𝐷(𝑠) converges absolutely in
the half-plane Re (𝑠) > 0. Hence, for Re (𝑠) > 0, one has
𝐷(𝑠) = 2
−1−𝑠
∞
∑
π‘š=1
π‘Š (2π‘š)π‘š−1−𝑠 = π‘Š0 (1)2−1−𝑠
∏
𝐷𝑝,0 (𝑠),
(45)
𝑝
with 𝐷𝑝,0 (𝑠) defined via (31).
We may now execute the argument following (31) to establish that the analogue of (33)
holds, save that in present circumstances 𝐷(𝑠)/π‘Š (1) is replaced by 21+𝑠 𝐷(𝑠)/π‘Š0 (1). It
follows, in particular, that the analogue of the asymptotic relation (38) holds. A modest
calculation reveals that
𝑅0 (0) = 41 𝑐0 (𝑓 )𝐻 2 log 𝐻 + Γ0 𝐻 2 ,
(46)
where 𝑐0 (𝑓 ) is defined via (40), and Γ0 = Γ0 (𝑓 ) is a certain (but different) real number,
the precise value of which is again of no importance in what follows. In addition, as a
consequence of the argument underlying (41), we have
)
)
∏(
∏(
𝜌0 (𝑝)
𝜌(𝑝)
𝑐0 (𝑓 ) =
1+
=2
1+
= 2𝑐(𝑓 ).
(47)
𝑝(𝑝
−
1)
𝑝(𝑝
−
1)
𝑝
𝑝>2
So far as 𝑅(−1) is concerned, the modification to the formula (33) leaves the relation
(42) essentially unchanged, save that Γ−1 will need adjustment invisible in the ensuing
discussion. We therefore deduce as before that the analogue of (43) holds, whence on
substituting (46) and (43) into (38) we conclude that
Θ𝑓 (𝐻) = 21 𝑐0 (𝑓 )𝐻 2 log 𝐻 + 2Γ0 𝐻 2 + 𝐻 log 𝐻 + 2Γ−1 𝐻 + 𝑂(𝐻 1/2 ).
Incorporating this, as before, into (29), and recalling (47), we arrive at the asymptotic
formula (44). In this way, we have established the asymptotic relation (44) in all circumstances.
14
J. BRÜDERN AND T. D. WOOLEY
6. Another exercise in multiplicative number theory
Our treatment of the sum Φ𝑓 defined in equation (7) takes the convolution
∑ πœ‡(π‘Ÿ)2
π‘ž
=
πœ™(π‘ž)
πœ™(π‘Ÿ)
π‘Ÿβˆ£π‘ž
as the starting point. By reversing the order of summation, we obtain
∑
𝑦0 (𝑓 )<π‘˜≤𝑧
∑ πœ‡(π‘Ÿ)2
𝑓 (π‘˜)
=
πœ™(𝑓 (π‘˜))
πœ™(π‘Ÿ)
π‘Ÿ≤𝑓 (𝑧)
∑
𝑦0 (𝑓 )<π‘˜≤𝑧
𝑓 (π‘˜)≡0 mod π‘Ÿ
)
∑ πœ‡(π‘Ÿ)2 ( 𝜌(π‘Ÿ)
𝑧 + 𝑂(𝜌(π‘Ÿ)) .
1=
πœ™(π‘Ÿ)
π‘Ÿ
π‘Ÿ≤𝑓 (𝑧)
Since 𝜌(𝑝) ≤ 𝑑 for primes 𝑝 large enough in terms of 𝑓 , we now deduce from (41) that
∑
𝑦0 (𝑓 )<π‘˜≤𝑧
𝑓 (π‘˜)
= 𝑐(𝑓 )𝑧 + 𝑂((log 𝑧)𝑑 ).
πœ™(𝑓 (π‘˜))
This relation may be summed by parts against the differentiable factor 𝑓 ′ (𝑑)/𝑓 (𝑑). For
large 𝑑, one has
𝑑 ′
(𝑓 (𝑑)/𝑓 (𝑑)) β‰ͺ 𝑑−2 ,
𝑑𝑑
and consequently we are led to the asymptotic formula
(
)
∑
𝑓 ′ (π‘˜)
(log 𝑧)𝑑
= 𝑐(𝑓 ) log 𝑓 (𝑧) + 𝑐˜(𝑓 ) + 𝑂
,
πœ™(𝑓 (π‘˜))
𝑧
𝑦0 (𝑓 )<π‘˜≤𝑧
where 𝑐˜(𝑓 ) is a real number, the value of which, yet again, plays no role in our conclusions.
Substituting this formula into the definition (7), we infer that whenever π‘₯1/𝑑 (log π‘₯)−𝐡 ≤
𝑦 ≤ π‘₯1/𝑑 , one has
(
)
𝑓 (𝑦)
Φ𝑓 (𝑦1 , 𝑦) = 𝑐(𝑓 ) log
+ 𝑂(π‘₯−1/𝑑 (log π‘₯)𝐡+𝑑 ).
(48)
𝑓 (𝑦1 )
We now substitute (44) and (48) into the asymptotic relation (12), thereby deducing
that
(
)
∑
𝑉𝑓 (π‘₯, 𝑦) = π‘₯𝑓 (𝑦) log 𝑓 (𝑦)+𝑓 (𝑦)
(log 𝑝)2 − π‘₯ log π‘₯
𝑝≤π‘₯
− 2Γ−1 π‘₯𝑓 (𝑦) + 𝑂(π‘₯1/2 𝑓 (𝑦)3/2 + π‘₯2 (log π‘₯)−𝐴 ).
By the Prime Number Theorem and partial summation, we therefore arrive at the asymptotic formula
𝑉𝑓 (π‘₯, 𝑦) = π‘₯𝑓 (𝑦) log 𝑓 (𝑦) − (2Γ−1 + 1)π‘₯𝑓 (𝑦) + 𝑂(π‘₯1/2 𝑓 (𝑦)3/2 + π‘₯2 (log π‘₯)−𝐴 ),
and this confirms the conclusion of our theorem with 𝐢(𝑓 ) = −(2Γ−1 + 1).
SPARSE VARIANCE FOR PRIMES IN ARITHMETIC PROGRESSION
15
References
[1] J. Brüdern and A. Perelli, Goldbach numbers in sparse sequences. Ann. Inst. Fourier (Grenoble) 48
(1998), 353–378.
[2] J. B. Friedlander and D. A. Goldston, Variance of distribution of primes in residue classes. Quart.
J. Math. Oxford Ser. (2) 47 (1996), 313–336.
[3] D. A. Goldston and R. C. Vaughan, On the Montgomery-Hooley asymptotic formula. Sieve methods,
exponential sums, and their applications in number theory (Cardiff, 1995), pp. 117–142, London
Math. Soc. Lecture Note Ser., vol. 237, Cambridge Univ. Press, Cambridge, 1997.
[4] G. H. Hardy and E. M. Wright, An introduction to the theory of numbers. Fifth edition. The Clarendon Press, Oxford University Press, New York, 1979.
[5] C. Hooley, On the Barban-Davenport-Halberstam theorem. I. J. Reine Angew. Math. 274/275 (1975),
206–223.
[6] C. Hooley, On the Barban-Davenport-Halberstam theorem. II. J. London Math. Soc. (2) 9 (1975),
625–636.
[7] H. Mikawa and T. P. Peneva, Primes in arithmetic progressions to spaced moduli. Arch. Math. (Basel)
84 (2005), 239–248.
[8] H. L. Montgomery, Primes in arithmetic progressions. Michigan Math. J. 17 (1970) 33–39.
[9] R. C. Vaughan, The Hardy-Littlewood method. Second edition. Cambridge Tracts in Mathematics,
vol. 125. Cambridge University Press, Cambridge, 1997.
[10] R. C. Vaughan, On a variance associated with the distribution of general sequences in arithmetic
progressions. I. R. Soc. Lond. Philos. Trans. Ser. A 356 (1998), 781–791.
[11] R. C. Vaughan, On a variance associated with the distribution of primes in arithmetic progressions.
Proc. London Math. Soc. (3) 82 (2001), 533–553.
JB: Institut für Algebra und Zahlentheorie, Universität Stuttgart, D-70511 Stuttgart,
Germany
TDW: School of Mathematics, University of Bristol, University Walk, Clifton, Bristol BS8 1TW, United Kingdom
Download