ON SUMS OF PRIMES FROM BEATTY SEQUENCES. Angel V. Kumchev 1 Department of Mathematics, Towson University, Towson, MD , U.S.A.

Similar documents
Prime numbers with Beatty sequences

Character sums with Beatty sequences on Burgess-type intervals

CONSECUTIVE PRIMES AND BEATTY SEQUENCES

The major arcs approximation of an exponential sum over primes

Goldbach's problem with primes in arithmetic progressions and in short intervals

On sums of squares of primes

Diophantine Approximation by Cubes of Primes and an Almost Prime

Congruences involving product of intervals and sets with small multiplicative doubling modulo a prime

On the Fractional Parts of a n /n

A Diophantine Inequality Involving Prime Powers

Density of non-residues in Burgess-type intervals and applications

Prime Numbers and Irrational Numbers

GAPS IN BINARY EXPANSIONS OF SOME ARITHMETIC FUNCTIONS, AND THE IRRATIONALITY OF THE EULER CONSTANT

ON THE CONVERGENCE OF SOME ALTERNATING SERIES

Series of Error Terms for Rational Approximations of Irrational Numbers

On Gauss sums and the evaluation of Stechkin s constant

On the distribution of consecutive square-free

Notes on Systems of Linear Congruences

PRIME NUMBERS YANKI LEKILI

Irrationality exponent and rational approximations with prescribed growth

MATH 271 Summer 2016 Practice problem solutions Week 1

Waring s problem, the declining exchange rate between small powers, and the story of 13,792

not to be republished NCERT REAL NUMBERS CHAPTER 1 (A) Main Concepts and Results

Journal of Combinatorics and Number Theory 1(2009), no. 1, ON SUMS OF PRIMES AND TRIANGULAR NUMBERS. Zhi-Wei Sun

Squares in products with terms in an arithmetic progression

GOLDBACH S PROBLEMS ALEX RICE

The Circle Method. Basic ideas

A HYBRID OF TWO THEOREMS OF PIATETSKI-SHAPIRO

Notes on Equidistribution

On the Subsequence of Primes Having Prime Subscripts

Solutions to Practice Final

SOME REMARKS ON ARTIN'S CONJECTURE

1 + lim. n n+1. f(x) = x + 1, x 1. and we check that f is increasing, instead. Using the quotient rule, we easily find that. 1 (x + 1) 1 x (x + 1) 2 =

Cullen Numbers in Binary Recurrent Sequences

Solving a linear equation in a set of integers II

A Smorgasbord of Applications of Fourier Analysis to Number Theory

On the existence of primitive completely normal bases of finite fields

Divisibility. 1.1 Foundations

Small gaps between primes

Goldbach and twin prime conjectures implied by strong Goldbach number sequence

On the number of elements with maximal order in the multiplicative group modulo n

E-SYMMETRIC NUMBERS (PUBLISHED: COLLOQ. MATH., 103(2005), NO. 1, )

Denser Egyptian fractions

Estimates for sums over primes

The arithmetic of binary representations of even positive integer 2n and its application to the solution of the Goldbach s binary problem

Carmichael numbers with a totient of the form a 2 + nb 2

Math Circle Beginners Group February 28, 2016 Euclid and Prime Numbers

A Remark on Sieving in Biased Coin Convolutions

PATTERNS OF PRIMES IN ARITHMETIC PROGRESSIONS

arxiv:math/ v1 [math.nt] 20 May 2004

Products of ratios of consecutive integers

4 Powers of an Element; Cyclic Groups

NOTES Edited by William Adkins. On Goldbach s Conjecture for Integer Polynomials

NORMAL NUMBERS AND UNIFORM DISTRIBUTION (WEEKS 1-3) OPEN PROBLEMS IN NUMBER THEORY SPRING 2018, TEL AVIV UNIVERSITY

Mathematics 228(Q1), Assignment 2 Solutions

The dichotomy between structure and randomness. International Congress of Mathematicians, Aug Terence Tao (UCLA)

1 Basic Combinatorics

Integers without large prime factors in short intervals and arithmetic progressions

Large Sieves and Exponential Sums. Liangyi Zhao Thesis Director: Henryk Iwaniec

On Roth's theorem concerning a cube and three cubes of primes. Citation Quarterly Journal Of Mathematics, 2004, v. 55 n. 3, p.

= 1 2x. x 2 a ) 0 (mod p n ), (x 2 + 2a + a2. x a ) 2

AN INVITATION TO ADDITIVE PRIME NUMBER THEORY. A. V. Kumchev, D. I. Tolev

Constructions of digital nets using global function fields

Arithmetic progressions in sumsets

3 The language of proof

A family of quartic Thue inequalities

arxiv:math/ v2 [math.gm] 4 Oct 2000

JEAN-MARIE DE KONINCK AND IMRE KÁTAI

arxiv: v1 [math.nt] 5 Mar 2015

Analytic Number Theory

On prime factors of subset sums

. As the binomial coefficients are integers we have that. 2 n(n 1).

The density of integral solutions for pairs of diagonal cubic equations

CHAPTER 1 REAL NUMBERS KEY POINTS

arxiv: v2 [math.nt] 4 Nov 2012

Roth s Theorem on Arithmetic Progressions

Incomplete exponential sums over finite fields and their applications to new inversive pseudorandom number generators

A New Sieve for the Twin Primes

On the distribution of lattice points on spheres and level surfaces of polynomials.

NOTES ON SIMPLE NUMBER THEORY

BURGESS INEQUALITY IN F p 2. Mei-Chu Chang

Prime Number Theory and the Riemann Zeta-Function

The Diophantine equation x n = Dy 2 + 1

ON SIMULTANEOUS DIAGONAL INEQUALITIES, II

A Remark on Sieving in Biased Coin Convolutions

Lecture 1: Small Prime Gaps: From the Riemann Zeta-Function and Pair Correlation to the Circle Method. Daniel Goldston

Florian Luca Instituto de Matemáticas, Universidad Nacional Autonoma de México, C.P , Morelia, Michoacán, México

Fermat numbers and integers of the form a k + a l + p α

2 Arithmetic. 2.1 Greatest common divisors. This chapter is about properties of the integers Z = {..., 2, 1, 0, 1, 2,...}.

The ranges of some familiar arithmetic functions

#A42 INTEGERS 10 (2010), ON THE ITERATION OF A FUNCTION RELATED TO EULER S

On the Density of Sequences of Integers the Sum of No Two of Which is a Square II. General Sequences

ON PRIMES IN QUADRATIC PROGRESSIONS & THE SATO-TATE CONJECTURE

Notations. Primary definition. Specific values. Traditional name. Traditional notation. Mathematica StandardForm notation. Specialized values

Les chiffres des nombres premiers. (Digits of prime numbers)

PARABOLAS INFILTRATING THE FORD CIRCLES SUZANNE C. HUTCHINSON THESIS

INTEGERS DIVISIBLE BY THE SUM OF THEIR PRIME FACTORS

REAL NUMBERS. Any positive integer a can be divided by another positive integer b in such a way that it leaves a remainder r that is smaller than b.

A NEW UPPER BOUND FOR FINITE ADDITIVE BASES

On Gelfond s conjecture on the sum-of-digits function

Transcription:

INTEGERS: ELECTRONIC JOURNAL OF COMBINATORIAL NUMBER THEORY 8 (2008), #A08 ON SUMS OF PRIMES FROM BEATTY SEQUENCES Angel V. Kumchev 1 Department of Mathematics, Towson University, Towson, MD 21252-0001, U.S.A. akumchev@towson.edu Received: 6/7/07, Accepted: 1/28/08, Published: 2/6/08 Abstract Let k 2 and α 1, β 1,..., α k, β k be reals such that the α i s are irrational and greater than 1. Suppose further that some ratio α i /α j is irrational. We study the representations of an integer n in the form p 1 + p 2 + + p k = n, where p i is a prime from the Beatty sequence B i = {n N : n = [α i m + β i ] for some m Z}. 1. Introduction Ever since the days of Euler and Goldbach, number-theorists have been fascinated by additive representations of the integers as sums of primes. The most famous result in this field is I.M. Vinogradov s three primes theorem [7], which states that every sufficiently large odd integer is the sum of three primes. Over the years, a number of authors have studied variants of the three primes theorem with prime numbers restricted to various sequences of arithmetic interest. For instance, a recent work by Banks, Güloğlu and Nevans [1] studies the question of representing integers as sums of primes from a Beatty sequence. Suppose that α and β are real numbers, with α > 1 and irrational. The Beatty sequence B α,β is defined by B α,β = {n N : n = [αm + β] for some m Z}. (Henceforth, [θ] represents the integer part of the real number θ.) Banks et al. proved that if k 3, then every sufficiently large integer n k (mod 2) can be expressed as the sum of k primes from the sequence B α,β, provided that α < k and α has a finite type (see below). In their closing remarks, the authors of [1] note that their method can be used to extend the main results of [1] to representations of an integer n in the form p 1 + p 2 + + p k = n, (1) where p i B α,βi. However, they remark that for a sequence α 1,..., α k of irrational numbers greater than 1, it appears to be much more difficult to estimate the number of representations 1 The author s research was partially supported by a Towson University summer research fellowship.

INTEGERS: ELECTRONIC JOURNAL OF COMBINATORIAL NUMBER THEORY 8 (2008), #A08 2 of n k (mod 2) in the form (1), where p i lies in the Beatty sequence B αi,β i. The main purpose of the present note is to address the latter question in the case when at least one of the ratios α i /α j, 1 i, j k, is irrational. Let α 1, β 1,..., α k, β k, k 2, be real numbers, and suppose that α 1,..., α k are irrational and greater than 1. For i = 1,..., k, we denote by B i the Beatty sequence B αi,β i. We write R(n) = R(n; α, β) = (log p 1 ) (log p k ), (2) p 1 + +p k =n p i B i where the summation is over the solutions of (1) in prime numbers p 1,..., p k such that p i B i. Similarly to [1], we shall use the Hardy Littlewood circle method to obtain an asymptotic formula for R(n). The circle method requires some quantitative measure of the irrationality of the α i s in the form of hypotheses on the rational approximations to the α i s. Let θ denote the distance from the real number θ to the nearest integer. We say that an s-tuple θ 1,..., θ s of real numbers is of a finite type, if there exists a real number η such that the inequality q 1 θ 1 + + q s θ s < max(1, q 1,..., q s ) η (3) has only finitely many solutions in q 1,..., q s Z. In particular, the reals in an s-tuple of a finite type are irrational and linearly independent over Q. Our main result can now be stated as follows. Theorem 1. Let k 3 and let α 1, β 1,..., α k, β k be real numbers, with α 1,..., α k > 1. Suppose that each individual α i is of a finite type and that at least one pair α 1 i, α 1 j is also of a finite type. Then, for any fixed A > 0 and any sufficiently large integer n, one has R(n; α, β) = S k (n)n k 1 α 1 α k (k 1)! + O( n k 1 (log n) A), (4) where R(n; α, β) is the quantity defined in (2) and S k (n) is given by S k (n) = ) ) (1 + (1 ( 1)k + ( 1)k+1. (5) (p 1) k 1 (p 1) k p n p n The implied constant in (4) depends at most on A, α, β. Since 1 S k (n) 1 when n k (mod 2), Theorem 1 has the following direct consequence. Corollary 1. Let k 3 and suppose that α 1, β 1,..., α k, β k are real numbers subject to the hypotheses of Theorem 1. Then, every sufficiently large integer n k (mod 2) can be represented in the form (1) with p i B i, 1 i k. After some standard adjustments, the techniques used in the proof of Theorem 1 yield also the following result on sums of two Beatty primes.

INTEGERS: ELECTRONIC JOURNAL OF COMBINATORIAL NUMBER THEORY 8 (2008), #A08 3 Theorem 2. Suppose that α 1, β 1, α 2, β 2 are real numbers, with α 1, α 2 > 1. Suppose further that the pair α1 1, α2 1 is of a finite type. Then, for any fixed A > 0, and all but O(x(log x) A ) integers n x, one has R(n; α, β) = (α 1 α 2 ) 1 S 2 (n)n + O ( n(log n) A), where R(n; α, β) is the quantity defined in (2) and S 2 (n) is given by (5) with k = 2. The implied constants depend at most on A, α, β. Since for even n, 1 S 2 (n) log log n, we have the following corollary to Theorem 2. Corollary 2. Suppose that α 1, β 1, α 2, β 2 are real numbers subject to the hypotheses of Theorem 2. Then, for any fixed A > 0, all but O(x(log x) A ) even integers n x can be represented as sums of a prime p 1 B 1 and a prime p 2 B 2. By making some adjustments in the proof of Theorem 2, we can call upon a celebrated theorem by Montgomery and Vaughan [5] to improve on Corollary 2. Corollary 3. Suppose that α 1, β 1, α 2, β 2 are real numbers subject to the hypotheses of Theorem 2. Then there exists an ε = ε(α) > 0 such that all but O(x 1 ε ) even integers n x can be represented as sums of a prime p 1 B 1 and a prime p 2 B 2. Comparing Theorems 1 and 2 with the main results in [1], one notes that our theorems include no hypotheses similar to the condition α < k required in [1]. The latter condition is necessary in the case α 1 = = α k = α, if all large integers n k (mod 2) are to be represented. However, it can be dispensed with when some pair α i, α j is linearly independent over Q. It seems that the natural hypotheses for the above theorems are that all α i s and some ratio α i /α j be irrational, but such generality is beyond the reach of our method. The finite type conditions above approximate these natural hypotheses without being too restrictive. For example, by a classical theorem of Khinchin s [4], almost all (in the sense of Lebesgue measure) real numbers are of a finite type. 2. Preliminaries 2.1. Notation For a real number θ, [θ], {θ} and θ denote, respectively, the integer part of θ, the fractional part of θ and the distance from θ to the nearest integer; also, e(θ) = e 2πiθ. For integers a and b, we write (a, b) and [a, b] for the greatest common divisor and the least common multiple of a and b. The letter p, with or without indices, is reserved for prime numbers. Finally, if x = (x 1,..., x s ), we write x = max( x 1,..., x s ).

INTEGERS: ELECTRONIC JOURNAL OF COMBINATORIAL NUMBER THEORY 8 (2008), #A08 4 2.2. For i = 1,..., k, we set γ i = α 1 i and δ i = α 1 i (1 β i ). It is not difficult to see that m B i if and only if 0 < {γ i m + δ i } < γ i. Thus, the characteristic function of the Beatty sequence B i is g i (γ i m + δ i ), where g i is the 1-periodic extension of the characteristic function of the interval (0, γ i ). Our analysis will require smooth approximations to g i. Suppose that 1 i k and is a real such that 0 < < 1 4 min(γ i, 1 γ i ). Then there exist 1-periodic C -functions g ± i such that: i) 0 g i (x) g i(x) g + i (x) 1 for all real x; ii) g ± i (x) = g i(x) when x γ i or γ i + x 1 ; iii) d r g ± i (x) dx r r r for r = 1, 2,.... Furthermore, the Fourier coefficients ĝ ± i (m) of g± i ĝ ± i (0) = γ i + O( ), ĝ ± i (m) r satisfy the bounds 1 r (1 + m ) r (r = 1, 2,... ), (6) where the latter bound follows from ii) and iii) above via partial integration. 2.3. The proofs of Theorems 1 and 2 use the following generalization of the classical bound for exponential sums over primes in Vaughan [6, Theorem 3.1]. Lemma 1. Suppose that α is real and a, q are integers, with (a, q) = 1 and q N. Then (log p)e(αp) ( Nq 1/2 + N 4/5 + (Nq) 1/2) ( 1 + q 2 θ ) (log 2N) 4, p N where θ = α a/q. The proof of the above lemma is essentially the same as that of [6, Theorem 3.1], which is the case θ q 2. The only adjustment one needs to make in the argument in [6] is to replace [6, Lemma 2.2] by the following variant. Lemma 2. Suppose that α, X, Y are real with X 1, Y 1, and a, q are integers with (a, q) = 1. Then min ( XY x 1, αx 1) ( XY q 1 + X + q ) ( 1 + q 2 θ ) (log 2Xq), x X where θ = α a/q.

INTEGERS: ELECTRONIC JOURNAL OF COMBINATORIAL NUMBER THEORY 8 (2008), #A08 5 2.4. In the next lemma, we use the finite type of an s-tuple θ 1,..., θ s to obtain rational approximations to linear combinations of θ 1,..., θ s. Lemma 3. Suppose that the s-tuple θ 1,..., θ s has a finite type and let η > 1 be such that (3) has finitely many solutions. Let 0 < ε < (2η) 1, let Q be sufficiently large, and let m = (m 1,..., m s ) Z s, with 0 < m Q ε. Then there exist integers a and q such that q(m 1 θ 1 + + m s θ s ) a Q 1, Q ε q Q, (a, q) = 1. Proof. Since the sum m 1 θ 1 + + m s θ s is irrational, it has an infinite continued fraction. Let q and q be the denominators of two consecutive convergents to that continued fraction, such that q Q < q. By the properties of continued fractions, q(m 1 θ 1 + + m s θ s ) (q ) 1 < Q 1. If 1 q Q ε, then Q 1 (q m ) 1/(2ε), and hence q(m 1 θ 1 + + m s θ s ) < (q m ) 1/(2ε) which contradicts the choice of η and ε. This completes the proof. 3. Proof of Theorem 1 Without loss of generality, we may assume that the pair of a finite type in the hypotheses of the theorem is α1 1, α2 1. We also note that if α i has a finite type, then so does γ i = α 1 i. Indeed, let q be a sufficiently large solution of qγ i < q η and let a be the nearest integer to qγ i. Then q a q and aα i q < α i q η a η < a η+ε. Thus, if qγ i < q η has an infinite number of solutions in positive integers q, then so does qα i < q η+ε. Recall the functions g i and g ± i given by We have R(n) = p 1 + +p k =n described in 2.2. We shall use those functions with = (log n) A. (7) (log p 1 ) (log p k )g 1 (γ 1 p 1 + δ 1 ) g k (γ k p k + δ k ),

INTEGERS: ELECTRONIC JOURNAL OF COMBINATORIAL NUMBER THEORY 8 (2008), #A08 6 so by the construction of g ± i, R (n) R(n) R + (n), (8) where R ± (n) = p 1 + +p k =n (log p 1 ) (log p k )g ± 1 (γ 1 p 1 + δ 1 ) g ± k (γ kp k + δ k ). We now proceed to evaluate the sums R + (n) and R (n). We shall focus on R + (n), the evaluation of R (n) being similar. Substituting the Fourier expansions of g + 1,..., g + k into the definition of R+ (n), we obtain R + (n) = m Z k ĝ + 1 (m 1 ) ĝ + k (m k)e(δ 1 m 1 + + δ k m k )R(n, m), (9) where m = (m 1,..., m k ) and R(n, m) = (log p 1 ) (log p k )e(γ 1 m 1 p 1 + + γ k m k p k ). p 1 + +p k =n We note for the record that when m = 0, we have R(n, 0) = S k(n)n k 1 (k 1)! + O ( n k 1 (log n) A). (10) When k = 3, this is due to Vinogradov [7] (see also Vaughan [6, Theorem 3.4]), and the result for k 4 can be proved similarly (see Hua [3]). We now set Combining (6), (7), (9) and (10), we deduce that M = 1 (log n) = (log n) A+1. (11) R + (n) = γ 1 γ k (k 1)! S k(n)n k 1 + O ( n k 1 (log n) A) + O(Σ 1 + Σ 2 ), (12) where Σ 1 = ĝ 1 + (m 1 ) ĝ + k (m k) R(n, m), 0< m M Σ 2 = ĝ 1 + (m 1 ) ĝ + k (m k) R(n, m). m >M We may use (6) to estimate Σ 2. It follows easily from the second bound in (6) that ĝ + i (m) m Z m M 1 1 + m + m >M 1 log M. (13) (1 + m ) 2

INTEGERS: ELECTRONIC JOURNAL OF COMBINATORIAL NUMBER THEORY 8 (2008), #A08 7 Hence, by (10), (11) and (6) with r = [A] + 3, Σ 2 R(n, 0)(log M) k 1 1 i k m >M ĝ + i n k 1 (log n)( M) 1 r n k 1 (log n) A. (m) (14) Next, we use a variant of the circle method to bound R(n, m) when 0 < m M. Define the exponential sum S(ξ) = p n(log p)e(ξp). By orthogonality, Put R(n, m) = 1 For j = 1,..., k, we write λ j = λ j (m) = γ j m j γ 1 m 1. Then R(n, m) = e(γ 1 m 1 n) 0 S(ξ + γ 1 m 1 ) S(ξ + γ k m k )e( nξ) dξ. (15) P = (log n) 2A+12, Q = np 1. (16) 1+1/Q 1/Q S(ξ)S(ξ + λ 2 ) S(ξ + λ k )e( nξ) dξ. We partition the interval [1/Q, 1 + 1/Q) into Farey arcs of order Q and write M(q, a) for the arc containing the Farey fraction a/q: if a /q and a /q are, respectively, the left and right neighbors of a/q in the Farey sequence, then [ a + a M(q, a) = q + q, a + ) a. q + q Thus, R(n, m) q Q 1 a q (a,q)=1 M(q,a) S(ξ)S(ξ + λ 2 ) S(ξ + λ k ) dξ. (17) When ξ M(q, a), with P < q Q, Lemma 1 yields S(ξ) np 1/2 (log n) 4 n(log n) A 2. Inserting this bound into the right side of (17), we obtain R(n, m) S(ξ)S(ξ + λ 2 ) S(ξ + λ k ) dξ + Σ 3, (18) q P M(q,a) 1 a q (a,q)=1

INTEGERS: ELECTRONIC JOURNAL OF COMBINATORIAL NUMBER THEORY 8 (2008), #A08 8 where Σ 3 n(log n) A 2 P <q Q 1 a q (a,q)=1 M(q,a) 1 n k 2 (log n) A 2 S(ξ + λ 2 )S(ξ + λ 3 ) dξ. By the Cauchy Schwarz inequality and Parseval s identity, 1 so we deduce from (19) that 0 0 S(ξ + λ 2 ) S(ξ + λ k ) dξ (19) S(ξ + λ i )S(ξ + λ j ) dξ n(log n) (1 i, j k), (20) Σ 3 n k 1 (log n) A 1. (21) To estimate the remaining sum on the right side of (18), we consider separately the cases m 1 = 0 and m 1 0. Case 1: m 1 = 0. Since m 0, we have m i 0 for some i = 2,..., k. By Lemma 3 with s = 1 and θ 1 = γ i, there exist an ε > 0 and integers b and r such that r(m i γ i ) b < Q 1/2, Q ε r Q 1/2, (b, r) = 1. Suppose that ξ M(q, a), where 1 q P. It follows that ξ + λ i a q b r 1 qq + 1 rq 2 1/2 rq. 1/2 Let the integers a 1, q 1 be such that a 1 q 1 = a q + b r, (a 1, q 1 ) = 1. Then q 1 divides [q, r] and is divisible by [q, r]/(q, r), so ξ + λ i a 1 2P q 1 Q, rp 1 q 1/2 1 rp, (a 1, q 1 ) = 1. Thus, Lemma 1 yields q 1 S(ξ + λ i ) ( nq 1/2 1 + n 4/5 + (nq 1 ) 1/2) ( 1 + q 1 P Q 1/2) (log n) 4 (22) ( nq 1/2 1 + n 4/5 P ) (log n) 4 n 1 ε/3. Combining (22) and (20), we easily get S(ξ)S(ξ + λ 2 ) S(ξ + λ k ) dξ n k 1 ε/4. (23) q P 1 a q (a,q)=1 M(q,a)

INTEGERS: ELECTRONIC JOURNAL OF COMBINATORIAL NUMBER THEORY 8 (2008), #A08 9 Case 2: m 1 0. Then we apply Lemma 3 to the pair γ 1, γ 2. It follows that there exist an ε > 0 and integers b and r such that r(m 2 γ 2 m 1 γ 1 ) b < Q 1/2, Q ε r Q 1/2, (b, r) = 1. Suppose that ξ M(q, a), where 1 q P. Arguing similarly to Case 1, we find that there exist integers a 1, q 1 such that ξ + λ 2 a 1 2P q 1 Q, rp 1 q 1/2 1 rp, (a 1, q 1 ) = 1. q 1 Using this rational approximation to ξ + λ 2, we can now apply Lemma 1 to show that S(ξ + λ 2 ) ( nq 1/2 1 + n 4/5 P ) (log n) 4 n 1 ε/3. We then derive (23) in a similar fashion to Case 1. We conclude that (23) holds for all vectors m with 0 < m M. Together, (18), (21) and (23) yield R(n, m) n k 1 (log n) A 1 for all 0 < m M, whence Finally, from (12), (14) and (24), Σ 1 n k 1 (log n) A 1 (log M) k n k 1 (log n) A. (24) R + (n) = γ 1 γ k (k 1)! S k(n)n k 1 + O ( n k 1 (log n) A). Since an analogous asymptotic formula holds for R (n), the conclusion of the theorem follows from (8). 4. Sketch of the Proof of Theorem 2 Let R + (n) and R (n) be the quantities defined in 3 with k = 2. To prove Theorem 2 it suffices to establish the inequality R ± (n) γ 1 γ 2 S 2 (n)n 2 x 3 (log x) 3A. (25) n x As in the proof of Theorem 1, we focus on the proof of the inequality for R + (n), the proof of the other inequality being similar. We use the notation introduced in 3 with k = 2 and A replaced by 2A + 1. When k = 2, the asymptotic formula (10) is not known, but we do have the upper bound (see [2]) R(n, 0) n ( ) p n log log n. p 1 p n

INTEGERS: ELECTRONIC JOURNAL OF COMBINATORIAL NUMBER THEORY 8 (2008), #A08 10 This bound suffices to show similarly to (9) (14) that R + (n) = γ 1 γ 2 R(n, 0) + O ( ) n(log n) 2A + Σ 1. Furthemore, by [6, Theorem 3.7], R(n, 0) S 2 (n)n 2 x 3 (log x) 3A. n x Thus, (25) for R + (n) follows from the inequality ĝ 1 + (m 1 )ĝ 2 + (m 2 ) R(n, m) n x 0< m M By (13) and Cauchy s inequality, the left side of (26) is (log M) 4 max R(n, m) 2, 0< m M n x 2 x 3 (log x) 3A. (26) so it suffices to show that R(n, m) 2 x(log x) 3A 1 (27) n x for all m, 0 < m M. By (15) and Bessel s inequality, R(n, m) 2 1 n x 0 S(ξ + γ 1 m 1 )S(ξ + γ 2 m 2 ) 2 dξ, (28) where the definition of the exponential sum S(ξ) has been altered to S(ξ) = p)e(ξp). p x(log We set and obtain similarly to (17) that As in 3, 1 0 P = (log x) 3A+10, Q = xp 1, (29) S(ξ + γ 1 m 1 )S(ξ + γ 2 m 2 ) 2 dξ q Q 1 a q (a,q)=1 M(q,a) min ( S(ξ), S(ξ + λ 2 ) ) xp 1/2 (log x) 4 for all ξ, so we deduce from (20), (29) and (30) that 1 0 S(ξ)S(ξ + λ 2 ) 2 dξ. (30) S(ξ + γ 1 m 1 )S(ξ + γ 2 m 2 ) 2 dξ x 3 P 1 (log x) 9 x 2 (log x) 3A 1. Inserting the last bound into (28), we obtain (27).

INTEGERS: ELECTRONIC JOURNAL OF COMBINATORIAL NUMBER THEORY 8 (2008), #A08 11 5. Closing Remarks In the proof of Theorem 1, we essentially showed that R(n) = γ 1 γ k R(n, 0) + error terms, (31) and then chose the parameters, M, P, Q so that the error terms were n k 1 (log n) A. It is possible to alter the above choices so that the error terms in (31) are n k 1 ε for some ε > 0 which depends only on the α i s. For example, if η > 1 is such that each of the inequalities q 1 α 1 1 + q 2 α 1 2 < max(1, q 1, q 2 ) η, qα i < q η (1 i k), has finitely many solutions, then one may choose = n ε, M = 1 n ε = n 2ε, P = n 3ε, Q = np 1, with 0 < ε < (20η) 1. Thus, the quality of the error term in (4) is determined solely by the quality of the error term in the asymptotic formula (10) for the number of representations of an integer n as the sum of k primes. However, since no improvements on (10) are known, the improved bounds for the error terms in (31) have no effect on Theorem 1. Similarly, when k = 2, a slight alteration of our choices in 4 yields the bound R(n) = γ 1 γ 2 R(n, 0) + O ( n 1 ε) (32) for all but O(x 1 ε ) values of n x. In this case, however, such a variation has a tangible effect: it yields Corollary 3. Indeed, by a well-known result of Montgomery and Vaughan [5], there is an absolute constant ω < 1 such that the right side of (32) is positive for all but O(x ω ) even integers n x. Finally, a comment regarding our finite type hypotheses. We say that an s-tuple θ 1,..., θ s of real numbers is of subexponential type, if for each fixed η > 0, the inequality q 1 θ 1 + + q s θ s < exp( q η ) has only finitely many solutions q = (q 1,..., q s ) Z s. Clearly, every s-tuple of a finite type is also of subexponential type, but not vice versa. It takes little effort to check that in the arguments in 3 and 4, it suffices to assume that each α i and some pair α 1 i, α 1 j are of subexponential type. Thus, our method reaches some Beatty sequences B α,β with α of an infinite type. On the other hand, under the weaker subexponential type hypotheses, we no longer have the improved remainder estimates in (31) and (32). In particular, we no longer have Corollary 3 (at least, not by the simple argument sketched above). This seems to be too steep a price to pay for such a modest gain in generality.

INTEGERS: ELECTRONIC JOURNAL OF COMBINATORIAL NUMBER THEORY 8 (2008), #A08 12 References [1] W. D. Banks, A. M. Güloğlu, and C. W. Nevans, Representations of integers as sums of primes from a Beatty sequence, preprint, arxiv:math/0701285v1, 2007. [2] H. Halberstam and H.-E. Richert, Sieve Methods, Academic Press, 1974. [3] L. K. Hua, Additive Theory of Prime Numbers, American Mathematical Society, 1965. [4] A. Ya. Khinchin, Zur metrischen Theorie der Diophantischen Approximationen, Math. Z. 24 (1926), 706 714. [5] H. L. Montgomery and R. C. Vaughan, The exceptional set in Goldbach s problem, Acta Arith. 27 (1975), 353 370. [6] R. C. Vaughan, The Hardy Littlewood Method, 2nd ed., Cambridge University Press, 1997. [7] I. M. Vinogradov, Representation of an odd number as the sum of three primes, Dokl. Akad. Nauk SSSR 15 (1937), 291 294, in Russian.