Loading [MathJax]/jax/element/mml/optable/MathOperators.js
Research article

The least common multiple of consecutive terms in a cubic progression

  • Received: 30 October 2019 Accepted: 07 February 2020 Published: 14 February 2020
  • MSC : 11B25, 11N13, 11A05

  • Let k be a positive integer and f(x) a polynomial with integer coefficients. Associated to the least common multiple lcm0ik{f(n+i)}, we define the function Gk,f for all positive integers nNZk,f by Gk,f(n):=ki=0|f(n+i)|lcm0ik{f(n+i)}, where Zk,f:=ki=0{nN:f(n+i)=0}. If f(x)=x, then Farhi showed in 2007 that Gk,f is periodic with k! as its period. Consequently, Hong and Yang improved Farhi's period k! to lcm(1,...,k). Later on, Farhi and Kane confirmed a conjecture of Hong and Yang and determined the smallest period of Gk,f. For the general linear polynomial f(x), Hong and Qian showed in 2011 that Gk,f is periodic and got a formula for its smallest period. In 2015, Hong and Qian characterized the quadratic polynomial f(x) such that Gk,f is almost periodic and also arrived at an explicit formula for the smallest period of Gk,f. If degf(x)3, then one naturally asks the following interesting question: Is the arithmetic function Gk,f almost periodic and, if so, what is the smallest period? In this paper, we asnwer this question for the case f(x)=x3+2. First of all, with the help of Hua's identity, we prove that Gk,x3+2 is periodic. Consequently, we use Hensel's lemma, develop a detailed p-adic analysis to Gk,x3+2 and particularly investigate arithmetic properties of the congruences x3+20(modpe) and x6+1080(modpe), and with more efforts, its smallest period is finally determined. Furthermore, an asymptotic formula for log lcm0ik{(n+i)3+2} is given.

    Citation: Zongbing Lin, Shaofang Hong. The least common multiple of consecutive terms in a cubic progression[J]. AIMS Mathematics, 2020, 5(3): 1757-1778. doi: 10.3934/math.2020119

    Related Papers:

    [1] Guangyan Zhu, Kaimin Cheng, Wei Zhao . Notes on Hong's conjecture on nonsingularity of power LCM matrices. AIMS Mathematics, 2022, 7(6): 10276-10285. doi: 10.3934/math.2022572
    [2] Naif Alotaibi, A. S. Al-Moisheer, Ibrahim Elbatal, Salem A. Alyami, Ahmed M. Gemeay, Ehab M. Almetwally . Bivariate step-stress accelerated life test for a new three-parameter model under progressive censored schemes with application in medical. AIMS Mathematics, 2024, 9(2): 3521-3558. doi: 10.3934/math.2024173
    [3] Kai Wang, Guicang Zhang . Curve construction based on quartic Bernstein-like basis. AIMS Mathematics, 2020, 5(5): 5344-5363. doi: 10.3934/math.2020343
    [4] Yulu Feng, Min Qiu . Some results on p-adic valuations of Stirling numbers of the second kind. AIMS Mathematics, 2020, 5(5): 4168-4196. doi: 10.3934/math.2020267
    [5] Keqiang Li, Shangjiu Wang . Multiple periodic solutions of nonlinear second order differential equations. AIMS Mathematics, 2023, 8(5): 11259-11269. doi: 10.3934/math.2023570
    [6] Hu Jiayuan, Chen Zhuoyu . On distribution properties of cubic residues. AIMS Mathematics, 2020, 5(6): 6051-6060. doi: 10.3934/math.2020388
    [7] Yan Yan . Multiplicity of positive periodic solutions for a discrete impulsive blood cell production model. AIMS Mathematics, 2023, 8(11): 26515-26531. doi: 10.3934/math.20231354
    [8] Shuangnian Hu, Rongquan Feng . The number of solutions of cubic diagonal equations over finite fields. AIMS Mathematics, 2023, 8(3): 6375-6388. doi: 10.3934/math.2023322
    [9] Long Chen, Shaofang Hong . Divisibility among determinants of power matrices associated with integer-valued arithmetic functions. AIMS Mathematics, 2020, 5(3): 1946-1959. doi: 10.3934/math.2020130
    [10] Yonghang Chang, Menglan Liao . Global well-posedness and scattering of the four dimensional cubic focusing nonlinear Schrödinger system. AIMS Mathematics, 2024, 9(9): 25659-25688. doi: 10.3934/math.20241254
  • Let k be a positive integer and f(x) a polynomial with integer coefficients. Associated to the least common multiple lcm0ik{f(n+i)}, we define the function Gk,f for all positive integers nNZk,f by Gk,f(n):=ki=0|f(n+i)|lcm0ik{f(n+i)}, where Zk,f:=ki=0{nN:f(n+i)=0}. If f(x)=x, then Farhi showed in 2007 that Gk,f is periodic with k! as its period. Consequently, Hong and Yang improved Farhi's period k! to lcm(1,...,k). Later on, Farhi and Kane confirmed a conjecture of Hong and Yang and determined the smallest period of Gk,f. For the general linear polynomial f(x), Hong and Qian showed in 2011 that Gk,f is periodic and got a formula for its smallest period. In 2015, Hong and Qian characterized the quadratic polynomial f(x) such that Gk,f is almost periodic and also arrived at an explicit formula for the smallest period of Gk,f. If degf(x)3, then one naturally asks the following interesting question: Is the arithmetic function Gk,f almost periodic and, if so, what is the smallest period? In this paper, we asnwer this question for the case f(x)=x3+2. First of all, with the help of Hua's identity, we prove that Gk,x3+2 is periodic. Consequently, we use Hensel's lemma, develop a detailed p-adic analysis to Gk,x3+2 and particularly investigate arithmetic properties of the congruences x3+20(modpe) and x6+1080(modpe), and with more efforts, its smallest period is finally determined. Furthermore, an asymptotic formula for log lcm0ik{(n+i)3+2} is given.


    The least common multiple of consecutive positive integers was investigated by Chebyshev for the first significant attempt to prove the prime number theorem in [2]. Since then, the least common multiple of any given sequence of positive integers has been an interesting and important topic. Hanson [9] and Nair [20] got the upper and lower bound of lcm1in{i}, respectively. Then Nair's lower bound was extended in [3,4,5,10,11,14,17,22,25]. Goutziers [8] investigated the asymptotic behavior on the least common multiple of a set of integers not exceeding N. Bateman, Kalb and Stenger [1] obtained an asymptotic estimate for the least common multiple of arithmetic progressions. Hong, Qian and Tan [15] got an asymptotic estimate for the least common multiple of the sequence of product of linear polynomials. Qian and Hong [23] studied the asymptotic behavior of the least common multiple of consecutive terms in arithmetic progressions. In [6], Farhi found an interesting relation between the least common multiple and the derivative of integr-valued polynomials. Actually, Farhi [6] showed that lcm1in{i} is the smallest positive integer cn satisfying the property: For any P(x)En, one has cnP(x)En, where En is the set of all the integer-valued polynomials of degree no more than n.

    Let k be a positive integer and f(x) be a polynomial with integer coefficients. Associated to the least common multiple lcm0ik{f(n+i)} of any k+1 consecutive terms in the sequence {f(n)}n=1, we define the arithmetic function Gk,f for all positive integers nNZk,f by

    Gk,f(n):=ki=0|f(n+i)|lcm0ik{f(n+i)},

    where

    Zk,f:=ki=0{nN:f(n+i)=0}.

    If f(x)=x, then Farhi [4] showed that Gk,f is periodic with k! as its period. Consequently, Hong and Yang [16] improved Farhi's period k! to lcm(1,...,k). Later on, Farhi and Kane [7] confirmed the Hong-Yang conjecture in [16] and determined the smallest period of Gk,f. For the general linear polynomial f(x), Hong and Qian [12] showed that Gk,f is periodic and got a formula for its smallest period. In 2012, Qian, Tan and Hong [24] proved that if f(x)=x2+1, then Gk,f is periodic with determining its exact period. In 2015, Hong and Qian [13] characterized the quadratic polynomial f(x) such that Gk,f is almost periodic and also arrived at an explicit formula for the smallest period of Gk,f. One naturally asks the following problem: Let degf(x)3. Is the arithmetic function Gk,f almost periodic and, if so, what is the smallest period? This is a nontrivial and interesting question. In general, it is hard to answer it because few are known about the roots of higher degree congruences modulo prime powers that play key roles in this topic. Among the higher degree sequences, {n3+2}n=1 is the first example one would like to understand. For instance, a renowned conjecture in number theory states that there are infinitely many primes in the cubic sequence {n3+2}n=1.

    In this paper, our main goal is to study the above problem for the cubic sequence {n3+2}n=1. In what follows, for brevity, we write Gk:=Gk,f when f(x)=x3+2. That is, we have

    Gk(n):=ki=0((n+i)3+2)lcm0ik{(n+i)3+2}. (1.1)

    Assume that Gk is periodic and let Pk denote its smallest period. Now we can use Pk to give a formula for lcm0ik{(n+i)3+2} as follows: For any positive integer n, one has

    lcm0ik{(n+i)3+2}=ki=0((n+i)3+2)Gk(nPk),

    where nPk means the least positive integer congruent to n modulo Pk. So it is significant to determine the exact value of the smallest period Pk.

    Before stating the main results of this paper, let us briefly explain the basic new ideas and key techniques in this paper. To deal with the cubic sequence {n3+2}n=1, we have to develop furthermore the techniques presented in previous works [7,12,16,24] and [13] that are far not enough for the current case. In fact, we use Hua's identity [18] to show the periodicity and establish the factorization result to reduce the computation of the smallest period Pk to that of the local periods Pp,k. For the small primes p=2 and 3, we calculate directly Pp,k. For the prime p5, we consider two cases p1(mod6) and p5(mod6). The former case is much more difficult to treat than the latter one. More complicated analysis will be needed in the former case. But for both cases, one needs to study in detail the arithmetic properties of roots of the congruences x3+20(modpe) and x6+1080(modpe), where e1 is an integer and the famous Hensel's lemma is used to lift the roots of the two congruences from modulo p to modulo pe.

    To state our main results, as usual, for any prime number p, we let vp be the normalized p-adic valuation of Q, that is, vp(a)=b if pba. We also let gcd(a,b) denote the greatest common divisor of any integers a and b. For any real number x, by x we denote the largest integer no more than x, we denote by x the smallest integer no less than x. Further, by {x} we denote the fractional part of x, i.e. {x}:=xx. Obviously, 0{x}<1. For any positive integer k, we define the positive integers Rk and Qk by

    Rk:=lcm1ik{i(i6+108)} (1.2)

    and

    Qk:=2(1)k+123{k+13}Rk2v2(Rk)3v3(Rk)p|Rkp1(mod6)2p13 (1.3)

    Evidently, v_p(Q_k) = v_p(R_k) for any prime p \equiv 5 \pmod 6 . The main result of this paper can be stated as follows.

    Theorem 1.1. Let k be a positive integer. The arithmetic function \mathcal{G}_k is periodic, and its smallest period is equal to R_1 = 109 if k = 1 , and if k\ge 2 , then its smallest period equals Q_k except that v_p(k+1)\ge v_p(Q_k)\ge 1 for at most one prime p\ge 5 , in which case its smallest period is equal to \frac{Q_k}{p^{v_p(Q_k)}} .

    Obviously, Theorem 1.1 answers affirmatively the above question for the case where f(x) = x^3+2 . Moreover, from Theorem 1.1 one can easily deduce the following interesting asymptotic result.

    Theorem 1.2. Let k be any given positive integer. Then the following asymptotic formula holds:

    {\rm log \ lcm}_{0 \le i \le k}\{(n+i)^3+2\} \sim 3(k+1){\rm log}n {\ \ as \ n \to \infty}.

    This paper is organized as follows. In Section 2, we first use an identity of Hua [18] to show that the arithmetic function \mathcal{G}_k is periodic with R_k as a period of it. Then we factor the smallest period P_k of \mathcal{G}_k into the product of local period P_{p, k} with p running over all prime divisors of R_k , and determine the local periods P_{2, k} and P_{3, k} . Later on, with a little more effort, we show that Q_k is a period of \mathcal{G}_k (see Theorem 2.7 below). Subsequently, we develop in Section 3 a p -adic analysis of the periodic function \mathcal{G}_k , and determine the local period P_{p, k} of \mathcal{G}_k for the case p\equiv 5 \pmod 6 . In Section 4, we evaluate the local period P_{p, k} when p\equiv 1 \pmod 6 and 2^{\frac{p-1}{3}}\equiv 1 \pmod p . In the process, we need to explore the arithmetic properties of the congruences x^3+2\equiv 0\pmod {p^e} and x^6+108\equiv 0\pmod {p^e} . Particularly, we express the smallest positive root of x^6+108\equiv0\pmod{p^e} in the terms of the roots of x^3+2\equiv 0 \pmod {p^e} . In the final section, we provide the proofs of Theorems 1.1 and 1.2. Two examples are also given to illustrate the validity of our main result.

    In this section, by using a well-known theorem of Hua in [18], we first prove that \mathcal{G}_k is periodic. In the meantime, we also arrive at a nontrivial period of \mathcal{G}_k .

    Lemma 2.1. The arithmetic function \mathcal{G}_k is periodic, and R_k is a period of \mathcal{G}_k .

    Proof. For any positive integer n , it follows from Theorem 7.3 of [18] (see page 11 of [18]) that

    \mathcal{G}_k(n) = \prod\limits_{t = 1}^k\prod\limits_{0\le i_0 \lt ... \lt i_t\le k} \big(\gcd\big((n+i_0)^3+2, ..., (n+i_t)^3+2 \big)\big)^{(-1)^{t-1}}

    and

    \mathcal{G}_k(n+R_k) = \prod\limits_{t = 1}^k\prod\limits_{0\le i_0 \lt ... \lt i_t\le k} \big(\gcd\big((n+i_0+R_k)^3+2, ..., (n+i_t+R_k)^3+2 \big)\big)^{(-1)^{t-1}}.

    Claim that \mathcal{G}_k(n+R_k) = \mathcal{G}_k(n) from which one can read that \mathcal{G}_k is periodic with R_k as its period. It remains to show the claim that will be done in what follows. It is clear that the claim follows if one can prove the following identity holds for all integers i and j with 0\le i < j\le k :

    \begin{align} \gcd\big( (n+i+R_k)^3+2, (n+j+R_k)^3+2\big) = \gcd\big( (n+i)^3+2, (n+j)^3+2\big). \end{align} (2.1)

    One can easily check that

    u(n, i, j)\big((n+i)^3+2\big)-u(n, j, i)\big((n+j)^3+2\big) = (j-i)((j-i)^6+108),

    where u(n, i, j): = -4(i-j)^4-3(n+2j-i)((2n+i+j)(i-j)^2-6). So one has

    \gcd\big( (n+i)^3+2, (n+j)^3+2\big)|(j-i)((j-i)^6+108).

    It then follows from (j-i)((j-i)^6+108)|R_k that

    \begin{equation} \gcd\big( (n+i)^3+2, (n+j)^3+2\big)|R_k. \end{equation} (2.2)

    This implies that

    \gcd\big( (n+i)^3+2, (n+j)^3+2\big)|(n+i\pm R_k)^3+2

    and

    \gcd\big((n+i)^3+2, (n+j)^3+2\big)|(n+j\pm R_k)^3+2.

    One then derives that

    \begin{equation} \gcd\big((n+i)^3+2, (n+j)^3+2\big)| \gcd\big((n+i+R_k)^3+2, (n+j+R_k)^3+2\big) \end{equation} (2.3)

    and

    \begin{equation} \gcd\big((n+i)^3+2, (n+j)^3+2\big)| \gcd\big((n+i-R_k)^3+2, (n+j-R_k)^3+2\big). \end{equation} (2.4)

    On the other hand, replacing n by n+R_k in (2.4) gives us that

    \begin{equation} \gcd\big( (n+i+R_k)^3+2, (n+j+R_k)^3+2\big)|\gcd\big( (n+i)^3+2, (n+j)^3+2\big). \end{equation} (2.5)

    Therefore (2.1) follows immediately from (2.3) and (2.5). Hence the claim is proved.

    This completes the proof of Lemma 2.1.

    For any given prime p , we define the arithmetic function \mathcal{G}_{p, k} for any positive integer n by \mathcal{G}_{p, k}(n): = v_{p}(\mathcal{G}_k(n)). Since \mathcal{G}_k is a periodic function, \mathcal{G}_{p, k} is periodic for each prime p and P_k is a period of \mathcal{G}_{p, k} . Let P_{p, k} be the smallest period of \mathcal{G}_{p, k} . Then P_{p, k} is called p -adic period of \mathcal{G}_k . All the p -adic periods P_{p, k} are called local period of \mathcal{G}_k . The following result says that P_{p, k} is a power of p , and P_k equals the product of all the p -adic periods P_{p, k} with p being a prime divisor of R_k .

    Lemma 2.2. For any prime p , P_{p, k} divides p^{v_p(R_k)} . Furthermore, one has

    P_k = \prod\limits_{p| R_k}P_{p, k}.

    Proof. At first, we show that for any prime p , p^{v_p(R_k)} is a period of \mathcal{G}_{p, k} . To do so, it is enough to prove that for any given positive integer n and integers i and j with 0\le i < j\le k , one has

    \begin{equation} v_{p}({\rm gcd} ((n+i+p^{v_p(R_k)})^3+2, (n+j+p^{v_p(R_k)})^3+2)) = v_{p}({\rm gcd} ((n+i)^3+2, (n+j)^3+2)). \end{equation} (2.6)

    In the following we show that (2.6) is true. Clearly, (2.2) infers that v_{p}({\rm gcd} ((n+i)^3+2, (n+j)^3+2))\le v_p(R_k) . In other words, we have v_p((n+i)^3+2)\le v_p(R_k) {\rm \ or \ }v_p((n+j)^3+2)\le v_p(R_k). From this one can deduce that v_p((n+i)^3+2)\le v_p((n+i\pm p^{v_p(R_k)})^3+2) or v_p((n+j)^3+2)\le v_p((n+j\pm p^{v_p(R_k)})^3+2). It follows that

    \begin{align} & v_{p}({\rm gcd} ((n+i)^3+2, (n+j)^3+2))\\ & = \min(v_p((n+i)^3+2), v_p((n+j)^3+2)))\\ &\le \min\{v_p ((n+i+p^{v_p(R_k)})^3+2), v_p( (n+j+p^{v_p(R_k)})^3+2) \}\\ & = v_{p}({\rm gcd}((n+i+p^{v_p(R_k)})^3+2, (n+j+p^{v_p(R_k)})^3+2)) \end{align} (2.7)

    and

    \begin{equation} v_{p}({\rm gcd} ((n+i)^3+2, (n+j)^3+2))\le v_{p}({\rm gcd} ((n+i-p^{v_p(R_k)})^3+2, (n+j-p^{v_p(R_k)})^3+2)). \end{equation} (2.8)

    Replacing n by n+p^{v_p(R_k)} in (2.8) gives us that

    \begin{equation} v_p(\gcd ((n+i+p^{v_p(R_k)})^3+2, (n+j+p^{v_p(R_k)})^3+2))\le v_p(\gcd((n+i)^3+2, (n+j)^3+2))). \end{equation} (2.9)

    Therefore (2.6) follows immediately from (2.7) and (2.9).

    Now using (2.6) and Hua's identity (Theorem 7.3 of [18]), we can deduce that for any given prime p , \mathcal{G}_{p, k}(n) = \mathcal{G}_{p, k}(n+p^{v_p(R_k)}) holds for any positive integer n . Hence p^{v_p(R_k)} is a period of \mathcal{G}_{p, k} which implies that P_{p, k}| p^{v_p(R_k)} . Therefore P_{p, k} are pairwise relatively prime for different prime numbers p and P_{p, k} = 1 for those primes p\nmid R_k . So \prod_{{\rm prime}\ q|R_k}P_{q, k}\Big|P_k since P_{p, k}|P_k for each prime p .

    On the other hand, by the definition of P_{p, k} , we know that for any positive integer n , v_p(\mathcal{G}_k(n+\prod_{{\rm prime}\ q|R_k}P_{q, k})) = v_p(\mathcal{G}_k(n)) holds for all primes p . Hence \mathcal{G}_k\big(n+\prod_{{\rm prime}\ q|R_k}P_{q, k}\big) = \mathcal{G}_k(n) holds for any positive integer n . That is, \prod_{p|R_k}P_{p, k} is a period of \mathcal{G}_k , which implies that P_k\Big|\prod_{{\rm prime} \ p|R_k}P_{p, k}. Hence P_k = \prod_{p|R_k}P_{p, k} as desired. Lemma 2.2 is proved.

    In order to determine the smallest period P_k of \mathcal{G}_k , by Lemma 2.2 it is enough to determine the exact value of P_{p, k} for all prime divisors p of R_k . In what follows, we show that Q_k is a period of \mathcal{G}_k . For arbitrary positive integers n and k , let

    \begin{equation} S_{k}(n): = \{n^3+2, (n+1)^3+2, ..., (n+k)^3+2\} \end{equation} (2.10)

    be the set of any k+1 consecutive terms in the cubic progression \{m^3+2\}_{m\in \mathbb{N}} , and

    \begin{equation} S_k^{(e)}(n): = \{m \in S_k(n):p^e|m\}. \end{equation} (2.11)

    Then

    \begin{align} \mathcal{G}_{p, k}(n)& = \sum\limits_{m\in S_{k}(n)}v_{p}(m) -\max\limits_{m\in S_{k}(n)}\{v_{p}(m)\} \end{align} (2.12)
    \begin{align} & = \sum\limits_{e = 1}^{\infty}|S_k^{(e)}(n)|-\sum\limits_{e = 1}^{\infty}(1 \ {\rm if} \ p^e|m \ {\rm for \ some} \ m\in S_{k}(n))\\ & = \sum\limits_{e = 1}^{\infty}{\rm max} (0, |S_k^{(e)}(n)|-1) . \end{align} (2.13)

    We denote l_p: = v_p(R_k) for any prime p . If there is at most one element divisible by p^{l_p+1} in S_{k}(n) for any positive integer n , then all the terms on the right-hand side of (2.13) are 0 if e\ge l_p+1 . It then follows that for any positive integer n , if there is at most one element divisible by p^{l_p+1} in S_{k}(n) , then

    \begin{equation} \mathcal{G}_{p, k}(n) = \sum\limits_{e = 1}^{l_p}f_{e}(n) = \sum\limits_{e = 1}^{l_p-1}f_{e}(n)+f_{l_p}(n), \end{equation} (2.14)

    where

    \begin{equation} f_{e}(n): = {\rm max} (0, |S_k^{(e)}(n)|-1). \end{equation} (2.15)

    Clearly, one has that f_{e}(n) = 0 if |S_k^{(e)}(n)|\le 1 , and f_{e}(n) = |S_k^{(e)}(n)|-1 if |S_k^{(e)}(n)| > 1 .

    Lemma 2.3. We have P_{2, k} = 2^{\frac{(-1)^{k}+1}{2}}.

    Proof. Clearly, for any odd integer n , we have v_2(n^3+2) = 0 . For any even integer n , letting n = 2m gives us that v_2(n^3+2) = v_2((2m)^3+2) = v_2(8m^3+2) = 1. So for any positive integer k , one has that \max_{m\in S_{k}(n)}\{v_2(m)\} = 1 and

    \begin{align} \sum\limits_{m\in S_{k}(n)}v_2(m) = &\sum\limits_{i = 0}^{k}v_2((n+i)^3+2)\\ = &v_2((n+k)^3+2)\delta _k+\sum\limits_{i = 0}^{\lfloor \frac{k-1}{2}\rfloor} (v_2((n+2i)^3+2)+v_2((n+2i+1)^3+2)). \end{align} (2.16)

    where \delta_k: = 1 if 2|k , and 0 otherwise. One can easily see that for any positive integers k and n , exactly one of v_2((n+2i)^3+2) and v_2((n+2i+1)^3+2) equals 1 and another one is 0 for all integers i with 0\le i\le \lfloor \frac{k-1}{2}\rfloor . This implies that v_2((n+2i)^3+2)+v_2((n+2i+1)^3+2) = 1. Also v_2((n+k)^3+2)\delta _k = 1 happens only when both of k and n are even, and 0 otherwise. Then by (2.12) and (2.16), we derive that

    \begin{array}{l} g_{2, k}(n) = &\sum\limits_{m\in S_{k}(n)}v_2(m)-{\rm max}_{m\in S_{k}(n)}\{v_{p}(m)\}\\ = &\Big\lfloor \frac{k-1}{2}\Big\rfloor+v_2((n+k)^3+2)\delta _k\\ = &{\left\{ \begin{array}{rl} \Big\lfloor \frac{k-1}{2}\Big\rfloor+1, & {\rm if} \ k \ {\rm and} \ n \ {\rm are \ even}, \\ \Big\lfloor \frac{k-1}{2}\Big\rfloor, & {\rm otherwise}\\ \end{array} \right.}\\ = &{\left\{ \begin{array}{rl} \frac{k-1}{2}, & {\rm if} \ k \ {\rm is \ odd}, \\ \frac{k}{2}, & {\rm if} \ k \ {\rm and} \ n \ {\rm are \ even}, \\ \frac{k}{2}-1, & {\rm otherwise}.\\ \end{array} \right.} \end{array}

    Therefore P_{2, k} = 1 if 2\nmid k , and if k is even, then for any positive integer n , one has g_{2, k}(n+2) = g_{2, k}(n) and g_{2, k}(n+1) \neq g_{2, k}(n) . It follows that P_{2, k} = 2 if 2|k as required. So Lemma 2.3 is proved.

    Lemma 2.4. We have that P_{3, 1} = 1 and P_{3, k} = 3^{\lceil\{ \frac{k+1}{3}\} \rceil} if k\ge 2 .

    Proof. First of all, we have

    \begin{array}{l} v_3(n^3+2) = {\left\{ \begin{array}{rl} 1, & {\rm if} \ n \equiv 1 \pmod 3, \\ 0, & {\rm otherwise}. \end{array} \right.} \end{array}

    Since 3 does not divide the difference of (n+1)^3+2 and n^3+2 , at most one of them is divisible by 3. This implies that g_{3, 1}(n) = 0 for all positive integers n . So P_{3, 1} = 1 as desired.

    Now let k\ge 2 . One has that {\rm max}_{m\in S_{k}(n)}\{v_3(m)\} = 1 and

    \begin{align} \sum\limits_{m\in S_{k}(n)}v_3(m) = &\sum\limits_{i = 0}^{k}v_3((n+i)^3+2)\\ = &\sum\limits_{i = 0}^{\lfloor \frac{k-2}{3}\rfloor} (v_3((n+3i)^3+2)+v_3((n+3i+1)^3+2)+v_3((n+3i+2)^3+2))\\ &+v_3((n+k)^3+2)\delta _k^{(1)}+v_3((n+k-1)^3+2)\delta _k^{(2)}, \end{align} (2.17)

    where

    \begin{array}{l} \delta_k^{(1)} = {\left\{ \begin{array}{rl} 1, & {\rm if \ } k \equiv 0 {\rm \ or \ } 1 \pmod 3, \\ 0, & {\rm otherwise} \end{array} \right.} \end{array}

    and

    \begin{array}{l} \delta_k^{(2)} = {\left\{ \begin{array}{rl} 1, & {\rm if \ } k \equiv 1 \pmod 3, \\ 0, & {\rm otherwise}. \end{array} \right.} \end{array}

    One can easily see that for any positive integers k and n , exactly one of v_3((n+3i)^3+2), v_3((n+3i+1)^3+2) and v_3((n+3i+2)^3+2) equals 1 and the remaining two are 0 for all integers i with 0\le i\le \lfloor \frac{k-2}{3}\rfloor . This implies that v_3((n+3i)^3+2)+v_3((n+3i+1)^3+2)+v_3((n+3i+2)^3+2 = 1. Then by (2.12) applied to p = 3 and (2.17), we derive that

    \begin{align} g_{3, k}(n) = \Big\lfloor \frac{k-2}{3}\Big\rfloor+v_3\big((n+k)^3+2\big) \delta _k^{(1)}+v_3\big((n+k-1)^3+2\big)\delta _k^{(2)}. \end{align} (2.18)

    If k\equiv 2 \pmod 3 , then \delta _k^{(1)} = \delta _k^{(2)} = 0 . So by (2.18), one has g_{3, k}(n) = \big\lfloor \frac{k-2}{3}\big\rfloor = \frac{k-2}{3} , which infers that P_{3, k} = 1 if k\equiv 2 \pmod 3 .

    If k\equiv 0 \pmod 3 , then \delta _k^{(1)} = 1 and \delta _k^{(2)} = 0 . If n\equiv 1 \pmod 3 , then v_3\big((n+k)^3+2\big) = 1 , and 0 otherwise. Again by (2.18), we have

    \begin{array}{l} g_{3, k}(n) = &\Big\lfloor \frac{k-2}{3}\Big\rfloor+v_3\big((n+k)^3+2\big)\\ = &{\left\{ \begin{array}{rl} \lfloor \frac{k-2}{3}\rfloor+1, & {\rm if} \ n \equiv 1 \pmod 3, \\ \lfloor \frac{k-2}{3}\rfloor, & {\rm otherwise}\\ \end{array} \right.}\\ = &{\left\{ \begin{array}{rl} \frac{k}{3}, & {\rm if} \ n \equiv 1 \pmod 3, \\ \frac{k}{3}-1, & {\rm otherwise}.\\ \end{array} \right.} \end{array}

    It follows that for any integer t , one has that g_{3, k}(t+3) = g_{3, k}(t) and g_{3, k}(3t) \neq g_{3, k}(3t+1) . That is P_{3, k} = 3 if k\equiv 0 \pmod 3 .

    If k\equiv 1 \pmod 3 , then \delta _k^{(1)} = \delta _k^{(2)} = 1 . If n\equiv 0 \ {\rm or} \ 1 \pmod 3 , then v_3\big((n+k-1)^3+2\big)+v_3\big((n+k)^3+2\big) = 1 , and 0 otherwise. Then by (2.18), one has

    \begin{array}{l} g_{3, k}(n) = &\Big\lfloor \frac{k-2}{3}\Big\rfloor+v_3((n+k-1)^3+2)+v_3((n+k)^3+2)\\ = &{\left\{ \begin{array}{rl} \lfloor \frac{k-2}{3}\rfloor+1, & {\rm if} \ n\equiv 0 {\rm \ or \ } 1 \pmod 3, \\ \lfloor \frac{k-2}{3}\rfloor, & {\rm otherwise}\\ \end{array} \right.}\\ = &{\left\{\begin{array}{rl} \frac{k-1}{3}, & {\rm if} \ n \equiv 0 {\rm \ or \ } 1 \pmod 3, \\ \frac{k-4}{3}, & {\rm otherwise}.\\ \end{array} \right.} \end{array}

    Hence for any integer t , one has that g_{3, k}(t+3) = g_{3, k}(t) and g_{3, k}(3t+1) \neq g_{3, k}(3t+2) . Therefore P_{3, k} = 3 if k\equiv 1 \pmod 3 .

    Finally, we can conclude that if k\ge 2 , then

    \begin{array}{l} P_{3, k} = {\left\{ \begin{array}{rl} 1, & {\rm if} \ k \equiv 2 \pmod 3, \\ 3, & {\rm otherwise}. \end{array} \right.} \end{array}

    Namely, P_{3, k} = 3^{\lceil\{ \frac{k+1}{3}\} \rceil} if k\ge 2 . This finishes the proof of Lemma 2.4.

    Lemma 2.5. Let p be a prime such that p\equiv 1\pmod 6 . Then each of the following is true:

    (ⅰ). The congruence x^3+2\equiv 0\pmod p is solvable if and only if 2^{\frac{p-1}{3}} \equiv 1 \pmod p .

    (ⅱ). The congruence x^6+108\equiv 0\pmod p is solvable if and only if 2^{\frac{p-1}{3}} \equiv 1 \pmod p .

    Proof. (ⅰ). Let m and a be positive integers with \gcd(a, p) = 1 . It is well known that x^m \equiv a \pmod p has a solution if and only if a^{\frac{p-1}{\gcd(p-1, m)}}\equiv 1\pmod p. Since p\equiv 1 \pmod 6 , letting m = 3 and a = -2 gives that x^3 \equiv -2 \pmod p has a solution if and only if 2^{\frac{p-1}{3}} \equiv 1 \pmod p . So part (i) is proved.

    (ⅱ). Likewise, x^6\equiv -108 \pmod p has a solution if and only if (-108)^{\frac{p-1}{\gcd(6, p-1)}} = (-108)^{\frac{p-1}{6}} \equiv 1 \pmod p. It is sufficient to show that (-108)^{\frac{p-1}{6}} \equiv 2^{\frac{p-1}{3}} \pmod p .

    Let (\frac{a}{p}) be the Legendre symbol. By the Gauss quadratic reciprocity law, one has

    \Bigg(\frac{p}{3}\Bigg) = (-1)^{\frac{p-1}{2}\cdot \frac{3-1}{2}} \Bigg(\frac{3}{p}\Bigg) = (-1)^{\frac{p-1}{2}}\Bigg(\frac{3}{p}\Bigg).

    Then by Euler's criterion, one has

    \begin{array}{l} (-3)^{\frac{p-1}{2}} = (-1)^{\frac{p-1}{2}}3^{\frac{p-1}{2}} \equiv (-1)^{\frac{p-1}{2}}\Bigg(\frac{3}{p}\Bigg) = \Bigg(\frac{p}{3}\Bigg)\equiv \Bigg(\frac{1}{3}\Bigg)\equiv 1\pmod p \end{array}

    since p\equiv 1 \pmod 6 . Therefore (-108)^{\frac{p-1}{6}} = 2^{\frac{p-1}{3}}\cdot(-3)^{\frac{p-1}{2}}\equiv 2^{\frac{p-1}{3}} \pmod p as desired. This completes the proof of Lemma 2.5.

    Lemma 2.6. If p is a prime number such that p\equiv 1 \pmod 6 and 2^{\frac{p-1}{3}} \not \equiv 1 \pmod p , then P_{p, k} = 1 .

    Proof. Since 2^{\frac{p-1}{3}} \not \equiv 1 \pmod p , by Lemma 2.5, x^3+2 \equiv 0 \pmod p has no solution. Thus for any positive integer n , one has v_p(n^3+2) = 0 . It then follows that \mathcal{G}_{p, k}(n) = v_p(\mathcal{G}_k(n)) = 0 . So P_{p, k} = 1 as desired. Lemma 2.6 is proved.

    Theorem 2.7. Let k be a positive integer with k\ge 2 . Then Q_k is a period of \mathcal{G}_k .

    Proof. By Lemmas 2.2-2.6, one has

    \begin{align} P_k = &P_{2, k}P_{3, k} \Bigg(\prod\limits_{p|R_k \atop{p \equiv 1\pmod 6 \atop 2^{\frac{p-1}{3}}\not\equiv 1 \pmod p}}P_{p, k}\Bigg) \Bigg(\prod\limits_{p|R_k \atop{p \equiv 5\pmod 6 }}P_{p, k}\Bigg) \Bigg(\prod\limits_{p|P_k \atop{p \equiv 1 \pmod 6 \atop 2^{\frac{p-1}{3}}\equiv 1 \pmod p}}P_{p, k}\Bigg) \\ = &2^{\frac{(-1)^{k}+1}{2}}3^{\lceil\{ \frac{k+1}{3}\} \rceil} \Bigg(\prod\limits_{p|R_k\atop {p \equiv 5 \pmod 6 }}P_{p, k}\Bigg) \Bigg(\prod\limits_{p|R_k \atop {p \equiv 1 \pmod 6 \atop 2^{\frac{p-1}{3}}\equiv 1 \pmod p}}P_{p, k}\Bigg). \end{align} (2.19)

    Since P_{p, k} is a power of p for each prime p , the product

    \Bigg(\prod\limits_{p|R_k\atop {p \equiv 5\pmod 6}}P_{p, k}\Bigg)\Bigg(\prod\limits_{p|R_k\atop {p \equiv 1 \pmod 6 \atop 2^{\frac{p-1}{3}}\equiv 1 \pmod p}}P_{p, k}\Bigg)

    divides the integer

    \frac{R_k}{2^{v_2(R_k)}3^{v_3(R_k)}\prod\limits_{p|R_k \atop{p\equiv1({\rm mod 6}) \atop {2^{\frac{p-1}{3}}\not\equiv1({\rm mod p}) }}}p^{v_p(R_k)}}.

    Thus P_k|Q_k and Q_k is a period of \mathcal{G}_k . This concludes the proof of Theorem 2.7.

    Consequently, we prove a result which may be of independently interest.

    Lemma 2.8. There is at most one prime p such that v_p(k+1)\ge v_p(R_k)\ge 1 .

    Proof. Suppose that there are two distinct primes p and q such that v_p(k+1)\ge v_p(R_k)\ge 1 and v_q(k+1)\ge v_q(R_k)\ge 1 . Then k+1\ge p and k+1\ge q . Furthermore, we have

    k+1\ge p^{v_p(R_k)}q^{v_q(R_k)}\ge \max (pq, p^{v_{p}(L_k)}q^{v_{q}(L_k)}),

    where L_k: = {\rm lcm}_{1\le i\le k}\{i\} .

    If v_{p}(L_k) = 0 or v_{q}(L_k) = 0 , then p\ge k+1 or q\ge k+1 . Thus k+1 = p\ {\rm or}\ q since k+1\ge p and k+1\ge q . We arrive at a contradiction since k+1\ge pq .

    If v_{p}(L_k)\ge 1 and v_{q}(L_k)\ge 1 , then

    \begin{align} k+1\ge p^{v_{p}(L_k)}q^{v_{q}(L_k)} \gt \min(p^{v_{p}(L_k)+1}, q^{v_{q}(L_k)+1}). \end{align} (2.20)

    But the inequality p^{v_{p}(L_k)+1} = p^{\lfloor\log_p k\rfloor +1} > p^{\log_p k} = k together with q^{v_{q}(L_k)+1} = q^{\lfloor\log_q k\rfloor+1} > q^{\log_q k} = k implies that

    \begin{align} \min(p^{v_{p}(L_k)+1}, q^{v_{q}(L_k)+1}) \gt k. \end{align} (2.21)

    Clearly, (2.20) is contradict to (2.21). So the assumption is not true. Thus Lemma 2.8 is proved.

    In the conclusion of this section, we state two well-known results that are also needed in this paper.

    Lemma 2.9. (Hensel's lemma) (see, for example, [19] or [21]) Let p be a prime and f(x) \in \mathbb{Z}[x] be a polynomial of degree n with leading coefficient not divisible by p . If there exists an integer x_1 such that f(x_1) \equiv 0 \pmod p and f'(x_1) \not \equiv 0 \pmod p , then for every integer k \ge 2 , there exists an integer x_k suck that f(x_k) \equiv 0 \pmod {p^k} and x_k \equiv x_{k-1} \pmod {p^{k-1}}.

    Lemma 2.10. (Theorem 7.2 of [18]) Let p be a prime and let a and m be positive integers such that \gcd(p, a) = 1 . Then either the congruence x^m\equiv a\pmod p has no solution or it has \gcd(m, p-1) solutions.

    Throughout this section, we always assume that p is a prime number such p|R_k and p\equiv 5\pmod 6 .

    Lemma 3.1. Then the congruence x^3+2\equiv 0\pmod p has exactly one solution in any complete residue system modulo p which is given by x\equiv\Big(\frac{-2}{p}\Big)(-2)^{\frac{p+1}{6}}\pmod p.

    Proof. By the Euler's criterion, one has (-2)^{\frac{p-1}{2}} \equiv \big(\frac{-2}{p}\big) \pmod p . Since -2 is coprime to p , one has \big(\frac{-2}{p}\big) = \pm 1 . Then by the Fermat's little theorem, we have

    \Big(\Bigg(\frac{-2}{p}\Bigg)(-2)^{\frac{p+1}{6}}\Big)^3 = \Bigg(\frac{-2}{p}\Bigg)(-2)^{\frac{p+1}{2}} \equiv (-2)^{\frac{p-1}{2}+\frac{p+1}{2}}\equiv (-2)^p \equiv (-2) \pmod p.

    So \Big(\frac{-2}{p}\Big)(-2)^{\frac{p+1}{6}} is a solution of the congruence x^3 \equiv -2 \pmod p .

    Since p\equiv 5\pmod 6 , one has \gcd(3, p-1) = 1 . Then by Lemma 2.10, one has that the congruence x^3 \equiv -2 \pmod p has only one solution. Hence Lemma 3.1 is proved.

    Lemma 3.2. The congruence x^6+108\equiv 0\pmod p has no solution.

    Proof. At first, we have \Big(\frac{-1}{p}\Big) = (-1)^{\frac{p-1}{2}} . Then by the Gauss quadratic reciprocity law, one has

    \Bigg(\frac{-108}{p}\Bigg) = \Bigg(\frac{-3}{p}\Bigg) = \Bigg(\frac{-1}{p}\Bigg)\Bigg(\frac{3}{p}\Bigg) = \Bigg(\frac{-1}{p}\Bigg)(-1)^{\frac{p-1}{2}}\Bigg(\frac{p}{3}\Bigg) = \Bigg(\frac{p}{3}\Bigg) = \Bigg(\frac{2}{3}\Bigg) = -1

    since p\equiv 5\pmod 6 . In other words, x^2+108 \equiv 0 \pmod p has no solution. This implies that the congruence x^6+108 \equiv 0 \pmod p has no solution. So Lemma 3.2 is proved.

    Lemma 3.3. Let e and n be any given positive integers. Each of the following holds:

    (ⅰ).There exists exactly one term divisible by p^e in any p^e consecutive terms of the cubic progression \{(n+i)^3+2\}_{i\in \mathbb{N}} .

    (ⅱ).There is at most one element divisible by p^{v_p(R_k)+1} in S_{k}(n) .

    Proof. (ⅰ). By Lemma 3.1, for any prime p\equiv 5\pmod 6 , x^3+2\equiv0\pmod p has exactly one solution in any given complete residue system modulo p . It then follows immediately from Lemma 2.9 (Hensel's lemma) that for any positive integer e , the congruence x^3+2\equiv0\pmod {p^e} has exactly one solution in a complete residue system modulo p^e . So there exist exactly one term divisible by p^e in any p^e consecutive terms of the cubic progression \{(n+i)^3+2\}_{i\in \mathbb{N}} . Part (ⅰ) is proved.

    (ⅱ). Let l_p: = v_p(R_k) . Then by Lemmas 3.1 and 3.2, one has

    \begin{equation} l_p = v_p({\rm lcm}_{1\le i \le k} \{i(i^6+108)\} = v_p({\rm lcm}_{1\le i \le k} \{i\}) = \max\limits_{1\le i \le k}\{v_p(i)\} \end{equation} (3.1)

    for all primes p\equiv 5\pmod 6 . By (3.1), one derives that

    \begin{align} p^{l_p}\le k \lt p^{l_p+1} \end{align} (3.2)

    Suppose that there exist integers i_1 and i_2 such that 0 \le i_1 < i_2 \le k and (n+i_1)^3+2 \equiv (n+i_2)^3+2 \equiv 0 \pmod {p^{l_p+1}} . By part (i), one has i_2-i_1 \ge p^{l_p+1} . This is a contradiction since 0\le i_1 < i_2\le k < p^{l_p+1} . Therefore there is at most one term divisible by p^{l_p+1} in S_{k}(n) . This finishes the proof of part (ⅱ).

    Lemma 3.4. Let p be a prime number such that p\equiv 5\pmod 6 and p|R_k . Then P_{p, k} = p^{v_p(R_k)} except that v_p(k+1)\ge v_p(R_k)\ge 1 , in which case one has P_{p, k} = 1 .

    Proof. First of all, part (ⅱ) of Lemma 3.3 tells us that there is at most one element divisible by p^{l_p+1} in S_{k}(n) for any positive integer n .

    Now let v_p(k+1)\ge v_p(R_k)\ge 1 . The inequality (3.2) implies that p^{l_p} < k+1\le p^{l_p+1} . Thus we can write k+1 = vp^{l_p} for some integer v with 2\le v\le p . For any positive integers n and e with 1\le e \le l_p , it follows from Lemma 3.3 (ⅰ) that |S_k^{(e)}(n)| = vp^{l_p-e} . Then |S_k^{(e)}(n)| = |S_k^{(e)}(n+1)| . By (2.15), one has f_e(n) = f_e(n+1) . Thus P_{p, k} = 1 as desired. Lemma 3.4 is true if v_p(k+1)\ge v_p(R_k) = l_p .

    In what follows, we let v_p(k+1) < v_p(R_k) = l_p . Then we can write k+1\equiv r \pmod {p^{l_p}} for some integer r with p^{v_p(k+1)}\leq r\leq p^{l_p}-1 . So one can write k+1 = vp^{l_p}+r for some integer v . By (3.2), one has 1\le v < p . Since for all positive integers n and t , (n+tp^{l_p-1})^3+2\equiv n^3+2\pmod {p^e} holds for any integer e with 1\le e \le l_p-1 , we deduce that |S_k^{(e)}(n)| = |S_k^{(e)}(n+tp^{l_p-1})| . It follows that

    \begin{equation} \sum\limits_{e = 1}^{l_p-1}f_e(n+tp^{l_p-1}) = \sum\limits_{e = 1}^{l_p-1}f_e(n). \end{equation} (3.3)

    By Lemma 2.2, P_{p, k}|p^{l_p} . Then P_{p, k} = p^{l_p} if and only if p^{l_p-1} is not a period of \mathcal{G}_{p, k} . So it is sufficient to prove that there exist a positive integers n_0 such that \mathcal{G}_{p, k}(n_0)\ne \mathcal{G}_{p, k}(n_0+p^{l_p-1}) . By Lemma 3.3, (2.14) and (3.3), this is equivalent to showing that

    \begin{equation} f_{l_p}(n_0) \ne f_{l_p}(n_0+p^{l_p-1}). \end{equation} (3.4)

    Clearly, to prove (3.4), it is enough to show that there is a positive integer n' such that

    \begin{equation} f_{l_p}(n') \gt f_{l_p}(n'+p^{l_p-1}). \end{equation} (3.5)

    By Lemma 3.3 (i), the congruence x^3+2\equiv0\pmod {p^{l_p}} has exactly one solution in a complete residue system modulo p^{l_p} . Let n_1 be a positive integer such that n_1^3+2\equiv 0\pmod {p^{l_p}} . Then any term divisible by p^{l_p} in the cubic progression \{n^3+2\}_{n\in \mathbb{N}} must be of the form (n_1+tp^{l_p})^3+2 for some integer t . In the following we show that (3.5) is true. For this purpose, we consider the following two cases:

    Case 1. p^{v_p(k+1)} \le r \le p^{l_p}-p^{l_p-1} . Then n_1+k = n_1+v{p^{l_p}}+r-1\ge n_1+vp^{l_p} . It infers that the set of all the terms divisible by p^{l_p} in the set S_k(n_1) contains the set \{n_1^3+2, (n_1+p^{l_p})^3+2, \cdots, (n_1+vp^{l_p})^3+2\} . Therefore f_{l_p}(n_1) = |S_k^{(l_p)}(n_1)|-1\ge (v+1)-1\ge v .

    On the other hand, since r\le p^{l_p}-p^{l_p-1} , we have n_1+p^{l_p-1}+k = n_1+vp^{l_p}+p^{l_p-1}+r-1 < n_1+(v+1)p^{l_p}. But n_1+p^{l_p-1} > n_1 . Hence the set of the terms divisible by p^{l_p} in the set S_k(n_1+p^{l_p-1}) is contained in the set \{(n_1+p^{l_p})^3+2, (n_1+2p^{l_p})^3+2, \cdots, (n_1+vp^{l_p})^3+2\} . So |S_k^{(l_p)}(n_1+p^{l_p-1})|\le v . It follows that f_{l_p}(n_1+p^{l_p-1}) = \max\{0, |S_k^{(l_p)}(n_1+p^{l_p-1})|-1\} \le v-1 . Then picking n' = n_1 gives us the desired result (3.5). The claim is proved in this case.

    Case 2. p^{l_p}-p^{l_p-1}+1 \le r \le p^{l_p}-1 . Let n_2: = n_1+p^{l_p}-p^{l_p-1}+1 . Then n_2\le n_1+p^{l_p} . Since p is a prime number with p\equiv 5 \pmod 6 , one has p^{l_p} > 2p^{l_p-1} . Then we have

    n_2+k = n_1+(v+1)p^{l_p}+r-p^{l_p-1}+1 \gt n_1+(v+2)p^{l_p}-2p^{l_p-1} \gt n_1+(v+1)p^{l_p}.

    So the set of all the terms divisible by p^{l_p} in the set S_k(n_2) contains the set \{(n_1+p^{l_p})^3+2, \cdots, (n_1+(v+1)p^{l_p})^3+2\} . Therefore f_{l_p}(n_2) = |S_k^{(l_p)}(n_2)|-1\ge (v+1)-1\ge v .

    However, one has n_2+p^{l_p-1} = n_1+p^{l_p}+1 > n_1+p^{l_p} and

    n_2+p^{l_p-1}+k = n_1+(v+1)p^{l_p}+r \le n_1+(v+1)p^{l_p}-1 \lt n_1+(v+2)p^{l_p}.

    Then the set of all the terms divisible by p^{l_p} in the set S_k(n_2+p^{l_p-1}) is contained in the set \{(n_1+2p^{l_p})^3+2, (n_1+3p^{l_p})^3+2, \cdots, (n_1+(v+1)p^{l_p})^3+2\} . It implies that |S_k^{(l_p)}(n_2+p^{l_p-1})|\le v . Thus f_{l_p}(n_2+p^{l_p-1}) = |S_k^{(l_p)}(n_2+p^{l_p-1})|-1 \le v-1 . Hence f_{l_p}(n_2) > f_{l_p}(n_2+p^{l_p-1}) . So letting n' = n_2 gives us the desired result (3.5). The claim is still true in this case.

    Finally, by (3.5), we have P_{p, k}\nmid p^{l_p-1} . That is, p^{l_p-1} is not a period of \mathcal{G}_{p, k} . Thus P_{p, k} = p^{v_p(R_k)} if v_p(k+1) < v_p(R_k) .

    The proof of Lemma 3.4 is complete.

    Throughout this section, we always assume that p is a prime number such that p\equiv 1\pmod 6 , p|R_k and 2^{\frac{p-1}{3}}\equiv 1\pmod p . Then \gcd(p, 108) = 1 . We begin with the following result.

    Lemma 4.1. Let e and n be positive integers. Each of the following is true:

    (ⅰ). There exist exactly three terms divisible by p^e in any p^e consecutive terms of the cubic progression \{(n+i)^3+2\}_{i\in \mathbb{N}} .

    (ⅱ). There exist exactly six terms divisible by p^e in any p^e consecutive terms of the cubic progression \{(n+i)^6+108\}_{i\in \mathbb{N}} .

    Proof. (ⅰ). Since p\equiv 1\pmod 6 and 2^{\frac{p-1}{3}}\equiv 1\pmod p , by Lemmas 2.5 and 2.10, the congruence x^3+2 \equiv0\pmod p has exactly \gcd(3, p-1) = 3 solutions. It then follows immediately from Lemma 2.9 (Hensel's lemma) that for any positive integer e , the congruence x^3+2\equiv0\pmod {p^e} has exactly 3 solutions. Hence there are exactly 3 terms divisible by p^e in any p^e consecutive terms of the cubic progression \{(n+i)^3+2\}_{i\in \mathbb{N}} . Part (ⅰ) is proved.

    (ⅱ). Likewise, by Lemmas 2.5 to 2.10 we know that the congruence x^6+108\equiv0\pmod{p^e} has exactly \gcd(6, p-1) = 6 solutions. Hence there are exactly 6 terms divisible by p^e in any p^e consecutive terms of the cubic progression \{(n+i)^3+2\}_{i\in \mathbb{N}} . Part (ⅱ) is proved.

    For any positive integer e , we define d_{p^e} to be the smallest positive root of x^6+108\equiv0\pmod{p^e}. By Lemma 4.1, one has d_{p^e} < p^e for any positive integer e . For the simplicity, we write r_0: = \max_{1\le i\le k}\{v_p(i)\} = v_p(L_k). Then d_{p^{r_0}} < p^{r_0}\le k and v_p((d_{p^{r_0}})^6+108)\ge {r_0} = \max_{1\le i\le k}\{v_p(i)\}. Since v_p(\gcd(i, i^6+108)) = v_p(\gcd(i, 108)) = 0 for any positive integer i and v_p((d_{p^{r_0}})^6+108) \le \max_{1\le i\le k}\{v_p(i^6+108)\}, we have

    \begin{align} l_p = &v_p(R_k) = \max\limits_{1\le i\le k}\{v_p(i(i^6+108))\}\\ = &\max\limits_{1\le i\le k}\{v_p(\gcd(i, i^6+108))+v_p({\rm lcm}(i, i^6+108))\}\\ = &\max\limits_{1\le i\le k}\{v_p({\rm lcm}(i, i^6+108))\}\\ = &\max\limits_{1\le i\le k}\{\max(v_p(i^6+108), v_p(i))\}\\ = &\max(\max\limits_{1\le i\le k}\{v_p(i^6+108)\}, \max\limits_{1\le i\le k}\{v_p(i)\})\\ = &\max\limits_{1\le i\le k}\{v_p(i^6+108)\}. \end{align} (4.1)

    So j^6+108\equiv 0\pmod {p^{l_p}} for some integer j with 1\le j\le k . By the definition of d_{p^{l_p}} , one has k\ge j\ge d_{p^{l_p}} and v_p(d_{p^{l_p}}^6+108)\ge l_p . Then by the fact that k\ge d_{p^{l_p}} and (4.1), we have v_p(d_{p^{l_p}}^6+108)\le l_p . Therefore l_p = v_p(d_{p^{l_p}}^6+108) .

    Since v_p(d_{p^{l_p+1}}^6+108)\ge l_p+1 and l_p = \max_{1\le i\le k}\{v_p(i^6+108)\} , one deduces that k < d_{p^{l_p+1}} . Thus we arrive at

    \begin{equation} d_{p^{l_p}}\le k \lt d_{p^{l_p+1}}. \end{equation} (4.2)

    Lemma 4.2. Let p be an odd prime number with 2^{\frac{p-1}{3}}\equiv 1\pmod p and e be a positive integer. If x_1 and x_2 are distinct roots of the congruence x^3+2\equiv 0 \pmod {p^e} in the interval [1, p^e] , then x_1^2 x_2+x_1 x_2^2 \equiv 2 \pmod {p^{e}} and x_2-x_1 is a root of the congruence x^6+108\equiv 0 \pmod {p^e} .

    Proof. Obviously, one has \gcd(x_2, p) = \gcd(x_1, p) = 1 .

    First, we show that \gcd(x_2-x_1, p) = 1 . Clearly, \gcd(x_2-x_1, p) = 1 if e = 1 . If e > 1 , then by Lemma 2.9 (Hensel's Lemma), there exist two distinct integers x_1' and x_2' in the interval [1, p] such that x_1-x_1' \equiv x_2-x_2' \equiv 0 \pmod p and x_1'^3+2\equiv x_2'^3+2 \equiv 0 \pmod p . Therefore one has that \gcd(x_2-x_1, p)\equiv \gcd(x_2'-x_1', p) \equiv 0 \pmod p . Thus we arrive at \gcd(x_2-x_1, p) = 1 .

    Subsequently, we prove that

    \begin{equation} x_1^2 x_2+x_1 x_2^2 \equiv 2 \pmod {p^e}. \end{equation} (4.3)

    In fact, by x_1^3+2\equiv x_2^3+2\equiv 0 \pmod {p^e} , we deduce that x_1(x_2^3-x_1^3) = x_1(x_2-x_1)(x_2^2+x_1x_2+x_1^2) \equiv (x_2-x_1)(x_1x_2^2+x_1^2x_2-2) \equiv 0 \pmod {p^e}. Since \gcd(x_2-x_1, p) = 1 , we have x_1x_2^2+x_1^2x_2-2 \equiv 0 \pmod {p^e} . This implies that (4.3) is true.

    Finally, we show that (x_2-x_1)^6+108\equiv 0 \pmod {p^e} . Actually, by (4.3) and noticing that x_1^3\equiv x_2^3\equiv -2\pmod {p^e} , one derives that

    \begin{array}{l} (x_2-x_1)^6+108& = x_2^6 - 6x_2^5 x_1+15x_2^4x_1^2-20x_2^3x_1^3 + 15x_2^2x_1^4 - 6x_2 x_1^5 + x_1^6 + 108\\ &\equiv 4 + 12 x_1 x_2^2 - 30 x_2 x_1^2 -80- 30 x_1 x_2^2 + 12 x_2 x_1^2 + 4 + 108\\ &\equiv 18 (2-x_1 x_2^2-x_1^2x_2)\equiv 0 \pmod {p^e}. \end{array}

    In other words, we have x_2-x_1 is a root of the congruence x^6+108\equiv 0\pmod {p^e} as desired. This concludes the proof of Lemma 4.2.

    Lemma 4.3. There is at most one element divisible by p^{l_p+1} in S_{k}(n) for any positive integer n .

    Proof. For any positive integer e , by Lemma 4.1, we know that the congruence x^3+2 \equiv 0 \pmod {p^e} holds three distinct roots in the interval [1, p^e] . We write x_{1e}, x_{2e} and x_{3e} as the three roots of x^3+2 \equiv 0 \pmod {p^e} in the interval [1, p^e] . Furthermore, by Lemma 2.9, we have x_{i1}\equiv x_{i2}\equiv \cdots \equiv x_{il_p} \pmod p for 1\le i \le 3 . We may let x_{11} < x_{21} < x_{31} . By Lemma 4.2, one has \gcd(x_{i_1 j}-x_{i_2 j}, p) = \gcd(x_{i_1j}, p) = 1 for 1\le i_1\ne i_2\le 3 and 1\le j \le l_p .

    Suppose that there exist positive integers n_1 and i_0 with 1\le i_0\le k such that

    \begin{equation} n_1^3+2 \equiv (n_1+i_0)^3+2 \equiv 0 \pmod {p^{l_p+1}}. \end{equation} (4.4)

    By Lemma 2.9 (Hensel's lemma), we know that the terms divisible by p^{l_p+1} in the cubic progression \{n^3+2\}_{n\in \mathbb{N}} must be of the form (x_{1p^{l_p}}+tp^{l_p})^3+2, (x_{2p^{l_p}}+tp^{l_p})^3+2 or (x_{3p^{l_p}}+tp^{l_p})^3+2 for some integer t . So there exist j_1, j_2 \in \{1, 2, 3\} and integers t_1, t_2\ge 0 such that n_1 = x_{j_1 l_p}+t_1 p^{l_p} and n_1+i_0 = x_{j_2 l_p}+t_2 p^{l_p} .

    First, we show that \gcd(i_0, p) = 1 . If j_1 = j_2 , then i_0 = (t_2-t_1)p^{l_p} = tp^{l_p} where t: = t_2-t_1 . By i_0\le k < d_{p^{l_p}+1} < p^{l_p+1} , one has tp^{l_p} < p^{l_p+1} which implies that t\le p-1 . Then (n_1+i_0)^3+2 = n_1^3+3n_1^2tp^{l_p}+3n_1t^2p^{2 l_p}+t^3p^{3 l_p}+2 \equiv 3n_1^2tp^{l_p}\not\equiv 0 \pmod {p^{l_p+1}}. This is a contradiction with the fact that (n_1+i_0)^3+2\equiv 0 \pmod {p^{l_p+1}} . Therefore we must have j_1\ne j_2 . Then by Lemma 4.2, one has \gcd(i_0, p) = \gcd(x_{j_2 l_p}-x_{j_1 l_p}, p) = 1. Thus n_1+i_0\not\equiv n_1 \pmod {p^{l_p+1}} and so n_1+i_0 and n_1 are distinct roots of the congruence x^3+2\equiv 0 \pmod {p^{l_p+1}} .

    By Lemma 4.2, one has i_0 is a root of the congruence x^6+108 \equiv 0 \pmod {p^{l_p+1}} , which implies i_0\ge d_{p^{l_p+1}} It contradicts with the assumption that 1\le i_0\le k < d_{p^{l_p+1}} . So the assumption that there exist positive integers n_1 and i_0 with 1\le i_0\le k such that (4.4) holds is false. This proves Lemma 4.3.

    Lemma 4.4. Let e be a positive integer and let x_{1e}, x_{2e} and x_{3e} be the three roots of x^3+2\equiv 0 \pmod {p^e} such that x_{1e} < x_{2e} < x_{3e} and x_{3e}-x_{1e} < p^e . Then d_{p^e} = \min(x_{2e}-x_{1e}, x_{3e}-x_{2e}, x_{1e}-x_{3e}+p^e) and d_{p^e}\le\frac{p^e-4}{3}.

    Proof. Let y_1: = x_{2e}-x_{1e} , y_2: = x_{3e}-x_{2e} , y_3: = x_{3e}-x_{1e} , y_4: = p^e-y_1 , y_5: = p^e-y_2 and y_6: = p^e-y_3 . Then 1\le y_i < p^e for any integer i with 1\le i\le 6 .

    By Lemma 4.2, one knows that y_1, y_2 and y_3 are the roots of the congruence x^6+108\equiv 0\pmod {p^e} . We then deduce that y_4, y_5 and y_6 are the roots of x^6+108 \equiv 0 \pmod {p^e} in the interval [1, p^e] . So all the y_i \ (1\le i\le 6) are the roots of the congruence x^6+108 \equiv 0 \pmod {p^e} in the interval [1, p^e] .

    Now we show that y_i\not\equiv y_j \pmod {p^e} for all integers i and j with 1\le i\ne j \le 6 . It is obvious that y_1\not \equiv y_3\pmod {p^e} , y_2\not\equiv y_3\pmod {p^e} , y_4\not \equiv y_6\pmod {p^e} and y_5\not\equiv y_6\pmod {p^e} . Now we show that y_1\not\equiv y_2\pmod {p^e} . If y_1\equiv y_2 \pmod {p^e} , then x_{1e}+x_{3e}\equiv 2x_{2e}\pmod {p^e} . By Lemma 4.2, one has x_{1e}^2{x_{3e}}+x_{1e} x_{3e}^2 \equiv 2 \pmod {p^e} . It follows that

    \begin{array}{l} (x_{1e}+x_{3e})^3-(2x_{2e})^3& = x_{1e}^3+x_{3e}^3-8x_{2e}^3+3x_{1e}^2x_{3e} +3x_{1e}x_{3e}^2 \\ &\equiv 12+3(x_{1e}^2{x_{3e}}+x_{1e} x_{3e}^2) \equiv 18 \pmod {p^e}. \end{array}

    In other words, 18|{p^e} , which contradicts with the assumption p \equiv 1\pmod 6 . So x_{1e}+x_{3e}\not\equiv 2x_{2e}\pmod {p^e} , i.e. y_1\not \equiv y_2 \pmod {p^e} . This implies that y_4\not \equiv y_5\pmod {p^e} .

    Likewise, one has x_{2e}+x_{3e} \not\equiv 2x_{1e} \pmod {p^e} . Then y_6-y_1 = p^e-(x_{3e}+x_{2e}-2x_{1e})\not\equiv 0\pmod {p^e} . Also we have y_4-y_1 = p^e-2(x_{2e}-x_{1e})\not\equiv 0\pmod {p^e} and y_5-y_1 = p^e-(x_{3e}-x_{1e})\not\equiv 0\pmod {p^e} . That is, y_1\not\equiv y_i \pmod {p^e} for any integer i with 4\le i\le 6 . Similarly, one has y_2\not\equiv y_i \pmod {p^e} and y_3\not\equiv y_i \pmod {p^e} for any integer i with 4\le i\le 6 . Hence y_i\not\equiv y_j \pmod {p^e} for any integers i and j with 1\le i\ne j \le 6 as one desires. So all the y_i \ (1\le i\le 6) are pairwise distinct solutions of the congruence x^6+108\equiv 0 \pmod {p^e} in the interval [1, p^e] .

    Consequently, we show that d_{p^e} = \min (y_1, y_2, y_6) . In fact, one has y_3 = y_1+y_2 and y_6 = y_4-y_2 = y_5-y_1 . It follows that \max(y_1, y_2) < y_3 and \min(y_4, y_5) > y_6 . Thus \min_{1\le i\le 6}\{y_i\} = \min(y_1, y_2, y_6) . So by the definition of d_{p^e} , we have

    d_{p^e} = \min\limits_{1\le i\le 6}\{y_i\} = \min (y_1, y_2, y_6) = \min (x_{2e}-x_{1e}, x_{3e}-x_{2e}, x_{1e}-x_{3e}+p^e)

    as required.

    Finally, since y_1, y_2 and y_6 are pairwise distinct and y_1+y_2+y_6 = p^e , we derive that

    p^e\ge d_{p^e}+(d_{p^e}+1)+(d_{p^e}+2) = 3d_{p^e}+3.

    It then follows that

    d_{p^e} = \min(y_1, y_2, y_6)\le\Big\lfloor \frac{p^e-3}{3}\Big\rfloor = \frac{p^e-4}{3}

    as one desires. This finishes the proof of Lemma 4.4.

    Lemma 4.5. Let k\ge 1 be an integer such that v_p(k+1) < v_p(R_k) . Let x_1, x_2 and x_3 be the three positive roots of x^3+2\equiv 0 \pmod {p^{l_p}} such that x_1 < x_2 < x_3 and x_3-x_1\le p^{l_p}-1 . If \max(x_2-x_1, x_3-x_2) < x_1+p^{l_p}-x_3 , then there exist positive integers n_1 and n_2 such that n_1\equiv n_2 \pmod {p^{l_p-1}} , |S_k^{(l_p)}(n_1)| > |S_k^{(l_p)}(n_2)| and |S_k^{(l_p)}(n_1)| \ge 2 .

    Proof. By Lemma 4.1 (ⅰ), there exist exactly 3 terms divisible by p^{l_p} in any p^{l_p} consecutive terms of the cubic progression \{(n+i)^3+2\}_{i\in \mathbb{N}} . Since x_i^3+2\equiv 0 \pmod {p^{l_p}} for 1\le i\le 3 with x_1 < x_2 < x_3 and x_3-x_1\le p^{l_p}-1 , we have x_0^3+2\not\equiv 0\pmod {p^{l_p}} for any integer x_0 in the interval (x_1, x_3) with x_0\ne x_2 . By Lemma 4.1, we know that the terms divisible by p^{l_p} in the cubic progression \{n^3+2\}_{n\in \mathbb{N}} must be of the form (x_1+tp^{l_p})^3+2, or (x_2+tp^{l_p})^3+2 , or (x_3+tp^{l_p})^3+2 with t being an integer. It then follows from Lemma 4.4 that

    \begin{equation} d_{p^{l_p}} \le\frac{p^{l_p}-4}{3} \lt \frac{p+2}{3}p^{l_p-1} \ {\rm and } \ d_{p^{l_p+1}} \lt \frac{p+2}{3}p^{l_p}. \end{equation} (4.5)

    So there exists a positive integer u\in [1, \frac{p+2}{3}] such that

    \begin{align} (u-1)p^{l_p-1}\le d_{p^{l_p}}\le up^{l_p-1}-1. \end{align} (4.6)

    Since v_p(k+1) < l_p , one has p^{l_p}\nmid (k+1) . Then we can write

    \begin{equation} k = vp^{l_p}+r \end{equation} (4.7)

    for some integers v and r with \ 0 \le v < \frac{p+2}{3} and 0\le r < p^{l_p}-1 . But \max(x_2-x_1, x_3-x_2) < x_1+p^{l_p}-x_3 and (x_2-x_1)+(x_3-x_2)+(x_1+p^{l_p}-x_3) = p^{l_p} . Hence

    \begin{align} x_1+p^{l_p}-x_3 \ge\Big\lceil \frac{p^{l_p}+3}{3}\Big\rceil = \frac{p^{l_p}+5}{3}. \end{align} (4.8)

    Further, by Lemma 4.4 and the hypothesis one can write d_{p^{l_p}} = x_{i_0+1}-x_{i_0} for i_0\in\{1, 2\} . Let x_4: = x_1+p^{l_p} . Then x_3 < x_4 . Now we show Lemma 4.5 by considering the following five cases.

    Case 1. 0\le r < d_{p^{l_p}} . Then by (4.2) and (4.7), we have r+1\le d_{p^{l_p}} \le vp^{l_p}+r , and so v\ge 1 . Let n_1: = x_3-\min(p^{l_p-1}, r+1)+1 . Then n_1\le x_3 and n_1+k = x_3+vp^{l_p}+(r+1)-\min(p^{l_p-1}, r+1)\ge x_3+vp^{l_p} . It follows that the set of all the terms divisible by p^{l_p} in the set S_k(n_1) contains the set \{x_3^3+2, (x_1+p^{l_p})^3+2, (x_2+p^{l_p})^3+2, (x_3+p^{l_p})^3+2, \cdots, (x_1+vp^{l_p})^3+2, (x_2+vp^{l_p})^3+2, (x_3+vp^{l_p})^3+2\} . Thus |S_k^{(l_p)}(n_1)|\ge 3v+1\ge 4 since v\ge 1 .

    Now picking n_2: = n_1+p^{l_p-1} gives us that n_2 = x_3+p^{l_p-1}-\min(p^{l_p-1}, r+1)+1 > x_3 . But from (4.8) it follows that x_1+p^{l_p}-x_3 > p^{l_p-1} . This together with the assumption x_1+p^{l_p}-x_3 > d_{p^{l_p}}\ge r+1 implies that x_1+p^{l_p}-x_3 > \max(p^{l_p-1}, r+1) . Thus x_3+\max(p^{l_p-1}, r+1) < x_1+p^{l_p} . We then deduce that n_2+k = x_3+vp^{l_p}+ p^{l_p-1}+(r+1)-\min(p^{l_p-1}, r+1) = x_3+vp^{l_p}+\max(p^{l_p-1}, r+1) < x_1+(v+1)p^{l_p}. Therefore the set of the terms divisible by p^{l_p} in the set S_k(n_2) is contained in the set \{(x_1+p^{l_p})^3+2, (x_2+p^{l_p})^3+2, (x_3+p^{l_p})^3+2, \cdots, (x_1+vp^{l_p})^3+2, (x_2+vp^{l_p})^3+2, (x_3+vp^{l_p})^3+2\} . Thus |S_k^{(l_p)}(n_2)| \le 3v which implies that |S_k^{(l_p)}(n_1)| > |S_k^{(l_p)}(n_2)| as required. So Lemma 4.5 is true if r\le d_{p^{l_p}} .

    Case 2. d_{p^{l_p}}\le r < p^{l_p}-up^{l_p-1} . Let n_1: = x_{i_0} . Since r\ge d_{p^{l_p}} and d_{p^{l_p}} = x_{i_0+1}-x_{i_0} , one has n_1+k = x_{i_0}+r+vp^{l_p}\ge x_{i_0}+d_{p^{l_p}}+vp^{l_p} = x_{{i_0}+1}+vp^{l_p} . It infers that the set of all the terms divisible by p^{l_p} in the set S_k(n_1) contains the set \{x_{i_0}^3+2, x_{{i_0}+1}^3+2, x_{{i_0}+2}^3+2, \cdots, (x_{i_0}+(v-1)p^{l_p})^3+2, (x_{{i_0}+1}+(v-1)p^{l_p})^3+2, (x_{{i_0}+2}+(v-1)p^{l_p})^3+2, (x_{i_0}+vp^{l_p})^3+2, (x_{{i_0}+1}+vp^{l_p})^3+2\} . Hence |S_k^{(l_p)}(n_1)| \ge 3v+2\ge 2 .

    Now let n_2: = x_{i_0}+up^{l_p-1} . By (4.6), one has n_2 > x_{i_0}+d_{p^{l_p}} = x_{{i_0}+1} . However, it follows from r < p^{l_p}-up^{l_p-1} that n_2+k = x_{i_0}+vp^{l_p}+up^{l_p-1}+r < x_{i_0}+vp^{l_p}+up^{l_p-1}+(p^{l_p}-up^{l_p-1}) = x_{i_0}+(v+1)p^{l_p}. Thus the set of the terms divisible by p^{l_p} in the set S_k(n_2) is contained in the set \{x_{{i_0}+2}^3+2, (x_{i_0}+p^{l_p})^3+2, (x_{{i_0}+1}+p^{l_p})^3+2, (x_{{i_0}+2}+p^{l_p})^3+2, \cdots, (x_{i_0}+vp^{l_p})^3+2, (x_{{i_0}+1}+vp^{l_p})^3+2, (x_{{i_0}+2}+vp^{l_p})^3+2\} . Thus |S_k^{(l_p)}(n_2)| \le 3v+1 . This implies that |S_k^{(l_p)}(n_2)| < |S_k^{(l_p)}(n_1)| . Lemma 4.5 holds in this case.

    Case 3. p^{l_p}-up^{l_p-1}\le r < p^{l_p}-d_{p^{l_p}}-1 . Let n_1: = x_{{i_0}+1}-up^{l_p-1}+1 . Since d_{p^{l_p}} = x_{i_0+1}-x_{i_0} , by (4.6) one has x_{i_0+1}-x_{i_0}\le up^{l_p-1}-1 , i.e. n_1\le x_{i_0} . But 1\le u \le \frac{p+2}{3} together with p > 4 shows that p^{l_p} > 2up^{l_p-1} . Then it follows from p^{l_p}-up^{l_p-1}\le r that n_1+k = x_{{i_0}+1}+vp^{l_p}-up^{l_p-1}+r+1 \ge x_{{i_0}+1}+vp^{l_p}-up^{l_p-1}+(p^{l_p}-up^{l_p-1}+1) = x_{{i_0}+1}+(v+1)p^{l_p}-2up^{l_p-1}+1 > x_{{i_0}+1}+vp^{l_p}+1 > x_{{i_0}+1}+vp^{l_p}. This concludes that the set of all the terms divisible by p^{l_p} in the set S_k(n_1) contains the set \{x_{i_0}^3+2, x_{{i_0}+1}^3+2, x_{{i_0}+2}^3+2, \cdots, (x_{i_0}+(v-1)p^{l_p})^3+2, (x_{{i_0}+1}+(v-1)p^{l_p})^3+2, (x_{{i_0}+2}+(v-1)p^{l_p})^3+2, (x_{i_0}+vp^{l_p})^3+2, (x_{{i_0}+1}+vp^{l_p})^3+2\} . Thus |S_k^{(l_p)}(n_1)|\ge 3v+2\ge 2 .

    On the other hand, let n_2: = x_{{i_0}+1}+1 . Then n_2 > x_{{i_0}+1} . Since r < p^{l_p}-d_{p^{l_p}}-1 = p^{l_p}-(x_{{i_0}+1}-x_{i_0})-1 , one has n_2+k = x_{{i_0}+1}+vp^{l_p}+r+1 < x_{{i_0}+1}+vp^{l_p}+(p^{l_p}-(x_{{i_0}+1}-x_{i_0})) = x_{i_0}+(v+1)p^{l_p}. Hence the set of the terms divisible by p^{l_p} in the set S_k(n_2) is contained in the set \{x_{{i_0}+2}^3+2, (x_{i_0}+p^{l_p})^3+2, (x_{{i_0}+1}+p^{l_p})^3+2, (x_{{i_0}+2}+p^{l_p})^3+2, \cdots, (x_{i_0}+vp^{l_p})^3+2, (x_{{i_0}+1}+vp^{l_p})^3+2, (x_{{i_0}+2}+vp^{l_p})^3+2\} . So |S_k^{(l_p)}(n_2)| \le 3v+1 . Thus Lemma 4.5 is proved in this case.

    Case 4. p^{l_p}-d_{p^{l_p}}-1 \le r < p^{l_p}-p^{l_p-1} . Let n_1: = x_{i_0} . We assert that x_{{i_0}+2}-x_{i_0} < p^{l_p}-d_{p^{l_p}} . In fact, if i_0 = 1 , then by Lemma 4.4, one has d_{p^{l_p}} = \min(x_2-x_1, x_3-x_2) since \max(x_2-x_1, x_3-x_2) < x_1+p^{l_p}-x_3 . So d_{p^{l_p}} < x_1+p^{l_p}-x_3 and the assertion follows immediately. If i_0 = 2 , then d_{p^{l_p}} = x_3-x_2\le x_2-x_1 . But the proof of Lemma 4.4 tells us that x_3-x_2\not\equiv x_2-x_1 \pmod{p^{l_p}} . Thus x_3-x_2 < x_2-x_1 that implies that d_{p^{l_p}} < x_2-x_1 . It then follows from x_4 = x_1+p^{l_p} that x_4-x_2 < p^{l_p}-d_{p^{l_p}} . The claim is proved.

    Now by the claim and the hypothesis that p^{l_p}-d_{p^{l_p}}-1 \le r , we derive that x_{{i_0}+2}-x_{i_0}+1\le p^{l_p}-d_{p^{l_p}} \le r+1. Then n_1+k = x_{i_0}+vp^{l_p}+r\ge x_{i_0}+vp^{l_p}+(x_{{i_0}+2}-x_{i_0}) = x_{{i_0}+2}+vp^{l_p}. It implies that the set of all the terms divisible by p^{l_p} in the set S_k(n_1) contains the set \{x_{i_0}^3+2, x_{{i_0}+1}^3+2, x_{{i_0}+2}^3+2, (x_{i_0}+p^{l_p})^3+2, \cdots, (x_{i_0}+vp^{l_p})^3+2, (x_{{i_0}+1}+vp^{l_p})^3+2, (x_{{i_0}+2}+vp^{l_p})^3+2\} . Thus |S_k^{(l_p)}(n_1)| \ge 3v+3 .

    Now let n_2: = x_{i_0}+p^{l_p-1} . Then n_2 > x_{i_0} . Since r < p^{l_p}-p^{l_p-1} , one has n_2+k = x_{i_0}+vp^{l_p}+p^{l_p-1}+r < x_{i_0}+vp^{l_p}+p^{l_p-1}+(p^{l_p}-p^{l_p-1}) = x_{i_0}+(v+1)p^{l_p}. Hence the set of the terms divisible by p^{l_p} in the set S_k(n_2) is contained in the set \{x_{{i_0}+1}^3+2, x_{{i_0}+2}^3+2, (x_{{i_0}}+p^{l_p})^3+2, (x_{{i_0}+1}+p^{l_p})^3+2, (x_{{i_0}+2}+p^{l_p})^3+2, \cdots, (x_{i_0}+vp^{l_p})^3+2, (x_{{i_0}+1}+vp^{l_p})^3+2, (x_{{i_0}+2}+vp^{l_p})^3+2\} . Therefore |S_k^{(l_p)}(n_2)|\le 3v+2 that infers that the desired result |S_k^{(l_p)}(n_2)| < |S_k^{(l_p)}(n_1)| is true. Hence Lemma 4.5 is proved in this case.

    Case 5. p^{l_p}-p^{l_p-1}\le r < p^{l_p}-1 . Let n_1: = x_3+p^{l_p-1}+1 . By (4.8), one has x_1+p^{l_p}-x_3 > p^{l_p-1}+1 . Then x_1+p^{l_p}-n_1 = (x_1+p^{l_p}-x_3)-(p^{l_p-1}+1) > 0, that is, n_1 < x_1+p^{l_p} . Moreover, the assumption p^{l_p}-p^{l_p-1}\le r implies that n_1+k = x_3+p^{l_p-1}+vp^{l_p}+r+1 > x_3+p^{l_p-1}+vp^{l_p}+(p^{l_p}-p^{l_p-1}) = x_3+(v+1)p^{l_p}. This infers that the set of all the terms divisible by p^{l_p} in the set S_k(n_1) contains the set \{(x_1+p^{l_p})^3+2, (x_2+p^{l_p})^3+2, (x_3+p^{l_p})^3+2, \cdots, (x_1+(v+1)p^{l_p})^3+2, (x_2+(v+1)p^{l_p})^3+2, (x_3+(v+1)p^{l_p})^3+2\} . Hence |S_k^{(l_p)}(n_1)| \ge 3v+3 .

    Let n_2: = x_3+1 . Then n_2 > x_3 . Since r < p^{l_p}-1 , we deduce that n_2+k = x_3+vp^{l_p}+r+1 < x_3+vp^{l_p}+(p^{l_p}-1)+1 = x_3+(v+1)p^{l_p}. Then the set of the terms divisible by p^{l_p} in the set S_k(n_2) is contained in the set \{(x_1+p^{l_p})^3+2, (x_2+p^{l_p})^3+2, (x_3+p^{l_p})^3+2, \cdots, (x_1+(v+1)p^{l_p})^3+2, (x_2+(v+1)p^{l_p})^3+2\} . Thus |S_k^{(l_p)}(n_2)| \le 3v+2 . So |S_k^{(l_p)}(n_2)| < |S_k^{(l_p)}(n_1)| as required.

    This completes the proof of Lemma 4.5.

    Lemma 4.6. Let p be a prime number such that p\equiv 1\pmod 6 , p|R_k and 2^{\frac{p-1}{3}}\equiv 1\pmod p . Then P_{p, k} = p^{v_p(R_k)} except that v_p(k+1)\ge v_p(R_k)\ge 1 , in which case one has P_{p, k} = 1 .

    Proof. First of all, we let v_p(k+1)\ge v_p(R_k): = l_p . Then p^{l_p}|(k+1) . But k+1\le p^{l_p+1} . So we can write k+1 = vp^{l_p} for some positive integer v with 1\le v\le p . For all positive integers n and e with 1\le e \le l_p , by Lemma 4.1 one deduces |S_k^{(e)}(n)| = 3vp^{l_p-e} . Thus |S_k^{(e)}(n)| = |S_k^{(e)}(n+1)| . By (2.15), we deduce that f_e(n) = f_e(n+1) . That is, P_{p, k} = 1 if v_p(k+1)\ge v_p(R_k) . So Lemma 4.6 is true if v_p(k+1)\ge v_p(R_k) .

    In what follows, we let v_p(k+1) < l_p . By Lemma 2.2, p^{l_p} is a period of \mathcal{G}_{p, k} . So P_{p, k} = p^{l_p} if and only if p^{l_p-1} is not a period of \mathcal{G}_{p, k} . In the following, we show that p^{l_p-1} is not a period of \mathcal{G}_{p, k} .

    Let x_1, x_2 and x_3 be the three positive roots of x^3+2\equiv 0 \pmod {p^{l_p}} such that x_1 < x_2 < x_3 and x_3-x_1\le p^{l_p}-1 . Then x_1+p^{l_p} and x_2+p^{l_p} are the roots of x^3+2\equiv 0 \pmod {p^{l_p}} and x_2 < x_3 < x_1+p^{l_p} < x_2+p^{l_p} . By the proof of Lemma 4.4, one knows that x_2-x_1, x_3-x_2 and x_1+p^{l_p}-x_3 are distinct roots of the congruence x^6+108\equiv 0 \pmod {p^{l_p}} . Then exactly one of the following three cases happens:

    (ⅰ). \max(x_2-x_1, x_3-x_2) < x_1+p^{l_p}-x_3 ,

    (ⅱ). \max(x_3-x_2, x_1+p^{l_p}-x_3) < x_2-x_1 , which is equivalent to \max(x_3-x_2, (x_1+p^{l_p})-x_3) < x_2+p^{l_p}-(x_1+p^{l_p}) ,

    (ⅲ). \max(x_2-x_1, x_1+p^{l_p}-x_3) < x_3-x_2 , which is equivalent to \max((x_1+p^{l_p})-x_3, (x_2+p^{l_p})-(x_1+p^{l_p})) < x_3+p^{l_p}-(x_2 +p^{l_p}) .

    So with Lemma 4.5 applied to the three roots x_1, x_2 and x_3 of x^3+2\equiv 0 \pmod {p^{l_p}} for case (ⅰ), and with Lemma 4.5 applied to the three roots x_2, x_3 and x_1+p^{l_p} of x^3+2\equiv 0 \pmod {p^{l_p}} for case (ⅱ), and with Lemma 4.5 applied to the three roots x_3, x_1+p^{l_p} and x_2+p^{l_p} of x^3+2\equiv 0 \pmod {p^{l_p}} for case (ⅲ), Lemma 4.5 tells us that there exist positive integers n_1 and n_2 such that n_1\equiv n_2 \pmod {p^{l_p-1}} , |S_k^{(l_p)}(n_1)| > |S_k^{(l_p)}(n_2)| and |S_k^{(l_p)}(n_1)|\ge 2 . Then one can deduce that

    \begin{equation} f_{l_p}(n_1) \gt f_{l_p}(n_2). \end{equation} (4.9)

    On the other hand, one has n_1^3+2\equiv n_2^3+2\pmod {p^e} for any integer e with 1\le e\le l_p-1 . Hence |S_k^{(e)}(n_1)| = |S_k^{(e)}(n_2)| . It then follows that

    \begin{equation} \sum\limits_{e = 1}^{l_p-1}f_e(n_1) = \sum\limits_{e = 1}^{l_p-1}f_e(n_2). \end{equation} (4.10)

    By Lemma 4.3, |S_k^{(l_p+1)}(n)|\le 1 for any positive integer n . Then we can use (2.14) and put (4.9) and (4.10) together to obtain that

    \mathcal{G}_{p, k}(n_1) = \sum\limits_{e = 1}^{l_p}f_e(n_1) \gt \sum\limits_{e = 1}^{l_p}f_e(n_2) = \mathcal{G}_{p, k}(n_2).

    Thus p^{l_p-1} is not a period of \mathcal{G}_{p, k} . It concludes that P_{p, k} = p^{l_p} if v_p(k+1) < v_p(R_k) = l_p .

    So Lemma 4.6 is proved.

    Using the lemmas presented in previous sections, we are now in a position to prove Theorems 1.1 and 1.2. We begin with the proof of Theorem 1.1.

    Proof of Theorem 1.1. First of all, by Lemma 2.1, we know that \mathcal{G}_k is periodic.

    Consequently, since R_1 = 109 is a prime number with 109 \equiv 1 \pmod 6 ,

    2^{\frac{109-1}{3}}\equiv 1 \pmod {109}

    and

    0 = v_{109}(2) \lt v_{109}(109) = 1,

    one derives from Lemma 4.6 that P_{109, 1} = 109 . It then follows from Lemma 2.2 that P_1 = 109 .

    Now let k \ge 2 . By Lemmas 2.3 and 2.4, one has

    P_{2, k} = 2^{\frac{(-1)^{k}+1}{2}}

    and

    P_{3, k} = 3^{\lceil\{ \frac{k+1}{3}\} \rceil}.

    Further, if p \equiv 1 \pmod 6 and 2^{\frac{p-1}{3}} \not \equiv 1 \pmod p , Lemma 2.6 tells us that P_{p, k} = 1 .

    By Lemma 2.8, we know that there is at most one prime p_0\ge 5 such that v_{p_0}(k+1)\ge v_{p_0}(R_k)\ge 1 . If p_0 \equiv 5 \pmod 6 , then by Lemma 3.4 one has P_{p_0, k} = 1 . If p_0 \equiv 1 \pmod 6 , then we can deduce from Lemmas 2.6 and 4.6 that P_{p_0, k} = 1 . For all other primes q\ge 5 with q|R_k , one derives from Lemmas 3.4 and 4.6 that P_{q, k} = q^{v_q(R_k)} .

    Finally, by Lemma 2.2 one then derives that the smallest period P_k of \mathcal{G}_k is equal to

    \begin{equation*} Q_k: = 2^{\frac{(-1)^{k}+1}{2}}3^{\lceil\{ \frac{k+1}{3}\} \rceil}\cdot \frac{R_k}{2^{v_2(R_k)}3^{v_3(R_k)}\prod\limits_{p|R_k \atop{ p\equiv1({\rm mod 6}) \atop {2^{\frac{p-1}{3}}\not\equiv1({\rm mod p}) }}}p^{v_p(R_k)}} \end{equation*}

    except that v_p(k+1)\ge v_p(Q_k)\ge 1 for at most one prime p\ge 5 , in which case its smallest period P_k equals \frac{Q_k}{p^{v_p(Q_k)}} .

    This finishes the proof of Theorem 1.1.

    Consequently, we give the proof of Theorem 1.2.

    Proof of Theorem 1.2. At first, \mathcal{G}_k is periodic by Theorem 1.1. Then for any positive integer n , one has \mathcal{G}_k(n) \le M: = {\rm max}_{1 \le m \le P_k}\{\mathcal{G}_k(m)\}. So one deduces that

    \begin{array}{l} {\rm log}\Big(\prod\limits_{i = 0}^k((n+i)^3+2)\Big)-{\rm log}M &\le {\rm log \ lcm}_{0 \le i \le k}\{(n+i)^3+2\}\\ &\le {\rm log}\Big( \prod\limits_{i = 0}^k ((n+i)^3+2)\Big). \end{array}

    However, one has

    \begin{array}{l} &{\rm log}\big( \prod\limits_{i = 0}^k ((n+i)^3+2) \big)-{\rm log}M\\ = &3(k+1){\rm log}n+\sum\limits_{i = 0}^k{\rm log}\Big(1+\frac{3i}{n} +\frac{3i^2}{n^2}+\frac{i^3+2}{n^3}\Big)-{\rm log}M. \end{array}

    It implies that

    \lim\limits_{n \to \infty}\frac{{\rm log} \big(\prod\limits_{i = 0}^k ((n+i)^3+2)\big)-{\rm log}M}{3(k+1){\rm log}n} = 1.

    On the other hand, we have

    \lim\limits_{n \to \infty}\frac{{\rm log} \big(\prod\limits_{i = 0}^k ((n+i)^3+2) \big)}{3(k+1){\rm log}n} = \lim\limits_{n \to \infty}\Big(1+\sum\limits_{i = 0}^k \frac{{\rm log}\big(1+\frac{3i}{n}+\frac{3i^2}{n^2} +\frac{i^3+2}{n^3}\big)}{3(k+1){\rm log}n}\Big) = 1.

    It then follows that

    \lim\limits_{n \to \infty} \frac{{\rm log \ lcm}_{0\le i\le k}\{(n+i)^3+2\}}{3(k+1){\rm log}n} = 1

    as one desires. The proof of Theorem 1.2 is complete.

    By Theorem 1.1, we can easily find infinitely many positive integers k such that P_k = Q_k as the following example shows.

    Example 5.1. If k+1 has no prime factors congruent to 5 modulo 6, and k+1 has no prime factors p such that p\equiv 1 \pmod 6 with 2^{\frac{p-1}{3}}\equiv 1 \pmod p , then P_k = Q_k by Theorem 1.1. For instance, if k+1 equals 6^r with r being a positive integer, then P_k = Q_k .

    On the other hand, there are also infinitely many positive integers k such that P_k equals Q_k divided by a power of one prime p . Moreover, by Theorem 1.1 we can present the following proposition that gives us such an example.

    Proposition 5.2. If k+1 is equal to tp^e for any positive integer e , where p is a prime number with p\equiv 5 \pmod 6 and t \in \{2, \cdots, p-1\} , then P_k = Q_k/p^e .

    Remark 5.3. Let k be a positive integer and f(x) be a polynomial with integer coefficients. Let the arithmetic function \mathcal{G}_{k, f} be defined as in the Introduction section. If f(x) is linear, then it was proved in 2011 by Hong and Qian [12] that \mathcal{G}_{k, f} is periodic with determination of its smallest period. In 2015, Hong and Qian [13] characterized the quadratic polynomial f(x) such that \mathcal{G}_{k, f} is almost periodic and an explicit and complicated formula for the smallest period of \mathcal{G}_{k, f} is obtained too. Now let \deg f(x)\ge 3 . If f(x) = x^3+2 , then by Theorem 1.1 of this paper, we know that \mathcal{G}_{k, f} is a periodic function. Furthermore, Theorem 1.1 gives us an explicit formula for its smallest period. By developing the methods presented in [12], [13] and [24] and in this paper, we can show that the arithmetic function \mathcal{G}_{k, f} is periodic when f(x) is irreducible. One can also find reducible polynomials f_1(x) and f_2(x) such that \mathcal{G}_{k, f_1} is almost periodic and \mathcal{G}_{k, f_2} is not almost periodic. For which reducible polynomials f(x) , the arithmetic function \mathcal{G}_{k, f} is almost periodic? What is the smallest period of \mathcal{G}_{k, f} if \mathcal{G}_{k, f} is almost periodic? They were answered in [12], [13] and in this paper when \deg f(x)\in\{1, 2\} and f(x) = x^3+2 . However, these problems are kept widely open when f(x)\ne x^3+2 and \deg f(x)\ge 3 .

    The authors would like to thank the anonymous referees for careful reading of the manuscript and helpful comments and corrections that improved the presentation of this paper. Hong was supported partially by National Science Foundation of China Grant # 11771304. Lin was supported partially by Research Fund of Panzhihua University Grant #2015YB41 and Doctoral Research Initiation Fund Project of Panzhihua University.

    We declare that we have no conflict of interest.



    [1] P. Bateman, J. Kalb and A. Stenger, A limit involving least common multiples, Amer. Math. Monthly, 109 (2002), 393-394.
    [2] P. L. Chebyshev, Memoire sur les nombres premiers, J. Math. Pures Appl., 17 (1852), 366-390.
    [3] B. Farhi, Minorations non triviales du plus petit commun multiple de certaines suites finies d'entiers, C. R. Acad. Sci. Paris, Ser. I, 341 (2005), 469-474. doi: 10.1016/j.crma.2005.09.019
    [4] B. Farhi, Nontrivial lower bounds for the least common multiple of some finite sequences of integers, J. Number Theory, 125 (2007), 393-411. doi: 10.1016/j.jnt.2006.10.017
    [5] B. Farhi, An identity involving the least common multiple of binomial coeffcients and its application, Amer. Math. Monthly, 116 (2009), 836-839. doi: 10.4169/000298909X474909
    [6] B. Farhi, On the derivatives of the integer-valued polynomials, arXiv:1810.07560.
    [7] B. Farhi and D. Kane, New results on the least common multiple of consecutive integers, Proc. Amer. Math. Soc., 137 (2009), 1933-1939.
    [8] C. J. Goutziers, On the least common multiple of a set of integers not exceeding N, Indag. Math., 42 (1980), 163-169.
    [9] D. Hanson, On the product of the primes, Canad. Math. Bull., 15 (1972), 33-37. doi: 10.4153/CMB-1972-007-7
    [10] S. F. Hong and W. D. Feng, Lower bounds for the least common multiple of finite arithmetic progressions, C. R. Acad. Sci. Paris, Ser. I, 343 (2006), 695-698. doi: 10.1016/j.crma.2006.11.002
    [11] S. F. Hong, Y. Y. Luo, G. Y. Qian, et al. Uniform lower bound for the least common multiple of a polynomial sequence, C.R. Acad. Sci. Paris, Ser. I, 351 (2013), 781-785. doi: 10.1016/j.crma.2013.10.005
    [12] S. F. Hong and G. Y. Qian, The least common multiple of consecutive arithmetic progression terms, Proc. Edinb. Math. Soc., 54 (2011), 431-441. doi: 10.1017/S0013091509000431
    [13] S. F. Hong and G. Y. Qian, The least common multiple of consecutive quadratic progression terms, Forum Math., 27 (2015), 3335-3396.
    [14] S. F. Hong and G. Y. Qian, New lower bounds for the least common multiple of polynomial sequences, J. Number Theory, 175 (2017), 191-199. doi: 10.1016/j.jnt.2016.11.026
    [15] S. F. Hong, G. Y. Qian and Q. R. Tan, The least common multiple of a sequence of products of linear polynomials, Acta Math. Hungar., 135 (2012), 160-167. doi: 10.1007/s10474-011-0173-4
    [16] S. F. Hong and Y. J. Yang, On the periodicity of an arithmetical function, C. R. Acad. Sci. Paris Sér. I, 346 (2008), 717-721.
    [17] S. F. Hong and Y. J. Yang, Improvements of lower bounds for the least common multiple of arithmetic progressions, Proc. Amer. Math. Soc., 136 (2008), 4111-4114. doi: 10.1090/S0002-9939-08-09565-8
    [18] L.-K. Hua, Introduction to number theory, Springer-Verlag, Berlin Heidelberg, 1982.
    [19] N. Koblitz, p-Adic numbers, p-adic analysis, and zeta-functions, Springer-Verlag, Heidelberg, 1977.
    [20] M. Nair, On Chebyshev-type inequalities for primes, Amer. Math. Monthly, 89 (1982), 126-129. doi: 10.1080/00029890.1982.11995398
    [21] J. Neukirch, Algebraic number theory, Springer-Verlag, 1999.
    [22] S. M. Oon, Note on the lower bound of least common multiple, Abstr. Appl. Anal., 2013.
    [23] G. Y. Qian and S. F. Hong, Asymptotic behavior of the least common multiple of consecutive arithmetic progression terms, Arch. Math., 100 (2013), 337-345. doi: 10.1007/s00013-013-0510-7
    [24] G. Y. Qian, Q. R. Tan and S. F. Hong, The least common multiple of consecutive terms in a quadratic progression, Bull. Aust. Math. Soc., 86 (2012), 389-404. doi: 10.1017/S0004972712000202
    [25] R. J. Wu, Q. R. Tan and S. F. Hong, New lower bounds for the least common multiple of arithmetic progressions, Chinese Annals of Mathematics, Series B, 34 (2013), 861-864. doi: 10.1007/s11401-013-0805-9
  • Reader Comments
  • © 2020 the Author(s), licensee AIMS Press. This is an open access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.0)
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

Metrics

Article views(3959) PDF downloads(364) Cited by(0)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog