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

On the number of integers which form perfect powers in the way of x(y21+y22+y23+y24)=zk

  • Let k2 be an integer. We studied the number of integers which form perfect k-th powers in the way of

    x(y21+y22+y23+y24)=zk.

    For k4, we established a unified asymptotic formula with a power-saving error term for the number of such integers of bounded size under Lindelöf hypothesis, and we also gave an unconditional result for k=2.

    Citation: Tingting Wen. On the number of integers which form perfect powers in the way of x(y21+y22+y23+y24)=zk[J]. AIMS Mathematics, 2024, 9(4): 8732-8748. doi: 10.3934/math.2024423

    Related Papers:

    [1] Zhen Pu, Kaimin Cheng . Consecutive integers in the form $ a^x+y^b $. AIMS Mathematics, 2023, 8(8): 17620-17630. doi: 10.3934/math.2023899
    [2] Ziyu Dong, Zhengjun Zhao . An application of $ p $-adic Baker method to a special case of Jeśmanowicz' conjecture. AIMS Mathematics, 2023, 8(5): 11617-11628. doi: 10.3934/math.2023588
    [3] Bingzhou Chen, Jiagui Luo . On the Diophantine equations $x^2-Dy^2=-1$ and $x^2-Dy^2=4$. AIMS Mathematics, 2019, 4(4): 1170-1180. doi: 10.3934/math.2019.4.1170
    [4] Jinyan He, Jiagui Luo, Shuanglin Fei . On the exponential Diophantine equation $ (a(a-l)m^{2}+1)^{x}+(alm^{2}-1)^{y} = (am)^{z} $. AIMS Mathematics, 2022, 7(4): 7187-7198. doi: 10.3934/math.2022401
    [5] Cheng Feng, Jiagui Luo . On the exponential Diophantine equation $ \left(\frac{q^{2l}-p^{2k}}{2}n\right)^x+(p^kq^ln)^y = \left(\frac{q^{2l}+p^{2k}}{2}n\right)^z $. AIMS Mathematics, 2022, 7(5): 8609-8621. doi: 10.3934/math.2022481
    [6] Xiao Jiang, Shaofang Hong . On the denseness of certain reciprocal power sums. AIMS Mathematics, 2019, 4(3): 412-419. doi: 10.3934/math.2019.3.412
    [7] Yahui Yu, Jiayuan Hu . On the generalized Ramanujan-Nagell equation $ x^2+(2k-1)^y = k^z $ with $ k\equiv 3 $ (mod 4). AIMS Mathematics, 2021, 6(10): 10596-10601. doi: 10.3934/math.2021615
    [8] Li Zhou, Liqun Hu . Sum of the triple divisor function of mixed powers. AIMS Mathematics, 2022, 7(7): 12885-12896. doi: 10.3934/math.2022713
    [9] Bai-Ni Guo, Dongkyu Lim, Feng Qi . Series expansions of powers of arcsine, closed forms for special values of Bell polynomials, and series representations of generalized logsine functions. AIMS Mathematics, 2021, 6(7): 7494-7517. doi: 10.3934/math.2021438
    [10] Cencen Dou, Jiagui Luo . Complete solutions of the simultaneous Pell's equations $ (a^2+2)x^2-y^2 = 2 $ and $ x^2-bz^2 = 1 $. AIMS Mathematics, 2023, 8(8): 19353-19373. doi: 10.3934/math.2023987
  • Let k2 be an integer. We studied the number of integers which form perfect k-th powers in the way of

    x(y21+y22+y23+y24)=zk.

    For k4, we established a unified asymptotic formula with a power-saving error term for the number of such integers of bounded size under Lindelöf hypothesis, and we also gave an unconditional result for k=2.



    Counting the number of integral solutions of a certain Diophantine equation is an interesting project in number theory. Let Nk(H) be the number of pairs of positive integers x1, x2H whose product x1x2 is a perfect k-th power. Tolev [1] first established an asymptotic formula for Nk(H). The proof combines Perron's formula with elementary ideas from the work of Heath-Brown and Moroz [2]. It is proved that for any integer k2,

    Nk(H)=ckH2/k(logH)k1+O(H2/k(logH)k2),

    where ck>0 is an explicit constant depending on k. De la Bretèche et al. [3] improved this result based on the multiple Dirichlet series theory and complex analysis. They showed that there exists a constant θk>0 such that

    Nk(H)=H2/kQ(logH)+O(H2/kθk+ε),

    where Q is a polynomial of degree k1 with leading coefficient ck. In fact, they considered a more general case and counted the number of tuples of n2 integers whose product is a perfect k-th power. Precisely, they proved that there exists a constant θn,k>0, such that

    Nn,k=|{(x1,,xn)[1,H]nNn:x1,,xn=zk,zN}|=Hn/kQn,k(logH)+O(Hn/kθn,k),

    where Qn,k is a polynomial of degree (n+k1k)n.

    It is natural to consider the analogue of the above results for polynomials. Liu and Niu [4] counted pairs of polynomials whose product is a cube over finite field and obtained an asymptotic formula (see [4, Theorem 1.2]) by contour integration. It is indicated that there are some differences and challenges between polynomials and integers.

    Returning to the case of integers, this problem is closely related to the integral points on algebraic surface

    S:x1x2x3=x30.

    This is a split toric surface and can also be regarded as the product of three integers forming a perfect cube. It has been studied by many authors including de la Bretèche [5], Fouvry [6], Heath-Brown and Moroz [2], and Salberger [7]. Denote by NS(H) the number of primitive integral points (i.e., gcd(x0,x1,x2,x3)=1) on S satisfying x00 and max0i3|xi|H. The sharpest unconditional result is proved by de la Bretèche [5], which states

    NS(H)=HP(logH)+O(H7/8exp(c(logH)3/5)(loglogH)1/5)),

    where P is a polynomial of degree 6 and c is a positive constant. Now, we look at the corresponding non-split toric surface

    S:x(y21+y22)=z3.

    It can be observed that S and S are isomorphic over Q(i). Denote by NS(H) the number of primitive integral points on S satisfying z0 and

    max{|x|,y21+y22,|z|}H.

    De la Bretèche et al. [8] studied this surface and proved that

    NS(H)=HP(logH)+O(H8/9+ε),

    where P is a cubic polynomial. Liu et al. [9] further considered the following case

    S4:x(y21+y22+y23+y24)=z3. (1.1)

    For ease of presentation, we let

    y=(y1,y2,y3,y4),y=y21+y22+y23+y24,

    and denote by N4(H) the number of integral tuples (x,y1,,y4), satisfying max{|x|,y}H, which form nonzero perfect cubes in the way of (1.1) since the variable z is completely determined by x and y. They showed that

    N4(H)=c4H3(logH)2+O(H3logH).

    In fact, they dealt with a more general case that the number of squares is a multiple of 4 (see [9, Theorem 7.1]). Zhai [10] improved this result by obtaining a power-saving error term that

    N4(H)=H3P4(logH)+O(H31/4+ε),

    where P4 is a quadratic polynomial. All above results are obtained by studying the corresponding multiple Dirichlet series. Liu et al. stated in [9] that the same idea can be applied to investigate the number of integral solutions of some higher-degree Diophantine equation like

    xd=(y21+y22+y23+y24)zd2,d4.

    They obtained an asymptotic formula in [11] for d=4, and Wen [12] established the asymptotic formulae for any integer d4 with power-saving error terms.

    We remark here that the Diophantine equations S, S, and S4 mentioned above are homogeneous, so there is an equivalent relation between rational points in projective space and integral points in affine space up to a scalar multiplication. They are actually related to another project in number theory, which is Manin's conjecture (see [2,5,6,7,8,9] for more details).

    Motivated by the above work, in this paper, we mainly focus on the Diophantine equation

    Sk4:x(y21+y22+y23+y24)=zk (1.2)

    for k2. Similar to (1.1), we denote by Nk4(H) the number of integral tuples (x,y1,,y4) satisfying max{|x|,y}H, which form nonzero perfect k-th powers in the way of (1.2). Recall that the Lindelöf hypothesis (LH in brief, [13, §Ⅱ.3.4]) states that

    ζ(1/2+it)(|t|+1)ε

    for any ε>0. Our main result is as follows.

    Theorem 1.1. Let k4 be any integer. Assuming LH, then for any ε>0, there exists a constant ϑk, such that

    Nk4(H)=H2+3/kP(logH)+O(H2+3/kϑk+ε),

    where

    ϑk=32k(12k/31[2k/3])>0,

    and [α] is the integral part of α, the implied constant only depends on k and ε, P is a polynomial of degree k1 given by (3.14) with leading coefficient 16Ck, and Ck is a positive constant given by (3.13).

    Remark 1.2. The assumption of LH is just for simplifying the calculation. We claim that the above asymptotic formulae still holds for k=4 unconditionally, since we can apply the fourth moment estimate of the Riemann zeta function instead of LH as in (4.8).

    The case of k=3 has been solved in [9,10] unconditionally as mentioned above. In fact, following the proof of Theorem 1.1, one can easily obtain an asymptotic formula for k=3 with a power-saving error term O(H31/5+ε). We shall leave that as an example to justify our result. We next give an unconditional result for k=2.

    Theorem 1.3. Unconditionally, we have

    N24(H)=16H2+3/2P(logH)+O(H2+3/21/3+ε),

    where P is a linear polynomial given by (4.7) and the implied constant only depends on ε.

    We first introduce the bivariate Perron's formula [10,Lemma 2.2], which plays an important role in our proof.

    Lemma 2.1. Suppose that f(n1,n2) is a bivariate arithmetic function and its Dirichlet series

    F(s1,s2)=n1=1n2=1f(n1,n2)ns11ns22

    is absolutely convergent for (sj)>σj(j=1,2) with some σ1, σ2>0. Let x1, x2, T1, T210 be parameters such that xjN, and define

    bj=σj+1/logxj,j=1,2.

    We have

    n1x1n2x2f(n1,n2)=1(2πi)2b1+iT1b1iT1b2+iT2b2iT2F(s1,s2)xs11xs22s1s2ds2ds1+O(xσ11xσ22E),

    where

    E:=2j=1n1=1n2=1|f(n1,n2)|nb11nb22min{1,1Tj|logxjnj|}.

    Recalling the definition of Nk4(H) before, we see that

    Nk4(H)=|{(x,y,z)Z6Sk4:max{|x|,y}H,z0}|.

    Let r4(n) be the number of representations of a positive integer n as the sum of four squares

    n=y21+y22+y23+y24

    with

    (y1,y2,y3,y4)Z4.

    It is well known that

    r4(n)=8r4(n) with r4(n)=dnd (2.1)

    and r^*_4(n) is a multiplicative arithmetic function. Let \mathbb{1}_k denote an indicator function of perfect k -th power defined by

    \begin{equation} \mathbb{1}_k(n) = \begin{cases} 1, &\text{ if } \;n {\text{ is a perfect }} k-{\text{th power}}, \\ 0, &\text{ otherwise.} \end{cases} \end{equation} (2.2)

    In view of the above problem, we can write

    \begin{equation} \begin{split} N_4^k(H)& = 2\sum\limits_{1 \leqslant m \leqslant H}\sum\limits_{1 \leqslant n \leqslant H^2}r_4(n)\mathbb{1}_k(n)\\& = 16\sum\limits_{1 \leqslant m \leqslant H}\sum\limits_{1 \leqslant n \leqslant H^2}r_4^*(n)\mathbb{1}_k(n).\end{split} \end{equation} (2.3)

    In order to deal with this double sum, we define the corresponding Dirichlet series

    \begin{equation} \mathcal F_k(s, w) = \sum\limits_{m, n = 1}^\infty\frac{r_4^*(n)\mathbb{1}_k(mn)}{m^sn^w}. \end{equation} (2.4)

    The following proposition gives the expression and convergence of \mathcal F_k(s, w) , which allows us to extend the double Dirichlet series to a suitable large region.

    Proposition 2.2. Let k \geqslant2 . If \Re (s) > 1/k and \Re (w) > 1+1/k , then

    \begin{equation} \mathcal F_k(s, w) = \prod\limits_{j = 0}^k\zeta \left((k-j)s+j(w-1) \right)\mathcal{H}_k(s, w), \end{equation} (2.5)

    where \mathcal{H}_k(s, w) is an Euler product given by (2.16), which is absolutely convergent if s and w satisfy the conditions

    \begin{equation} \min\limits_{1 \leqslant i \leqslant k-1}\Re((k-i)s+i(w-1)) \geqslant 1/2+\varepsilon \ \mathit{\text{and}}\ \Re(w) \geqslant 1+\varepsilon. \end{equation} (2.6)

    Furthermore, we have

    \begin{equation} \mathcal{H}_k(s, w)\ll 1 \end{equation} (2.7)

    in the above region, and the implied constant is absolute.

    Proof. Note that r_4^*(n) is multiplicative, and (2.1) yields r_4^*(1) = 1 and

    \begin{equation} r_4^*(p^\nu) = \begin{cases} \dfrac{1-p^{\nu+1}}{1-p}, & \text{ if } p \geqslant 3, \\ 3, & \text{ if } p = 2, \end{cases} \end{equation} (2.8)

    for any integer \nu \geqslant1 , then we can rewrite \mathcal F_k(s, w) in (2.4) as the Euler product

    \begin{equation} \begin{split} \mathcal F_k(s, w)& = \prod\limits_p\sum\limits_{\mu \geqslant0}\sum\limits_{\nu \geqslant0}\frac{r_4^*(p^\nu)\mathbb{1}_k(p^{\mu+\nu})}{p^{\mu s+\nu w}}\\& = \prod\limits_p\sum\limits_{d \geqslant0}p^{-kds}\sum\limits_{0 \leqslant\nu \leqslant kd}\frac{r_4^*(p^\nu)}{p^{\nu(w-s)}}\\& = :\prod\limits_p\mathcal F_{k, p}(s, w).\end{split} \end{equation} (2.9)

    Here, we let \mu+\nu = kd for d \geqslant0 , according to the definition of \mathbb{1}_k in (2.2). On the other hand, a simple formal calculation shows

    \begin{equation} \begin{split} & \sum\limits_{d \geqslant0}x^d\sum\limits_{0 \leqslant\nu \leqslant kd}y^\nu\frac{1-z^{kd+1}}{1-z} = \frac{1}{1-z}\sum\limits_{d \geqslant0}x^d \left(\frac{1-y^{kd+1}}{1-y}-z\frac{1-(yz)^{kd+1}}{1-yz} \right) \\ & = \frac{1}{1-z} \left(\frac{1}{1-y} \left(\frac{1}{1-x}-\frac{y}{1-xy^k} \right)-\frac{z}{1-yz} \left(\frac{1}{1-x}-\frac{yz}{1-x(yz)^k} \right) \right)\\ & = \frac{G_k(x, y, z)}{(1-x)(1-xy^k)(1-x(yz)^k)} \end{split} \end{equation} (2.10)

    with

    G_2(x, y, z) = 1+xy(1+z)+xy^2z

    and

    \begin{split} G_k(x, y, z)& = 1+xy(1+z)+xy^2(1+z+z^2)+\cdots\cdots+xy^{k-1}(1+z+\cdots+z^{k-1})\\ &\quad+xy^k(z+z^2+\cdots+z^{k-1})+x^2y^{k+1}(z^2+z^3+\cdots+z^{k-1})\\ &\quad+x^2y^{k+2}(z^3+z^4+\cdots+z^{k-1})+\cdots\cdots+x^2y^{2k-2}z^{k-1} \end{split}

    for k \geqslant3 . Similarly, we have

    \begin{equation} \begin{split} 1+\sum\limits_{d \geqslant1}x^d \left(1+3\sum\limits_{1 \leqslant\nu \leqslant kd}y^\nu \right)& = \frac{1}{1-x}+\frac{3}{1-y}\sum\limits_{d \geqslant1}x^d \left(y-y^{kd+1} \right) \\ & = \frac{1}{1-x}+\frac{3(xy-xy^{k+1})}{(1-y)(1-x)(1-xy^k)}\\& = \frac{1+3xy(1+y+\cdots+y^{k-2})+2xy^k}{(1-x)(1-xy^k)}. \end{split} \end{equation} (2.11)

    When p \geqslant3 , in view of (2.8), we apply (2.10) with

    (x, y, z) = (p^{-ks}, p^{-(w-s)}, p)

    to deduce that

    \begin{equation} \mathcal F_{k, p}(s, w) = \prod\limits_{0 \leqslant j \leqslant k} \left(1-\frac{1}{p^{(k-j)s+j(w-1)}} \right)^{-1}\mathcal H_{k, p}(s, w), \end{equation} (2.12)

    where

    \begin{equation} \mathcal H_{k, p}(s, w) = G_k \left(p^{-ks}, p^{-(w-s)}, p \right) \left(1-\frac{1}{p^{kw}} \right)^{-1}\prod\limits_{1 \leqslant j \leqslant k-1} \left(1-\frac{1}{p^{(k-j)s+j(w-1)}} \right). \end{equation} (2.13)

    Meanwhile, for p = 2 , the formula (2.11) with

    (x, y) = (2^{-ks}, 2^{-(w-s)})

    gives us

    \begin{equation} \mathcal F_{k, 2}(s, w) = \prod\limits_{0 \leqslant j \leqslant k} \left(1-\frac{1}{2^{(k-j)s+j(w-1)}} \right)^{-1}\mathcal H_{k, 2}(s, w), \end{equation} (2.14)

    where

    \begin{aligned} \mathcal H_{k, 2}(s, w) = \left(1+\frac{3}{2^{(k-1)s+w}} \left(1+\frac{1}{2^{w-s}}+\cdots+\frac{1}{2^{(k-2)(w-s)}} \right)+\frac{1}{2^{kw-1}} \right) \times \left(1-\frac{1}{2^{kw}} \right)^{-1}\prod\limits_{1 \leqslant j \leqslant k} \left(1-\frac{1}{2^{(k-j)s+j(w-1)}} \right). \end{aligned} (2.15)

    It can be observed that \mathcal F_{k, p}(s, w)/\mathcal H_{k, p}(s, w) will give the Euler product of Riemann zeta functions in (2.12) and (2.14). Combining (2.9) and (2.12)–(2.15), we get (2.5) with

    \begin{equation} \mathcal H_k(s, w) = \mathcal H_{k, 2}(s, w)\prod\limits_{p \geqslant3}\mathcal H_{k, p}(s, w). \end{equation} (2.16)

    Next, we shall discuss the convergence. In view of the expression of \mathcal H_{k, p}(s, w) in (2.13) and (2.15), when expanding \mathcal H_{k, p}(s, w) into 1 plus some monomials about p , we see that the power of p in the denominator of each monomial is great than 1 if

    \min\limits_{1 \leqslant i \leqslant k-1}\Re((k-i)s+i(w-1)) \geqslant 1/2+\varepsilon \ \text{ and }\ \Re(w) \geqslant 1+\varepsilon,

    so we have

    \mathcal H_{k, p}(s, w) = 1+O(p^{-1-\varepsilon}),

    and \mathcal H_k(s, w) is absolutely convergent in this region, which implies (2.7). This completes the proof.

    We next make use of bivariate Perron's formula. Suppose H\notin \mathbb{N} and let T\in[10, H^{{1}/{2}}] be a parameter to be chosen later. The analytic property of \mathcal F_k(s, w) allows us to set

    \begin{equation*} \begin{split} x_1 = H, \ &\ x_2 = H^2, \ \ s_1 = s = \sigma + \mathrm{i} t, \ \ s_2 = w = u+ \mathrm{i} v, \ \ b_1 = 1/k+\varepsilon, \ \ b_2 = 1/k+1+\varepsilon \end{split} \end{equation*}

    in Lemma 2.1, then we get

    \begin{equation*} \begin{split} N_4^k(H)& = \frac{16}{(2\pi \mathrm{i})^2}\int_{b_1- \mathrm{i} T}^{b_1+ \mathrm{i} T}\int_{b_2-10 \mathrm{i} T}^{b_2+10 \mathrm{i} T}\frac{\mathcal F_k(s, w)H^{s+2w}}{ws}\, \mathrm{d} w\, \mathrm{d} s +O\Big({H^{2+\frac{3}{k}+\varepsilon}}\big(\mathfrak{J}_1(H, T)+\mathfrak{J}_2(H, T)\big)\Big) \end{split} \end{equation*}

    with

    \begin{equation*} \begin{split} \mathfrak{J}_1(H, T) &: = \sum\limits_{m, n = 1}^\infty \frac{r_4^*(n)\mathbb{1}_k(mn)}{m^{b_1}n^{b_2}}\min\Bigg(1, \frac{1}{T |\log \frac{H}{m}|}\Bigg), \\ \mathfrak{J}_2(H, T) &: = \sum\limits_{m, n = 1}^\infty\frac{r_4^*(n)\mathbb{1}_k(mn)}{m^{b_1}n^{b_2}}\min\Bigg(1, \frac{1}{T |\log \frac{H^{2}}{n}|}\Bigg). \end{split} \end{equation*}

    Noticing that

    r_4^*(n)\ll n\tau(n),

    and following the arguments in [10], one can easily get

    \begin{equation*} \label{0000} \begin{split} \mathfrak{J}_1(H, T)&\ll\sum\limits_{m, n = 1}^\infty\frac{\tau(n)\mathbb{1}_k(mn)}{(mn)^{b_1}}\min\Bigg(1, \frac{1}{T |\log \frac{H}{m}|}\Bigg)\\&\ll \frac{H^\varepsilon}{T}, \end{split} \end{equation*}

    and the same result holds for \mathfrak{J}_2(H, T) . It follows that

    \begin{equation} N_4^k(H) = \frac{16}{(2\pi \mathrm{i})^2}\int_{b_1- \mathrm{i} T}^{b_1+ \mathrm{i} T}\int_{b_2-10 \mathrm{i} T}^{b_2+10 \mathrm{i} T}\frac{\mathcal F_k(s, w)H^{s+2w}}{ws}\, \mathrm{d} w\, \mathrm{d} s+O\bigg(\frac{H^{2+3/k+\varepsilon}}{T}\bigg). \end{equation} (2.17)

    It suffices to evaluate the double integral on the righthand side of (2.17).

    In this section, we shall prove the main theorem. The proof is divided into three steps. We shall first turn the double integral into some usual single integrals by Cauchy's residue theorem and then deal with the single integrals. The final step is choosing the suitable parameters.

    In this subsection, we shall apply Cauchy's residue theorem to evaluate the inner integral over w in (2.17) and derive the following result.

    Lemma 3.1. Let k \geqslant2 and N_4^k(H) be defined as in (2.3). Assuming LH, we have

    \begin{equation*} \label{expression of S with 3 error terms} \begin{split} N_4^k(H) = 16\sum\limits_{j = 1}^kI_j(H, T)+O \left({H^{2+3/k+\varepsilon}}/{T}+H^{2+2/k+\varepsilon} \right), \end{split} \end{equation*}

    where

    \begin{equation} \begin{split} &I_j(H, T): = \frac{1}{2\pi \mathrm{i}}\int_{b_1- \mathrm{i} T}^{b_1+ \mathrm{i} T}\underset{{w = w_j} }{\mathop{{\rm{res}} }}\,\frac{\mathcal F_k(s, w)}{ws}H^{s+2w_j}\, \mathrm{d} s \end{split} \end{equation} (3.1)

    with

    w_j = (1-(k-j)s)/j+1

    for 1 \leqslant j \leqslant k .

    Proof. In terms of the inner integral in (2.17), we consider the domain formed by four points

    w = b_2 \pm 10 \mathrm{i} T, \quad w = 1/2k+1 \pm 10 \mathrm{i} T.

    In this domain, from Proposition 2.2, we easily see that the integrand function \mathcal F_k(s, w)/(ws) has k simple poles:

    w_j = \tfrac{1}{j}(1-(k-j)s)+1

    for 1 \leqslant j \leqslant k.

    Using the residue theorem for the variable w , we get

    \begin{equation} N_4^k(H) = 16\sum\limits_{j = 1}^k I_j(H, T)+ R_1(H, T)+R_2(H, T)-R_3(H, T)+O \left({H^{2+3/k+\varepsilon}}/{T} \right), \end{equation} (3.2)

    where I_j(H, T) is given by (3.1) and

    \begin{equation*} \begin{split} &R_1(H, T): = \frac{16}{(2\pi \mathrm{i})^2}\int_{b_1- \mathrm{i} T}^{b_1+ \mathrm{i} T}\int_{\frac{1}{2k}+1+10 \mathrm{i} T}^{b_2+10 \mathrm{i} T}\frac{\mathcal F_k(s, w)H^{s+2w}}{ws}\, \mathrm{d} w\, \mathrm{d} s, \\ &R_2(H, T): = \frac{16}{(2\pi \mathrm{i})^2}\int_{b_1- \mathrm{i} T}^{b_1+ \mathrm{i} T}\int_{\frac{1}{2k}+1-10 \mathrm{i} T}^{\frac{1}{2k}+1+10 \mathrm{i} T}\frac{\mathcal F_k(s, w)H^{s+2w}}{ws}\, \mathrm{d} w\, \mathrm{d} s, \\ &R_3(H, T): = \frac{16}{(2\pi \mathrm{i})^2}\int_{b_1- \mathrm{i} T}^{b_1+ \mathrm{i} T}\int_{\frac{1}{2k}+1-10 \mathrm{i} T}^{b_2-10 \mathrm{i} T}\frac{\mathcal F_k(s, w)H^{s+2w}}{ws}\, \mathrm{d} w\, \mathrm{d} s. \end{split} \end{equation*}

    Set s = b_1+ \mathrm{i} t with -T \leqslant t \leqslant T and w = u+10 \mathrm{i} T with 1/2k+1 \leqslant u \leqslant b_2 . By Proposition 2.2, one has \mathcal{H}_k(s, w) \ll 1 , since the conditions in (2.6) are satisfied clearly. On the other hand, recall that LH states

    \zeta(1/2+it)\ll (|t|+1)^\varepsilon,

    then we deduce from the Phragmén-Lindelöf principle that the subconvexity bound of the Riemann zeta function under LH satisfies

    \begin{equation} \zeta(\sigma+ \mathrm{i} t)\ll \begin{cases} 1, &\text{ if } \sigma > 1, \\ (|t|+1)^{\max\{\frac12-\sigma, \, 0\}+\varepsilon}, &\text{ if } 0 < \sigma \leqslant1. \end{cases} \end{equation} (3.3)

    This will be used several times below, and it follows that

    \mathcal F_k(b_1+ \mathrm{i} t, u+10 \mathrm{i} T)\ll T^{\varepsilon}

    holds uniformly for 1/2k+1 \leqslant u \leqslant b_2 . This implies

    \begin{equation*} \begin{split} R_1(H, T) &\ll \int_{-T}^T\frac{ \mathrm{d} t}{|t|+1}\int_{\frac{1}{2k}+1}^{b_2}T^{-1+\varepsilon}H^{b_1+2u}\, \mathrm{d} u \\& \ll {H^{2+3/k+\varepsilon}}/T, \end{split} \end{equation*}

    and the same bound holds for R_3(H, T) . Similarly, for u = 1/2k+1 , we have

    \mathcal F_k(b_1+ \mathrm{i} t, 1/2k+1+ \mathrm{i} v)\ll (|t|+|v|+1)^{\varepsilon},

    and R_2(H, T) can be estimated as

    \begin{align*} \begin{split} R_2(H, T)&\ll H^{b_1+2(1/2k+1)} \int_{-T}^T(|t|+1)^{-1+\varepsilon}\, \mathrm{d} t\int_{-10T}^{10T}(|v|+1)^{-1+\varepsilon}\, \mathrm{d} v\\&\ll H^{2+2/k+\varepsilon}. \end{split} \end{align*}

    Inserting the upper bounds of R_i(H, T) (i = 1, 2, 3) into (3.2), we obtain the required formula.

    In this subsection, we shall evaluate I_j (H, T) for 1 \leqslant j \leqslant k , and our main idea is Cauchy's residue theorem.

    Lemma 3.2. Let k \geqslant4 and I_j(H, T) for 1 \leqslant j \leqslant k be defined as in (3.1), then we have the following estimates under LH for each I_j(H, T) :

    \begin{equation*} \label{all estimate of I_123} \begin{split} I_j(H, T) &\ll{H^{2+\frac{2}{k}+\frac{j-2}{2k(k-j-1)}+\varepsilon}}+{H^{2+\frac{3}{k}+\varepsilon}}/{T}, \quad \mathit{\text{for}} 1 \leqslant j < 2k/3, \\ I_{2k/3}(H, T)& = \mathcal{C}_{2k/3}H^{2+3/k}+O \left({H^{2+\frac{3}{k}+\varepsilon}}/{T} \right), \\ I_j(H, T) & = H^{2+3/k}P_j(\log H)+O \left(H^{2+\frac{3}{2k}+\frac{2k-3}{2k(j-1)}+\varepsilon}+H^{2+\frac{3}{k}+\varepsilon}/T \right), \quad \mathit{\text{for}}\, 2k/3 < j < k, \\ I_k(H, T)& = H^{2+3/k}P_k(\log H)+O\Big(H^{2+\frac{2}{k}+\frac{k-2}{2k(k-1)}+\varepsilon}T^{\frac{1}{2(k-1)}}+H^{2+\frac{3}{k}+\varepsilon}/T\Big), \end{split} \end{equation*}

    where \mathcal C_{3k/2}, P_j(\log H) for 2k/3 < j < k and P_k(\log H) are defined in (3.6), (3.9), and (3.12), respectively.

    Remark 3.3. In fact, the above result still holds for k = 3 apart from the third formula, since the third case vanishes if k = 3 .

    Proof. Recall the definition of I_j(H, T) in (3.1). We deduce from Proposition 2.2 that

    \underset{{w = w_j}}{\mathop{{\rm{res}} }}\,\frac{\mathcal F_k(s, w)}{ws} = \frac{\mathcal H_k(s, w_j) }{j\cdot w_js}\prod\limits_{\substack{0 \leqslant \ell \leqslant k\\ \ell\neq j}}\zeta \left(\frac{\ell}{j}+ \left(1-\frac{\ell}{j} \right)ks \right) = :G_j(s).

    It follows that

    \begin{equation*} I_j(H, T) = \frac{1}{2\pi \mathrm{i}}\int_{b_1- \mathrm{i} T}^{b_1+ \mathrm{i} T}G_j(s)H^{2+\frac{2}{j}+(3-\frac{2k}{j})s} \mathrm{d} s. \end{equation*}

    The proof of this lemma is divided into four parts based on whether 3-2k/j is positive, negative, or zero. Keep in mind that the letter j is always an integer.

    If 1 \leqslant j < 2k/3 , the condition (2.6) becomes

    \min\limits_{1 \leqslant i \leqslant k-1}\Re((k-i)s+i(w_j-1)) \geqslant 1/2+\varepsilon,

    which implies

    \Re(s) < (2k-2-j)/2k(k-1-j).

    So, we shift the line of integration from b_1 to

    \Re (s) = \sigma_j = {(2k-2-j)}/{2k(k-1-j)}-\varepsilon.

    By Proposition 2.2, the Euler product \mathcal{H}_k(s, w_j) is absolutely convergent in the region b_1 \leqslant \Re(s) \leqslant \sigma_j , and we have \mathcal{H}_k(s, w_j)\ll 1 . It follows from this and (3.3) that

    \begin{equation} \begin{split} G_j(\sigma+ \mathrm{i} t)&\ll \frac{|\mathcal{H}_k(s, w_j)|}{(|t|+1)^2}\prod\limits_{\substack{0 \leqslant \ell \leqslant k\\ \ell\neq j}} \left|\zeta \left(\frac{\ell}{j}+ \left(1-\frac{\ell}{j} \right)k(\sigma+ \mathrm{i} t) \right) \right|\\ &\ll \left|\zeta \left(\tfrac{k}{j}+\big(1-\tfrac{k}{j}\big)k(\sigma+ \mathrm{i} t) \right) \right|\big/(|t|+1)^{2-\varepsilon}\\ & \ll (|t|+1)^{1/2(k-1-j)-2+\varepsilon}\\&\ll (|t|+1)^{-3/2+\varepsilon} \end{split} \end{equation} (3.4)

    for b_1 \leqslant \sigma \leqslant \sigma_j , since the real part of each zeta function is

    \Re(\ell/j+(1-\ell/j)ks) > 1/2

    for 0 \leqslant\ell \leqslant k-1 and \ell\neq j . The above bound allows us to extend the integral of I_j(H, T) on [-T, T] to (-\infty, \infty) , so we have

    \begin{equation*} \begin{split} I_j(H, T)& = \frac{1}{2\pi \mathrm{i}}\int_{(b_1)}G_j(s)H^{2+\frac{2}{j}-(\frac{2k}{j}-3)s}\, \mathrm{d} s+O \left(H^{2+\frac{3}{k}+\varepsilon}/T \right), \end{split} \end{equation*}

    where \int_{(b_1)} means \int_{b_1-i\infty}^{b_1+i\infty} . Note that there is no pole in the region b_1 \leqslant \Re(s) \leqslant \sigma_j . The residue theorem tells us that

    \begin{equation} \begin{split} I_j(H, T)& = \frac{1}{2\pi \mathrm{i}}\int_{(\sigma_j)}G_j(s)H^{2+\frac{2}{j}-(\frac{2k}{j}-3)s}\, \mathrm{d} s+O \left(H^{2+\frac{3}{k}+\varepsilon}/T \right). \end{split} \end{equation} (3.5)

    By (3.4), the integral on the righthand side can be estimated as

    \begin{align*} H^{2+\frac{2}{j}-(\frac{2k}{j}-3)\sigma_j}\int_{-\infty}^\infty|G_j(\sigma_j+ \mathrm{i} t)|\, \mathrm{d} t &\ll H^{2+\frac{2}{k}+\frac{j-2}{2k(k-j-1)}+\varepsilon}\int_{-\infty}^\infty\frac{ \mathrm{d} t}{(|t|+1)^{\frac{3}{2}-\varepsilon}}\\& \ll H^{2+\frac{2}{k}+\frac{j-2}{2k(k-j-1)}+\varepsilon}. \end{align*}

    Combining this with (3.5), we obtain the first assertion in Lemma 3.2.

    If j = 2k/3 is an integer, we find that

    I_j(H, T) = \frac{1}{2\pi \mathrm{i}}\int_{b_1- \mathrm{i} T}^{b_1+ \mathrm{i} T}G_j(s)H^{2+\frac{2}{j}} \mathrm{d} s.

    Otherwise, this case does not exist. Note that

    G_{2k/3}(s)\ll (|t|+1)^{-(2-\varepsilon)}

    for s = b_1+ \mathrm{i} t , which shows that G_{2k/3}(b_1+ \mathrm{i} t, \chi) has a good convergence as a function in t . It follows that

    \begin{equation*} \begin{split} I_{2k/3}(H, T) & = \frac{1}{2\pi \mathrm{i}}\int_{(b_1)} G_{2k/3}(s)H^{2+\frac{3}{k}}\, \mathrm{d} s+ O \left({H^{2+3/k+\varepsilon} }/{T} \right)\\ & = :\mathcal{C}_{2k/3}H^{2+3/k}+O \left({H^{2+3/k+\varepsilon}}/{T} \right), \end{split} \end{equation*}

    where

    \begin{equation} \mathcal C_{2k/3} = \frac{1}{2\pi \mathrm{i}}\int_{(\frac{1}{k}+\varepsilon)}G_{2k/3}(s)\, \mathrm{d} s \end{equation} (3.6)

    is an absolute constant. This gives the second assertion.

    If 2k/3 < j < k , we shift the line of integration from b_1 to

    \Re (s) = \sigma_j' = (j-2)/2k(j-1)+\varepsilon.

    It is easy to check that (2.6) is exactly satisfied, so we have \mathcal{H}_k(s, w_j)\ll 1 . It follows from (3.3) again that

    \begin{equation} \begin{split} G_j(\sigma+ \mathrm{i} t)&\ll \frac{|\mathcal{H}_k(s, w_j)|}{|w_js|}\prod\limits_{\substack{0 \leqslant \ell \leqslant k\\ \ell\neq j}} \left|\zeta \left(\frac{\ell}{j}+ \left(1-\frac{\ell}{j} \right)k(\sigma+ \mathrm{i} t) \right) \right|\\ &\ll \frac{|\zeta(k(\sigma+ \mathrm{i} t))|}{(|t|+1)^2}\\& \ll (|t|+1)^{\max\{1/2-k\sigma, \, 0\}-2+\varepsilon}\\&\ll (|t|+1)^{-3/2+\varepsilon} \end{split} \end{equation} (3.7)

    for \sigma_j' < \sigma \leqslant b_1 , which yields

    \begin{split} I_j(H, T)& = \frac{1}{2\pi \mathrm{i}}\int_{(b_1)}G_j(s)H^{2+\frac{2}{j}+(3-\frac{2k}{j})s}\, \mathrm{d} s+O \left(H^{2+\frac{3}{k}+\varepsilon}/T \right). \end{split}

    Note that

    0 < \Re(\ell/j+(1-\ell/j)ks) < 1

    for 0 \leqslant\ell < j , so there is only a pole s = 1/k of order j given by G_j(s) in the strip \sigma_j' \leqslant \sigma \leqslant b_1 . The residue theorem gives us

    \begin{equation} \begin{split} I_j(H, T) & = \underset{s = 1/k}{\mathop{{\rm{res}} }}\,G_j(s)H^{2+3/k}+\frac{1}{2\pi \mathrm{i}}\int_{(\sigma_j')}G_j(s)H^{2+\frac{2}{j}+(3-\frac{2k}{j})s}\, \mathrm{d} s+O \left(H^{2+\frac{3}{k}+\varepsilon}/T \right). \end{split} \end{equation} (3.8)

    We can compute the residue in the main term as

    \begin{equation} \begin{split} \underset{s = 1/k}{\mathop{{\rm{res}} }}\,G_j(s)& = \frac{1}{(j-1)!}\lim\limits_{s\to 1/k}\frac{ \mathrm{d}^{j-1}}{ \mathrm{d} s^{j-1}}\left( {\big(s-\tfrac{1}{k}\big)^jG_j(s)H^{s-\frac{1}{k}}} \right)\\& = :P_j(\log H).\end{split} \end{equation} (3.9)

    This is a polynomial of degree j-1 with leading coefficient

    \frac{\mathcal H_k(\frac{1}{k}, \frac{1}{k}+1)(j/k)^{j-2}}{(k+1)((j-1)!)^2}.

    The integral in (3.8) can be bounded by

    H^{2+\frac{2}{j}+(3-\frac{2k}{3})\sigma_j'}\int_{-\infty}^\infty |G_j(\sigma_j'+ \mathrm{i} t)|\, \mathrm{d} t \ll H^{2+\frac{3}{2k}+\frac{2k-3}{2k(j-1)}+\varepsilon}.

    Inserting (3.9) and the above estimate into (3.8), we get the third formula in Lemma 3.2.

    Finally, it remains to deal with the case j = k . Now, we shift the line of integration from b_1 to

    \Re (s) = \sigma_k = (k-2)/2k(k-1)+\varepsilon.

    Noticing that w_k = 1/k+1 is no longer dependent on s , it follows from (3.3) that

    \begin{equation} \begin{split} G_k(\sigma+ \mathrm{i} t)& = \frac{\mathcal{H}_k(s, w_k)}{k\cdot w_ks}\prod\limits_{\substack{0 \leqslant\ell < k}}\zeta \left(\frac{\ell}{k}+ \left(1-\frac{\ell}{k} \right)k(\sigma+ \mathrm{i} t) \right)\\ &\ll \frac{|\zeta(k(\sigma+ \mathrm{i} t))|}{(|t|+1)^{1-\varepsilon}}\\&\ll \begin{cases} (|t|+1)^{(1/2-k\sigma)-1+\varepsilon}, &\text{if $\sigma_k \leqslant \sigma \leqslant 1/2k$, }\\ (|t|+1)^{-1+\varepsilon}, &\text{if $1/2k < \sigma \leqslant b_1$, } \end{cases} \end{split} \end{equation} (3.10)

    since \mathcal{H}_k(s, w_k) is absolutely convergent for \sigma_k \leqslant\Re(s) \leqslant b_1 according to (2.6). Note that there is only one pole s = 1/k of order k given by G_k(s) in the rectangle formed by \sigma_k\pm \mathrm{i} T and b_1\pm \mathrm{i} T . It follows from the residue theorem that

    \begin{equation} \begin{split} I_k(H, T) & = \underset{s = 1/k}{\mathop{{\rm{res}} }}\,G_k(s)H^{2+3/k}+\left(\int_{b_1- \mathrm{i} T}^{\sigma_k- \mathrm{i} T}+\int_{\sigma_k- \mathrm{i} T}^{\sigma_k+ \mathrm{i} T}+\int_{\sigma_k+ \mathrm{i} T}^{b_1+ \mathrm{i} T} \right)G_k(s)H^{2+\frac{2}{k}+s}\, \mathrm{d} s\\ & = H^{2+3/k}P_k(\log H)+ K_{1}(H, T)+ K_{2}(H, T) +K_{3}(H, T), \end{split} \end{equation} (3.11)

    where K_{i}(H, T) for i = 1, 2, 3 corresponds to the above three integrals, respectively, and P_k is a polynomial of degree k-1 given by

    \begin{equation} P_k(\log H) = \frac{1}{(k-1)!}\lim\limits_{s\to 1/k}\frac{ \mathrm{d}^{k-1}}{ \mathrm{d} s^{k-1}}\left( {\big(s-\tfrac{1}{k}\big)^kG_k(s)H^{s-\frac{1}{k}}} \right) \end{equation} (3.12)

    with leading coefficient

    \begin{equation} \mathscr C_k = \frac{\mathcal{H}_k(\frac{1}{k}, \frac{1}{k}+1)}{(k+1)((k-1)!)^2}. \end{equation} (3.13)

    This is a positive constant depending on k . By (3.10), using the same method as before, we can estimate K_{1}(H, T) by

    \begin{align*} K_{1}(H, T)&\ll\int_{\sigma_k}^{\frac{1}{2k}}\frac{H^{2+{2}/{k}+\sigma}}{T^{1/2+k\sigma-\varepsilon}}\, \mathrm{d}\sigma+\int^{b_1}_{\frac{1}{2k}}\frac{H^{2+{2}/{k}+\sigma}}{T^{1-\varepsilon}}\, \mathrm{d}\sigma\\ &\ll H^{2+\frac{2}{k}+\frac{k-2}{2k(k-1)}+\varepsilon}\big/T^\frac{2k-3}{2(k-1)}+H^{2+\frac{3}{k}+\varepsilon}/T, \end{align*}

    so does K_{3}(H, T) . As for K_{2}(H, T) , noting that

    G_k(\sigma_k+ \mathrm{i} t)\ll (|t|+1)^{-(2k-3)/2(k-1)+\varepsilon},

    we get

    \begin{align*} K_{2}(H, T)&\ll H^{2+\frac{2}{k}+\sigma_k}\int_{-T}^T(|t|+1)^{-(2k-3)/2(k-1)+\varepsilon}\, \mathrm{d} t\\&\ll H^{2+\frac{2}{k}+\frac{k-2}{2k(k-1)}+\varepsilon}T^{\frac{1}{2(k-1)}}. \end{align*}

    The last formula in Lemma 3.2 is obtained followed from the above two estimates and (3.11). This completes the proof.

    Combining Lemmas 3.1 and 3.2, we get

    \begin{equation*} \begin{split} \frac{N_4^k(H)}{H^2}& = 16H^\frac{3}{k}\Bigg(\mathbb{1}_{j = {2k}/{3}}\mathcal{C}_{2k/3}+\sum\limits_{2k/3 < j \leqslant k}P_j(\log H)\Bigg)+O \left(H^{\frac{2}{k}+\varepsilon}+{H^{\frac{3}{k}+\varepsilon}}/{T} \right)\\ &\quad+O \Bigg(\sum\limits_{1 \leqslant j < 2k/3}{H^{\frac{2}{k}+\frac{j-2}{2k(k-j-1)}+\varepsilon}} + \sum\limits_{2k/3 < j < k} {H^{\frac{3}{2k}+\frac{2k-3}{2k(j-1)}+\varepsilon}} +H^{\frac{2}{k}+\frac{k-2}{2k(k-1)}+\varepsilon}T^{\frac{1}{2(k-1)}}\Bigg) \end{split} \end{equation*}

    for k \geqslant4 . It can be simplified to

    {N_4^k(H)}/{H^2} = H^\frac{3}{k}\mathcal{P}(\log H)+O \left({H^{\frac{3}{k}+\varepsilon}}/{T}+H^{\frac{3}{2k}+\frac{2k-3}{2k[2k/3]}+\varepsilon}+H^{\frac{2}{k}+\frac{k-2}{2k(k-1)}+\varepsilon}T^{\frac{1}{2(k-1)}} \right),

    where [\alpha] is the integral part of \alpha and

    \begin{equation} \mathcal{P}(\log H) = 16 \Bigg(\mathbb{1}_{j = 2k/3}\mathcal{C}_{2k/3}+\sum\limits_{2k/3 < j \leqslant k}P_j(\log H)\Bigg). \end{equation} (3.14)

    This is a polynomial of degree k-1 with leading coefficient 16\mathscr C_k . It suffices to choose a suitable T to balance the error terms. It can be observed that the error term is actually dominated by H^{\frac{3}{2k}+\frac{2k-3}{2k[2k/3]}+\varepsilon} for k \geqslant4 . Just letting H^{9/4k^2}\ll T\ll H^{1/k} , one can get the second formula in Theorem 1.1.

    Following the proof of Theorem 1.1, we give a sketch of the proof for k = 2 . According to Proposition 2.2, the corresponding Dirichlet series can be written as

    \begin{equation*} \label{expression of F_2} \mathcal F_2(s, w) = \zeta(2s)\zeta(s+w-1)\zeta(2(w-1))\mathcal H_2(s, w) \end{equation*}

    with

    \begin{split} \mathcal H_2(s, w)& = \left(1-\frac{1}{2^{2w-2}} \right) \prod\limits_{p} \left(1+\frac{1}{p^{s+w}}+\frac{1}{p^{s+w-1}}+\frac{1}{p^{2w-1}} \right) \left(1-\frac{1}{p^{2w}} \right)^{-1} \left(1-\frac{1}{p^{s+w-1}} \right), \end{split}

    which is absolutely convergent and satisfies \mathcal H_2(s, w)\ll 1 in the region

    \begin{equation} \Re(s+w) \geqslant 3/2+\varepsilon, \quad\Re(w) \geqslant 1+\varepsilon. \end{equation} (4.1)

    Applying Perron's formula, we get

    \begin{equation} N_4^2(H) = \frac{16}{(2\pi \mathrm{i})^2}\int_{b_1- \mathrm{i} T}^{b_1+ \mathrm{i} T}\int_{b_2-10 \mathrm{i} T}^{b_2+10 \mathrm{i} T}\frac{\mathcal F_2(s, w)H^{s+2w}}{ws}\, \mathrm{d} w\, \mathrm{d} s+O\bigg(\frac{H^{2+3/2+\varepsilon}}{T}\bigg) \end{equation} (4.2)

    for b_1 = 1/2+\varepsilon and b_2 = 3/2+\varepsilon .

    Following the arguments in Section 3, we still shift the path of integration over w to \Re(w) = 5/4 . Clearly, (4.1) is satisfied. Applying the residue theorem to evaluate the inner integral in (4.2), we can get (3.2) for k = 2 . Note that the assumption of LH is used to bound \mathcal F_k(s, w) and G_j(s) in some region before. Unconditionally, we have the following well-known estimate for the Riemann zeta function (see [13, Theorem Ⅱ.3.8] for example):

    \begin{equation} \zeta(\sigma+ \mathrm{i} t)\ll \begin{cases} 1, &\text{ if }\sigma > 1, \\ (|t|+1)^{\frac{1-\sigma}{3}+\varepsilon}, &\text{ if } 1/2 \leqslant\sigma \leqslant 1, \\ (|t|+1)^{\frac{3-4\sigma}{6}+\varepsilon}, &\text{ if } 0 < \sigma < 1/2. \end{cases} \end{equation} (4.3)

    Just replacing (3.3) by (4.3) and using the same method as in §3.1, we can get

    \begin{equation} N_4^2(H) = 16\big(I_1(H, T)+I_2(H, T)\big)+O \left({H^{2+\frac{3}{2}+\varepsilon}}/T+H^{3+\varepsilon}T^\frac{1}{2} \right) \end{equation} (4.4)

    with

    \begin{split} I_1(H, T)& = \frac{1}{2\pi \mathrm{i}}\int_{b_1- \mathrm{i} T}^{b_1+ \mathrm{i} T}\frac{\zeta(2s)\zeta(2-2s)}{(2-s)s}\mathcal H_2(s, 2-s)H^{4-s}\, \mathrm{d} s, \\ I_2(H, T)& = \frac{1}{2\pi \mathrm{i}}\int_{b_1- \mathrm{i} T}^{b_1+ \mathrm{i} T}\frac{\zeta(2s)\zeta(s+1/2)}{3s}\mathcal H_2(s, 3/2)H^{3+s}\, \mathrm{d} s. \end{split}

    The treatments of I_1(H, T) and I_2(H, T) are a little different from before. There are only two Riemann zeta functions involved here, which can be treated more carefully. As for I_1(H, T) , we shift the line of integration from b_1 to \sigma_1 = 1-\varepsilon due to (4.1). It follows from (4.3) that

    \zeta(2s)\zeta(2-2s)\mathcal{H}_2(s, 2-s)\ll |\zeta(2-2s)|\ll (|t|+1)^{\max\{(2\sigma-1)/3, \, (8\sigma-5)/6\}+\varepsilon}

    for s = \sigma+ \mathrm{i} t and b_1 \leqslant\sigma \leqslant \sigma_1 . Noticing that there is no pole for the integrand function in this strip, similar to (3.5), we deduce from the residue theorem that

    I_1(H, T) = \frac{1}{2\pi \mathrm{i}}\int_{(\sigma_1)}\frac{\zeta(2s)\zeta(2-2s)}{(2-s)s}\mathcal H_2(s, 2-s)H^{4-s}\, \mathrm{d} s+O \left(H^{2+\frac{3}{2}+\varepsilon}/T \right).

    For s = \sigma_1+ \mathrm{i} t , the integral above can be bounded by

    H^{4-\sigma_1}\int_{-\infty}^\infty\frac{\zeta(2\varepsilon-2 \mathrm{i} t)}{(|t|+1)^2}\, \mathrm{d} t\ll H^{3+\varepsilon}\int_{-\infty}^{\infty}\frac{ \mathrm{d} t}{(|t|+1)^{3/2-\varepsilon}}\ll H^{3+\varepsilon}.

    It follows that

    \begin{equation} I_1(H, T)\ll H^{3+\varepsilon}+H^{2+3/2+\varepsilon}/T. \end{equation} (4.5)

    It suffices to deal with I_2(H, T) . In view of (4.1), we shift the path of integration from b_1\pm \mathrm{i} T to \varepsilon\pm \mathrm{i} T . Note that there is a pole s = 1/2 of order 2 in this region. It follows from the residue theorem that

    \begin{equation} I_2(H, T) = H^{2+\frac{3}{2}}P(\log H)+ \left(\int_{b_1- \mathrm{i} T}^{\varepsilon- \mathrm{i} T}+\int_{\varepsilon- \mathrm{i} T}^{\varepsilon+ \mathrm{i} T}+\int_{\varepsilon+ \mathrm{i} T}^{b_1+ \mathrm{i} T} \right)\frac{\zeta(2s)\zeta(s+1/2)}{3s}\mathcal H_2(s, 3/2)H^{3+s}\, \mathrm{d} s, \end{equation} (4.6)

    where P is a linear polynomial given by

    \begin{equation} \begin{split} P(\log H)& = \lim\limits_{s\to 1/2}\frac{ \mathrm{d}}{ \mathrm{d} s}\left( \frac{1}{3s}{(s-1/2)^2} {\zeta(2s)\zeta(s+1/2)}\mathcal H_2(s, 3/2) H^{s-\frac{1}{2}} \right)\\ & = \frac{1}{3}\mathcal{H}_2(1/2, 3/2)\log H+(\gamma-2/3)\mathcal{H}_2(1/2, 3/2)+\frac{1}{3}\mathcal{H}_2'(1/2, 3/2) \end{split} \end{equation} (4.7)

    and \gamma is Euler's constant. We derive from (4.3) that the first and third integrals in (4.6) can be bounded by

    \int_\varepsilon^{1/4} T^{\frac{2-5\sigma}{3}-1+\varepsilon}H^{3+\sigma} \mathrm{d}\sigma+\int_{1/4}^{b_1}T^{1/2-\sigma-1+\varepsilon}H^{3+\sigma} \mathrm{d} \sigma\ll H^{3+\varepsilon}/T^{\frac{1}{3}}+H^{3+\frac{1}{4}+\varepsilon}/T^{\frac{3}{4}}+H^{3+\frac{1}{2}+\varepsilon}/T.

    Recall the fourth moment estimate of the Riemann zeta function (see [14, Theorem 5.1])

    \begin{equation*} \label{moment of zeta} \int_{-T}^T|\zeta(\sigma + \mathrm{i} t)|^4\, \mathrm{d} t\ll T\log^4T \end{equation*}

    for 1/2 \leqslant \sigma < 1 and the functional equation, which states

    \zeta(s) = \chi(s)\zeta(1-s)\; \; \text{with}\; \; \chi(s)\sim (|t|+1)^{1/2-\sigma}

    for \sigma \leqslant1/2 . By Hölder's inequality and partial integration, we can estimate the second integral in (4.6) as follows

    \begin{equation} \begin{split} & H^{3+\varepsilon}\int_{-T}^T\frac{|\zeta(2\varepsilon+2 \mathrm{i} t)\zeta(1/2+\varepsilon+ \mathrm{i} t)|}{|t|+1}\, \mathrm{d} t\\ &\ll H^{3+\varepsilon} \left(\int_{-T}^T\frac{|\zeta(2\varepsilon+2 \mathrm{i} t)|^4}{|t|+1}\, \mathrm{d} t \right)^{1/4} \left(\int_{-T}^T\frac{|\zeta(1/2+\varepsilon+ \mathrm{i} t)|^4}{|t|+1}\, \mathrm{d} t \right)^{1/4} \left(\int_{-T}^T\frac{ \mathrm{d} t}{|t|+1} \right)^{1/2}\\ &\ll H^{3+\varepsilon} \left(\int_{-T}^T\frac{ \left|(|t|+1)^{1/2-2\varepsilon}\zeta(1-2\varepsilon-2 \mathrm{i} t) \right|^4}{|t|+1}\, \mathrm{d} t \right)^{1/4}\\&\ll H^{3+\varepsilon}T^{1/2}. \end{split} \end{equation} (4.8)

    Combining all the above, we get

    \begin{equation} I_2(H, T) = H^{2+\frac{3}{2}}P(\log H)+O \left(H^{3+\varepsilon}T^{1/2}+H^{3+1/4+\varepsilon}/T^{3/4}+H^{3+1/2+\varepsilon}/T \right). \end{equation} (4.9)

    Finally, inserting (4.5) and (4.9) into (4.2), we get

    N_4^2(H) = 16H^{2+\frac{3}{2}}P(\log H)+O \left(H^{2+\frac{3}{2}+\varepsilon}/T+H^{3+\varepsilon}T^{\frac{1}{2}}+H^{3+\frac{1}{4}+\varepsilon}/T^\frac{3}{4} \right).

    Choosing T = H^{1/3} , the error terms are balanced to H^{2+3/2-1/3+\varepsilon} . This gives the required formula in Theorem 1.3.

    In this paper, we study the number of integers which form perfect powers in the way of

    x(y_1^2+y_2^2+y_3^2+y_4^2) = z^k

    and the proof relies on techniques coming from complex analysis. We point out that this problem was never done before except for the case k = 3 . It is not easy to establish a unified asymptotic formula with power-saving error terms for large k , so we assume the Lindelöf hypothesis for k \geqslant4 . Theorem 1.1 gives an asymptotic formula with a power-saving error term for the number of such integers of bounded size under LH. This is the novelty of this paper. Moreover, Theorem 1.3 gives an unconditional result for k = 2 .

    The author declares she has not used Artificial Intelligence (AI) tools in the creation of this article.

    The author would like to thank the anonymous referees for many useful comments on the manuscript.

    The author declares no conflict of interest.



    [1] D. I. Tolev, On the number of pairs of positive integers x_1, x_2 \leqslant H such that x_1x_2 is a k-th power, Pacific J. Math., 249 (2011), 495–507. https://doi.org/10.2140/pjm.2011.249.495 doi: 10.2140/pjm.2011.249.495
    [2] D. R. Heath-Brown, B. Z. Moroz, The density of rational points on the cubic surface X_0^3 = X_1X_2X_3, Math. Proc. Cambridge Philos. Soc., 125 (1999), 385–395. https://doi.org/10.1017/S0305004198003089 doi: 10.1017/S0305004198003089
    [3] R. de la Bretèche, P. Kurlberg, I. E. Shparlinski, On the number of products which form perfect powers and discriminants of multiquadratic extensions, Int. Math. Res. Not., 22 (2021), 17140–17169. https://doi.org/10.1093/imrn/rnz316 doi: 10.1093/imrn/rnz316
    [4] K. Liu, W. Niu, Counting pairs of polynomials whose product is a cube, J. Number Theory, 256 (2024), 170–194. https://doi.org/10.1016/j.jnt.2023.09.009 doi: 10.1016/j.jnt.2023.09.009
    [5] R. de la Bretèche, Sur le nombre de points de hauteur bornée d'une certaine surface cubique singulière, Astérisque, 1998, 51–77. https://doi.org/10.24033/ast.410
    [6] É. Fouvry, Sur la hauteur des points d'une certaine surface cubique singulière, Astérisque, 1998, 31–49. https://doi.org/10.24033/ast.409
    [7] P. Salberger, Tamagawa measures on universal torsors and points of bounded height on Fano varieties, Astérisque, 1998, 91–258. https://doi.org/10.24033/ast.412
    [8] R. de la Bretèche, K. Destagnol, J. Liu, J. Wu, Y. Zhao, On a certain non-split cubic surface, Sci. China Math., 62 (2019), 2435–2446. https://doi.org/10.1007/s11425-018-9543-8 doi: 10.1007/s11425-018-9543-8
    [9] J. Liu, J. Wu, Y. Zhao, Manin's conjecture for a class of singular cubic hypersurfaces, Int. Math. Res. Not., 2019 (2019), 2008–2043. https://doi.org/10.1093/imrn/rnx179 doi: 10.1093/imrn/rnx179
    [10] W. Zhai, Manin's conjecture for a class of singular cubic hypersurfaces, Front. Math., 17 (2022), 1089–1132. https://doi.org/10.1007/s11464-021-0945-2 doi: 10.1007/s11464-021-0945-2
    [11] J. Liu, J. Wu, Y. Zhao, On a senary quartic form, Period. Math. Hungar., 80 (2020), 237–248. https://doi.org/10.1007/s10998-019-00308-y doi: 10.1007/s10998-019-00308-y
    [12] T. Wen, On the number of rational points on a class of singular hypersurfaces, Period. Math. Hungar., 86 (2023), 621–636. https://doi.org/10.1007/s10998-022-00495-1 doi: 10.1007/s10998-022-00495-1
    [13] G. Tenenbaum, Introduction to analytic and probabilistic number theory, 3 Eds., American Mathematical Society, 2015. https://doi.org/10.1090/gsm/163
    [14] A. Ivić, The Riemann zeta-function: the theory of Riemann zeta-function with applications, New York: John Wiley & Sons, Inc., 1985.
  • Reader Comments
  • © 2024 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(1260) PDF downloads(91) Cited by(0)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog