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

High power sums of Fourier coefficients of holomorphic cusp forms and their applications

  • Let λf(n) be the nth normalized Fourier coefficient of a holomorphic cusp form f for the full modular group. In this paper, we established asymptotic formulae for high power sums of Fourier coefficients of cusp forms and further improved previous results. Moreover, as an application, we studied the signs of the sequences {λf(n)} and {λf(n)λg(n)} in short intervals, and presented some quantitative results for the number of sign changes for nx.

    Citation: Guangwei Hu, Huixue Lao, Huimin Pan. High power sums of Fourier coefficients of holomorphic cusp forms and their applications[J]. AIMS Mathematics, 2024, 9(9): 25166-25183. doi: 10.3934/math.20241227

    Related Papers:

    [1] Fei Hou, Bin Chen . On triple correlation sums of Fourier coefficients of cusp forms. AIMS Mathematics, 2022, 7(10): 19359-19371. doi: 10.3934/math.20221063
    [2] Huafeng Liu, Rui Liu . The sum of a hybrid arithmetic function over a sparse sequence. AIMS Mathematics, 2024, 9(2): 4830-4843. doi: 10.3934/math.2024234
    [3] Fei Hou . On the exponential sums estimates related to Fourier coefficients of $ GL_3 $ Hecke-Maaß forms. AIMS Mathematics, 2023, 8(4): 7806-7816. doi: 10.3934/math.2023392
    [4] Tingting Jiang, Jiantao Jiang, Jing An . An efficient Fourier spectral method and error analysis for the fourth order problem with periodic boundary conditions and variable coefficients. AIMS Mathematics, 2023, 8(4): 9585-9601. doi: 10.3934/math.2023484
    [5] Stefano Bonaccorsi, Bernard Hanzon, Giulia Lombardi . A generalized Budan-Fourier approach to generalized Gaussian and exponential mixtures. AIMS Mathematics, 2024, 9(10): 26499-26537. doi: 10.3934/math.20241290
    [6] Yannan Sun, Wenchao Qian . Fast algorithms for nonuniform Chirp-Fourier transform. AIMS Mathematics, 2024, 9(7): 18968-18983. doi: 10.3934/math.2024923
    [7] Andrey Muravnik . Nonclassical dynamical behavior of solutions of partial differential-difference equations. AIMS Mathematics, 2025, 10(1): 1842-1858. doi: 10.3934/math.2025085
    [8] S. Akansha, Aditya Subramaniam . Exploring Chebyshev polynomial approximations: Error estimates for functions of bounded variation. AIMS Mathematics, 2025, 10(4): 8688-8706. doi: 10.3934/math.2025398
    [9] Siqin Tang, Hong Li . A Legendre-tau-Galerkin method in time for two-dimensional Sobolev equations. AIMS Mathematics, 2023, 8(7): 16073-16093. doi: 10.3934/math.2023820
    [10] Feiting Fan, Xingwu Chen . Dynamical behavior of traveling waves in a generalized VP-mVP equation with non-homogeneous power law nonlinearity. AIMS Mathematics, 2023, 8(8): 17514-17538. doi: 10.3934/math.2023895
  • Let λf(n) be the nth normalized Fourier coefficient of a holomorphic cusp form f for the full modular group. In this paper, we established asymptotic formulae for high power sums of Fourier coefficients of cusp forms and further improved previous results. Moreover, as an application, we studied the signs of the sequences {λf(n)} and {λf(n)λg(n)} in short intervals, and presented some quantitative results for the number of sign changes for nx.



    Let Hk denote the set of all primitive holomorphic cusp forms of even integral weight k2 for the full modular group SL(2,Z). Every fHk has a Fourier expansion at the cusp of the type

    f(z)=n=1λf(n)nk12e2πinz.

    The Fourier coefficient λf(n) satisfies the multiplicative relation

    λf(m)λf(n)=d(m,n)λf(mnd2).

    In 1974, Deligne [1] proved the Ramanujan-Petersson conjecture

    |λf(n)|d(n)nϵ, (1.1)

    where d(n) is the divisor function.

    For fHk, we define the i-th symmetric power L-function attached to f as

    L(symif,s)=pim=0(1αf(p)imβf(p)mps)1 (1.2)

    provided that s>1, where αf(p) and βf(p) are two complex numbers satisfying

    αf(p)βf(p)=|αf(p)|=|βf(p)|=1,λf(p)=αf(p)+βf(p).

    We can express it as a Dirichlet series:

    L(symif,s)=n=1λsymif(n)ns=p(1+v=1λsymif(pv)pvs), (1.3)

    where λsymif(n) is a real multiplicative function, and

    λsymif(p)=im=0αf(p)imβf(p)m=λf(pi).

    It's easy to see that

    {L(sym0f,s)=ζ(s),L(sym1f,s)=L(f,s).

    Let f,gHk be two different cusp forms. The Rankin-Selberg L-function attached to symif and symjg is defined by

    L(symif×symjg,s)=pim=0jn=0(1αf(p)imβf(p)mαg(p)jnβg(p)nps)1

    for s>1. Further, this can also be written as

    L(symif×symjg,s)=n=1λsymif×symjg(n)ns=p(1+v=1λsymif×symjg(pv)pvs). (1.4)

    Then, we get

    λsymif×symjg(p)=im=0jn=0αf(p)imβf(p)mαf(p)jnβf(p)n=λsymif(p)λsymjg(p),

    where i,j1 are integers. In particular, we have

    {L(sym1f×sym1g,s)=L(f×g,s),L(sym1f×symjg,s)=L(f×symjg,s).

    For a more comprehensive investigation on basic properties of symmetric power L-functions and Rankin-Selberg L-functions, the interested readers can refer to [2, Chapter 13].

    There are many hidden structures underlying the Fourier coefficients λf(n). In analytic number theory, it is a classical problem to estimate the sums of the type

    nxλf(n)lλg(n)m, (1.5)

    where l,mN. The study of O-results on the sum (1.5) is of great significance and has attracted much attention of many number theorists. For l=1 and m=0, the best result to date was given by Wu [3]. Rankin [4] and Selberg [5] studied the average behavior of the power sum for the case l=2 and m=0. More recently, Huang [6] established the better result

    nxλf(n)2=Cx+O(x351560+ϵ).

    Fomenko [7] solved the problem when l=3,4 and m=0. Lü [8] improved Fomenko's result and successfully established the results with l=6,8 and m=0 for the first time. Shortly afterward, Lau et al. [9] considered more general cases and obtained better results. Recently, Newton and Thorne [10] proved in a general setting that symif is automorphic for i1. On the basis of the deep results of Newton and Thorne, by applying some techniques of analytic number theory, Xu [11] and Liu [12] investigated the average behavior of the power sums (1.5) with lN and m=0. Hua [13] focused on the sum (1.4) with l9,m=0 over indices that are sums of two squares.

    For the power sum (1.5) with m>0, Ogg [14] first established an asymptotic formula for l,m=1. Subsequently, Fomenko [7] considered the sum of coefficient of the Rankin-Selberg L-function, and then successfully attained the O-results of the case of l=1,m=2 and l=2,m=2. In 2014, Lü[15] showed that

    nxλf(n)λg(n)x35logx23(183π)

    and

    nxλf(n)2λg(n)2=Cx+O(x78+ϵ), (1.6)

    which improved the results of Fomenko [7]. The current best known estimate for (1.6) is due to He [16], who showed that

    nxλf(n)2λg(n)2=Cx+O(x1315+ϵ).

    [17] also investigated the cases of l=m=3,l=4,m=2, and l=4,m=4.

    The first purpose of this paper is to further improve the upper bounds on the error term of the sum (1.5) with m=0 and m2, respectively. The result is formulated in the following theorem:

    Theorem 1.1. Let fHk and gHk be two different nonzero cusp forms:

    (ⅰ) For l=2r6, we have

    nxλf(n)l=xPl(logx)+O(xθl+ϵ),

    where Pl(y) denotes a polynomial in y of degree (lr)(lr1)1, and

    θl={32713391,   l=6,1θ12,r,    l8.

    Here,

    θ2,r=1342r(lr1)+185(r1)(lr2)+12(r2n=1(l2n+1)2n(ln1)+l+1)314.

    (ⅱ) For l2, m2, we have

    nxλf(n)lλg(n)m=xPl,m(logx)+O(xθl,m+ϵ),

    where Pl,m(y) denotes a polynomial in y of degree ((l[l2])(l[l2]1))((m[m2])(l[m2]1))1 for 2l and 2m; otherwise, 0. In the O-term, we have

    θ2,2=773893=0.865621 θ2,3=θ3,2=206221=0.932126 θ2,4=θ4,2=349361=0.966759
    θ2,6=θ6,2=1461114731=0.991853 θ4,4=36773707=0.991907 θ4,6=θ6,4=2994130001=0.998000

    θl,m={1θ12,2,r,ˉr,       l=2r,m=2ˉr,l=m=6,l8 or m8,1θ11,1,t,ˉt,       l=2t+1,m=2ˉt+1,l,m3,1θ11,2,t,ˉr,     l=2t+1,m=2ˉr,l3 or m2,1θ12,1,r,ˉt,      l=2r,m=2ˉt+1,l2 or m3.

    Here,

    θ2,2,r,ˉr=13(lr1)(mˉr1)42rˉr+18(lr2)(mˉr1)5ˉr(r1)+18(lr1)(mˉr2)5r(ˉr1)+r2n1=1(l2n1+1)2ˉrn1(ln11)(mˉr1)+(l+1)1ˉr(mˉr1)2+ˉr2n2=1(m2n2+1)2rn2(lr1)(mn21)(m+1)1r(lr1)2+(l+1)(m+1)2+r1n1=1(l2n1+1)2(m+1)n1(ln11)2+ˉr1n2=1(m2n2+1)2(l+1)n2(mn21)2+r1n1=1ˉr1n2=1(l2n1+1)2(m2n2+1)2n1n2(ln11)(mn21)2314,θ1,1,t,ˉt=(l+1)(m+1)2+tn1=1ˉtn2=1(l2n1+1)2(m2n2+1)2n1n2(ln11)(mn21)2+tn1=1(l2n1+1)2(m+1)n1(ln11)2+ˉtn2=1(m2n2+1)2(l+1)n2(mn21)2,θ1,2,t,ˉr=4(lt1)(mˉr1)3tˉr+81(lt1)(mˉr2)5t(ˉr1)+4(m+1)t(lt1)+ˉr2n2=14(m2n2+1)2tn2(lt1)(mn21)2+(l+1)(m+1)2+t1n1=1(l2n1+1)2(m+1)n1(ln11)2+ˉrn2=1(m2n2+1)2(l+1)n2(mn21)2+t1n1=1ˉrn2=1(l2n1+1)2(m2n2+1)2n1n2(ln11)(mn21)2,θ2,1,r,ˉt=4(lr1)(mˉt1)3rˉt+81(lr2)(mˉt1)5ˉt(r1)+4(l+1)ˉt(mˉt1)+t2n1=14(l2n1+1)2ˉtn1(ln11)(mˉt1)2+(l+1)(m+1)2+rn1=1(l2n1+1)2(m+1)n1(ln11)2+ˉt1n2=1(m2n2+1)2(l+1)n2(mn21)2+rn1=1ˉt1n2=1(l2n1+1)2(m2n2+1)2n1n2(ln11)(mn21)2.

    Remark 1.2 Compared with Liu [12, Theorem 1.1], He [16, Proposition 4.2], and [17, Theorems 1.4 and 1.5], we improve the previous results. In fact, we have

    Old θ6=0.9647 θ8=0.9914 θ2,2=0.8666 θ4,2=0.9687 θ4,4=0.9921
    New θ6=0.9646 θ8=0.9913 θ2,2=0.8656 θ4,2=0.9667 θ4,4=0.9919

    As an application of Theorem 1.1, we then investigate quantitative results of the sign changes of λf(n) and λf(n)λg(n). The sign changes of the sequence of Fourier coefficients in short intervals was first investigated by Murty [18]. Later, Meher and Murty [19] established a lower bound for the number of sign changes of the sequence {λf(n)}. In addition, the analogous questions of simultaneous sign changes of λf(n)λg(n) was considered by Kumari and Murty [20], where f and g are two different cusp forms. In 2019, He[16] improved the result of Kumari and Murty [20]. Here, we obtain the better quantitative results for sign changes of the sequences {λf(n)} and {λf(n)λg(n)}.

    Theorem 1.3. Let λf(n) and λg(n) be the coefficients of L(f,s) and L(g,s), respectively.

    (ⅰ) Suppose fHk. Then, for any r1 with 67112<r1<1, the sequence {λf(n)} has at least one sign change for n(x,x+xr1]. Moreover, the number of sign changes for nx is x1r1 for sufficiently large x.

    (ⅱ) Suppose f,gHk. Then, for any r2 with 773893<r2<1, the sequence {λf(n)λg(n)} has at least one sign change for n(x,x+xr2]. Moreover, the number of sign changes for nx is x1r2 for sufficiently large x.

    Remark 1.4 In view of

    35=0.6>0.5982=67112,   1315=0.8666>0.8656=773893,

    we improve the results of Meher and Murty [19, Theorem 1.2] and He [16, Theorem 1.5].

    In this section, we will recall and establish some preliminary results which are used to prove the main theorems in this paper.

    We define

    Fl(s)=n=1λf(n)lns    and    Fl,m(s)=n=1λf(n)lλg(n)mns. (2.1)

    Lemma 2.1. Let fHk, then we have

    Fl(s)=Gl(s)Hl(s),

    where

    Gl(s)=[l2]n=0L(syml2nf,s)((ln)(ln1)).

    (ln) is the binomial coefficient with the convention that (ln)=0 if n<0, and the function Hl(s) admits a Dirichlet series convergent absolutely in s>1/2. Noting Hl(s)0 for s=1.

    Proof. This can be found in Xu [11, Lemma 5].

    Based on Ivić [21, Theorem 8.4], Bourgain [22, Theorem 5], and Ramachandra and Sankaranarayanan [23, Lemma 2], we give the following lemma:

    Lemma 2.2. For any ϵ>0, we have

    T0|ζ(57+iτ)|12dτϵT1+ϵ (2.2)

    uniformly for T1 and

    ζ(σ+iτ)(|τ+1|)max{1342(1σ),0}+ϵ (2.3)

    uniformly for 1/2σ2 and |τ|1. Moreover, for U>U0, where U0 is sufficiently large, there exists a T(U,2U) such that

    maxσ12|ζ(σ+iT)|ϵexp(C(loglogU)2). (2.4)

    Lemma 2.3. Let fHk, then we have

    T0|L(f,58+iτ)|4dτf,ϵT1+ϵ (2.5)

    uniformly for T1 and

    L(f,σ+iτ)f,ϵ(|τ+1|)max{23(1σ),0}+ϵ, (2.6)
    L(sym2f,σ+iτ)f,ϵ(|τ+1|)max{65(1σ),0}+ϵ, (2.7)
    L(f×sym2g, σ+iτ)f,g,ϵ(|τ+1|)2710(1σ)+ϵ (2.8)

    uniformly for 12σ2 and |τ|1.

    Proof. These are Ivić [24, Theorem 2], Good [25, Corollary 3], Lin et al. [26, Corollary 1.2], and Lin and Sun [27, Corollary 1.3], respectively.

    For general L-functions, we have the following averaged or individual convexity bounds (see [28]):

    Lemma 2.4. Suppose that L(s) is a general L-function of degree m. Then, for any ϵ>0, we have

    2TT|L(σ+iτ)|2dτTm(1σ)+ϵ (2.9)

    uniformly for 1/2σ2 and T1, and

    L(σ+iτ)(|τ|+1)max{m2(1σ),0}+ϵ (2.10)

    uniformly for 1/2σ1+ϵ and |τ|1.

    Remark 2.5 According to the Euler product (1.2), the degree of L(symjf,s) is j+1. In the proof of Theorem 1.1, we take m=j+1 in Lemma 2.4 for L(symjf,s),j3. Similarly, take m=(i+1)(j+1) for the Rankin-Selberg L-function L(symif×symjg,s).

    By the Perron formula ([29, Proposition 5.54]) with (1.1), we obtain

    nxλf(n)l=12πi1+ϵ+iT1+ϵiTFl(s)xssds+Of,ϵ(x1+ϵT)

    uniformly for 2Tx, where the implied constant depends only on f and ϵ. From Lemma 2.1, we can easily get that the point s=1 is the only pole of the integrand in the region σ0σ1+ϵ and |τ|T for any σ0[1/2+ϵ,1). Using the Cauchy residue theorem, we get

    nxλf(n)l=Ress=1Fl(s)xss+12πi(1+ϵ+iTσ0+iT+1+ϵiTσ0iT+σ0+iTσ0iT)Fl(s)xssds+Of,ϵ(x1+ϵT).

    The factorization expression of Fl(s) in Lemma 2.1 contains ζ(s)(ln)(ln1), which means s=1 is a pole of order (ln)(ln1) of Fl(s) in the half-plane s>1/2. Thus, by standard argument in complex analysis, we know the residue at s=1 is equal to xPl(logx), where Pl(logx) is a polynomial of degree (lr)(lr1)1 for l=2r; otherwise, 0. Thus, we get

    nxλf(n)l=xPl(logx)+12πi(1+ϵ+iTσ0+iT+1+ϵiTσ0iT+σ0+iTσ0iT)Fl(s)xssds+Of,ϵ(x1+ϵT). (3.1)

    The absolute convergence of Hl(s) for s>1/2+ϵ yields Hl(s)1. Hence, (3.1) can be written as

    nxλf(n)l=xPl(logx)+Of,ϵ(x1+ϵT+Rhl+Rvl), (3.2)

    where

    Rhl=1T1+ϵσ0|Gl(σ+iT)|xσdσ

    and

    Rvl=xσ0T1|Gl(σ0+iτ)|dττxσ0+ϵsup1T1T1T12T1T1|Gl(σ0+iτ)|dτ.

    Our goal is to test for constraints on Rhl and Rvl, which can certify Rhlx1+ϵ/T and Rvlx1+ϵ/T.

    Let us consider 2l,l6. When the power of ζ(s) is less than 12, (2.2) cannot be used directly. In order to get better results, we consider it separately.

    Case 1. For l=6, according to Lemma 2.1, we have

    G6(s)=ζ(s)5L(sym2f,s)9L(sym4f,s)5L(sym6f,s).

    Taking U=x1203391 in (2.4), there must exist a T(U,2U) such that

    ζ(σ+iT)ϵexp(C(loglogU)2)Uϵ.

    Suppose that T=δU with 1<δ<2. Now, we choose

    σ0=57,  T=T=δU=δx1203391

    in (3.2). Then, we obtain

    Rh61T1+ϵ57(Tδ)5ϵT(9×65+5×52+72)(1σ)+ϵxσdσT1295+ϵ1+ϵ57(xT1345)σdσx1+ϵT+x57+ϵT23335+ϵ. (3.3)

    Applying the Hölder's inequality, we obtain

    Rv6x57+ϵsup1T1T1T12T1T1|G6(57+iτ)|dτx57+ϵsup1T1T1T1(2T1T1|ζ(57+iτ)|12dτ)512(2T1T1|L(sym4f,57+iτ)|607dτ)712×L(sym2f,57+iT1)9L(sym6f,57+iT1).

    By (2.2), (2.7), and Lemma 2.4, we have

    Rv6x57+ϵT1+512(1+ϵ)+(5×712+52×467×712+9×65+72)(157)+ϵx57+ϵT2971420+ϵ. (3.4)

    Combining (3.2)–(3.4), we obtain

    nxλf(n)6=xP6(logx)+O(x1+ϵT+x57+ϵT2971420+ϵ).

    Recall that

    T=δx1203391

    with 1<δ<2. Thus, we get the required result.

    Case 2. When l=2r8, according to Lemma 2.1, we obtain

    Gl(s)=rn=0L(syml2nf,s)((ln)(ln1)).

    Take U=xθ12,r in (2.4), where

    θ2,r=1342r(lr1)+185(r1)(lr2)+r2n=1(l2n+1)2n(ln1)+l+12314.

    Then, there must exist a T(U,2U) such that

    ζ(σ+iT)ϵexp(C(loglogU)2)Uϵ.

    Suppose that T=δU with 1<δ<2. Now, we choose

    σ0=57,  T=T=δU=δxθ12,r

    in (3.2). Then, we obtain

    Rhl1T1+ϵ57(Tδ)1r(lr1)ϵT(185(r1)(lr2)+r2n=1(l2n+1)2n(ln1)+l+12)(1σ)+ϵxσdσTθ2,r1342r(lr1)+3141+ϵ1+ϵ57(xTθ2,r1342r(lr1)+314)σdσx1+ϵT+x57+ϵT27(θ2,r1342r(lr1)+314)1+ϵ. (3.5)

    Applying the Hölder inequality and (2.2), we obtain

    Rvlx57+ϵsup1T1T1T12T1T1|Gl(57+iτ)|dτx57+ϵsup1T1T1T1(2T1T1|ζ(57+iτ)|12dτ)ζ(57+iT1)1341r(lr1)12×r2n=0L(syml2nf,57+iT1)((ln)(ln1))x57+ϵT27θ2,r1+ϵ. (3.6)

    Combining (3.2), (3.5), and (3.6), we obtain

    nxλf(n)l=xPl(logx)+O(x1+ϵT+x57+ϵT27θ2,r1+ϵ).

    Recall that T=δxθ12,r with 1<δ<2. Thus, we get the required result.

    To prove these results, we will use the following proposition:

    Proposition 3.1. Let f,gHk, then we have

    Fl,m(s)=Gl,m(s)Hl,m(s),

    where

    Gl,m(s)=[l2]n1=0[m2]n2=0L(syml2n1f×symm2n2g,s)((ln1)(ln11))((mn2)(mn21)), (3.7)

    (ln1) and (mn2) are the binomial coefficients with the convention that (ln1)=0 and (mn2)=0 if n1,n2<0, and the function Hl,m(s) admits a Dirichlet series convergent absolutely in s>1/2. Note Hl,m(s)0 for s=1.

    Proof. For fHk, we know

    Fl,m(s)=n=1λf(n)lλg(n)mns=p(1+v1λf(pv)lλg(pv)mpvs).

    By [11, Lemma 5], we get that

    λf(p)l=[l2]n1=0((ln1)(ln11))λsyml2n1f(p),λg(p)m=[m2]n2=0((mn2)(mn21))λsymm2n2g(p).

    Hence, the coefficient of ps is

    λf(p)lλg(p)m=([l2]n1=0((ln1)(ln11))λsyml2n1f(p))×([m2]n2=0((mn2)(mn21))λsymm2n2g(p))=[l2]n1=0[m2]n2=0((ln1)(ln11))((mn2)(mn21))λsyml2n1f×symm2n2g(p).

    We define

    Hl,m(s)=Fl,m(s)/Gl,m(s),

    and its p-local factor is of the form 1+O(p2s). So, the Euler product (hence, the Dirichlet series) of Hl,m(s) converges absolutely in s>1/2.

    Utilizing the similar method in Section 3.1 and the decomposition in Proposition 3.1, we can get the claim easily.

    Case 1. When l=m=2, according to Proposition 3.1, we have

    G2,2(s)=ζ(s)L(sym2f,s)L(sym2g,s)L(sym2f×sym2g,s).

    Taking U=x120/893 in (2.4), there must exist a T(U,2U) such that

    ζ(σ+iT)ϵexp(C(loglogU)2)Uϵ.

    Suppose that T=δU with 1<δ<2. Now, we choose

    σ0=57,  T=T=δU=δx120893

    in (3.2). Then,

    Rh2,21T1+ϵ57(Tδ)ϵT(65+65+92)(1σ)+ϵxσdσT5910+ϵ1+ϵ57(xT6910)σdσx1+ϵT+x57+ϵT3435+ϵ. (3.8)

    Applying Hölder's inequality, we obtain

    Rv2,2x57+ϵsup1T1T1T12T1T1|G2,2(57+iτ)|dτx57+ϵsup1T1T1T1(2T1T1|ζ(57+iτ)|12dτ)112(2T1T1|L(sym2f×sym2g,57+iτ)|2dτ)12×(2T1T1|L(sym2f,57+iτ)|125dτ)512L(sym2g,57+iT1).

    By (2.2), (2.7), and Lemma 2.4, we have

    Rv2,2x57+ϵT1+112(1+ϵ)+(92+25×512×65+512×3+65)(157)+ϵx57+ϵT473420+ϵ. (3.9)

    Combining (3.2), (3.8), and (3.9), we obtain

    nxλf(n)2λg(n)2=xP2,2(logx)+O(x1+ϵT+x57+ϵT473420+ϵ).

    Recall that T=δx120893 with 1<δ<2, Thus, we get the required result.

    Utilizing the similar method of λf(n)2λg(n)2, we can get the results of λf(n)2λg(n)4, λf(n)4λg(n)2, λf(n)4λg(n)4, λf(n)4λg(n)6, and λf(n)6λg(n)4 easily.

    Case 2. When l=m=6, l8, or m8, we have

    Gl,m(s)=rn1=0ˉrn2=0L(syml2n1f×symm2n2g,s)((ln1)(ln11))((mn2)(mn21))=ζ(s)((lr)(lr1))((mˉr)(mˉr1))L(sym2f,s)((lr1)(lr2))((mˉr)(mˉr1))L(sym2g,s)((lr)(lr1))((mˉr1)(mˉr2))×r2n1=0L(syml2n1f,s)((ln1)(ln11))((mˉr)(mˉr1))ˉr2n2=0L(symm2n2g,s)((lr)(lr1))((mn2)(mn21))×r1n1=0ˉr1n2=0L(syml2n1f×symm2n2g,s)((ln1)(ln11))((mn2)(mn21)),

    where l=2r, m=2ˉr.

    Taking U=xθ12,2,r,ˉr in (2.4), there must exist a T(U,2U) such that

    ζ(σ+iT)ϵexp(C(loglogU)2)Uϵ.

    Suppose that T=δU with 1<δ<2. Now, we choose

    σ0=57,  T=T=δU=δxθ12,2,r,ˉr,

    where

    θ2,2,r,ˉr=13(lr1)(mˉr1)42rˉr+18(lr2)(mˉr1)5ˉr(r1)+18(lr1)(mˉr2)5r(ˉr1)+r2n1=1(l2n1+1)2ˉrn1(ln11)(mˉr1)+(l+1)1ˉr(mˉr1)2+ˉr2n2=1(m2n2+1)2rn2(lr1)(mn21)(m+1)1r(lr1)2+(l+1)(m+1)2+r1n1=1(l2n1+1)2(m+1)n1(ln11)2+ˉr1n2=1(m2n2+1)2(l+1)n2(mn21)2+r1n1=1ˉr1n2=1(l2n1+1)2(m2n2+1)2n1n2(ln11)(mn21)2314

    in (3.2). Then,

    Rhl,m1T1+ϵ57(Tδ)1rˉr(lr1)(mˉr1)ϵT(θ2,2,r,ˉr1342rˉr(lr1)(mˉr1)+314)(1σ)xσdσT(θ2,2,r,ˉr1342rˉr(lr1)(mˉr1)+314)1+ϵ1+ϵ57(xT(θ2,2,r,ˉr1342rˉr(lr1)(mˉr1)+314))σdσx1+ϵT+x57+ϵT27(θ2,2,r,ˉr1342rˉr(lr1)(mˉr1)+314)1+ϵ. (3.10)

    Applying the Hölder's inequality, we obtain

    Rvl,mx57+ϵsup1T1T1T12T1T1|Gl,m(57+iτ)|dτx57+ϵsup1T1T1T1(2T1T1|ζ(57+iτ)|12dτ)T(θ2,2,r,ˉr72)(157)+ϵ1x57+ϵT27θ2,2,r,ˉr1+ϵ. (3.11)

    Combining (3.2), (3.10), and (3.11), we obtain

    nxλf(n)lλg(n)m=xP1(logx)+O(x1+ϵT+x57+ϵT27θ2,2,r,ˉr1+ϵ).

    Recall that T=δxθ12,2,r,ˉr with 1<δ<2. Thus, we get the required result.

    From (3.7), we have

    Gl,m(s)=tn1=0ˉtn2=0L(syml2n1f×symm2n2g,s)((ln1)(ln11))((mn2)(mn21))=tn1=1L(syml2n1f×symmg,s)((ln1)(ln11))ˉtn2=1L(symlf×symm2n2g,s)((mn2)(mn21))×L(symlf×symmg)tn1=1ˉtn2=1L(syml2n1f×symm2n2g,s)((ln1)(ln11))((mn2)(mn21)),

    where l=2t+1, m=2ˉt+1. Then,

    Rhl,m1T1+ϵ12Tθ1,1,t,ˉt(1σ)xσdσTθ1,1,t,ˉt11+ϵ12(xTθ1,1,t,ˉt)σdσx1+ϵT+x12+ϵT12θ1,1,t,ˉt1, (3.12)

    where

    θ1,1,t,ˉt=(l+1)(m+1)2+tn1=1(l2n1+1)2(m+1)n1(ln11)2+ˉtn2=1(m2n2+1)2(l+1)n2(mn21)2+tn1=1ˉtn2=1(l2n1+1)2(m2n2+1)2n1n2(ln11)(mn21)2.

    Applying the Cauchy inequality and (2.9), we obtain

    Rvl,mx12+ϵsup1T1T1T12T1T1|Gl,m(12+iτ)|dτx12+ϵsup1T1T1T12T1T1|L(syml2f×symmg,s)|2dτ×T(θ1,1,t,ˉt2(l1)(m+1)2)(112)+ϵ1x12+ϵT12θ1,1,t,ˉt1+ϵ. (3.13)

    Combining (3.2), (3.12), and (3.13), we obtain

    nxλf(n)lλg(n)m=xPl,m(logx)+O(x1+ϵT+x12+ϵT12θ1,1,t,ˉt1+ϵ).

    By taking a fixed T=xθ11,1,t,ˉt, we obtain the result.

    When 2l,2m, because (2.5) cannot be used directly when the power of L(f,s) is less than 4, we need to think about this case separately.

    Case 1. When l=3,m=2, according to Proposition 3.1 and taking σ0=58, we have

    G3,2(s)=L(f,s)2L(f×sym2g,s)2L(sym3f,s)L(sym3f×sym2g,s).

    Then, we get

    Rh3,2=1T1+ϵ58|L(f,s)2L(f×sym2g,s)2L(sym3f,s)L(sym3f×sym2g,s)|xσdσ1T1+ϵ58T(2×23+2×2710+42+122)(1σ)+ϵxσdσT20615+ϵ1+ϵ58(xT22115)σdσx1+ϵT+x58+ϵT18140+ϵ. (3.14)

    In order to estimate Rv3,2, we apply the Cauchy-Schwarz inequality to obtain

    Rv3,2x58+ϵsup1T1T1T12T1T1|G3,2(58+iτ)|dτx58+ϵsup1T1T1T1T1(2×2710+122)(158)+ϵ(2T1T1|L(symf,58+iτ)|4dτ)12×(2T1T1|L(sym3f,58+iτ)|2dτ)12x58+ϵsup1T1T1T1T1(2×2710+122)(158)+ϵT112+ϵT112×4×(158)+ϵx58+ϵT18140+ϵ. (3.15)

    Combining (3.14) and (3.15) with (3.2) and T=x15221, we get the required result.

    Case 2. When l is odd, m is even, and l3 or m2. From (3.7), we have

    Gl,m(s)=tn1=0ˉrn2=0L(syml2n1f×symm2n2g,s)((ln1)(ln11))((mn2)(mn21))=L(f,s)((lt)(lt1))((mˉr)(mˉr1))t1n1=0ˉrn2=0L(syml2n1f×symm2n2g,s)((ln1)(ln11))((mn2)(mn21))×L(f×sym2g,s)((lt)(lt1))((mˉr1)(mˉr2))ˉr2n2=0L(f×symm2n2g,s)((lt)(lt1))((mn2)(mn21)),

    where l=2t+1, m=2ˉr. By taking σ0=58, then

    Rhl,m1T1+ϵ58Tθ1,2,t,ˉr(1σ)xσdσTθ1,2,t,ˉr11+ϵ58(xTθ1,2,t,ˉr)σdσx1+ϵT+x58+ϵT38θ1,2,t,ˉr1, (3.16)

    where

    θ1,2,t,ˉr=4(lt1)(mˉr1)3tˉr+81(lt1)(mˉr2)5t(ˉr1)+4(m+1)t(lt1)+ˉr2n2=14(m2n2+1)2tn2(lt1)(mn21)2+(l+1)(m+1)2+t1n1=1(l2n1+1)2(m+1)n1(ln11)2+ˉrn2=1(m2n2+1)2(l+1)n2(mn21)2+t1n1=1ˉrn2=1(l2n1+1)2(m2n2+1)2n1n2(ln11)(mn21)2.

    Applying the Hölder inequality, we obtain

    Rvl,mx58+ϵsup1 (3.17)

    Combining (3.16) and (3.17) with (3.2) and T = x^{\theta_{1, 2, t, \bar{r}}^{-1}} , we get the required result.

    The proof of 2\mid l, 2\nmid m is similar to 2\nmid l, 2\mid m , so it can be estimated in a similar way. In order to avoid repetition, we shall not prove it verbatim here.

    In order to prove Theorem 1.3, we state the following result of Meher and Murty [19, Theorem 1.1] to detect sign changes:

    Lemma 4.1. Suppose a sequence of real numbers \{a(n)\} satisfies

    (1) a(n) = O(x^{\alpha+\epsilon}),

    (2) \sum\limits_{n\leq x}a(n) = O(x^{\beta+\epsilon}),

    (3) \sum\limits_{n\leq x}a(n)^2 = cx+O(x^{\gamma+\epsilon}),

    where \alpha, \beta, \gamma , and c are positive real constants. If \alpha+\beta < 1 , then for any r with \max\{\alpha+\beta, \gamma\} < r < 1 , the sequence \{a(n)\} has at least one sign change for n\in (x, x+x^r] for sufficiently large x . Moreover, the number of sign changes of \{a(n)\} for n\leq x is \gg x^{1-r} .

    (ⅰ) Let f\in H_k^* . From Wu [3, Theorem 2] and Huang [6, Theorem 1], we have

    \begin{align} \sum\limits_{n\leq x} \lambda_f(n)\ll x^\frac{1}{3}(\log x)^{-0.118}\ \ \ \ \text{and}\ \ \ \ \sum\limits_{n\leq x} \lambda_f(n)^{2} = Cx + O\big(x^{\frac{3}{5}-\frac{1}{560}+\epsilon} \big). \end{align} (4.1)

    Combining (1.1) and (4.1) in Lemma 4.1, we know that \alpha = 0, \ \beta = {1}/{3}, \ \gamma = {67}/{112} , which means

    \max\{\alpha+\beta,\gamma\} = {67}/{112} < 1.

    (ⅱ) Let f, g\in H_k^* . In 2014, Lü[15] proved that

    \sum\limits_{n\leq x}\lambda_f(n)\lambda_g(n)\ll x^{\frac{3}{5}}\log x ^{-\frac{2}{3}(1-\frac{8}{3\pi})}.

    From Theorem 1.1, we improve the error term for the sharp-cut sum (1.5) with l = m = 2 from O(x^{{13}/{15}+\epsilon}) to O(x^{{773}/{893}+\epsilon}) . With Lemma 4.1, we obtain \alpha = 0, \ \beta = {3}/{5}, \ \gamma = {773}/{893} , which means

    \max\{\alpha+\beta,\gamma\} = {773}/{893} < 1.

    Thus, we finish the proof of Theorem 1.3.

    In this paper, we study the distribution of Fourier coefficients of holomorphic cusp forms. Let \lambda_f(n) be the n th normalized Fourier coefficient of a holomorphic cusp form f for the full modular group. Combining the classical analytic method with property of some primitive automorphic L-functions, we establish asymptotic formulae for high power sums of Fourier coefficients of cusp forms. As an application, we also use a general criteria to detect the signs of \lambda_f(n) and \lambda_f(n)\lambda_g(n) , and obtain some quantitative results for the number of sign changes for n\leq x . We are able to improve or extend previous results.

    Guangwei Hu: writing-review and editing, supervision, validation, methodology, formal analysis, conceptualization, funding acquisition; Huixue Lao: writing-review and editing, resources, methodology, supervision, validation, formal analysis; Huimin Pan: writing-original draft, methodology, validation, formal analysis. All authors read and approved the final manuscript.

    The authors declare they have not used Artificial Intelligence (AI) tools in the creation of this article.

    The authors are very grateful to the referees for some extremely helpful remarks. This work is supported by National Natural Science Foundation of China (Grant No. 12201363) and Natural Science Foundation of Shandong Province (Grant No. ZR2022QA047).

    The authors declare that they have no conflicts of interest.



    [1] P. Deligne, La conjecture de Weil, Publ. Math. Inst. Hautes Etudes Sci., 43 (1974), 273–307. https://doi.org/10.1007/BF02684373
    [2] H. Iwaniec, Topics in classical automorphic forms, American Mathematical Society, 1997. https://doi.org/10.1090/gsm/017
    [3] J. Wu, Power sums of Hecke eigenvalues and application, Acta Arith., 137 (2009), 333–344. https://doi.org/10.4064/aa137-4-3 doi: 10.4064/aa137-4-3
    [4] R. A. Rankin, Contributions to the theory of Ramanujan's function \tau(n) and similar arithmetical functions. Ⅱ. The order of the Fourier coefficients of integral modular forms, Math. Proc. Cambridge Phios. Soc., 35 (1939), 357–372. https://doi.org/10.1017/S0305004100021101 doi: 10.1017/S0305004100021101
    [5] A. Selberg, Bemerkungen über eine Dirichletsche Reihe, die mit der theorie der modulformen nahe verbunden ist, Arch. Math. Naturvid, 43 (1940), 47–50.
    [6] B. R. Huang, On the Rankin-Selberg problem, Math. Ann., 381 (2021), 1217–1251. https://doi.org/10.1007/s00208-021-02186-7 doi: 10.1007/s00208-021-02186-7
    [7] O. M. Fomenko, Fourier coefficients of parabolic forms and automorphic L-functions, J. Math. Sci., 95 (1999), 2295–2316. https://doi.org/10.1007/BF02172473 doi: 10.1007/BF02172473
    [8] G. S. Lü, The sixth and eighth moments of Fourier coefficiehts of cusp forms, J. Number Theory, 129 (2009), 2790–2800. https://doi.org/10.1016/j.jnt.2009.01.019 doi: 10.1016/j.jnt.2009.01.019
    [9] Y. K. Lau, G. S. Lü, J. Wu, Integral power sums of Hecke eigenvalues, Acta. Arith., 150 (2011), 193–207. https://doi.org/10.4064/aa150-2-7 doi: 10.4064/aa150-2-7
    [10] J. Newton, J. A. Thorne, Symmetric power functoriality for holomorphic modular forms, Publ. Math. Inst. Hautes Etudes Sci., 134 (2021), 1–116. https://doi.org/10.1007/s10240-021-00127-3 doi: 10.1007/s10240-021-00127-3
    [11] C. R. Xu, General asymptotic formula of Fourier coefficients of cusp forms over sum of two squares, J. Number Theory, 236 (2022), 214–229. https://doi.org/10.1016/j.jnt.2021.07.017 doi: 10.1016/j.jnt.2021.07.017
    [12] H. F. Liu, On the asymptotic distribution of Fourier coefficients of cusp forms, Bull. Braz. Math. Soc. New Ser., 54 (2023), 21. https://doi.org/10.1007/s00574-023-00335-x doi: 10.1007/s00574-023-00335-x
    [13] G. D. Hua, On the higher power moments of cusp form coefficients over sums of two squares, Czech. Math. J., 72 (2022), 1089–1104. https://doi.org/10.21136/CMJ.2022.0358-21 doi: 10.21136/CMJ.2022.0358-21
    [14] A. P. Ogg, On a convolution of L-series, Invent. Math., 7 (1969), 297–312. https://doi.org/10.1007/BF01425537 doi: 10.1007/BF01425537
    [15] G. S. Lü, Sums of absolute values of cusp form coefficients and their application, J. Number Theory, 139 (2014), 29–43. https://doi.org/10.1016/j.jnt.2013.12.011 doi: 10.1016/j.jnt.2013.12.011
    [16] X. G. He, On sign change of Fourier coefficients of cusp forms, Ph.D. thesis, Shandong University, 2019.
    [17] G. S. Lü, On higher moments of Fourier coefficients of holomorphic cusp form, Canad. J. Math., 63 (2011), 643–647. https://doi.org/10.4153/CJM-2011-010-5 doi: 10.4153/CJM-2011-010-5
    [18] M. R. Murty, Oscillations of Fourier coefficients of modular forms, Math. Ann., 262 (1983), 431–446. https://doi.org/10.1007/BF01456059 doi: 10.1007/BF01456059
    [19] J. Meher, M. R. Murty, Sign changes of Fourier coefficients of half-integral weight cusp forms, Int. J. Number Theory, 10 (2014), 905–914. https://doi.org/10.1142/S1793042114500067 doi: 10.1142/S1793042114500067
    [20] M. Kumari, M. R. Murty, Simultaneous non-vanishing and sign changes of Fourier coefficients of modular forms, Int. J. Number Theory, 14 (2018), 2291–2301. https://doi.org/10.1142/S1793042118501397 doi: 10.1142/S1793042118501397
    [21] A. Ivić, Exponent pairs and the zeta function of Riemann, Stud. Sci. Math. Hung., 15 (1980), 157–181.
    [22] J. Bourgain, Decoupling, exponential sums and the Riemann zeta function, J. Amer. Math. Soc., 30 (2017), 205–224. https://doi.org/10.1090/jams/860 doi: 10.1090/jams/860
    [23] K. Ramachandra, A. Sankaranarayanan, Notes on the Riemann zeta-function, J. Indian Math. Soc., 57 (1991), 67–77.
    [24] A. Ivić, On zeta-functions associated with Fourier coefficients of cusp forms, Proceedings of the Amalifi Conference on Analytic Number Theory, 1989,231–246.
    [25] A. Good, The square mean of Dirichlet series associated with cusp forms, Mathematika, 29 (1982), 278–295. https://doi.org/10.1112/S0025579300012377 doi: 10.1112/S0025579300012377
    [26] Y. Lin, R. Nunes, Z. Qi, Strong subconvexity for self-dual GL(3) L-functions, Int. Math. Res. Not., 2023 (2023), 11453–11470. https://doi.org/10.1093/imrn/rnac153 doi: 10.1093/imrn/rnac153
    [27] Y. Lin, Q. Sun, Analytic twists of GL_3\times GL_2 automorphic forms, Int. Math. Res. Not., 2021 (2021), 15143–15208. https://doi.org/10.1093/imrn/rnaa348 doi: 10.1093/imrn/rnaa348
    [28] A. Perelli, General L-functions, Ann. Mat. Pura Appl., 130 (1982), 287–306. https://doi.org/10.1007/BF01761499
    [29] H. Iwaniec, E. Kowalski, Analytic number theory, American Mathematical Society, 2004.
  • 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(793) PDF downloads(64) Cited by(0)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog