This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

On distinguishing Siegel cusp forms of degree two

Zhining Wei and Shaoyun Yi
Abstract

In this work we establish several results on distinguishing Siegel cusp forms of degree two. In particular, a Hecke eigenform of level one can be determined by its second Hecke eigenvalue under a certain assumption. Moreover, we can also distinguish two Hecke eigenforms of level one by using LL-functions.

2020 Mathematics Subject Classification: Primary 11F46, 11F60, 11F66
      Key words and phrases. Siegel cusp forms, Hecke eigenforms, LL-functions.

1 Introduction

One of the fundamental problems in the theory of automorphic forms is whether we can distinguish them by a set of eigenvalues. It is well known that in the elliptic modular forms case, this question is equivalent to asking how many Fourier coefficients are sufficient to determine a normalized eigenform. This question is answered first by the classical result of Sturm [25]. In 2011, Ghitza [6] obtains a result by considering two cuspidal Hecke eigenforms of distinct weights, which improves a result of Ram Murty [18]. Later, Vilardi and Xue [26] give a much stronger result that two normalized eigenforms of full level can be determined by their second coefficients under the assumption of Maeda’s conjecture for the Hecke operator T2T_{2}. Recently, Xue and Zhu [28] generalize this result in terms of their third coefficients under a similar assumption.

However, distinguishing Siegel cusp forms is a long-standing unanswered problem and only recently Schmidt [24], in a remarkable paper, gives an affirmative answer to this question for normalized eigenvalues of a Siegel cuspidal eigenform of degree two. This result has been improved by Kumar, Meher and Shankhadhar [13] in the full level case, in which they essentially show that any set of eigenvalues (normalized or non-normalized) at primes pp of positive upper density are sufficient to determine the Siegel cuspidal eigenform. In this work we further investigate the question on distinguishing Siegel cusp forms of degree two from various aspects with several improved results. We point out that we have also discussed the similar question for paramodular forms with the combination of the methods from both of automorphic side and Galois side in [27].

Let 𝒮k(Γ0(N))\mathcal{S}_{k}(\Gamma_{0}(N)) be the space of Siegel cusp form of level Γ0(N)\Gamma_{0}(N) and weight kk, where Γ0(N)\Gamma_{0}(N) is the Siegel congruence subgroup of level NN defined as in (11). Let F𝒮k(Γ0(N))F\in\mathcal{S}_{k}(\Gamma_{0}(N)) be a Hecke eigenform with eigenvalue λF(n)\lambda_{F}(n) for (n,N)=1(n,N)=1. Then our first main result is as follows.

Theorem 1.1.

Let k1,k2k_{1},k_{2} be distinct positive integers. Let F𝒮k1(Γ0(N))F\in\mathcal{S}_{k_{1}}(\Gamma_{0}(N)) and G𝒮k2(Γ0(N))G\in\mathcal{S}_{k_{2}}(\Gamma_{0}(N)) be Hecke eigenforms. Then we can find nn satisfying

n(2logN+2)4n\leq(2\log N+2)^{4} (1)

such that λF(n)λG(n)\lambda_{F}(n)\neq\lambda_{G}(n).

Remark 1.

It is shown in [9, Corollary 5.3] that there exists some nn satisfying n(2logN+2)6n\leq(2\log N+2)^{6} such that λF(n)λG(n)\lambda_{F}(n)\neq\lambda_{G}(n). In particular, we obtain an improved bound for nn in Theorem 1.1.

Next, we assume that N=1N=1, and let Γ2=Sp(4,)\Gamma_{2}={\rm Sp}(4,{\mathbb{Z}}). It is well known that the space 𝒮k(Γ2)\mathcal{S}_{k}(\Gamma_{2}) has a natural decomposition into orthogonal subspaces

𝒮k(Γ2)=𝒮k(𝐏)(Γ2)𝒮k(𝐆)(Γ2)\mathcal{S}_{k}(\Gamma_{2})=\mathcal{S}_{k}^{\mathbf{(P)}}(\Gamma_{2})\oplus\mathcal{S}_{k}^{\mathbf{(G)}}(\Gamma_{2}) (2)

with respect to the Petersson inner product. Here, 𝒮k(𝐏)(Γ2)\mathcal{S}_{k}^{\mathbf{(P)}}(\Gamma_{2}) is the subspace of Saito-Kurokawa liftings, and 𝒮k(𝐆)(Γ2)\mathcal{S}_{k}^{\mathbf{(G)}}(\Gamma_{2}) is the subspace of non-liftings. We refer the reader to [24, § 2.1] for more details about this type decomposition. For our purpose, let

𝒦()(2)={k:The characteristic polynomial of Tk(2) for F𝒮k()(Γ2) is irreducible},\mathcal{K}^{\mathbf{(*)}}(2)=\{k\in{\mathbb{Z}}\colon\text{The characteristic polynomial of }T_{k}(2)\text{ for }F\in\mathcal{S}_{k}^{\mathbf{(*)}}(\Gamma_{2})\text{ is irreducible}\}, (3)

where 𝒮k()(Γ2)\mathcal{S}_{k}^{\mathbf{(*)}}(\Gamma_{2}) is the set of those F𝒮k(Γ2)F\in\mathcal{S}_{k}(\Gamma_{2}) of type ()\mathbf{(*)} as in (2). Then we can prove the following result.

Theorem 1.2.

Let k1,k2𝒦(𝐏)(2)𝒦(𝐆)(2)k_{1},k_{2}\in\mathcal{K}^{\mathbf{(P)}}(2)\cap\mathcal{K}^{\mathbf{(G)}}(2) be two even positive integers, where k1k_{1} and k2k_{2} may equal. Let F𝒮k1(Γ2)F\in\mathcal{S}_{k_{1}}(\Gamma_{2}) and G𝒮k2(Γ2)G\in\mathcal{S}_{k_{2}}(\Gamma_{2}) be Hecke eigenforms. If λF(2)=λG(2)\lambda_{F}(2)=\lambda_{G}(2), then F=cGF=c\cdot G for some non-zero constant cc.

Remark 2.

We point out that 𝒦()(2)\mathcal{K}^{\mathbf{(*)}}(2) defined as in (3) is a weak version of the generalized Maeda’s conjecture for Γ2=Sp(4,)\Gamma_{2}={\rm Sp}(4,{\mathbb{Z}}). In fact, Maeda’s conjecture for Γ1=SL(2,)\Gamma_{1}={\rm SL}(2,{\mathbb{Z}}) would imply that 𝒦(𝐏)(2)={k:k even and k10}\mathcal{K}^{\mathbf{(P)}}(2)=\{k\colon\mbox{$k$ even and $k\geq 10$}\}. Moreover, it is expected that the set 𝒦(𝐆)(2)\mathcal{K}^{\mathbf{(G)}}(2) has the natural density of 11. See [10, 8] for more discussions about Maeda’s conjecture.

In addition, we can also distinguish Hecke eigenforms in each type by using LL-functions with different methods. First, recall that Saito-Kurokawa liftings of level one and weight k2>0k\in 2{\mathbb{Z}}_{>0} can be obtained from elliptic cusp forms of level one and weight 2k22k-2. More precisely, let f𝒮2k2(Γ1)f\in\mathcal{S}_{2k-2}(\Gamma_{1}) be a Hecke eigenform, and let πf\pi_{f} be the cuspidal automorphic representation of GL(2,𝔸){\rm GL}(2,{\mathbb{A}}) associated to ff. Here, 𝔸{\mathbb{A}} is the ring of adeles of {\mathbb{Q}}. Then the resulting Saito-Kurokawa lifting is in 𝒮k(Γ2)\mathcal{S}_{k}(\Gamma_{2}), denoted by FfF_{f}. It is well known that FfF_{f} is also a Hecke eigenform; see [15, 17] for more details about the classical Saito-Kurokawa liftings. The normalized spinor LL-function of FfF_{f} and the normalized LL-function of ff are connected by the following relation:

L(s,πFf,ρ4)=ζ(s+1/2)ζ(s1/2)L(s,πf),L(s,\pi_{F_{f}},\rho_{4})=\zeta(s+1/2)\zeta(s-1/2)L(s,\pi_{f}), (4)

where ρ4\rho_{4} is the 44-dimensional irreducible representation of Sp(4,){\rm Sp}(4,{\mathbb{C}}), and πFf\pi_{F_{f}} is the cuspidal automorphic representation of GSp(4,𝔸){\rm GSp}(4,{\mathbb{A}}) corresponding to FfF_{f}. Let ξ\xi be a primitive Dirichlet character, and let χ\chi be the corresponding Hecke character of ×\𝔸×{\mathbb{Q}}^{\times}\backslash{\mathbb{A}}^{\times}. Let σ1\sigma_{1} be the standard representation of the dual group GL(1,)=×{\rm GL}(1,{\mathbb{C}})={\mathbb{C}}^{\times}. Then we can define the twisted LL-function by

L(s,πFf×χ,ρ4σ1)=L(s+1/2,χ)L(s1/2,χ)L(s,πf×χ).L(s,\pi_{F_{f}}\times\chi,\rho_{4}\otimes\sigma_{1})=L(s+1/2,\chi)L(s-1/2,\chi)L(s,\pi_{f}\times\chi). (5)
Proposition 1.3.

Let k1,k2k_{1},k_{2} be even positive integers and f𝒮2k12(Γ1),g𝒮2k22(Γ1)f\in\mathcal{S}_{2k_{1}-2}(\Gamma_{1}),g\in\mathcal{S}_{2k_{2}-2}(\Gamma_{1}) be normalized Hecke eigenforms. Suppose that there exists a non-zero constant cc such that

L(1/2,πFf×χd,ρ4σ1)=cL(1/2,πFg×χd,ρ4σ1)L(1/2,\pi_{F_{f}}\times\chi_{d},\rho_{4}\otimes\sigma_{1})=c\cdot L(1/2,\pi_{F_{g}}\times\chi_{d},\rho_{4}\otimes\sigma_{1}) (6)

for almost all quadratic Hecke characters χd\chi_{d} of ×\𝔸×{\mathbb{Q}}^{\times}\backslash{\mathbb{A}}^{\times}, which are corresponding to primitive quadratic Dirichlet characters ξd\xi_{d} of conductor dd. Then k1=k2k_{1}=k_{2} and Ff=FgF_{f}=F_{g}.

To prove this proposition, it suffices to show that f=gf=g, which is due to [16, Theorem B].

Finally, we will distinguish Hecke eigenforms of type (𝐆)\mathbf{(G)} by using Rankin-Selberg LL-functions under the Generalized Riemann hypothesis. Let k1,k2k_{1},k_{2} be even integers. Let F𝒮k1(𝐆)(Γ2)F\in\mathcal{S}_{k_{1}}^{\mathbf{(G)}}(\Gamma_{2}) and G𝒮k2(𝐆)(Γ2)G\in\mathcal{S}_{k_{2}}^{\mathbf{(G)}}(\Gamma_{2}) be Hecke eigenforms, and let πF\pi_{F} (resp. πG\pi_{G}) be the cuspidal automorphic representation of GSp(4,𝔸){\rm GSp}(4,{\mathbb{A}}) corresponding to FF (resp. GG). Then we can define the Rankin-Selberg LL-function of FF and GG, denoted by L(s,πF×πG,ρiρj)L(s,\pi_{F}\times\pi_{G},\rho_{i}\otimes\rho_{j}) with i,j{4,5}i,j\in\{4,5\}; see [20, (271)]. Note that the LL-function here is actually the finite part of LL-functions in [20]. Moreover, L(s,πF×πG,ρiρj)L(s,\pi_{F}\times\pi_{G},\rho_{i}\otimes\rho_{j}) has a simple pole at s=1s=1 if and only if i=j,k1=k2i=j,k_{1}=k_{2} and F=cGF=c\cdot G for some non-zero constant cc; see [20, Theorem 5.2.3]. In this case, we can prove the following result:

Theorem 1.4.

Assume the notations above. Suppose that L(s,πF×πG,ρ4ρ4)L(s,\pi_{F}\times\pi_{G},\rho_{4}\otimes\rho_{4}) and L(s,πF×πF,ρ4ρ4)L(s,\pi_{F}\times\pi_{F},\rho_{4}\otimes\rho_{4}) satisfy the Generalized Riemann hypothesis. If FF is not a scalar multiplication of GG, then there exists an integer

n(logk1k2)2(loglogk1k2)4n\ll(\log k_{1}k_{2})^{2}(\log\log k_{1}k_{2})^{4} (7)

such that λ~F(n)λ~G(n)\tilde{\lambda}_{F}(n)\neq\tilde{\lambda}_{G}(n). Here, λ~F(n)=n3/2k1λF(n)\tilde{\lambda}_{F}(n)=n^{3/2-k_{1}}\lambda_{F}(n) (resp. λ~G(n)=n3/2k2λG(n)\tilde{\lambda}_{G}(n)=n^{3/2-k_{2}}\lambda_{G}(n)) is the normalized Hecke eigenvalue for FF (resp. GG).

To prove above theorem, we will apply the method in [5]. Then combine with Lemma 5.1 and we will conclude Theorem 1.4.

Acknowledgements

We thank Wenzhi Luo, Kimball Martin, Ralf Schmidt, Biao Wang, Pan Yan and Liyang Yang for their helpful discussions and comments. We thank Ariel Weiss for drawing our attention to the low weights case (ki=1,2k_{i}=1,2) in Theorem 1.1. We thank Biplab Paul for forwarding us their paper [7]. Shaoyun Yi is supported by the National Natural Science Foundation of China (No. 12301016) and the Fundamental Research Funds for the Central Universities (No. 20720230025).

2 Preliminaries

We consider the symplectic similitude group

GSp(4){gGL(4):tgJg=μ(g)J,μ(g)GL(1)},{\rm GSp}(4)\coloneqq\{g\in{\rm GL}(4)\colon\>^{t}gJg=\mu(g)J,\>\mu(g)\in{\rm GL}(1)\}, (8)

which is an algebraic {\mathbb{Q}}-group. Here, J=[1111]J=\begin{bmatrix}&&&1\\ &&1&\\ &-1&&\\ -1&&&\end{bmatrix}. The function μ\mu is called the multiplier homomorphism. The kernel of this function is the symplectic group Sp(4){\rm Sp}(4). Let Z\mathrm{Z} be the center of GSp(4){\rm GSp}(4) and PGSp(4)=GSp(4)/Z{\rm PGSp}(4)={\rm GSp}(4)/\mathrm{Z}. When speaking about Siegel modular forms of degree two, it is more convenient to realize symplectic groups using the symplectic form J=[012120]J=\begin{bmatrix}0&1_{2}\\ -1_{2}&0\end{bmatrix}. The Siegel upper half plane of degree 2 is defined by

2{ZMat2():tZ=Z,Im(Z)>0}.\mathbb{H}_{2}\coloneqq\{Z\in\mathrm{Mat}_{2}({\mathbb{C}})\colon\,^{t}Z=Z,\mathrm{Im}(Z)>0\}. (9)

The group GSp(4,)+{gGSp(4,):μ(g)>0}{\rm GSp}(4,\mathbb{R})^{+}\coloneqq\{g\in{\rm GSp}(4,{\mathbb{R}})\colon\mu(g)>0\} acts on 2\mathbb{H}_{2} by

gZ(AZ+B)(CZ+D)1for g=[ABCD]GSp(4,)+ and Z2.g\langle Z\rangle\coloneqq(AZ+B)(CZ+D)^{-1}\quad\text{for }g=\begin{bmatrix}A&B\\ C&D\end{bmatrix}\in{\rm GSp}(4,\mathbb{R})^{+}\text{ and }Z\in\mathbb{H}_{2}. (10)

Let Γ2=Sp(4,)\Gamma_{2}={\rm Sp}(4,{\mathbb{Z}}). In general, for a positive integer NN we let

Γ0(N){[ABCD]Sp(4,):C0(modN)}\Gamma_{0}(N)\coloneqq\left\{\begin{bmatrix}A&B\\ C&D\end{bmatrix}\in{\rm Sp}(4,{\mathbb{Z}})\colon C\equiv 0\pmod{N}\right\} (11)

be the Siegel congruence subgroup of level NN. It is clear that Γ2=Γ0(1)\Gamma_{2}=\Gamma_{0}(1).

Let k(Γ0(N))\mathcal{M}_{k}(\Gamma_{0}(N)) be the space of Siegel modular form of weight kk with respect to Γ0(N)\Gamma_{0}(N), and let 𝒮k(Γ0(N))\mathcal{S}_{k}(\Gamma_{0}(N)) be the subspace of cusp forms. That is to say, for any function Fk(Γ0(N))F\in\mathcal{M}_{k}(\Gamma_{0}(N)), it is a holomorphic {\mathbb{C}}-valued function on 2\mathbb{H}_{2} satisfying (F|kγ)(Z)=F(Z)\big{(}F|_{k}\gamma\big{)}(Z)=F(Z) for all γΓ0(N)\gamma\in\Gamma_{0}(N). Here,

(F|kg)(Z)μ(g)kj(g,Z)kF(gZ)for g=[ABCD]GSp(4,)+ and Z2,\big{(}F|_{k}g\big{)}(Z)\coloneqq\mu(g)^{k}j(g,Z)^{-k}F(g\langle Z\rangle)\quad\text{for }g=\begin{bmatrix}A&B\\ C&D\end{bmatrix}\in{\rm GSp}(4,\mathbb{R})^{+}\text{ and }Z\in\mathbb{H}_{2}, (12)

where j(g,Z)det(CZ+D)j(g,Z)\coloneqq\operatorname{det}(CZ+D) is the automorphy factor. We remark that this operator differs from the classical one used in [1] by a factor. We do so to make the center of GSp(4,)+{\rm GSp}(4,\mathbb{R})^{+} act trivially.

Let F𝒮k(Γ0(N))F\in\mathcal{S}_{k}(\Gamma_{0}(N)) be a Hecke eigenform, i.e., it is an eigenvector for all the Hecke operator T(n),(n,N)=1T(n),(n,N)=1. Denote by λF(n)\lambda_{F}(n) the eigenvalue of FF under T(n)T(n) when (n,N)=1(n,N)=1. For any prime pNp\nmid N, we let αp,0,αp,1,αp,2\alpha_{p,0},\alpha_{p,1},\alpha_{p,2} be the classical Satake parameters of FF at pp. It is well known that

αp,02αp,1αp,2=p2k3.\alpha_{p,0}^{2}\alpha_{p,1}\alpha_{p,2}=p^{2k-3}. (13)

In particular, let N=1N=1 and F𝒮k(Γ2)F\in\mathcal{S}_{k}(\Gamma_{2}) be a Hecke eigenform, we can define the LL-series

H(s)=n=1λF(n)ns.H(s)=\sum_{n=1}^{\infty}\frac{\lambda_{F}(n)}{n^{s}}. (14)

This can be written as a Euler product

H(s)=pHp(s)=p(1+λF(p)ps+λF(p2)p2s+)H(s)=\prod_{p}H_{p}(s)=\prod_{p}\left(1+\frac{\lambda_{F}(p)}{p^{s}}+\frac{\lambda_{F}(p^{2})}{p^{2s}}+\cdots\right) (15)

provided (s)>k\Re(s)>k. Moreover, one can show that

Hp(s)=(1p2k42s)Lp(s,F,spin),H_{p}(s)=\big{(}1-p^{2k-4-2s}\big{)}L_{p}(s,F,\operatorname{spin}), (16)

where Lp(s,F,spin)L_{p}(s,F,\operatorname{spin}) is the local spinor LL-factor of FF at pp and it can be given by

Lp(s,F,spin)1=(1αp,0ps)(1αp,0αp,1ps)(1αp,0αp,2ps)(1αp,0αp,1αp,2ps).L_{p}(s,F,\operatorname{spin})^{-1}=(1-\alpha_{p,0}p^{-s})(1-\alpha_{p,0}\alpha_{p,1}p^{-s})(1-\alpha_{p,0}\alpha_{p,2}p^{-s})(1-\alpha_{p,0}\alpha_{p,1}\alpha_{p,2}p^{-s}). (17)

On the other hand, by [1, pp. 62, 69] one can see that

Lp(s,F,spin)1=1λF(p)ps+(λF(p)2λF(p2)p2k4)p2sλF(p)p2k33s+p4k64s.L_{p}(s,F,\operatorname{spin})^{-1}=1-\lambda_{F}(p)p^{-s}+(\lambda_{F}(p)^{2}-\lambda_{F}(p^{2})-p^{2k-4})p^{-2s}-\lambda_{F}(p)p^{2k-3-3s}+p^{4k-6-4s}. (18)

In this case, we can define the spinor LL-function

L(s,F,spin)=pLp(s,F,spin).L(s,F,\operatorname{spin})=\prod_{p}L_{p}(s,F,\operatorname{spin}). (19)

Let αp=p3/2kαp,0\alpha_{p}=p^{3/2-k}\alpha_{p,0} and βp=αpαp,1\beta_{p}=\alpha_{p}\alpha_{p,1}. By comparing (17) with (18), we obtain (also see [19, Proposition 4.1])

λF(p)\displaystyle\lambda_{F}(p) =pk3/2(αp+αp1+βp+βp1),\displaystyle=p^{k-3/2}(\alpha_{p}+\alpha_{p}^{-1}+\beta_{p}+\beta_{p}^{-1}), (20)
λF(p2)\displaystyle\lambda_{F}(p^{2}) =p2k3((αp+αp1)2+(αp+αp1)(βp+βp1)+(βp+βp1)221/p).\displaystyle=p^{2k-3}\left((\alpha_{p}+\alpha_{p}^{-1})^{2}+(\alpha_{p}+\alpha_{p}^{-1})(\beta_{p}+\beta_{p}^{-1})+(\beta_{p}+\beta_{p}^{-1})^{2}-2-1/p\right). (21)

Let ρ4\rho_{4} be the 44-dimensional irreducible representation of Sp(4,){\rm Sp}(4,{\mathbb{C}}). In fact, ρ4\rho_{4} is the natural representation of Sp(4,){\rm Sp}(4,{\mathbb{C}}) on 4{\mathbb{C}}^{4}, which is also called the spin representation. For later use, we would normalize the spinor LL-function. More precisely, the normalized spinor LL-function L(s,πF,ρ4)L(s,\pi_{F},\rho_{4}) is defined as follows

L(s,πF,ρ4)=L(s+k3/2,F,spin)=n=1aF(n)ns.L(s,\pi_{F},\rho_{4})=L(s+k-3/2,F,\operatorname{spin})=\sum_{n=1}^{\infty}\frac{a_{F}(n)}{n^{s}}. (22)

Note that this is the finite part of the completed LL-function of πF\pi_{F}, where πF\pi_{F} is the cuspidal automorphic representation of GSp(4,𝔸){\rm GSp}(4,{\mathbb{A}}) associated to FF. For more details about the connection between Siegel modular forms of degree two and automorphic representations of GSp(4,𝔸){\rm GSp}(4,{\mathbb{A}}); for example see [2] and [23, Section 4.2]. Moreover, let λ~F(n)=n3/2kλF(n)\tilde{\lambda}_{F}(n)=n^{3/2-k}\lambda_{F}(n) be the normalized eigenvalues. It follows that

n=1λ~F(n)ns=ζ(2s+1)1L(s,πF,ρ4).\sum_{n=1}^{\infty}\frac{\tilde{\lambda}_{F}(n)}{n^{s}}=\zeta(2s+1)^{-1}L(s,\pi_{F},\rho_{4}). (23)

Here, ζ\zeta is the Riemann zeta function. Note that if F𝒮k(Γ0(N))F\in\mathcal{S}_{k}(\Gamma_{0}(N)) with N>1N>1, we still can define the partial spinor LL-functions by Euler products for all primes pp not dividing NN. In particular, the local factor at pp with (p,N)=1(p,N)=1 is defined in the same way as above.

Similarly, let ρ5\rho_{5} be the 55-dimensional irreducible representation of Sp(4,){\rm Sp}(4,{\mathbb{C}}). An explicit formula for the representation ρ5\rho_{5} as a map Sp(4,)SO(5,){\rm Sp}(4,{\mathbb{C}})\to{\rm SO}(5,{\mathbb{C}}) is given in [21, Appendix A.7]. The standard LL-function associated to FF is defined as

L(s,πF,ρ5)=pLp(s,F,std)=n=1bF(n)ns,L(s,\pi_{F},\rho_{5})=\prod_{p}L_{p}(s,F,\mathrm{std})=\sum_{n=1}^{\infty}\frac{b_{F}(n)}{n^{s}}, (24)

where

Lp(s,F,std)1=(1ps)(1αp,1ps)(1αp,2ps)(1αp,11ps)(1αp,21ps).L_{p}(s,F,\mathrm{std})^{-1}=(1-p^{-s})(1-\alpha_{p,1}p^{-s})(1-\alpha_{p,2}p^{-s})(1-\alpha_{p,1}^{-1}p^{-s})(1-\alpha_{p,2}^{-1}p^{-s}). (25)

Again, for F𝒮k(Γ0(N))F\in\mathcal{S}_{k}(\Gamma_{0}(N)) with N>1N>1, we can define the partial standard LL-functions by Euler products for all primes pp not dividing NN in the same way.

3 Proof of Theorem 1.1

First, by (16) and (18) we have

λF(p3)=2λF(p)λF(p2)λF(p)3+λF(p)(p2k3+p2k4),\lambda_{F}(p^{3})=2\lambda_{F}(p)\lambda_{F}(p^{2})-\lambda_{F}(p)^{3}+\lambda_{F}(p)(p^{2k-3}+p^{2k-4}), (26)

and

λF(p4)=λF(p)4+λF(p)2λF(p2)+λF(p2)2+λF(p)2p2k4+λF(p2)p2k4+2λF(p)2p2k3p4k6.\lambda_{F}(p^{4})=-\lambda_{F}(p)^{4}+\lambda_{F}(p)^{2}\lambda_{F}(p^{2})+\lambda_{F}(p^{2})^{2}+\lambda_{F}(p)^{2}p^{2k-4}+\lambda_{F}(p^{2})p^{2k-4}+2\lambda_{F}(p)^{2}p^{2k-3}-p^{4k-6}. (27)

Then we can prove the following result:

Theorem 3.1.

Let k1,k2k_{1},k_{2} be distinct positive integers. Let F𝒮k1(Γ0(N))F\in\mathcal{S}_{k_{1}}(\Gamma_{0}(N)) and G𝒮k2(Γ0(N))G\in\mathcal{S}_{k_{2}}(\Gamma_{0}(N)) be Hecke eigenforms. Then for any prime pp not dividing NN, we can find i{1,2,3,4}i\in\{1,2,3,4\} such that

λF(pi)λG(pi).\lambda_{F}(p^{i})\neq\lambda_{G}(p^{i}).

To prove this theorem, we need the following lemma.

Lemma 3.2.

Let F𝒮k(Γ0(N))F\in\mathcal{S}_{k}(\Gamma_{0}(N)) be a Hecke eigenform with k>0k\in{\mathbb{Z}}_{>0}. Let pNp\nmid N be a prime. If λF(p)=0\lambda_{F}(p)=0, then

|λF(p2)|p2k2+2p2k4.|\lambda_{F}(p^{2})|\leq p^{2k-2}+2p^{2k-4}. (28)
Proof.

By (20)-(21) and λF(p)=0\lambda_{F}(p)=0, we have

λF(p2)\displaystyle\lambda_{F}(p^{2}) =λF(p)2p2k3((αp+αp1)(βp+βp1)+2+1/p)\displaystyle=\lambda_{F}(p)^{2}-p^{2k-3}((\alpha_{p}+\alpha_{p}^{-1})(\beta_{p}+\beta_{p}^{-1})+2+1/p)
=p2k3((αp+αp1)(αp+αp1)21/p)\displaystyle=p^{2k-3}((\alpha_{p}+\alpha_{p}^{-1})(\alpha_{p}+\alpha_{p}^{-1})-2-1/p)
=p2k3(αp2+αp21/p)=p2k3(βp2+βp21/p).\displaystyle=p^{2k-3}(\alpha_{p}^{2}+\alpha_{p}^{-2}-1/p)=p^{2k-3}(\beta_{p}^{2}+\beta_{p}^{-2}-1/p).

Hence the desired assertion follows immediately from the bound of αp,βp\alpha_{p},\beta_{p} as below:

p1/2|αp|,|βp|p1/2.p^{-1/2}\leq|\alpha_{p}|,|\beta_{p}|\leq p^{1/2}. (29)

To see (29), we separate two cases: i) If FF is not of type (G), then (29) follows from [24, Table 1] (note that the type (F) cannot occur) and the Ramanujan conjecture for elliptic modular forms; ii) If FF is of type (G), then (29) follows from the fact that the cuspidal automorphic representation πF\pi_{F} of PGSp(4,𝔸){\rm PGSp}(4,{\mathbb{A}}) associated to FF admits a functorial transfer to a unitary, cuspidal automorphic representation of GL(4,𝔸){\rm GL}(4,{\mathbb{A}}) and the Jacquet-Shalika bound for GL(4){\rm GL}(4) (see [12, Corollary 2.5]). ∎

Proof of Theorem 3.1.

Assume that there exists a prime pNp\nmid N such that λF(pi)=λG(pi)\lambda_{F}(p^{i})=\lambda_{G}(p^{i}) for i=1,2,3,4i=1,2,3,4; we will obtain a contradiction. More precisely, we consider the following two cases:

(1) If λF(p)0\lambda_{F}(p)\neq 0, then by (26) and λF(pi)=λG(pi)\lambda_{F}(p^{i})=\lambda_{G}(p^{i}) (i=1,2,3i=1,2,3), we have

λF(p)(p2k13+p2k14)=λG(p)(p2k23+p2k24).\lambda_{F}(p)(p^{2k_{1}-3}+p^{2k_{1}-4})=\lambda_{G}(p)(p^{2k_{2}-3}+p^{2k_{2}-4}). (30)

This yields the contradiction k1=k2k_{1}=k_{2}.

(2) If λF(p)=0\lambda_{F}(p)=0, then λG(p)=0\lambda_{G}(p)=0 by assumption. By Lemma 3.2 we have

|λF(p2)|p2k12+2p2k14and|λG(p2)|p2k22+2p2k24.|\lambda_{F}(p^{2})|\leq p^{2k_{1}-2}+2p^{2k_{1}-4}\quad\text{and}\quad|\lambda_{G}(p^{2})|\leq p^{2k_{2}-2}+2p^{2k_{2}-4}. (31)

Without loss of generality, we assume that k1k2+1k_{1}\geq k_{2}+1. Since λF(p2)=λG(p2)\lambda_{F}(p^{2})=\lambda_{G}(p^{2}), we have

|λF(p2)|=|λG(p2)|p2k22+2p2k24.|\lambda_{F}(p^{2})|=|\lambda_{G}(p^{2})|\leq p^{2k_{2}-2}+2p^{2k_{2}-4}. (32)

On the other hand, it follows from (27) and λF(pi)=λG(pi),i=1,2,3,4\lambda_{F}(p^{i})=\lambda_{G}(p^{i}),i=1,2,3,4, that

λF(p2)p2k14p4k16=λG(p2)p2k24p4k26.\lambda_{F}(p^{2})p^{2k_{1}-4}-p^{4k_{1}-6}=\lambda_{G}(p^{2})p^{2k_{2}-4}-p^{4k_{2}-6}. (33)

Then we have λF(p2)(p2k14p2k24)=p4k16p4k26\lambda_{F}(p^{2})(p^{2k_{1}-4}-p^{2k_{2}-4})=p^{4k_{1}-6}-p^{4k_{2}-6}. Multiplying p2p^{2} both sides we obtain

λF(p2)(p2k12p2k22)=p4k14p4k24=(p2k12p2k22)(p2k12+p2k22).\lambda_{F}(p^{2})(p^{2k_{1}-2}-p^{2k_{2}-2})=p^{4k_{1}-4}-p^{4k_{2}-4}=(p^{2k_{1}-2}-p^{2k_{2}-2})(p^{2k_{1}-2}+p^{2k_{2}-2}). (34)

It follows that λF(p2)=p2k12+p2k22\lambda_{F}(p^{2})=p^{2k_{1}-2}+p^{2k_{2}-2}. This equality leads to a contradiction due to (32) and k1k2+1k_{1}\geq k_{2}+1. ∎

Then Theorem 1.1 immediately follows from Theorem 3.1 and Lemma 3.3 below.

Lemma 3.3 ([6, cf. § 2] ).

Let N1N\geq 1 be a positive integer, then we can find a prime pp such that (p,N)=1(p,N)=1 and p2logN+2p\leq 2\log N+2.

4 Proof of Theorem 1.2

In this section, we only consider Γ2=Sp(4,)\Gamma_{2}={\rm Sp}(4,{\mathbb{Z}}). Let mk=dim𝒮k(Γ2)m_{k}=\dim_{{\mathbb{C}}}\mathcal{S}_{k}(\Gamma_{2}). For (){(𝐆),(𝐏)}\mathbf{(*)}\in\{\mathbf{(G)},\mathbf{(P)}\}, recall that the set 𝒦()(2)\mathcal{K}^{\mathbf{(*)}}(2) of integers is defined as in (3). Moreover, we let mk()=dim𝒮k()(Γ2)m_{k}^{\mathbf{(*)}}=\dim_{\mathbb{C}}\mathcal{S}_{k}^{\mathbf{(*)}}(\Gamma_{2}). Evidently, mk=mk(𝐏)+mk(𝐆)m_{k}=m_{k}^{\mathbf{(P)}}+m_{k}^{\mathbf{(G)}}. It is well known that mk(𝐏)>0m_{k}^{\mathbf{(P)}}>0 only if k10k\in{\mathbb{Z}}_{\geq 10} is even and mk(𝐆)>0m_{k}^{\mathbf{(G)}}>0 only if k20k\in{\mathbb{Z}}_{\geq 20}; for example see [22, Theorem 3.1]. We also note that mk(𝐏)=dim𝒮2k2(Γ1)m_{k}^{(\mathbf{P})}=\dim_{\mathbb{C}}\mathcal{S}_{2k-2}(\Gamma_{1}).

Proof of Theorem 1.2.

We are going to separate into the following three cases.

Case I: If FF and GG both are Saito-Kurokawa liftings, say F𝒮k1(𝐏)(Γ2)F\in\mathcal{S}_{k_{1}}^{\mathbf{(P)}}(\Gamma_{2}) and G𝒮k2(𝐏)(Γ2)G\in\mathcal{S}_{k_{2}}^{\mathbf{(P)}}(\Gamma_{2}) with k1,k2𝒦(𝐏)(2)k_{1},k_{2}\in\mathcal{K}^{\mathbf{(P)}}(2), then we can write F=FfF=F_{f} and G=FgG=F_{g}, which are lifts from f𝒮2k12(Γ1)f\in\mathcal{S}_{2k_{1}-2}(\Gamma_{1}) and g𝒮2k22(Γ1)g\in\mathcal{S}_{2k_{2}-2}(\Gamma_{1}), respectively. Recall that if ff and gg are Hecke eigenforms, then both FfF_{f} and FgF_{g} are also Hecke eigenforms. Let T2k2(1)(2)T_{2k-2}^{(1)}(2) be the Hecke operator on f𝒮2k2(Γ1)f\in\mathcal{S}_{2k-2}(\Gamma_{1}) with Hecke eigenvalue λf(2)\lambda_{f}(2), and let Tk(2)T_{k}(2) be the Hecke operator on Ff𝒮k(Γ2)F_{f}\in\mathcal{S}_{k}(\Gamma_{2}) with Hecke eigenvalue λFf(2)\lambda_{F_{f}}(2). Then we have

λFf(2)=2k1+2k2+λf(2).\lambda_{F_{f}}(2)=2^{k-1}+2^{k-2}+\lambda_{f}(2). (35)

Moreover, let P(Tk(𝐏)(2),t)P(T_{k}^{(\mathbf{P})}(2),t) be the characteristic polynomial of Tk(2)T_{k}(2) on 𝒮k(𝐏)(Γ2)\mathcal{S}_{k}^{\mathbf{(P)}}(\Gamma_{2}), which is irreducible if k𝒦(𝐏)(2)k\in\mathcal{K}^{\mathbf{(P)}}(2). For k𝒦(𝐏)(2)k\in\mathcal{K}^{\mathbf{(P)}}(2) we can see that the characteristic polynomial P(T2k2(1)(2),t)P(T_{2k-2}^{(1)}(2),t) of T2k2(1)(2)T_{2k-2}^{(1)}(2) is irreducible as well since P(Tk(𝐏)(2),t)=P(T2k2(1)(2),t2k12k2)P(T_{k}^{(\mathbf{P})}(2),t)=P(T_{2k-2}^{(1)}(2),t-2^{k-1}-2^{k-2}). We can further assume that ff is not a constant multiple of gg since the Saito-Kurokawa lifting is injective.

  1. (i)

    If k1=k2=kk_{1}=k_{2}=k, then λFf(2)λFg(2)\lambda_{F_{f}}(2)\neq\lambda_{F_{g}}(2) due to the fact that the irreducible characteristic polynomial P(Tk(𝐏)(2),t)P(T_{k}^{(\mathbf{P})}(2),t) has distinct roots.

  2. (ii)

    If mk1(𝐏)mk2(𝐏)m_{k_{1}}^{(\mathbf{P})}\neq m_{k_{2}}^{(\mathbf{P})}, then degP(Tk1(𝐏)(2),t)degP(Tk2(𝐏)(2),t)\deg P(T_{k_{1}}^{(\mathbf{P})}(2),t)\neq\deg P(T_{k_{2}}^{(\mathbf{P})}(2),t). Recall that both of them are irreducible, it follows that P(Tk1(𝐏)(2),t)P(T_{k_{1}}^{(\mathbf{P})}(2),t) and P(Tk2(𝐏)(2),t)P(T_{k_{2}}^{(\mathbf{P})}(2),t) have distinct roots. Hence, λFf(2)λFg(2)\lambda_{F_{f}}(2)\neq\lambda_{F_{g}}(2).

  3. (iii)

    If mk1(𝐏)=mk2(𝐏)1m_{k_{1}}^{(\mathbf{P})}=m_{k_{2}}^{(\mathbf{P})}\geq 1 and k1k2k_{1}\neq k_{2}, it is clear that 2k12,2k22182k_{1}-2,2k_{2}-2\geq 18. Additionally, we can show that there exists n1n\geq 1 such that 2k12,2k22{12n+6,12n+10,12n+14}2k_{1}-2,2k_{2}-2\in\{12n+6,12n+10,12n+14\} since k1,k2k_{1},k_{2} are even and mk1(𝐏)=mk2(𝐏)m_{k_{1}}^{(\mathbf{P})}=m_{k_{2}}^{(\mathbf{P})}. On the other hand, we can show that

    TrTk1(𝐏)(2)=mk1(𝐏)(2k11+2k12)+TrT2k12(1)(2).\operatorname{Tr}T_{k_{1}}^{(\mathbf{P})}(2)=m^{(\mathbf{P})}_{k_{1}}(2^{k_{1}-1}+2^{k_{1}-2})+\operatorname{Tr}T_{2k_{1}-2}^{(1)}(2). (36)

    Assume that k1=k2+lk_{1}=k_{2}+l with l>0l>0. Then by the choice of k1,k2k_{1},k_{2}, we know that l{2,4}l\in\{2,4\}. Let l=2ml=2^{m} as in [26, Corollary 3.4], and so m{1,2}m\in\{1,2\}. It follows from mk1(𝐏)=mk2(𝐏)m_{k_{1}}^{(\mathbf{P})}=m_{k_{2}}^{(\mathbf{P})} and (36) that

    TrTk1(𝐏)(2)TrTk2(𝐏)(2)=2k22(mk2(𝐏)(2k1k2+1+2k1k23)+ak21,l2m+5k2ck21,l),\operatorname{Tr}T_{k_{1}}^{(\mathbf{P})}(2)-\operatorname{Tr}T_{k_{2}}^{(\mathbf{P})}(2)=2^{k_{2}-2}\left(m^{(\mathbf{P})}_{k_{2}}(2^{k_{1}-k_{2}+1}+2^{k_{1}-k_{2}}-3)+a_{k_{2}-1,l}-2^{m+5-k_{2}}c_{k_{2}-1,l}\right),

    where ak21,la_{k_{2}-1,l} is an integer and ck21,lc_{k_{2}-1,l} is an odd integer as in the proof of [26, Corollary 3.4]. However, we know that m+5k23m+5-k_{2}\leq-3 since m2m\leq 2 and k210k_{2}\geq 10. Therefore, 2m+5k2ck21,l2^{m+5-k_{2}}c_{k_{2}-1,l} is not an integer and hence TrTk1(𝐏)(2)TrTk2(𝐏)(2)\operatorname{Tr}T_{k_{1}}^{(\mathbf{P})}(2)\neq\operatorname{Tr}T_{k_{2}}^{(\mathbf{P})}(2). By irreducibility of characteristic polynomials TrTk1(𝐏)(2)=TrλFf(2)\operatorname{Tr}T_{k_{1}}^{(\mathbf{P})}(2)=\operatorname{Tr}\lambda_{F_{f}}(2) and TrTk2(𝐏)(2)=TrλFg(2)\operatorname{Tr}T_{k_{2}}^{(\mathbf{P})}(2)=\operatorname{Tr}\lambda_{F_{g}}(2), which implies that λFf(2)λFg(2)\lambda_{F_{f}}(2)\neq\lambda_{F_{g}}(2).

Case II: If FF and GG both are non-liftings, say F𝒮k1(𝐆)(Γ2)F\in\mathcal{S}_{k_{1}}^{\mathbf{(G)}}(\Gamma_{2}) and G𝒮k2(𝐆)(Γ2)G\in\mathcal{S}_{k_{2}}^{\mathbf{(G)}}(\Gamma_{2}), we can apply the similar arguments as in the above Case I. More precisely, if k1=k2𝒦(𝐆)(2)k_{1}=k_{2}\in\mathcal{K}^{\mathbf{(G)}}(2), as the characteristic polynomial of Tk1(𝐆)(2)T_{k_{1}}^{(\mathbf{G})}(2) is irreducible by assumption, then all of its roots are distinct. Thus if FcGF\neq c\cdot G for any non-zero constant cc, then λF(2)λG(2)\lambda_{F}(2)\neq\lambda_{G}(2). On the other hand, if k1k2k_{1}\neq k_{2}, then it follows from straightforward computations by using [22, Theorem 3.1] that mk1(𝐆)>mk2(𝐆)m_{k_{1}}^{\mathbf{(G)}}>m_{k_{2}}^{\mathbf{(G)}} for any k1>k2k_{1}>k_{2} and k1,k240k_{1},k_{2}\geq 40. Finally, if k1k2k_{1}\neq k_{2} and mk1(𝐆)=mk2(𝐆)m_{k_{1}}^{\mathbf{(G)}}=m^{\mathbf{(G)}}_{k_{2}}, then for we can just use [3] to see that TrTk1(𝐆)(2)TrTk2(𝐆)(2)\operatorname{Tr}T_{k_{1}}^{(\mathbf{G})}(2)\neq\operatorname{Tr}T_{k_{2}}^{(\mathbf{G})}(2) for all small even weights k1,k240k_{1},k_{2}\leq 40. Hence the assertion follows.

Case III: If one of FF and GG is a Saito-Kurokawa lifting and the other one is non-lifting, say F𝒮k1(𝐏)(Γ2)F\in\mathcal{S}_{k_{1}}^{\mathbf{(P)}}(\Gamma_{2}) and G𝒮k2(𝐆)(Γ2)G\in\mathcal{S}_{k_{2}}^{\mathbf{(G)}}(\Gamma_{2}). It follows from (20) and (35) that if k1k26k_{1}-k_{2}\geq 6, then we must have λF(2)>λG(2)\lambda_{F}(2)>\lambda_{G}(2). Next, we only need to consider k1k24k_{1}-k_{2}\leq 4 cases. Again, by [22, Theorem 3.1] we can easily to see that mk1(𝐏)mk2(𝐆)m_{k_{1}}^{\mathbf{(P)}}\neq m_{k_{2}}^{\mathbf{(G)}} unless k2S{20,22,24,26,28,30,32}k_{2}\in S\coloneqq\{20,22,24,26,28,30,32\}. By [4], we know that the Hecke eigenvalues λF(n)>0\lambda_{F}(n)>0 for all nn. Then by irreducibility of characteristic polynomials TrTk1(𝐏)(2)=TrλF(2)>0\operatorname{Tr}T_{k_{1}}^{\mathbf{(P)}}(2)=\operatorname{Tr}\lambda_{F}(2)>0. On the other hand, for every k2Sk_{2}\in S, by [3] we can see that TrTk2(𝐆)(2)<0\operatorname{Tr}T_{k_{2}}^{\mathbf{(G)}}(2)<0 and so TrλG(2)=TrTk2(𝐆)(2)TrTk1(𝐏)(2)\operatorname{Tr}\lambda_{G}(2)=\operatorname{Tr}T_{k_{2}}^{\mathbf{(G)}}(2)\neq\operatorname{Tr}T_{k_{1}}^{\mathbf{(P)}}(2). In particular, we have λF(2)λG(2)\lambda_{F}(2)\neq\lambda_{G}(2). Hence the assertion follows. ∎

Since Saito-Kurokawa liftings only happen for even weights, there is no need to discuss the odd weights situation for Case I and Case III in the proof above. However, we still can consider the Case II, i.e., both of FF and GG are non-liftings with k1k_{1} and k2k_{2} being odd integers. In particular, with a similar argument, we can show the following result.

Corollary 4.1.

Let k1,k2𝒦(𝐆)(2)k_{1},k_{2}\in\mathcal{K}^{\mathbf{(G)}}(2) be two odd positive integers, where k1k_{1} and k2k_{2} may equal. Let F𝒮k1(Γ2)F\in\mathcal{S}_{k_{1}}(\Gamma_{2}) and G𝒮k2(Γ2)G\in\mathcal{S}_{k_{2}}(\Gamma_{2}) be Hecke eigenforms. If λF(2)=λG(2)\lambda_{F}(2)=\lambda_{G}(2), then F=cGF=c\cdot G for some non-zero constant cc.

Remark 3.

Our approach cannot apply for the case that k1k_{1} and k2k_{2} have the different parity. It would be interesting to work out a general result of Theorem 1.2 without any restriction of weights.

5 Distinguishing Hecke eigenforms by using LL-functions

In this section, our main goal is to prove Proposition 1.3 and Theorem 1.4 in Section 1.

Proof of Proposition 1.3.

The proof is essentially based on [16, Theorem B]. In fact, it suffices to assume that d<0d<0. For a Saito-Kurokawa lifting FfF_{f}, by (5) we have

L(1/2,πFf×χd,ρ4σ1)=L(0,χd)L(1,χd)L(1/2,πf×χd),L(1/2,\pi_{F_{f}}\times\chi_{d},\rho_{4}\otimes\sigma_{1})=L(0,\chi_{d})L(1,\chi_{d})L(1/2,\pi_{f}\times\chi_{d}), (37)

where σ1\sigma_{1} is the standard representation of the dual group ×{\mathbb{C}}^{\times}. By the well-known result of Dirichlet, we have L(1,χd)0L(1,\chi_{d})\neq 0. Then by the functional equation of L(s,χd)L(s,\chi_{d}), we can see that L(0,χd)0L(0,\chi_{d})\neq 0. Moreover, we have ξd(1)=1\xi_{d}(-1)=-1 since d<0d<0. This gives

L(1/2,πf×χd)=cL(1/2,πg×χd)L(1/2,\pi_{f}\times\chi_{d})=c\cdot L(1/2,\pi_{g}\times\chi_{d}) (38)

for almost all quadratic Hecke characters χd\chi_{d} of ×\𝔸×{\mathbb{Q}}^{\times}\backslash{\mathbb{A}}^{\times}, which are corresponding to primitive Dirichlet quadratic characters ξd\xi_{d} of conductor d<0d<0. Recall that ff is of weight 2k122k_{1}-2 and gg is of weight 2k222k_{2}-2 with k1,k2k_{1},k_{2} even, then the root numbers of the cuspidal automorphic representation associated to ff and gg are 1-1. Similar to [16, Theorem B], we also have the set

𝒟ω={d:ωd>0, and dv2(mod4M) for some v coprime to 4M and M is an integer},\mathcal{D}^{\omega}=\{d\colon\mbox{$\omega d>0$, and $d\equiv v^{2}\pmod{4M}$ for some $v$ coprime to $4M$ and $M$ is an integer}\}, (39)

where ω\omega is the root number. This is exactly our case since we assume that d<0d<0 and ω=1\omega=-1. In this case, for any d𝒟ωd\in\mathcal{D}^{\omega}, we can find a non-zero constant cc such that

L(1/2,πf×χd)=cL(1/2,πg×χd),L(1/2,\pi_{f}\times\chi_{d})=c\cdot L(1/2,\pi_{g}\times\chi_{d}), (40)

By the virtual of [16, Theorem B], we have k1=k2k_{1}=k_{2} and f=gf=g. Therefore, Ff=FgF_{f}=F_{g} as desired. ∎

Proof of Theorem 1.4.

We would consider the following integral

12πi(2)(xs12x12ss12)2(ZZ(s))𝑑s,\frac{1}{2\pi i}\int_{(2)}\left(\frac{x^{s-\frac{1}{2}}-x^{\frac{1}{2}-s}}{s-\frac{1}{2}}\right)^{2}\left(-\frac{Z^{\prime}}{Z}(s)\right)\,ds, (41)

where later we will choose Z(s)Z(s) to be L(s,πF×πF,ρ4ρ4)L(s,\pi_{F}\times\pi_{F},\rho_{4}\otimes\rho_{4}) and L(s,πF×πG,ρ4ρ4)L(s,\pi_{F}\times\pi_{G},\rho_{4}\otimes\rho_{4}). Assume that

LL(s,πF×πF,ρ4ρ4)=n=1ΛF×F(n)nsandLL(s,πF×πG,ρ4ρ4)=n=1ΛF×G(n)ns.-\frac{L^{\prime}}{L}(s,\pi_{F}\times\pi_{F},\rho_{4}\otimes\rho_{4})=\sum_{n=1}^{\infty}\frac{\Lambda_{F\times F}(n)}{n^{s}}\quad\text{and}\quad-\frac{L^{\prime}}{L}(s,\pi_{F}\times\pi_{G},\rho_{4}\otimes\rho_{4})=\sum_{n=1}^{\infty}\frac{\Lambda_{F\times G}(n)}{n^{s}}. (42)

Following the idea of [5], for x>0x>0 we can show that

2n<x2ΛF×F(n)n12log(x2n)=8(x2+x1)4γsin2(γlogx)γ2+J1,2\sum_{n<x^{2}}\frac{\Lambda_{F\times F}(n)}{n^{\frac{1}{2}}}\log\left(\frac{x^{2}}{n}\right)=8(x-2+x^{-1})-4\sum_{\gamma}\frac{\sin^{2}(\gamma\log x)}{\gamma^{2}}+J_{1}, (43)

where 12+iγ\frac{1}{2}+i\gamma runs over the non-trivial zeros of L(s,πF×πF,ρ4ρ4)L(s,\pi_{F}\times\pi_{F},\rho_{4}\otimes\rho_{4}) and

J1=12πi(1/2)(G1G1(s)+G1G1(1s))(xs12x12ss12)2𝑑s.J_{1}=\frac{1}{2\pi i}\int_{(1/2)}\left(\frac{G_{1}^{\prime}}{G_{1}}(s)+\frac{G_{1}^{\prime}}{G_{1}}(1-s)\right)\left(\frac{x^{s-\frac{1}{2}}-x^{\frac{1}{2}-s}}{s-\frac{1}{2}}\right)^{2}\,ds. (44)

Here, G1(s)G_{1}(s) is the archimedean part of L(s,πF×πF,ρ4ρ4)L(s,\pi_{F}\times\pi_{F},\rho_{4}\otimes\rho_{4}); see Proposition A.2 with k1=k2k_{1}=k_{2}. Note that we moved the integration line to Re(s)=1/2\operatorname{Re}(s)=1/2 since there exist no poles of G1(s)G_{1}(s) when 1/4Re(s)3/41/4\leq\operatorname{Re}(s)\leq 3/4 by Proposition A.2 in Appendix Appendix A: Archimedean factors associated to certain LL-functions.

Similarly, in the case of L(s,πF×πG,ρ4ρ4)L(s,\pi_{F}\times\pi_{G},\rho_{4}\otimes\rho_{4}), we have

2n<x2ΛF×G(n)n12log(x2n)=4γsin2(γlogx)(γ)2+J2,2\sum_{n<x^{2}}\frac{\Lambda_{F\times G}(n)}{n^{\frac{1}{2}}}\log\left(\frac{x^{2}}{n}\right)=-4\sum_{\gamma^{\prime}}\frac{\sin^{2}(\gamma^{\prime}\log x)}{(\gamma^{\prime})^{2}}+J_{2}, (45)

where 12+iγ\frac{1}{2}+i\gamma^{\prime} runs over the non-trivial zeros of L(s,πF×πG,ρ4ρ4)L(s,\pi_{F}\times\pi_{G},\rho_{4}\otimes\rho_{4}) and

J2=12πi(1/2)(G2G2(s)+G2G2(1s))(xs12x12ss12)2𝑑s.J_{2}=\frac{1}{2\pi i}\int_{(1/2)}\left(\frac{G_{2}^{\prime}}{G_{2}}(s)+\frac{G_{2}^{\prime}}{G_{2}}(1-s)\right)\left(\frac{x^{s-\frac{1}{2}}-x^{\frac{1}{2}-s}}{s-\frac{1}{2}}\right)^{2}\,ds. (46)

Here, G2(s)G_{2}(s) is the archimedean part of L(s,πF×πG,ρ4ρ4)L(s,\pi_{F}\times\pi_{G},\rho_{4}\otimes\rho_{4}). By [11, Proposition 5.7], we can show that

γsin2(γlogx)γ2,γsin2(γlogx)(γ)2log(k1k2)(logx)2.\sum_{\gamma}\frac{\sin^{2}(\gamma\log x)}{\gamma^{2}},\hskip 11.38109pt\sum_{\gamma^{\prime}}\frac{\sin^{2}(\gamma^{\prime}\log x)}{(\gamma^{\prime})^{2}}\ll\log(k_{1}k_{2})(\log x)^{2}. (47)

By Stirling’s formula, we can show that

J1,J2O(log(k1k2)(logx)2).J_{1},J_{2}\ll O(\log(k_{1}k_{2})(\log x)^{2}). (48)

Suppose that ΛF×F(n)=ΛF×G(n)\Lambda_{F\times F}(n)=\Lambda_{F\times G}(n) for all n<x2n<x^{2}. Subtracting (45) from (43) implies

0=8(x2+x1)+O((logk1k2)(logx)2).0=8(x-2+x^{-1})+O((\log k_{1}k_{2})(\log x)^{2}). (49)

This will give a contradiction when x(logk1k2)(loglogk1k2)2x\gg(\log k_{1}k_{2})(\log\log k_{1}k_{2})^{2}. That is, if FF is not a multiple of GG, then we can find a sufficiently large CC such that, for some integer nC(logk1k2)2(loglogk1k2)4n\leq C(\log k_{1}k_{2})^{2}(\log\log k_{1}k_{2})^{4}, ΛF×F(n)ΛF×G(n)\Lambda_{F\times F}(n)\neq\Lambda_{F\times G}(n). Then Theorem 1.4 can be deduced by Lemma 5.1 below. ∎

Lemma 5.1.

Assume the notations above. Suppose that we can find AA such that ΛF×F(n)ΛF×G(n)\Lambda_{F\times F}(n)\neq\Lambda_{F\times G}(n) for some nAn\leq A. Then we can find nAn\leq A such that aF(n)aG(n)a_{F}(n)\neq a_{G}(n). Moreover, for such nAn\leq A we have λ~F(n)λ~G(n)\tilde{\lambda}_{F}(n)\neq\tilde{\lambda}_{G}(n).

Proof.

As for the first assertion, notice that ΛF×F(n)\Lambda_{F\times F}(n) and ΛF×G(n)\Lambda_{F\times G}(n) are arithmetic functions supported on prime powers. So there exist a prime number pp and a positive integer rr such that prAp^{r}\leq A, and ΛF×F(pr)ΛF×G(pr)\Lambda_{F\times F}(p^{r})\neq\Lambda_{F\times G}(p^{r}). We consider two cases: when pA1/2+1p\geq\lfloor A^{1/2}\rfloor+1 and pA1/2p\leq\lfloor A^{1/2}\rfloor. Here, x\lfloor x\rfloor denotes the greatest integer less than or equal to xx.

In the first case, p2>Ap^{2}>A and hence ΛF×F(p)ΛF×G(p)\Lambda_{F\times F}(p)\neq\Lambda_{F\times G}(p). It can be shown that ΛF×F(p)=aF(p)2logp\Lambda_{F\times F}(p)=a_{F}(p)^{2}\log p and ΛF×G(p)=aF(p)aG(p)logp\Lambda_{F\times G}(p)=a_{F}(p)a_{G}(p)\log p. Therefore, we have aF(p)aG(p)a_{F}(p)\neq a_{G}(p).

In the second case, we prove by contradiction. Suppose that aF(n)=aG(n)a_{F}(n)=a_{G}(n) for nAn\leq A. Then for pA1/2p\leq\lfloor A^{1/2}\rfloor, we have aF(pi)=aG(pi)a_{F}(p^{i})=a_{G}(p^{i}) for i=1,2i=1,2. This implies that FF and GG have the same Satake parameters at pp (up to permutation), which can be obtained by (23) and (20)-(21). This shows that ΛF×F(pr)=ΛF×G(pr)\Lambda_{F\times F}(p^{r})=\Lambda_{F\times G}(p^{r}) for any rr, which is a contradiction.

As for the second assertion, it follows from the relation between aF(n)a_{F}(n) and λ~F(n)\tilde{\lambda}_{F}(n); see (23). ∎

Remark 4.

Suppose that

L(s,πF,ρ5)=n=1bF(n)nsandL(s,πG,ρ5)=n=1bG(n)ns.L(s,\pi_{F},\rho_{5})=\sum_{n=1}^{\infty}\frac{b_{F}(n)}{n^{s}}\quad\text{and}\quad L(s,\pi_{G},\rho_{5})=\sum_{n=1}^{\infty}\frac{b_{G}(n)}{n^{s}}.

Assume that L(s,πF×πG,ρ5ρ5)L(s,\pi_{F}\times\pi_{G},\rho_{5}\otimes\rho_{5}) and L(s,πF×πF,ρ5ρ5)L(s,\pi_{F}\times\pi_{F},\rho_{5}\otimes\rho_{5}) satisfies the Generalized Riemann Hypothesis. A similar argument will show that if FF is not a scalar multiplication of GG, then there exists an integer

n(logk1k2)2(loglogk1k2)4n\ll(\log k_{1}k_{2})^{2}(\log\log k_{1}k_{2})^{4}

such that bF(n)bG(n)b_{F}(n)\neq b_{G}(n). Indeed, a direct calculation will show that, {bF(pr)}r=1\{b_{F}(p^{r})\}_{r=1}^{\infty} will determined by {bF(p),bF(p2)}\{b_{F}(p),b_{F}(p^{2})\} and hence we can obtain a result similar to the first assertion as in Lemma 5.1.

Appendix A: Archimedean factors associated to certain LL-functions

In this section, we briefly discuss the calculation of the archimedean factors associated to the LL-functions shown up in the proof of Theorem 1.4 as in Section 5.

Let F𝒮k1(Γ2)F\in\mathcal{S}_{k_{1}}(\Gamma_{2}) and G𝒮k2(Γ2)G\in\mathcal{S}_{k_{2}}(\Gamma_{2}) be Hecke eigenforms. Then we can associate the cuspidal automorphic representations πF\pi_{F} (resp. πG\pi_{G}) for FF (resp. GG) of GSp(4,𝔸){\rm GSp}(4,{\mathbb{A}}). For πF\pi_{F}, we can associate the completed spinor LL-function and the completed standard LL-function, denoted by Λ(s,πF,ρ4)\Lambda(s,\pi_{F},\rho_{4}) and Λ(s,πF,ρ5)\Lambda(s,\pi_{F},\rho_{5}), respectively. Moreover, via the Langlands transfer (see [20, § 5.1]), we can find Π4F\Pi_{4}^{F} (resp. Π5F\Pi_{5}^{F}), which is an cuspidal automorphic representation of GL(4,𝔸){\rm GL}(4,{\mathbb{A}}) (resp. GL(5,𝔸){\rm GL}(5,{\mathbb{A}})) such that

Λ(s,πF,ρ4)=Λ(s,Π4F)andΛ(s,πF,ρ5)=Λ(s,Π5F).\Lambda(s,\pi_{F},\rho_{4})=\Lambda(s,\Pi_{4}^{F})\quad\text{and}\quad\Lambda(s,\pi_{F},\rho_{5})=\Lambda(s,\Pi_{5}^{F}).

In this case, the Rankin-Selberg LL-function Λ(s,πF×πG,ρ4ρ4)\Lambda(s,\pi_{F}\times\pi_{G},\rho_{4}\otimes\rho_{4}) and Λ(s,πF×πG,ρ5ρ5)\Lambda(s,\pi_{F}\times\pi_{G},\rho_{5}\otimes\rho_{5}) is defined by the Rankin-Selberg convolutions on GL(4)×GL(4){\rm GL}(4)\times{\rm GL}(4) and GL(5)×GL(5){\rm GL}(5)\times{\rm GL}(5), respectively, i.e.,

Λ(s,πF×πG,ρ4ρ4)=Λ(s,Π4F×Π4G)andΛ(s,πF×πG,ρ5ρ5)=Λ(s,Π5F×Π5G).\Lambda(s,\pi_{F}\times\pi_{G},\rho_{4}\otimes\rho_{4})=\Lambda(s,\Pi_{4}^{F}\times\Pi_{4}^{G})\quad\text{and}\quad\Lambda(s,\pi_{F}\times\pi_{G},\rho_{5}\otimes\rho_{5})=\Lambda(s,\Pi_{5}^{F}\times\Pi_{5}^{G}). (50)

To calculate the associated archimedean factors, we recall some basic facts regarding the real Weil group W=×j×W_{\mathbb{R}}=\mathbb{C}^{\times}\sqcup j\mathbb{C}^{\times}. Here, the multiplication on ×{\mathbb{C}}^{\times} is standard, and jj is an element satisfying j2=1j^{2}=-1 and jzj1=z¯jzj^{-1}=\bar{z} (complex conjugation) for z×z\in{\mathbb{C}}^{\times}. More precisely, we are considering representations of WW_{\mathbb{R}}, which are continuous homomorphisms WGL(n,)W_{\mathbb{R}}\to{\rm GL}(n,{\mathbb{C}}) for some nn with the image consisting of semisimple elements. By [14], every finite-dimensional semisimple representation of WW_{\mathbb{R}} is completely recucible, and each irreducible representation is either one- or two-dimensional. The complete list of one-dimensional representations is as follows:

φ+,t\displaystyle\varphi_{+,t} :reiθr2t,j1,\displaystyle\colon re^{i\theta}\longmapsto r^{2t},\quad j\mapsto 1, (51)
φ,t\displaystyle\varphi_{-,t} :reiθr2t,j1,\displaystyle\colon re^{i\theta}\longmapsto r^{2t},\quad j\mapsto-1, (52)

where tt\in{\mathbb{C}}, and we write any non-zero complex number zz as reiθre^{i\theta} with r>0r\in{\mathbb{R}}_{>0} and θ/2π\theta\in{\mathbb{R}}/2\pi{\mathbb{Z}}. The two-dimensional representations are precisely

φ,t:reiθ[r2teiθr2teiθ],j[(1)1],\varphi_{\ell,t}\colon re^{i\theta}\mapsto\left[\begin{smallmatrix}r^{2t}e^{i\ell\theta}\\ &r^{2t}e^{-i\ell\theta}\end{smallmatrix}\right],\qquad j\mapsto\left[\begin{smallmatrix}&(-1)^{\ell}\\ 1\end{smallmatrix}\right], (53)

where >0\ell\in{\mathbb{Z}}_{>0} and tt\in{\mathbb{C}}. And the corresponding LL-factors, i.e., the archimedean factors, are given as follows:

L(s,φ)={Γ(s+t)if φ=φ+,t,Γ(s+t+1)if φ=φ,t,Γ(s+t+2)if φ=φ,t.L_{\infty}(s,\varphi)=\begin{cases}\Gamma_{\mathbb{R}}(s+t)&\mbox{if $\varphi=\varphi_{+,t}$},\\ \Gamma_{\mathbb{R}}(s+t+1)&\mbox{if $\varphi=\varphi_{-,t}$},\\ \Gamma_{\mathbb{C}}(s+t+\frac{\ell}{2})&\mbox{if $\varphi=\varphi_{\ell,t}$}.\end{cases} (54)

Here,

Γ(s)πs/2Γ(s2),Γ(s)2(2π)sΓ(s),\Gamma_{\mathbb{R}}(s)\coloneqq\pi^{-s/2}\Gamma\left(\frac{s}{2}\right),\qquad\Gamma_{\mathbb{C}}(s)\coloneqq 2(2\pi)^{-s}\Gamma(s), (55)

where Γ(s)\Gamma(s) is the usual gamma function. By a direct calculations we have the following lemma:

Lemma A.1.

For ,1,2>0\ell,\ell_{1},\ell_{2}\in{\mathbb{Z}}_{>0} and t1,t2t_{1},t_{2}\in{\mathbb{C}}, we have

φ+,t1φ+,t2=φ,t1φ,t2=\displaystyle\varphi_{+,t_{1}}\otimes\varphi_{+,t_{2}}=\varphi_{-,t_{1}}\otimes\varphi_{-,t_{2}}= φ+,t1+t2\displaystyle\varphi_{+,t_{1}+t_{2}} (56)
φ+,t1φ,t2=φ,t1φ+,t2=\displaystyle\varphi_{+,t_{1}}\otimes\varphi_{-,t_{2}}=\varphi_{-,t_{1}}\otimes\varphi_{+,t_{2}}= φ,t1+t2\displaystyle\varphi_{-,t_{1}+t_{2}} (57)
φ±,t1φ,t2=\displaystyle\varphi_{\pm,t_{1}}\otimes\varphi_{\ell,t_{2}}= φ,t1+t2\displaystyle\varphi_{\ell,t_{1}+t_{2}} (58)
φ1,t1φ2,t2=\displaystyle\varphi_{\ell_{1},t_{1}}\otimes\varphi_{\ell_{2},t_{2}}= {φ1+2,t1+t2φ|12|,t1+t2if 12,φ1+2,t1+t2φ+,t1+t2φ,t1+t2if 1=2.\displaystyle\begin{cases}\varphi_{\ell_{1}+\ell_{2},t_{1}+t_{2}}\oplus\varphi_{|\ell_{1}-\ell_{2}|,t_{1}+t_{2}}&\text{if }\ell_{1}\neq\ell_{2},\\ \varphi_{\ell_{1}+\ell_{2},t_{1}+t_{2}}\oplus\varphi_{+,t_{1}+t_{2}}\oplus\varphi_{-,t_{1}+t_{2}}&\text{if }\ell_{1}=\ell_{2}.\end{cases} (59)
Remark 5.

Recall that Γ(s)=Γ(s)Γ(s+1)\Gamma_{\mathbb{C}}(s)=\Gamma_{\mathbb{R}}(s)\Gamma_{\mathbb{R}}(s+1), the second case in (59) looks precisely like the first case in (59), if we allow 1=2\ell_{1}=\ell_{2}.

For our purpose, we will only consider the case t=0t=0; in this case, we write φ±\varphi_{\pm} instead of φ±,0\varphi_{\pm,0} and φ\varphi_{\ell} instead of φ,0\varphi_{\ell,0}. It follows from [23, § 3.2] (observing that λ1=k11\lambda_{1}=k_{1}-1 and λ2=k12\lambda_{2}=k_{1}-2) and [20, Theorem 5.1.2] that the LL-parameter of Π4F\Pi_{4}^{F} at the archimedean place is given by:

φ2k13φ1.\varphi_{2k_{1}-3}\oplus\varphi_{1}. (60)

Then we have the following proposition proved by Gun, Kohnen and Paul in [7, pp. 56-57]. In their proof, they considered two separated cases (k1>k2k_{1}>k_{2} and k1=k2k_{1}=k_{2}). Observing Remark 5, these two cases can be combined as follows:

Proposition A.2.

Assume the notations above. The archimedean factor of Λ(s,πF×πG,ρ4ρ4)\Lambda(s,\pi_{F}\times\pi_{G},\rho_{4}\otimes\rho_{4}) is given by:

Γ(s+k1+k23)Γ(s+k11)Γ(s+k21)Γ(s+k12)Γ(s+k22)Γ(s+1)Γ(s+|k1k2|)Γ(s)Γ(s+1).\begin{split}&\Gamma_{\mathbb{C}}\left(s+k_{1}+k_{2}-3\right)\Gamma_{\mathbb{C}}\left(s+k_{1}-1\right)\Gamma_{\mathbb{C}}(s+k_{2}-1)\Gamma_{\mathbb{C}}\left(s+k_{1}-2\right)\Gamma_{\mathbb{C}}\left(s+k_{2}-2\right)\\ &\Gamma_{\mathbb{C}}(s+1)\Gamma_{\mathbb{C}}(s+|k_{1}-k_{2}|)\Gamma_{\mathbb{R}}(s)\Gamma_{\mathbb{R}}(s+1).\end{split}

Next, we consider the archimedean factor of the Rankin-Selberg LL-function Λ(s,πF×πG,ρ5ρ5)\Lambda(s,\pi_{F}\times\pi_{G},\rho_{5}\otimes\rho_{5}). It follows from the construction of the standard LL-function that the LL-parameter of Π5F\Pi_{5}^{F} is

φ2k12φ2k14φ+.\varphi_{2k_{1}-2}\oplus\varphi_{2k_{1}-4}\oplus\varphi_{+}. (61)

Here, we require that k12k_{1}\geq 2; see [23, Table 5]. Then by (61) and Lemma A.1 we have

Proposition A.3.

Assume the notations above. The archimedean factor of Λ(s,πF×πG,ρ5ρ5)\Lambda(s,\pi_{F}\times\pi_{G},\rho_{5}\otimes\rho_{5}) is given by:

Γ(s+k1+k22)Γ(s+|k1k2|)2Γ(s+k1+k23)2Γ(s+|k2k11|)Γ(s+k11)Γ(s+|k1k21|)Γ(s+k1+k24)Γ(s+k12)Γ(s+k21)Γ(s+k22)Γ(s).\begin{split}&\Gamma_{\mathbb{C}}\left(s+k_{1}+k_{2}-2\right)\Gamma_{\mathbb{C}}\left(s+|k_{1}-k_{2}|\right)^{2}\Gamma_{\mathbb{C}}\left(s+k_{1}+k_{2}-3\right)^{2}\Gamma_{\mathbb{C}}\left(s+|k_{2}-k_{1}-1|\right)\Gamma_{\mathbb{C}}\left(s+k_{1}-1\right)\\ &\Gamma_{\mathbb{C}}\left(s+|k_{1}-k_{2}-1|\right)\Gamma_{\mathbb{C}}\left(s+k_{1}+k_{2}-4\right)\Gamma_{\mathbb{C}}\left(s+k_{1}-2\right)\Gamma_{\mathbb{C}}\left(s+k_{2}-1\right)\Gamma_{\mathbb{C}}\left(s+k_{2}-2\right)\Gamma_{\mathbb{R}}(s).\end{split}

Again, by Remark 5 we can write Γ(s+0)\Gamma_{\mathbb{C}}\left(s+0\right) as Γ(s)Γ(s+1)\Gamma_{\mathbb{R}}\left(s\right)\Gamma_{\mathbb{R}}\left(s+1\right) if happens.

References

  • And [74] A. N. Andrianov. Euler products that correspond to Siegel’s modular forms of genus 22. Uspehi Mat. Nauk, 29(3 (177)):43–110, 1974.
  • AS [01] Mahdi Asgari and Ralf Schmidt. Siegel modular forms and representations. Manuscripta Math., 104(2):173–200, 2001.
  • BCFvdG [17] Jonas Bergstro¨\ddot{\rm{o}}m, Fabien Cle´\acute{\rm{e}}ry, Carel Faber, and Gerard van der Geer. Siegel modular forms of degree two and three. 2017. Retrieved [May 2022].
  • Bre [99] Stefan Breulmann. On Hecke eigenforms in the Maaßspace. Math. Z., 232(3):527–530, 1999.
  • GH [93] Dorian Goldfeld and Jeffrey Hoffstein. On the number of Fourier coefficients that determine a modular form. In A tribute to Emil Grosswald: number theory and related analysis, volume 143 of Contemp. Math., pages 385–393. Amer. Math. Soc., Providence, RI, 1993.
  • Ghi [11] Alexandru Ghitza. Distinguishing Hecke eigenforms. Int. J. Number Theory, 7(5):1247–1253, 2011.
  • GKP [21] Sanoli Gun, Winfried Kohnen, and Biplab Paul. Arithmetic behaviour of Hecke eigenvalues of Siegel cusp forms of degree two. Ramanujan J., 54(1):43–62, 2021.
  • GM [12] Alexandru Ghitza and Angus McAndrew. Experimental evidence for Maeda’s conjecture on modular forms. Tbil. Math. J., 5(2):55–69, 2012.
  • GS [14] Alexandru Ghitza and Robert Sayer. Hecke eigenvalues of Siegel modular forms of “different weights”. J. Number Theory, 143:125–141, 2014.
  • HM [97] Haruzo Hida and Yoshitaka Maeda. Non-abelian base change for totally real fields. Number Special Issue, pages 189–217. 1997. Olga Taussky-Todd: in memoriam.
  • IK [04] Henryk Iwaniec and Emmanuel Kowalski. Analytic number theory, volume 53 of American Mathematical Society Colloquium Publications. American Mathematical Society, Providence, RI, 2004.
  • JS [81] H. Jacquet and J. A. Shalika. On Euler products and the classification of automorphic representations. I. Amer. J. Math., 103(3):499–558, 1981.
  • KMS [21] Arvind Kumar, Jaban Meher, and Karam Deo Shankhadhar. Strong multiplicity one for Siegel cusp forms of degree two. Forum Math., 33(5):1157–1167, 2021.
  • Kna [94] A. W. Knapp. Local Langlands correspondence: the Archimedean case. In Motives (Seattle, WA, 1991), volume 55 of Proc. Sympos. Pure Math., pages 393–410. Amer. Math. Soc., Providence, RI, 1994.
  • Kur [78] Nobushige Kurokawa. Examples of eigenvalues of Hecke operators on Siegel cusp forms of degree two. Invent. Math., 49(2):149–165, 1978.
  • LR [97] Wenzhi Luo and Dinakar Ramakrishnan. Determination of modular forms by twists of critical LL-values. Invent. Math., 130(2):371–398, 1997.
  • Maa [79] Hans Maass. Über eine Spezialschar von Modulformen zweiten Grades. I, II, III. Invent. Math., 53(3):95–104, 249–253, 255–265, 1979.
  • Mur [97] M. Ram Murty. Congruences between modular forms. In Analytic number theory (Kyoto, 1996), volume 247 of London Math. Soc. Lecture Note Ser., pages 309–320. Cambridge Univ. Press, Cambridge, 1997.
  • PS [09] Ameya Pitale and Ralf Schmidt. Ramanujan-type results for Siegel cusp forms of degree 2. J. Ramanujan Math. Soc., 24(1):87–111, 2009.
  • PSS [14] Ameya Pitale, Abhishek Saha, and Ralf Schmidt. Transfer of Siegel cusp forms of degree 2. Mem. Amer. Math. Soc., 232(1090):vi+107, 2014.
  • RS [07] Brooks Roberts and Ralf Schmidt. Local newforms for GSp(4), volume 1918 of Lecture Notes in Mathematics. Springer, Berlin, 2007.
  • RSY [21] Manami Roy, Ralf Schmidt, and Shaoyun Yi. On counting cuspidal automorphic representations for GSp(4)\rm GSp(4). Forum Math., 33(3):821–843, 2021.
  • Sch [17] Ralf Schmidt. Archimedean aspects of Siegel modular forms of degree 2. Rocky Mountain J. Math., 47(7):2381–2422, 2017.
  • Sch [18] Ralf Schmidt. Packet structure and paramodular forms. Trans. Amer. Math. Soc., 370(5):3085–3112, 2018.
  • Stu [87] Jacob Sturm. On the congruence of modular forms. In Number theory (New York, 1984–1985), volume 1240 of Lecture Notes in Math., pages 275–280. Springer, Berlin, 1987.
  • VX [18] Trevor Vilardi and Hui Xue. Distinguishing eigenforms of level one. Int. J. Number Theory, 14(1):31–36, 2018.
  • WWYY [23] Xiyuan Wang, Zhining Wei, Pan Yan, and Shaoyun Yi. Some remarks on strong multiplicity one for paramodular forms. arXiv preprint arXiv:2310.17144, 2023.
  • XZ [22] Hui Xue and Daozhou Zhu. Uniqueness of Fourier coefficients of eigenforms. Ramanujan J., 57(2):487–505, 2022.

Department of Mathematics, Brown University, Providence, RI 02912, USA.

E-mail address: [email protected]

School of Mathematical Sciences, Xiamen University, Xiamen, Fujian 361005, China.

E-mail address: [email protected]