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

On the classification of (𝔤,K)(\mathfrak{g},K)-modules generated by nearly holomorphic Hilbert-Siegel modular forms and projection operators

Shuji Horinaga
Abstract.

We classify the (𝔤,K)(\mathfrak{g},K)-modules generated by nearly holomorphic Hilbert-Siegel modular forms by the global method. As an application, we study the image of projection operators on the space of nearly holomorphic Hilbert-Siegel modular forms with respect to infinitesimal characters in terms of (𝔤,K)(\mathfrak{g},K)-modules.

1. Introduction

1.1. Algebraicity of special LL values

The arithmeticity of special LL values is a central problem in modern number theory. In the motivic setting, Deligne [Del79] conjectured the algebraicity of critical LL values up to the period. For the critical values attached to scalar valued Hilbert-Siegel modular forms and Hermitian modular forms, Shimura proved the arithmeticity of them up to suitable periods in [Shi00] by using of nearly holomorphic modular forms. The period can be expressed by Petersson inner product times some power of π\pi. Recently, in [HPSS21], Pitale, Saha, Schmidt and the author prove the arithmeticity of them attached to vector valued Siegel modular forms under the parity condition of weights. The purpose of this paper is to prepare to remove the parity condition by investigating the (𝔤,K)(\mathfrak{g},K)-modules generated by nearly holomorphic Hilbert-Siegel modular forms.

1.2. (𝔤,K)(\mathfrak{g},K)-modules generated by nearly holomorphic Siegel modular forms

Let FF be a totally real field with degree dd and 𝐚\mathbf{a} the set of embeddings of FF into \mathbb{R}. Put Gn=ResF/Sp2nG_{n}=\mathop{\mathrm{Res}}_{F/\mathbb{Q}}\mathrm{Sp}_{2n}. Here Res\mathop{\mathrm{Res}} is the Weil restriction and Sp2n\mathrm{Sp}_{2n} is the symplectic group of rank nn. Let n\mathfrak{H}_{n} be the Siegel upper half space of degree nn. Put 𝔤n=Lie(Gn())\mathfrak{g}_{n}=\mathrm{Lie}(G_{n}(\mathbb{R}))\otimes_{\mathbb{R}}\mathbb{C}. We denote by Kn,K_{n,\infty} and 𝒵n\mathcal{Z}_{n} the stabilizer of 𝐢=(1 1n,,1 1n)nd\mathbf{i}=(\sqrt{-1}\,\mathbf{1}_{n},\ldots,\sqrt{-1}\,\mathbf{1}_{n})\in\mathfrak{H}_{n}^{d} and the center of the universal enveloping algebra 𝒰(𝔤n)\mathcal{U}(\mathfrak{g}_{n}), respectively. Let Kn,K_{n,\mathbb{C}} be the complexification of Kn,K_{n,\infty}. Set 𝔨n=Lie(Kn,)\mathfrak{k}_{n}=\mathrm{Lie}(K_{n,\infty})\otimes_{\mathbb{R}}\mathbb{C}. We then have the well-known decomposition:

𝔤n=𝔨n𝔭n,+𝔭n,.\mathfrak{g}_{n}=\mathfrak{k}_{n}\oplus\mathfrak{p}_{n,+}\oplus\mathfrak{p}_{n,-}.

Here 𝔭n,+\mathfrak{p}_{n,+} (resp. 𝔭n,\mathfrak{p}_{n,-}) is the Lie subalgebra of 𝔤n\mathfrak{g}_{n} corresponding to the holomorphic tangent space (resp. anti-holomorphic tangent space) of nd\mathfrak{H}_{n}^{d} at 𝐢\mathbf{i}. We take a Cartan subalgebra of 𝔨n\mathfrak{k}_{n}. Then it is a Cartan subalgebra of 𝔤n\mathfrak{g}_{n}. The root system Φ\Phi of 𝔰𝔭2n()\mathfrak{sp}_{2n}(\mathbb{C}) is

Φ={±(ei+ej),±(eke),1ijn,1k<n}.\Phi=\{\;\pm(e_{i}+e_{j}),\;\pm(e_{k}-e_{\ell}),\quad 1\leq i\leq j\leq n,1\leq k<\ell\leq n\;\}.

We consider the set

Φ+={(ei+ej),eke,1ijn,1k<n}\Phi^{+}=\{\;-(e_{i}+e_{j}),\;e_{k}-e_{\ell},\quad 1\leq i\leq j\leq n,1\leq k<\ell\leq n\;\}

to be a positive root system. Let ρ\rho be half the sum of positive roots. Note that 𝔤n=v𝐚𝔰𝔭2n()\mathfrak{g}_{n}=\bigoplus_{v\in\mathbf{a}}\mathfrak{sp}_{2n}(\mathbb{C}). We say that a weight λ=(λ1,v,,λn,v)v𝐚\lambda=(\lambda_{1,v},\ldots,\lambda_{n,v})_{v\in\mathbf{a}} which lies in v𝐚n\bigoplus_{v\in\mathbf{a}}\mathbb{C}^{n} is 𝔨n\mathfrak{k}_{n}-dominant if λi,vλi+1,v0\lambda_{i,v}-\lambda_{i+1,v}\in\mathbb{Z}_{\geq 0} for any 1in11\leq i\leq n-1 and v𝐚v\in\mathbf{a}. We also say that a 𝔨n\mathfrak{k}_{n}-dominant integral weight λ=(λ1,v,,λn,v)v𝐚\lambda=(\lambda_{1,v},\ldots,\lambda_{n,v})_{v\in\mathbf{a}} is anti-dominant if λnn\lambda_{n}\geq n. For any 𝔨n\mathfrak{k}_{n}-dominant integral weight λ\lambda, there exist the (parabolic) Verma module N(λ)N(\lambda) with respect to a parabolic subalgebra 𝔭n,𝔨n\mathfrak{p}_{n,-}\oplus\mathfrak{k}_{n} and a unique irreducible highest weight (𝔤n,Kn,)(\mathfrak{g}_{n},K_{n,\infty})-module L(λ)L(\lambda) of highest weight λ\lambda. Then, L(λ)L(\lambda) is the unique irreducible quotient of N(λ)N(\lambda). For a (𝔤n,Kn,)(\mathfrak{g}_{n},K_{n,\infty})-module π\pi, the symbol π\pi^{\vee} denotes the contragredient of π\pi in the sense of [Hum08].

For an automorphic form φ\varphi on Gn(𝔸)G_{n}(\mathbb{A}_{\mathbb{Q}}), we say that φ\varphi is nearly holomorphic if φ\varphi is 𝔭n,\mathfrak{p}_{n,-}-finite, i.e., 𝒰(𝔭n,)φ\mathcal{U}(\mathfrak{p}_{n,-})\cdot\varphi is finite-dimensional. The goal of this paper is to classify the indecomposable (𝔤n,Kn,)(\mathfrak{g}_{n},K_{n,\infty})-modules generated by nearly holomorphic automorphic forms.

Theorem 1.2.1 (Theorem 6.5.1).

Let π\pi be an indecomposable (𝔤n,Kn,)(\mathfrak{g}_{n},K_{n,\infty})-module generated by a nearly holomorphic automorphic form on Gn(𝔸)G_{n}(\mathbb{A}_{\mathbb{Q}}). If FF\neq\mathbb{Q}, π\pi is irreducible. If F=F=\mathbb{Q}, the length of π\pi is at most two. More precisely, if π\pi is reducible, there exists an odd integer ii and (λ1,,λni)ni(\lambda_{1},\ldots,\lambda_{n-i})\in\mathbb{Z}^{n-i} with λ1λnin(i3)/2\lambda_{1}\geq\cdots\geq\lambda_{n-i}\geq n-(i-3)/2 such that πN(λ1,,λni,n(i3)/2,,n(i3)/2)\pi\cong N(\lambda_{1},\ldots,\lambda_{n-i},n-(i-3)/2,\ldots,n-(i-3)/2)^{\vee}.

This result is a generalization of [PSS21]. The key idea of proof is the harmonic analysis of the space of nearly holomorphic automorphic forms on Gn(𝔸)G_{n}(\mathbb{A}_{\mathbb{Q}}), which is investigated in [Hor20b].

1.3. Projection operators

Fix a weight ρ\rho and a congruence subgroup Γ\Gamma. Let Nρ(Γ)N_{\rho}(\Gamma) be the space of nearly holomorphic Hilbert-Siegel modular forms of weight ρ\rho with respect to Γ\Gamma. For an infinitesimal character χ\chi of 𝒵n\mathcal{Z}_{n}, we can define the projection operator 𝔭χEnd(Nρ(Γ))\mathfrak{p}_{\chi}\in\mathrm{End}(N_{\rho}(\Gamma)) associated to χ\chi. Then, the projection operator 𝔭χ\mathfrak{p}_{\chi} commutes with the Aut()\mathrm{Aut}(\mathbb{C}) action as follows:

Theorem 1.3.1 (Theorem 7.2.2).

For any fNρ(Γ)f\in N_{\rho}(\Gamma) and σAut()\sigma\in\mathrm{Aut}(\mathbb{C}), we have

𝔭χ(fσ)=𝔭χσ(f).\mathfrak{p}_{\chi}({{}^{\sigma}f})={{}^{\sigma}\mathfrak{p}_{\chi}(f)}.

For a 𝔨n\mathfrak{k}_{n}-dominant integral weight λ\lambda and v𝐚v\in\mathbf{a}, put jv(λ)=#{jλ1,vλj,v(mod 2)}j_{v}(\lambda)=\#\{j\mid\lambda_{1,v}\equiv\lambda_{j,v}\,(\mathrm{mod}\,2)\}. Set

jv(λ)=jv(λ)stdGLn(),ρv=detλ1,v1jvstdGLn(),andρ=v𝐚ρv,\wedge^{j_{v}(\lambda)}=\wedge^{j_{v}(\lambda)}\mathrm{std}_{\mathrm{GL}_{n}(\mathbb{C})},\quad\rho_{v}=\mathrm{det}^{\lambda_{1,v}-1}\otimes\wedge^{j_{v}}\mathrm{std}_{\mathrm{GL}_{n}(\mathbb{C})},\quad\text{and}\quad\rho=\bigotimes_{v\in\mathbf{a}}\rho_{v},

where stdGLn()\mathrm{std}_{\mathrm{GL}_{n}(\mathbb{C})} is the standard representation of GLn()\mathrm{GL}_{n}(\mathbb{C}) and jv(λ)stdGLn()\wedge^{j_{v}(\lambda)}\mathrm{std}_{\mathrm{GL}_{n}(\mathbb{C})} is the jv(λ)j_{v}(\lambda)-th exterior product of stdGLn()\mathrm{std}_{\mathrm{GL}_{n}(\mathbb{C})}.

Theorem 1.3.2 (Theorem 7.2.3).

Let λ=(λ1,v,,λn,v)v\lambda=(\lambda_{1,v},\ldots,\lambda_{n,v})_{v} be a regular anti-dominant integral weight. Put ρ=v𝐚(detλ1,v1jv(λ))\rho=\bigotimes_{v\in\mathbf{a}}(\det^{\lambda_{1,v}-1}\otimes\wedge^{j_{v}(\lambda)}) and Nρ(Γ,χλ)=𝔭χλ(Nρ(Γ))N_{\rho}(\Gamma,\chi_{\lambda})=\mathfrak{p}_{\chi_{\lambda}}(N_{\rho}(\Gamma)). If F=F=\mathbb{Q} and λn,v=n+1\lambda_{n,v}=n+1, any modular form in Nρ(Γ,χλ)N_{\rho}(\Gamma,\chi_{\lambda}) generates L(λ)L(\lambda) or N(λ1,,λn1,n1)N(\lambda_{1},\ldots,\lambda_{n-1},n-1)^{\vee}. If not, any modular form in Nρ(Γ,χλ)N_{\rho}(\Gamma,\chi_{\lambda}) generates L(λ)L(\lambda).

The following is the analogue of holomorphic projection.

Corollary 1.3.3.

Let λ=(λ1,v,,λn,v)v\lambda=(\lambda_{1,v},\ldots,\lambda_{n,v})_{v} be an anti-dominant 𝔨n\mathfrak{k}_{n}-dominant integral weight and ρ\rho the irreducible highest weight representation of Kn,K_{n,\mathbb{C}} with highest weight λ\lambda. Suppose λ1,vλn,v1\lambda_{1,v}-\lambda_{n,v}\leq 1 and λn,vn+1\lambda_{n,v}\geq n+1 for any v𝐚v\in\mathbf{a}. If FF\neq\mathbb{Q} or λn,vn+1\lambda_{n,v}\neq n+1 for some v𝐚v\in\mathbf{a}, the projection 𝔭χ\mathfrak{p}_{\chi} defines a projection onto Mρ(Γ)M_{\rho}(\Gamma), the subspace of holomorphic modular forms.

We then characterize the nearly holomorphic Hilbert-Siegel modular forms which generate a holomorphic discrete series representation in terms of projections 𝔭χ\mathfrak{p}_{\chi} under a mild assumption. This gives a generalization of Shimura’s holomorphic projection.

Notation

We denote by Matm,n\mathrm{Mat}_{m,n} the set of m×nm\times n-matrices. Put Matn=Matn,n\mathrm{Mat}_{n}=\mathrm{Mat}_{n,n} with the unit 𝟏n\mathbf{1}_{n}. Let GLn\mathrm{GL}_{n} and Sp2n\mathrm{Sp}_{2n} be the algebraic groups defined by

GLn(R)={gMatndetgR×}\mathrm{GL}_{n}(R)=\{g\in\mathrm{Mat}_{n}\mid\det g\in R^{\times}\}

and

Sp2n(R)={gGL2n(R)|gt(0n𝟏n𝟏n0n)g=(0n𝟏n𝟏n0n)}\mathrm{Sp}_{2n}(R)=\left\{g\in\mathrm{GL}_{2n}(R)\,\middle|\,{{}^{t}{g}}\begin{pmatrix}0_{n}&-\mathbf{1}_{n}\\ \mathbf{1}_{n}&0_{n}\end{pmatrix}g=\begin{pmatrix}0_{n}&-\mathbf{1}_{n}\\ \mathbf{1}_{n}&0_{n}\end{pmatrix}\right\}

for a ring RR, respectively. Set Symn={gMatngt=g}\mathrm{Sym}_{n}=\{g\in\mathrm{Mat}_{n}\mid{{}^{t}g}=g\}. Let BnB_{n} be the subgroup of Sp2n\mathrm{Sp}_{2n} defined by

Bn={(a0a1t)|a is a upper triangular matrix.}.B_{n}=\left\{\begin{pmatrix}a&*\\ 0&{{}^{t}a^{-1}}\end{pmatrix}\,\middle|\,\text{$a$ is a upper triangular matrix.}\right\}.

The group BnB_{n} is a Borel subgroup of Sp2n\mathrm{Sp}_{2n} with the Levi decomposition Bn=TnNnB_{n}=T_{n}N_{n}. Here TnBnT_{n}\subset B_{n} is the maximal diagonal torus of Sp2n\mathrm{Sp}_{2n}. A parabolic subgroup PP of Sp2n\mathrm{Sp}_{2n} is called standard if PP contains BnB_{n}. Let APA_{P} be the split component of PP and APA_{P}^{\infty} the identity component of AP()A_{P}(\mathbb{R}). We denote by Pi,nP_{i,n} and Qi,nQ_{i,n} the standard parabolic groups of Sp2n\mathrm{Sp}_{2n} with the Levi subgroups GLi×Sp2(ni)\mathrm{GL}_{i}\times\mathrm{Sp}_{2(n-i)} and (GL1)i×Sp2(ni)(\mathrm{GL}_{1})^{i}\times\mathrm{Sp}_{2(n-i)}, respectively. Set Pn=Pn,nP_{n}=P_{n,n}. For a parabolic subgroup PP, let δP\delta_{P} be the modulus character of PP.

For n1n\in\mathbb{Z}_{\geq 1}, set

n={zSymn()Im(z) is positive definite}.\mathfrak{H}_{n}=\{z\in\mathrm{Sym}_{n}(\mathbb{C})\mid\text{$\mathrm{Im}(z)$ is positive definite}\}.

The space n\mathfrak{H}_{n} is called the Siegel upper half space of degree nn. The Lie group Sp2n()\mathrm{Sp}_{2n}(\mathbb{R}) acts on n\mathfrak{H}_{n} by the rule

(abcd)(z)=(az+b)(cz+d)1,(abcd)Sp2n(),zn.\begin{pmatrix}a&b\\ c&d\end{pmatrix}(z)=(az+b)(cz+d)^{-1},\qquad\begin{pmatrix}a&b\\ c&d\end{pmatrix}\in\mathrm{Sp}_{2n}(\mathbb{R}),\,z\in\mathfrak{H}_{n}.

Put

Kn,={g=(abcd)Sp2n()|a=d,c=b}.K_{n,\infty}=\left\{g=\begin{pmatrix}a&b\\ c&d\end{pmatrix}\in\mathrm{Sp}_{2n}(\mathbb{R})\,\middle|\,a=d,\,c=-b\right\}.

Then Kn,K_{n,\infty} is the group of stabilizers of 𝐢=1 1nn\mathbf{i}=\sqrt{-1}\,\mathbf{1}_{n}\in\mathfrak{H}_{n}. For simplicity the notation, the symbol 𝐢\mathbf{i} also denotes the element (1 1n,,1 1n)nd(\sqrt{-1}\,\mathbf{1}_{n},\ldots,\sqrt{-1}\,\mathbf{1}_{n})\in\mathfrak{H}_{n}^{d}. Since the action of Sp2n()\mathrm{Sp}_{2n}(\mathbb{R}) on n\mathfrak{H}_{n} is transitive, we have nSp2n()/Kn,\mathfrak{H}_{n}\cong\mathrm{Sp}_{2n}(\mathbb{R})/K_{n,\infty}.

Let FF be a totally real field with degree dd. Let 𝐚={1,,d}\mathbf{a}=\{\infty_{1},\ldots,\infty_{d}\} be the set of embeddings of FF into \mathbb{R}. We denote by 𝔸F\mathbb{A}_{F} and 𝔸F,fin\mathbb{A}_{F,\mathrm{fin}} the adele ring of FF and the finite part of 𝔸F\mathbb{A}_{F}, respectively. For a place vv, let FvF_{v} be the vv-completion of FF. Put F=v𝐚FvF_{\infty}=\prod_{v\in\mathbf{a}}F_{v}. For a non-archimedean place vv, let 𝒪Fv\mathcal{O}_{F_{v}} be the ring of integers of FvF_{v}.

Set Gn=ResF/Sp2nG_{n}=\mathop{\mathrm{Res}}_{F/\mathbb{Q}}\mathrm{Sp}_{2n} where Res\mathop{\mathrm{Res}} is the Weil restriction. We define the standard parabolic subgroups Pi,n,Qi,nP_{i,n},Q_{i,n} and BnB_{n} of GnG_{n} by the Weil restriction of parabolic subgroups Pi,n,Qi,nP_{i,n},Q_{i,n} and BnB_{n} of Sp2n\mathrm{Sp}_{2n}, respectively. Let WnW_{n} be the Weyl group of Sp2n\mathrm{Sp}_{2n}. For an archimedean place vv, set Kn,v=Kn,K_{n,v}=K_{n,\infty}. For the sake of simplicity, the symbol Kn,K_{n,\infty} denotes the maximal compact subgroup v𝐚Kn,v\prod_{v\in\mathbf{a}}K_{n,v} of Gn()G_{n}(\mathbb{R}). Let Kn,K_{n,\mathbb{C}} be the complexification of v𝐚Kn,v\prod_{v\in\mathbf{a}}K_{n,v}. Put 𝔤n=Lie(Gn())\mathfrak{g}_{n}=\mathrm{Lie}(G_{n}(\mathbb{R}))\otimes_{\mathbb{R}}\mathbb{C} and 𝔨n=Lie(v𝐚Kn,v)\mathfrak{k}_{n}=\mathrm{Lie}(\prod_{v\in\mathbf{a}}K_{n,v})\otimes_{\mathbb{R}}\mathbb{C}. Set Kv=Sp2n(𝒪Fv)K_{v}=\mathrm{Sp}_{2n}(\mathcal{O}_{F_{v}}) for a non-archimedean place vv. We denote by 𝒵n\mathcal{Z}_{n} the center of the universal enveloping algebra 𝒰(𝔤n)\mathcal{U}(\mathfrak{g}_{n}). We then obtain the well-known decomposition

𝔤n=𝔨n𝔭n,+𝔭n,\mathfrak{g}_{n}=\mathfrak{k}_{n}\oplus\mathfrak{p}_{n,+}\oplus\mathfrak{p}_{n,-}

where 𝔭n,+\mathfrak{p}_{n,+} (resp. 𝔭n,\mathfrak{p}_{n,-}) is the Lie subalgebra of 𝔤n\mathfrak{g}_{n} corresponding to the holomorphic tangent space (resp. anti-holomorphic tangent space) of nd\mathfrak{H}_{n}^{d} at 𝐢\mathbf{i}. It is well-known that the Lie algebras 𝔤n\mathfrak{g}_{n} and 𝔨n\mathfrak{k}_{n} have the same Cartan subalgebra. We fix such a Cartan subalgebra. Then the root system of 𝔰𝔭2n()\mathfrak{sp}_{2n}(\mathbb{C}) is

Φ={±(ei+ej),±(eke),1ijn,1k<n}.\Phi=\{\;\pm(e_{i}+e_{j}),\;\pm(e_{k}-e_{\ell}),\quad 1\leq i\leq j\leq n,1\leq k<\ell\leq n\;\}.

We consider the set

Φ+={(ei+ej),eke,1ijn,1k<n}\displaystyle\Phi^{+}=\{\;-(e_{i}+e_{j}),\;e_{k}-e_{\ell},\quad 1\leq i\leq j\leq n,1\leq k<\ell\leq n\;\}

to be a positive root system. Let ρ\rho be half the sum of positive roots. Put ρi,n=n(i1)/2\rho_{i,n}=n-(i-1)/2 and ρn=ρn,n\rho_{n}=\rho_{n,n}. This corresponds to half the sum of roots in the unipotent subgroup of Pi,nP_{i,n}. For λ=(λ1,v,,λn,v)vv𝐚n\lambda=(\lambda_{1,v},\ldots,\lambda_{n,v})_{v}\in\bigoplus_{v\in\mathbf{a}}\mathbb{C}^{n}, we say that λ\lambda is a weight if λi,vλi+1,v\lambda_{i,v}-\lambda_{i+1,v}\in\mathbb{Z} for any vv and 1in11\leq i\leq n-1. The weight λ\lambda is 𝔨n\mathfrak{k}_{n}-dominant if λi,vλi+1,v0\lambda_{i,v}-\lambda_{i+1,v}\geq 0 for any vv and 1in11\leq i\leq n-1. For a 𝔨n\mathfrak{k}_{n}-dominant weight λ\lambda, let ρλ\rho_{\lambda} be an irreducible finite-dimensional representation of 𝔨n\mathfrak{k}_{n}. When λ\lambda is integral, i.e., any entry of λ\lambda is an integer, we identify ρλ\rho_{\lambda} as the derivative of an irreducible finite-dimensional representation of Kn,K_{n,\mathbb{C}} with highest weight λ\lambda. We then write the representation of Kn,K_{n,\mathbb{C}} by the same ρλ\rho_{\lambda}.

We fix a non-trivial additive character ψ=vψv\psi=\bigotimes_{v}\psi_{v} of F\𝔸FF\backslash\mathbb{A}_{F} as follows: If F=F=\mathbb{Q}, let

ψp(x)\displaystyle\psi_{p}(x) =exp(2π1y),xp,\displaystyle=\exp(-2\pi\sqrt{-1}\,y),\qquad x\in\mathbb{Q}_{p},
ψ(x)\displaystyle\psi_{\infty}(x) =exp(2π1x),x,\displaystyle=\exp(2\pi\sqrt{-1}\,x),\qquad x\in\mathbb{R},

where ym=1pmy\in\cup_{m=1}^{\infty}p^{-m}\mathbb{Z} such that xypx-y\in\mathbb{Z}_{p}. In general, for an archimedean place vv of FF, put ψv=ψ\psi_{v}=\psi_{\infty} and for a non-archimedean place vv with the rational prime pp divisible by vv, put ψv(x)=ψp(TrFv/p(x))\psi_{v}(x)=\psi_{p}(\mathrm{Tr}_{F_{v}/\mathbb{Q}_{p}}(x)).

For a function ff on a group GG, let rr be the right translation, i.e., r(g)f(h)=f(hg)r(g)f(h)=f(hg) for any g,hGg,h\in G. For a subset HH of GG, we denote by f|Hf|_{H} the restriction of ff to HH. Let GG be a Lie group with the Lie algebra 𝔤\mathfrak{g}. For a smooth function ff on GG and X𝔤X\in\mathfrak{g}, put

Xf(g)=ddt|t=0f(gexp(tX)),gG.X\cdot f(g)=\left.\frac{d}{dt}\right|_{t=0}f(g\exp(tX)),\qquad g\in G.

For the action of Gn(𝔸)G_{n}(\mathbb{A}_{\mathbb{Q}}), we mean the Gn(𝔸,fin)×(𝔤n,Kn,)G_{n}(\mathbb{A}_{\mathbb{Q},\mathrm{fin}})\times(\mathfrak{g}_{n},K_{n,\infty})-action.

2. Nearly holomorphic Hilbert-Siegel modular forms and automorphic forms

In this section, we review the definition and arithmeticity of nearly holomorphic Hilbert-Siegel modular forms. We also recall some properties of nearly holomorphic automorphic forms on Gn(𝔸)G_{n}(\mathbb{A}_{\mathbb{Q}}) and basic terminologies of automorphic forms.

2.1. Differential operators on the Siegel upper half space

We recall the differential operators on n\mathfrak{H}_{n}. For details, see [Shi00, §12]. Fix a basis on Symn()\mathrm{Sym}_{n}(\mathbb{C}) by {(1+δi,j)1(ei,j+ej,i)1ijn}\{(1+\delta_{i,j})^{-1}(e_{i,j}+e_{j,i})\mid 1\leq i\leq j\leq n\}. We denote the basis by {εν}\{\varepsilon_{\nu}\}. For uSymn()u\in\mathrm{Sym}_{n}(\mathbb{C}), write u=νuνενu=\sum_{\nu}u_{\nu}\varepsilon_{\nu} with uνu_{\nu}\in\mathbb{C} and for znz\in\mathfrak{H}_{n}, write z=νzνενz=\sum_{\nu}z_{\nu}\varepsilon_{\nu} with zνz_{\nu}\in\mathbb{C}. For a non-negative integer ee and a finite-dimensional vector space VV, let Se(Symn(),V)S_{e}(\mathrm{Sym}_{n}(\mathbb{C}),V) be the space of VV-valued homogeneous polynomial maps of degree ee on Symn()\mathrm{Sym}_{n}(\mathbb{C}) and Mle(Symn(),V)\mathrm{Ml}_{e}(\mathrm{Sym}_{n}(\mathbb{C}),V) the space of ee-multilinear maps on Symn()e\mathrm{Sym}_{n}(\mathbb{C})^{e} to VV. Note that Se(Symn(),V)S_{e}(\mathrm{Sym}_{n}(\mathbb{C}),V) can be viewed as the space of symmetric elements of Mle(Symn(),V)\mathrm{Ml}_{e}(\mathrm{Sym}_{n}(\mathbb{C}),V). For a representation ρ\rho of GLn()\mathrm{GL}_{n}(\mathbb{C}) on VV, we define representations ρτe\rho\otimes\tau^{e} and ρσe\rho\otimes\sigma^{e} on Mle(Symn(),V)\mathrm{Ml}_{e}(\mathrm{Sym}_{n}(\mathbb{C}),V) by

((ρτe)(a)h)(u1,,ue)=ρ(a)h(atu1a,,atuea)((\rho\otimes\tau^{e})(a)h)(u_{1},\ldots,u_{e})=\rho(a)h({{}^{t}a}u_{1}a,\ldots,{{}^{t}a}u_{e}a)

and

((ρσe)(a)h)(u1,,ue)=ρ(a)h(a1u1a1t,,a1uea1t),((\rho\otimes\sigma^{e})(a)h)(u_{1},\ldots,u_{e})=\rho(a)h({a}^{-1}u_{1}{{}^{t}a^{-1}},\ldots,{a}^{-1}u_{e}{{}^{t}a^{-1}}),

respectively. Here, hMle(Symn(),V)h\in\mathrm{Ml}_{e}(\mathrm{Sym}_{n}(\mathbb{C}),V), aGLn()a\in\mathrm{GL}_{n}(\mathbb{C}) and (u1,,ue)Symn()e(u_{1},\ldots,u_{e})\in\mathrm{Sym}_{n}(\mathbb{C})^{e}. The symbols ρτe\rho\otimes\tau^{e} and ρσe\rho\otimes\sigma^{e} also denote the restrictions to the representations space Se(Symn(),V)S_{e}(\mathrm{Sym}_{n}(\mathbb{C}),V).

For fC(n,V)f\in C^{\infty}(\mathfrak{H}_{n},V), we define functions Df,D¯f,Cf,E,fDf,\overline{D}f,Cf,E,f on C(n,S1(Symn(),V))C^{\infty}(\mathfrak{H}_{n},S_{1}(\mathrm{Sym}_{n}(\mathbb{C}),V)) by

((Df)(z))(u)=νuνfzν(z)\displaystyle((Df)(z))(u)=\sum_{\nu}u_{\nu}\frac{\partial f}{\partial z_{\nu}}(z) ,((D¯f)(z))(u)=νuνfzν¯(z),\displaystyle,\qquad((\overline{D}f)(z))(u)=\sum_{\nu}u_{\nu}\frac{\partial f}{\partial\overline{z_{\nu}}}(z),
((Cf)(z))(u)=4((Df)(z))(yuy)\displaystyle((Cf)(z))(u)=4((Df)(z))(yuy) ,((Ef)(z))(u)=4((D¯f)(z))(yuy).\displaystyle,\qquad((Ef)(z))(u)=4((\overline{D}f)(z))(yuy).

Here, u=νuνενSymn(),z=νzνενnu=\sum_{\nu}u_{\nu}\varepsilon_{\nu}\in\mathrm{Sym}_{n}(\mathbb{C}),z=\sum_{\nu}z_{\nu}\varepsilon_{\nu}\in\mathfrak{H}_{n} and y=Im(z)y=\mathrm{Im}(z). For fC(n,V)f\in C^{\infty}(\mathfrak{H}_{n},V), we say that ff is nearly holomorphic if there exists ee such that Eef=0E^{e}f=0.

2.2. Definition

Let FF be the fixed totally real field. For an integral ideal 𝔫\mathfrak{n} of FF, set

Γ(𝔫)={γSp2n(𝒪F)|γ𝟏2nMat2n(𝔫)}.\Gamma(\mathfrak{n})=\left\{\gamma\in\mathrm{Sp}_{2n}(\mathcal{O}_{F})\,\middle|\,\gamma-\mathbf{1}_{2n}\in\mathrm{Mat}_{2n}(\mathfrak{n})\right\}.

The group Γ(𝔫)\Gamma(\mathfrak{n}) is called the principal congruence subgroup of Gn()G_{n}(\mathbb{Q}) of level 𝔫\mathfrak{n}. We say that a subgroup Γ\Gamma of Gn()G_{n}(\mathbb{Q}) is a congruence subgroup if there exists an integral ideal 𝔫\mathfrak{n} such that Γ\Gamma contains Γ(𝔫)\Gamma(\mathfrak{n}) and [Γ:Γ(𝔫)]<[\Gamma\colon\Gamma(\mathfrak{n})]<\infty. In this subsection, we regard Gn()G_{n}(\mathbb{Q}) as a subgroup of Gn()=v𝐚Sp2n(Fv)G_{n}(\mathbb{R})=\prod_{v\in\mathbf{a}}\mathrm{Sp}_{2n}(F_{v}) by γ(1(γ),,d(γ))\gamma\longmapsto(\infty_{1}(\gamma),\ldots,\infty_{d}(\gamma)). Similarly, we regard a congruence subgroup Γ\Gamma of Gn()G_{n}(\mathbb{Q}) as a subgroup of Gn()G_{n}(\mathbb{R}).

We define the factor of automorphy j:Gn()×ndGLn()dj\colon G_{n}(\mathbb{R})\times\mathfrak{H}_{n}^{d}\longrightarrow\mathrm{GL}_{n}(\mathbb{C})^{d} by

j(g,z)=(cvzv+dv)vv𝐚GLn()=GLn()d,g=((avbvcvdv))vGn(),z=(zv)vnd.j(g,z)=(c_{v}z_{v}+d_{v})_{v}\in\prod_{v\in\mathbf{a}}\mathrm{GL}_{n}(\mathbb{C})=\mathrm{GL}_{n}(\mathbb{C})^{d},\qquad g=\left(\begin{pmatrix}a_{v}&b_{v}\\ c_{v}&d_{v}\end{pmatrix}\right)_{v}\in G_{n}(\mathbb{R}),\quad z=(z_{v})_{v}\in\mathfrak{H}_{n}^{d}.

For a representation ρ\rho of Kn,K_{n,\mathbb{C}} on VV, set jρ=ρjj_{\rho}=\rho\circ j. For gGn()g\in G_{n}(\mathbb{R}), we define the slash operator |ρg|_{\rho}g on C(nd,V)C^{\infty}(\mathfrak{H}_{n}^{d},V) by

(f|ρg)(z1,,zd)=jρ(g,z)1f(γ(z1,,zd)),(f|_{\rho}g)(z_{1},\ldots,z_{d})=j_{\rho}(g,z)^{-1}f(\gamma(z_{1},\ldots,z_{d})),

for fC(nd,V)f\in C^{\infty}(\mathfrak{H}_{n}^{d},V) and (z1,,zd)nd(z_{1},\ldots,z_{d})\in\mathfrak{H}^{d}_{n}. Let Γ\Gamma be a congruence subgroup of Gn()G_{n}(\mathbb{Q}). Suppose that a function fC(nd,V)f\in C^{\infty}(\mathfrak{H}_{n}^{d},V) satisfies the automorphy f|ργ=ff|_{\rho}\gamma=f for any γΓ\gamma\in\Gamma. Then, ff has the Fourier expansion

(f|ργ)(z)=hSymn(F)cf(h,y,γ)𝐞(tr(hz)),znd,y=Im(z)(f|_{\rho}\gamma)(z)=\sum_{h\in\mathrm{Sym}_{n}(F)}c_{f}(h,y,\gamma)\mathbf{e}(\mathrm{tr}({hz})),\qquad z\in\mathfrak{H}_{n}^{d},\,y=\mathrm{Im}(z)

where 𝐞(tr(hz))=exp(2π1j=1htr(j(h)zj))\mathbf{e}(\mathrm{tr}(hz))=\exp(2\pi\sqrt{-1}\,\sum_{j=1}^{h}\mathrm{tr}(\infty_{j}(h)z_{j})) for (z1,,zd)nd(z_{1},\ldots,z_{d})\in\mathfrak{H}_{n}^{d} and hSymn(F)h\in\mathrm{Sym}_{n}(F). We consider the following condition: If cf(h,y,γ)0c_{f}(h,y,\gamma)\neq 0, the matrix hh is positive semi-definite. We call this condition the cusp condition. We say that a VV-valued CC^{\infty}-function ff on nd\mathfrak{H}_{n}^{d} is a nearly holomorphic Hilbert-Siegel modular form of weight ρ\rho with respect to Γ\Gamma if ff satisfies the following conditions:

  • ff is a nearly holomorphic function.

  • f|ργ=ff|_{\rho}\gamma=f for all γΓ\gamma\in\Gamma.

  • ff satisfies the cusp condition.

We denote by Nρ(Γ)N_{\rho}(\Gamma) the space of nearly holomorphic Hilbert-Siegel modular forms of weight ρ\rho with respect to Γ\Gamma. In the following, for modular forms, we mean a (nearly holomorphic) Hilbert-Siegel modular forms. By Köecher principle, we can remove the cusp condition if n>1n>1 or FF\neq\mathbb{Q}. For the proof, see [Hor20a, Proposition 4.1] for n>1n>1. We can give the same proof for the case of FF\neq\mathbb{Q}. For simplicity, if ρ=detk\rho=\det^{k}, we say that a modular form of weight detk\det^{k} is a modular form of weight kk.

2.3. Aut()\mathrm{Aut}(\mathbb{C}) action for nearly holomorphic Hilbert-Siegel modular forms and the holomorphic projection

Let ff be a nearly holomorphic modular form of weight ρ\rho with respect to Γ\Gamma. Take a model VV of ρ\rho and fix a rational structure of VV. Then, Shimura introduced the Aut()\mathrm{Aut}(\mathbb{C})-action on ff. For details, see [Shi00, §14.11] and [HPSS21, §3.3]. For σAut()\sigma\in\mathrm{Aut}(\mathbb{C}), we denote by fσ{{}^{\sigma}f} the action of σ\sigma on ff. For a weight ρ=v𝐚ρv\rho=\bigotimes_{v\in\mathbf{a}}\rho_{v}, put ρσ=v𝐚ρσv{{}^{\sigma}\rho}=\bigotimes_{v\in\mathbf{a}}\rho_{\sigma\circ v}. The following theorem is proved in [Shi00, Theorem 14.12].

Theorem 2.3.1.

For fNρ(Γ)f\in N_{\rho}(\Gamma) and σAut()\sigma\in\mathrm{Aut}(\mathbb{C}), one has fσNρσ(Γ){{}^{\sigma}f}\in N_{{}^{\sigma}\rho}(\Gamma).

Let Mρ(Γ)M_{\rho}(\Gamma) be the space of holomorphic functions in Nρ(Γ)N_{\rho}(\Gamma). Set Nρp(Γ)=Nρ(pv)v(Γ)={fNρ(Γ)Evpv+1f=0 for any v𝐚}N^{p}_{\rho}(\Gamma)=N_{\rho}^{(p_{v})_{v}}(\Gamma)=\{f\in N_{\rho}(\Gamma)\mid\text{$E_{v}^{p_{v}+1}f=0$ for any $v\in\mathbf{a}$}\}. The, Nρ0(Γ)=Mρ(Γ)N^{0}_{\rho}(\Gamma)=M_{\rho}(\Gamma). Let ρ=vρv\rho=\bigotimes_{v}\rho_{v} be a character of Kn,K_{n,\mathbb{C}} with the weight (kv)v𝐚(k_{v})_{v\in\mathbf{a}}. Take non-negative integers pvp_{v} satisfies kv>n+pvk_{v}>n+p_{v} or kv<n+(3pv)/2k_{v}<n+(3-p_{v})/2 for any v𝐚v\in\mathbf{a}. Put p=(pv)vp=(p_{v})_{v}. Then, in [Shi00, §15.3], Shimura introduced a projection 𝔄:Nρp(Γ)Mρ(Γ)\mathfrak{A}\colon N_{\rho}^{p}(\Gamma)\longrightarrow M_{\rho}(\Gamma). The projection 𝔄\mathfrak{A} is called the holomorphic projection. By Shimura [Shi00, Proposition 15.3], it commutes with the Aut()\mathrm{Aut}(\mathbb{C}) actions as follows:

Theorem 2.3.2.

With the above notation, for any σAut()\sigma\in\mathrm{Aut}(\mathbb{C}) and fNρ(Γ)f\in N_{\rho}(\Gamma), one has 𝔄(fσ)=𝔄σ(f)\mathfrak{A}({{}^{\sigma}f})={{}^{\sigma}\mathfrak{A}(f)}.

In [HPSS21, §3.4], we define other projection operators 𝔭χ\mathfrak{p}_{\chi} associated to infinitesimal characters χ\chi of 𝒵n\mathcal{Z}_{n}. This can be viewed as a generalization of the holomorphic projection 𝔄\mathfrak{A}. In this paper, we study the image of 𝔭χ\mathfrak{p}_{\chi} in terms of (𝔤n,Kn,)(\mathfrak{g}_{n},K_{n,\infty})-modules.

2.4. Automorphic forms on Gn(𝔸)G_{n}(\mathbb{A}_{\mathbb{Q}})

Let P=MNP=MN be a standard parabolic subgroup of GnG_{n}. For a smooth function ϕ:N(𝔸)M()\Gn(𝔸)\phi:N(\mathbb{A}_{\mathbb{Q}})M(\mathbb{Q})\backslash G_{n}(\mathbb{A}_{\mathbb{Q}})\longrightarrow\mathbb{C}, we say that ϕ\phi is automorphic if it satisfies the following conditions:

  • ϕ\phi is right KnK_{n}-finite.

  • ϕ\phi is 𝒵n\mathcal{Z}_{n}-finite.

  • ϕ\phi is slowly increasing.

We denote by 𝒜(P\Gn)\mathcal{A}(P\backslash G_{n}) the space of automorphic forms on N(𝔸)M()\Gn(𝔸)N(\mathbb{A}_{\mathbb{Q}})M(\mathbb{Q})\backslash G_{n}(\mathbb{A}_{\mathbb{Q}}). For simplicity, we write 𝒜(Gn)\mathcal{A}(G_{n}) when P=GnP=G_{n}. The space 𝒜(P\Gn)\mathcal{A}(P\backslash G_{n}) is stable under the action of Gn(𝔸)G_{n}(\mathbb{A}_{\mathbb{Q}}).

For parabolic subgroups PP and QQ of GnG_{n}, we say that PP and QQ are associate if the split components APA_{P} and AQA_{Q} are Gn()G_{n}(\mathbb{Q})-conjugate. We denote by {P}\{P\} the associated class of the parabolic subgroup PP. For a locally integrable function φ\varphi on NP()\Gn(𝔸)N_{P}(\mathbb{Q})\backslash G_{n}(\mathbb{A}_{\mathbb{Q}}), set

φP(g)=NP()\NP(𝔸)φ(ng)𝑑n\varphi_{P}(g)=\int_{N_{P}(\mathbb{Q})\backslash N_{P}(\mathbb{A}_{\mathbb{Q}})}\varphi(ng)\,dn

where P=MPNPP=M_{P}N_{P} is the Levi decomposition of PP and the Haar measure dndn is normalized by

NP()\NP(𝔸)𝑑n=1.\int_{N_{P}(\mathbb{Q})\backslash N_{P}(\mathbb{A}_{\mathbb{Q}})}\,dn=1.

The function φP\varphi_{P} is called the constant term of φ\varphi along PP. If φ\varphi lies in 𝒜(P\Gn)\mathcal{A}(P\backslash G_{n}), φQ\varphi_{Q} is an automorphic form on NQ(𝔸)MQ()\Gn(𝔸)N_{Q}(\mathbb{A}_{\mathbb{Q}})M_{Q}(\mathbb{Q})\backslash G_{n}(\mathbb{A}_{\mathbb{Q}}) for a parabolic subgroup QPQ\subset P. We call φ\varphi cuspidal if φQ\varphi_{Q} is zero for any standard parabolic subgroup QQ of GG with QPQ\subsetneq P. We denote by 𝒜cusp(P\Gn)\mathcal{A}_{\mathrm{cusp}}(P\backslash G_{n}) the space of cusp forms in 𝒜(P\Gn)\mathcal{A}(P\backslash G_{n}). For a character ξ\xi of the split component APA_{P}^{\infty}, put

𝒜(P\G)ξ={φ𝒜(P\Gn)φ(ag)=aξ+ρPφ(g) for any gGn(𝔸) and aAP}.\mathcal{A}(P\backslash G)_{\xi}=\{\varphi\in\mathcal{A}(P\backslash G_{n})\mid\text{$\varphi(ag)=a^{\xi+\rho_{P}}\varphi(g)$ for any $g\in G_{n}(\mathbb{A}_{\mathbb{Q}})$ and $a\in A_{P}^{\infty}$}\}.

Here, ρP\rho_{P} is the character of APA_{P}^{\infty} corresponding to half the sum of roots of NPN_{P} relative to APA_{P}. We define 𝒜cusp(P\G)ξ\mathcal{A}_{\mathrm{cusp}}(P\backslash G)_{\xi} similarly. Set

𝒜(P\Gn)Z=ξ𝒜(P\Gn)ξ,𝒜cusp(P\Gn)Z=ξ𝒜cusp(P\Gn)ξ.\mathcal{A}(P\backslash G_{n})_{Z}=\bigoplus_{\xi}\mathcal{A}(P\backslash G_{n})_{\xi},\qquad\mathcal{A}_{\mathrm{cusp}}(P\backslash G_{n})_{Z}=\bigoplus_{\xi}\mathcal{A}_{\mathrm{cusp}}(P\backslash G_{n})_{\xi}.

Here, ξ\xi runs over all the characters of APA_{P}^{\infty}. Let 𝔞P\mathfrak{a}_{P} be the real vector space generated by coroots associated to the root system of GnG_{n} relative to APA_{P}. Then, by [MW95, Lemma I.3.2], there exist canonical isomorphisms

(2.4.1) [𝔞P]𝒜(P\Gn)Z𝒜(P\G),[𝔞P]𝒜cusp(P\Gn)Z𝒜cusp(P\Gn).\displaystyle\mathbb{C}[\mathfrak{a}_{P}]\otimes\mathcal{A}(P\backslash G_{n})_{Z}\cong\mathcal{A}(P\backslash G),\qquad\mathbb{C}[\mathfrak{a}_{P}]\otimes\mathcal{A}_{\mathrm{cusp}}(P\backslash G_{n})_{Z}\cong\mathcal{A}_{\mathrm{cusp}}(P\backslash G_{n}).

For a standard Levi subgroup MM, set

M(𝔸)1=χHomconti(M(𝔸),×)Ker(|χ|).M(\mathbb{A}_{\mathbb{Q}})^{1}=\bigcap_{\chi\in\mathop{\mathrm{Hom}}_{\mathrm{conti}}(M(\mathbb{A}_{\mathbb{Q}}),\mathbb{C}^{\times})}\mathrm{Ker}(|\chi|).

For a function ff on Gn(𝔸)G_{n}(\mathbb{A}_{\mathbb{Q}}) and gGn(𝔸)g\in G_{n}(\mathbb{A}_{\mathbb{Q}}), let fgf_{g} be the function on MP(𝔸)1M_{P}(\mathbb{A}_{\mathbb{Q}})^{1} defined by mmρPf(mg)m\longmapsto m^{-\rho_{P}}f(mg). Put

𝒜(Gn){P}={φ𝒜(G)|φQ,ak is orthogonal to all cusp forms on MQ(𝔸Q)1for any aAQ,kKn, and Q{P}}.\mathcal{A}(G_{n})_{\{P\}}=\left\{\varphi\in\mathcal{A}(G)\,\middle|\,\begin{matrix}\text{$\varphi_{Q,ak}$ is orthogonal to all cusp forms on $M_{Q}(\mathbb{A}_{Q})^{1}$}\\ \text{for any $a\in A_{Q},k\in K_{n}$, and $Q\not\in\{P\}$}\end{matrix}\right\}.

By [MW95, Lemma I.3.4], 𝒜(Gn){G}\mathcal{A}(G_{n})_{\{G\}} is equal to 𝒜cusp(Gn)\mathcal{A}_{\mathrm{cusp}}(G_{n}). More precisely, Langlands [Lan06] had proven the following result:

Theorem 2.4.1.

With the above notation, we have

𝒜(Gn)={P}𝒜(Gn){P},\mathcal{A}(G_{n})=\bigoplus_{\{P\}}\mathcal{A}(G_{n})_{\{P\}},

where {P}\{P\} runs through all associated classes of parabolic subgroups.

Let MM be a standard Levi subgroup of GnG_{n} and τ\tau an irreducible cuspidal automorphic representation of M(𝔸)M(\mathbb{A}_{\mathbb{Q}}). We say that a cuspidal datum is a pair (M,τ)(M,\tau) such that MM is a Levi subgroup of GnG_{n} and that τ\tau is an irreducible cuspidal automorphic representation of M(𝔸)M(\mathbb{A}_{\mathbb{Q}}). Take wWnw\in W_{n}. Put Mw=wMw1M^{w}=wMw^{-1} and let Pw=MwNwP^{w}=M^{w}N^{w} be the standard parabolic subgroup with Levi subgroup MwM^{w}. The irreducible cuspidal automorphic representation τw\tau^{w} of Mw(𝔸)M^{w}(\mathbb{A}_{\mathbb{Q}}) is defined by τw(m)=τ(w1mw)\tau^{w}(m^{\prime})=\tau(w^{-1}m^{\prime}w) for mMw(𝔸)m^{\prime}\in M^{w}(\mathbb{A}_{\mathbb{Q}}). Two cuspidal data (M,τ)(M,\tau) and (M,τ)(M^{\prime},\tau^{\prime}) are called equivalent if there exists wW(M)w\in W(M) such that M=MwM^{\prime}=M^{w} and that τ=τw\tau^{\prime}=\tau^{w}. Here we put

W(M)={wW|wMw1 is a standard Levi subgroup of Gnand w has a minimal length in wWM}W(M)=\left\{w\in W\,\middle|\,\begin{matrix}\text{$wMw^{-1}$ is a standard Levi subgroup of $G_{n}$}\\ \text{and $w$ has a minimal length in $wW_{M}$}\end{matrix}\right\}

where WMW_{M} is the Weyl group of MM.

Let 𝒜(Gn)(M,τ)\mathcal{A}(G_{n})_{(M,\tau)} is the subspace of automorphic forms in 𝒜(Gn)\mathcal{A}(G_{n}) with the cuspidal support (M,τ)(M,\tau). For the definition, see [MW95, §III.2.6]. Then the following result is well-known. For example, see [MW95, Theorem III.2.6].

Theorem 2.4.2.

The space 𝒜(Gn)\mathcal{A}(G_{n}) is decomposed as

𝒜(Gn)=(M,τ)𝒜(Gn)(M,τ).\mathcal{A}(G_{n})=\bigoplus_{(M,\tau)}\mathcal{A}(G_{n})_{(M,\tau)}.

Here, (M,τ)(M,\tau) runs through all equivalence classes of cuspidal data.

Let PP be a standard parabolic subgroup of GnG_{n} with standard Levi subgroup MM and π\pi an irreducible cuspidal automorphic representation of M(𝔸)M(\mathbb{A}_{\mathbb{Q}}). Put

𝒜cusp(P\Gn)π={φ𝒜(P\Gn)φk𝒜cusp(M)π for any kKn}.\mathcal{A}_{\mathrm{cusp}}(P\backslash G_{n})_{\pi}=\{\varphi\in\mathcal{A}(P\backslash G_{n})\mid\text{$\varphi_{k}\in\mathcal{A}_{\mathrm{cusp}}(M)_{\pi}$ for any $k\in K_{n}$}\}.

Here, 𝒜cusp(M)π\mathcal{A}_{\mathrm{cusp}}(M)_{\pi} is the π\pi-isotypic component of 𝒜cusp(M)\mathcal{A}_{\mathrm{cusp}}(M). For an automorphic form φ\varphi, there exists a finite correction of cuspidal data (M,τ)(M,\tau) such that

φ(M,τ)𝒜(Gn)(M,τ)\varphi\in\bigoplus_{(M,\tau)}\mathcal{A}(G_{n})_{(M,\tau)}

by Theorem 2.4.2. Let φPcusp\varphi_{P}^{\mathrm{cusp}} be the cuspidal part of φP\varphi_{P}. Then, there exists a finite number of irreducible cuspidal automorphic representations π1,,π\pi_{1},\ldots,\pi_{\ell} of MP(𝔸)M_{P}(\mathbb{A}_{\mathbb{Q}}) such that

φPcuspj=1[𝔞P]𝒜cusp(P\Gn)πj.\varphi_{P}^{\mathrm{cusp}}\in\bigoplus_{j=1}^{\ell}\mathbb{C}[\mathfrak{a}_{P}]\otimes\mathcal{A}_{\mathrm{cusp}}(P\backslash G_{n})_{\pi_{j}}.

We say that a set M{χπ1,,χπ}\cup_{M}\{\chi_{\pi_{1}},\ldots,\chi_{\pi_{\ell}}\} is the set of cuspidal exponents of φ\varphi. Here, χπj\chi_{\pi_{j}} is the central character of πj\pi_{j}. For a character χ\chi of the center of MP(𝔸)M_{P}(\mathbb{A}_{\mathbb{Q}}), we call the restriction of χ\chi to APA_{P}^{\infty} the real part of χ\chi.

Let us now introduce the notion for some induced representations on Gn(𝔸)G_{n}(\mathbb{A}_{\mathbb{Q}}) and Sp2n(Fv)\mathrm{Sp}_{2n}(F_{v}). For a character μ\mu of GLn(𝔸F)\mathrm{GL}_{n}(\mathbb{A}_{F}), we mean an automorphic character, i.e., GLn(F)\mathrm{GL}_{n}(F) is contained in the kernel of μ\mu. Let μ\mu be a character of GLi(𝔸F)\mathrm{GL}_{i}(\mathbb{A}_{F}) and an irreducible cuspidal automorphic representation π\pi of Gni(𝔸)G_{n-i}(\mathbb{A}_{\mathbb{Q}}). We define the space IndPi,n(𝔸)Gn(𝔸)(μ||sπ)\mathrm{Ind}_{P_{i,n}(\mathbb{A}_{\mathbb{Q}})}^{G_{n}(\mathbb{A}_{\mathbb{Q}})}(\mu|\cdot|^{s}\boxtimes\pi) by the space of smooth functions φ\varphi on NPi,n(𝔸)Pi,n()\Gn(𝔸)N_{P_{i,n}}(\mathbb{A}_{\mathbb{Q}})P_{i,n}(\mathbb{Q})\backslash G_{n}(\mathbb{A}_{\mathbb{Q}}) such that

  • φ\varphi is an automorphic form.

  • For any kKnk\in K_{n}, the function φk\varphi_{k} lies in the μ||sπ\mu|\cdot|^{s}\boxtimes\pi-isotypic component of Ldisc2(MPi,n(𝔸))L^{2}_{\mathrm{disc}}(M_{P_{i,n}}(\mathbb{A}_{\mathbb{Q}})).

We write

Ii,n(s,μ,π)=IndPi,n(𝔸)Gn(𝔸)(μ||sπ)andIn(s,μ)=IndPn(𝔸)Gn(𝔸)μ||s.I_{i,n}(s,\mu,\pi)=\mathrm{Ind}_{P_{i,n}(\mathbb{A}_{\mathbb{Q}})}^{G_{n}(\mathbb{A}_{\mathbb{Q}})}\left(\mu|\cdot|^{s}\boxtimes\pi\right)\quad\mathrm{and}\quad I_{n}(s,\mu)=\mathrm{Ind}_{P_{n}(\mathbb{A}_{\mathbb{Q}})}^{G_{n}(\mathbb{A}_{\mathbb{Q}})}\mu|\cdot|^{s}.

For a place vv of FF, we similarly write

Ii,n,v(s,μv,πv)=IndPi,n(Fv)Gn(Fv)(μv||sπv)andIn,v(s,μv)=IndPn(Fv)Gn(Fv)μv||s.I_{i,n,v}(s,\mu_{v},\pi_{v})=\mathrm{Ind}_{P_{i,n}(F_{v})}^{G_{n}(F_{v})}\left(\mu_{v}|\cdot|^{s}\boxtimes\pi_{v}\right)\quad\mathrm{and}\quad I_{n,v}(s,\mu_{v})=\mathrm{Ind}_{P_{n}(F_{v})}^{G_{n}(F_{v})}\mu_{v}|\cdot|^{s}.

Here, μv\mu_{v} is a character of GLi(Fv)\mathrm{GL}_{i}(F_{v}) and πv\pi_{v} is an irreducible representation of Sp2(ni)(Fv)\mathrm{Sp}_{2(n-i)}(F_{v}).

2.5. Nearly holomorphic automorphic forms

For an automorphic form φ\varphi on Gn(𝔸)G_{n}(\mathbb{A}_{\mathbb{Q}}), we say that φ\varphi is nearly holomorphic if φ\varphi is 𝔭n,\mathfrak{p}_{n,-}-finite. The symbol 𝒩(Gn)\mathcal{N}(G_{n}) denotes the space of nearly holomorphic automorphic forms on Gn(𝔸)G_{n}(\mathbb{A}_{\mathbb{Q}}). Put 𝒩(Gn)(M,τ)=𝒩(Gn)𝒜(Gn)(M,τ)\mathcal{N}(G_{n})_{(M,\tau)}=\mathcal{N}(G_{n})\cap\mathcal{A}(G_{n})_{(M,\tau)}. We say that an irreducible cuspidal automorphic representation π=vπv\pi=\bigotimes_{v}\pi_{v} of Gn(𝔸)G_{n}(\mathbb{A}_{\mathbb{Q}}) is holomorphic if πv\pi_{v} is an irreducible unitary highest weight representation of Sp2n(Fv)\mathrm{Sp}_{2n}(F_{v}) for any v𝐚v\in\mathbf{a}. In [Hor20b, Theorem 1.2], we determine the cuspidal components of nearly holomorphic automorphic forms as follows:

Proposition 2.5.1.

Let PP be a standard parabolic subgroup of GnG_{n} with the standard Levi subgroup MM.

  1. (1)

    With the above notation, the space 𝒩(Gn)(M,π)\mathcal{N}(G_{n})_{(M,\pi)} is non-zero only if PP is associated to Qi,nQ_{i,n} for some ii.

  2. (2)

    Let Π=μ1μiπ\Pi=\mu_{1}\boxtimes\cdots\boxtimes\mu_{i}\boxtimes\pi be an irreducible cuspidal automorphic representation of MQi,n(𝔸)=(ResF/GL1)(𝔸F)i×Gni(𝔸)M_{Q_{i,n}}(\mathbb{A}_{\mathbb{Q}})=(\mathop{\mathrm{Res}}_{F/\mathbb{Q}}\mathrm{GL}_{1})(\mathbb{A}_{F})^{i}\times G_{n-i}(\mathbb{A}_{\mathbb{Q}}). If the space 𝒩(Gn)(Qi,n,Π)\mathcal{N}(G_{n})_{(Q_{i,n},\Pi)} is non-zero, we have

    • μ1==μi\mu_{1}=\cdots=\mu_{i}.

    • π\pi is a holomorphic cuspidal automorphic representation of Gni(𝔸)G_{n-i}(\mathbb{A}_{\mathbb{Q}}).

Let μ\mu be a character of GL1(𝔸F)\mathrm{GL}_{1}(\mathbb{A}_{F}). For simplicity the notation, we denote by μ\mu the character μμ\mu\boxtimes\cdots\boxtimes\mu of GL1(𝔸F)i\mathrm{GL}_{1}(\mathbb{A}_{F})^{i}. In [Hor20b], we determine the structure of the space 𝒩(Gn)(M,τ)\mathcal{N}(G_{n})_{(M,\tau)} explicitly under several assumptions.

2.6. Modular forms and automorphic forms

We recall the correspondence of modular forms on the Siegel upper half space and automorphic forms on Gn(𝔸)G_{n}(\mathbb{A}_{\mathbb{Q}}). Fix a weight ρ\rho and a congruence subgroup Γ\Gamma. We embed Γ\Gamma into Gn(𝔸,fin)G_{n}(\mathbb{A}_{\mathbb{Q},\mathrm{fin}}) diagonally. Let KΓK_{\Gamma} be the closure of Γ\Gamma in Gn(𝔸,fin)G_{n}(\mathbb{A}_{\mathbb{Q},\mathrm{fin}}). Then, KΓK_{\Gamma} is an open compact subgroup of Gn(𝔸,fin)G_{n}(\mathbb{A}_{\mathbb{Q},\mathrm{fin}}).

By the strong approximation, one has Gn(𝔸)=Gn()Gn()KΓG_{n}(\mathbb{A}_{\mathbb{Q}})=G_{n}(\mathbb{Q})G_{n}(\mathbb{R})K_{\Gamma}. For fNρ(Γ)f\in N_{\rho}(\Gamma) and vρv^{*}\in\rho^{*}, the dual of ρ\rho, put

φf,v(γgk)=(f|ρg)(𝐢),v,γgkGn()Gn()KΓ=Gn(𝔸).\varphi_{f,v^{*}}(\gamma g_{\infty}k)=\langle(f|_{\rho}g_{\infty})(\mathbf{i}),v^{*}\rangle,\qquad\gamma g_{\infty}k\in G_{n}(\mathbb{Q})G_{n}(\mathbb{R})K_{\Gamma}=G_{n}(\mathbb{A}_{\mathbb{Q}}).

This is well-defined. The map fvφf,vf\otimes v^{*}\longmapsto\varphi_{f,v^{*}} induces the inclusion

(2.6.1) Nρ(Γ)ρ𝒩(Gn).\displaystyle N_{\rho}(\Gamma)\otimes\rho^{*}\longrightarrow\mathcal{N}(G_{n}).

Put

𝒩(Gn)ρKΓ={φ𝒩(Gn)|φ generates ρ under the action of Kn, andφ(gk)=φ(g) for any gGn(𝔸) and kKΓ}.\mathcal{N}(G_{n})_{\rho}^{K_{\Gamma}}=\left\{\varphi\in\mathcal{N}(G_{n})\,\middle|\,\begin{matrix}\text{$\varphi$ generates $\rho$ under the action of $K_{n,\infty}$ and}\\ \text{$\varphi(gk)=\varphi(g)$ for any $g\in G_{n}(\mathbb{A}_{\mathbb{Q}})$ and $k\in K_{\Gamma}$}\end{matrix}\right\}.

By the choice of embedding U(n)GLn()\mathrm{U}(n)\xhookrightarrow{\quad}\mathrm{GL}_{n}(\mathbb{C}), the map (2.6.1) induces the isomorphism

(2.6.2) Nρ(Γ)ρ𝒩(Gn)ρKΓ.\displaystyle N_{\rho}(\Gamma)\otimes\rho^{*}\xrightarrow{\,\,\sim\,\,}\mathcal{N}(G_{n})_{\rho}^{K_{\Gamma}}.

For details, see [HPSS21, §3.2]. For a representation generated by fNρ(Γ)f\in N_{\rho}(\Gamma), we mean the representation generated by φf,v\varphi_{f,v^{*}} with 0vρ0\neq v^{*}\in\rho^{*}. Note that the representation is independent of the choice of v0v^{*}\neq 0.

3. Computations of unitary highest weight modules with a regular integral infinitesimal character

In this section, we introduce the parabolic BGG category 𝒪𝔭\mathcal{O}^{\mathfrak{p}} and unitarizable modules in this category. For later use, we compute extensions of certain modules and multiplicities of Kn,K_{n,\infty}-types.

3.1. parabolic BGG category

For simplicity the notation, throughout this section, we assume F=F=\mathbb{Q}. Let 𝔫\mathfrak{n} be a nilpotent subalgebra of 𝔤n\mathfrak{g}_{n}. For a 𝔤n\mathfrak{g}_{n}-module MM, we say that MM is locally 𝔫\mathfrak{n}-finite if 𝒰(𝔫)v\mathcal{U}(\mathfrak{n})\cdot v is finite-dimensional for any vMv\in M.

We consider the parabolic subalgebra 𝔭=𝔨n𝔭n,\mathfrak{p}=\mathfrak{k}_{n}\oplus\mathfrak{p}_{n,-}. We define the full subcategory 𝒪𝔭\mathcal{O}^{\mathfrak{p}} of the category of 𝔤n\mathfrak{g}_{n}-modules whose objects MM satisfy the following three conditions:

  • MM is finitely generated.

  • MM decomposes as a direct sum of irreducible finite-dimensional representations of 𝔨n\mathfrak{k}_{n}.

  • MM is locally 𝔭n,\mathfrak{p}_{n,-}-finite.

The category 𝒪𝔭\mathcal{O}^{\mathfrak{p}} is called the parabolic BGG category 𝒪𝔭\mathcal{O}^{\mathfrak{p}} with respect to 𝔭\mathfrak{p}. For further properties of the BGG category 𝒪\mathcal{O} and a parabolic BGG category 𝒪𝔭\mathcal{O}^{\mathfrak{p}}, see [Hum08].

Let us introduce the Verma modules. For a 𝔨n\mathfrak{k}_{n}-dominant weight λ\lambda, let VλV_{\lambda} be a model of ρλ\rho_{\lambda}. We regard VλV_{\lambda} as a 𝔭\mathfrak{p}-module by letting 𝔭n,\mathfrak{p}_{n,-} act trivially. Put

N(λ)=𝒰(𝔤n)𝒰(𝔭)Vλ.N(\lambda)=\mathcal{U}(\mathfrak{g}_{n})\otimes_{\mathcal{U}(\mathfrak{p})}V_{\lambda}.

Then, N(λ)N(\lambda) has the canonical left 𝔤n\mathfrak{g}_{n}-module structure. The module N(λ)N(\lambda) is called the (parabolic) Verma module of weight λ\lambda. Since N(λ)N(\lambda) is generated by a highest weight vector, N(λ)N(\lambda) has the unique irreducible quotient L(λ)L(\lambda). Note that N(λ)N(\lambda) and L(λ)L(\lambda) are objects in 𝒪𝔭\mathcal{O}^{\mathfrak{p}}.

For a 𝔤n\mathfrak{g}_{n}-module MM, we say that MM is a highest weight module if there exists a highest weight vector vMv\in M such that vv generates MM. By definition, Verma modules are highest weight modules. Moreover, N(λ)N(\lambda) has the following universality: For a highest weight module MM with the highest weight λ\lambda, there exists a surjective homomorphism N(λ)MN(\lambda)\twoheadrightarrow M.

For a weight λ\lambda, let χλ\chi_{\lambda} be the infinitesimal character with the Harish-Chandra parameter λ+ρ\lambda+\rho. Then, the Verma module N(λ)N(\lambda) has the infinitesimal character χλ\chi_{\lambda}. Note that for χ0\chi_{0}, we mean the infinitesimal character of the trivial representation. The infinitesimal characters χλ\chi_{\lambda} and χμ\chi_{\mu} are the same if and only if there exists wWnw\in W_{n} such that λ=wμ\lambda=w\cdot\mu. Here \cdot is the dot action defined by wμ=w(μ+ρ)ρw\cdot\mu=w(\mu+\rho)-\rho. For a weight λ\lambda, put 𝒪λ={wλwWn}\mathcal{O}_{\lambda}=\{w\cdot\lambda\mid w\in W_{n}\}. We say that λ\lambda is (dot-)regular if #𝒪λ=#Wn\#\mathcal{O}_{\lambda}=\#W_{n}. If λ\lambda is not of (dot-)regular, we say that λ\lambda is (dot-)singular.

For a nearly holomorphic automorphic form φ\varphi, we consider the 𝔤n\mathfrak{g}_{n}-module MM generated by φ\varphi under the right translation. Then, MM is a (𝔤n,Kn,)(\mathfrak{g}_{n},K_{n,\infty})-module. By the definition of the parabolic BGG category 𝒪𝔭\mathcal{O}^{\mathfrak{p}}, the 𝔤n\mathfrak{g}_{n}-module MM is an object in 𝒪𝔭\mathcal{O}^{\mathfrak{p}}.

3.2. First reduction point and unitarizability

We recall the definition of the first reduction point in the sense of [EHW83]. Let λ=(λ1,,λn)\lambda=(\lambda_{1},\ldots,\lambda_{n}) be a 𝔨n\mathfrak{k}_{n}-dominant weight with λn=n\lambda_{n}=n. We say that a real number r0=r0(λ)r_{0}=r_{0}(\lambda) is the first reduction point if the module N(λ+r0(1,,1))N(\lambda+r_{0}(-1,\ldots,-1)) is reducible and N(λ+r(1,,1))N(\lambda+r(-1,\ldots,-1)) is irreducible for r<r0r<r_{0}. Set p(λ)=#{iλi=λn}p(\lambda)=\#\{i\mid\lambda_{i}=\lambda_{n}\} and q(λ)=#{iλi=λn+1}q(\lambda)=\#\{i\mid\lambda_{i}=\lambda_{n}+1\}. One can compute the first reduction point explicitly by the result of Enright-Howe-Wallach [EHW83, Theorem 2.10].

Theorem 3.2.1.

Let λ=(λ1,,λn)\lambda=(\lambda_{1},\ldots,\lambda_{n}) be a 𝔨n\mathfrak{k}_{n}-dominant weight with λn=n\lambda_{n}=n. Then, the first reduction point r0r_{0} equals to (p(λ)+q(λ)+1)/2(p(\lambda)+q(\lambda)+1)/2.

Let r0r_{0} be the first reduction point. Then for r<r0r<r_{0}, the irreducible representation L(λ+r(1,,1))L(\lambda+r(-1,\ldots,-1)) is unitarizable. More precisely, we have the following by [EHW83, Theorem 2.8]:

Theorem 3.2.2.

With the same notation as in Theorem 3.2.1, L(λ+r(1,,1))L(\lambda+r(-1,\ldots,-1)) is unitarizable if and only if either of the following conditions holds:

  • r(p(λ)+q(λ)+1)/2r\leq(p(\lambda)+q(\lambda)+1)/2.

  • λ(1/2)n\lambda\in(1/2)\mathbb{Z}^{n} and rp(λ)+q(λ)/2r\leq p(\lambda)+q(\lambda)/2.

3.3. Dot-orbits of regular integral weights and unitary highest weight modules

Let λ=(λ1,,λn)n\lambda=(\lambda_{1},\ldots,\lambda_{n})\in\mathbb{Z}^{n} be a 𝔨n\mathfrak{k}_{n}-dominant integral weight. Let |λ||\lambda| be a multiset {|λ11|,|λ22|,,|λnn|}\{|\lambda_{1}-1|,|\lambda_{2}-2|,\ldots,|\lambda_{n}-n|\}. Then, the multiset is invariant under the dot-action, i.e., |λ|=|wλ||\lambda|=|w\cdot\lambda| for any wWnw\in W_{n}. We then say that λ\lambda is anti-dominant if λnn\lambda_{n}\geq n. We compute the dot-orbits of regular anti-dominant integral weights. Note that for any regular integral weight λ\lambda, there exists σW\sigma\in W such that σλ\sigma\cdot\lambda is anti-dominant. Moreover, such an anti-dominant weight is unique in the dot-orbit 𝒪λ\mathcal{O}_{\lambda}.

Lemma 3.3.1.

Let λ=(λ1,,λn)\lambda=(\lambda_{1},\ldots,\lambda_{n}) be a regular anti-dominant integral weight and σ\sigma an element of the Weyl group WnW_{n}. Suppose that the weight σλ\sigma\cdot\lambda is 𝔨n\mathfrak{k}_{n}-dominant and L(σλ)L(\sigma\cdot\lambda) is unitarizable. If σλλ\sigma\cdot\lambda\neq\lambda, one has λn=n+1\lambda_{n}=n+1.

Proof.

Put ω=σλ=(ω1,,ωn)\omega=\sigma\cdot\lambda=(\omega_{1},\ldots,\omega_{n}). Suppose that ωλ\omega\neq\lambda and L(ω)L(\omega) is unitarizable. By ωλ\omega\neq\lambda and the uniqueness of anti-dominant weights in 𝒪λ\mathcal{O}_{\lambda}, one has ωn<n\omega_{n}<n. Set p=p(ω)p=p(\omega) and q=q(ω)q=q(\omega). Since L(ω)L(\omega) is unitarizable, one has

(3.3.1) npq/2ωn<n.\displaystyle n-p-q/2\leq\omega_{n}<n.

If ωn>np\omega_{n}>n-p, there exists np+1jnn-p+1\leq j\leq n such that ωjj=0\omega_{j}-j=0. Then, ω\omega is singular. This is contradiction. Similarly, if ωn<np\omega_{n}<n-p, one has q>0q>0 by (3.3.1). By (3.3.1) and the unitarizability of L(ω)L(\omega), either of the following statements holds:

  • There exists jj such that ωj=j\omega_{j}=j.

  • There exists i<ji<j such that ωii=jωj\omega_{i}-i=j-\omega_{j}.

Thus, ω\omega is singular. This is contradiction. Hence, one has ωn=np\omega_{n}=n-p and in particular 1|ω|=|λ|1\in|\omega|=|\lambda|. Indeed, |ω||ωnp+1(np+1)|=|np(np+1)|=1|\omega|\ni|\omega_{n-p+1}-(n-p+1)|=|n-p-(n-p+1)|=1. Since λ\lambda is anti-dominant, we obtain λn=n+1\lambda_{n}=n+1. This completes the proof. ∎

For a 𝔨n\mathfrak{k}_{n}-dominant integral weight λ\lambda, we put

𝒪λunit={μ𝒪λμ is 𝔨n-dominant and L(μ) is unitarizable}.\mathcal{O}_{\lambda}^{\mathrm{unit}}=\{\mu\in\mathcal{O}_{\lambda}\mid\text{$\mu$ is $\mathfrak{k}_{n}$-dominant and $L(\mu)$ is unitarizable}\}.

By the proof of the above lemma, we obtain the following corollary:

Corollary 3.3.2.

Let λ\lambda be a regular anti-dominant integral weight.

  1. (1)

    If λn>n+1\lambda_{n}>n+1, one has 𝒪λunit={λ}\mathcal{O}_{\lambda}^{\mathrm{unit}}=\{\lambda\}.

  2. (2)

    If λn=n+1\lambda_{n}=n+1, one has 𝒪λunit={λ(0),,λ(p(λ))}\mathcal{O}_{\lambda}^{\mathrm{unit}}=\{\lambda^{(0)},\ldots,\lambda^{(p(\lambda))}\}, where

    λ(j)=(λ1,,λnj,nj,,nj).\lambda^{(j)}=(\lambda_{1},\ldots,\lambda_{n-j},n-j,\ldots,n-j).
Proof.

If #𝒪λunit>1\#\mathcal{O}_{\lambda}^{\mathrm{unit}}>1, one has λn=n+1\lambda_{n}=n+1 by Lemma 3.3.1. We may assume λn=n+1\lambda_{n}=n+1. In this case, the representation L(λ())L(\lambda^{(\ell)}) is unitary for any 1p(λ)1\leq\ell\leq p(\lambda). Thus, {λ(0),,λp(λ)}𝒪λunit\{\lambda^{(0)},\ldots,\lambda^{p(\lambda)}\}\subset\mathcal{O}_{\lambda}^{\mathrm{unit}}. We prove the converse. Take μ=(μ1,,μn)𝒪λunit\mu=(\mu_{1},\ldots,\mu_{n})\in\mathcal{O}_{\lambda}^{\mathrm{unit}}. By the proof of Lemma 3.3.1, we obtain μn=np(μ)\mu_{n}=n-p(\mu). Since λ\lambda is regular, the multiset |λ||\lambda| is a set. Note that {|λ11|,,|λnn|}={|μ11|,,|μnn|}\{|\lambda_{1}-1|,\ldots,|\lambda_{n}-n|\}=\{|\mu_{1}-1|,\ldots,|\mu_{n}-n|\}. By the 𝔨n\mathfrak{k}_{n}-dominance of λ\lambda and μ\mu, we obtain λii=μii\lambda_{i}-i=\mu_{i}-i for 1inp(μ)1\leq i\leq n-p(\mu). Thus, one has λ(p(μ))=μ\lambda^{(p(\mu))}=\mu. This completes the proof. ∎

3.4. Multiplicities of certain KK-types

In this subsection, we distinguish L(λ)L(\lambda) in terms of Kn,K_{n,\infty}-types in the orbit 𝒪λunit\mathcal{O}_{\lambda}^{\mathrm{unit}}. For this, we first recall the embeddings of highest weight modules into principal series representations.

Theorem 3.4.1 ([Yam89]).

For a principal series representation

IndBn()Gn()(μ1||s1μn||sn)\mathrm{Ind}_{B_{n}(\mathbb{R})}^{G_{n}(\mathbb{R})}\left(\mu_{1}|\cdot|^{s_{1}}\boxtimes\cdots\boxtimes\mu_{n}|\cdot|^{s_{n}}\right)

with unitary characters μi\mu_{i} of ×\mathbb{R}^{\times}, the induced representation contains a highest weight representation of weight (λ1,λn)(\lambda_{1},\ldots\lambda_{n}) if and only if we have

si=λni+1n+i1,μi=sgnλni+1s_{i}=\lambda_{n-i+1}-n+i-1,\mu_{i}=\mathrm{sgn}^{\lambda_{n-i+1}}

for any 1in1\leq i\leq n.

For 0jn0\leq j\leq n, let j\wedge^{j} be the jj-th exterior product of the standard representation of 𝔨n\mathfrak{k}_{n}. This is an irreducible representation of 𝔨n\mathfrak{k}_{n} with highest weight (1,,1j,0,,0)(\overbrace{1,\ldots,1}^{j},0,\ldots,0). Put

j(λ)=#{λ1λmod2}.j(\lambda)=\#\{\ell\mid\lambda_{1}\equiv\lambda_{\ell}\mod 2\}.

The following statement follows from the Littlewood-Richardson rule.

Lemma 3.4.2.

For a 𝔨n\mathfrak{k}_{n}-dominant integral weight λ\lambda, one has

Hom𝔨n(j(λ)detλ11,N(λ)|𝔨n)0.\mathrm{Hom}_{\mathfrak{k}_{n}}(\wedge^{j(\lambda)}\otimes\mathrm{det}^{\lambda_{1}-1},N(\lambda)|_{\mathfrak{k}_{n}})\neq 0.
Proof.

For an integral weight ω=(ω1,,ωn)\omega=(\omega_{1},\ldots,\omega_{n}), we consider the following two step operation:

  • Step 1.

    Put ω1=ω1\omega^{\prime}_{1}=\omega_{1}. For 2in2\leq i\leq n, set

    ωi={ωi1if ωi1ωi is evenωi11if ωi1ωi is odd.\omega_{i}^{\prime}=\begin{cases}\omega_{i-1}&\text{if $\omega_{i-1}-\omega_{i}$ is even}\\ \omega_{i-1}-1&\text{if $\omega_{i-1}-\omega_{i}$ is odd}.\end{cases}
  • Step 2.

    Consider the set X=X(ω)={i2in,ωi1ωi}X=X(\omega)=\{i\mid 2\leq i\leq n,\omega_{i-1}\neq\omega_{i}^{\prime}\}. Let aa be the maximal element in XX. We define a new set X=X(ω)X^{\prime}=X^{\prime}(\omega) by X=XX^{\prime}=X if #X\#X is even and by X=X{a}X^{\prime}=X\setminus\{a\} if #X\#X is odd. Put

    {ωi′′=ωiif iXωi′′=ωi+1if iX.\begin{cases}\omega_{i}^{\prime\prime}=\omega_{i}^{\prime}&\text{if $i\not\in X^{\prime}$}\\ \omega_{i}^{\prime\prime}=\omega_{i}^{\prime}+1&\text{if $i\in X^{\prime}$}.\end{cases}

We define a map g:nng\colon\mathbb{Z}^{n}\longrightarrow\mathbb{Z}^{n} by g((ω1,,ωn))=(ω1′′,,ωn′′)g((\omega_{1},\ldots,\omega_{n}))=(\omega_{1}^{\prime\prime},\ldots,\omega_{n}^{\prime\prime}). Note that the image of 𝔨n\mathfrak{k}_{n}-dominant weight is 𝔨n\mathfrak{k}_{n}-dominant. We denote by gg^{\ell} the \ell-th composite of gg. Set g(λ)=(λ1,,,λn,)g^{\ell}(\lambda)=(\lambda_{1,\ell},\ldots,\lambda_{n,\ell}) and a=i=1n(λi,λi,+1)a_{\ell}=\sum_{i=1}^{n}(\lambda_{i,\ell}-\lambda_{i,\ell+1}). Then, by definition, ai2a_{i}\in 2\mathbb{Z}. By the well-known correspondence of young diagrams and irreducible finite-dimensional representations of 𝔨n\mathfrak{k}_{n}, one can show that (a1,,an)(a_{1},\ldots,a_{n}) is 𝔨n\mathfrak{k}_{n}-dominant. By the definition of gg and the Littlewood-Richardson rule, the irreducible representation of 𝔨n\mathfrak{k}_{n} with highest weight gn1(λ)g^{n-1}(\lambda) occurs in the tensor product representation ρλρ(a1,,an)\rho_{\lambda}\otimes\rho_{(a_{1},\ldots,a_{n})} of 𝔨n\mathfrak{k}_{n}.

We next compute the weight gn1(λ)=(λ1,n1,,λn,n1)g^{n-1}(\lambda)=(\lambda_{1,n-1},\ldots,\lambda_{n,n-1}). By the construction, gn1(λ)g^{n-1}(\lambda) is of the form (λ1,,λ1,λ11,,λ11)(\lambda_{1},\ldots,\lambda_{1},\lambda_{1}-1,\ldots,\lambda_{1}-1). Indeed, by induction on \ell, one has λ1λ1+,1\lambda_{1}-\lambda_{1+\ell,\ell}\leq 1 for any 11\leq\ell. Thus, λ1λn,n11\lambda_{1}-\lambda_{n,n-1}\leq 1. We claim j(ω)=j(g(ω))j(\omega)=j(g(\omega)) for any 𝔨n\mathfrak{k}_{n}-dominant weight ω=(ω1,,ωn)\omega=(\omega_{1},\ldots,\omega_{n}). Set g(ω)=(ω1′′,,ωn′′)g(\omega)=(\omega_{1}^{\prime\prime},\ldots,\omega_{n}^{\prime\prime}). We write X(ω)={x1,,x2m}X^{\prime}(\omega)=\{x_{1},\ldots,x_{2m}\} with x1<x2<<x2mx_{1}<x_{2}<\cdots<x_{2m}. Then, for any xixi+1x_{i}\leq\ell\leq x_{i+1} with 0i2m10\leq i\leq 2m-1, one has ω1ω+(1+(1)i+1)/2\omega_{1}\equiv\omega_{\ell}+(1+(-1)^{i+1})/2 mod 22. Here, x0=1x_{0}=1. In particular, for any 1m1\leq\ell\leq m, we have ωx21ωx2\omega_{x_{2\ell-1}}\not\equiv\omega_{x_{2\ell}} mod 22. By ωxωx+1\omega_{x}\equiv\omega_{x}^{\prime}+1 mod 22 for any xX(ω)x\in X^{\prime}(\omega), we have j(ω)=j(g(ω))j(\omega)=j(g(\omega)). Hence we obtain

gn1(λ)=(λ1,,λ1j(λ),λ11,,λ11).g^{n-1}(\lambda)=(\overbrace{\lambda_{1},\ldots,\lambda_{1}}^{j(\lambda)},\lambda_{1}-1,\ldots,\lambda_{1}-1).

By the claim, we see that ρgn1(λ)=j(λ)detλ11\rho_{g^{n-1}(\lambda)}=\wedge^{j(\lambda)}\otimes\det^{\lambda_{1}-1} occurs in ρλρ(a1,,an)\rho_{\lambda}\otimes\rho_{(a_{1},\ldots,a_{n})}. The space 𝒰(𝔭n,+)\mathcal{U}(\mathfrak{p}_{n,+}) decomposes as

𝒰(𝔭n,+)=(b1,,bn)(2)n,b1bnρ(b1,,bn)\mathcal{U}(\mathfrak{p}_{n,+})=\bigoplus_{(b_{1},\ldots,b_{n})\in(2\mathbb{Z})^{n},b_{1}\geq\cdots\geq b_{n}}\rho_{(b_{1},\ldots,b_{n})}

as a representation of 𝔨n\mathfrak{k}_{n}. Thus, ρ(a1,,an)\rho_{(a_{1},\ldots,a_{n})} occurs in 𝒰(𝔭n,+)\mathcal{U}(\mathfrak{p}_{n,+}). Note that the restriction of N(λ)N(\lambda) to 𝔨n\mathfrak{k}_{n} is semisimple and N(λ)|𝔨n=𝒰(𝔭n,+)|𝔨nρλN(\lambda)|_{\mathfrak{k}_{n}}=\mathcal{U}(\mathfrak{p}_{n,+})|_{\mathfrak{k}_{n}}\otimes_{\mathbb{C}}\rho_{\lambda}. We then have

Hom𝔨n(j(λ)detλ11,N(λ)|𝔨n)0.\mathrm{Hom}_{\mathfrak{k}_{n}}(\wedge^{j(\lambda)}\otimes\mathrm{det}^{\lambda_{1}-1},N(\lambda)|_{\mathfrak{k}_{n}})\neq 0.

This completes the proof. ∎

Since a weight of \wedge^{\ell} is a permutation of (1,,1j,0,,0)(\overbrace{1,\ldots,1}^{j},0,\ldots,0), one has the following:

Proposition 3.4.3.

For a regular anti-dominant integral weight λ\lambda and 1jp(λ)1\leq j\leq p(\lambda), one has

dimHom𝔨n(j(λ)detλ11,L(λ)|𝔨n)=1\dim_{\mathbb{C}}\mathrm{Hom}_{\mathfrak{k}_{n}}(\wedge^{j(\lambda)}\otimes\mathrm{det}^{\lambda_{1}-1},L(\lambda)|_{\mathfrak{k}_{n}})=1

and

Hom𝔨n(j(λ)detλ11,L(λ(j))|𝔨n)=0.\mathrm{Hom}_{\mathfrak{k}_{n}}(\wedge^{j(\lambda)}\otimes\mathrm{det}^{\lambda_{1}-1},L(\lambda^{(j)})|_{\mathfrak{k}_{n}})=0.
Proof.

Let In(μ1,,μn)I_{n}(\mu_{1},\ldots,\mu_{n}) be the principal series representation

IndBn()Gn()(μ1μn).\mathrm{Ind}_{B_{n}(\mathbb{R})}^{G_{n}(\mathbb{R})}(\mu_{1}\boxtimes\cdots\boxtimes\mu_{n}).

Here, μi\mu_{i} are real valued characters of ×\mathbb{R}^{\times}. Take εj{0,1}\varepsilon_{j}\in\{0,1\} such that μj(1)=(1)εj\mu_{j}(-1)=(-1)^{\varepsilon_{j}}. Through the weight structure of j\wedge^{j}, one can find that the Hom space

Hom𝔨n(j,In(μ1,,μn)|𝔨n)\mathrm{Hom}_{\mathfrak{k}_{n}}(\wedge^{j},I_{n}(\mu_{1},\ldots,\mu_{n})|_{\mathfrak{k}_{n}})

is non-zero if and only if =1nε=j\sum_{\ell=1}^{n}\varepsilon_{\ell}=j by the Frobenius reciprocity. By the Frobenius reciprocity, one has the multiplicity free, i.e., dimHom𝔨n(j,In(μ1,,μn)|𝔨n)1\dim_{\mathbb{C}}\mathrm{Hom}_{\mathfrak{k}_{n}}(\wedge^{j},I_{n}(\mu_{1},\ldots,\mu_{n})|_{\mathfrak{k}_{n}})\leq 1.

For any ω𝒪λunit\omega\in\mathcal{O}^{\mathrm{unit}}_{\lambda}, the highest weight module L(ω)L(\omega) occurs in constituents of the induced representation In(sgnλn||λnn,,sgnλ1||λ11)I_{n}(\mathrm{sgn}^{\lambda_{n}}|\cdot|^{\lambda_{n}-n},\ldots,\mathrm{sgn}^{\lambda_{1}}|\cdot|^{\lambda_{1}-1}) by Theorem 3.4.1. Then, the statement follows from Lemma 3.4.2 and the above multiplicity free. This completes the proof. ∎

For a 𝔨n\mathfrak{k}_{n}-type σ\sigma, put

𝒪λunit(σ)={π𝒪λunitHom𝔨n(σ,π|𝔨n)0}.\mathcal{O}_{\lambda}^{\mathrm{unit}}(\sigma)=\{\pi\in\mathcal{O}_{\lambda}^{\mathrm{unit}}\mid\mathrm{Hom}_{\mathfrak{k}_{n}}(\sigma,\pi|_{\mathfrak{k}_{n}})\neq 0\}.
Corollary 3.4.4.

For a regular anti-dominant integral weight λ\lambda, one has

𝒪λunit(detλ11j(λ))={L(λ)}.\mathcal{O}_{\lambda}^{\mathrm{unit}}(\mathrm{det}^{\lambda_{1}-1}\otimes\wedge^{j(\lambda)})=\{L(\lambda)\}.
Proof.

The statement follows immediately from Corollary 3.3.2 and Proposition 3.4.3. ∎

3.5. Extensions of certain modules

Fix an odd integer ii. Let λ=(λ1,,λn)\lambda=(\lambda_{1},\ldots,\lambda_{n}) be a 𝔨n\mathfrak{k}_{n}-dominant integral weight such that λni+1==λn=n(i3)/2\lambda_{n-i+1}=\cdots=\lambda_{n}=n-(i-3)/2. Put λ=(λ1,,λn)=(λ1,,λni,n(i+1)/2,,n(i+1)/2)\lambda^{\prime}=(\lambda_{1}^{\prime},\ldots,\lambda_{n}^{\prime})=(\lambda_{1},\ldots,\lambda_{n-i},n-(i+1)/2,\ldots,n-(i+1)/2). By |λ|=|λ||\lambda|=|\lambda^{\prime}| as the multisets, the weights λ\lambda and λ\lambda^{\prime} have the same dot-orbit.

Lemma 3.5.1.

One has

dimExt𝒪𝔭(L(λ),L(λ))=1.\dim_{\mathbb{C}}\mathrm{Ext}_{\mathcal{O}^{\mathfrak{p}}}(L(\lambda^{\prime}),L(\lambda))=1.

Moreover an indecomposable module MM with a non-trivial exact sequence

0L(λ)ML(λ)00\longrightarrow L(\lambda)\longrightarrow M\longrightarrow L(\lambda^{\prime})\longrightarrow 0

is isomorphic to N(λ)N(\lambda^{\prime}).

Proof.

Set λ′′=λ+((i+1)/2,,(i+1)/2)\lambda^{\prime\prime}=\lambda^{\prime}+((i+1)/2,\ldots,(i+1)/2). Then, λ′′\lambda^{\prime\prime} satisfies the condition as in Theorem 3.2.1, i.e., the nn-th entry of λ′′\lambda^{\prime\prime} is nn. By p(λ′′)=ip(\lambda^{\prime\prime})=i and q(λ′′)=0q(\lambda^{\prime\prime})=0, the weight λ\lambda^{\prime} corresponds to the first reduction point r0(λ′′)r_{0}(\lambda^{\prime\prime}). Thus, the Verma module N(λ)N(\lambda^{\prime}) is reducible and N(λ)N(\lambda) is irreducible. Let ω=(ω1,,ωn)\omega=(\omega_{1},\ldots,\omega_{n}) be a 𝔨n\mathfrak{k}_{n}-dominant integral weight such that L(ω)L(\omega) is a constituent of N(λ)N(\lambda^{\prime}). Then, there exists wWnw\in W_{n} such that ω=wλ\omega=w\cdot\lambda^{\prime} and ωλ\omega\leq\lambda^{\prime}. We then have λjωj\lambda^{\prime}_{j}\leq\omega_{j} for any jj. The multiplicity of |λjj||\lambda_{j}-j| in the multiset |λ||\lambda| is one if and only if 1jni+21\leq j\leq n-i+2 or j=n(i3)/2j=n-(i-3)/2. Thus, ω\omega satisfies the following conditions:

  • ωj=λj\omega_{j}=\lambda_{j} for 1jni1\leq j\leq n-i.

  • For jni+3j\geq n-i+3, there exist kk and \ell such that ωkk=ω=|λjj|\omega_{k}-k=\ell-\omega_{\ell}=|\lambda_{j}-j|.

  • For j=ni+1,ni+2j=n-i+1,n-i+2, there exists kk such that |ωkk|=|λjj||\omega_{k}-k|=|\lambda_{j}-j|.

Indeed, it suffices to check the first condition. Suppose that there exist jnij\leq n-i and kk such that ωkk=(λjj)\omega_{k}-k=-(\lambda_{j}-j). Then, ωk=k+jλjk+ni(n(i3)/2)=k(i+3)/2<n(i+1)/2λn\omega_{k}=k+j-\lambda_{j}\leq k+n-i-(n-(i-3)/2)=k-(i+3)/2<n-(i+1)/2\leq\lambda_{n}^{\prime}. This contradicts to λkωk\lambda_{k}^{\prime}\leq\omega_{k}. We note that λ\lambda is a minimal element in 𝒪λunit\mathcal{O}_{\lambda}^{\mathrm{unit}}. The candidates of ω\omega are

λ,λ,(λ1,,λni,n(i1)/2,n(i+1)/2,,n(i+1)/2)\lambda,\qquad\lambda^{\prime},\qquad(\lambda_{1},\ldots,\lambda_{n-i},n-(i-1)/2,n-(i+1)/2,\ldots,n-(i+1)/2)

and

(λ1,,λni,n(i+1)/2,,n(i+1)/2,n(i1)/2).(\lambda_{1},\ldots,\lambda_{n-i},n-(i+1)/2,\ldots,n-(i+1)/2,n-(i-1)/2).

Since L(ω)L(\omega) occurs in a constituent of N(λ)N(\lambda^{\prime}), the representation ρω\rho_{\omega} occur in the restriction of N(λ)N(\lambda^{\prime}) to 𝔨n\mathfrak{k}_{n}. However, the last two candidates of ω\omega do not occur in the restriction N(λ)|𝔨nN(\lambda)|_{\mathfrak{k}_{n}}. Thus, any constituent of N(λ)N(\lambda) is of the form L(λ)L(\lambda) and L(λ)L(\lambda^{\prime}). The irreducible representations ρλ\rho_{\lambda} and ρλ\rho_{\lambda^{\prime}} of 𝔨n\mathfrak{k}_{n} have the multiplicity one in N(λ)N(\lambda^{\prime}). Hence the multiplicities of L(λ)L(\lambda) and L(λ)L(\lambda^{\prime}) in the constituent of N(λ)N(\lambda^{\prime}) are at most one. Since N(λ)N(\lambda^{\prime}) is reducible and L(λ)L(\lambda^{\prime}) is the irreducible quotient of N(λ)N(\lambda^{\prime}), we obtain a non-split exact sequence

0L(λ)N(λ)L(λ)0.0\longrightarrow L(\lambda)\longrightarrow N(\lambda^{\prime})\longrightarrow L(\lambda^{\prime})\longrightarrow 0.

By applying Hom𝒪𝔭(,L(λ))\mathrm{Hom}_{\mathcal{O}^{\mathfrak{p}}}(\,\cdot\,,L(\lambda)) to this short exact sequence, we obtain the following long exact sequence

0Hom𝒪𝔭(L(λ),L(λ))Hom𝒪𝔭(N(λ),L(λ))Hom𝒪𝔭(L(λ),L(λ))\displaystyle 0\longrightarrow\mathrm{Hom}_{\mathcal{O}^{\mathfrak{p}}}(L(\lambda^{\prime}),L(\lambda))\longrightarrow\mathrm{Hom}_{\mathcal{O}^{\mathfrak{p}}}(N(\lambda^{\prime}),L(\lambda))\longrightarrow\mathrm{Hom}_{\mathcal{O}^{\mathfrak{p}}}(L(\lambda),L(\lambda))
Ext𝒪𝔭1(L(λ),L(λ))Ext𝒪𝔭1(N(λ),L(λ))Ext𝒪𝔭1(L(λ),L(λ)).\displaystyle\qquad\longrightarrow\mathrm{Ext}^{1}_{\mathcal{O}^{\mathfrak{p}}}(L(\lambda^{\prime}),L(\lambda))\longrightarrow\mathrm{Ext}^{1}_{\mathcal{O}^{\mathfrak{p}}}(N(\lambda^{\prime}),L(\lambda))\longrightarrow\mathrm{Ext}^{1}_{\mathcal{O}^{\mathfrak{p}}}(L(\lambda),L(\lambda))\longrightarrow\cdots.

By definition, we have Hom𝒪𝔭(N(λ),L(λ))=0\mathrm{Hom}_{\mathcal{O}^{\mathfrak{p}}}(N(\lambda^{\prime}),L(\lambda))=0. Consider an extension

0L(λ)MN(λ)0.0\longrightarrow L(\lambda)\longrightarrow M\longrightarrow N(\lambda^{\prime})\longrightarrow 0.

Let vv be a weight vector in MM of weight λ\lambda^{\prime} such that the image of vv in N(λ)N(\lambda^{\prime}) is a non-zero highest weight vector. Since the weight of vv is highest in MM, there exists a splitting N(λ)MN(\lambda^{\prime})\longrightarrow M by the universality of N(λ)N(\lambda^{\prime}). Thus, the short exact sequence splits, i.e., Ext𝒪𝔭1(N(λ),L(λ))=0\mathrm{Ext}^{1}_{\mathcal{O}^{\mathfrak{p}}}(N(\lambda^{\prime}),L(\lambda))=0. We then obtain Ext𝒪𝔭1(L(λ),L(λ))Hom𝒪𝔭(L(λ),L(λ))\mathrm{Ext}^{1}_{\mathcal{O}^{\mathfrak{p}}}(L(\lambda^{\prime}),L(\lambda))\cong\mathrm{Hom}_{\mathcal{O}^{\mathfrak{p}}}(L(\lambda),L(\lambda)). This is of dimension one. This completes the proof. ∎

4. Siegel Eisenstein series

In this section, we compute the cuspidal components and exponents of Siegel Eisenstein series via the Siegel-Weil formula. We also show the near holomorphy of certain Siegel Eisenstein series.

4.1. Siegel-Weil formula

Let mm be a positive even integer. Set s0=(mn1)/2s_{0}=(m-n-1)/2. Let VV be a mm-dimensional quadratic space over FF and 𝒮(V(𝔸F)n)\mathcal{S}(V(\mathbb{A}_{F})^{n}) the space of Schwartz functions on V(𝔸F)nV(\mathbb{A}_{F})^{n}. We denote by ω(V)=ωψ(V)=vωψ,v(Vv)\omega(V)=\omega_{\psi}(V)=\bigotimes_{v}\omega_{\psi,v}(V_{v}) the Weil representation of Gn(𝔸)×O(V)(𝔸F)G_{n}(\mathbb{A}_{\mathbb{Q}})\times\mathrm{O}(V)(\mathbb{A}_{F}) on 𝒮(V(𝔸F)n)\mathcal{S}(V(\mathbb{A}_{F})^{n}). Here VvV_{v} is the vv-completion of VV for a place vv of FF. For φ𝒮(V(𝔸F)n)\varphi\in\mathcal{S}(V(\mathbb{A}_{F})^{n}), set

θ(g,h;φ)=vV(F)nω(g)φ(vh),gGn(𝔸),hO(V)(𝔸F).\theta(g,h;\varphi)=\sum_{v\in V(F)^{n}}\omega(g)\varphi(v\cdot h),\qquad g\in G_{n}(\mathbb{A}_{\mathbb{Q}}),h\in\mathrm{O}(V)(\mathbb{A}_{F}).

The function θ\theta is called the theta function. Put

I(g,φ)=O(V)(F)\O(V)(𝔸F)θ(g,h;φ)𝑑h,gGn(𝔸).I(g,\varphi)=\int_{\mathrm{O}(V)(F)\backslash\mathrm{O}(V)(\mathbb{A}_{F})}\theta(g,h;\varphi)\,dh,\qquad g\in G_{n}(\mathbb{A}_{\mathbb{Q}}).

The following condition (W) is called the Weil’s convergence condition:

(W) {V is anisotropicmr>n+1,\displaystyle\begin{cases}\text{$V$ is anisotropic}\\ \text{$m-r>n+1$},\end{cases}

where rr is the Witt index of VV. If VV satisfies the condition (W), the theta integral I(,φ)I(\,\cdot\,,\varphi) converges absolutely.

For φ𝒮(V(𝔸F)n)\varphi\in\mathcal{S}(V(\mathbb{A}_{F})^{n}), set

fφ(g)=ωψ(g)φ(0).f_{\varphi}(g)=\omega_{\psi}(g)\varphi(0).

Let χV\chi_{V} be the quadratic character associated to VV. Then fφf_{\varphi} is an element of In(s0,χV)I_{n}(s_{0},\chi_{V}). We denote by fs,φf_{s,\varphi} the standard section of In(s,χV)I_{n}(s,\chi_{V}) such that

fs0,φ=fφ.f_{s_{0},\varphi}=f_{\varphi}.

For a standard section fsf_{s} of In(s,μ)I_{n}(s,\mu), put

E(g,s,f)=γPn()\Gn()fs(γg),E(g,s,φ)=E(g,s,fφ).E(g,s,f)=\sum_{\gamma\in P_{n}(\mathbb{Q})\backslash G_{n}(\mathbb{Q})}f_{s}(\gamma g),\qquad E(g,s,\varphi)=E(g,s,f_{\varphi}).

The Siegel-Weil formula states the relationship between E(g,s0,φ)E(g,s_{0},\varphi) and I(g,φ)I(g,\varphi). In this paper, we use the following Siegel-Weil formula due to Kudla-Rallis [KR88].

Theorem 4.1.1.

Suppose that VV satisfies the condition (W). One has

E(s0,φ)=cI(g,φ)E(s_{0},\varphi)=cI(g,\varphi)

and

c={1If m>n+12If mn+1.c=\begin{cases}1&\text{If $m>n+1$}\\ 2&\text{If $m\leq n+1$}.\end{cases}

4.2. The representation RnR_{n}

Let VvV_{v} be a mm-dimensional quadratic space over FvF_{v} with m20m\in 2\mathbb{Z}_{\geq 0}. The character χV\chi_{V} denotes the quadratic character associated to VV. The map φfφ\varphi\longmapsto f_{\varphi} induces a Sp2n(Fv)\mathrm{Sp}_{2n}(F_{v})-intertwining map

ωψ,v(Vv)In(s0,χ).\omega_{\psi,v}(V_{v})\longrightarrow I_{n}(s_{0},\chi).

We denote by Rn(Vv)R_{n}(V_{v}) the image of the intertwining map. Then, Rn(Vv)R_{n}(V_{v}) can be viewed as the O(Vv)(Fv)\mathrm{O}(V_{v})(F_{v})-coinvariants of ωψ,v(Vv)\omega_{\psi,v}(V_{v}).

Proposition 4.2.1.

With the above notation, we obtain the following:

  1. (1)

    For a non-archimedean place vv, if χ2=𝟏\chi^{2}=\mathbf{1} and s00s_{0}\geq 0, one has

    In,v(s0,χ)=Rn(V1)+Rn(V2).I_{n,v}(s_{0},\chi)=R_{n}(V_{1})+R_{n}(V_{2}).

    Here V1V_{1} and V2V_{2} are the mm-dimensional inequivalent quadratic spaces over FvF_{v} with χ=χV1=χV2\chi=\chi_{V_{1}}=\chi_{V_{2}}.

  2. (2)

    For an archimedean place vv of FF, let VV be a mm-dimensional quadratic space over Fv=F_{v}=\mathbb{R} with the signature (m,0)(m,0) or (m2,2)(m-2,2). If s0>0s_{0}>0, the representation Rn(Vv)R_{n}(V_{v}) contains L(m,,m)L(m,\ldots,m).

  3. (3)

    For v𝐚v\in\mathbf{a} and s0>0s_{0}>0, the space of 𝔭n,\mathfrak{p}_{n,-}-finite vectors in In,v(s0,χ)I_{n,v}(s_{0},\chi) forms

    {0if χ=sgnm+1L(m,,m)if χ=sgnm.\begin{dcases}0&\text{if $\chi=\mathrm{sgn}^{m+1}$}\\ L(m,\ldots,m)&\text{if $\chi=\mathrm{sgn}^{m}$}.\end{dcases}
Proof.

The statement (1) and (2) are proved in [KR94, Proposition 5.3] and [KR90, Proposition 2.1], respectively. For the last statement, see [Hor20b, Corollary 6.5]. ∎

For a quadratic space VV over FF and a place vv of FF, let VvV_{v} be the vv-completion of FF. Put

Rn(V)=vRn(Vv)R_{n}(V)=\bigotimes_{v}R_{n}(V_{v})

where vv runs over all places of FF.

4.3. Holomorphy of Siegel Eisenstein series for s0>1s_{0}>1 or FF\neq\mathbb{Q}

Let fs=vfv,sf_{s}=\bigotimes_{v}f_{v,s} be a standard section of In(s,μ)I_{n}(s,\mu). For a representation MM of 𝔤n\mathfrak{g}_{n}, we denote by M𝔭n,finM_{\mathfrak{p}_{n,-}{\rm\mathchar 45\relax fin}} the space of 𝔭n,\mathfrak{p}_{n,-}-finite vectors in MM. Put s0=(mn1)/2s_{0}=(m-n-1)/2 with non-negative even integer mm.

Lemma 4.3.1.

For s0>1s_{0}>1, the map

In(s0,μ)𝔭n,fin𝒜(Gn)I_{n}(s_{0},\mu)_{\mathfrak{p}_{n,-}{\rm\mathchar 45\relax fin}}\longrightarrow\mathcal{A}(G_{n})

defined by

fsE(,s0,f)f_{s}\longmapsto E(\,\cdot\,,s_{0},f)

is injective and intertwining under the action of Gn(𝔸)G_{n}(\mathbb{A}_{\mathbb{Q}}). If FF\neq\mathbb{Q} or μ2𝟏\mu^{2}\neq\mathbf{1}, the same statement holds for s0>0s_{0}>0.

Proof.

Suppose μ2𝟏\mu^{2}\neq\mathbf{1}. This case is clear by the holomorphy of intertwining operators as in [Ike92].

Next, we suppose μ2=𝟏\mu^{2}=\mathbf{1} and s0>1s_{0}>1. By Proposition 4.2.1 (3), we have

(4.3.1) In(s0,μ)𝔭n,fin=(v<In,v(s0,μv))L(m/2,,m/2).\displaystyle I_{n}(s_{0},\mu)_{\mathfrak{p}_{n,-}{\rm\mathchar 45\relax fin}}=\left(\mathop{\bigotimes}_{v<\infty}I_{n,v}(s_{0},\mu_{v})\right)\otimes L(m/2,\ldots,m/2).

We claim that the representation In(s0,μ)𝔭n,finI_{n}(s_{0},\mu)_{\mathfrak{p}_{n,-}{\rm\mathchar 45\relax fin}} is contained in VRn(V)\sum_{V}R_{n}(V) where VV runs through the mm-dimensional quadratic spaces over FF such that VV satisfies the Weil’s convergence condition (W) and χV=μ\chi_{V}=\mu. Take a function f=vfvf=\bigotimes_{v}f_{v} in In(s0,μ)𝔭n,finI_{n}(s_{0},\mu)_{\mathfrak{p}_{n,-}{\rm\mathchar 45\relax fin}}. We may assume that each local functions fvf_{v} lie in Rn(Vv)R_{n}(V_{v}) for some quadratic space VvV_{v} over FvF_{v} by Proposition 4.2.1 (1). Let εv\varepsilon_{v} be the Hasse invariant of VvV_{v}. We denote by V(a,b)V(a,b) the non-degenerate real quadratic space with the signature (a,b)(a,b). The Hasse invariants of V(m,0)V(m,0) and V(m2,2)V(m-2,2) are 11 and 1-1, respectively. Fix an archimedean place ww. Then, there exists the quadratic space W=vWvW=\bigotimes_{v}W_{v} over FF such that WvVvW_{v}\cong V_{v} for any non-archimedean place vv, WvV(m,0)W_{v}\cong V(m,0) for any archimedean place vwv\neq w and

Ww={V(m,0)if v<εv=1V(m2,2)if v<εv=1.W_{w}=\begin{cases}V(m,0)&\text{if $\prod_{v<\infty}\varepsilon_{v}=1$}\\ V(m-2,2)&\text{if $\prod_{v<\infty}\varepsilon_{v}=-1$}.\end{cases}

By Proposition 4.2.1 (2) and m>n+3m>n+3, the quadratic space WW satisfies the condition (W) and fRn(W)f\in R_{n}(W). Hence the claim holds. By the claim, the theta integral converges absolutely. This states that the theta integral is an intertwining map under the action of Gn(𝔸)G_{n}(\mathbb{A}_{\mathbb{Q}}). Hence we obtain the following diagram:

𝒮(W(𝔸F)n)𝒩(Gn)Rn(W)In(s0,χW)𝔭n,fin.\begin{CD}\mathcal{S}(W(\mathbb{A}_{F})^{n})@>{}>{}>\mathcal{N}(G_{n})\\ @V{}V{}V@A{}A{}A\\ R_{n}(W)@>{}>{}>I_{n}(s_{0},\chi_{W})_{\mathfrak{p}_{n,-}{\rm\mathchar 45\relax fin}}.\end{CD}

Here the upper horizontal line is given by φI(,φ)\varphi\longmapsto I(\,\cdot\,,\varphi), the left vertical line is the canonical surjective morphism and the right vertical line is given by fE(,s0,f)f\longmapsto E(\,\cdot\,,s_{0},f). By the definition of the theta integral, it factors through the O(W)(𝔸F)\mathrm{O}(W)(\mathbb{A}_{F})-coinvariants Rn(W)R_{n}(W) of ωψ\omega_{\psi}. By the Siegel-Weil formula Theorem 4.1.1, the diagram is commutative. Hence the right vertical map is intertwining under the action of Gn(𝔸)G_{n}(\mathbb{A}_{\mathbb{Q}}). For the injectivity, we consider the constant term of I(,φ)I(\,\cdot\,,\varphi) along PnP_{n}. By the straightforward computation, one has I(,φ)Pn(g)=ω(g)φ(0)I(\,\cdot\,,\varphi)_{P_{n}}(g)=\omega(g)\varphi(0). Thus the right vertical line is injective.

For the case FF\neq\mathbb{Q}, it suffices to show that the space of induced representations (4.3.1) are contained in VRn(V)\sum_{V}R_{n}(V) where VV runs over all quadratic spaces over FF with dimension mm such that VV satisfies the condition (W). The proof is similar. Take f=fvf=\bigotimes f_{v} in the induced representation (4.3.1). We may assume fvRn(Vv)f_{v}\in R_{n}(V_{v}) for any place vv. Let εv\varepsilon_{v} be the Hasse invariant of VvV_{v}. Fix an archimedean place ww. If v<εv=1\prod_{v<\infty}\varepsilon_{v}=1, we can find a positive definite quadratic space WW over FF such that fvRn(Wv)f_{v}\in R_{n}(W_{v}). If v<εv=1\prod_{v<\infty}\varepsilon_{v}=-1, we can find a quadratic space WW over FF such that WvVvW_{v}\cong V_{v} for any non-archimedean place vv, WvW_{v} is positive definite for any archimedean place vwv\neq w, WwW_{w} is of signature (m2,2)(m-2,2) and fRn(W)f\in R_{n}(W). Then, WW is anisotropic. We obtain the claim. This completes the proof. ∎

In the following of this section, we assume F=F=\mathbb{Q}.

4.4. Near holomorphy of Siegel Eisenstein series for s=1s=1

Proposition 4.4.1.

Let fs=vfv,sf_{s}=\bigotimes_{v}f_{v,s} be a standard section of In(s,μ)I_{n}(s,\mu) such that f1In(1,μ)𝔭n,finf_{1}\in I_{n}(1,\mu)_{\mathfrak{p}_{n,-}{\rm\mathchar 45\relax fin}}. Then the Eisenstein series E(,1,f)E(\,\cdot\,,1,f) is nearly holomorphic, i.e., there exists \ell such that 𝔭n,E(,1,f)=0\mathfrak{p}_{n,-}^{\ell}\cdot E(\,\cdot\,,1,f)=0.

Proof.

We may assume that there exists a (n+3)/2(n+3)/2-dimensional quadratic space WvW_{v} over FvF_{v} such that fv,1R(Wv)f_{v,1}\in R(W_{v}) for any place vv of FF. Here WvW_{v} is positive definite for the archimedean place vv. Let εv\varepsilon_{v} be the Hasse invariant of WvW_{v}. If vεv=1\prod_{v}\varepsilon_{v}=1, the corresponding Eisenstein series E(,1,f)E(\,\cdot\,,1,f) generates the representation vR(Wv)\bigotimes_{v}R(W_{v}), by the same method as in the proof of Lemma 4.3.1. Then, the archimedean component is a highest weight representation. In particular, E(,1,f)E(\,\cdot\,,1,f) is nearly holomorphic.

Suppose vεv=1\prod_{v}\varepsilon_{v}=-1. Let VV be the quadratic space over F=F=\mathbb{Q} with dimension (n+3)/2(n+3)/2 such that VvWvV_{v}\cong W_{v} for non-archimedean place vv. Then, for the archimedean place vv, we may assume that the quadratic space VvV_{v} has the signature (n+1,2)(n+1,2). We consider the constant term of E(,1,f)E(\,\cdot\,,1,f) along P1,nP_{1,n}. For (t,g)GL1(𝔸F)×Gn1(𝔸)(t,g)\in\mathrm{GL}_{1}(\mathbb{A}_{F})\times G_{n-1}(\mathbb{A}_{\mathbb{Q}}) and kKnk\in K_{n}, one has

(4.4.1) E((t,g)k,s,f)P1,n\displaystyle E((t,g)k,s,f)_{P_{1,n}} =μ(t)|t|s+(n+1)/2E(g,s+12,ιr(k)f)\displaystyle=\mu(t)|t|^{s+(n+1)/2}E\left(g,s+\frac{1}{2},\iota^{*}r(k)f\right)
+μ(t)1|t|s+(n+1)/2E(g,s12,ιU(s,μ)r(k)f)\displaystyle\qquad+\mu(t)^{-1}|t|^{-s+(n+1)/2}E\left(g,s-\frac{1}{2},\iota^{*}U(s,\mu)r(k)f\right)

where ι\iota is the embedding Gn1P1,nGnG_{n-1}\xhookrightarrow{\quad}P_{1,n}\xhookrightarrow{\quad}G_{n} and U(s,μ)U(s,\mu) is the intertwining integral defined by

U(s,μ)fs=U1(𝔸F)fs(w1ug)𝑑uU(s,\mu)f_{s}=\int_{U_{1}(\mathbb{A}_{F})}f_{s}(w_{1}ug)\,du

for

w1=(00100𝟏n1001000000𝟏n1),U1={u=(1xy00𝟏n100001000xt𝟏n1)|x𝔸Fn1,y𝔸F}.w_{1}=\begin{pmatrix}0&0&-1&0\\ 0&\mathbf{1}_{n-1}&0&0\\ 1&0&0&0\\ 0&0&0&\mathbf{1}_{n-1}\end{pmatrix},\qquad U_{1}=\left\{u=\begin{pmatrix}1&x&y&0\\ 0&\mathbf{1}_{n-1}&0&0\\ 0&0&1&0\\ 0&0&-{{}^{t}x}&\mathbf{1}_{n-1}\end{pmatrix}\,\middle|\,x\in\mathbb{A}_{F}^{n-1},y\in\mathbb{A}_{F}\right\}.

We denote by Uv(s,μ)fv,sU_{v}(s,\mu)f_{v,s} the local intertwining integral so that vUv(s,μ)fv,s=U(s,μ)fs\bigotimes_{v}U_{v}(s,\mu)f_{v,s}=U(s,\mu)f_{s}. Note that the local intertwining integral Uv(s,μ)U_{v}(s,\mu) converges absolutely for Re(s)0\mathrm{Re}(s)\gg 0. Moreover, it is holomorphic and non-zero for Re(s)>0\mathrm{Re}(s)>0. See [PSR87, pp. 91]. Let \infty be the archimedean place of \mathbb{Q}. Let FsF_{s} be the standard section of In1(s,μ)I_{n-1}(s,\mu) such that Fs0=ιU(s,μ)r(k)fs|s=s0F_{s_{0}^{\prime}}=\iota^{*}U(s,\mu)r(k)f_{s}|_{s=s_{0}^{\prime}} for some s0s_{0}^{\prime} with Re(s0)0\mathrm{Re}(s_{0}^{\prime})\gg 0. We claim that there exists a non-zero constant cc such that

Ress=1/2E(g,s,F)=cE(g,1/2,ιU(s,μ)r(k)f).\mathrm{Res}_{s=1/2}E(g,s,F)=cE(g,1/2,\iota^{*}U(s,\mu)r(k)f).

For a non-zero standard section hsh_{s} of I(s,μ)I_{\infty}(s,\mu) of weight kk, the integral U(s,μ)hs(1)U_{\infty}(s,\mu)h_{s}(1) is a non-zero multiple of

(4.4.2) Γ(s)Γ((s+k+(n+1)/2)/s)Γ((s+(n+1)/2k)/2)\displaystyle\frac{\Gamma(s)}{\Gamma((s+k+(n+1)/2)/s)\Gamma((s+(n+1)/2-k)/2)}

by [KR94, (1.22)] and [KR88, Lemma 4.6]. Substitute k=(n+3)/2k=(n+3)/2. Then, U(s,μ)hsU_{\infty}(s,\mu)h_{s} has a simple zero at s=1/2s=1/2. Hence the integral U(s,μ)fs,U_{\infty}(s,\mu)f_{s,\infty} has a simple zero at s=1s=1. Indeed, at s=s0s=s_{0}, fsf_{s} can be written as a sum of right translations of a non-zero function of weight (n+3)/2(n+3)/2. Put

Uv(s,μ)fs={ιUv(s,μ)fv,sif v is non-archimedean.Γ((s+n+4)/2)Γ((s1)/2)Γ(s)ιUv(s,μ)fv,sif v is archimedean.U^{*}_{v}(s,\mu)f_{s}=\begin{dcases}\iota^{*}U_{v}(s,\mu)f_{v,s}&\text{if $v$ is non-archimedean.}\\ \frac{\Gamma((s+n+4)/2)\Gamma((s-1)/2)}{\Gamma(s)}\iota^{*}U_{v}(s,\mu)f_{v,s}&\text{if $v$ is archimedean.}\end{dcases}

For an unramified place vv, by [KR94, (1.23)], one has

(4.4.3) ιU(s,μ)fv,s(1)=Lv(s+(n1)/2,μ)Lv(2s,μ2)Lv(s+(n+1)/2,μ)Lv(2s+n1,μ2)fv,s(1)\displaystyle\iota^{*}U(s,\mu)f^{\circ}_{v,s}(1)=\frac{L_{v}(s+(n-1)/2,\mu)L_{v}(2s,\mu^{2})}{L_{v}(s+(n+1)/2,\mu)L_{v}(2s+n-1,\mu^{2})}f_{v,-s}^{\circ}(1)

where fv,sf_{v,s}^{\circ} is the unramified section of In,v(s,μ)I_{n,v}(s,\mu). Thus, the meromorphic section U(s,μ)fsU^{*}(s,\mu)f_{s} is holomorphic for s=1s=1. By Lemma 4.3.1 and [KR94, Theorem 1.1], E(,s1/2,ιU(s,μ)r(k)f)E(\,\cdot\,,s-1/2,\iota^{*}U(s,\mu)r(k)f) has at most simple pole at s=1s=1. We then have

lims1E(g,s1/2,ιU(s,μ)f)\displaystyle\lim_{s\rightarrow 1}E(g,s-1/2,\iota^{*}U(s,\mu)f) =lims1Γ(s)(s1)Γ((s+n+4)/2)Γ((s1)/2)(s1)1E(g,s1/2,U(s,μ)f)\displaystyle=\lim_{s\rightarrow 1}\frac{\Gamma(s)(s-1)}{\Gamma((s+n+4)/2)\Gamma((s-1)/2)}(s-1)^{-1}E(g,s-1/2,U^{*}(s,\mu)f)
=c2Γ((n+5)/2)Ress=1E(g,s1/2,F)\displaystyle=\frac{c}{2\cdot\Gamma((n+5)/2)}\mathrm{Res}_{s=1}E(g,s-1/2,F)

with some non-zero constant cc. Hence the claim holds. Let V0V_{0} be the complementary space of VV in the sense of [KR94, pp. 34]. By [KR94, Corollary 6.3], the constant term of Ress=1E(g,s1/2,F)\mathrm{Res}_{s=1}E(g,s-1/2,F) along Pn1P_{n-1} lies in Rn1(V0)In1(1/2,μ)R_{n-1}(V_{0})\subset I_{n-1}(-1/2,\mu). Thus, the constant term of E(g,s,f)E(g,s,f) along the Borel subgroup is an element of weight kk in a direct sum of principal series representations. Comparing the scalar Kn,K_{n,\infty}-types of principal series representations and degenerate principal series representations, the constant term lies in

In(1,χV)In(1,χV)I_{n}(1,\chi_{V})\oplus I_{n}(-1,\chi_{V})

of weight (n+3)/2(n+3)/2. Note that the Kn,K_{n,\infty}-type with highest weight ((n+3)/2,,(n+3)/2)((n+3)/2,\ldots,(n+3)/2) occur in In,(1,μ)𝔭n,finI_{n,\infty}(-1,\mu)_{\mathfrak{p}_{n,-}{\rm\mathchar 45\relax fin}} by [Hor20b, Lemma 3.5]. We also note that E(,s,f)E(\,\cdot\,,s,f) concentrates on the Borel subgroup. Hence the Eisenstein series E(,1,f)E(\,\cdot\,,1,f) is nearly holomorphic. This completes the proof. ∎

Remark 4.4.2.

In the above proof, we use the formula of dn,v(s,)d_{n,v}(s,\ell) as in [KR88, Lemma 4.6]. We should note that there is a typo in this formula. The correct one is

dn,v(s,)=(1)nk2ns(2π)nπn(n1)/2Γn(s)Γn((s+ρn+)/2)Γn((s+ρn)/2).d_{n,v}(s,\ell)=(\sqrt{-1})^{nk}2^{-ns}(2\pi)^{n}\pi^{n(n-1)/2}\frac{\Gamma_{n}(s)}{\Gamma_{n}((s+\rho_{n}+\ell)/2)\Gamma_{n}((s+\rho_{n}-\ell)/2)}.

Indeed, by the straightforward computation, dn,v(s,)d_{n,v}(s,\ell) equals to a non-zero constant multiple of a confluent hypergeometric function ξ(1,0;(s+ρn+)/2,(s+ρn)/2)\xi(1,0;(s+\rho_{n}+\ell)/2,(s+\rho_{n}-\ell)/2). For the explicit formulas of ξ\xi, see [Shi82] and [Shi00, pp. 140].

4.5. Cuspidal components of Siegel Eisenstein series at s=0s=0

We recall the properties of Siegel Eisenstein series at s=0s=0. If the rank nn is odd and μ\mu is quadratic, one has

In(0,μ)=VRn(V)𝒞Rn(𝒞),I_{n}(0,\mu)=\bigoplus_{V}R_{n}(V)\oplus\bigoplus_{\mathcal{C}}R_{n}(\mathcal{C}),

where VV runs over all the quadratic spaces of dimension n+1n+1 over FF such that μ=χV\mu=\chi_{V} and 𝒞={Wv}v\mathcal{C}=\{W_{v}\}_{v} runs through all incoherent families such that μv=χWv\mu_{v}=\chi_{W_{v}} for any place vv of FF. For the definition of incoherent family, see [KR94, pp. 7]. By [KR94, Theorem 4.10], one can identify a certain subspace of automorphic forms as the space of Eisenstein series at s=0s=0 as follows:

Theorem 4.5.1.

The following statements hold.

  1. (1)

    For a quadratic space VV of dimension n+1n+1 over FF, one has

    dimHomGn(𝔸)(Rn(V),𝒜(Gn))=1.\dim\mathrm{Hom}_{G_{n}(\mathbb{A}_{\mathbb{Q}})}(R_{n}(V),\mathcal{A}(G_{n}))=1.

    Moreover, the normalized Eisenstein series at s=0s=0 gives the non-trivial intertwining map Rn(V)𝒜(Gn(𝔸))R_{n}(V)\longrightarrow\mathcal{A}(G_{n}(\mathbb{A}_{\mathbb{Q}})).

  2. (2)

    For an incoherent family 𝒞\mathcal{C}, one has

    dimHomGn(𝔸)(Rn(𝒞),𝒜(Gn))=0.\dim\mathrm{Hom}_{G_{n}(\mathbb{A}_{\mathbb{Q}})}(R_{n}(\mathcal{C}),\mathcal{A}(G_{n}))=0.

    Moreover, for a standard section fsf_{s} with f0Rn(𝒞)f_{0}\in R_{n}(\mathcal{C}), one has E(g,0,f)=0E(g,0,f)=0.

The following statement follows from the theorem immediately.

Corollary 4.5.2.

Let fsf_{s} be a standard section of In(s,μ)I_{n}(s,\mu). The candidates of real parts of non-zero cuspidal exponents of E(,0,f)E(\,\cdot\,,0,f) are only ((n1)/2,(n3)/2,,(1n)/2)((n-1)/2,(n-3)/2,\ldots,(1-n)/2).

Proof.

By Theorem 4.5.1, the constant term of Eisenstein series along BnB_{n} lies in the direct sum of induced representations of the form In(0,μ)I_{n}(0,\mu). The lemma then follows from E(,s,f)𝒜(Gn){B}E(\,\cdot\,,s,f)\in\mathcal{A}(G_{n})_{\{B\}} and the definition of cuspidal exponents. ∎

5. Pullback formula

In this section, we compute the pullback formulas of Siegel Eisenstein series. As an application, we show the holomorphy and non-vanishing of Klingen Eisenstein series.

5.1. The formal identity and meromorphic sections

For mnm\leq n, we define the embeddings ιm,n\iota_{m,n}^{\uparrow} and ιm,n\iota_{m,n}^{\downarrow} of GmG_{m} into GnG_{n} by

ιm,n((abcd))=(ab𝟏nmcd𝟏nm),ιm,n((abcd))=(𝟏nmab𝟏nmcd).\iota_{m,n}^{\uparrow}\left(\begin{pmatrix}a&b\\ c&d\end{pmatrix}\right)=\left(\begin{array}[]{cc|cc}a&&b&\\ &\mathbf{1}_{n-m}&&\\ \hline\cr c&&d&\\ &&&\mathbf{1}_{n-m}\end{array}\right),\qquad\iota_{m,n}^{\downarrow}\left(\begin{pmatrix}a&b\\ c&d\end{pmatrix}\right)=\left(\begin{array}[]{cc|cc}\mathbf{1}_{n-m}&&&\\ &a&&b\\ \hline\cr&&\mathbf{1}_{n-m}&\\ &c&&d\end{array}\right).

Put Gm=ιm,n(Gm)G_{m}^{\uparrow}=\iota_{m,n}^{\uparrow}(G_{m}) and Gm=ιm,n(Gm)G_{m}^{\downarrow}=\iota_{m,n}^{\downarrow}(G_{m}). Take n,r>0n,r\in\mathbb{Z}_{>0}. For gGn+r(𝔸)g\in G_{n+r}(\mathbb{A}_{\mathbb{Q}}) and hGn(𝔸)h\in G_{n}(\mathbb{A}_{\mathbb{Q}}), put

g×h=(agbgcgdg)×(ahbhchdh)=ιn+r,2n+r(g)ιn,2n+r(h)=(agbgahbhcgdgchdh)G2n+r(𝔸).g\times h=\begin{pmatrix}a_{g}&b_{g}\\ c_{g}&d_{g}\end{pmatrix}\times\begin{pmatrix}a_{h}&b_{h}\\ c_{h}&d_{h}\end{pmatrix}=\iota_{n+r,2n+r}^{\uparrow}(g)\cdot\iota_{n,2n+r}^{\downarrow}(h)=\left(\begin{array}[]{cc|cc}a_{g}&&b_{g}&\\ &a_{h}&&b_{h}\\ \hline\cr c_{g}&&d_{g}&\\ &c_{h}&&d_{h}\end{array}\right)\in G_{2n+r}(\mathbb{A}_{\mathbb{Q}}).

Set

H=Gn+r×GnG2n+r,g^=(0𝟏n𝟏n0)g(0𝟏n𝟏n0),gGn.H=G_{n+r}^{\uparrow}\times G_{n}^{\downarrow}\subset G_{2n+r},\qquad\widehat{g}=\begin{pmatrix}0&\mathbf{1}_{n}\\ \mathbf{1}_{n}&0\end{pmatrix}g\begin{pmatrix}0&\mathbf{1}_{n}\\ \mathbf{1}_{n}&0\end{pmatrix},\qquad g\in G_{n}.

Let fsf_{s} be a standard section of I2n+r(s,μ)I_{2n+r}(s,\mu). For a cusp form φ\varphi on Gn(𝔸)G_{n}(\mathbb{A}_{\mathbb{Q}}) and gGn+r(𝔸)g\in G_{n+r}(\mathbb{A}_{\mathbb{Q}}), we consider the zeta integral

E(g,s;f,φ)=Gn()\Gn(𝔸)E((g×h^),s,f)φ(h)𝑑h.E(g,s;f,\varphi)=\int_{G_{n}(\mathbb{Q})\backslash G_{n}(\mathbb{A}_{\mathbb{Q}})}E((g\times\widehat{h}),s,f)\varphi(h)\,dh.

Put

fj=(000𝟏j)Matn+r,n,τj=(𝟏n+r𝟏nfj𝟏n+rfjt𝟏n)f_{j}=\begin{pmatrix}0&0\\ 0&\mathbf{1}_{j}\end{pmatrix}\in\mathrm{Mat}_{n+r,n},\qquad\tau_{j}=\left(\begin{array}[]{cc|cc}\mathbf{1}_{n+r}&&&\\ &\mathbf{1}_{n}&&\\ \hline\cr&f_{j}&\mathbf{1}_{n+r}&\\ {{}^{t}f_{j}}&&&\mathbf{1}_{n}\end{array}\right)

for 0jn0\leq j\leq n. Note that for any gGj(𝔸)g\in G_{j}(\mathbb{A}_{\mathbb{Q}}) and hG2n+r(𝔸)h\in G_{2n+r}(\mathbb{A}_{\mathbb{Q}}), one has

fs(τj((𝟏2(n+rj)×g)×(𝟏2(nj)×g^))h)=fs(τjh).f_{s}\left(\tau_{j}((\mathbf{1}_{2(n+r-j)}\times g)\times(\mathbf{1}_{2(n-j)}\times\widehat{g}))h\right)=f_{s}(\tau_{j}h).

The following double coset decomposition is well-known. For example, see [Shi00, Lemma 24.1].

Lemma 5.1.1.

One has the decomposition

G2n+r()=0jnP2n+r()τjH().G_{2n+r}(\mathbb{Q})=\bigsqcup_{0\leq j\leq n}P_{2n+r}(\mathbb{Q})\tau_{j}H(\mathbb{Q}).

Moreover, P2n+r()τjH()=ξ,β,γP2n+r()τj(((𝟏2(n+rj)×ξ)×𝟏2n)(β×γ))P_{2n+r}(\mathbb{Q})\tau_{j}H(\mathbb{Q})=\coprod_{\xi,\beta,\gamma}P_{2n+r}(\mathbb{Q})\tau_{j}(((\mathbf{1}_{2(n+r-j)}\times\xi)\times\mathbf{1}_{2n})\cdot(\beta\times\gamma)), where ξ\xi runs over Gj()G_{j}(\mathbb{Q}), β\beta over Pn+rj,n+r()\Gn+r()P_{n+r-j,n+r}(\mathbb{Q})\backslash G_{n+r}(\mathbb{Q}), and γ\gamma over Pnj,n()\Gn()P_{n-j,n}(\mathbb{Q})\backslash G_{n}(\mathbb{Q}).

By the lemma, we compute the integral E(g,s;f,φ)E(g,s;f,\varphi) as follows:

Gn()\Gn(𝔸)E((g×h^),s,f)φ(h)𝑑h\displaystyle\int_{G_{n}(\mathbb{Q})\backslash G_{n}(\mathbb{A}_{\mathbb{Q}})}E((g\times\widehat{h}),s,f)\varphi(h)\,dh
=Gn()\Gn(𝔸)γP2n+r()\G2n+r()fs(γ(g×h^))φ(h)dh\displaystyle=\int_{G_{n}(\mathbb{Q})\backslash G_{n}(\mathbb{A}_{\mathbb{Q}})}\sum_{\gamma\in P_{2n+r}(\mathbb{Q})\backslash G_{2n+r}(\mathbb{Q})}f_{s}(\gamma(g\times\widehat{h}))\varphi(h)\,dh
=Gn()\Gn(𝔸)0jnγP2n+r(F)\P2n+r()τjH()fs(γ(g×h^))φ(h)dh\displaystyle=\int_{G_{n}(\mathbb{Q})\backslash G_{n}(\mathbb{A}_{\mathbb{Q}})}\sum_{0\leq j\leq n}\sum_{\gamma\in P_{2n+r}(F)\backslash P_{2n+r}(\mathbb{Q})\tau_{j}H(\mathbb{Q})}f_{s}(\gamma(g\times\widehat{h}))\varphi(h)\,dh
=0jnGn()\Gn(𝔸)ξGj(𝔸)βPn+rj,n+r()\Gn+r()γPnj,n()\Gn()\displaystyle=\sum_{0\leq j\leq n}\int_{G_{n}(\mathbb{Q})\backslash G_{n}(\mathbb{A}_{\mathbb{Q}})}\sum_{\xi\in G_{j}(\mathbb{A}_{\mathbb{Q}})}\sum_{\beta\in P_{n+r-j,n+r}(\mathbb{Q})\backslash G_{n+r}(\mathbb{Q})}\sum_{\gamma\in P_{n-j,n}(\mathbb{Q})\backslash G_{n}(\mathbb{Q})}
fs(τj((𝟏2(n+rj)×ξ)×𝟏2n)(βg×γh^))φ(h)dh\displaystyle\hskip 199.16928ptf_{s}(\tau_{j}((\mathbf{1}_{2(n+r-j)}\times\xi)\times\mathbf{1}_{2n})\cdot(\beta g\times\gamma\widehat{h}))\varphi(h)\,dh
=0jnξGj(𝔸)βPn+rj,n+r()\Gn+r()Gn()\Gn(𝔸)γPnj,n()\Gn()\displaystyle=\sum_{0\leq j\leq n}\sum_{\xi\in G_{j}(\mathbb{A}_{\mathbb{Q}})}\sum_{\beta\in P_{n+r-j,n+r}(\mathbb{Q})\backslash G_{n+r}(\mathbb{Q})}\int_{G_{n}(\mathbb{Q})\backslash G_{n}(\mathbb{A}_{\mathbb{Q}})}\sum_{\gamma\in P_{n-j,n}(\mathbb{Q})\backslash G_{n}(\mathbb{Q})}
fs(τj((𝟏2(n+rj)×ξ)×𝟏2n)(βg×γh^))φ(h)dh.\displaystyle\hskip 199.16928ptf_{s}(\tau_{j}((\mathbf{1}_{2(n+r-j)}\times\xi)\times\mathbf{1}_{2n})\cdot(\beta g\times\gamma\widehat{h}))\varphi(h)\,dh.

If j<nj<n, we claim that the integral

Gn()\Gn(𝔸)γPnj,n()\Gn()fs(τj((𝟏2(n+rj)×ξ)×𝟏2n)(βg×γh^))φ(h)dh\int_{G_{n}(\mathbb{Q})\backslash G_{n}(\mathbb{A}_{\mathbb{Q}})}\sum_{\gamma\in P_{n-j,n}(\mathbb{Q})\backslash G_{n}(\mathbb{Q})}f_{s}(\tau_{j}((\mathbf{1}_{2(n+r-j)}\times\xi)\times\mathbf{1}_{2n})\cdot(\beta g\times\gamma\widehat{h}))\varphi(h)\,dh

vanishes. Put Pnj,n={p^pPnj,n}P_{n-j,n}^{\prime}=\{\widehat{p}\mid p\in P_{n-j,n}\}. We write 𝟏2(n+rj)×ξ\mathbf{1}_{2(n+r-j)}\times\xi by ξ\xi for simplicity. Then, it equals to

Pnj,n()\Gn(𝔸)fs(τj(ξβg×h^))φ(h)𝑑h\displaystyle\int_{P_{n-j,n}^{\prime}(\mathbb{Q})\backslash G_{n}(\mathbb{A}_{\mathbb{Q}})}f_{s}(\tau_{j}(\xi\beta g\times\widehat{h}))\varphi(h)\,dh
=Pnj,n()NPnj,n(𝔸)\Gn(𝔸)NPnj,n()\NPnj,n(𝔸)fs(τj(ξβg×nh^))φ(nh)𝑑n𝑑h\displaystyle=\int_{P_{n-j,n}^{\prime}(\mathbb{Q})N_{P_{n-j,n}^{\prime}}(\mathbb{A}_{\mathbb{Q}})\backslash G_{n}(\mathbb{A}_{\mathbb{Q}})}\int_{N_{P^{\prime}_{n-j,n}}(\mathbb{Q})\backslash N_{P^{\prime}_{n-j,n}}(\mathbb{A}_{\mathbb{Q}})}f_{s}(\tau_{j}(\xi\beta g\times\widehat{nh}))\varphi(nh)\,dndh
=Pnj,n()NPnj,n(𝔸)\Gn(𝔸)NPnj,n()\NPnj,n(𝔸)fs(τj(ξβg×h^))φ(nh)𝑑n𝑑h\displaystyle=\int_{P_{n-j,n}^{\prime}(\mathbb{Q})N_{P_{n-j,n}^{\prime}}(\mathbb{A}_{\mathbb{Q}})\backslash G_{n}(\mathbb{A}_{\mathbb{Q}})}\int_{N_{P^{\prime}_{n-j,n}}(\mathbb{Q})\backslash N_{P^{\prime}_{n-j,n}}(\mathbb{A}_{\mathbb{Q}})}f_{s}(\tau_{j}(\xi\beta g\times\widehat{h}))\varphi(nh)\,dndh
=Pnj,n()NPnj,n(𝔸)\Gn(𝔸)fs(τj(ξβg×γh^))(NPnj,n()\NPnj,n(𝔸)φ(nh)𝑑n)𝑑h\displaystyle=\int_{P_{n-j,n}^{\prime}(\mathbb{Q})N_{P_{n-j,n}^{\prime}}(\mathbb{A}_{\mathbb{Q}})\backslash G_{n}(\mathbb{A}_{\mathbb{Q}})}f_{s}(\tau_{j}(\xi\beta g\times\gamma\widehat{h}))\left(\int_{N_{P^{\prime}_{n-j,n}}(\mathbb{Q})\backslash N_{P^{\prime}_{n-j,n}}(\mathbb{A}_{\mathbb{Q}})}\varphi(nh)\,dn\right)dh
=0.\displaystyle=0.

Hence, we obtain

E(g,s;f,φ)\displaystyle E(g,s;f,\varphi) =ξGn()βPr,n+r()\Gn+r()Gn()\Gn(𝔸)fs(τn(𝟏2(n+r)×ξ^)(βg×h^))φ(h)𝑑h\displaystyle=\sum_{\xi\in G_{n}(\mathbb{Q})}\sum_{\beta\in P_{r,n+r}(\mathbb{Q})\backslash G_{n+r}(\mathbb{Q})}\int_{G_{n}(\mathbb{Q})\backslash G_{n}(\mathbb{A}_{\mathbb{Q}})}f_{s}(\tau_{n}(\mathbf{1}_{2(n+r)}\times\widehat{\xi})\cdot(\beta g\times\widehat{h}))\varphi(h)\,dh
=βPr,n+r()\Gn+r()Gn()\Gn(𝔸)ξGn(𝔸)fs(τn(βg×ξh^))φ(h)dh\displaystyle=\sum_{\beta\in P_{r,n+r}(\mathbb{Q})\backslash G_{n+r}(\mathbb{Q})}\int_{G_{n}(\mathbb{Q})\backslash G_{n}(\mathbb{A}_{\mathbb{Q}})}\sum_{\xi\in G_{n}(\mathbb{A}_{\mathbb{Q}})}f_{s}(\tau_{n}(\beta g\times\widehat{\xi h}))\varphi(h)\,dh
=βPr,n+r()\Gn+r()Gn(𝔸)fs(τn(βg×h^))φ(h)𝑑h.\displaystyle=\sum_{\beta\in P_{r,n+r}(\mathbb{Q})\backslash G_{n+r}(\mathbb{Q})}\int_{G_{n}(\mathbb{A}_{\mathbb{Q}})}f_{s}(\tau_{n}(\beta g\times\widehat{h}))\varphi(h)\,dh.

Put

Z(g,s;f,φ)=Gn(𝔸)fs(τn(g×h^))φ(h)𝑑h,gGn+r(𝔸).Z(g,s;f,\varphi)=\int_{G_{n}(\mathbb{A}_{\mathbb{Q}})}f_{s}(\tau_{n}(g\times\widehat{h}))\varphi(h)\,dh,\qquad g\in G_{n+r}(\mathbb{A}_{\mathbb{Q}}).

We then have

E(g,s;f,φ)=γPr,n+r()\Gn+r()Z(γg,s;f,φ).E(g,s;f,\varphi)=\sum_{\gamma\in P_{r,n+r}(\mathbb{Q})\backslash G_{n+r}(\mathbb{Q})}Z(\gamma g,s;f,\varphi).
Lemma 5.1.2.

The integral Z(g,s;f,φ)Z(g,s;f,\varphi) converges absolutely for ss\in\mathbb{C} with Re(s)0\mathrm{Re}(s)\gg 0 and can be meromorphically continued to whole ss-plane.

Proof.

Since E(g,s,f)E(g,s,f) converges absolutely for ss with Re(s)0\mathrm{Re}(s)\gg 0, the integral also converges absolutely for such ss. When r=0r=0, the meromorphic continuation follows from the meromorphic continuation of E(g,s,f)E(g,s,f). In general, we write g=n(t,m)kg=n(t,m)k for nNPr,n+r(𝔸),(t,m)GLr(𝔸F)×Gn(𝔸)n\in N_{P_{r,n+r}}(\mathbb{A}_{\mathbb{Q}}),(t,m)\in\mathrm{GL}_{r}(\mathbb{A}_{F})\times G_{n}(\mathbb{A}_{\mathbb{Q}}) and kKn+rk\in K_{n+r}. Then, one has

Z(g,s;f,φ)\displaystyle Z(g,s;f,\varphi) =μ(t)|dett|s+(2n+r+1)/2Z(m,s;r(k)f,φ)\displaystyle=\mu(t)|\det t|^{s+(2n+r+1)/2}Z(m,s;r(k)f,\varphi)
=μ(t)|dett|s+(2n+r+1)/2Z(m,s+r/2;ι2n,2n+r,r(k)f,φ).\displaystyle=\mu(t)|\det t|^{s+(2n+r+1)/2}Z(m,s+r/2;\iota_{2n,2n+r}^{\downarrow,*}r(k)f,\varphi).

Thus, the meromorphic continuation follows from the case r=0r=0. ∎

The section Z(,s;f,φ)Z(\,\cdot\,,s;f,\varphi) is then a meromorphic section of

Ir,n+r(s,μ,𝒜cusp(Gn)).I_{r,n+r}\left(s,\mu,\mathcal{A}_{\mathrm{cusp}}(G_{n})\right).

Indeed, let PP be a parabolic subgroup of GnG_{n} with the unipotent radical NN. It suffices to prove that the constant term of Z(,s;f,φ)Z(\,\cdot\,,s;f,\varphi) along PP is zero. It equals to

Z(g,s;f,φ)P\displaystyle Z(g,s;f,\varphi)_{P} =N()\N(𝔸)Z(ng,s;f,φ)𝑑n\displaystyle=\int_{N(\mathbb{Q})\backslash N(\mathbb{A}_{\mathbb{Q}})}Z(ng,s;f,\varphi)\,dn
=N()\N(𝔸)Gn(𝔸)fs(τn(ng×h^))φ(h)𝑑h𝑑n\displaystyle=\int_{N(\mathbb{Q})\backslash N(\mathbb{A}_{\mathbb{Q}})}\int_{G_{n}(\mathbb{A}_{\mathbb{Q}})}f_{s}(\tau_{n}(ng\times\widehat{h}))\varphi(h)\,dhdn
=Gn(𝔸)N()\N(𝔸)fs(τn(g×n1h^))φ(h)𝑑n𝑑h\displaystyle=\int_{G_{n}(\mathbb{A}_{\mathbb{Q}})}\int_{N(\mathbb{Q})\backslash N(\mathbb{A}_{\mathbb{Q}})}f_{s}(\tau_{n}(g\times\widehat{n^{-1}h}))\varphi(h)\,dndh
=Gn(𝔸)fs(τn(g×h^))(N()\N(𝔸)φ(nh)𝑑n)𝑑h\displaystyle=\int_{G_{n}(\mathbb{A}_{\mathbb{Q}})}f_{s}(\tau_{n}(g\times\widehat{h}))\left(\int_{N(\mathbb{Q})\backslash N(\mathbb{A}_{\mathbb{Q}})}\varphi(nh)\,dn\right)dh
=0,\displaystyle=0,

by the cuspidality of φ\varphi. Take a cusp form ϕ\phi on Gn(𝔸)G_{n}(\mathbb{A}_{\mathbb{Q}}). For any kKn+rk\in K_{n+r}, one has

Z(k,s;f,φ),ϕ\displaystyle\langle Z(k,s;f,\varphi),\phi\rangle =Gn()\Gn(𝔸)Z((𝟏r×x)k,s;f,φ)ϕ(x)¯𝑑x\displaystyle=\int_{G_{n}(\mathbb{Q})\backslash G_{n}(\mathbb{A}_{\mathbb{Q}})}Z((\mathbf{1}_{r}\times x)k,s;f,\varphi)\overline{\phi(x)}\,dx
=Gn()\Gn(𝔸)Gn(𝔸)fs(τn(k×x1h^))φ(h)𝑑hϕ(x)¯𝑑x\displaystyle=\int_{G_{n}(\mathbb{Q})\backslash G_{n}(\mathbb{A}_{\mathbb{Q}})}\int_{G_{n}(\mathbb{A}_{\mathbb{Q}})}f_{s}(\tau_{n}(k\times\widehat{x^{-1}h}))\varphi(h)\,dh\,\overline{\phi(x)}\,dx
=Gn(𝔸)fs(τn(k×h^))(Gn()\Gn(𝔸)φ(xh)ϕ(x)¯𝑑x)𝑑h\displaystyle=\int_{G_{n}(\mathbb{A}_{\mathbb{Q}})}f_{s}(\tau_{n}(k\times\widehat{h}))\left(\int_{G_{n}(\mathbb{Q})\backslash G_{n}(\mathbb{A}_{\mathbb{Q}})}\varphi(xh)\overline{\phi(x)}\,dx\right)dh
=Gn(𝔸)fs(τn(k×h^))r(h)φ,ϕ𝑑h.\displaystyle=\int_{G_{n}(\mathbb{A}_{\mathbb{Q}})}f_{s}(\tau_{n}(k\times\widehat{h}))\langle r(h)\varphi,\phi\rangle\,dh.

The pairing Z(g,s;f,φ),ϕ\langle Z(g,s;f,\varphi),\phi\rangle is zero unless ϕ\phi lies in the πφ\pi_{\varphi}-isotypic component of 𝒜cusp(Gn)\mathcal{A}_{\mathrm{cusp}}(G_{n}). Here the representation πφ\pi_{\varphi} is the representation of Gn(𝔸)G_{n}(\mathbb{A}_{\mathbb{Q}}) generated by φ\varphi. For any kKn+rk\in K_{n+r}, the function mZ(mk,s;f,φ)m\longmapsto Z(mk,s;f,\varphi) on Gn(𝔸)G_{n}(\mathbb{A}_{\mathbb{Q}}) lies in the πφ\pi_{\varphi}-isotypic component. Hence, the section Z(,s;f,φ)Z(\,\cdot\,,s;f,\varphi) is a section of Ir,n+r(s,μ,πφ)I_{r,n+r}(s,\mu,\pi_{\varphi}).

Let π=vπv\pi=\bigotimes_{v}\pi_{v} be an irreducible cuspidal automorphic representation of Gn(𝔸)G_{n}(\mathbb{A}_{\mathbb{Q}}). By the above computations, we define a meromorphic section Z(,s;f,φ)Z(\,\cdot\,,s;f,\varphi) of Ir,n+r(s,μ,π)I_{r,n+r}(s,\mu,\pi) for φπ=vπv\varphi\in\pi=\bigotimes_{v}\pi_{v}. For fs=vfv,sf_{s}=\bigotimes_{v}f_{v,s} and φ=vφvvπv\varphi=\bigotimes_{v}\varphi_{v}\in\bigotimes_{v}\pi_{v}, set

Zv(g,s;fv,φv)=Sp2n(Fv)fv,s(τn(g×h^))πv(h)φv𝑑h.Z_{v}(g,s;f_{v},\varphi_{v})=\int_{\mathrm{Sp}_{2n}(F_{v})}f_{v,s}(\tau_{n}(g\times\widehat{h}))\pi_{v}(h)\varphi_{v}\,dh.

Then,

Z(g,s;f,φ)=vZv(g,s;fv,φv).Z(g,s;f,\varphi)=\prod_{v}Z_{v}(g,s;f_{v},\varphi_{v}).

In the following, we first consider the relationship between the constant terms of E(,s;f,φ)E(\,\cdot\,,s;f,\varphi) and the global section Z(,s;f,φ)Z(\,\cdot\,,s;f,\varphi). After that, we compute the local sections Z(,s;fv,φv)Z(\,\cdot\,,s;f_{v},\varphi_{v}).

5.2. Near holomorphy of Klingen Eisenstein series

We prove the near holomorphy of Eisenstein series E(,s0;f,φ)E(\,\cdot\,,s_{0};f,\varphi) on Gn+r(𝔸)G_{n+r}(\mathbb{A}_{\mathbb{Q}}) as follows:

Proposition 5.2.1.

Fix r,nr,n with 1rn1\leq r\leq n and s00s_{0}\geq 0 with s0+(2n+r+1)/2s_{0}\in\mathbb{Z}+(2n+r+1)/2. For a character μ\mu of GL2n+r(𝔸F)\mathrm{GL}_{2n+r}(\mathbb{A}_{F}), let fsf_{s} be a standard section of I2n+r(s,μ)I_{2n+r}(s,\mu). We assume

  • fs0f_{s_{0}} is 𝔭2n+r,\mathfrak{p}_{2n+r,-}-finite.

  • If F=F=\mathbb{Q} and s0=1/2s_{0}=1/2, there exists a quadratic space VV over FF with dimension (n+2)/2(n+2)/2 such that WW satisfies the condition (W) and fs0Rn(V)f_{s_{0}}\in R_{n}(V).

Then, for a cusp form φ\varphi on Gn(𝔸)G_{n}(\mathbb{A}_{\mathbb{Q}}), the Eisenstein series E(,s0;f,φ)E(\,\cdot\,,s_{0};f,\varphi) on Gn+r(𝔸)G_{n+r}(\mathbb{A}_{\mathbb{Q}}) is nearly holomorphic.

Proof.

Under the assumptions, Siegel Eisenstein series E(,s,f)E(\,\cdot\,,s,f) is nearly holomorphic at s=s0s=s_{0} by the proof of Lemma 4.3.1 and Proposition 4.4.1. Take an integer 0\ell\gg 0 so that 𝔭2n+r,E(,s0,f)=0\mathfrak{p}_{2n+r,-}^{\ell}\cdot E(\,\cdot\,,s_{0},f)=0. Since the integral

E(g,s0;f,φ)=Gn+r()\Gn+r(𝔸)E((g×h^),s0,f)φ(h)𝑑h,gGn(𝔸)E(g,s_{0};f,\varphi)=\int_{G_{n+r}(\mathbb{Q})\backslash G_{n+r}(\mathbb{A}_{\mathbb{Q}})}E((g\times\widehat{h}),s_{0},f)\varphi(h)\,dh,\qquad g\in G_{n}(\mathbb{A}_{\mathbb{Q}})

converges absolutely, one has 𝔭n+r,E(g,s0;f,φ)=0\mathfrak{p}_{n+r,-}^{\ell}\cdot E(g,s_{0};f,\varphi)=0. This completes the proof. ∎

We next compute the constant term of E(,s0;f,φ)E(\,\cdot\,,s_{0};f,\varphi) along Pr,n+rP_{r,n+r}. Let UU be the subgroup of G2n+rG_{2n+r} in which elements of the form

(𝟏r𝟏n𝟏n𝟏r𝟏n𝟏n).\left(\begin{array}[]{ccc|ccc}\mathbf{1}_{r}&&&*&&\\ &\mathbf{1}_{n}&&&&\\ &&\mathbf{1}_{n}&&&\\ \hline\cr&&&\mathbf{1}_{r}&&\\ &&&&\mathbf{1}_{n}&\\ &&&&&\mathbf{1}_{n}\end{array}\right).

We may regard the group UU as a subgroup of Gn+rG_{n+r}^{\uparrow}. Then, it is a subgroup of the unipotent radical of Pr,n+rP_{r,n+r}. Set

E(g,s0,f)U=U()\U(𝔸)E(ug,s0,f)𝑑n.E(g,s_{0},f)_{U}=\int_{U(\mathbb{Q})\backslash U(\mathbb{A}_{\mathbb{Q}})}E(ug,s_{0},f)\,dn.

We compute E(g,s0,f)UE(g,s_{0},f)_{U} as follows:

Lemma 5.2.2.

Let fsf_{s} be a standard section of I2n+r(s,μ)I_{2n+r}(s,\mu). Suppose that fsf_{s} satisfies the conditions as in Proposition 5.2.1 and moreover if F=F=\mathbb{Q}, assume s0>1s_{0}>1. We then have

E((t,m),s0,f)U=μ(t)|dett|s0+(2n+r+1)/2E(m,s0+r/2,ι2n,2n+r,f)E((t,m),s_{0},f)_{U}=\mu(t)|\det t|^{s_{0}+(2n+r+1)/2}E(m,s_{0}+r/2,\iota_{2n,2n+r}^{\downarrow,*}f)

for (t,m)GLr(𝔸F)×G2n(𝔸)=MPr,2n+r(𝔸)(t,m)\in\mathrm{GL}_{r}(\mathbb{A}_{F})\times G_{2n}(\mathbb{A}_{\mathbb{Q}})=M_{P_{r,2n+r}}(\mathbb{A}_{\mathbb{Q}}).

Proof.

By the near holomorphy of E(g,s0,f)E(g,s_{0},f) and [Hor20b, Lemma 5.10], we have

E(g,s0,f)U=E(g,s0,f)Qr,2n+r.E(g,s_{0},f)_{U}=E(g,s_{0},f)_{Q_{r,2n+r}}.

Thus, for (t,m)=(t1,,tr,m)GL1(𝔸F)××GL1(𝔸F)×G2n(𝔸)=MQr,2n+r(t,m)=(t_{1},\ldots,t_{r},m)\in\mathrm{GL}_{1}(\mathbb{A}_{F})\times\cdots\times\mathrm{GL}_{1}(\mathbb{A}_{F})\times G_{2n}^{\downarrow}(\mathbb{A}_{\mathbb{Q}})=M_{Q_{r,2n+r}}, by taking the constant terms successively, we obtain

E((t,m),s0,f)U=(((E((t,m),s0,f)Q1,2n+r|G2n+r1(𝔸))Q1,2n+r1|Gn+r2(𝔸))Q1,2n+1|G2n(𝔸).E((t,m),s_{0},f)_{U}=\left(\cdots\left(\left(E((t,m),s_{0},f\right)_{Q_{1,2n+r}}|_{G_{2n+r-1}^{\downarrow}(\mathbb{A}_{\mathbb{Q}})}\right)_{Q_{1,2n+r-1}}|_{G_{n+r-2}^{\downarrow}(\mathbb{A}_{\mathbb{Q}})}\cdots\right)_{Q_{1,2n+1}}|_{G_{2n}^{\downarrow}(\mathbb{A}_{\mathbb{Q}})}.

We tacitly assume r=1r=1. By (4.4.1), one has

E((t,m),s0,f)U=μ(t)|t|s0+n+(r+1)/2E(m,s0+1/2,ι,f)+μ(t)1|t|s0+n+1E(m,s01/2,ι,U(s,μ)f).E((t,m),s_{0},f)_{U}=\mu(t)|t|^{s_{0}+n+(r+1)/2}E(m,s_{0}+1/2,\iota^{\downarrow,*}f)+\mu(t)^{-1}|t|^{-s_{0}+n+1}E(m,s_{0}-1/2,\iota^{\downarrow,*}U(s,\mu)f).

Then, for s=s0s=s_{0} and v𝐚v\in\mathbf{a}, the archimedean component Uv(s,μ)fvU_{v}(s,\mu)f_{v} has at least simple zero. Hence, by assumptions, the Eisenstein series E(m,s1/2,ιU(s,μ)f)E(m,s-1/2,\iota^{*}U(s,\mu)f) is zero at s=s0s=s_{0}. For general rr, we thus have

E((t,m),s0,f)U=(j=1rμ(tj)|tj|s0+(2n+r+1)/2)E(m,s0+r/2,ι,f).E((t,m),s_{0},f)_{U}=\left(\prod_{j=1}^{r}\mu(t_{j})|t_{j}|^{s_{0}+(2n+r+1)/2}\right)E(m,s_{0}+r/2,\iota^{\downarrow,*}f).

Let SLr\mathrm{SL}_{r} be the derived subgroup of GLrMPr,2n+r\mathrm{GL}_{r}\subset M_{P_{r,2n+r}}. It suffices to show that E(,s0,f)E(\,\cdot\,,s_{0},f) is left SLr(𝔸F)\mathrm{SL}_{r}(\mathbb{A}_{F}) invariant. It follows from [Hor20b, Lemma 5.7] by the near holomorphy of Eisenstein series. This completes the proof. ∎

Proposition 5.2.3.

With the notation as in Proposition 5.2.1, suppose s0>1s_{0}>1 if F=F=\mathbb{Q}. Then, the constant term of E(g,s0;f,φ)E(g,s_{0};f,\varphi) along Pr,n+rP_{r,n+r} equals to the zeta integral Z(g,s0;f,φ)Z(g,s_{0};f,\varphi) for any gGn+r(𝔸)g\in G_{n+r}(\mathbb{A}_{\mathbb{Q}}).

Proof.

Since Kn+r,K_{n+r,\infty} normalizes 𝔭n+r,\mathfrak{p}_{n+r,-}, the right translation r(k)fsr(k)f_{s} satisfies the same condition as in Proposition 5.2.1. Thus, for any (t,m)GLr(𝔸)×Gn(𝔸)=MPr,n+r(𝔸)(t,m)\in\mathrm{GL}_{r}(\mathbb{A}_{\mathbb{Q}})\times G_{n}(\mathbb{A}_{\mathbb{Q}})=M_{P_{r,n+r}}(\mathbb{A}_{\mathbb{Q}}) and kKn+rk\in K_{n+r}, we have

E((t,m)k,s0;f,φ)Pr,n+r\displaystyle E((t,m)k,s_{0};f,\varphi)_{P_{r,n+r}} =U()\U(𝔸)Gn()\Gn(𝔸)E((u(t,m)×h^),s0,r(k)f)φ(h)𝑑h𝑑u\displaystyle=\int_{U(\mathbb{Q})\backslash U(\mathbb{A}_{\mathbb{Q}})}\int_{G_{n}(\mathbb{Q})\backslash G_{n}(\mathbb{A}_{\mathbb{Q}})}E((u(t,m)\times\widehat{h}),s_{0},r(k)f)\varphi(h)\,dhdu
=μ(t)|dett|s0+ρ2n+rGn()\Gn(𝔸)E((m×h^),s0+r/2,ι2n,2n+r,r(k)f)φ(h)𝑑h\displaystyle=\mu(t)|\det t|^{s_{0}+\rho_{2n+r}}\int_{G_{n}(\mathbb{Q})\backslash G_{n}(\mathbb{A}_{\mathbb{Q}})}E((m\times\widehat{h}),s_{0}+r/2,\iota_{2n,2n+r}^{\downarrow,*}r(k)f)\varphi(h)\,dh
=μ(t)|dett|s0+ρ2n+rZ(m,s0+r/2;ι2n,2n+r,r(k)f,φ)\displaystyle=\mu(t)|\det t|^{s_{0}+\rho_{2n+r}}Z(m,s_{0}+r/2;\iota_{2n,2n+r}^{\downarrow,*}r(k)f,\varphi)
=Z((t,m),s0;r(k)f,φ)\displaystyle=Z((t,m),s_{0};r(k)f,\varphi)
=Z((t,m)k,s0;f,φ).\displaystyle=Z((t,m)k,s_{0};f,\varphi).

For the first and second equality, we use Lemma 5.2.2. Hence we see E(g,s0;f,φ)Pr,n+r=Z(g,s0;f,φ)E(g,s_{0};f,\varphi)_{P_{r,n+r}}=Z(g,s_{0};f,\varphi). This completes the proof. ∎

Corollary 5.2.4.

With the notation as in Proposition 5.2.1, suppose s0>1s_{0}>1 if F=F=\mathbb{Q}. Then, the zeta integral Z(g,s;f,φ)Z(g,s;f,\varphi) is holomorphic at s=s0s=s_{0}.

Proof.

The statement follows immediately from the definition of zeta integral and the holomorphy of E(,s,f)E(\,\cdot\,,s,f) at s=s0s=s_{0}. ∎

We next consider the local zeta integrals Zv(,s;fv,φv)Z_{v}(\,\cdot\,,s;f_{v},\varphi_{v}).

5.3. Unramified computations

We first compute Zv(g,s;f,φ)Z_{v}(g,s;f,\varphi) at unramified places.

Lemma 5.3.1.

Let μv\mu_{v} be an unramified character of GL2n+r(Fv)\mathrm{GL}_{2n+r}(F_{v}), fv,sf_{v,s} be an unramified standard section of I2n+r,v(s,μv)I_{2n+r,v}(s,\mu_{v}) with fv,s(1)=1f_{v,s}(1)=1 and πv\pi_{v} be an irreducible unramified representation of Sp2n(Fv)\mathrm{Sp}_{2n}(F_{v}) with an invariant inner product ,\langle\,,\,\rangle. Take an unramified vector φvπv\varphi_{v}\in\pi_{v} so that φv,φv=1\langle\varphi_{v},\varphi_{v}\rangle=1. We then have

Z(1,s;fv,φv)=Lv(s+(r+1)/2,πv,μv)Lv(s+n+(r+1)/2,μv)j=1nLv(2s+2n+r+12j,μv2)×φv.Z(1,s;f_{v},\varphi_{v})=\frac{L_{v}(s+(r+1)/2,\pi_{v},\mu_{v})}{L_{v}(s+n+(r+1)/2,\mu_{v})\prod_{j=1}^{n}L_{v}(2s+2n+r+1-2j,\mu^{2}_{v})}\times\varphi_{v}.
Proof.

The restriction ι,fs+r/2\iota^{\downarrow,*}f_{s+r/2} of fv,sf_{v,s} to G2nG^{\downarrow}_{2n} is a standard unramified section of I2n(s+r/2,μ)I_{2n}(s+r/2,\mu). Since Z(1,s;fv,φv)Z(1,s;f_{v},\varphi_{v}) is an unramified vector, it is a constant multiple of φv\varphi_{v}. By definition of local zeta integral, we have

Zv(1,s;fv,φv),φv\displaystyle\langle Z_{v}(1,s;f_{v},\varphi_{v}),\varphi_{v}\rangle =Sp2n(Fv)fv,s(τn(1×h^))πv(h)φv,φv𝑑h\displaystyle=\int_{\mathrm{Sp}_{2n}(F_{v})}f_{v,s}(\tau_{n}(1\times\widehat{h}))\langle\pi_{v}(h)\varphi_{v},\varphi_{v}\rangle\,dh
=Sp2n(Fv)ι2n,2n+r,fv,s+r/2(τn(1×h^))πv(h)φv,φv𝑑h.\displaystyle=\int_{\mathrm{Sp}_{2n}(F_{v})}\iota_{2n,2n+r}^{\downarrow,*}f_{v,s+r/2}(\tau_{n}(1\times\widehat{h}))\langle\pi_{v}(h)\varphi_{v},\varphi_{v}\rangle\,dh.

By [KR94, (7.2.8)], one has

Z(1,s;fv,φv)=Lv(s+(r+1)/2,πv,μv)Lv(s+n+(r+1)/2,μv)j=1nLv(2s+2n+r+12j,μv2)×φv.Z(1,s;f_{v},\varphi_{v})=\frac{L_{v}(s+(r+1)/2,\pi_{v},\mu_{v})}{L_{v}(s+n+(r+1)/2,\mu_{v})\prod_{j=1}^{n}L_{v}(2s+2n+r+1-2j,\mu^{2}_{v})}\times\varphi_{v}.

This completes the proof. ∎

5.4. Computations of ramified places

Fix a non-archimedean place vv. In this subsection, we compute the zeta integrals at the non-archimedean ramified place vv. We then show the following lemma.

Lemma 5.4.1.

Let αs\alpha_{s} be a standard section of Ir,n+r,v(s,μv,πv)I_{r,n+r,v}(s,\mu_{v},\pi_{v}). There exists a finite number of standard sections fv,s,1,,fv,s,f_{v,s,1},\ldots,f_{v,s,\ell} of I2n+r,v(s,μ)I_{2n+r,v}(s,\mu) and vectors φv,1,,φv,πv\varphi_{v,1},\ldots,\varphi_{v,\ell}\in\pi_{v} such that

j=1Zv(g,s;fv,j,φv,j)=αs(g),gSp2n(Fv).\sum_{j=1}^{\ell}Z_{v}(g,s;f_{v,j},\varphi_{v,j})=\alpha_{s}(g),\qquad g\in\mathrm{Sp}_{2n}(F_{v}).
Proof.

Put

Kn,v(𝔭va)={kKn,vk𝟏n mod 𝔭va}.K_{n,v}(\mathfrak{p}_{v}^{a})=\{k\in K_{n,v}\mid\text{$k\equiv\mathbf{1}_{n}$ mod $\mathfrak{p}_{v}^{a}$}\}.

Let \ell be a positive integer such that αs\alpha_{s} is fixed by Kn+r,v(𝔭v)K_{n+r,v}(\mathfrak{p}_{v}^{\ell}). We write K=Kn+r,v(𝔭v)K=K_{n+r,v}(\mathfrak{p}_{v}^{\ell}). Let {γ1,,γ}Kn+r,v\{\gamma_{1},\ldots,\gamma_{\ell}\}\subset K_{n+r,v} be a set of complete representatives of Pr,n+r(Fv)\Sp2(n+r)(Fv)/KP_{r,n+r}(F_{v})\backslash\mathrm{Sp}_{2(n+r)}(F_{v})/K. We may assume γ1=1\gamma_{1}=1. Put φj=αs(γj)\varphi_{j}=\alpha_{s}(\gamma_{j}) and Kφj=StabKn,v(φj)K_{\varphi_{j}}=\mathrm{Stab}_{K_{n,v}}(\varphi_{j}). We claim that for any jj, one has prn(Kn+r,vPr,n+r(Fv))Kφj\mathrm{pr}_{n}(K_{n+r,v}\cap P_{r,n+r}(F_{v}))\subset K_{\varphi_{j}}. Here, prn\mathrm{pr}_{n} is the projection map prn:Pr,n+r(Fv)GLr×Sp2nSp2n\mathrm{pr}_{n}\colon P_{r,n+r}(F_{v})\longrightarrow\mathrm{GL}_{r}\times\mathrm{Sp}_{2n}\longrightarrow\mathrm{Sp}_{2n}. Indeed, take kprn(Kn+r,vPr,n+r(Fv))k\in\mathrm{pr}_{n}(K_{n+r,v}\cap P_{r,n+r}(F_{v})). Fix kKk^{\prime}\in K such that prn(k)=k\mathrm{pr}_{n}(k^{\prime})=k. By the choice of KK, one has πv(k)φj=αs(kγj)\pi_{v}(k)\varphi_{j}=\alpha_{s}(k^{\prime}\gamma_{j}). Since KK is a normal subgroup of Kn+r,vK_{n+r,v}, one obtains αs(kγj)=αs(γjγj1kγj)=αs(γj)\alpha_{s}(k^{\prime}\gamma_{j})=\alpha_{s}(\gamma_{j}\gamma_{j}^{-1}k^{\prime}\gamma_{j})=\alpha_{s}(\gamma_{j}). Thus, πv(k)φj=φj\pi_{v}(k)\varphi_{j}=\varphi_{j} and kKφjk\in K_{\varphi_{j}}.

Let fv,s,jf_{v,s,j} be a standard section of I2n+r(s,μ)I_{2n+r}(s,\mu) such that

  • supp(fv,s,j)P2n+r(F)τn(K×Kφj)\mathrm{supp}(f_{v,s,j})\subset P_{2n+r}(F)\tau_{n}(K\times K_{\varphi_{j}}^{\prime}).

  • fv,s,j(pτn(k1×k2))=1vol(Kφj)μ(p)|p|s+(2n+r+1)/2f_{v,s,j}(p\tau_{n}(k_{1}\times k_{2}))=\frac{1}{\mathrm{vol}(K_{\varphi_{j}})}\mu(p)|p|^{s+(2n+r+1)/2} for pP2n+r(Fv)p\in P_{2n+r}(F_{v}) and k1×k2K×Kφjk_{1}\times k_{2}\in K\times K_{\varphi_{j}}.

Here, Kφj={k^kKφj}K_{\varphi_{j}}^{\prime}=\{\widehat{k}\mid k\in K_{\varphi_{j}}\}. Let kKk\in K. By the claim, if τ(k×h^)supp(fv,s,j)\tau(k\times\widehat{h})\in\mathrm{supp}(f_{v,s,j}), one has hKφjh\in K_{\varphi_{j}}. Thus, we have

Zv(k,s;fv,j,φj)\displaystyle Z_{v}(k,s;f_{v,j},\varphi_{j}) =Sp2n(Fv)fv,s,j(τn(k×h^))πv(h)(φj)𝑑h\displaystyle=\int_{\mathrm{Sp}_{2n}(F_{v})}f_{v,s,j}(\tau_{n}(k\times\widehat{h}))\pi_{v}(h)(\varphi_{j})\,dh
=Kφjfv,s,j(τn(k×h^))πv(h)(φj)𝑑h\displaystyle=\int_{K_{\varphi_{j}}}f_{v,s,j}(\tau_{n}(k\times\widehat{h}))\pi_{v}(h)(\varphi_{j})\,dh
=φj.\displaystyle=\varphi_{j}.

Next, we compute the support of the section. For gSp2(n+r)(Fv)g\in\mathrm{Sp}_{2(n+r)}(F_{v}), we assume Zv(g,s;f,φj)0Z_{v}(g,s;f,\varphi_{j})\neq 0. Suppose that gg lies in Pr,n+r(Fv)γqKqP_{r,n+r}(F_{v})\gamma_{q}K_{q} for some q1q\neq 1. Then, by the definition of fv,s,jf_{v,s,j}, one has

fv,s,j(τn(g×h^))=fv,s,j(τn(h1g×1))f_{v,s,j}(\tau_{n}(g\times\widehat{h}))=f_{v,s,j}(\tau_{n}(h^{-1}g\times 1))

for any hh. By h1gPr,n+r(Fv)γqKh^{-1}g\in P_{r,n+r}(F_{v})\gamma_{q}K with q1q\neq 1, we get fv,s,j(τn(g×h^))=0f_{v,s,j}(\tau_{n}(g\times\widehat{h}))=0. Hence we obtain

Zv(g,s;fv,j,φj)=0Z_{v}(g,s;f_{v,j},\varphi_{j})=0

and supp(r(γj1)Zv(,s;fv,j,φj))=Pr,n+r(Fv)Kγj=Pr,n+rγjK\mathop{\mathrm{supp}}(r(\gamma_{j}^{-1})Z_{v}(\,\cdot\,,s;f_{v,j},\varphi_{j}))=P_{r,n+r}(F_{v})K\gamma_{j}=P_{r,n+r}\gamma_{j}K. We then have

αs(g)=j=1r(γj1)Zv(g,s;f,φj).\alpha_{s}(g)=\sum_{j=1}^{\ell}r(\gamma_{j}^{-1})Z_{v}(g,s;f,\varphi_{j}).

This completes the proof. ∎

5.5. Computations of archimedean places

In this subsection, we assume F=F=\mathbb{Q} for simplicity. Let vv be the archimedean place of F=F=\mathbb{Q}. Let π\pi be a holomorphic discrete series representation of Gn()G_{n}(\mathbb{R}) with highest weight λ=(λ1,v,,λn,v)v\lambda=(\lambda_{1,v},\ldots,\lambda_{n,v})_{v}. For a standard section fsf_{s} of I2n+r(s,μ)I_{2n+r}(s,\mu), put

Zv(g,s;f,φ,φ)=Zv(g,s;f,φ),φZ_{v}(g,s;f,\varphi,\varphi^{\prime})=\langle Z_{v}(g,s;f,\varphi),\varphi^{\prime}\rangle

for gGn+r(Fv)g\in G_{n+r}(F_{v}) and φπ\varphi^{\prime}\in\pi. Here, ,\langle\,\cdot\,,\,\cdot\,\rangle is an invariant inner product on π\pi.

Lemma 5.5.1.

With the above notation, suppose that a real number s0s_{0} satisfies s0+ρ2n+rs_{0}\in\mathbb{Z}+\rho_{2n+r} and r/2<s0λn,vρ2n+r-r/2<s_{0}\leq\lambda_{n,v}-\rho_{2n+r} for any v𝐚v\in\mathbf{a}. Let fsf_{s} be a standard section of I2n+r(s,μ)I_{2n+r}(s,\mu) such that fs0f_{s_{0}} is 𝔭2n+r,\mathfrak{p}_{2n+r,-}-finite. Then, the integral Zv(g,s;f,φ,φ)Z_{v}(g,s;f,\varphi,\varphi^{\prime}) converges absolutely at s=s0s=s_{0} for any gGn+r(Fv)g\in G_{n+r}(F_{v}) and v,vπv,v^{\prime}\in\pi. Moreover, we may choose g,fsg,f_{s} and φ,φπ\varphi,\varphi^{\prime}\in\pi so that Zv(g,f;s,φ,φ)Z_{v}(g,f;s,\varphi,\varphi^{\prime}) is non-zero at s=s0s=s_{0}.

Proof.

For mGn(Fv)m\in G_{n}(F_{v}), one has

Zv((𝟏r×m)k,s;f,φ,φ)=Zv(1,s;r(k)f,φ,π(m1)φ).Z_{v}((\mathbf{1}_{r}\times m)k,s;f,\varphi,\varphi^{\prime})=Z_{v}(1,s;r(k)f,\varphi,\pi(m^{-1})\varphi^{\prime}).

Since the standard section r(k)fsr(k)f_{s} satisfies the assumption as in the statement, we may assume g=1g=1. Then, the integral equals to

Zv(1,s;f,v,v)=Gn(Fv)fs(τn(𝟏n+r×h^))π(h)v,v𝑑h=Gn(Fv)fs(τn((𝟏r×h)×𝟏n))π(h)v,v𝑑h.Z_{v}(1,s;f,v,v^{\prime})=\int_{G_{n}(F_{v})}f_{s}(\tau_{n}(\mathbf{1}_{n+r}\times\widehat{h}))\langle\pi(h)v,v^{\prime}\rangle\,dh=\int_{G_{n}(F_{v})}f_{s}(\tau_{n}((\mathbf{1}_{r}\times h)\times\mathbf{1}_{n}))\langle\pi(h)v,v^{\prime}\rangle\,dh.

Consider the restriction of fsf_{s} to the subgroup G2n(Fv)G_{2n}^{\downarrow}(F_{v}). The restriction ι2n,2n+r,fs\iota_{2n,2n+r}^{*,\downarrow}f_{s} to G2n(Fv)G^{\downarrow}_{2n}(F_{v}) is a standard section of I2n(s+r/2,μ)I_{2n}(s+r/2,\mu). The restriction map ι2n,2n+r,\iota_{2n,2n+r}^{\downarrow,*} induces a non-zero intertwining map

I2n+r(s0,μ)𝔭2n+r,finI2n(s0+r/2,μ)𝔭2n,fin=L(m/2,,m/2).I_{2n+r}(s_{0},\mu)_{\mathfrak{p}_{2n+r,-}{\rm\mathchar 45\relax fin}}\longrightarrow I_{2n}(s_{0}+r/2,\mu)_{\mathfrak{p}_{2n,-}{\rm\mathchar 45\relax fin}}=L(m/2,\ldots,m/2).

Thus, ZvZ_{v} induces

L(m/2,,m/2)ππ.L(m/2,\ldots,m/2)\otimes\pi\otimes\pi\longrightarrow\mathbb{C}.

Here, m=2s0+2n+r+12n+2m=2s_{0}+2n+r+1\geq 2n+2. This map is the same as in [Liu20, (4.3.4)]. Hence, the lemma follows from [Liu20, Proposition 4.3.1]. This completes the proof. ∎

Corollary 5.5.2.

With the above notation, the zeta integral at s=s0s=s_{0} induces a non-zero intertwining map

I2n+r(μ,s0)𝔭2n+r,finπIr,n+r(s0,μ,π)𝔭n+r,fin.I_{2n+r}(\mu,s_{0})_{\mathfrak{p}_{2n+r,-}{\rm\mathchar 45\relax fin}}\otimes\pi\longrightarrow I_{r,n+r}(s_{0},\mu,\pi)_{\mathfrak{p}_{n+r,-}{\rm\mathchar 45\relax fin}}.
Proof.

By Lemma 5.5.1, the zeta integral defines a non-zero intertwining map

I2n+r(μ,s0)𝔭2n+r,finπIr,n+r(s0,μ,π).I_{2n+r}(\mu,s_{0})_{\mathfrak{p}_{2n+r,-}{\rm\mathchar 45\relax fin}}\otimes\pi\longrightarrow I_{r,n+r}(s_{0},\mu,\pi).

Since the integral is intertwining, the image is contained in the 𝔭n+r,\mathfrak{p}_{n+r,-}-finite vectors. This completes the proof. ∎

6. Structure theorem of the space of nearly holomorphic automorphic forms

In this section, we compare the space of nearly holomorphic automorphic forms with the space of Eisenstein series.

6.1. Parametrization of infinitesimal characters

For an infinitesimal character χ\chi of 𝒵n\mathcal{Z}_{n}, put

𝒩(Gn,χ)={φ𝒩(Gn)(zχ(z))φ=0 for any z𝒵n}.\mathcal{N}(G_{n},\chi)=\{\varphi\in\mathcal{N}(G_{n})\mid\text{$(z-\chi(z))\varphi=0$ for any $z\in\mathcal{Z}_{n}$}\}.

By [Hor20b, Proposition 5.15], we have

(6.1.1) 𝒩(Gn)=χ𝒩(Gn,χ),\displaystyle\mathcal{N}(G_{n})=\bigoplus_{\chi}\mathcal{N}(G_{n},\chi),

where χ\chi runs over all integral infinitesimal characters of 𝒵n\mathcal{Z}_{n}. We define 𝒩(Gn,χ){P}\mathcal{N}(G_{n},\chi)_{\{P\}} and 𝒩(Gn,χ)(M,π)\mathcal{N}(G_{n},\chi)_{(M,\pi)} similarly. By [Hor20b, Proposition 5.9], the constant term along Qi,nQ_{i,n} induces an embedding of the space 𝒩(Gn,χ)(MQi,n,μπ)\mathcal{N}(G_{n},\chi)_{(M_{Q_{i,n}},\mu\boxtimes\pi)} into the direct sum

s0Ii,n(s0,μ,π).\bigoplus_{s_{0}}I_{i,n}(s_{0},\mu,\pi).

Here s0s_{0} runs over all real numbers such that the induced representation Ii,n(s0,μ,π)I_{i,n}(s_{0},\mu,\pi) has the integral infinitesimal character χ\chi. Take a real number tt. We define the projection 𝐩𝐫t{\mathbf{pr}}_{t} by

𝐩𝐫t:s0Ii,n(s0,μ,π)Ii,n(t,μ,π).\mathbf{pr}_{t}\colon\bigoplus_{s_{0}}I_{i,n}(s_{0},\mu,\pi)\longrightarrow I_{i,n}(t,\mu,\pi).

The infinitesimal character of the induced representation Ii,n(s0,μ,π)I_{i,n}(s_{0},\mu,\pi) has the Harish-Chandra parameter

(λ1,v,,λni,v,s0+n(i1)/2,,s0+n(i1)/2)+ρ.(\lambda_{1,v},\ldots,\lambda_{n-i,v},s_{0}+n-(i-1)/2,\ldots,s_{0}+n-(i-1)/2)+\rho.

Here, (λ1,v,,λni,v)(\lambda_{1,v},\ldots,\lambda_{n-i,v}) is the highest weight of πv\pi_{v} for any v𝐚v\in\mathbf{a}. For χs0\chi_{s_{0}}, we mean the infinitesimal character of the induced representation. Note that χs0\chi_{s_{0}} depends on λ\lambda and ii.

Lemma 6.1.1.

With the above notation, fix s0s_{0}. Let {t1,,t}\{t_{1},\ldots,t_{\ell}\} be the set of real numbers such that 𝐩𝐫tj(φQi,n)0\mathbf{pr}_{t_{j}}(\varphi_{Q_{i,n}})\neq 0 for some φ𝒩(Gn,χs0)(μπ,MQi,n)\varphi\in\mathcal{N}(G_{n},\chi_{s_{0}})_{(\mu\boxtimes\pi,M_{Q_{i,n}})}. Then, for any jj, the highest weight submodule of Ii,n(tj,μ,π)I_{i,n}(t_{j},\mu,\pi) is unitarizable.

Proof.

We may assume t1<<tρi,n+s0t_{1}<\cdots<t_{\ell}\leq\rho_{i,n}+s_{0}. Note that the highest weight of Ii,n(tj,μ,π)I_{i,n}(t_{j},\mu,\pi) is of the form aj=(λ1,v,,λni,v,ρi,n+tj,,ρi,n+tj)a_{j}=(\lambda_{1,v},\ldots,\lambda_{n-i,v},\rho_{i,n}+t_{j},\ldots,\rho_{i,n}+t_{j}). Then, a1a_{1} is maximal in {a1,,a}\{a_{1},\ldots,a_{\ell}\}. By assumption, there exists φ𝒩(Gn,χs0)(μπ,MQi,n)\varphi\in\mathcal{N}(G_{n},\chi_{s_{0}})_{(\mu\boxtimes\pi,M_{Q_{i,n}})} such that φ\varphi is of weight a1a_{1}. Note that by maximality of a1a_{1}, for j1j\neq 1, the Kn,K_{n,\infty}-type ρa1\rho_{a_{1}} does not occur in Ii,n(tj,μ,π)𝔭n,finI_{i,n}(t_{j},\mu,\pi)_{\mathfrak{p}_{n,-}{\rm\mathchar 45\relax fin}}. Then, φQi,n\varphi_{Q_{i,n}} lies in Ii,n(t1,μ,π)I_{i,n}(t_{1},\mu,\pi). By [Hor20b, Corollary 7.3], the module generated by φQi,n\varphi_{Q_{i,n}} is isomorphic to L(a1)L(a_{1}). By [MW95, I.4.11], if t1<0t_{1}<0, the automorphic form φ\varphi is of square-integrable. Thus, the highest weight module L(a1)L(a_{1}) is unitarizable. If t10t_{1}\geq 0, the highest weight module L(a1)L(a_{1}) is unitarizable by Theorem 3.2.2. Since the highest weight submodule of Ii,n(tj,μ,π)I_{i,n}(t_{j},\mu,\pi) is irreducible with integral weight aja_{j}, the highest weight submodules L(aj)L(a_{j}) are unitarizable by Theorem 3.2.2 for all jj. This completes the proof. ∎

If the induced representation Ii,n(s0,μ,π)I_{i,n}(s_{0},\mu,\pi) contains a unitary highest weight representation, one has s0+n(i1)/2s_{0}\in\mathbb{Z}+n-(i-1)/2 with nis0+n(i1)/2λn,vn-i\leq s_{0}+n-(i-1)/2\leq\lambda_{n,v}. The following statement follows from the straightforward computation. For details, see [Hor20b, Proposition 6.4].

Lemma 6.1.2.

With the above notation, suppose for simplicity F=F=\mathbb{Q}. Let a,ba,b be real numbers so that a,b+n(i1)/2,nia+n(i1)/2λn,va,b\in\mathbb{Z}+n-(i-1)/2,n-i\leq a+n-(i-1)/2\leq\lambda_{n,v} and nib+n(i1)/2λn,vn-i\leq b+n-(i-1)/2\leq\lambda_{n,v}. Then, one has χa=χb\chi_{a}=\chi_{b} if and only if |a|=|b||a|=|b|.

Put

𝒩2(Gn,χ)(M,π)={φ𝒩(Gn,χ)(M,π)φ is square-integrable}.\mathcal{N}^{2}(G_{n},\chi)_{(M,\pi)}=\{\varphi\in\mathcal{N}(G_{n},\chi)_{(M,\pi)}\mid\text{$\varphi$ is square-integrable}\}.

In the following of this section, we study 𝒩(Gn,χ)(M,π)\mathcal{N}(G_{n},\chi)_{(M,\pi)} in terms of 𝒩2(Gn,χ)(M,π)\mathcal{N}^{2}(G_{n},\chi)_{(M,\pi)} and induced representations.

6.2. Constant terms of nearly holomorphic automorphic forms

Toward the classification of (𝔤n,Kn,)(\mathfrak{g}_{n},K_{n,\infty})-modules generated by nearly holomorphic automorphic forms on Gn(𝔸)G_{n}(\mathbb{A}_{\mathbb{Q}}), we investigate the embedding of 𝒩(Gn)(M,π)\mathcal{N}(G_{n})_{(M,\pi)} into a direct sum of induced representations. Fix a positive integer ini\leq n. Let μ\mu be a character of GL1(𝔸F)\mathrm{GL}_{1}(\mathbb{A}_{F}) and π\pi an irreducible holomorphic cuspidal automorphic representation on Gni(𝔸)G_{n-i}(\mathbb{A}_{\mathbb{Q}}) with πv=L(λv)=L(λ1,v,,λni,v)\pi_{v}=L(\lambda_{v})=L(\lambda_{1,v},\ldots,\lambda_{n-i,v}) for v𝐚v\in\mathbf{a}. Put Π=μπ\Pi=\mu\boxtimes\pi. For the notation, see §2.5. We consider the space 𝒩(Gn,χs0)(MQi,n,Π)\mathcal{N}(G_{n},\chi_{s_{0}})_{(M_{Q_{i,n}},\Pi)}. By Lemma 6.1.2, the constant term along Pi,nP_{i,n} induces the embedding

(6.2.1) 𝒩(Gn,χs0)(MQi,n,Π){(Ii,n(s0,μ,π)Ii,n(s0,μ,π))𝔭n,finif s00Ii,n(s0,μ,π)𝔭n,finif s0=0.\displaystyle\mathcal{N}(G_{n},\chi_{s_{0}})_{(M_{Q_{i,n}},\Pi)}\xhookrightarrow{\,\quad\,}\begin{dcases}\left(I_{i,n}(-s_{0},\mu,\pi)\oplus I_{i,n}(s_{0},\mu,\pi)\right)_{\mathfrak{p}_{n,-}{\rm\mathchar 45\relax fin}}&\text{if $s_{0}\neq 0$}\\ I_{i,n}(s_{0},\mu,\pi)_{\mathfrak{p}_{n,-}{\rm\mathchar 45\relax fin}}&\text{if $s_{0}=0$}.\end{dcases}

Note that 𝒩(Gn,χs0)(MQi,n,Π)=𝒩2(Gn,χs0)(MQi,n,Π)\mathcal{N}(G_{n},\chi_{s_{0}})_{(M_{Q_{i,n}},\Pi)}=\mathcal{N}^{2}(G_{n},\chi_{s_{0}})_{(M_{Q_{i,n}},\Pi)} if and only if the image of (6.2.1) is contained in Ii,n(s0,μ,π)I_{i,n}(-s_{0},\mu,\pi). In this case, the space 𝒩(Gn,χs0)(MQi,n,Π)\mathcal{N}(G_{n},\chi_{s_{0}})_{(M_{Q_{i,n}},\Pi)} is semisimple as (𝔤n,Kn,)(\mathfrak{g}_{n},K_{n,\infty})-modules. The highest weights of the right hand side of (6.2.1) are of the form

(λ1,v,,λni,v,ρi,n+s0,,ρi,n+s0)v(\lambda_{1,v},\ldots,\lambda_{n-i,v},\rho_{i,n}+s_{0},\ldots,\rho_{i,n}+s_{0})_{v}

and

(λ1,v,,λni,v,ρi,ns0,,ρi,ns0)v(\lambda_{1,v},\ldots,\lambda_{n-i,v},\rho_{i,n}-s_{0},\ldots,\rho_{i,n}-s_{0})_{v}

if exist. If λni,v<ρi,n\lambda_{n-i,v}<\rho_{i,n} for some vv, one has 𝒩(Gn,χs0)(MQi,n,Π)=𝒩2(Gn,χs0)(MQi,n,Π)\mathcal{N}(G_{n},\chi_{s_{0}})_{(M_{Q_{i,n}},\Pi)}=\mathcal{N}^{2}(G_{n},\chi_{s_{0}})_{(M_{Q_{i,n}},\Pi)}. Thus, for the classification, it suffices to consider the case where λ\lambda satisfies λni,vρi,n\lambda_{n-i,v}\geq\rho_{i,n} for any v𝐚v\in\mathbf{a}. In this case, we may assume 0s0minv𝐚{λni,vρi,n}0\leq s_{0}\leq\mathrm{min}_{v\in\mathbf{a}}\{\lambda_{n-i,v}-\rho_{i,n}\}.

Lemma 6.2.1.

Under the above assumption, if s0=0s_{0}=0, the space 𝒩(Gn,χs0)(MQi,n,Π)\mathcal{N}(G_{n},\chi_{s_{0}})_{(M_{Q_{i,n}},\Pi)} is isotypic for

v𝐚L(λ1,v,λni,v,ρi,n,,ρi,n)\boxtimes_{v\in\mathbf{a}}L(\lambda_{1,v},\ldots\lambda_{n-i,v},\rho_{i,n},\ldots,\rho_{i,n})

as (𝔤n,Kn,)(\mathfrak{g}_{n},K_{n,\infty})-modules.

Proof.

By (6.2.1), one has

𝒩(Gn,χs0)(MQi,n,Π)Ii,n(0,μ,π).\mathcal{N}(G_{n},\chi_{s_{0}})_{(M_{Q_{i,n}},\Pi)}\xhookrightarrow{\,\quad\,}I_{i,n}(0,\mu,\pi).

Consider the induced representation

Ii,n,v(0,μv,L(λv))I_{i,n,v}(0,\mu_{v},L(\lambda_{v}))

for v𝐚v\in\mathbf{a} and a unitary character μv\mu_{v}. Since this induced representation lies in the unitary axis, it is unitary by the unitarizability of L(λ)L(\lambda). Thus, it is semisimple as (𝔤n,Kn,)(\mathfrak{g}_{n},K_{n,\infty})-modules. Highest weights in it are of the form (λ1,v,λni,v,ρi,n,,ρi,n)(\lambda_{1,v},\ldots\lambda_{n-i,v},\rho_{i,n},\ldots,\rho_{i,n}). We then have

Ii,n,v(0,μv,L(λv))𝔭n,finL(λ1,v,λni,v,ρi,n,,ρi,n).I_{i,n,v}(0,\mu_{v},L(\lambda_{v}))_{\mathfrak{p}_{n,-}{\rm\mathchar 45\relax fin}}\subset L(\lambda_{1,v},\ldots\lambda_{n-i,v},\rho_{i,n},\ldots,\rho_{i,n}).

This completes the proof. ∎

In the following of this section, we assume λni,v>ρi,n\lambda_{n-i,v}>\rho_{i,n} for any v𝐚v\in\mathbf{a}, s0+ρi,ns_{0}\in\mathbb{Z}+\rho_{i,n} and 0<s0minv𝐚{λni,vρi,n}0<s_{0}\leq\mathrm{min}_{v\in\mathbf{a}}\{\lambda_{n-i,v}-\rho_{i,n}\}. We then have

𝒩2(Gn,χs0)(Qi,n,μπ)\𝒩(Gn,χs0)(Qi,n,μπ)Ii,n(s0,μ,π)𝔭n,fin.\mathcal{N}^{2}(G_{n},\chi_{s_{0}})_{(Q_{i,n},\mu\boxtimes\pi)}\backslash\mathcal{N}(G_{n},\chi_{s_{0}})_{(Q_{i,n},\mu\boxtimes\pi)}\xhookrightarrow{\qquad}I_{i,n}(s_{0},\mu,\pi)_{\mathfrak{p}_{n,-}{\rm\mathchar 45\relax fin}}.

6.3. Structure theorem for i=ni=n

Proposition 6.3.1.

We assume that either of the following conditions holds:

  • FF\neq\mathbb{Q} and s0>0s_{0}>0.

  • μ2𝟏\mu^{2}\neq\mathbf{1} and s0>0s_{0}>0.

We then have

𝒩(Gn,χs0)(B,μ)𝒩2(Gn,χs0)(B,μ)In(s0,μ)𝔭n,fin.\mathcal{N}(G_{n},\chi_{s_{0}})_{(B,\mu)}\cong\mathcal{N}^{2}(G_{n},\chi_{s_{0}})_{(B,\mu)}\oplus I_{n}(s_{0},\mu)_{\mathfrak{p}_{n,-}{\rm\mathchar 45\relax fin}}.
Proof.

By (6.2.1), the constant term along PnP_{n} induces the injective map

𝒩2(Gn,χs0)(B,μ)\𝒩(Gn,χs0)(B,μ)In(s0,μ)𝔭n,fin.\mathcal{N}^{2}(G_{n},\chi_{s_{0}})_{(B,\mu)}\backslash\mathcal{N}(G_{n},\chi_{s_{0}})_{(B,\mu)}\xhookrightarrow{\qquad}I_{n}(s_{0},\mu)_{\mathfrak{p}_{n,-}{\rm\mathchar 45\relax fin}}.

By Lemma 4.3.1, the Eisenstein series at s=s0s=s_{0} gives the splitting

In(s0,μ)𝔭n,fin𝒩(Gn,χs0)(B,μ).I_{n}(s_{0},\mu)_{\mathfrak{p}_{n,-}{\rm\mathchar 45\relax fin}}\xhookrightarrow{\qquad}\mathcal{N}(G_{n},\chi_{s_{0}})_{(B,\mu)}.

Hence the statement follows. ∎

Next we treat the case F=F=\mathbb{Q}.

Proposition 6.3.2.

The following statements hold.

  1. (1)

    For s0>1s_{0}>1, one has

    𝒩(Gn,χs0)(B,μ)𝒩2(Gn,χs0)(B,μ)In(s0,μ)𝔭n,fin.\mathcal{N}(G_{n},\chi_{s_{0}})_{(B,\mu)}\cong\mathcal{N}^{2}(G_{n},\chi_{s_{0}})_{(B,\mu)}\oplus I_{n}(s_{0},\mu)_{\mathfrak{p}_{n,-}{\rm\mathchar 45\relax fin}}.
  2. (2)

    For s0=1s_{0}=1, one has

    𝒩2(Gn,χs0)(B,μ)\𝒩(Gn,χs0)(B,μ)In(1,μ)𝔭n,fin.\mathcal{N}^{2}(G_{n},\chi_{s_{0}})_{(B,\mu)}\backslash\mathcal{N}(G_{n},\chi_{s_{0}})_{(B,\mu)}\cong I_{n}(1,\mu)_{\mathfrak{p}_{n,-}{\rm\mathchar 45\relax fin}}.

    Moreover, there are no splitting In(1,μ)𝔭n,fin𝒩(Gn,χs0)(B,μ)I_{n}(1,\mu)_{\mathfrak{p}_{n,-}{\rm\mathchar 45\relax fin}}\longrightarrow\mathcal{N}(G_{n},\chi_{s_{0}})_{(B,\mu)} if In(1,μ)𝔭n,fin0I_{n}(1,\mu)_{\mathfrak{p}_{n,-}{\rm\mathchar 45\relax fin}}\neq 0.

  3. (3)

    For s0=1/2s_{0}=1/2, one has

    𝒩(Gn,χs0)(B,μ)=𝒩2(Gn,χs0)(B,μ),if μvsgn(n+2)/2 for any v𝐚\mathcal{N}(G_{n},\chi_{s_{0}})_{(B,\mu)}=\mathcal{N}^{2}(G_{n},\chi_{s_{0}})_{(B,\mu)},\qquad\text{if $\mu_{v}\neq\mathrm{sgn}^{(n+2)/2}$ for any $v\in\mathbf{a}$}

    and

    𝒩(Gn,χs0)(B,μ)In(1/2,μ)𝔭n,fin,if μv=sgn(n+2)/2 for any v𝐚.\mathcal{N}(G_{n},\chi_{s_{0}})_{(B,\mu)}\subset I_{n}(1/2,\mu)_{\mathfrak{p}_{n,-}{\rm\mathchar 45\relax fin}},\qquad\text{if $\mu_{v}=\mathrm{sgn}^{(n+2)/2}$ for any $v\in\mathbf{a}$}.
Proof.

The proof of (1) is the same as the proof of Proposition 6.3.1.

Next we show (2). If μvsgn(n+3)/2\mu_{v}\neq\mathrm{sgn}^{(n+3)/2} for some v𝐚v\in\mathbf{a}, one has 𝒩(Gn,χs0)(B,μ)=0\mathcal{N}(G_{n},\chi_{s_{0}})_{(B,\mu)}=0 and In(1,μ)𝔭n,fin=0I_{n}(1,\mu)_{\mathfrak{p}_{n,-}{\rm\mathchar 45\relax fin}}=0. We may assume μv=sgn(n+3)/2\mu_{v}=\mathrm{sgn}^{(n+3)/2} for any v𝐚v\in\mathbf{a}. Take f=vfvIn(1,μ)f=\bigotimes_{v}f_{v}\in I_{n}(1,\mu) such that fvf_{v} lies in Rn(Wv)R_{n}(W_{v}) for some WvW_{v} and fVRn(V)f\not\in\sum_{V}R_{n}(V). Here VV runs over all positive definite quadratic forms over FF of dimension n+3n+3. Let fsf_{s} be the standard section of In(s,μ)I_{n}(s,\mu) such that f1=ff_{1}=f. We assume that there exists a nearly holomorphic automorphic form φ𝒩(Gn){B}\varphi\in\mathcal{N}(G_{n})_{\{B\}} such that φPn=f\varphi_{P_{n}}=f. By Proposition 4.4.1, the difference φE(,1,f)\varphi-E(\,\cdot\,,1,f) is non-zero and square integrable. However, for v𝐚v\in\mathbf{a}, the Kn,vK_{n,v}-type ((n+3)/2,,(n+3)/2)((n+3)/2,\ldots,(n+3)/2) in In,v(1,μv)I_{n,v}(-1,\mu_{v}) generates a reducible indecomposable representation of Sp2n(Fv)\mathrm{Sp}_{2n}(F_{v}). This contradicts to the square integrability. Hence there are no automorphic form φ\varphi such that φPn=f\varphi_{P_{n}}=f. Recall that the constant term along PnP_{n} induces the inclusion

𝒩2(Gn,χs0)(B,μ)\𝒩(Gn,χs0)(B,μ)In(1,μ)𝔭n,fin.\mathcal{N}^{2}(G_{n},\chi_{s_{0}})_{(B,\mu)}\backslash\mathcal{N}(G_{n},\chi_{s_{0}})_{(B,\mu)}\xhookrightarrow{\,\quad\,}I_{n}(1,\mu)_{\mathfrak{p}_{n,-}{\rm\mathchar 45\relax fin}}.

The image of E(,1,f)E(\,\cdot\,,1,f) is the same as ff. Hence the above inclusion is surjective and there are no splitting. This completes the proof of (2).

For (3), we assume μvsgn(n+2)/2\mu_{v}\neq\mathrm{sgn}^{(n+2)/2} for any v𝐚v\in\mathbf{a}. Note that if In,v(1/2,μv)I_{n,v}(1/2,\mu_{v}) has a highest weight vector, one has μv=sgn(n+2)/2\mu_{v}=\mathrm{sgn}^{(n+2)/2}. Thus, the constant term (6.2.1) induces the embedding

𝒩(Gn,χs0)(B,μ)In(1/2,μ)𝔭n,fin.\mathcal{N}(G_{n},\chi_{s_{0}})_{(B,\mu)}\xhookrightarrow{\,\quad\,}I_{n}(-1/2,\mu)_{\mathfrak{p}_{n,-}{\rm\mathchar 45\relax fin}}.

We then have 𝒩(Gn,χs0)(B,μ)=𝒩2(Gn,χs0)(B,μ)\mathcal{N}(G_{n},\chi_{s_{0}})_{(B,\mu)}=\mathcal{N}^{2}(G_{n},\chi_{s_{0}})_{(B,\mu)}. The last statement follows immediately from (6.2.1). This completes the proof. ∎

6.4. Structure theorem for PBP\neq B

Fix ii. We consider the case P=Pi,nP=P_{i,n} Let μ\mu be a character of GLi(𝔸F)\mathrm{GL}_{i}(\mathbb{A}_{F}) and π\pi an irreducible holomorphic cuspidal representation of Gni(𝔸)G_{n-i}(\mathbb{A}_{\mathbb{Q}}) and s0+ρi,ns_{0}\in\mathbb{Z}+\rho_{i,n}. Suppose s0>0s_{0}>0. Let SS be a finite set of places such that 𝐚S\mathbf{a}\subset S and for vSv\not\in S, the representations μv\mu_{v} and πv\pi_{v} are unramified. Set

LS(s,π,μ)=vSL(s,πv,μv).L^{S}(s,\pi,\mu)=\prod_{v\not\in S}L(s,\pi_{v},\mu_{v}).
Lemma 6.4.1.

Let α=vαvIi,n(s0,μ,π)𝔭n,fin\alpha=\bigotimes_{v}\alpha_{v}\in I_{i,n}(s_{0},\mu,\pi)_{\mathfrak{p}_{n,-}{\rm\mathchar 45\relax fin}} and SS the finite set of places such that for vSv\not\in S, the function αv\alpha_{v} is unramified. Then, there exists finite number of standard sections f1,,ff_{1},\ldots,f_{\ell} of I2n+r(s,μ)I_{2n+r}(s,\mu) and φ1,φπ\varphi_{1},\ldots\varphi_{\ell}\in\pi such that

limss01LS(s+(r+1)/2,π,μ)j=1Z(g,s;fj,φj)=α(g),gGn(𝔸).\lim_{s\rightarrow s_{0}}\frac{1}{L^{S}(s+(r+1)/2,\pi,\mu)}\sum_{j=1}^{\ell}Z(g,s;f_{j},\varphi_{j})=\alpha(g),\qquad g\in G_{n}(\mathbb{A}_{\mathbb{Q}}).
Proof.

The statement follows from Lemma 5.3.1, Lemma 5.4.1 and Corollary 5.5.2. ∎

Proposition 6.4.2.

Suppose s0>1s_{0}>1 if F=F=\mathbb{Q}. We then have

𝒩(Gn,χs0)(MQi,n,μπ)𝒩2(Gn,χs0)(MQi,n,μπ)Ii,n(s0,μ,π)𝔭n,fin.\mathcal{N}(G_{n},\chi_{s_{0}})_{(M_{Q_{i,n}},\mu\boxtimes\pi)}\cong\mathcal{N}^{2}(G_{n},\chi_{s_{0}})_{(M_{Q_{i,n}},\mu\boxtimes\pi)}\oplus I_{i,n}(s_{0},\mu,\pi)_{\mathfrak{p}_{n,-}{\rm\mathchar 45\relax fin}}.
Proof.

It suffices to show that for any 𝔭n,\mathfrak{p}_{n,-}-finite function αIi,n(s0,μ,π)\alpha\in I_{i,n}(s_{0},\mu,\pi), there exists a nearly holomorphic automorphic form φ𝒩(Gn,χs0)(MQi,n,μπ)\varphi\in\mathcal{N}(G_{n},\chi_{s_{0}})_{(M_{Q_{i,n}},\mu\boxtimes\pi)} such that φPi,n=α\varphi_{P_{i,n}}=\alpha. This follows immediately from Proposition 5.2.3 and Lemma 6.4.1. This completes the proof. ∎

In the following, we give partial results.

Proposition 6.4.3.

Assume F=F=\mathbb{Q}. Let Π=μπ\Pi=\mu\boxtimes\pi be an irreducible holomorphic cuspidal automorphic representation of MPi,n(𝔸)M_{P_{i,n}}(\mathbb{A}_{\mathbb{Q}}). Suppose that highest weights of the archimedean component v𝐚πv=v𝐚L(λ1,v,,λni,v)\bigotimes_{v\in\mathbf{a}}\pi_{v}=\bigotimes_{v\in\mathbf{a}}L(\lambda_{1,v},\ldots,\lambda_{n-i,v}) satisfies λni,vρi,n+s0\lambda_{n-i,v}\geq\rho_{i,n}+s_{0}. We then obtain the following result:

  1. (1)

    For s0=1/2s_{0}=1/2, the space 𝒩(Gn,χs0)(M,Π)\mathcal{N}(G_{n},\chi_{s_{0}})_{(M,\Pi)} is v𝐚L(λ1,v,,λni,v,ρi,n+ε,,ρi,n+ε)\bigotimes_{v\in\mathbf{a}}L(\lambda_{1,v},\ldots,\lambda_{n-i,v},\rho_{i,n}+\varepsilon,\ldots,\rho_{i,n}+\varepsilon)-isotypic. Here, ε{±1/2}\varepsilon\in\{\pm 1/2\} is defined so that sgnρi,n+ε=μv\mathrm{sgn}^{\rho_{i,n}+\varepsilon}=\mu_{v} for any vv.

  2. (2)

    For s0=1s_{0}=1, the space 𝒩(Gn,χs0)(MQi,n,Π)\mathcal{N}(G_{n},\chi_{s_{0}})_{(M_{Q_{i,n}},\Pi)} is contained in

    Ii,n(1,μ,π)Ii,n(1,μ,π).I_{i,n}(-1,\mu,\pi)\oplus I_{i,n}(1,\mu,\pi).
Proof.

The statements follow from (6.2.1) immediately. ∎

6.5. Classification of (𝔤n,Kn,)(\mathfrak{g}_{n},K_{n,\infty})-module generated by nearly holomorphic automorphic forms

We finally show the following classification:

Theorem 6.5.1.

Let MM be an indecomposable reducible (𝔤n,Kn,)(\mathfrak{g}_{n},K_{n,\infty})-module generated by a nearly holomorphic modular form. Then, the length of MM is at most two. Moreover, if FF\neq\mathbb{Q}, MM is irreducible. If F=F=\mathbb{Q} and MM is reducible, let L(a1,,an)L(a_{1},\ldots,a_{n}) be the socle of MM and L(b1,,bn)L(b_{1},\ldots,b_{n}) the irreducible quotient of MM. Then, there exists ii such that

  • aj=bja_{j}=b_{j} for j=1,,nij=1,\ldots,n-i.

  • ani+1==an=ρi,n1a_{n-i+1}=\cdots=a_{n}=\rho_{i,n}-1 and bni+1==bn=ρi,n+1b_{n-i+1}=\cdots=b_{n}=\rho_{i,n}+1.

  • MN(a1,,an)M\cong N(a_{1},\ldots,a_{n})^{\vee}.

Moreover, if a reducible module MM has a regular infinitesimal character, one has i=1i=1.

Proof.

We may assume MM is reducible. There exists s0(1/2)0s_{0}\in(1/2)\mathbb{Z}_{\geq 0}, a positive integer ii, a character μ\mu of GLi(𝔸F)\mathrm{GL}_{i}(\mathbb{A}_{F}) and an irreducible cuspidal automorphic representation π\pi of Gni(𝔸)G_{n-i}(\mathbb{A}_{\mathbb{Q}}) such that the indecomposable reducible module MM can be embedded into 𝒩(Gn,χs0)(MQi,n,μπ)\mathcal{N}(G_{n},\chi_{s_{0}})_{(M_{Q_{i,n}},\mu\boxtimes\pi)}. By Lemma 6.2.1, Proposition 6.3.1, Proposition 6.3.2, Proposition 6.4.2 and Proposition 6.4.3, since MM is reducible, one has F=F=\mathbb{Q} and s0=1s_{0}=1. In the following, we assume F=F=\mathbb{Q} and s0=1s_{0}=1.

Put

M1=M𝒩2(Gn,χs0)(MQi,n,Π).M_{1}=M\cap\mathcal{N}^{2}(G_{n},\chi_{s_{0}})_{(M_{Q_{i,n}},\Pi)}.

Then, the submodule M1M_{1} is semisimple. Since the submodule M1M_{1} occurs in

Ii,n(1,μ,π)𝔭n,fin,I_{i,n}(-1,\mu,\pi)_{\mathfrak{p}_{n,-}{\rm\mathchar 45\relax fin}},

the module M1M_{1} is isomorphic to L(λ1,,λni,ρi,n1,,ρi,n1)L(\lambda_{1},\ldots,\lambda_{n-i},\rho_{i,n}-1,\ldots,\rho_{i,n}-1) with some multiplicities. Put M2=M/M1M_{2}=M/M_{1}. Then, one obtains that M2M_{2} is isomorphic to L(λ1,,λni,ρi,n+1,,ρn,i+1)L(\lambda_{1},\ldots,\lambda_{n-i},\rho_{i,n}+1,\ldots,\rho_{n,i}+1) with some multiplicities by Proposition 6.3.2 (2) and Proposition 6.4.3. By Lemma 3.5.1, the module MM is isomorphic to N(λ1,,λni,ρi,n1,,ρn,i1)N(\lambda_{1},\ldots,\lambda_{n-i},\rho_{i,n}-1,\ldots,\rho_{n,i}-1)^{\vee}.

If MM has a regular infinitesimal character, the socle L(a1,,an)L(a_{1},\ldots,a_{n}) has a regular infinitesimal character. Then, one has i=1i=1. This completes the proof. ∎

Remark 6.5.2.

A typical example of nearly holomorphic modular form that generates an indecomposable reducible module is E2E_{2}. Here, E2E_{2} is defined by

E2(z)=3πy1+24n=1(0<d|nd)exp(2π1nz),z1.E_{2}(z)=\frac{3}{\pi y}-1+24\sum_{n=1}^{\infty}\left(\sum_{0<d|n}d\right)\exp(2\pi\sqrt{-1}nz),\qquad z\in\mathfrak{H}_{1}.

Then, E2E_{2} generates N(0)N(0)^{\vee}. For details, see [Hor21].

Corollary 6.5.3.

Let λ\lambda be a regular anti-dominant integral weight and χ=χλ\chi=\chi_{\lambda}. Let 𝒩Repn(χ)\mathcal{N}\mathrm{Rep}_{n}(\chi) be the set of isomorphism classes of indecomposable (𝔤n,Kn,)(\mathfrak{g}_{n},K_{n,\infty})-modules with the regular integral infinitesimal character χ\chi generated by nearly holomorphic Siegel modular forms of degree nn. For a Kn,K_{n,\infty}-type σ\sigma, put 𝒩Repn(χ,σ)={π𝒩Repn(χ)π has the Kn,-type σ}\mathcal{N}\mathrm{Rep}_{n}(\chi,\sigma)=\{\pi\in\mathcal{N}\mathrm{Rep}_{n}(\chi)\mid\text{$\pi$ has the $K_{n,\infty}$-type $\sigma$}\}. We then have

𝒩Repn(χ){{L(λ(0)),,L(λ(p)),N(λ(1))}if λn=n+1{L(λ)}if λnn+1\mathcal{N}\mathrm{Rep}_{n}(\chi)\subset\begin{cases}\{L(\lambda^{(0)}),\ldots,L(\lambda^{(p)}),N(\lambda^{(1)})^{\vee}\}&\text{if $\lambda_{n}=n+1$}\\ \{L(\lambda)\}&\text{if $\lambda_{n}\neq n+1$}\end{cases}

and

𝒩Repn(χ,detλ11j(λ)){{L(λ(0)),N(λ(1))}if λn=n+1{L(λ)}if λnn+1.\mathcal{N}\mathrm{Rep}_{n}(\chi,\mathrm{det}^{\lambda_{1}-1}\otimes\wedge^{j(\lambda)})\subset\begin{cases}\{L(\lambda^{(0)}),N(\lambda^{(1)})^{\vee}\}&\text{if $\lambda_{n}=n+1$}\\ \{L(\lambda)\}&\text{if $\lambda_{n}\neq n+1$}.\end{cases}
Proof.

The statement follows immediately from Proposition 3.4.4 and Theorem 6.5.1. ∎

7. Projection operators

In this section, we investigate projection operators associated to infinitesimal characters.

7.1. Generators of 𝒵n\mathcal{Z}_{n}

In this subsection, we assume F=F=\mathbb{Q} for simplicity. It is well-known that 𝒵n\mathcal{Z}_{n} is generated by nn generators. We give generators explicitly. We first define matrices Bi,jB_{i,j} and E±,i,j=E±,j,iE_{\pm,i,j}=E_{\pm,j,i} as follows:

Bi,j=(12(ei,jej,i)12(ei,j+ej,i)12(ei,j+ej,i)12(ei,jej,i)),E±,i,j=(12(ei,j+ej,i)±12(ei,j+ej,i)±12(ei,j+ej,i)12(ei,j+ej,i))B_{i,j}=\begin{pmatrix}\frac{1}{2}(e_{i,j}-e_{j,i})&\frac{-\sqrt{-1}}{2}(e_{i,j}+e_{j,i})\\ \frac{\sqrt{-1}}{2}(e_{i,j}+e_{j,i})&\frac{1}{2}(e_{i,j}-e_{j,i})\end{pmatrix},\qquad E_{\pm,i,j}=\begin{pmatrix}\frac{1}{2}(e_{i,j}+e_{j,i})&\frac{\pm\sqrt{-1}}{2}(e_{i,j}+e_{j,i})\\ \frac{\pm\sqrt{-1}}{2}(e_{i,j}+e_{j,i})&\frac{-1}{2}(e_{i,j}+e_{j,i})\end{pmatrix}

Then, {Bi,j1i,jn}\{B_{i,j}\mid 1\leq i,j\leq n\} and {E±,i,j1ijn}\{E_{\pm,i,j}\mid 1\leq i\leq j\leq n\} are basis of 𝔨n\mathfrak{k}_{n} and 𝔭n,±\mathfrak{p}_{n,\pm}, respectively. We define BMatn(Mat2n())B\in\mathrm{Mat}_{n}(\mathrm{Mat}_{2n}(\mathbb{C})) and E±Symn(Mat2n())E_{\pm}\in\mathrm{Sym}_{n}(\mathrm{Mat}_{2n}(\mathbb{C})) by

B=(Bk,)k,,E±=(E±,k,)k,Symn(Mat2n()).B=(B_{k,\ell})_{k,\ell},\qquad E_{\pm}=(E_{\pm,k,\ell})_{k,\ell}\in\mathrm{Sym}_{n}(\mathrm{Mat}_{2n}(\mathbb{C})).

Put B=(Bj,i)i,jB^{*}=(B_{j,i})_{i,j}, the transpose of BB. Let w=X1Xmw=X_{1}\cdots X_{m} be a word with letters B,BB,B^{*} and E±E_{\pm}. We assume the word ww satisfies the following five conditions:

  • E+E_{+} is followed by EE_{-} or BB^{*}.

  • EE_{-} is followed by E+E_{+} or BB.

  • BB is followed by E+E_{+} or BB.

  • BB^{*} is followed by EE_{-} or BB^{*}.

  • E+E_{+} and EE_{-} occur with the same multiplicity.

For a word, let tr(w)Mat2n()\mathrm{tr}(w)\in\mathrm{Mat}_{2n}(\mathbb{C}) be the trace as the Mat2n()\mathrm{Mat}_{2n}(\mathbb{C})-valued matrix. We may identify tr(w)\mathrm{tr}(w) as an element of 𝒰(𝔤n)\mathcal{U}(\mathfrak{g}_{n}). Let L(w)L(w) be the sum of number of times EBE_{-}B and BE+BE_{+} occur isolatedly in ww counted cyclicly. For example, L(EBE+)=0,L(EBE+B)=1,L(E+EBB)=L(EBBE+)=2L(E_{-}BE_{+})=0,L(E_{-}BE_{+}B^{*})=1,L(E_{+}E_{-}BB)=L(E_{-}BBE_{+})=2. Put

D2r=w(1)L(w)tr(w)D_{2r}=\sum_{w}(-1)^{L(w)}\mathrm{tr}(w)

where ww runs over all words of length 2r2r with the above five conditions.

Theorem 7.1.1 ([Mau12]).

The algebra 𝒵n\mathcal{Z}_{n} is generated by elements D2,,D2nD_{2},\ldots,D_{2n} as an algebra over \mathbb{C}.

7.2. Projection operators

Fix an infinitesimal character χ\chi, a weight ρ\rho and a congruence subgroup Γ\Gamma. Let KΓK_{\Gamma} be the closure of Γ\Gamma in Gn(𝔸,fin)G_{n}(\mathbb{A}_{\mathbb{Q},\mathrm{fin}}). We now define a projection on 𝒩(Gn)ρKΓ\mathcal{N}(G_{n})_{\rho}^{K_{\Gamma}}. Let λ\lambda be the highest weight of ρ\rho. We define a set X(ρ)X(\rho) of 𝔨n\mathfrak{k}_{n}-dominant weights by the set of 𝔨n\mathfrak{k}_{n}-dominant weights μ\mu such that μ\mu satisfies the following three conditions:

  • L(μ)L(\mu) is unitarizable.

  • L(μ)L(\mu) has the Kn,K_{n,\infty}-type ρ\rho.

  • λμ\lambda\leq\mu.

Then, X(ρ)X(\rho) is finite. Put χ(ρ)={χμμX(ρ)}\chi(\rho)=\{\chi_{\mu}\mid\mu\in X(\rho)\}. For infinitesimal characters χ\chi and ω\omega, we define Dχ,ωD_{\chi,\omega} as follows: Let v𝐚v\in\mathbf{a}. If the local components χv\chi_{v} and ωv\omega_{v} are the same, put Dχ,ω,v=1D_{\chi,\omega,v}=1. If χvωv\chi_{v}\neq\omega_{v}, there exists ii such that χv(D2i)ωv(D2i)\chi_{v}(D_{2i})\neq\omega_{v}(D_{2i}). Then, put Dχ,ω,v=D2iωv(D2i)D_{\chi,\omega,v}=D_{2i}-\omega_{v}(D_{2i}). Set Dχ,ω=v𝐚Dχ,ω,vD_{\chi,\omega}=\bigotimes_{v\in\mathbf{a}}D_{\chi,\omega,v}. By definition, for an ω\omega-eigenvector vv with ωχ(ρ)\omega\in\chi(\rho), we have

1χ(Dχ,ω)Dχ,ωv={vif χ=ω0if χω.\frac{1}{\chi(D_{\chi,\omega})}D_{\chi,\omega}\cdot v=\begin{cases}v&\text{if $\chi=\omega$}\\ 0&\text{if $\chi\neq\omega$}.\end{cases}

We now can define the projection 𝔭χEnd(𝒩(Gn)ρKΓ)\mathfrak{p}_{\chi}\in\mathrm{End}_{\mathbb{C}}(\mathcal{N}(G_{n})_{\rho}^{K_{\Gamma}}) by

𝔭χ(f)=1ωχ(ρ)χ(Dχ,ω)ωX(ρ)Dχ,ωf.\mathfrak{p}_{\chi}(f)=\frac{1}{\prod_{\omega\in\chi(\rho)}\chi(D_{\chi,\omega})}\prod_{\omega\in X(\rho)}D_{\chi,\omega}\cdot f.

By (6.1.1), 𝔭χ\mathfrak{p}_{\chi} defines a projection onto the χ\chi-eigen subspace of 𝒩(Gn)ρKΓ\mathcal{N}(G_{n})_{\rho}^{K_{\Gamma}} associated to χ\chi.

By Lemma 2.6.2, one has

Nρ(Γ)ρ𝒩(Gn)ρKΓ.N_{\rho}(\Gamma)\otimes\rho^{*}\cong\mathcal{N}(G_{n})_{\rho}^{K_{\Gamma}}.

The projection defines an endomorphism on Nρ(Γ)ρN_{\rho}(\Gamma)\otimes\rho^{*}.

Lemma 7.2.1.

The projection 𝔭χ\mathfrak{p}_{\chi} defines a projection on Nρ(Γ)N_{\rho}(\Gamma).

Proof.

We have the map

Nρ(Γ)HomKn,(ρ,Nρ(Γ)ρ)N_{\rho}(\Gamma)\longrightarrow\mathrm{Hom}_{K_{n,\infty}}(\rho^{*},N_{\rho}(\Gamma)\otimes\rho^{*})

by

f(vfv).f\longmapsto(v\longmapsto f\otimes v).

Since it is injective, it is isomorphism by comparing the dimensions. We identify Nρ(Γ)ρN_{\rho}(\Gamma)\otimes\rho^{*} as 𝒩(Gn)ρKΓ\mathcal{N}(G_{n})_{\rho}^{K_{\Gamma}}. Let (Nρ(Γ)ρ)χ(N_{\rho}(\Gamma)\otimes\rho^{*})_{\chi} be the χ\chi-eigen subspace of Nρ(Γ)ρN_{\rho}(\Gamma)\otimes\rho^{*} associated to an infinitesimal character χ\chi. Since the χ\chi-isotypic component of 𝒩(Gn)ρKΓ\mathcal{N}(G_{n})_{\rho}^{K_{\Gamma}} is Kn,K_{n,\infty}-stable, the corresponding space (Nρ(Γ)ρ)χ(N_{\rho}(\Gamma)\otimes\rho^{*})_{\chi} is Kn,K_{n,\infty}-stable. Thus we can define the subspace

HomKn,(ρ,(Nρ(Γ)ρ)χ)\mathrm{Hom}_{K_{n,\infty}}(\rho^{*},(N_{\rho}(\Gamma)\otimes\rho^{*})_{\chi})

of HomKn,(ρ,Nρ(Γ)ρ)\mathrm{Hom}_{K_{n,\infty}}(\rho^{*},N_{\rho}(\Gamma)\otimes\rho^{*}) and of Nρ(Γ)N_{\rho}(\Gamma). We denote the subspace of Nρ(Γ)N_{\rho}(\Gamma) by Nρ(Γ,χ)N_{\rho}(\Gamma,\chi). Since 𝒩(Gn)ρKΓ\mathcal{N}(G_{n})_{\rho}^{K_{\Gamma}} decomposes as the direct sum of χ\chi-eigen spaces, one has Nρ(Γ)=χNρ(Γ,χ)N_{\rho}(\Gamma)=\bigoplus_{\chi}N_{\rho}(\Gamma,\chi). By the map F𝔭χFF\longmapsto\mathfrak{p}_{\chi}\circ F, one obtains a map HomKn,(ρ,Nρ(Γ)ρ)HomKn,(ρ,(Nρ(Γ)ρ)χ)\mathrm{Hom}_{K_{n,\infty}}(\rho^{*},N_{\rho}(\Gamma)\otimes\rho^{*})\longrightarrow\mathrm{Hom}_{K_{n,\infty}}(\rho^{*},(N_{\rho}(\Gamma)\otimes\rho^{*})_{\chi}) and thus it induces the map Nρ(Γ)Nρ(Γ,χ)N_{\rho}(\Gamma)\longrightarrow N_{\rho}(\Gamma,\chi). It suffices to show that this map is a projection. For fNρ(Γ,χ)f\in N_{\rho}(\Gamma,\chi), one can regard ff as an element FHomKn,(ρ,(Nρ(Γ)ρ)χ)F\in\mathrm{Hom}_{K_{n,\infty}}(\rho^{*},(N_{\rho}(\Gamma)\otimes\rho^{*})_{\chi}). Since 𝔭χ\mathfrak{p}_{\chi} is projection, one has 𝔭χF=F\mathfrak{p}_{\chi}\circ F=F. It shows that ff is invariant under the map Nρ(Γ)Nρ(Γ,χ)N_{\rho}(\Gamma)\longrightarrow N_{\rho}(\Gamma,\chi). Thus, this map is an idempotent and hence a projection. This completes the proof. ∎

We denote by the same letter 𝔭χ\mathfrak{p}_{\chi} the projection on Nρ(Γ)N_{\rho}(\Gamma) as in the above lemma. Thus we have 𝔭χ(fv)=𝔭χ(f)v\mathfrak{p}_{\chi}(f\otimes v^{*})=\mathfrak{p}_{\chi}(f)\otimes v^{*} for fNρ(Γ)f\in N_{\rho}(\Gamma) and vρv^{*}\in\rho^{*}. Set Nρ(Γ,χ)=𝔭χ(Nρ(Γ))N_{\rho}(\Gamma,\chi)=\mathfrak{p}_{\chi}(N_{\rho}(\Gamma)).

Theorem 7.2.2.

The projection 𝔭χ\mathfrak{p}_{\chi} on Nρ(Γ)N_{\rho}(\Gamma) commutes with the Aut()\mathrm{Aut}(\mathbb{C})-action.

Proof.

The case where F=F=\mathbb{Q} is proved in [HPSS21, Proposition 3.16]. The general case is similar. We omit the details. ∎

For an integral weight λ=(λ1,v,,λn,v)v\lambda=(\lambda_{1,v},\ldots,\lambda_{n,v})_{v}, put jv(λ)=#{iλ1,vλi,v mod 2}j_{v}(\lambda)=\#\{i\mid\text{$\lambda_{1,v}\equiv\lambda_{i,v}$ mod $2$}\}.

Theorem 7.2.3.

Let λ=(λ1,v,,λn,v)v\lambda=(\lambda_{1,v},\ldots,\lambda_{n,v})_{v} be a regular anti-dominant 𝔨n\mathfrak{k}_{n}-dominant integral weight. Put ρ=v𝐚(detλ1,v1jv(λ))\rho=\bigotimes_{v\in\mathbf{a}}(\det^{\lambda_{1,v}-1}\otimes\wedge^{j_{v}(\lambda)}). If F=F=\mathbb{Q} and λn,v=n+1\lambda_{n,v}=n+1, any modular form in Nρ(Γ,χλ)N_{\rho}(\Gamma,\chi_{\lambda}) generates L(λ)L(\lambda) or N(λ(1))N(\lambda^{(1)})^{\vee}. If not, any modular form in Nρ(Γ,χλ)N_{\rho}(\Gamma,\chi_{\lambda}) generates L(λ)L(\lambda).

Proof.

Take fNρ(Γ,χλ)f\in N_{\rho}(\Gamma,\chi_{\lambda}). Then the (𝔤n,Kn,)(\mathfrak{g}_{n},K_{n,\infty})-module generated by ff is a direct sum of modules in 𝒩Repn(χλ,ρ)\mathcal{N}\mathrm{Rep}_{n}(\chi_{\lambda},\rho). Thus, the statement follows from Corollary 6.5.3. ∎

We finally give an analogue of holomorphic projections.

Corollary 7.2.4.

Let λ=(λ1,v,,λn,v)v\lambda=(\lambda_{1,v},\ldots,\lambda_{n,v})_{v} be a regular anti-dominant integral weight with λ1,vλn,v1\lambda_{1,v}-\lambda_{n,v}\leq 1 for any v𝐚v\in\mathbf{a} and ρ\rho the irreducible highest weight representation of Kn,K_{n,\mathbb{C}} with highest weight λ\lambda. If FF\neq\mathbb{Q} or λn,vn+1\lambda_{n,v}\neq n+1 for some v𝐚v\in\mathbf{a}, the projection 𝔭χ\mathfrak{p}_{\chi} defines a projection onto Mρ(Γ)M_{\rho}(\Gamma).

Proof.

By λ1,vλn,v1\lambda_{1,v}-\lambda_{n,v}\leq 1 and Theorem 7.2.3, any modular form ff in Nρ(Γ,χλ)N_{\rho}(\Gamma,\chi_{\lambda}) generates L(λ)L(\lambda). Since ff is of weight λ\lambda, ff corresponds to a highest weight vector. Thus, ff is holomorphic and Nρ(Γ,χλ)=Mρ(Γ)N_{\rho}(\Gamma,\chi_{\lambda})=M_{\rho}(\Gamma). This completes the proof. ∎

References

  • [Del79] P. Deligne. Valeurs de fonctions LL et périodes d’intégrales. In Automorphic forms, representations and LL-functions (Proc. Sympos. Pure Math., Oregon State Univ., Corvallis, Ore., 1977), Part 2, Proc. Sympos. Pure Math., XXXIII, pages 313–346. Amer. Math. Soc., Providence, R.I., 1979. With an appendix by N. Koblitz and A. Ogus.
  • [EHW83] Thomas Enright, Roger Howe, and Nolan Wallach. A classification of unitary highest weight modules. In Representation theory of reductive groups, pages 97–143. Springer, 1983.
  • [Hor20a] Shuji Horinaga. Constructions of nearly holomorphic siegel modular forms of degree two. International Journal of Mathematics, 31(01):2050002, 2020.
  • [Hor20b] Shuji Horinaga. Nearly holomorphic automorphic forms on Sp2n\mathrm{Sp}_{2n} with sufficiently regular infinitesimal characters and applications. to appear in Pasific J. of Math, 2020.
  • [Hor21] Shuji Horinaga. Nearly holomorphic automorphic forms on SL2\mathrm{SL}_{2}. Journal of Number Theory, 219:247–282, 2021.
  • [HPSS21] Shuji Horinaga, Ameya Pitale, Abhishek Saha, and Ralf Schmidt. The special values of the standard ll-functions for GSp2n×GL1\mathrm{GSp}_{2n}\times\mathrm{GL}_{1}. to appear in Trans. of Amer. Math. Soc., 2021.
  • [Hum08] James E Humphreys. Representations of Semisimple Lie Algebras in the BGG Category 𝒪\mathcal{O}, volume 94. American Mathematical Soc., 2008.
  • [Ike92] Tamotsu Ikeda. On the location of poles of the triple ll-functions. Compositio Mathematica, 83(2):187–237, 1992.
  • [KR88] Stephen Kudla and Stephen Rallis. On the weil-siegel formula. J. reine angew. Math, 387(1-68):33, 1988.
  • [KR90] Stephen S Kudla and Stephen Rallis. Degenerate principal series and invariant distributions. Israel Journal of Mathematics, 69(1):25–45, 1990.
  • [KR94] Stephen S Kudla and Stephen Rallis. A regularized siegel-weil formula: the first term identity. Annals of Mathematics, 140(1):1–80, 1994.
  • [Lan06] Robert P Langlands. On the functional equations satisfied by Eisenstein series, volume 544. Springer, 2006.
  • [Liu20] Zheng Liu. pp-adic ll-functions for ordinary families on symplectic groups. Journal of the Institute of Mathematics of Jussieu, 19(4):1287–1347, 2020.
  • [Mau12] Kathrin Maurischat. Casimir operators for symplectic groups. International Journal of Number Theory, 8(04):923–932, 2012.
  • [MW95] Colette Moeglin and Jean-Loup Waldspurger. Spectral decomposition and Eisenstein series: a paraphrase of the scriptures. Number 113. Cambridge University Press, 1995.
  • [PSR87] Ilya Piatetski-Shapiro and Stephen Rallis. Rankin triple ll functions. Compositio Mathematica, 64(1):31–115, 1987.
  • [PSS21] Ameya Pitale, Abhishek Saha, and Ralf Schmidt. Lowest weight modules of Sp4()\mathrm{Sp}_{4}(\mathbb{R}) and nearly holomorphic siegel modular forms. Kyoto Journal of Mathematics, 61(4):745–814, 2021.
  • [Shi82] Goro Shimura. Confluent hypergeometric functions on tube domains. Mathematische Annalen, 260(3):269–302, 1982.
  • [Shi00] Goro Shimura. Arithmeticity in the theory of automorphic forms. Number 82. American Mathematical Soc., 2000.
  • [Yam89] Hiroshi Yamashita. Highest weight vectors for the principal series of semisimple lie groups and embeddings of highest weight modules. Journal of Mathematics of Kyoto University, 29(1):165–173, 1989.