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

Formal series of Jacobi forms

Hiroki Aoki, Tomoyoshi Ibukiyama, Cris Poor Department of Mathematics, Faculty of Science and Technology, Tokyo University of Science, Noda, Chiba, 278-8510 Japan aoki_\_\hskip 1.084pthiroki_\_\hskip 1.084ptmath@nifty.com Department of Mathematics, Graduate School of Mathematics, Osaka University, Machikaneyama 1-1, Toyonaka, Osaka, 560-0043 Japan ibukiyam@math.sci.osaka-u.ac.jp Department of Mathematics, Fordham University, Bronx, NY 10458 USA poor@fordham.edu
Abstract.

We prove for general paramodular level that formal series of scalar Jacobi forms with an involution condition necessarily converge and are therefore the Fourier-Jacobi expansions at the standard 1-cusp of paramodular Fricke eigenforms.

Key words and phrases:
Formal series, Jacobi form, Paramodular form
2020 Mathematics Subject Classification:
Primary: 11F46, 11F50

1. Introduction.

A parmodular form is a Siegel modular form of degree two for the following discrete group

K(N)=Sp(4,)(N1NNNNN)=StabSp(4,)(N).K(N)=\operatorname{Sp}(4,{\mathbb{Q}})\cap\begin{pmatrix}{\mathbb{Z}}&N{\mathbb{Z}}&{\mathbb{Z}}&{\mathbb{Z}}\\ {\mathbb{Z}}&{\mathbb{Z}}&{\mathbb{Z}}&\frac{1}{N}{\mathbb{Z}}\\ {\mathbb{Z}}&N{\mathbb{Z}}&{\mathbb{Z}}&{\mathbb{Z}}\\ N{\mathbb{Z}}&N{\mathbb{Z}}&N{\mathbb{Z}}&{\mathbb{Z}}\end{pmatrix}=\operatorname{Stab}_{\operatorname{Sp}(4,{\mathbb{Q}})}\begin{pmatrix}{\mathbb{Z}}\\ {\mathbb{Z}}\\ {\mathbb{Z}}\\ N{\mathbb{Z}}\end{pmatrix}.

Paramodular forms Mk(K(N))M_{k}\left(K(N)\right) are a natural generalization of elliptic modular forms Mk(Γ0(N))M_{k}\left(\Gamma_{0}(N)\right) and are interesting in many ways. Roberts and Schmidt [35, 36] gave a sophisticated theory of local and global paramodular newforms. Paramodular newforms have applications to modularity; weight two to abelian surfaces [8, 9], and weight three to nonrigid Calabi-Yau threefolds [15] of Hodge type (1,1,1,1)(1,1,1,1). Conjectures of Ibukiyama [22, 23], and of Ibukiyama and Kitayama [25], connecting paramodular forms to algebraic modular forms on a compact twist of GSp(4)\operatorname{GSp}(4), motivated mainly by independent calculations of dimension formulae, have recently been proven [39, 37]. In [10], this connection was generalized to a correspondence with Fricke eigenspaces. Utilizing results of [10], dimension formulae for Fricke plus and minus spaces of paramodular forms for prime level were computed in [24]. For weights k3k\geq 3, these recent proofs allow paramodular Hecke eigensystems to also be computed using orthogonal modular forms [10, 3]. Finally, Gritsenko lifts and Borcherds products provide concrete examples of paramodular forms [19].

Paramodular forms fMk(K(N))f\in M_{k}\left(K(N)\right) have Fourier expansions in three variables but the natural generalization of the Fourier series of an elliptic modular form is perhaps the Fourier-Jacobi expansion of a paramodular form, which, using the components Ω=(τzzω)2\Omega=\left(\begin{smallmatrix}{\tau}&{z}\\ {z}&{\omega}\end{smallmatrix}\right)\in\mathcal{H}_{2}, recollects the Fourier series in powers of e(ω)=e2πiωe(\omega)=e^{2\pi i\omega},

f(τzzω)=m=0ϕm(τ,z)e(Nmω).f\left(\begin{smallmatrix}{\tau}&{z}\\ {z}&{\omega}\end{smallmatrix}\right)=\sum_{m=0}^{\infty}\phi_{m}(\tau,z)e\left(Nm\omega\right).

Each coefficient is a Jacobi form, ϕmJk,Nm\phi_{m}\in J_{k,Nm}, and has its own Fourier expansion ϕm(τ,z)=n0,rc(n,r;ϕm)e(nτ+rz)\phi_{m}(\tau,z)=\sum_{n\geq 0,r\in{\mathbb{Z}}}c(n,r;\phi_{m})e(n\tau+rz). The Fourier-Jacobi expansion thus defines a map to formal series of Jacobi forms FJ:Mk(K(N))𝕄(k,N)=m=0Jk,Nm{\operatorname{FJ}}:M_{k}\left(K(N)\right)\to\mathbb{M}(k,N)=\prod_{m=0}^{\infty}J_{k,Nm}. This map is not surjective because we cannot freely select a sequence of Jacobi forms and obtain the convergent Fourier-Jacobi expansion of a paramodular form. One source of consistency conditions among the Fourier-Jacobi coefficients (ϕm)(\phi_{m}) arises from a normalizing involution of K(N)K(N), the paramodular Fricke involution μN=(FN1t00FN)\mu_{N}=\left(\begin{smallmatrix}{{{}^{t}F_{N}^{-1}}}&{0}\\ {0}&{F_{N}}\end{smallmatrix}\right), where FN=1N(01N0)F_{N}={\tiny\frac{1}{\sqrt{N}}}\left(\begin{smallmatrix}{0}&{1}\\ {-N}&{0}\end{smallmatrix}\right) is the Fricke involution on Γ0(N)\Gamma_{0}(N). This involution splits Mk(K(N))M_{k}\left(K(N)\right) into plus and minus forms, Mk(K(N))=Mk(K(N))+Mk(K(N))M_{k}\left(K(N)\right)=M_{k}\left(K(N)\right)^{+}\oplus M_{k}\left(K(N)\right)^{-}. The block diagonal form of μN\mu_{N} gives a simple action on the Fourier series and consequently gives the following involution condition on the Fourier-Jacobi coefficients of any fMk(K(N))ϵf\in M_{k}\left(K(N)\right)^{\epsilon}, ϵ{±1}\epsilon\in\{\pm 1\}:

(1) For all semidefinite (nr/2r/2Nm)\left(\begin{smallmatrix}{n}&{r/2}\\ {r/2}&{Nm}\end{smallmatrix}\right) with n,r,mn,r,m\in{\mathbb{Z}},
c(n,r;ϕm)=(1)kϵc(m,r;ϕn).\displaystyle\hskip 50.58878ptc(n,r;\phi_{m})=(-1)^{k}\epsilon\,c(m,r;\phi_{n}).

Let 𝕄(k,N,ϵ)\mathbb{M}(k,N,\epsilon) be the subspace of formal series of Jacobi forms satifying the involution condition:

𝕄(k,N,ϵ)={𝔣𝕄(k,N):𝔣 satisfies condition (1)}.\mathbb{M}(k,N,\epsilon)=\{\mathfrak{f}\in\mathbb{M}(k,N):\text{\rm$\mathfrak{f}$ satisfies condition~{}{(\ref{eqinvcondintro})}}\}.

The Fourier-Jacobi expansion gives FJ:Mk(K(N))ϵ𝕄(k,N,ϵ){\operatorname{FJ}}:M_{k}\left(K(N)\right)^{\epsilon}\to\mathbb{M}(k,N,\epsilon). It seems a bit audacious to hope that the involution condition alone forces the convergence of a formal series to a paramodular Fricke eigenform, but theoretical results for low level NN and computed examples suggest that this is true, and we prove it here.

Theorem 1.1 (Main Theorem).

Let NN\in{\mathbb{N}}, k0k\in{\mathbb{N}}_{0}, and ϵ{±1}\epsilon\in\{\pm 1\}. The map FJ:Mk(K(N))ϵ𝕄(k,N,ϵ){\operatorname{FJ}}:M_{k}\left(K(N)\right)^{\epsilon}\to\mathbb{M}(k,N,\epsilon) is an isomorphism.

The case k=0k=0 is not difficult, and neither is the case k=1k=1, when both the domain and codomain are trivial by a result of Skoruppa [38], but k=2k=2 already has important examples. The case N=1N=1 of Theorem 1.1 was proven by the first author [1], the cases 2N42\leq N\leq 4 by Yuen and the second and third authors [26]. These results raised the question, made explicit in [26], of whether the involution condition alone implies the convergence of a formal series of Jacobi forms. Theorem 1.1 resolves this question in the affirmative for the first time.

Bruinier [6], Raum [40], and Bruinier and Raum [7, 5] have approached the theory of formal series of Jacobi forms in a more general setting but with a different set of hypotheses. Bruinier [6] and Raum [40] independently extended Aoki’s result for Sp(4,)\operatorname{Sp}(4,{\mathbb{Z}}) to vector valued formal series of Jacobi forms, and thereby resolved a conjecture of Kudla [27, 28, 29] concerning generating series of special cycles for degree n=2n=2. In [7], Bruinier and Raum proved that formal series of Jacobi forms for Sp(2n,)\operatorname{Sp}(2n,{\mathbb{Z}}) converge under the assumption of all GL(n,){\operatorname{GL}}(n,{\mathbb{Z}}) symmetries, thereby proving Kudla’s modularity conjecture for general nn. Pollack [32] has also given a proof of automatic convergence for cuspidal automorphic forms on Sp(2n)\operatorname{Sp}(2n) that takes Jacobi and Levi symmetries as hypotheses. Bruinier and Raum [5] have recently developed a theory of formal series for subgroups commensurable with Sp(2n,)\operatorname{Sp}(2n,{\mathbb{Z}}); they prove that compatible formal series of Jacobi forms at every 11-cusp and Levi symmetries imply convergence. They interpret formal series as sections of a line bundle over a formal complex space. Our Main Theorem 1.1 for paramodular groups in degree two and levels N>3N>3 is not a consequence of any of these results. We only use formal series of Jacobi forms at the standard 11-cusp and only assume symmetry under a single involution.

Our main result has applications to computations. Jacobi restriction [26, 4, 33] is a method that attempts to rigorously compute the space FJ(Sk(K(N))ϵ){\operatorname{FJ}}\left(S_{k}\left(K(N)\right)^{\epsilon}\right) by imposing necessary linear relations on the space of Jacobi forms m=1dJk,Nmcusp\prod_{m=1}^{d}J_{k,Nm}^{\rm cusp}. Jacobi restriction has provided rigorous upper bounds for dimSk(K(N))ϵ\dim S_{k}\left(K(N)\right)^{\epsilon}, even though the authors could not guarantee in advance that the method would work. The following corollary proves that any schema for computing spaces of paramodular forms that spans spaces of Jacobi forms and imposes the involution condition is in principle sound.

Corollary 1.2.

For dd\in{\mathbb{N}}, define the {\mathbb{C}}-vector space 𝕄(k,N,ϵ)[d]\mathbb{M}(k,N,\epsilon)[d] as {(ϕm)m=0dJk,Nm:(ϕm) satisfies (1) for all n,md }\{(\phi_{m})\in\prod_{m=0}^{d}J_{k,Nm}:\text{$(\phi_{m})$ satisfies~{}{\rm(\ref{eqinvcondintro})} for all $n,m\leq d$ }\}. The sequence dim𝕄(k,N,ϵ)[d]\dim_{{\mathbb{C}}}\mathbb{M}(k,N,\epsilon)[d] is monotonically decreasing for d16Nkd\geq\frac{1}{6}Nk, and we have limd+dim𝕄(k,N,ϵ)[d]=dimMk(K(N))ϵ\lim_{d\to+\infty}\dim_{{\mathbb{C}}}\mathbb{M}(k,N,\epsilon)[d]=\dim_{{\mathbb{C}}}M_{k}\left(K(N)\right)^{\epsilon}.

In particular, we have dimMk(K(N))ϵdim𝕄(k,N,ϵ)[d]\dim_{{\mathbb{C}}}M_{k}\left(K(N)\right)^{\epsilon}\leq\dim_{{\mathbb{C}}}\mathbb{M}(k,N,\epsilon)[d] for d>16Nkd>\frac{1}{6}Nk, and we have equality for sufficiently large dd.

The first author was supported by JSPS KAKENHI Grant Numbers JP19K03429 and JP23K03039. The second author was supported by the following grants: JSPS KAKENHI Grant Number JP19K03424, JP20H00115 and JP23K03031. The authors thank the American Institute of Mathematics for its critical support of this research.

2. Notation.

We denote the natural numbers by ={1,2,3,}{\mathbb{N}}=\{1,2,3,\ldots\} and the whole numbers by 0={0,1,2,}{\mathbb{N}}_{0}=\{0,1,2,\ldots\}. Throughout this article NN\in{\mathbb{N}} denotes a level, k0k\in{\mathbb{N}}_{0} a weight, and ϵ{±1}\epsilon\in\{\pm 1\} a sign. We write σ\sigma^{\prime} for the transpose of a matrix σ\sigma, and σ\sigma^{*} for the transpose inverse. Let t[σ]=σtσt[\sigma]=\sigma^{\prime}t\sigma for compatibly sized matrices, and a,b=tr(ab)\langle a,b\rangle={\operatorname{tr}}(ab) for a,bMn×nsym()a,b\in M_{n\times n}^{\text{\rm sym}}({\mathbb{C}}). For zz\in{\mathbb{C}}, set e(z)=e2πize(z)=e^{2\pi iz}.

For the theory of Jacobi forms see [11]. The upper half plane is 1={τ:Im(τ)>0}\mathcal{H}_{1}=\{\tau\in{\mathbb{C}}:{\rm Im}(\tau)>0\}. For a Jacobi form ϕJk,m\phi\in J_{k,m} of weight kk and index m0m\in{\mathbb{N}}_{0} with ϕ:1×\phi:\mathcal{H}_{1}\times{\mathbb{C}}\to{\mathbb{C}} we write the Fourier expansion as ϕ(τ,z)=n0,rc(n,r;ϕ)qnζr\phi(\tau,z)=\sum_{n\in{\mathbb{N}}_{0},\,r\in{\mathbb{Z}}}c(n,r;\phi)q^{n}\zeta^{r} with q=e(τ)q=e(\tau) and ζ=e(z)\zeta=e(z). The order of a nonzero ϕ\phi is ordϕ=min{n0:r:c(n,r;ϕ)0}{\operatorname{ord}}\phi=\min\{n\in{\mathbb{N}}_{0}:\exists r\in{\mathbb{Z}}:c(n,r;\phi)\neq 0\}. For ν0\nu\in{\mathbb{N}}_{0}, we set Jk,m(ν)={ϕJk,m:ordϕν}J_{k,m}(\nu)=\{\phi\in J_{k,m}:{\operatorname{ord}}\phi\geq\nu\}.

For the theory of Siegel modular forms we refer to [12]. For a ring RR\subseteq{\mathbb{R}} define the positive definite cone with entries in RR by 𝒫n(R)={sMn×nsym(R):s>0}\mathcal{P}_{n}(R)=\{s\in M_{n\times n}^{\text{\rm sym}}(R):s>0\}, and the positive semidefinite cone by 𝒫¯n(R)={sMn×nsym(R):s0}\bar{\mathcal{P}}_{n}(R)=\{s\in M_{n\times n}^{\text{\rm sym}}(R):s\geq 0\}. The Siegel upper half space is n={ΩMn×nsym():Im(Ω)𝒫n()}\mathcal{H}_{n}=\{\Omega\in M_{n\times n}^{\text{\rm sym}}({\mathbb{C}}):{\rm Im}(\Omega)\in\mathcal{P}_{n}({\mathbb{R}})\}. An element σ=(abcd)\sigma=\left(\begin{smallmatrix}{a}&{b}\\ {c}&{d}\end{smallmatrix}\right) in the real symplectic group Sp(2n,)\operatorname{Sp}(2n,{\mathbb{R}}) acts on the symmetric space n\mathcal{H}_{n} by σΩ=(aΩ+b)(cΩ+d)1\sigma\langle\Omega\rangle=(a\Omega+b)(c\Omega+d)^{-1}, and the Siegel factor of automorphy j:Sp(2n,)×n×j:\operatorname{Sp}(2n,{\mathbb{R}})\times\mathcal{H}_{n}\to{\mathbb{C}}^{\times} is j(σ,Ω)=det(cΩ+d)j(\sigma,\Omega)=\det(c\Omega+d). For a function f:nf:\mathcal{H}_{n}\to{\mathbb{C}} the slash kk-action (f|kσ)(Ω)=j(σ,Ω)kf(σΩ)(f|_{k}\sigma)(\Omega)=j(\sigma,\Omega)^{-k}f\left(\sigma\langle\Omega\rangle\right) defines another function f|kσ:nf|_{k}\sigma:\mathcal{H}_{n}\to{\mathbb{C}}. For a discrete group ΓSp(2n,)\Gamma\subseteq\operatorname{Sp}(2n,{\mathbb{R}}) commensurable with Sp(2n,)\operatorname{Sp}(2n,{\mathbb{Z}}), write Mk(Γ)M_{k}\left(\Gamma\right) for the {\mathbb{C}}-vector space of Siegel modular forms of weight kk, and Sk(Γ)S_{k}\left(\Gamma\right) for the subspace of cusp forms.

For the paramodular group K(N)K(N) in degree n=2n=2, write the Fourier expansion for fMk(K(N))f\in M_{k}\left(K(N)\right) as f(Ω)=t𝒳¯(N)a(t;f)e(Ω,t)f(\Omega)=\sum_{t\in{\bar{\mathcal{X}}}(N)}a(t;f)e\left(\langle\Omega,t\rangle\right), where 𝒳¯(N)={(nr/2r/2Nm)𝒫¯2():n,r,m}{\bar{\mathcal{X}}}(N)=\{\left(\begin{smallmatrix}{n}&{r/2}\\ {r/2}&{Nm}\end{smallmatrix}\right)\in{\bar{\mathcal{P}}}_{2}({\mathbb{Q}}):n,r,m\in{\mathbb{Z}}\}. For the Fourier-Jacobi expansion of ff, we write Ω=(τzzω)\Omega=\left(\begin{smallmatrix}{\tau}&{z}\\ {z}&{\omega}\end{smallmatrix}\right) and collect the Fourier expansion in e(ω)e(\omega) to obtain f(τzzω)=m=0ϕm(τ,z)e(Nmω)f\left(\begin{smallmatrix}{\tau}&{z}\\ {z}&{\omega}\end{smallmatrix}\right)=\sum_{m=0}^{\infty}\phi_{m}(\tau,z)e(Nm\omega) with Fourier-Jacobi coeffcients ϕmJk,Nm\phi_{m}\in J_{k,Nm}. The paramodular group has a normalizing involution, the paramodular Fricke involution, given by μN=(FN1t00FN)Sp(4,)\mu_{N}=\left(\begin{smallmatrix}{{{}^{t}F_{N}^{-1}}}&{0}\\ {0}&{F_{N}}\end{smallmatrix}\right)\in\operatorname{Sp}(4,{\mathbb{R}}), with FN=1N(01N0)SL(2,)F_{N}=\frac{1}{\sqrt{N}}\left(\begin{smallmatrix}{0}&{1}\\ {-N}&{0}\end{smallmatrix}\right)\in{\operatorname{SL}}(2,{\mathbb{R}}). Define the Fricke paramodular group K(N)+=K(N),μNK(N)^{+}=\langle K(N),\mu_{N}\rangle. For ϵ{±1}\epsilon\in\{\pm 1\}, let Mk(K(N))ϵ={fMk(K(N)):f|kμN=ϵf}M_{k}\left(K(N)\right)^{\epsilon}=\{f\in M_{k}\left(K(N)\right):f|_{k}\mu_{N}=\epsilon f\} be the Fricke eigenspaces. There are graded rings, M(K(N))=k0Mk(K(N))M\left(K(N)\right)=\oplus_{k\in{\mathbb{N}}_{0}}M_{k}\left(K(N)\right), M±(K(N))=k0,ϵ{±1}Mk(K(N))ϵM_{\pm}\left(K(N)\right)=\oplus_{k\in{\mathbb{N}}_{0},\epsilon\in\{\pm 1\}}M_{k}\left(K(N)\right)^{\epsilon}, as well as the graded ring of Fricke plus forms M(K(N)+)=k0Mk(K(N)+)M\left(K(N)^{+}\right)=\oplus_{k\in{\mathbb{N}}_{0}}M_{k}\left(K(N)^{+}\right).

\bullet N(η)={τ1:Im(τ)>η}N_{\infty}(\eta)=\{\tau\in\mathcal{H}_{1}:{\rm Im}(\tau)>\eta\} for η0\eta\geq 0.

\bullet P2,1()=Sp(4,)(00000).\operatorname{P}_{2,1}({\mathbb{Z}})=\operatorname{Sp}(4,{\mathbb{Z}})\cap\begin{pmatrix}{\mathbb{Z}}&0&{\mathbb{Z}}&{\mathbb{Z}}\\ {\mathbb{Z}}&{\mathbb{Z}}&{\mathbb{Z}}&{\mathbb{Z}}\\ {\mathbb{Z}}&0&{\mathbb{Z}}&{\mathbb{Z}}\\ 0&0&0&{\mathbb{Z}}\end{pmatrix}.

3. Formal Series.

A paramodular form fMk(K(N))f\in M_{k}\left(K(N)\right) has a Fourier series f(Ω)=t𝒳¯(N)a(t;f)e(Ω,t)f(\Omega)=\sum_{t\in{\bar{\mathcal{X}}}(N)}a(t;f)e\left(\langle\Omega,t\rangle\right) that converges absolutely and uniformly on compact subsets of 2\mathcal{H}_{2}. The absolute convergence allows rearrangement into a Fourier-Jacobi series f(Ω)=m=0ϕm(τ,z)e(Nmω)f(\Omega)=\sum_{m=0}^{\infty}\phi_{m}(\tau,z)e\left(Nm\omega\right) where each Fourier-Jacobi coefficient is a Jacobi form ϕmJk,Nm\phi_{m}\in J_{k,Nm}. This follows from the fact that the Fourier-Jacobi expansion is term-by-term invariant under P2,1()K(N)\operatorname{P}_{2,1}({\mathbb{Z}})\subseteq K(N), see [18]. We define the formal Fourier-Jacobi series FJ(f){\operatorname{FJ}}(f) of a paramodular form ff by

FJ:Mk(K(N))\displaystyle{\operatorname{FJ}}:M_{k}\left(K(N)\right) 𝕄(k,N)=m=0Jk,Nm\displaystyle\to\mathbb{M}(k,N)=\prod_{m=0}^{\infty}J_{k,Nm}
f\displaystyle f m=0ϕmξNm,\displaystyle\mapsto\sum_{m=0}^{\infty}\phi_{m}\xi^{Nm},

where ξ\xi is a place-holding variable. The Fourier coefficients of a paramodular form also satisfy symmetries determined by the following group:

{σGL(2,):(σ00σ)K(N)}\displaystyle\{\sigma\in{\operatorname{GL}}(2,{\mathbb{R}}):\left(\begin{smallmatrix}{\sigma}&{0}\\ {0}&{\sigma^{*}}\end{smallmatrix}\right)\in K(N)\} =Γ0(N),diag(1,1)=Γ±0(N).\displaystyle=\langle\Gamma^{0}(N),\operatorname{diag}(1,-1)\rangle=\Gamma^{0}_{\pm}(N).

We call equation (2) below the Γ0(N)\Gamma^{0}(N)-symmetries,

(2) σΓ±0(N),t𝒳¯(N),a(t[σ];f)=det(σ)ka(t;f).\forall\sigma\in\Gamma^{0}_{\pm}(N),\forall t\in{\bar{\mathcal{X}}}(N),\,a\left(t[\sigma];f\right)=\det(\sigma)^{k}a\left(t;f\right).

The Fourier coefficients of ff are related to the Fourier coefficients of the Jacobi forms ϕm\phi_{m} in the Fourier-Jacobi expansion of ff by

a((nr/2r/2Nm);f)=c(n,r;ϕm).a\left(\left(\begin{smallmatrix}{n}&{r/2}\\ {r/2}&{Nm}\end{smallmatrix}\right);f\right)=c(n,r;\phi_{m}).

Accordingly, we define a subspace of 𝕄(k,N)\mathbb{M}(k,N) that satisfies corresponding symmetries. Let ΓΓ±0(N)\Gamma\subseteq\Gamma^{0}_{\pm}(N) be a subgroup.

𝕄(k,N;Γ)\displaystyle\mathbb{M}(k,N;\Gamma) ={𝔣=m=0ϕmξNm𝕄(k,N):𝔣 satisfies equation (3)},\displaystyle=\{\mathfrak{f}=\sum_{m=0}^{\infty}\phi_{m}\xi^{Nm}\in\mathbb{M}(k,N):\text{\rm$\mathfrak{f}$ satisfies equation~{}{(\ref{eqgammaoNsymformal})}}\},
(3) σΓ,t1=(n1r1/2r1/2Nm1),t2=(n2r2/2r2/2Nm2)𝒳¯(N),\displaystyle\forall\sigma\in\Gamma,\forall t_{1}=\left(\begin{smallmatrix}{n_{1}}&{r_{1}/2}\\ {r_{1}/2}&{Nm_{1}}\end{smallmatrix}\right),t_{2}=\left(\begin{smallmatrix}{n_{2}}&{r_{2}/2}\\ {r_{2}/2}&{Nm_{2}}\end{smallmatrix}\right)\in{\bar{\mathcal{X}}}(N),
if t1[σ]=t2 then c(n2,r2;ϕm2)=det(σ)kc(n1,r1;ϕm1).\displaystyle\text{if }t_{1}[\sigma]=t_{2}\text{ then }c(n_{2},r_{2};\phi_{m_{2}})=\det(\sigma)^{k}c(n_{1},r_{1};\phi_{m_{1}}).

Thus we have FJ:Mk(K(N))𝕄(k,N;Γ){\operatorname{FJ}}:M_{k}\left(K(N)\right)\to\mathbb{M}(k,N;\Gamma).

We write ϕm=ϕm(𝔣)\phi_{m}=\phi_{m}(\mathfrak{f}) if we need to indicate the dependence of this Jacobi form on the formal series 𝔣\mathfrak{f}. It is often helpful to reformulate the definition of 𝕄(k,N;Γ)\mathbb{M}(k,N;\Gamma) by using the notation a(t;𝔣)=c(n,r;ϕm(𝔣))a(t;\mathfrak{f})=c(n,r;\phi_{m}(\mathfrak{f})) for t=(nr/2r/2Nm)𝒳¯(N)t=\left(\begin{smallmatrix}{n}&{r/2}\\ {r/2}&{Nm}\end{smallmatrix}\right)\in{\bar{\mathcal{X}}}(N). In this way equation (3) shortens to

(4) σΓ,t𝒳¯(N),a(t[σ];𝔣)=det(σ)ka(t;𝔣).\forall\sigma\in\Gamma,\forall t\in{\bar{\mathcal{X}}}(N),\,a\left(t[\sigma];\mathfrak{f}\right)=\det(\sigma)^{k}a\left(t;\mathfrak{f}\right).

We may sometimes avoid tracking boundary conditions in summations by setting a(t;𝔣)=0a\left(t;\mathfrak{f}\right)=0 for tM2×2sym()𝒳¯(N)t\in M_{2\times 2}^{\rm sym}({\mathbb{Q}})\setminus{\bar{\mathcal{X}}}(N).

Lemma 3.1.

Let ΓΓ±0(N)\Gamma\subseteq\Gamma^{0}_{\pm}(N) be a subgroup. The Cauchy product gives 𝕄(N)=k=0𝕄(k,N)\mathbb{M}(N)=\oplus_{k=0}^{\infty}\mathbb{M}(k,N) the structure of a graded ring, and 𝕄(N;Γ)=k=0𝕄(k,N;Γ)\mathbb{M}(N;\Gamma)=\oplus_{k=0}^{\infty}\mathbb{M}(k,N;\Gamma) a graded subring. In particular, 𝕄(k1,N)𝕄(k2,N)𝕄(k1+k2,N)\mathbb{M}(k_{1},N)\mathbb{M}(k_{2},N)\subseteq\mathbb{M}(k_{1}+k_{2},N), and 𝕄(k1,N;Γ)𝕄(k2,N;Γ)𝕄(k1+k2,N;Γ)\mathbb{M}(k_{1},N;\Gamma)\mathbb{M}(k_{2},N;\Gamma)\subseteq\mathbb{M}(k_{1}+k_{2},N;\Gamma). The sets 𝕊(N)=k=0m=1Jk,Nmcusp\mathbb{S}(N)=\oplus_{k=0}^{\infty}\prod_{m=1}^{\infty}J_{k,Nm}^{\rm cusp}, and 𝕊(N;Γ)=𝕄(N;Γ)𝕊(N)\mathbb{S}(N;\Gamma)=\mathbb{M}(N;\Gamma)\cap\mathbb{S}(N) are graded ideals in 𝕄(N)\mathbb{M}(N), and 𝕄(N;Γ)\mathbb{M}(N;\Gamma), respectively. The Fourier-Jacobi map FJ:M(K(N))𝕄(N;Γ){\operatorname{FJ}}:M\left(K(N)\right)\to\mathbb{M}(N;\Gamma) is an injective ring homomorphism of graded rings that sends S(K(N))S\left(K(N)\right) to 𝕊(N;Γ)\mathbb{S}(N;\Gamma), and the inverse image of 𝕊(N;Γ)\mathbb{S}(N;\Gamma) is S(K(N))S\left(K(N)\right).

Proof.

When the formal series 𝔣j𝕄(kj,N)\mathfrak{f}_{j}\in\mathbb{M}(k_{j},N), for j=1,2j=1,2, are written as 𝔣j=m=0ϕm(𝔣j)ξNm\mathfrak{f}_{j}=\sum_{m=0}^{\infty}\phi_{m}(\mathfrak{f}_{j})\xi^{Nm} then by the definition of the Cauchy product we have 𝔣1𝔣2=m=0ϕm(𝔣1𝔣2)ξNm\mathfrak{f}_{1}\mathfrak{f}_{2}=\sum_{m=0}^{\infty}\phi_{m}(\mathfrak{f}_{1}\mathfrak{f}_{2})\xi^{Nm} where

ϕm(𝔣1𝔣2)=m1,m20:m1+m2=mϕm1(𝔣1)ϕm2(𝔣2).\phi_{m}(\mathfrak{f}_{1}\mathfrak{f}_{2})=\sum_{m_{1},m_{2}\in{\mathbb{N}}_{0}:\,m_{1}+m_{2}=m}\phi_{m_{1}}(\mathfrak{f}_{1})\phi_{m_{2}}(\mathfrak{f}_{2}).

Thus 𝕄(k1,N)𝕄(k2,N)𝕄(k1+k2,N)\mathbb{M}(k_{1},N)\mathbb{M}(k_{2},N)\subseteq\mathbb{M}(k_{1}+k_{2},N) follows from the grading Jk1,Nm1Jk2,Nm2Jk1+k2,N(m1+m2)J_{k_{1},Nm_{1}}J_{k_{2},Nm_{2}}\subseteq J_{k_{1}+k_{2},N(m_{1}+m_{2})} on Jacobi forms, which shows that 𝕄(N)\mathbb{M}(N) is a graded ring. To derive a similar result for 𝕄(N;Γ)\mathbb{M}(N;\Gamma) we reformulate the Cauchy product on 𝕄(N)\mathbb{M}(N) as

(5) t𝒳¯(N),a(t;𝔣1𝔣2)=t1,t2𝒳¯(N):t1+t2=ta(t1;𝔣1)a(t2;𝔣2).\forall t\in{\bar{\mathcal{X}}}(N),\,a(t;\mathfrak{f}_{1}\mathfrak{f}_{2})=\sum_{t_{1},t_{2}\in{\bar{\mathcal{X}}}(N):\,t_{1}+t_{2}=t}a(t_{1};\mathfrak{f}_{1})a(t_{2};\mathfrak{f}_{2}).

To see that (5) follows from the Cauchy product, note the following four equalities.

a(t;𝔣1𝔣2)=c(n,r;ϕm(𝔣1𝔣2))=m1+m2=mc(n,r;ϕm1(𝔣1)ϕm2(𝔣2))\displaystyle a(t;\mathfrak{f}_{1}\mathfrak{f}_{2})=c(n,r;\phi_{m}\left(\mathfrak{f}_{1}\mathfrak{f}_{2})\right)=\sum_{m_{1}+m_{2}=m}c\left(n,r;\phi_{m_{1}}(\mathfrak{f}_{1})\phi_{m_{2}}(\mathfrak{f}_{2})\right)
=\displaystyle= m1,m20:m1+m2=mn1,n20:n1+n2=nr1,r2:r1+r2=rc(n1,r1;ϕm1(𝔣1))c(n2,r2;ϕm2(𝔣2))\displaystyle\sum_{\begin{subarray}{c}m_{1},m_{2}\in{\mathbb{N}}_{0}:\\ m_{1}+m_{2}=m\end{subarray}}\,\sum_{\begin{subarray}{c}n_{1},n_{2}\in{\mathbb{N}}_{0}:\\ n_{1}+n_{2}=n\end{subarray}}\,\sum_{\begin{subarray}{c}r_{1},r_{2}\in{\mathbb{Z}}:\\ r_{1}+r_{2}=r\end{subarray}}c\left(n_{1},r_{1};\phi_{m_{1}}(\mathfrak{f}_{1})\right)c\left(n_{2},r_{2};\phi_{m_{2}}(\mathfrak{f}_{2})\right)
=\displaystyle= t1+t2=ta(t1;𝔣1)a(t2;𝔣2).\displaystyle\sum_{t_{1}+t_{2}=t}a(t_{1};\mathfrak{f}_{1})a(t_{2};\mathfrak{f}_{2}).

Conversely, equation (5) implies the second equality, which is equivalent to the Cauchy product. If we assume 𝔣j𝕄(kj,N;Γ)\mathfrak{f}_{j}\in\mathbb{M}(k_{j},N;\Gamma) for j=1,2j=1,2, then, for any σΓ\sigma\in\Gamma, a(t[σ];𝔣1𝔣2)=t1+t2=t[σ]a(t1;𝔣1)a(t2;𝔣2)a(t[\sigma];\mathfrak{f}_{1}\mathfrak{f}_{2})=\sum_{t_{1}+t_{2}=t[\sigma]}a(t_{1};\mathfrak{f}_{1})a(t_{2};\mathfrak{f}_{2}). We use {(t1,t2)𝒳¯(N)2:t1+t2=t[σ]}={(s1[σ],s2[σ])𝒳¯(N)2:s1+s2=t}\{(t_{1},t_{2})\in{\bar{\mathcal{X}}}(N)^{2}:t_{1}{+}t_{2}{=}t[\sigma]\}=\{(s_{1}[\sigma],s_{2}[\sigma])\in{\bar{\mathcal{X}}}(N)^{2}:s_{1}{+}s_{2}{=}t\} to change the indices of summation,

a(t[σ];𝔣1𝔣2)\displaystyle a(t[\sigma];\mathfrak{f}_{1}\mathfrak{f}_{2}) =s1+s2=ta(s1[σ];𝔣1)a(s2[σ];𝔣2)\displaystyle=\sum_{s_{1}+s_{2}=t}a(s_{1}[\sigma];\mathfrak{f}_{1})a(s_{2}[\sigma];\mathfrak{f}_{2})
=s1+s2=tdet(σ)k1a(s1;𝔣1)det(σ)k2a(s2;𝔣2)\displaystyle=\sum_{s_{1}+s_{2}=t}\det(\sigma)^{k_{1}}a(s_{1};\mathfrak{f}_{1})\det(\sigma)^{k_{2}}a(s_{2};\mathfrak{f}_{2})
=det(σ)k1+k2a(t;𝔣1𝔣2).\displaystyle=\det(\sigma)^{k_{1}+k_{2}}a(t;\mathfrak{f}_{1}\mathfrak{f}_{2}).

Thus 𝕄(k1,N;Γ)𝕄(k2,N;Γ)𝕄(k1+k2,N;Γ)\mathbb{M}(k_{1},N;\Gamma)\mathbb{M}(k_{2},N;\Gamma)\subseteq\mathbb{M}(k_{1}+k_{2},N;\Gamma) and 𝕄(N;Γ)\mathbb{M}(N;\Gamma) is a graded ring, noting that a sum of paramodular forms of distinct weights is zero as a holomorphic function if and only if each summand is zero as a holomorphic function. The sets 𝕊(N)\mathbb{S}(N), and 𝕊(N;Γ)\mathbb{S}(N;\Gamma) are graded ideals simply because Jk1,Nm1cuspJk2,Nm2Jk1+k2,N(m1+m2)cuspJ_{k_{1},Nm_{1}}^{\rm cusp}J_{k_{2},Nm_{2}}\subseteq J_{k_{1}+k_{2},N(m_{1}+m_{2})}^{\text{\rm cusp}}. The convergence of the Fourier-Jacobi expansion shows that FJ{\operatorname{FJ}} is injective on each graded piece and hence on M(K(N))M\left(K(N)\right). The absolute convergence of the Fourier-Jacobi expansion for fjMkj(K(N))f_{j}\in M_{k_{j}}\left(K(N)\right) proves that FJ:M(K(N))𝕄(N;Γ){\operatorname{FJ}}:M\left(K(N)\right)\to\mathbb{M}(N;\Gamma) is a homomorphism because for Ω=(τzzω)2\Omega=\left(\begin{smallmatrix}{\tau}&{z}\\ {z}&{\omega}\end{smallmatrix}\right)\in\mathcal{H}_{2},

f1(Ω)f2(Ω)=\displaystyle f_{1}(\Omega)f_{2}(\Omega)=
m=0(m1+m2=mϕm1(FJ(f1))(τ,z)ϕm2(FJ(f2))(τ,z))e(Nmω),\displaystyle\sum_{m=0}^{\infty}\left(\sum_{m_{1}+m_{2}=m}\phi_{m_{1}}({\operatorname{FJ}}(f_{1}))(\tau,z)\,\phi_{m_{2}}({\operatorname{FJ}}(f_{2}))(\tau,z)\right)e(Nm\omega),

and therefore

FJ(f1f2)\displaystyle{\operatorname{FJ}}(f_{1}f_{2}) =m=0(m1+m2=mϕm1(FJ(f1))ϕm2(FJ(f2)))ξNm\displaystyle=\sum_{m=0}^{\infty}\left(\sum_{m_{1}+m_{2}=m}\phi_{m_{1}}({\operatorname{FJ}}(f_{1}))\,\phi_{m_{2}}({\operatorname{FJ}}(f_{2}))\right)\xi^{Nm}
=FJ(f1)FJ(f2).\displaystyle={\operatorname{FJ}}(f_{1}){\operatorname{FJ}}(f_{2}).

For a cusp form fS(K(N))f\in S\left(K(N)\right), the Siegel Φ\Phi map gives the containment supp(f)={t𝒳¯(N):a(t;f)0}𝒳(N)\operatorname{supp}(f)=\{t\in{\bar{\mathcal{X}}}(N):a(t;f)\neq 0\}\subseteq{\mathcal{X}}(N). By examining the support of ϕm\phi_{m},

supp(ϕm)\displaystyle\operatorname{supp}(\phi_{m}) ={(n,r)0×:c(n,r;ϕm)0}\displaystyle=\{(n,r)\in{\mathbb{N}}_{0}\times{\mathbb{Z}}:c(n,r;\phi_{m})\neq 0\}
={(n,r)0×:a((nr/2r/2Nm);f)0}\displaystyle=\{(n,r)\in{\mathbb{N}}_{0}\times{\mathbb{Z}}:a\left(\left(\begin{smallmatrix}{n}&{r/2}\\ {r/2}&{Nm}\end{smallmatrix}\right);f\right)\neq 0\}
{(n,r)×:4Nmnr2>0},\displaystyle\subseteq\{(n,r)\in{\mathbb{N}}\times{\mathbb{Z}}:4Nmn-r^{2}>0\},

we see that ϕmJk,Nmcusp\phi_{m}\in J_{k,Nm}^{\text{\rm cusp}}, and that FJ(f)𝕊(N){\operatorname{FJ}}(f)\in\mathbb{S}(N), from which FJ(f)𝕊(N;Γ){\operatorname{FJ}}(f)\in\mathbb{S}(N;\Gamma) follows. The assertion that {fMk(K(N)):FJ(f)𝕊(N;Γ)}=Sk(K(N))\{f\in M_{k}\left(K(N)\right):{\operatorname{FJ}}(f)\in\mathbb{S}(N;\Gamma)\}=S_{k}\left(K(N)\right) relies on special properties of the paramodular group. For general subgroups commensurable with Sp(4,)\operatorname{Sp}(4,{\mathbb{Z}}) it is not true that the Fourier-Jacobi expansion of a Siegel modular form with coefficients that are all cusp forms necessarily comes from a Siegel modular cusp form; however, the inference is valid for paramodular forms because Reefschläger’s double coset decomposition [34] of K(N)\Sp(4,)/P2,1()K(N)\backslash\operatorname{Sp}(4,{\mathbb{Q}})/\operatorname{P}_{2,1}({\mathbb{Q}}) has representatives of the form (u00u)\left(\begin{smallmatrix}{u}&{0}\\ {0}&{u^{*}}\end{smallmatrix}\right) with uGL(2,)u\in{\operatorname{GL}}(2,{\mathbb{Q}}); see also Corollary 2.5 of Gritsenko [18]. ∎

The involution μN=(FN00FN)\mu_{N}=\left(\begin{smallmatrix}{F_{N}^{*}}&{0}\\ {0}&{F_{N}}\end{smallmatrix}\right) decomposes paramodular forms into plus and minus forms Mk(K(N))=Mk(K(N))+Mk(K(N))M_{k}\left(K(N)\right)=M_{k}\left(K(N)\right)^{+}\oplus M_{k}\left(K(N)\right)^{-} where Mk(K(N))ϵ={fMk(K(N)):f|kμN=ϵf}M_{k}\left(K(N)\right)^{\epsilon}=\{f\in M_{k}\left(K(N)\right):f|_{k}\mu_{N}=\epsilon f\}. The action of FNF_{N}^{*} on 𝒳¯(N){\bar{\mathcal{X}}}(N) is FN1tFN=(mr/2r/2Nn)F_{N}^{-1}tF_{N}^{*}=\left(\begin{smallmatrix}{m}&{-r/2}\\ {-r/2}&{Nn}\end{smallmatrix}\right) if t=(nr/2r/2Nm)𝒳¯(N)t=\left(\begin{smallmatrix}{n}&{r/2}\\ {r/2}&{Nm}\end{smallmatrix}\right)\in{\bar{\mathcal{X}}}(N). Therefore the Fourier coefficients of a Fricke eigenform fMk(K(N))ϵf\in M_{k}\left(K(N)\right)^{\epsilon} satisfy the involution condition

(6) t𝒳¯(N),a(FN1tFN;f)=ϵa(t;f).\forall t\in{\bar{\mathcal{X}}}(N),\,a\left(F_{N}^{-1}tF_{N}^{*};f\right)=\epsilon\,a\left(t;f\right).

Accordingly, we define a subspace of formal series satisfying the corresponding involution condition

𝕄(k,N,ϵ)\displaystyle\mathbb{M}(k,N,\epsilon) ={𝔣=m=0ϕmξNm𝕄(k,N):𝔣 satisfies equation (7)},\displaystyle=\{\mathfrak{f}=\sum_{m=0}^{\infty}\phi_{m}\xi^{Nm}\in\mathbb{M}(k,N):\text{\rm$\mathfrak{f}$ satisfies equation~{}{(\ref{eqinvcondformal})}}\},
(7) (nr/2r/2Nm)𝒳¯(N),c(n,r;ϕm)=(1)kϵc(m,r;ϕn).\displaystyle\forall\left(\begin{smallmatrix}{n}&{r/2}\\ {r/2}&{Nm}\end{smallmatrix}\right)\in{\bar{\mathcal{X}}}(N),\,c(n,r;\phi_{m})=(-1)^{k}\epsilon\,c(m,r;\phi_{n}).

The map FJ:Mk(K(N))ϵ𝕄(k,N,ϵ){\operatorname{FJ}}:M_{k}\left(K(N)\right)^{\epsilon}\to\mathbb{M}(k,N,\epsilon) injects. It is helpful to rewrite equation (7) as t𝒳¯(N),a(t[FN];𝔣)=(1)kϵa(t;𝔣)\forall t\in{\bar{\mathcal{X}}}(N),\,a\left(t[F_{N}^{*}];\mathfrak{f}\right)=(-1)^{k}\epsilon\,a\left(t;\mathfrak{f}\right). If we write 𝕄(k,N,ϵ;Γ)=𝕄(k,N;Γ)𝕄(k,N,ϵ)\mathbb{M}(k,N,\epsilon;\Gamma)=\mathbb{M}(k,N;\Gamma)\cap\mathbb{M}(k,N,\epsilon) we also have FJ:Mk(K(N))ϵ𝕄(k,N,ϵ;Γ){\operatorname{FJ}}:M_{k}\left(K(N)\right)^{\epsilon}\to\mathbb{M}(k,N,\epsilon;\Gamma). We set 𝕊(k,N,ϵ)=𝕄(k,N,ϵ)𝕊(k,N)\mathbb{S}(k,N,\epsilon)=\mathbb{M}(k,N,\epsilon)\cap\mathbb{S}(k,N), and the more frequently used 𝕊(k,N,ϵ;Γ)=𝕄(k,N,ϵ;Γ)𝕊(k,N)\mathbb{S}(k,N,\epsilon;\Gamma)=\mathbb{M}(k,N,\epsilon;\Gamma)\cap\mathbb{S}(k,N).

Lemma 3.2.

Let ΓΓ±0(N)\Gamma\subseteq\Gamma^{0}_{\pm}(N) be a subgroup. The Cauchy product gives 𝕄±(N)=k,ϵ𝕄(k,N,ϵ){\mathbb{M}}_{\pm}(N)=\oplus_{k,\epsilon}\mathbb{M}(k,N,\epsilon), 𝕄±(N;Γ)=k,ϵ𝕄(k,N,ϵ;Γ){\mathbb{M}}_{\pm}(N;\Gamma)=\oplus_{k,\epsilon}\mathbb{M}(k,N,\epsilon;\Gamma), and 𝕄(N,+)=k=0𝕄(k,N,+)\mathbb{M}(N,+)=\oplus_{k=0}^{\infty}\mathbb{M}(k,N,+) the structure of graded rings. In particular, we have 𝕄(k1,N,ϵ1)𝕄(k2,N,ϵ2)𝕄(k1+k2,N,ϵ1ϵ2)\mathbb{M}(k_{1},N,\epsilon_{1})\mathbb{M}(k_{2},N,\epsilon_{2})\subseteq\mathbb{M}(k_{1}+k_{2},N,\epsilon_{1}\epsilon_{2}), and 𝕄(k1,N,ϵ1;Γ)𝕄(k2,N,ϵ2;Γ)𝕄(k1+k2,N,ϵ1ϵ2;Γ)\mathbb{M}(k_{1},N,\epsilon_{1};\Gamma)\mathbb{M}(k_{2},N,\epsilon_{2};\Gamma)\subseteq\mathbb{M}(k_{1}+k_{2},N,\epsilon_{1}\epsilon_{2};\Gamma). The subsets 𝕊±(N)=k,ϵ𝕊(k,N,ϵ){\mathbb{S}}_{\pm}(N)=\oplus_{k,\epsilon}\mathbb{S}(k,N,\epsilon), and 𝕊±(N;Γ)=𝕄±(N;Γ)𝕊±(N){\mathbb{S}}_{\pm}(N;\Gamma)={\mathbb{M}}_{\pm}(N;\Gamma)\cap{\mathbb{S}}_{\pm}(N) are graded ideals in 𝕄±(N){\mathbb{M}}_{\pm}(N) and 𝕄±(N;Γ){\mathbb{M}}_{\pm}(N;\Gamma), respectively. The Fourier-Jacobi expansion map FJ:M±(K(N))𝕄±(N;Γ){\operatorname{FJ}}:M_{\pm}\left(K(N)\right)\to{\mathbb{M}}_{\pm}(N;\Gamma) is an injective homomorphism of graded rings that sends S±(K(N))S_{\pm}\left(K(N)\right) to 𝕊±(N;Γ){\mathbb{S}}_{\pm}(N;\Gamma), and the inverse image of 𝕊±(N;Γ){\mathbb{S}}_{\pm}(N;\Gamma) is S±(K(N))S_{\pm}\left(K(N)\right).

Proof.

The proof is very similar to that of Lemma 3.1. ∎

Lemma 3.3.

Let NN\in{\mathbb{N}}. The graded ring of formal series of Jacobi forms, 𝕄(N)=k=0𝕄(k,N)\mathbb{M}(N)=\oplus_{k=0}^{\infty}\mathbb{M}(k,N), is an integral domain.

Proof.

An element in 𝕄(N)\mathbb{M}(N) is zero precisely when each graded piece is zero. If a product of nonzero elements from 𝕄(N)\mathbb{M}(N) is zero then the product of the nonzero graded pieces of highest weight in each factor must be zero. Each nonzero element in each graded piece has a leading term ϕmJk,Nm𝒪(1×)\phi_{m}\in J_{k,Nm}\subseteq{\mathcal{O}}(\mathcal{H}_{1}\times{\mathbb{C}}). The product of these leading terms cannot be zero because the ring 𝒪(1×){\mathcal{O}}(\mathcal{H}_{1}\times{\mathbb{C}}) is an integral domain. ∎

4. Vanishing Order of Jacobi Forms.

In Corollary 3.1 of [2], Aoki significantly improved the known bounds on vanishing orders of Jacobi forms [11, 16], which were O(m)O(m) in the index mm for a fixed weight kk. He proved that a nontrivial Jacobi form of weight kk and index mm cannot have a vanishing order ν\nu greater than k+16(2m+1+1)\frac{k+1}{6}\left(\sqrt{2m+1}+1\right). As a consequence the {\mathbb{C}}-vector spaces of formal series 𝕄(k,N,ϵ)\mathbb{M}(k,N,\epsilon) are finite dimensional. In Theorem 4.3 we state Aoki’s main theorem from [2] in the case of even weights and use it to improve the estimate concerning when Jk,Nm(m)={0}J_{k,Nm}(m)=\{0\} enough to get the desired bound on the growth in kk for dimensions of spaces of formal series of Jacobi forms, namely, dim𝕄(k,N,ϵ)O(N3k3)\dim\mathbb{M}(k,N,\epsilon)\in O(N^{3}k^{3}) for N,kN,k\in{\mathbb{N}}.

Definition 4.1.

For jj\in{\mathbb{N}} we define ψj:+\psi_{j}:{\mathbb{N}}\to{\mathbb{Q}}^{+} by

ψj(u)={1if u=1,pu:p prime (11pj)if u2.\psi_{j}(u)=\begin{cases}1&\text{if $u=1$,}\\ {\prod}_{\begin{subarray}{c}p\mid u:\,\,\text{\rm$p$ prime }\end{subarray}}\left(1-\frac{1}{p^{j}}\right)&\text{if $u\geq 2$.}\end{cases}

It is clear that ata\mid t implies ψj(a)ψj(t)\psi_{j}(a)\geq\psi_{j}(t).

Definition 4.2.

We define ψ:{1}+\psi:{\mathbb{N}}\setminus\{1\}\to{\mathbb{Q}}^{+} by

ψ(t)={t2ψ2(t)=3if t=2,12t2ψ2(t)=12t2pt:p prime (11p2)if t3.\psi(t)=\begin{cases}t^{2}\psi_{2}(t)=3&\text{if $t=2$,}\\ \frac{1}{2}t^{2}\psi_{2}(t)=\frac{1}{2}t^{2}{\prod}_{\begin{subarray}{c}p\mid t:\,\,\text{\rm$p$ prime }\end{subarray}}\left(1-\frac{1}{p^{2}}\right)&\text{if $t\geq 3$.}\end{cases}

It is easy to check that ψ(t)3\psi(t)\geq 3 for t2t\geq 2.

Theorem 4.3 ([2]).

Let k2,mk\in 2{\mathbb{N}},m\in{\mathbb{N}}. Let ϕJk,m\phi\in J_{k,m} have ord(ϕ)=μ{\operatorname{ord}}(\phi)=\mu. We have

min(m,m6μ+k2)t:()ψ(t),\min\left(m,m-6\mu+\frac{k}{2}\right)\geq\sum_{t:(\spadesuit)}\psi(t),

where tt runs over all natural numbers satisfying the condition

t1 and c=0t1ψ1(gcd(t,c))max(μmc(tc)t2,0)>kt12ψ2(t).()t\neq 1\text{ and }\sum_{c=0}^{t-1}\psi_{1}\left(\gcd(t,c)\right)\max\left(\mu-\frac{mc(t-c)}{t^{2}},0\right)>\frac{kt}{12}\psi_{2}(t).\,\,(\spadesuit)
Lemma 4.4.

Let t,μ,mt,\mu,m\in{\mathbb{N}}. If μ<m\mu<m, we have

c=0t1max(μmc(tc)t2,0)>μ2t2m.\sum_{c=0}^{t-1}\max\left(\mu-\frac{mc(t-c)}{t^{2}},0\right)>\frac{\mu^{2}t}{2m}.
Proof.

The condition μ<m\mu<m is used in the first equality.

c=0t1max(μmc(tc)t2,0)c=0t1max(μmct,0)\displaystyle\sum_{c=0}^{t-1}\max\left(\mu-\frac{mc(t-c)}{t^{2}},0\right)\geq\sum_{c=0}^{t-1}\max\left(\mu-\frac{mc}{t},0\right)
=\displaystyle= c=0floor(μtm)(μmct)=(μm2tfloor(μtm))(floor(μtm)+1)\displaystyle\sum_{c=0}^{\operatorname{floor}(\frac{\mu t}{m})}\left(\mu-\frac{mc}{t}\right)=\left(\mu-\frac{m}{2t}\operatorname{floor}(\frac{\mu t}{m})\right)\left(\operatorname{floor}(\frac{\mu t}{m})+1\right)
\displaystyle\geq (μμ2)(floor(μtm)+1)=μ2(floor(μtm)+1)>μ2t2m.\displaystyle\left(\mu-\frac{\mu}{2}\right)\left(\operatorname{floor}(\frac{\mu t}{m})+1\right)=\frac{\mu}{2}\left(\operatorname{floor}(\frac{\mu t}{m})+1\right)>\frac{\mu^{2}t}{2m}.

Proposition 4.5.

Let k,N,νk,N,\nu\in{\mathbb{N}}. If ν>16Nk\nu>\frac{1}{6}Nk then Jk,Nν(ν)={0}J_{k,N\nu}(\nu)=\{0\}.

Proof.

The case of odd kk follows from that of even kk. If ϕJk,Nν(ν)\phi\in J_{k,N\nu}(\nu) then ϕ2J2k,N(2ν)(2ν)\phi^{2}\in J_{2k,N(2\nu)}(2\nu). Since ν>16Nk\nu>\frac{1}{6}Nk we have 2ν>16N(2k)2\nu>\frac{1}{6}N(2k) and so ϕ2=0\phi^{2}=0 assuming the result for even weights; hence ϕ=0\phi=0.

Suppose that there is a nontrivial φJk,Nν(ν)\varphi\in J_{k,N\nu}(\nu) with vanishing order μ=ordφν>16Nk\mu={\operatorname{ord}}\varphi\geq\nu>\frac{1}{6}Nk and kk even. We will obtain a contradiction to Theorem 4.3. For m=Nνm=N\nu, we will contradict

min(m,m6μ+k2)t:()ψ(t).\min\left(m,m-6\mu+\frac{k}{2}\right)\geq\sum_{t:(\spadesuit)}\psi(t).

Since ψ\psi has a positive minimum, it suffices to show that the set of positive integers satisfying ()(\spadesuit) is infinite.

We show that all sufficiently large primes pp satisfy

()c=0p1ψ1(gcd(p,c))max(μmc(pc)p2,0)>kp12ψ2(p).(\spadesuit)\qquad\sum_{c=0}^{p-1}\psi_{1}\left(\gcd(p,c)\right)\max\left(\mu-\frac{mc(p-c)}{p^{2}},0\right)>\frac{kp}{12}\psi_{2}(p).

We know that ata\mid t implies ψ1(a)ψ1(t)\psi_{1}(a)\geq\psi_{1}(t). Since gcd(p,c)p\gcd(p,c)\mid p, we have ψ1(gcd(p,c))ψ1(p)\psi_{1}(\gcd(p,c))\geq\psi_{1}(p). Therefore

c=0p1\displaystyle\sum_{c=0}^{p-1} ψ1(gcd(p,c))max(μmc(pc)p2,0)\displaystyle\psi_{1}\left(\gcd(p,c)\right)\max\left(\mu-\frac{mc(p-c)}{p^{2}},0\right)
\displaystyle\geq ψ1(p)c=0p1max(μmc(pc)p2,0).\displaystyle\psi_{1}(p)\sum_{c=0}^{p-1}\max\left(\mu-\frac{mc(p-c)}{p^{2}},0\right).

Lemma 4.4 says that for t,m,μt,m,\mu\in{\mathbb{N}} with μ<m\mu<m we have

c=0t1max(μmc(tc)t2,0)>μ2t2m.\sum_{c=0}^{t-1}\max\left(\mu-\frac{mc(t-c)}{t^{2}},0\right)>\frac{\mu^{2}t}{2m}.

We use the linear bound ordφk+2m12{\operatorname{ord}}\varphi\leq\frac{k+2m}{12} of [16], Proposition 3.2, to check the hypothesis μ<m\mu<m:

μ=ordφk+2m12<6νN+2m12=6N2+212m<m.\mu={\operatorname{ord}}\varphi\leq\frac{k+2m}{12}<\frac{\frac{6\nu}{N}+2m}{12}=\frac{\frac{6}{N^{2}}+2}{12}\,m<m.

Thus by Lemma 4.4 we have

c=0p1\displaystyle\sum_{c=0}^{p-1} ψ1(gcd(p,c))max(μmc(pc)p2,0)\displaystyle\psi_{1}\left(\gcd(p,c)\right)\max\left(\mu-\frac{mc(p-c)}{p^{2}},0\right)
>\displaystyle> ψ1(p)μ2p2mψ1(p)ν2p2m=ψ1(p)νp2N.\displaystyle\psi_{1}(p)\frac{\mu^{2}p}{2m}\geq\psi_{1}(p)\frac{\nu^{2}p}{2m}=\psi_{1}(p)\frac{\nu p}{2N}.

Thus a sufficient condition for a prime pp to satisfy ()(\spadesuit) is

ψ1(p)νp2N>kp12ψ2(p).\psi_{1}(p)\frac{\nu p}{2N}>\frac{kp}{12}\psi_{2}(p).

This simplifies to ν>Nk6(1+1p)\nu>\frac{Nk}{6}\left(1+\frac{1}{p}\right), or ν16Nk>Nk61p\nu-\frac{1}{6}Nk>\frac{Nk}{6}\frac{1}{p}, which is true for all sufficiently large pp because ν16Nk\nu-\frac{1}{6}Nk is positive. ∎

Corollary 4.6.

For k,Nk,N\in{\mathbb{N}} we have dim𝕄(k,N,ϵ)O(N3k3)\dim\mathbb{M}(k,N,\epsilon)\in O(N^{3}k^{3}).

Proof.

We have dim𝕄(k,N,ϵ)j=0dimJk,Nj(j+δ)\dim\mathbb{M}(k,N,\epsilon)\leq\sum_{j=0}^{\infty}\dim J_{k,Nj}(j+\delta), where δ\delta is 0 or 11 as (1)kϵ(-1)^{k}\epsilon is 11 or 1-1, by Lemma 3.2 of [26]. By Proposition 4.5 we may cap the summation at floor(16Nk){\operatorname{floor}(\frac{1}{6}Nk)} and, by Theorem 2.3 in [11], the codimension of Jk,mcuspJ_{k,m}^{\text{\rm cusp}} in Jk,mJ_{k,m} is at most floor(b/2)+1\operatorname{floor}(b/2)+1 where bb is the largest integer with b2mb^{2}\mid m. Therefore

dim𝕄(k,N,ϵ)\displaystyle\dim\mathbb{M}(k,N,\epsilon) j=0floor(16Nk)dimJk,Nj(j)j=0floor(16Nk)dimJk,Nj\displaystyle\leq\sum_{j=0}^{\operatorname{floor}(\frac{1}{6}Nk)}\dim J_{k,Nj}(j)\leq\sum_{j=0}^{\operatorname{floor}(\frac{1}{6}Nk)}\dim J_{k,Nj}
dimJk,0+j=1floor(16Nk)dimJk,Njcusp+12Nj+1\displaystyle\leq\dim J_{k,0}+\sum_{j=1}^{\operatorname{floor}(\frac{1}{6}Nk)}\dim J_{k,Nj}^{\text{\rm cusp}}+\tfrac{1}{2}\sqrt{Nj}+1
k+1212+23(16Nk+1)3/2+16Nk+j=1floor(16Nk)dimJk,Njcusp.\displaystyle\leq\tfrac{k+12}{12}+\tfrac{2}{3}\left(\tfrac{1}{6}Nk+1\right)^{3/2}+\tfrac{1}{6}Nk+\sum_{j=1}^{\operatorname{floor}(\frac{1}{6}Nk)}\dim J_{k,Nj}^{\text{\rm cusp}}.

It is known that for k,mk,m\in{\mathbb{N}}, dimJk,mcuspO(km)\dim J_{k,m}^{\text{\rm cusp}}\in O(km). An easy estimate from the dimension formula [11] is dimJk,mcuspm+124(2k+35)\dim J_{k,m}^{\text{\rm cusp}}\leq\frac{m+1}{24}(2k+35). From dimJk,NjcuspO(kNj)\dim J_{k,Nj}^{\text{\rm cusp}}\in O(kNj) we have j=1floor(16Nk)dimJk,NjcuspO(N3k3)\sum_{j=1}^{\operatorname{floor}(\frac{1}{6}Nk)}\dim J_{k,Nj}^{\text{\rm cusp}}\in O(N^{3}k^{3}). ∎

Proposition 4.7.

The integral domain 𝕄(N,+)=k=0𝕄(k,N,+)\mathbb{M}(N,+)=\oplus_{k=0}^{\infty}\mathbb{M}(k,N,+) is algebraic over its subring FJ(M(K(N)+))=k=0FJ(Mk(K(N)+)){\operatorname{FJ}}\left(M(K(N)^{+})\right)=\oplus_{k=0}^{\infty}{\operatorname{FJ}}\left(M_{k}(K(N)^{+})\right). For k0k\in{\mathbb{N}}_{0}, each 𝔣𝕄(k,N,+)\mathfrak{f}\in\mathbb{M}(k,N,+) satisfies a polynomial relation of the type

FJ(f0)𝔣d+FJ(f1)𝔣d1++FJ(fj)𝔣dj++FJ(fd)=0,{\operatorname{FJ}}(f_{0})\mathfrak{f}^{d}+{\operatorname{FJ}}(f_{1})\mathfrak{f}^{d-1}+\cdots+{\operatorname{FJ}}(f_{j})\mathfrak{f}^{d-j}+\cdots+{\operatorname{FJ}}(f_{d})=0,

for some dd\in{\mathbb{N}}, k00k_{0}\in{\mathbb{N}}_{0}, and some fjMk0+jk(K(N)+)f_{j}\in M_{k_{0}+jk}(K(N)^{+}) with f0f_{0} not identically zero.

Proof.

It suffices to prove the second statement. The group Γ=K(N)+\Gamma=K(N)^{+} is commensurable with Sp(4,)\operatorname{Sp}(4,{\mathbb{Z}}), so by Theorem 6.11 in Freitag [12], the homogeneous quotient field of M(Γ)M(\Gamma), 𝒦(Γ){\mathcal{K}}(\Gamma), has transcendence degree three. Take three meromorphic functions f1,f2,f3𝒦(Γ)f_{1},f_{2},f_{3}\in{\mathcal{K}}(\Gamma) that are algebraically independent over {\mathbb{C}} and select a common denominator DMk(Γ)D\in M_{k^{*}}\left(\Gamma\right). We must have kk^{*}\in{\mathbb{N}} because the fjf_{j} are not constant. We obtain four paramodular forms g1=Df1g_{1}=Df_{1}, g2=Df2g_{2}=Df_{2}, g3=Df3g_{3}=Df_{3}, and g4=Dg_{4}=D in Mk(Γ)M_{k^{*}}\left(\Gamma\right) that are algebraically independent over {\mathbb{C}}. This follows because we may reduce to the case where any putative polynomial relation is homogeneous in the gjg_{j}.

Take 𝔣𝕄(k,N,+)\mathfrak{f}\in\mathbb{M}(k,N,+). We may assume that 𝔣\mathfrak{f} is nontrivial because otherwise 𝔣1=0\mathfrak{f}^{1}=0 satisfies the conclusion with k0=0k_{0}=0, f0=1M0(Γ)f_{0}=1\in M_{0}(\Gamma), and f1=0Mk(Γ)f_{1}=0\in M_{k}(\Gamma). Similarly, if k=0k=0 then 𝔣=c\mathfrak{f}=c is constant and 𝔣1c=0\mathfrak{f}^{1}-c=0 satisfies the conclusion with k0=0k_{0}=0, f0=1M0(Γ)f_{0}=1\in M_{0}(\Gamma), and f1=cM0(Γ)f_{1}=-c\in M_{0}(\Gamma). So we assume k>0k>0 as well. The four formal series FJ(gjk)𝕄(kk,N,+){\operatorname{FJ}}(g_{j}^{k})\in\mathbb{M}(k^{*}k,N,+) are algebraically independent over {\mathbb{C}} because FJ{\operatorname{FJ}} is a monomorphism. Consider the list of five formal series FJ(gjk){\operatorname{FJ}}(g_{j}^{k}), 𝔣k𝕄(kk,N,+)\mathfrak{f}^{k^{*}}\in\mathbb{M}(k^{*}k,N,+). For any μ\mu\in{\mathbb{N}}, there are (μ+44)\binom{\mu+4}{4} distinct monomials x1i1x5i5x_{1}^{i_{1}}\cdots x_{5}^{i_{5}} in five variables with i1++i5=μi_{1}+\cdots+i_{5}=\mu. By substitution of the five formal series into these monomials, we have (μ+44)\binom{\mu+4}{4} elements FJ(g1ki1)FJ(g4ki4)𝔣ki5𝕄(μkk,N,+){\operatorname{FJ}}(g_{1}^{ki_{1}})\cdots{\operatorname{FJ}}(g_{4}^{ki_{4}})\mathfrak{f}^{k^{*}i_{5}}\in\mathbb{M}(\mu k^{*}k,N,+). By Corollary 4.6, however, we have dim𝕄(μkk,N,+)O((Nkk)3μ3)\dim\mathbb{M}(\mu k^{*}k,N,+)\in O((Nk^{*}k)^{3}\mu^{3}), so for sufficiently large μ\mu there is a nontrivial {\mathbb{C}}-linear dependence relation among these (μ+44)\binom{\mu+4}{4} elements. At least one supported monomial in the dependence relation must contain a positive power of 𝔣\mathfrak{f} because the remaining four formal series are algebraically independent. For the same reason, when a positive power of 𝔣\mathfrak{f} is supported then its coefficient, after collecting like terms in powers of 𝔣\mathfrak{f}, must be nontrivial. If dd\in{\mathbb{N}} is the highest power of 𝔣\mathfrak{f} that is supported in the dependence relation, then we may write this relation as j=0dFJ(fj)𝔣dj=0\sum_{j=0}^{d}{\operatorname{FJ}}(f_{j})\mathfrak{f}^{d-j}=0 where the fjf_{j} are {\mathbb{C}}-linear combinations of monomials in the four g1,,g4g_{1},\ldots,g_{4}. Since 𝔣d\mathfrak{f}^{d} is supported, we have f0f_{0} nontrivial. Let the weight of f0f_{0} be k00k_{0}\in{\mathbb{N}}_{0} so that f0Mk0(Γ)f_{0}\in M_{k_{0}}(\Gamma). The terms FJ(fj)𝔣dj{\operatorname{FJ}}(f_{j})\mathfrak{f}^{d-j} all have the same weight, k0+kdk_{0}+kd, which is the weight of FJ(f0)𝔣d{\operatorname{FJ}}(f_{0})\mathfrak{f}^{d}. Therefore we may take fjMk0+kj(Γ)f_{j}\in M_{k_{0}+kj}(\Gamma), as required. ∎


5. Invariance under subgroups of finite index in Γ0(N)\Gamma^{0}(N).

In Proposition 4.7 of the previous section we saw that a formal series of Jacobi forms 𝔣\mathfrak{f} possessing the involution condition for ϵ=+1\epsilon=+1 satisfies a polynomial P(X)=0P(X)=0 whose coefficients are formal Fourier-Jacobi expansions of paramodular forms. As a consequence of this polynomial relation we will show in this section that 𝔣\mathfrak{f} is invariant under a subgroup Γ\Gamma of finite index in Γ0(N)\Gamma^{0}(N). These arguments best take place inside the ring of formal Fourier series.

The ring structure on the ring of formal Fourier series, 𝒳¯(N){\mathbb{C}}^{{\bar{\mathcal{X}}}(N)}, is defined by the Cauchy product, noting that for every t𝒳¯(N)t\in{\bar{\mathcal{X}}}(N) the set {(t1,t2)𝒳¯(N)×𝒳¯(N):t1+t2=t}\{(t_{1},t_{2})\in{\bar{\mathcal{X}}}(N)\times{\bar{\mathcal{X}}}(N):t_{1}+t_{2}=t\} is finite. We accordingly use a place-holding variable qq to write an element ψ𝒳¯(N)\psi\in{\mathbb{C}}^{{\bar{\mathcal{X}}}(N)} as

ψ=t𝒳¯(N)a(t;ψ)qt.\psi=\sum_{t\in{\bar{\mathcal{X}}}(N)}a(t;\psi)q^{t}.

The ring of formal Fourier series, 𝒳¯(N){\mathbb{C}}^{{\bar{\mathcal{X}}}(N)} is also an integral domain, which will be proven in Corollary 5.2. The Fourier expansion of a paramodular form defines a map

FS:Mk(K(N))\displaystyle{\operatorname{FS}}:M_{k}\left(K(N)\right) 𝒳¯(N)\displaystyle\to{\mathbb{C}}^{{\bar{\mathcal{X}}}(N)}
f\displaystyle f t𝒳¯(N)a(t;f)qt.\displaystyle\mapsto\sum_{t\in{\bar{\mathcal{X}}}(N)}a(t;f)q^{t}.

Given a formal series of Jacobi forms 𝔣𝕄(k,N)\mathfrak{f}\in\mathbb{M}(k,N), we may define the associated formal Fourier series, AFS(𝔣)𝒳¯(N){\operatorname{AFS}}(\mathfrak{f})\in{\mathbb{C}}^{{\bar{\mathcal{X}}}(N)}, by

AFS:𝕄(k,N)\displaystyle{\operatorname{AFS}}:\mathbb{M}(k,N) 𝒳¯(N)\displaystyle\to{\mathbb{C}}^{{\bar{\mathcal{X}}}(N)}
𝔣=m=0ϕmξNm\displaystyle\mathfrak{f}=\sum_{m=0}^{\infty}\phi_{m}\xi^{Nm} t=(nr/2r/2Nm)𝒳¯(N)c(n,r;ϕm)qt=t𝒳¯(N)a(t;𝔣)qt.\displaystyle\mapsto\sum_{t={\tiny\left(\begin{smallmatrix}{n}&{r/2}\\ {r/2}&{Nm}\end{smallmatrix}\right)}\in{\bar{\mathcal{X}}}(N)}c(n,r;\phi_{m})q^{t}=\sum_{t\in{\bar{\mathcal{X}}}(N)}a(t;\mathfrak{f})q^{t}.

Extending by linearity we have a map AFS:𝕄(N)𝒳¯(N){\operatorname{AFS}}:\mathbb{M}(N)\to{\mathbb{C}}^{{\bar{\mathcal{X}}}(N)} from the ring 𝕄(N)=k0𝕄(k,N)\mathbb{M}(N)=\oplus_{k\in{\mathbb{N}}_{0}}\mathbb{M}(k,N). That AFS{\operatorname{AFS}} is a ring homomorphism is an exercise using the Cauchy product similar to the computations demonstrating equation (5). For fMk(K(N))f\in M_{k}\left(K(N)\right) we have the compatibility AFS(FJ(f))=FS(f){\operatorname{AFS}}\left({\operatorname{FJ}}(f)\right)={\operatorname{FS}}(f).

Formal Fourier series share some properties with formal power series [[a,b,c]]{\mathbb{C}}[[a,b,c]] in three variables due to the following monomorphism that sends qtq^{t} to the monomial as1,tbs2,tcs3,ta^{\langle s_{1},t\rangle}b^{\langle s_{2},t\rangle}c^{\langle s_{3},t\rangle}.

Lemma 5.1.

Set s1=(1000)s_{1}=\left(\begin{smallmatrix}{1}&{0}\\ {0}&{0}\end{smallmatrix}\right), s2=(1111)s_{2}=\left(\begin{smallmatrix}{1}&{1}\\ {1}&{1}\end{smallmatrix}\right), s3=(0001)s_{3}=\left(\begin{smallmatrix}{0}&{0}\\ {0}&{1}\end{smallmatrix}\right). The map

ι:𝒳¯(N)\displaystyle\iota:{\mathbb{C}}^{{\bar{\mathcal{X}}}(N)} [[a,b,c]]\displaystyle\to{\mathbb{C}}[[a,b,c]]
ψ=t𝒳¯(N)a(t;ψ)qt\displaystyle\psi=\sum_{t\in{\bar{\mathcal{X}}}(N)}a(t;\psi)q^{t} i,j,k0(t𝒳¯(N):s1,t=i,s2,t=j,s3,t=ka(t;ψ))aibjck\displaystyle\mapsto\sum_{i,j,k\in{\mathbb{N}}_{0}}\left(\sum_{\begin{subarray}{c}t\in{\bar{\mathcal{X}}}(N):\\ \langle s_{1},t\rangle=i,\langle s_{2},t\rangle=j,\langle s_{3},t\rangle=k\end{subarray}}a(t;\psi)\right)a^{i}b^{j}c^{k}

is a ring homomorphism satisfying the following properties.

  1. (1)

    ι(ψ)=i,j,k0a((ijik2jik2k);ψ)aibjck\iota(\psi)=\sum_{i,j,k\in{\mathbb{N}}_{0}}a\left(\left(\begin{smallmatrix}{i}&{\frac{j-i-k}{2}}\\ {\frac{j-i-k}{2}}&{k}\end{smallmatrix}\right);\psi\right)a^{i}b^{j}c^{k},

  2. (2)

    ι(ψ)=t𝒳¯(N)a(t;ψ)as1,tbs2,tcs3,t\iota(\psi)=\sum_{t\in{\bar{\mathcal{X}}}(N)}a\left(t;\psi\right)a^{\langle s_{1},t\rangle}b^{\langle s_{2},t\rangle}c^{\langle s_{3},t\rangle},

  3. (3)

    ι\iota is injective.

Proof.

Setting t=(nr/2r/2Nm)𝒳¯(N)t=\left(\begin{smallmatrix}{n}&{r/2}\\ {r/2}&{Nm}\end{smallmatrix}\right)\in{\bar{\mathcal{X}}}(N) and solving the system of linear equations s1,t=i\langle s_{1},t\rangle=i, s2,t=j\langle s_{2},t\rangle=j, and s3,t=k\langle s_{3},t\rangle=k, the unique solution is n=in=i, r=jik2r=\frac{j-i-k}{2}, and Nm=kNm=k. This proves formula (1)(1) if we understand that a(t;ψ)=0a(t;\psi)=0 for t𝒳¯(N)t\not\in{\bar{\mathcal{X}}}(N). Formula (2)(2) follows from (1)(1) since rearrangements of formal power series are equal. The ring homomorphism property is then formal because the s,t\langle s_{\ell},t\rangle are linear in tt. The injectivity (3)(3) follows from formula (1)(1). ∎

Corollary 5.2.

The ring 𝒳¯(N){\mathbb{C}}^{{\bar{\mathcal{X}}}(N)} is an integral domain.

Proof.

A ring with a monomorphism to an integral domain is an integral domain. The monomorphism here is ι:𝒳¯(N)[[a,b,c]]\iota:{\mathbb{C}}^{{\bar{\mathcal{X}}}(N)}\to{\mathbb{C}}[[a,b,c]]. ∎

There is a copy of Γ0(N)\Gamma^{0}(N) inside the group of automorphisms of formal Fourier series.

Lemma 5.3.

For σΓ0(N)\sigma\in\Gamma^{0}(N) define

j(σ):𝒳¯(N)\displaystyle\operatorname{j}(\sigma):{\mathbb{C}}^{{\bar{\mathcal{X}}}(N)} 𝒳¯(N)\displaystyle\to{\mathbb{C}}^{{\bar{\mathcal{X}}}(N)}
t𝒳¯(N)a(t;ψ)qt\displaystyle\sum_{t\in{\bar{\mathcal{X}}}(N)}a(t;\psi)q^{t} t𝒳¯(N)a(t[σ];ψ)qt.\displaystyle\mapsto\sum_{t\in{\bar{\mathcal{X}}}(N)}a(t[\sigma];\psi)q^{t}.

The map j(σ)\operatorname{j}(\sigma) is an automorphism of 𝒳¯(N){\mathbb{C}}^{{\bar{\mathcal{X}}}(N)} and the map

j:Γ0(N)\displaystyle\operatorname{j}:\Gamma^{0}(N) Aut(𝒳¯(N))\displaystyle\to{\operatorname{Aut}}\left({\mathbb{C}}^{{\bar{\mathcal{X}}}(N)}\right)
σ\displaystyle\sigma j(σ)\displaystyle\mapsto\operatorname{j}(\sigma)

is a homomorphism. For 𝔣𝕄(k,N)\mathfrak{f}\in\mathbb{M}(k,N), we have j(σ)AFS(𝔣)=AFS(𝔣)\operatorname{j}(\sigma){\operatorname{AFS}}(\mathfrak{f})={\operatorname{AFS}}(\mathfrak{f}) if and only if a(t[σ];𝔣)=a(t;𝔣)a(t[\sigma];\mathfrak{f})=a(t;\mathfrak{f}) for all t𝒳¯(N)t\in{\bar{\mathcal{X}}}(N). For fMk(K(N))f\in M_{k}\left(K(N)\right) we have j(σ)FS(f)=FS(f)\operatorname{j}(\sigma){\operatorname{FS}}(f)={\operatorname{FS}}(f) for all σΓ0(N)\sigma\in\Gamma^{0}(N).

Proof.

We show j(σ)\operatorname{j}(\sigma) is an automorphism. The map j(σ)\operatorname{j}(\sigma) has j(σ1)\operatorname{j}(\sigma^{-1}) as an inverse and the additivity of j(σ)\operatorname{j}(\sigma) is clear so that it suffices to prove j(σ)(ψ1ψ2)=(j(σ)ψ1)(j(σ)ψ2)\operatorname{j}(\sigma)\left(\psi_{1}\psi_{2}\right)=\left(\operatorname{j}(\sigma)\psi_{1}\right)\left(\operatorname{j}(\sigma)\psi_{2}\right). We have

j(σ)(ψ1ψ2)\displaystyle\operatorname{j}(\sigma)\left(\psi_{1}\psi_{2}\right) =j(σ)t𝒳¯(N)(t1,t2𝒳¯(N):t1+t2=ta(t1;ψ1)a(t2;ψ2))qt\displaystyle=\operatorname{j}(\sigma)\sum_{t\in{\bar{\mathcal{X}}}(N)}\left(\sum_{t_{1},t_{2}\in{\bar{\mathcal{X}}}(N):\,t_{1}+t_{2}=t}a(t_{1};\psi_{1})a(t_{2};\psi_{2})\right)q^{t}
=t𝒳¯(N)(t1,t2𝒳¯(N):t1+t2=t[σ]a(t1;ψ1)a(t2;ψ2))qt.\displaystyle=\sum_{t\in{\bar{\mathcal{X}}}(N)}\left(\sum_{t_{1},t_{2}\in{\bar{\mathcal{X}}}(N):\,t_{1}+t_{2}=t[\sigma]}a(t_{1};\psi_{1})a(t_{2};\psi_{2})\right)q^{t}.

We use the equality {(t1,t2)𝒳¯(N)×𝒳¯(N):t1+t2=t[σ]}={(s1[σ],s2[σ])𝒳¯(N)×𝒳¯(N):s1+s2=t}\{(t_{1},t_{2})\in{\bar{\mathcal{X}}}(N)\times{\bar{\mathcal{X}}}(N):t_{1}+t_{2}=t[\sigma]\}=\{(s_{1}[\sigma],s_{2}[\sigma])\in{\bar{\mathcal{X}}}(N)\times{\bar{\mathcal{X}}}(N):s_{1}+s_{2}=t\} to change the index of summation.

j(σ)(ψ1ψ2)\displaystyle\operatorname{j}(\sigma)\left(\psi_{1}\psi_{2}\right) =t𝒳¯(N)(s1,s2𝒳¯(N):s1+s2=ta(s1[σ];ψ1)a(s2[σ];ψ2))qt\displaystyle=\sum_{t\in{\bar{\mathcal{X}}}(N)}\left(\sum_{s_{1},s_{2}\in{\bar{\mathcal{X}}}(N):\,s_{1}+s_{2}=t}a(s_{1}[\sigma];\psi_{1})a(s_{2}[\sigma];\psi_{2})\right)q^{t}
=s1𝒳¯(N)a(s1[σ];ψ1)qs1s2𝒳¯(N)a(s2[σ];ψ2)qs2\displaystyle=\sum_{s_{1}\in{\bar{\mathcal{X}}}(N)}a(s_{1}[\sigma];\psi_{1})q^{s_{1}}\sum_{s_{2}\in{\bar{\mathcal{X}}}(N)}a(s_{2}[\sigma];\psi_{2})q^{s_{2}}
=(j(σ)ψ1)(j(σ)ψ2).\displaystyle=\left(\operatorname{j}(\sigma)\psi_{1}\right)\left(\operatorname{j}(\sigma)\psi_{2}\right).

We show that j:Γ0(N)Aut(𝒳¯(N))\operatorname{j}:\Gamma^{0}(N)\to{\operatorname{Aut}}\left({\mathbb{C}}^{{\bar{\mathcal{X}}}(N)}\right) is a homomorphism. Take σ1,σ2Γ0(N)\sigma_{1},\sigma_{2}\in\Gamma^{0}(N) and ψ=t𝒳¯(N)a(t;ψ)qt𝒳¯(N)\psi=\sum_{t\in{\bar{\mathcal{X}}}(N)}a(t;\psi)q^{t}\in{\mathbb{C}}^{{\bar{\mathcal{X}}}(N)}. We have

j(σ2)ψ\displaystyle\operatorname{j}(\sigma_{2})\psi =ta(t[σ2];ψ)qt\displaystyle=\sum_{t}a(t[\sigma_{2}];\psi)q^{t}
(j(σ1)j(σ2))ψ=j(σ1)(j(σ2)ψ)\displaystyle\left(\operatorname{j}(\sigma_{1})\operatorname{j}(\sigma_{2})\right)\psi=\operatorname{j}(\sigma_{1})\left(\operatorname{j}(\sigma_{2})\psi\right) =ta(t[σ1][σ2];ψ)qt=j(σ1σ2)ψ.\displaystyle=\sum_{t}a(t[\sigma_{1}][\sigma_{2}];\psi)q^{t}=\operatorname{j}(\sigma_{1}\sigma_{2})\psi.

For 𝔣=m=0ϕmξNm𝕄(k,N)\mathfrak{f}=\sum_{m=0}^{\infty}\phi_{m}\xi^{Nm}\in\mathbb{M}(k,N) we have AFS(𝔣)=ta(t;𝔣)qt{\operatorname{AFS}}(\mathfrak{f})=\sum_{t}a(t;\mathfrak{f})q^{t} where a(t;𝔣)=c(n,r;ϕm)a(t;\mathfrak{f})=c(n,r;\phi_{m}) for t=(nr/2r/2Nm)𝒳¯(N)t=\left(\begin{smallmatrix}{n}&{r/2}\\ {r/2}&{Nm}\end{smallmatrix}\right)\in{\bar{\mathcal{X}}}(N). We then have j(σ)AFS(𝔣)=ta(t[σ];𝔣)qt\operatorname{j}(\sigma){\operatorname{AFS}}(\mathfrak{f})=\sum_{t}a(t[\sigma];\mathfrak{f})q^{t} and, by definition of formal series, this equals AFS(𝔣){\operatorname{AFS}}(\mathfrak{f}) if and only if a(t[σ];𝔣)=a(t;𝔣)a(t[\sigma];\mathfrak{f})=a(t;\mathfrak{f}) for all t𝒳¯(N)t\in{\bar{\mathcal{X}}}(N). The final assertion, j(σ)FS(f)=FS(f)\operatorname{j}(\sigma){\operatorname{FS}}(f)={\operatorname{FS}}(f) for all σΓ0(N)\sigma\in\Gamma^{0}(N), follows from the Γ0(N)\Gamma^{0}(N)-symmetries of fMk(K(N))f\in M_{k}\left(K(N)\right) in equation (2). ∎

The next proposition shows that a formal series of Jacobi forms satisfying the involution condition necessarily has additional symmetries.

Proposition 5.4.

Let 𝔣𝕄(k,N,+)\mathfrak{f}\in\mathbb{M}(k,N,+). There is a subgroup Γ\Gamma of finite index in Γ0(N)\Gamma^{0}(N) such that 𝔣𝕄(k,N,+;Γ)\mathfrak{f}\in\mathbb{M}(k,N,+\,;\Gamma).

Proof.

By Proposition 4.7, 𝔣\mathfrak{f} satisfies a polynomial relation of the type

FJ(f0)𝔣d++FJ(fj)𝔣dj++FJ(fd)=0,{\operatorname{FJ}}(f_{0})\mathfrak{f}^{d}+\cdots+{\operatorname{FJ}}(f_{j})\mathfrak{f}^{d-j}+\cdots+{\operatorname{FJ}}(f_{d})=0,

for some dd\in{\mathbb{N}}, k00k_{0}\in{\mathbb{N}}_{0}, and some fjMk0+jk(K(N)+)f_{j}\in M_{k_{0}+jk}(K(N)^{+}) with f0f_{0} not identically zero. Apply the monomorphism AFS:𝕄(N)𝒳¯(N){\operatorname{AFS}}:\mathbb{M}(N)\to{\mathbb{C}}^{{\bar{\mathcal{X}}}(N)} to obtain

FS(f0)(AFS(𝔣))d++FS(fj)(AFS(𝔣))dj++FS(fd)=0.{\operatorname{FS}}(f_{0})({\operatorname{AFS}}(\mathfrak{f}))^{d}+\cdots+{\operatorname{FS}}(f_{j})({\operatorname{AFS}}(\mathfrak{f}))^{d-j}+\cdots+{\operatorname{FS}}(f_{d})=0.

The polynomial j=0dFS(fj)Xdj𝒳¯(N)[X]\sum_{j=0}^{d}{\operatorname{FS}}(f_{j})X^{d-j}\in{\mathbb{C}}^{{\bar{\mathcal{X}}}(N)}[X] has a root AFS(𝔣){\operatorname{AFS}}(\mathfrak{f}). Noting by Lemma 5.3 that j(σ)FS(fj)=FS(fj)\operatorname{j}(\sigma){\operatorname{FS}}(f_{j})={\operatorname{FS}}(f_{j}) for every σΓ0(N)\sigma\in\Gamma^{0}(N), each element of the orbit j(Γ0(N))AFS(𝔣)\operatorname{j}\left(\Gamma^{0}(N)\right){\operatorname{AFS}}(\mathfrak{f}) is a root. However, the ring 𝒳¯(N){\mathbb{C}}^{{\bar{\mathcal{X}}}(N)} is an integral domain and a polynomial of positive degree dd over an integral domain has at most dd roots. If we want to be definite we can find u1,,ud1Γ0(N)u_{1},\ldots,u_{d_{1}}\in\Gamma^{0}(N) with d1dd_{1}\leq d such that j(Γ0(N))AFS(𝔣)={j(u1)AFS(𝔣),,j(ud1)AFS(𝔣)}\operatorname{j}\left(\Gamma^{0}(N)\right){\operatorname{AFS}}(\mathfrak{f})=\{\operatorname{j}(u_{1}){\operatorname{AFS}}(\mathfrak{f}),\ldots,\operatorname{j}(u_{d_{1}}){\operatorname{AFS}}(\mathfrak{f})\}. The natural homomorphism ρ\rho from Γ0(N)\Gamma^{0}(N) to permutations of the orbit j(Γ0(N))AFS(𝔣)\operatorname{j}\left(\Gamma^{0}(N)\right){\operatorname{AFS}}(\mathfrak{f}) is specified in this labeling by j(σ)j(ui)AFS(𝔣)=j(uρ(σ)i)AFS(𝔣)\operatorname{j}(\sigma)\operatorname{j}(u_{i}){\operatorname{AFS}}(\mathfrak{f})=\operatorname{j}(u_{\rho(\sigma)i}){\operatorname{AFS}}(\mathfrak{f}) for i=1,,d1i=1,\ldots,d_{1}. Let Γ=ker(ρ)\Gamma=\ker(\rho) be the kernel of ρ:Γ0(N)Sd1\rho:\Gamma^{0}(N)\to S_{d_{1}}. Then Γ\Gamma is a normal subgroup of Γ0(N)\Gamma^{0}(N) of index at most d1!d_{1}!. For every σΓ\sigma\in\Gamma we have j(σ)AFS(𝔣)=AFS(𝔣)\operatorname{j}(\sigma){\operatorname{AFS}}(\mathfrak{f})={\operatorname{AFS}}(\mathfrak{f}) so that, by Lemma 5.3, a(t[σ];𝔣)=a(t;𝔣)a(t[\sigma];\mathfrak{f})=a(t;\mathfrak{f}) for all t𝒳¯(N)t\in{\bar{\mathcal{X}}}(N). Thus 𝔣𝕄(k,N,+)𝕄(k,N;Γ)=𝕄(k,N,+;Γ)\mathfrak{f}\in\mathbb{M}(k,N,+)\cap\mathbb{M}(k,N;\Gamma)=\mathbb{M}(k,N,+\,;\Gamma) as claimed. ∎

6. Specialization.

Let ΓΓ0(N)\Gamma\subseteq\Gamma^{0}(N) be a subgroup of finite index. A formal series of Jacobi forms 𝔣=m=1ϕmξNm𝕊(k,N;Γ)\mathfrak{f}=\sum_{m=1}^{\infty}\phi_{m}\xi^{Nm}\in\mathbb{S}(k,N;\Gamma) has the defining symmetries a(t[σ];𝔣)=a(t;𝔣)a(t[\sigma];\mathfrak{f})=a(t;\mathfrak{f}) for all σΓ\sigma\in\Gamma and all t𝒳¯(N)t\in{\bar{\mathcal{X}}}(N). We will construct a dense subset 𝒲1(Γ)\mathcal{W}_{1}(\Gamma) of 1×\mathcal{H}_{1}\times{\mathbb{C}} where the formal series 𝔣\mathfrak{f} specializes to a holomorphic function of one variable. More precisely, for each (τ1,z1)𝒲1(Γ)(\tau_{1},z_{1})\in\mathcal{W}_{1}(\Gamma), the series m=1ϕm(τ1,z1)e(Nmω)\sum_{m=1}^{\infty}\phi_{m}(\tau_{1},z_{1})e\left(Nm\omega\right) will converge to a holomorphic function H(τ1,z1,𝔣)(ω)H(\tau_{1},z_{1},\mathfrak{f})(\omega) on the neighborhood of infinity {ω1:(τ1z1z1ω)2}\{\omega\in\mathcal{H}_{1}:\left(\begin{smallmatrix}{\tau_{1}}&{z_{1}}\\ {z_{1}}&{\omega}\end{smallmatrix}\right)\in\mathcal{H}_{2}\}. The formal series 𝔣\mathfrak{f} thus converges on a dense subset of 2\mathcal{H}_{2}.

Definition 6.1.

For xx\in{\mathbb{Q}}, let Denom(x)=min{n:nx}\operatorname{Denom}(x)=\min\{n\in{\mathbb{N}}:nx\in{\mathbb{Z}}\} be the minimal positive denominator of xx.

Definition 6.2.

Let ΓSL(2,)\Gamma\subseteq{\operatorname{SL}}(2,{\mathbb{Z}}) be a subgroup of finite index. Let U=(0110)GL(2,)U=\left(\begin{smallmatrix}{0}&{1}\\ {1}&{0}\end{smallmatrix}\right)\in{\operatorname{GL}}(2,{\mathbb{Z}}). Define

𝒲0(Γ)\displaystyle\mathcal{W}_{0}(\Gamma) ={(x,y)2:x(UΓU)-orbit() and yDenom(x)},\displaystyle=\{(x,y)\in{\mathbb{Q}}^{2}:x\in\text{$(U\Gamma U)$-orbit$(\infty)$ and }y\operatorname{Denom}(x)\in{\mathbb{Z}}\},
𝒲1(Γ)\displaystyle\mathcal{W}_{1}(\Gamma) ={(τ,z)1×:(x,y)𝒲0(Γ):z=xτ+y},\displaystyle=\{(\tau,z)\in\mathcal{H}_{1}\times{\mathbb{C}}:\exists(x,y)\in\mathcal{W}_{0}(\Gamma):z=x\tau+y\},
𝒲(Γ)\displaystyle\mathcal{W}(\Gamma) ={(τzzω)2:(τ,z)𝒲1(Γ)}.\displaystyle=\{\left(\begin{smallmatrix}{\tau}&{z}\\ {z}&{\omega}\end{smallmatrix}\right)\in\mathcal{H}_{2}:(\tau,z)\in\mathcal{W}_{1}(\Gamma)\}.

We follow Lehner [30] for the theory of Fuchsian groups. We view the Riemann sphere 1()={}\mathbb{P}^{1}({\mathbb{C}})={\mathbb{C}}\cup\{\infty\} as the extended complex plane. The Riemann sphere is the disjoint union 1()=1()1¯1\mathbb{P}^{1}({\mathbb{C}})=\mathbb{P}^{1}({\mathbb{R}})\amalg\mathcal{H}_{1}\amalg{\overline{\mathcal{H}}_{1}} of the extended real numbers, 1()={}\mathbb{P}^{1}({\mathbb{R}})={\mathbb{R}}\cup\{\infty\}, and the upper and lower half planes. The groups SL(2,){\operatorname{SL}}(2,{\mathbb{R}}) and PSL(2,)=SL(2,)/{±I}{\operatorname{PSL}}(2,{\mathbb{R}})={\operatorname{SL}}(2,{\mathbb{R}})/\{\pm I\} act on 1()\mathbb{P}^{1}({\mathbb{C}}) by Möbius transformations and preserve this disjoint union. We will only consider subgroups ΓSL(2,)\Gamma\subseteq{\operatorname{SL}}(2,{\mathbb{R}}) and the corresponding transformation groups Γ¯=Γ,I/{±I}PSL(2,){\bar{\Gamma}}=\langle\Gamma,-I\rangle/\{\pm I\}\subseteq{\operatorname{PSL}}(2,{\mathbb{R}}).

Definition 6.3.

Let ΓSL(2,)\Gamma\subseteq{\operatorname{SL}}(2,{\mathbb{R}}). The limit set Λ(Γ)\Lambda(\Gamma) of Γ\Gamma is the set of z1()z\in\mathbb{P}^{1}({\mathbb{C}}) such that there exists a w1()w\in\mathbb{P}^{1}({\mathbb{C}}) and a sequence of distinct γnΓ\gamma_{n}\in\Gamma with limn+γnw=z\lim_{n\to+\infty}\gamma_{n}\langle w\rangle=z.

For subgroups Γ1,Γ2SL(2,)\Gamma_{1},\Gamma_{2}\subseteq{\operatorname{SL}}(2,{\mathbb{R}}), if Γ1\Gamma_{1} has finite index in Γ2\Gamma_{2} then Λ(Γ1)=Λ(Γ2)\Lambda(\Gamma_{1})=\Lambda(\Gamma_{2}); this is the theorem in section 2C of [30], page 11. The corollary in section 2F, page 14, is that either Λ(Γ)=1()\Lambda(\Gamma)=\mathbb{P}^{1}({\mathbb{C}}) or Λ(Γ)1()\Lambda(\Gamma)\subseteq\mathbb{P}^{1}({\mathbb{R}}). We mention some terminology to assist readers who use a different reference. A group Γ¯PSL(2,)\bar{\Gamma}\subseteq{\operatorname{PSL}}(2,{\mathbb{R}}) is Fuchsian when Λ(Γ)1()\Lambda(\Gamma)\subseteq\mathbb{P}^{1}({\mathbb{R}}). The theorem in section 2F of [30], page 13, characterizes Fuchsian groups as the discrete subgroups of PSL(2,){\operatorname{PSL}}(2,{\mathbb{R}}). For us, the salient result is Theorem 3 in section 3E, page 21.

Theorem 6.4 ([30]).

Let ΓSL(2,)\Gamma\subseteq{\operatorname{SL}}(2,{\mathbb{R}}) be a subgroup. If F1()F\subseteq\mathbb{P}^{1}({\mathbb{C}}) is a closed set containing at least two points, such that ΓFF\Gamma F\subseteq F, then FΛ(Γ)F\supseteq\Lambda(\Gamma).

As Lehner comments on page 21, this theorem may be rephrased: when Λ(Γ)\Lambda(\Gamma) has more than one point, Λ(Γ)\Lambda(\Gamma) is the smallest closed Γ\Gamma-invariant set containing at least two points.

An example of a Fuchsian group is PSL(2,){\operatorname{PSL}}(2,{\mathbb{Z}}). The orbit of \infty is 1()\mathbb{P}^{1}({\mathbb{Q}}) and Λ(SL(2,))=1()\Lambda({\operatorname{SL}}(2,{\mathbb{Z}}))=\mathbb{P}^{1}({\mathbb{R}}). This implies that any subgroup ΓSL(2,)\Gamma\subseteq{\operatorname{SL}}(2,{\mathbb{Z}}) of finite index also has Λ(Γ)=1()\Lambda(\Gamma)=\mathbb{P}^{1}({\mathbb{R}}). For a ring RR\subseteq{\mathbb{R}}, set 𝒫10(R)={(abcd)SL(2,R):c=0}\mathcal{P}_{10}(R)=\{\left(\begin{smallmatrix}{a}&{b}\\ {c}&{d}\end{smallmatrix}\right)\in{\operatorname{SL}}(2,R):c=0\}. The stabilizer of \infty in SL(2,){\operatorname{SL}}(2,{\mathbb{R}}) is 𝒫10()\mathcal{P}_{10}({\mathbb{R}}) and we have an orbit-stabilizer bijection SL(2,)/𝒫10()1(){\operatorname{SL}}(2,{\mathbb{Z}})/\mathcal{P}_{10}({\mathbb{Z}})\leftrightarrow\mathbb{P}^{1}({\mathbb{Q}}) given by sending γ𝒫10()γ\gamma\mathcal{P}_{10}({\mathbb{Z}})\mapsto\gamma\langle\infty\rangle. This bijection shows that a subgroup of finite index in SL(2,){\operatorname{SL}}(2,{\mathbb{Z}}) cannot stabilize \infty. We require the following ergodic corollary of Theorem 6.4.

Lemma 6.5.

Let ΓSL(2,)\Gamma\subseteq{\operatorname{SL}}(2,{\mathbb{Z}}) be a subgroup of finite index. The Γ\Gamma-orbit of \infty is dense in Λ(Γ)=1()\Lambda(\Gamma)=\mathbb{P}^{1}({\mathbb{R}}).

Proof.

We have Λ(Γ)=Λ(SL(2,))=1()\Lambda(\Gamma)=\Lambda({\operatorname{SL}}(2,{\mathbb{Z}}))=\mathbb{P}^{1}({\mathbb{R}}) because Γ\Gamma has finite index in SL(2,){\operatorname{SL}}(2,{\mathbb{Z}}). Furthermore, Γ\Gamma cannot stabilize \infty for the same reason. Therefore, the Γ\Gamma-orbit of \infty has at least two points, as does its closure F=Γ-orbit()¯1()F=\overline{\text{$\Gamma$-orbit$(\infty)$}}\subseteq\mathbb{P}^{1}({\mathbb{R}}). We check that the closure FF remains Γ\Gamma-invariant. Take zFz\in F and γΓ\gamma\in\Gamma. We have z=limn+γnz=\lim_{n\to+\infty}\gamma_{n}\langle\infty\rangle for some γnΓ\gamma_{n}\in\Gamma, so that γz=limn+(γγn)\gamma\langle z\rangle=\lim_{n\to+\infty}(\gamma\gamma_{n})\langle\infty\rangle for γγnΓ\gamma\gamma_{n}\in\Gamma. Hence γzF\gamma\langle z\rangle\in F and FF is Γ\Gamma-invariant. Thus, by Theorem 6.4 we have F=Λ(Γ)=1()F=\Lambda(\Gamma)=\mathbb{P}^{1}({\mathbb{R}}). The Γ\Gamma-orbit of \infty is thus dense in 1()\mathbb{P}^{1}({\mathbb{R}}). ∎

Lemma 6.6.

Let ΓSL(2,)\Gamma\subseteq{\operatorname{SL}}(2,{\mathbb{Z}}) be a subgroup of finite index. The sets 𝒲0(Γ)\mathcal{W}_{0}(\Gamma), 𝒲1(Γ)\mathcal{W}_{1}(\Gamma), and 𝒲(Γ)\mathcal{W}(\Gamma) are dense in 2{\mathbb{R}}^{2}, 1×\mathcal{H}_{1}\times{\mathbb{C}}, and 2\mathcal{H}_{2}, respectively.

Proof.

It suffices to prove 𝒲0(Γ)\mathcal{W}_{0}(\Gamma) is dense in 2{\mathbb{R}}^{2}. Take (ξ,η)2(\xi,\eta)\in{\mathbb{R}}^{2} and any neighborhoods XX of ξ\xi and YY of η\eta. We just need to find an element of 𝒲0(Γ)\mathcal{W}_{0}(\Gamma) in X×YX\times Y. Pick an irrational number ξoX\xi_{o}\in X.

The group UΓUU\Gamma U has finite index in SL(2,){\operatorname{SL}}(2,{\mathbb{Z}}) and so Λ(UΓU)=1()\Lambda(U\Gamma U)=\mathbb{P}^{1}({\mathbb{R}}). By Lemma 6.5, the (UΓU)(U\Gamma U)-orbit()(\infty) is dense in Λ(UΓU)=1()\Lambda(U\Gamma U)=\mathbb{P}^{1}({\mathbb{R}}), and hence (UΓU){\mathbb{Q}}\cap(U\Gamma U)-orbit()=(UΓU)(\infty)=(U\Gamma U)-orbit(){}(\infty)\setminus\{\infty\} is dense in {\mathbb{R}}. Accordingly there is a sequence xj(UΓU)x_{j}\in{\mathbb{Q}}\cap(U\Gamma U)-orbit()(\infty) with limjxj=ξo\lim_{j}x_{j}={\xi_{o}}, and we have xjXx_{j}\in X for all sufficiently large jj. Set Dj=Denom(xj)D_{j}=\operatorname{Denom}(x_{j}). Since ξo{\xi_{o}} is irrational we have limjDj=+\lim_{j}D_{j}=+\infty. Set yj=1Djfloor(ηDj)y_{j}=\frac{1}{D_{j}}\operatorname{floor}(\eta D_{j}) so that (xj,yj)𝒲0(Γ)(x_{j},y_{j})\in\mathcal{W}_{0}(\Gamma) and limjyj=η\lim_{j}y_{j}=\eta. For all sufficiently large jj we have (xj,yj)X×Y(x_{j},y_{j})\in X\times Y, which shows that 𝒲0(Γ)(X×Y)\mathcal{W}_{0}(\Gamma)\cap(X\times Y) is nonempty, and hence that 𝒲0(Γ)\mathcal{W}_{0}(\Gamma) is dense in 2{\mathbb{R}}^{2}. ∎

Let Γ\Gamma be a subgroup of SL(2,){\operatorname{SL}}(2,{\mathbb{Z}}) with finite index. For fSk(Γ)f\in S_{k}\left(\Gamma\right), define the Hecke bound HB(f)=supτ1(Im(τ))k/2|f(τ)|\operatorname{HB}(f)=\sup_{\tau\in\mathcal{H}_{1}}({\rm Im}(\tau))^{k/2}|f(\tau)|. The Hecke bound is a norm on the vector space Sk(Γ)S_{k}\left(\Gamma\right) and the Fourier coefficients a(t,):Sk(Γ)a(t,\cdot):S_{k}\left(\Gamma\right)\to{\mathbb{C}}, given by fa(t;f)f\mapsto a(t;f), become bounded linear functionals. Hence the Fourier coefficients are continuous in ff in the topology on Sk(Γ)S_{k}\left(\Gamma\right) induced by this norm. For Jacobi cusp forms ϕJk,mcusp\phi\in J_{k,m}^{\text{\rm cusp}} the Hecke bound [11], compare page 2727, is given by

HB(ϕ)=sup(τ,z)1×(Im(τ))k/2e2πm(Im(z))2Im(τ)|ϕ(τ,z)|.\operatorname{HB}(\phi)=\sup_{(\tau,z)\in\mathcal{H}_{1}\times{\mathbb{C}}}({\rm Im}(\tau))^{k/2}\,e^{-2\pi m\frac{({\rm Im}(z))^{2}}{{\rm Im}(\tau)}}\,|\phi(\tau,z)|.
Lemma 6.7.

Let ϕJk,mcusp\phi\in J_{k,m}^{\text{\rm cusp}} with mm\in{\mathbb{N}}. For all (n,r)×(n,r)\in{\mathbb{N}}\times{\mathbb{Z}} we have the bound |c(n,r;ϕ)|HB(ϕ)(eπmk|4mnr2|)k/2|c(n,r;\phi)|\leq\operatorname{HB}(\phi)\left(\frac{e\pi}{mk}|4mn-r^{2}|\right)^{k/2}.

Proof.

Let ϕ(τ,z)=n,rc(n,r;ϕ)e(nτ+rz)\phi(\tau,z)=\sum_{n\in{\mathbb{N}},\,r\in{\mathbb{Z}}}c(n,r;\phi)\,e(n\tau+rz) be the Fourier series with τ=u+iv1\tau=u+iv\in\mathcal{H}_{1} and z=x+iyz=x+iy\in{\mathbb{C}}. By the familiar formula for the Fourier coefficients, we have

c(n,r;ϕ)\displaystyle c(n,r;\phi) =[0,1]2ϕ(τ,z)e(nτrz)𝑑u𝑑x\displaystyle=\iint_{[0,1]^{2}}\phi(\tau,z)e(-n\tau-rz)\,du\,dx
|c(n,r;ϕ)|\displaystyle|c(n,r;\phi)| [0,1]2|ϕ(τ,z)|e2πnve2πry𝑑u𝑑x\displaystyle\leq\iint_{[0,1]^{2}}|\phi(\tau,z)|\,e^{2\pi nv}e^{2\pi ry}\,du\,dx
[0,1]2HB(ϕ)vk/2e2πmy2ve2π(nv+ry)𝑑u𝑑x\displaystyle\leq\iint_{[0,1]^{2}}\operatorname{HB}(\phi)v^{-k/2}e^{2\pi m\frac{y^{2}}{v}}e^{2\pi(nv+ry)}\,du\,dx
=HB(ϕ)vk/2e2π(my2v+ry+nv).\displaystyle=\operatorname{HB}(\phi)v^{-k/2}e^{2\pi(m\frac{y^{2}}{v}+ry+nv)}.

Choosing y=rv2my=-\frac{rv}{2m} we have my2v+ry+nv=4mnr24mvm\frac{y^{2}}{v}+ry+nv=\frac{4mn-r^{2}}{4m}v. Thus we have |c(n,r;ϕ)|HB(ϕ)vk/2e2π4mnr24mv|c(n,r;\phi)|\leq\operatorname{HB}(\phi)v^{-k/2}e^{2\pi\frac{4mn-r^{2}}{4m}v}. For 4mnr2>04mn-r^{2}>0 this is minimized by v=kmπ(4mnr2)v=\frac{km}{\pi(4mn-r^{2})} proving |c(n,r;ϕ)|HB(ϕ)(eπmk(4mnr2))k/2|c(n,r;\phi)|\leq\operatorname{HB}(\phi)\left(\frac{e\pi}{mk}(4mn-r^{2})\right)^{k/2}. This proves the result when 4mnr2>04mn-r^{2}>0, and when 4mnr204mn-r^{2}\leq 0 we have c(n,r;ϕ)=0c(n,r;\phi)=0 so the Lemma’s conclusion holds as well. ∎

The following proposition gives a specialization of a cuspidal formal series 𝔣𝕊(k,N;Γ)\mathfrak{f}\in\mathbb{S}(k,N;\Gamma) for each (x,y)𝒲0(Γ)(x,y)\in\mathcal{W}_{0}(\Gamma). Each specialization depends only upon a finite number of the Fourier-Jacobi coefficients of 𝔣\mathfrak{f}, although this finite number may increase with the denominator of xx.

Proposition 6.8.

Let ΓΓ0(N)\Gamma\subseteq\Gamma^{0}(N) be a subgroup of finite index. Let 𝔣=m=1ϕmξNm𝕊(k,N;Γ)\mathfrak{f}=\sum_{m=1}^{\infty}\phi_{m}\xi^{Nm}\in\mathbb{S}(k,N;\Gamma). For (x,y)𝒲0(Γ)(x,y)\in\mathcal{W}_{0}(\Gamma), set D=Denom(x)D=\operatorname{Denom}(x). We have fm(τ):=e(Nmx2τ)ϕm(τ,xτ+y)Sk(Γ(D2)).f_{m}(\tau):=e\left(Nmx^{2}\tau\right)\phi_{m}(\tau,x\tau+y)\in S_{k}\left(\Gamma(D^{2})\right). The sequence of Hecke bounds satisfies

HB(fm)O(mk+12),\operatorname{HB}(f_{m})\in O\left(m^{\frac{k+1}{2}}\right),

where the implied constant depends only on 𝔣\mathfrak{f} and xx.

Proof.

Take 𝔣=m=1ϕmξNm𝕊(k,N;Γ)\mathfrak{f}=\sum_{m=1}^{\infty}\phi_{m}\xi^{Nm}\in\mathbb{S}(k,N;\Gamma). For each mm\in{\mathbb{N}} there is a positive constant KmK_{m} such that |c(n,;ϕm)|Km(4Nmn24Nm)k/2|c(n,\ell;\phi_{m})|\leq K_{m}\left(\frac{4Nmn-\ell^{2}}{4Nm}\right)^{k/2} for all (n,)supp(ϕm)(n,\ell)\in\operatorname{supp}(\phi_{m}). Here we use the assumption that each ϕm\phi_{m} is a cusp form and Lemma 6.7. By Proposition 2.2 of [20] or Theorem 4.2 of [2], for example, we have fmSk(Γ(D2))f_{m}\in S_{k}\left(\Gamma(D^{2})\right) because Dx,DyDx,Dy\in{\mathbb{Z}}. This elliptic modular cusp form has the Fourier expansion

fm(τ)=v>0:D2va(v;fm)qv;a(v;fm)=(n,)ΛNm,x,ve(y)c(n,;ϕm),f_{m}(\tau)=\sum_{v>0:\,D^{2}v\in{\mathbb{Z}}}a(v;f_{m})q^{v};\ a(v;f_{m})=\sum_{(n,\ell)\in\Lambda_{Nm,x,v}}e(y\ell)c(n,\ell;\phi_{m}),

for ΛNm,x,v={(n,)×:Nmx2+x+n=v, 4Nmn>2}\Lambda_{Nm,x,v}=\{(n,\ell)\in{\mathbb{N}}\times{\mathbb{Z}}:Nmx^{2}+x\ell+n=v,\,4Nmn>\ell^{2}\}. Next we check |ΛNm,x,v|4Nmv+1|\Lambda_{Nm,x,v}|\leq 4\sqrt{Nmv}+1. For any (n,)ΛNm,x,v(n,\ell)\in\Lambda_{Nm,x,v} we first note that nn is determined by \ell, and that \ell satisfies

(2Nmx+)2+4Nmn2=4Nm(Nmx2+x+n)=4Nmv(2Nmx+\ell)^{2}+4Nmn-\ell^{2}=4Nm(Nmx^{2}+\ell x+n)=4Nmv.

Thus |2Nmx+|2Nmv|2Nmx+\ell|\leq 2\sqrt{Nmv} and the number of integers \ell in this interval is at most the length plus one, showing |ΛNm,x,v|4Nmv+1|\Lambda_{Nm,x,v}|\leq 4\sqrt{Nmv}+1. Pick an LL\in{\mathbb{N}} such that the Fourier coefficients a(v;f)a(v;f) for vLv\leq L determine fSk(Γ(D2))f\in S_{k}\left(\Gamma(D^{2})\right). By virtue of the valence inequality, the choice L=L(k,D)=floor(k12[SL(2,):Γ(D2)])L=L(k,D)=\operatorname{floor}\left(\frac{k}{12}[{\operatorname{SL}}(2,{\mathbb{Z}}):\Gamma(D^{2})]\right) will work. The following supremum exists because the denominator is nonzero, the quotient is continuous in ff, and the supremum may be taken over a compact sphere.

C(k,D)=supfSk(Γ(D2)){0}HB(f)0<vL(k,D)|a(v;f)|C(k,D)=\sup_{f\in S_{k}\left(\Gamma(D^{2})\right)\setminus\{0\}}\dfrac{\operatorname{HB}(f)}{\sum_{0<v\leq L(k,D)}|a(v;f)|}

Therefore we have

HB(fm)\displaystyle\operatorname{HB}(f_{m}) C(k,D)0<vL(k,D)|a(v;fm)|\displaystyle\leq C(k,D)\sum_{0<v\leq L(k,D)}|a(v;f_{m})|
C(k,D)0<vL(k,D)(n,)ΛNm,x,v|c(n,;ϕm)|\displaystyle\leq C(k,D)\sum_{0<v\leq L(k,D)}\sum_{(n,\ell)\in\Lambda_{Nm,x,v}}|c(n,\ell;\phi_{m})|
C(k,D)(n,)Λ~Nm,x,L|c(n,;ϕm)|\displaystyle\leq C(k,D)\sum_{(n,\ell)\in{\tilde{\Lambda}}_{Nm,x,L}}|c(n,\ell;\phi_{m})|

for Λ~Nm,x,L={(n,)×:Nmx2+x+nL, 4Nmn>2}{\tilde{\Lambda}}_{Nm,x,L}=\{(n,\ell)\in{\mathbb{N}}\times{\mathbb{Z}}:Nmx^{2}+x\ell+n\leq L,\,4Nmn>\ell^{2}\} or, equivalently, Λ~Nm,x,L=0<vLΛNm,x,v{\tilde{\Lambda}}_{Nm,x,L}=\cup_{0<v\leq L}\,\Lambda_{Nm,x,v}. Each ΛNm,x,v\Lambda_{Nm,x,v} has at most 4Nmv+14\sqrt{Nmv}+1 elements and there are at most D2LD^{2}L of them so we have |Λ~Nm,x,L|D2L(4NmL+1)|{\tilde{\Lambda}}_{Nm,x,L}|\leq D^{2}L(4\sqrt{NmL}+1). Note that (n,)Λ~Nm,x,L(n,\ell)\in{\tilde{\Lambda}}_{Nm,x,L} implies that 4Nmn24NmL4Nmn-\ell^{2}\leq 4NmL. This is because

4Nmn2(2Nmx+)2+4Nmn2=4Nm(Nmx2+x+n)4NmL4Nmn-\ell^{2}\leq(2Nmx+\ell)^{2}+4Nmn-\ell^{2}=4Nm(Nmx^{2}+\ell x+n)\leq 4NmL.

We now come to the main point where we use the Γ\Gamma-conditions for 𝔣\mathfrak{f}. Our assumption (x,y)𝒲0(Γ)(x,y)\in\mathcal{W}_{0}(\Gamma) gives us x(UΓU)x\in(U\Gamma U)-orbit()(\infty). Accordingly there exists a matrix (aηcξ)UΓU\left(\begin{smallmatrix}{a}&{\eta}\\ {c}&{\xi}\end{smallmatrix}\right)\in U\Gamma U with x=a/cx=a/c. Letting D=Denom(x)D=\operatorname{Denom}(x), the conditions gcd(a,c)=1\gcd(a,c)=1 and x=a/cx=a/c imply c=±Dc=\pm D and a=cxa=cx. Moving this matrix to Γ\Gamma we define

σ=(ξcηa)=U(aηcξ)UΓ.\sigma=\left(\begin{smallmatrix}{\xi}&{c}\\ {\eta}&{a}\end{smallmatrix}\right)=U\left(\begin{smallmatrix}{a}&{\eta}\\ {c}&{\xi}\end{smallmatrix}\right)U\in\Gamma.

We note that ΓΓ0(N)\Gamma\subseteq\Gamma^{0}(N) implies NDN\mid D. For (n/2/2Nm)𝒳(N)\left(\begin{smallmatrix}{n}&{\ell/2}\\ {\ell/2}&{Nm}\end{smallmatrix}\right)\in{\mathcal{X}}(N), compute σ(n/2/2Nm)σ=(n11/21/2Nm1)𝒳(N)\sigma^{\prime}\left(\begin{smallmatrix}{n}&{\ell/2}\\ {\ell/2}&{Nm}\end{smallmatrix}\right)\sigma=\left(\begin{smallmatrix}{n_{1}}&{\ell_{1}/2}\\ {\ell_{1}/2}&{Nm_{1}}\end{smallmatrix}\right)\in{\mathcal{X}}(N), where

n1\displaystyle n_{1} =Nmη2+ξη+nξ2,\displaystyle=Nm\eta^{2}+\ell\xi\eta+n\xi^{2},
1\displaystyle\ell_{1} =c(2nξ+(η+ξx)+2Nmxη),\displaystyle=c\left(2n\xi+\ell(\eta+\xi x)+2Nmx\eta\right),
m1\displaystyle m_{1} =DND(Nmx2+x+n).\displaystyle=\tfrac{D}{N}D\left(Nmx^{2}+\ell x+n\right).

For (n,)Λ~Nm,x,L(n,\ell)\in{\tilde{\Lambda}}_{Nm,x,L} we have Nmx2+x+nLNmx^{2}+\ell x+n\leq L so that m1D2NLm_{1}\leq\frac{D^{2}}{N}L, as well as (n/2/2Nm)\left(\begin{smallmatrix}{n}&{\ell/2}\\ {\ell/2}&{Nm}\end{smallmatrix}\right) and (n11/21/2Nm1)𝒳(N)\left(\begin{smallmatrix}{n_{1}}&{\ell_{1}/2}\\ {\ell_{1}/2}&{Nm_{1}}\end{smallmatrix}\right)\in{\mathcal{X}}(N). By the Γ\Gamma-conditions of equation (3) for 𝔣\mathfrak{f} we see

|c(n,;ϕm)|=|c(n1,1;ϕm1)|Km1(4Nm1n1124Nm1)k/2\displaystyle|c(n,\ell;\phi_{m})|=|c(n_{1},\ell_{1};\phi_{m_{1}})|\leq K_{m_{1}}\left(\frac{4Nm_{1}n_{1}-\ell_{1}^{2}}{4Nm_{1}}\right)^{k/2}
=Km1(4Nmn24Nm1)k/2Km1(4NmL4Nm1)k/2Km1(Lm)k/2,\displaystyle=K_{m_{1}}\left(\frac{4Nmn-\ell^{2}}{4Nm_{1}}\right)^{k/2}\leq K_{m_{1}}\left(\frac{4NmL}{4Nm_{1}}\right)^{k/2}\leq K_{m_{1}}\left(Lm\right)^{k/2},

if we remember that (n,)Λ~Nm,x,L(n,\ell)\in{\tilde{\Lambda}}_{Nm,x,L} implies 0<4Nmn24NmL0<4Nmn-\ell^{2}\leq 4NmL and note m11m_{1}\geq 1. Therefore, for (n,)Λ~Nm,x,L(n,\ell)\in{\tilde{\Lambda}}_{Nm,x,L}, we have

|c(n,;ϕm)|\displaystyle|c(n,\ell;\phi_{m})| Km1(Lm)k/2(max1jD2NLKj)Lk/2mk/2\displaystyle\leq K_{m_{1}}\left(Lm\right)^{k/2}\leq\left(\max_{1\leq j\leq\frac{D^{2}}{N}L}K_{j}\right)L^{k/2}m^{k/2}
=KD,N,LLk/2mk/2\displaystyle=K_{D,N,L}L^{k/2}m^{k/2}

since m1D2NLm_{1}\leq\frac{D^{2}}{N}L. The constant KD,N,LK_{D,N,L} grows with the denominator of (x,y)𝒲0(Γ)(x,y)\in\mathcal{W}_{0}(\Gamma) but in each individual case controls the growth of all HB(fm)\operatorname{HB}(f_{m}) in terms of the finite number of HB(ϕj)\operatorname{HB}(\phi_{j}) with jD2NLj\leq\frac{D^{2}}{N}L. Therefore we have

HB(fm)\displaystyle\operatorname{HB}(f_{m}) C(k,D)(n,)Λ~Nm,x,L|c(n,;ϕm)|\displaystyle\leq C(k,D)\sum_{(n,\ell)\in{\tilde{\Lambda}}_{Nm,x,L}}|c(n,\ell;\phi_{m})|
C(k,D)(n,)Λ~Nm,x,LKD,N,LLk/2mk/2\displaystyle\leq C(k,D)\sum_{(n,\ell)\in{\tilde{\Lambda}}_{Nm,x,L}}K_{D,N,L}L^{k/2}m^{k/2}
=C(k,D)|Λ~Nm,x,L|KD,N,LLk/2mk/2\displaystyle=C(k,D)\,|{\tilde{\Lambda}}_{Nm,x,L}|\,K_{D,N,L}L^{k/2}m^{k/2}
C(k,D)KD,N,LLk/2D2L(4NmL+1)mk/2\displaystyle\leq C(k,D)K_{D,N,L}L^{k/2}D^{2}L(4\sqrt{NmL}+1)m^{k/2}
C2(k,D,N,L)(4NmL+1)mk/2.\displaystyle\leq C_{2}(k,D,N,L)(4\sqrt{NmL}+1)m^{k/2}.

Thus HB(fm)O(mk+12)\operatorname{HB}(f_{m})\in O\left(m^{\frac{k+1}{2}}\right) as was to be shown. ∎

Consider a formal series 𝔣=m=0ϕmξNm𝕄(k,N)\mathfrak{f}=\sum_{m=0}^{\infty}\phi_{m}\xi^{Nm}\in\mathbb{M}(k,N). For a point Ωo=(τozozoωo)2\Omega_{o}=\left(\begin{smallmatrix}{\tau_{o}}&{z_{o}}\\ {z_{o}}&{\omega_{o}}\end{smallmatrix}\right)\in\mathcal{H}_{2}, if the series m=0ϕm(τo,zo)e(Nmωo)\sum_{m=0}^{\infty}\phi_{m}(\tau_{o},z_{o})e(Nm\omega_{o}) converges absolutely, we say that the formal series 𝔣\mathfrak{f} converges absolutely at Ωo\Omega_{o}.

Corollary 6.9.

Let ΓΓ0(N)\Gamma\subseteq\Gamma^{0}(N) be a subgroup of finite index. Let 𝔣=m=1ϕmξNm𝕊(k,N;Γ)\mathfrak{f}=\sum_{m=1}^{\infty}\phi_{m}\xi^{Nm}\in\mathbb{S}(k,N;\Gamma). The formal series 𝔣\mathfrak{f} converges absolutely on 𝒲(Γ)\mathcal{W}(\Gamma). For (τ1,z1)𝒲1(Γ)(\tau_{1},z_{1})\in\mathcal{W}_{1}(\Gamma) set η1=(Im(z1))2/Im(τ1)\eta_{1}=({\rm Im}(z_{1}))^{2}/{{\rm Im}(\tau_{1})}. The function H(τ1,z1,𝔣):N(η1)H(\tau_{1},z_{1},\mathfrak{f}):N_{\infty}(\eta_{1})\to{\mathbb{C}} defined by

H(τ1,z1,𝔣)(ω)=m=1ϕm(τ1,z1)e(Nmω)H(\tau_{1},z_{1},\mathfrak{f})(\omega)=\sum_{m=1}^{\infty}\phi_{m}(\tau_{1},z_{1})e\left(Nm\omega\right)

is holomorphic on N(η1)N_{\infty}(\eta_{1}).

Proof.

Pick Ω=(τxτ+yxτ+yω)𝒲(Γ)\Omega=\left(\begin{smallmatrix}{\tau}&{x\tau+y}\\ {x\tau+y}&{\omega}\end{smallmatrix}\right)\in\mathcal{W}(\Gamma) with (x,y)𝒲0(Γ)(x,y)\in\mathcal{W}_{0}(\Gamma). By Proposition 6.8, HB(fm)C(𝔣,x)mk+12\operatorname{HB}(f_{m})\leq C(\mathfrak{f},x)m^{\frac{k+1}{2}} for some constant C(𝔣,x)C(\mathfrak{f},x) and the elliptic forms fmSk(Γ(D2))f_{m}\in S_{k}\left(\Gamma(D^{2})\right) given by fm(τ)=e(Nmx2τ)ϕm(τ,xτ+y)f_{m}(\tau)=e\left(Nmx^{2}\tau\right)\phi_{m}(\tau,x\tau+y). From (Im(τ))k/2|fm(τ)|HB(fm)C(𝔣,x)mk+12({\rm Im}(\tau))^{k/2}|f_{m}(\tau)|\leq\operatorname{HB}(f_{m})\leq C(\mathfrak{f},x)m^{\frac{k+1}{2}} we infer

|ϕm(τ,xτ+y)|\displaystyle|\phi_{m}(\tau,x\tau+y)| C(𝔣,x)(Im(τ))k/2e2πNmx2(Im(τ))mk+12,\displaystyle\leq C(\mathfrak{f},x)({\rm Im}(\tau))^{-k/2}\,e^{2\pi Nmx^{2}({\rm Im}(\tau))}m^{\frac{k+1}{2}},
lim supm+|ϕm(τ,xτ+y)|1m\displaystyle\limsup_{m\to+\infty}|\phi_{m}(\tau,x\tau+y)|^{\frac{1}{m}} e2πNx2(Im(τ)).\displaystyle\leq e^{2\pi Nx^{2}({\rm Im}(\tau))}.

The radius of convergence RR of the series mϕm(τ,xτ+y)ξNm\sum_{m\in{\mathbb{N}}}\phi_{m}(\tau,x\tau+y)\xi^{Nm} thus satisfies Re2πNx2(Im(τ))R\geq e^{-2\pi Nx^{2}({\rm Im}(\tau))}, and so this series converges absolutely for |ξ|N<e2πNx2(Im(τ))2|\xi|^{N}<e^{-2\pi Nx^{2}({\rm Im}(\tau))^{2}}. In particular the series converges at Ω\Omega if Im(ω)>x2(Im(τ))=(Im(xτ+y))2Im(τ){\rm Im}(\omega)>x^{2}({\rm Im}(\tau))=\frac{({\rm Im}(x\tau+y))^{2}}{{\rm Im}(\tau)}. The condition Im(ω)Im(τ)>(Im(xτ+y))2{\rm Im}(\omega){\rm Im}(\tau)>({\rm Im}(x\tau+y))^{2}, however, is just Ω=(τxτ+yxτ+yω)2\Omega=\left(\begin{smallmatrix}{\tau}&{x\tau+y}\\ {x\tau+y}&{\omega}\end{smallmatrix}\right)\in\mathcal{H}_{2}.

Consider (τ1,z1)𝒲1(Γ)(\tau_{1},z_{1})\in\mathcal{W}_{1}(\Gamma) so that (τ1,z1)=(τ1,xτ1+y)(\tau_{1},z_{1})=(\tau_{1},x\tau_{1}+y) for some (x,y)𝒲0(Γ)(x,y)\in\mathcal{W}_{0}(\Gamma). We have seen that the power series in e(ω)e(\omega),

(8) mϕm(τ1,z1)e(Nmω),\sum_{m\in{\mathbb{N}}}\phi_{m}(\tau_{1},z_{1})e\left(Nm\omega\right),

converges when (τ1z1z1ω)2\left(\begin{smallmatrix}{\tau_{1}}&{z_{1}}\\ {z_{1}}&{\omega}\end{smallmatrix}\right)\in\mathcal{H}_{2}. A power series that converges on an open set is holomorphic there. Thus H(τ1,z1,𝔣)H(\tau_{1},z_{1},\mathfrak{f}) is holomorphic on the open set {ω1:(τ1z1z1ω)2}=N(η1)\{\omega\in\mathcal{H}_{1}:\left(\begin{smallmatrix}{\tau_{1}}&{z_{1}}\\ {z_{1}}&{\omega}\end{smallmatrix}\right)\in\mathcal{H}_{2}\}=N_{\infty}(\eta_{1}). ∎


7. Locally Bounded.

Locally bounded families of holomorphic functions possess remarkable convergence properties. In Theorem 7.4 we show that, for subgroups ΓΓ0(N)\Gamma\subseteq\Gamma^{0}(N) of finite index, the partial sums of the formal series 𝔤=m=1ϕmξNm𝕊(k,N,ϵ;Γ)\mathfrak{g}=\sum_{m=1}^{\infty}\phi_{m}\xi^{Nm}\in\mathbb{S}(k,N,\epsilon;\Gamma) are locally bounded on 2\mathcal{H}_{2} if 𝔤\mathfrak{g} is integral over FJ(𝕄±(N;Γ)){\operatorname{FJ}}\left({\mathbb{M}}_{\pm}(N;\Gamma)\right). The following homomorphism respects the ring structure but forgets the grading on 𝕄±(N){\mathbb{M}}_{\pm}(N).

Lemma 7.1.

For each (τ1,z1)1×(\tau_{1},z_{1})\in\mathcal{H}_{1}\times{\mathbb{C}} there is a ring homomorphism A(τ1,z1):𝕄±(N)[[ξ]]A(\tau_{1},z_{1}):{\mathbb{M}}_{\pm}(N)\to{\mathbb{C}}[[\xi]] defined by sending 𝔣=m=0ϕmξNm𝕄(k,N,ϵ)\mathfrak{f}=\sum_{m=0}^{\infty}\phi_{m}\xi^{Nm}\in\mathbb{M}(k,N,\epsilon) to A(τ1,z1)𝔣=m=0ϕm(τ1,z1)ξNm[[ξ]]A(\tau_{1},z_{1})\mathfrak{f}=\sum_{m=0}^{\infty}\phi_{m}(\tau_{1},z_{1})\xi^{Nm}\in{\mathbb{C}}[[\xi]] and extending additively to 𝕄±(N)=k,ϵ𝕄(k,N,ϵ){\mathbb{M}}_{\pm}(N)=\oplus_{k,\epsilon}\mathbb{M}(k,N,\epsilon).

Proof.

The ring structures are defined by the Cauchy product rule, so substitution is a homomorphism. ∎

Let D={z:|z|<r}D=\{z\in{\mathbb{C}}:|z|<r\} be the open disk about the origin in {\mathbb{C}} with positive radius rr. Let 𝒮(D)={w[[ξ]]:w converges on D}{\mathcal{S}}(D)=\{w\in{\mathbb{C}}[[\xi]]:\text{$w$ converges on~{}$D$}\} be the ring of power series that converge on DD. The map T0:𝒪(D)𝒮(D)T_{0}:{\mathcal{O}}(D)\to{\mathcal{S}}(D) that sends a holomorphic function on DD to its Taylor series about the origin is a ring isomomorphism. For ξoD\xi_{o}\in D and w=n=0wnξn𝒮(D)w=\sum_{n=0}^{\infty}w_{n}\xi^{n}\in{\mathcal{S}}(D), let w[ξo]=n=0wnξonw[\xi_{o}]=\sum_{n=0}^{\infty}w_{n}\xi_{o}^{n}\in{\mathbb{C}} be the sum of ww at ξ:=ξo\xi:=\xi_{o}. The evaluation map [ξo]:𝒮(D)[\xi_{o}]:{\mathcal{S}}(D)\to{\mathbb{C}} sending ww to w[ξo]w[\xi_{o}] is a ring homomorphism with the property that (T01w)(ξo)=w[ξo]\left(T_{0}^{-1}w\right)(\xi_{o})=w[\xi_{o}].

Lemma 7.2.

For each (τ1,z1)1×(\tau_{1},z_{1})\in\mathcal{H}_{1}\times{\mathbb{C}}, define η1=(Im(z1))2Im(τ1)\eta_{1}=\frac{({\rm Im}(z_{1}))^{2}}{{\rm Im}(\tau_{1})}, a radius r1=e2πη1r_{1}=e^{-2\pi\eta_{1}}, and a disk D1={ξ:|ξ|<r1}D_{1}=\{\xi\in{\mathbb{C}}:|\xi|<r_{1}\}. The ring homomorphism A(τ1,z1):𝕄±(N)[[ξ]]A(\tau_{1},z_{1}):{\mathbb{M}}_{\pm}(N)\to{\mathbb{C}}[[\xi]] sends FJ(M±(N)){\operatorname{FJ}}\left(M_{\pm}(N)\right) into 𝒮(D1){\mathcal{S}}(D_{1}). For ωN(η1)\omega\in N_{\infty}(\eta_{1}), set Ω1(ω)=(τ1z1z1ω)2\Omega_{1}(\omega)=\left(\begin{smallmatrix}{\tau_{1}}&{z_{1}}\\ {z_{1}}&{\omega}\end{smallmatrix}\right)\in\mathcal{H}_{2}. For fMk(K(N))ϵf\in M_{k}\left(K(N)\right)^{\epsilon} and ωN(η1)\omega\in N_{\infty}(\eta_{1}), we have (A(τ1,z1)FJ(f))[e(ω)]=f(Ω1(ω))\left(A(\tau_{1},z_{1}){\operatorname{FJ}}(f)\right)[e(\omega)]{=}f\left(\Omega_{1}(\omega)\right).

Let ΓΓ0(N)\Gamma\subseteq\Gamma^{0}(N) be a subgroup of finite index. For (τ1,z1)𝒲1(Γ)(\tau_{1},z_{1})\in\mathcal{W}_{1}(\Gamma), A(τ1,z1)A(\tau_{1},z_{1}) sends 𝕊±(N;Γ){\mathbb{S}}_{\pm}(N;\Gamma) into 𝒮(D1){\mathcal{S}}(D_{1}). For 𝔣𝕊(k,N,ϵ;Γ)\mathfrak{f}\in\mathbb{S}(k,N,\epsilon;\Gamma) and ωN(η1)\omega\in N_{\infty}(\eta_{1}) we have (A(τ1,z1)𝔣)[e(ω)]=H(τ1,z1,𝔣)(ω)\left(A(\tau_{1},z_{1})\mathfrak{f}\right)[e(\omega)]=H(\tau_{1},z_{1},\mathfrak{f})(\omega) for the holomorphic function H(τ1,z1,𝔣)H(\tau_{1},z_{1},\mathfrak{f}) defined in Corollary 6.9. The map H(τ1,z1):𝕊±(N;Γ)𝒪(N(η1))H(\tau_{1},z_{1}):{\mathbb{S}}_{\pm}(N;\Gamma)\to{\mathcal{O}}(N_{\infty}(\eta_{1})) that sends 𝔣\mathfrak{f} to H(τ1,z1,𝔣)H(\tau_{1},z_{1},\mathfrak{f}) is linear and multiplicative.

Proof.

We know that A(τ1,z1)A(\tau_{1},z_{1}) is a ring homomorphism from Lemma 7.1. We need to check that the relevant power series converge in D1D_{1} and correctly label their values. For fMk(K(N))ϵf\in M_{k}\left(K(N)\right)^{\epsilon} the Fourier-Jacobi expansion f(Ω)=m=0ϕm(τ,z)e(Nmω)f(\Omega)=\sum_{m=0}^{\infty}\phi_{m}(\tau,z)e(Nm\omega) converges for Ω=(τzzω)2\Omega=\left(\begin{smallmatrix}{\tau}&{z}\\ {z}&{\omega}\end{smallmatrix}\right)\in\mathcal{H}_{2}. The formal series is FJ(f)=m=0ϕmξNm{\operatorname{FJ}}(f)=\sum_{m=0}^{\infty}\phi_{m}\xi^{Nm} and the value of A(τ1,z1)A(\tau_{1},z_{1}) on it is A(τ1,z1)FJ(f)=m=0ϕm(τ1,z1)ξNm[[ξ]]A(\tau_{1},z_{1}){\operatorname{FJ}}(f)=\sum_{m=0}^{\infty}\phi_{m}(\tau_{1},z_{1})\xi^{Nm}\in{\mathbb{C}}[[\xi]]. For any ξoD1\xi_{o}\in D_{1}, there is an ωoN(η1)\omega_{o}\in N_{\infty}(\eta_{1}) with ξo=e(ωo)\xi_{o}=e(\omega_{o}) and Ω1(ωo)2\Omega_{1}(\omega_{o})\in\mathcal{H}_{2}. Accordingly, f(Ω1(ωo))=m=0ϕm(τ1,z1)ξoNmf(\Omega_{1}(\omega_{o}))=\sum_{m=0}^{\infty}\phi_{m}(\tau_{1},z_{1})\xi_{o}^{Nm} is convergent and, since ξoD1\xi_{o}\in D_{1} was arbitrary, A(τ1,z1)FJ(f)𝒮(D1)A(\tau_{1},z_{1}){\operatorname{FJ}}(f)\in{\mathcal{S}}(D_{1}). For any ωN(η1)\omega\in N_{\infty}(\eta_{1}) we have e(ω)D1e(\omega)\in D_{1} and (A(τ1,z1)FJ(f))[e(ω)]=m=0ϕm(τ1,z1)e(Nmω)=f(Ω1(ω))\left(A(\tau_{1},z_{1}){\operatorname{FJ}}(f)\right)[e(\omega)]=\sum_{m=0}^{\infty}\phi_{m}(\tau_{1},z_{1})e(Nm\omega)=f\left(\Omega_{1}(\omega)\right).

Assume that (τ1,z1)𝒲1(Γ)(\tau_{1},z_{1})\in\mathcal{W}_{1}(\Gamma). Since 𝔣𝕊(k,N,ϵ;Γ)𝕊(k,N;Γ)\mathfrak{f}\in\mathbb{S}(k,N,\epsilon;\Gamma)\subseteq\mathbb{S}(k,N;\Gamma), we can apply Corollary 6.9 to assure the convergence of the series H(τ1,z1,𝔣)(ω)=m=1ϕm(τ1,z1)e(Nmω)H(\tau_{1},z_{1},\mathfrak{f})(\omega){=}\sum_{m=1}^{\infty}\phi_{m}(\tau_{1},z_{1})e(Nm\omega) for ωN(η1)\omega\in N_{\infty}(\eta_{1}). Any ξD1\xi\in D_{1} may be written ξ=e(ω)\xi=e(\omega) for ωN(η1)\omega\in N_{\infty}(\eta_{1}), so the power series A(τ1,z1)𝔣=m=1ϕm(τ1,z1)ξNmA(\tau_{1},z_{1})\mathfrak{f}=\sum_{m=1}^{\infty}\phi_{m}(\tau_{1},z_{1})\xi^{Nm} converges on D1D_{1} and so A(τ1,z1)𝔣𝒮(D1)A(\tau_{1},z_{1})\mathfrak{f}\in{\mathcal{S}}(D_{1}). The value of the convergent series at e(ω)D1e(\omega)\in D_{1} is given by (A(τ1,z1)𝔣)[e(ω)]=m=1ϕm(τ1,z1)e(Nmω)=H(τ1,z1,𝔣)(ω)\left(A(\tau_{1},z_{1})\mathfrak{f}\right)[e(\omega)]=\sum_{m=1}^{\infty}\phi_{m}(\tau_{1},z_{1})e(Nm\omega)=H(\tau_{1},z_{1},\mathfrak{f})(\omega).

Both A(τ1,z1)A(\tau_{1},z_{1}) and the evaluation map [e(ω)][e(\omega)] are ring homomorphisms so their composition, restricted to the ideal 𝕊±(N;Γ){\mathbb{S}}_{\pm}(N;\Gamma) is linear and multiplicative. This shows, for all ωN(η1)\omega\in N_{\infty}(\eta_{1}), that we have

H(τ1,z1,𝔣1𝔣2)(ω)=(A(τ1,z1)𝔣1𝔣2)[e(ω)]=\displaystyle H(\tau_{1},z_{1},\mathfrak{f}_{1}\mathfrak{f}_{2})(\omega)=\left(A(\tau_{1},z_{1})\mathfrak{f}_{1}\mathfrak{f}_{2}\right)[e(\omega)]=
(A(τ1,z1)𝔣1)[e(ω)](A(τ1,z1)𝔣2)[e(ω)]=H(τ1,z1,𝔣1)(ω)H(τ1,z1,𝔣2)(ω),\displaystyle\left(A(\tau_{1},z_{1})\mathfrak{f}_{1}\right)[e(\omega)]\left(A(\tau_{1},z_{1})\mathfrak{f}_{2}\right)[e(\omega)]=H(\tau_{1},z_{1},\mathfrak{f}_{1})(\omega)H(\tau_{1},z_{1},\mathfrak{f}_{2})(\omega),

which is what H(τ1,z1,𝔣1𝔣2)=H(τ1,z1,𝔣1)H(τ1,z1,𝔣2)H(\tau_{1},z_{1},\mathfrak{f}_{1}\mathfrak{f}_{2})=H(\tau_{1},z_{1},\mathfrak{f}_{1})H(\tau_{1},z_{1},\mathfrak{f}_{2}) means. ∎

Lemma 7.3.

Let dd\in{\mathbb{N}} and zz\in{\mathbb{C}} be given. Let a monic polynomial P[X]P\in{\mathbb{C}}[X] of degree dd be given by P(X)=Xd+j=1dajXdjP(X)=X^{d}+\sum_{j=1}^{d}a_{j}X^{d-j} for a1,,ada_{1},\ldots,a_{d}\in{\mathbb{C}}. If P(z)=0P(z)=0 then |z|1+j=1d|aj||z|\leq 1+\sum_{j=1}^{d}|a_{j}|.

Proof.

Apply the Triangle Inequality to the case |z|1|z|\geq 1. ∎

Theorem 7.4.

Let N,k,dN,k,d\in{\mathbb{N}}. Let ϵ{±1}\epsilon\in\{\pm 1\}. Let ΓΓ0(N)\Gamma\subseteq\Gamma^{0}(N) be a subgroup of finite index. Let 𝔤=m=1ϕmξNm𝕊(k,N,ϵ;Γ)\mathfrak{g}=\sum_{m=1}^{\infty}\phi_{m}\xi^{Nm}\in\mathbb{S}(k,N,\epsilon;\Gamma) satisfy the monic polynomial relation P(𝔤)=0P(\mathfrak{g})=0 in 𝕄(kd,N,ϵd;Γ)\mathbb{M}(kd,N,\epsilon^{d};\Gamma) for

P(X)=Xd++FJ(gj)Xdj++FJ(gd)P(X)=X^{d}+\cdots+{\operatorname{FJ}}(g_{j})X^{d-j}+\cdots+{\operatorname{FJ}}(g_{d})

with gjMkj(K(N))ϵjg_{j}\in M_{kj}\left(K(N)\right)^{\epsilon^{j}} for j=1,,dj=1,\ldots,d.

Then the sequence of partial sums m=1Mϕm(τ,z)e(Nmω)\sum_{m=1}^{M}\phi_{m}(\tau,z)e\left(Nm\omega\right) for MM\in{\mathbb{N}} is locally bounded on (τzzω)2\left(\begin{smallmatrix}{\tau}&{z}\\ {z}&{\omega}\end{smallmatrix}\right)\in\mathcal{H}_{2}.

Proof.

Pick Ωo2\Omega_{o}\in\mathcal{H}_{2}. We need to make a neighborhood B1B_{1} of Ωo\Omega_{o} and a positive constant A1A_{1} such that for all MM\in{\mathbb{N}} and all (τzzω)B1\left(\begin{smallmatrix}{\tau}&{z}\\ {z}&{\omega}\end{smallmatrix}\right)\in B_{1} we have |m=1Mϕm(τ,z)e(Nmω)|A1|\sum_{m=1}^{M}\phi_{m}(\tau,z)e\left(Nm\omega\right)|\leq A_{1}. Let B1B_{1} be an open Euclidean ball centered at Ωo\Omega_{o} with closure B¯12\overline{B}_{1}\subseteq\mathcal{H}_{2}. We can push the closed ball B¯1\overline{B}_{1} down by i(000ϵ)-i\left(\begin{smallmatrix}{0}&{0}\\ {0}&{\epsilon}\end{smallmatrix}\right) so that this translation is still contained in 2\mathcal{H}_{2}. We do this in detail. For any ϵ>0\epsilon>0, the translated ball lies in the space L={(τzzω)M2×2sym():τ1}L=\{\left(\begin{smallmatrix}{\tau}&{z}\\ {z}&{\omega}\end{smallmatrix}\right)\in M_{2\times 2}^{\text{\rm sym}}({\mathbb{C}}):\tau\in\mathcal{H}_{1}\}. Define the function h:Lh:L\to{\mathbb{R}} for any Ω=(τzzω)L\Omega=\left(\begin{smallmatrix}{\tau}&{z}\\ {z}&{\omega}\end{smallmatrix}\right)\in L by

h(Ω)=det(Im(Ω))Im(Ω),(1000)=Im(ω)(Im(z))2Im(τ).h(\Omega)=\frac{\det({\rm Im}(\Omega))}{\langle{\rm Im}(\Omega),\left(\begin{smallmatrix}{1}&{0}\\ {0}&{0}\end{smallmatrix}\right)\rangle}={\rm Im}(\omega)-\frac{({\rm Im}(z))^{2}}{{\rm Im}(\tau)}.

The function hh is continuous on LL and positive on 2L\mathcal{H}_{2}\subseteq L and so has a positive minimum on B¯1\overline{B}_{1}. Set ϵ=12minΩB¯1h(Ω)>0\epsilon=\frac{1}{2}\min_{\Omega\in\overline{B}_{1}}h(\Omega)>0. We remark that ϵ\epsilon only depends on the choice of the ball B1B_{1}. Translate the ball B1B_{1} by i(000ϵ)-i\left(\begin{smallmatrix}{0}&{0}\\ {0}&{\epsilon}\end{smallmatrix}\right) to obtain the ball B2={Ωi(000ϵ)L:ΩB1}B_{2}=\{\Omega-i\left(\begin{smallmatrix}{0}&{0}\\ {0}&{\epsilon}\end{smallmatrix}\right)\in L:\Omega\in B_{1}\} with closure in LL given by B¯2={Ωi(000ϵ)L:ΩB¯1}\overline{B}_{2}=\{\Omega-i\left(\begin{smallmatrix}{0}&{0}\\ {0}&{\epsilon}\end{smallmatrix}\right)\in L:\Omega\in\overline{B}_{1}\}. We have

infΩB¯2h(Ω)=infΩB¯1(h(Ω)ϵ)=2ϵϵ=ϵ>0.\inf_{\Omega\in\overline{B}_{2}}h(\Omega)=\inf_{\Omega\in\overline{B}_{1}}\left(h(\Omega)-\epsilon\right)=2\epsilon-\epsilon=\epsilon>0.

Therefore det(Im(Ω))>0\det({\rm Im}(\Omega))>0 for ΩB¯2\Omega\in\overline{B}_{2} and B¯22\overline{B}_{2}\subseteq\mathcal{H}_{2}. By this process, B2B_{2} is completely determined by B1B_{1}. Next define a compact set K2={Ω+(000μ)2:ΩB¯2,0μ1}K_{2}=\{\Omega+\left(\begin{smallmatrix}{0}&{0}\\ {0}&{\mu}\end{smallmatrix}\right)\in\mathcal{H}_{2}:\Omega\in\overline{B}_{2},0\leq\mu\leq 1\} that contains B¯2\overline{B}_{2}. Use the continuity of the gjg_{j} to define A0=A0(B1,P)=supΩK2(1+j=1d|gj(Ω)|)A_{0}=A_{0}(B_{1},P)=\sup_{\Omega\in K_{2}}\left(1+\sum_{j=1}^{d}|g_{j}(\Omega)|\right). We will show that A1=A0e2πNϵ1A_{1}=\frac{A_{0}}{e^{2\pi N\epsilon}-1} works as the local bound at B1B_{1}.

The main step will be to show |m=1Mϕm(τ1,z1)e(Nmω1)|A1|\sum_{m=1}^{M}\phi_{m}(\tau_{1},z_{1})e\left(Nm\omega_{1}\right)|\leq A_{1} for every MM\in{\mathbb{N}} and every Ω1=(τ1z1z1ω1)B1𝒲(Γ)\Omega_{1}=\left(\begin{smallmatrix}{\tau_{1}}&{z_{1}}\\ {z_{1}}&{\omega_{1}}\end{smallmatrix}\right)\in B_{1}\cap\mathcal{W}(\Gamma). The conclusion |m=1Mϕm(τ,z)e(Nmω)|A1|\sum_{m=1}^{M}\phi_{m}(\tau,z)e\left(Nm\omega\right)|\leq A_{1} on the neighborhood B1B_{1} then follows because m=1Mϕm(τ,z)e(Nmω)\sum_{m=1}^{M}\phi_{m}(\tau,z)e\left(Nm\omega\right) is continuous on B1B_{1} and B1𝒲(Γ)B_{1}\cap\mathcal{W}(\Gamma) is dense in B1B_{1}. The assumption Ω1B1𝒲(Γ)\Omega_{1}\in B_{1}\cap\mathcal{W}(\Gamma) implies (τ1,z1)𝒲1(Γ)(\tau_{1},z_{1})\in\mathcal{W}_{1}(\Gamma). We may rewrite the hypothesis

𝔤d++FJ(gj)𝔤dj++FJ(gd)=0\mathfrak{g}^{d}+\cdots+{\operatorname{FJ}}(g_{j})\mathfrak{g}^{d-j}+\cdots+{\operatorname{FJ}}(g_{d})=0

as j=0dFJ(gj)𝔤dj=0\sum_{j=0}^{d}{\operatorname{FJ}}(g_{j})\mathfrak{g}^{d-j}=0 where g0=1M0(K(N))+g_{0}=1\in M_{0}(K(N))^{+}. Apply the ring homomorphism A(τ1,z1):𝕄±(N)[[ξ]]A(\tau_{1},z_{1}):{\mathbb{M}}_{\pm}(N)\to{\mathbb{C}}[[\xi]] to the given relation j=0dFJ(gj)𝔤dj=0\sum_{j=0}^{d}{\operatorname{FJ}}(g_{j})\mathfrak{g}^{d-j}=0 in 𝕄(kd,N,ϵd;Γ)\mathbb{M}(kd,N,\epsilon^{d};\Gamma) to obtain a relation among formal power series

(9) j=0d(A(τ1,z1)FJ(gj))(A(τ1,z1)𝔤)dj=0.\sum_{j=0}^{d}\left(A(\tau_{1},z_{1}){\operatorname{FJ}}(g_{j})\right)\left(A(\tau_{1},z_{1})\mathfrak{g}\right)^{d-j}=0.

By Lemma 7.2, we know A(τ1,z1)𝔤𝒮(D1)A(\tau_{1},z_{1})\mathfrak{g}\in{\mathcal{S}}(D_{1}) for η1=(Im(z1))2Im(τ1)\eta_{1}=\frac{({\rm Im}(z_{1}))^{2}}{{\rm Im}(\tau_{1})}, the radius r1=e2πη11r_{1}=e^{-2\pi\eta_{1}}\leq 1, and the disk D1={z:|z|<r1}D_{1}=\{z\in{\mathbb{C}}:|z|<r_{1}\}. Moreover, the power series A(τ1,z1)FJ(gj)A(\tau_{1},z_{1}){\operatorname{FJ}}(g_{j}) converges in D1D_{1}. Therefore the formal power series in equation (9) is in the subring 𝒮(D1){\mathcal{S}}(D_{1}) of convergent power series on D1D_{1}.

For ωN(η1)\omega\in N_{\infty}(\eta_{1}), we have ξ1=e(ω)D1\xi_{1}=e(\omega)\in D_{1} and so we may use the evaluation homomorphism [ξ1]:𝒮(D1)[\xi_{1}]:{\mathcal{S}}(D_{1})\to{\mathbb{C}} to obtain a relation among complex numbers

(10) j=0d(A(τ1,z1)FJ(gj))[ξ1]((A(τ1,z1)𝔤)[ξ1])dj=0.\sum_{j=0}^{d}\left(A(\tau_{1},z_{1}){\operatorname{FJ}}(g_{j})\right)[\xi_{1}]\left(\left(A(\tau_{1},z_{1})\mathfrak{g}\right)[\xi_{1}]\right)^{d-j}=0.

For ωN(η1)\omega\in N_{\infty}(\eta_{1}), define Ω1(ω)=(τ1z1z1ω)2\Omega_{1}(\omega)=\left(\begin{smallmatrix}{\tau_{1}}&{z_{1}}\\ {z_{1}}&{\omega}\end{smallmatrix}\right)\in\mathcal{H}_{2}. By Lemma 7.2, for any ωN(η1)\omega\in N_{\infty}(\eta_{1}), we have both (A(τ1,z1)FJ(gj))[ξ1]=gj(Ω1(ω))\left(A(\tau_{1},z_{1}){\operatorname{FJ}}(g_{j})\right)[\xi_{1}]=g_{j}\left(\Omega_{1}(\omega)\right) and (A(τ1,z1)𝔤)[ξ1]=H(τ1,z1,𝔤)(ω)\left(A(\tau_{1},z_{1})\mathfrak{g}\right)[\xi_{1}]=H(\tau_{1},z_{1},\mathfrak{g})(\omega). We rewrite equation (10) as

(H(τ1,z1,𝔤)(ω))d+j=1dgj(Ω1(ω))(H(τ1,z1,𝔤)(ω))dj=0.\left(H(\tau_{1},z_{1},\mathfrak{g})(\omega)\right)^{d}+\sum_{j=1}^{d}g_{j}\left(\Omega_{1}(\omega)\right)\left(H(\tau_{1},z_{1},\mathfrak{g})(\omega)\right)^{d-j}=0.

By Lemma 7.3 we have

(11) |H(τ1,z1,𝔤)(ω)|1+j=1d|gj(Ω1(ω))|.|H(\tau_{1},z_{1},\mathfrak{g})(\omega)|\leq 1+\sum_{j=1}^{d}|g_{j}\left(\Omega_{1}(\omega)\right)|.

By the definition of A0=supΩK2(1+j=1d|gj(Ω)|)A_{0}=\sup_{\Omega\in K_{2}}\left(1+\sum_{j=1}^{d}|g_{j}(\Omega)|\right), we have

(12) |H(τ1,z1,𝔤)(ω)|A0,if Ω1(ω)K2.|H(\tau_{1},z_{1},\mathfrak{g})(\omega)|\leq A_{0},\quad\text{if $\Omega_{1}(\omega)\in K_{2}$.}

We have ω1iϵN(η1)={ω1:Ω1(ω)2}\omega_{1}-i\epsilon\in N_{\infty}(\eta_{1})=\{\omega\in\mathcal{H}_{1}:\Omega_{1}(\omega)\in\mathcal{H}_{2}\} because Ω1(ω1iϵ)B22\Omega_{1}\left(\omega_{1}-i\epsilon\right)\in B_{2}\subseteq\mathcal{H}_{2}. Furthermore, for 0μ10\leq\mu\leq 1, we have ω1iϵ+μN(η1)\omega_{1}-i\epsilon+\mu\in N_{\infty}(\eta_{1}) because Ω1(ω1iϵ+μ)K22\Omega_{1}\left(\omega_{1}-i\epsilon+\mu\right)\in K_{2}\subseteq\mathcal{H}_{2}, and the function H(τ1,z1,𝔤)H(\tau_{1},z_{1},\mathfrak{g}) is holomorphic on ω1iϵ+[0,1]N(η1)\omega_{1}-i\epsilon+[0,1]\subseteq N_{\infty}(\eta_{1}). Thus we may compute the Fourier coefficients of

H(τ1,z1,𝔤)(ω)=m=1ϕm(τ1,z1)e(Nmω)H(\tau_{1},z_{1},\mathfrak{g})(\omega)=\sum_{m=1}^{\infty}\phi_{m}(\tau_{1},z_{1})e\left(Nm\omega\right)

by the familiar formula

ϕm(τ1,z1)=01H(τ1,z1,𝔤)(ω1iϵ+μ)e(Nm(ω1iϵ+μ))𝑑μ.\phi_{m}(\tau_{1},z_{1})=\int_{0}^{1}H(\tau_{1},z_{1},\mathfrak{g})(\omega_{1}-i\epsilon+\mu)e\left(-Nm(\omega_{1}-i\epsilon+\mu)\right)\,d\mu.

We calculate, for any MM\in{\mathbb{N}},

|m=1Mϕm(τ1,z1)e(Nmω1)|=\displaystyle|\sum_{m=1}^{M}\phi_{m}(\tau_{1},z_{1})e\left(Nm\omega_{1}\right)|=
|m=1M(01H(τ1,z1,𝔤)(ω1iϵ+μ)e(Nm(ω1iϵ+μ))𝑑μ)e(Nmω1)|\displaystyle|\sum_{m=1}^{M}\left(\int_{0}^{1}H(\tau_{1},z_{1},\mathfrak{g})(\omega_{1}{-}i\epsilon{+}\mu)e\left(-Nm(\omega_{1}{-}i\epsilon{+}\mu)\right)\,d\mu\right)e\left(Nm\omega_{1}\right)|
m=1Me2πNmIm(ω1)01|H(τ1,z1,𝔤)(ω1iϵ+μ)|e2πNm(Im(ω1)ϵ)𝑑μ\displaystyle\leq\sum_{m=1}^{M}e^{-2\pi Nm\,{\rm Im}(\omega_{1})}\int_{0}^{1}|H(\tau_{1},z_{1},\mathfrak{g})(\omega_{1}-i\epsilon+\mu)|e^{2\pi Nm({\rm Im}(\omega_{1})-\epsilon)}\,d\mu
=m=1Me2πNmϵ01|H(τ1,z1,𝔤)(ω1iϵ+μ)|𝑑μ.\displaystyle=\sum_{m=1}^{M}e^{-2\pi Nm\epsilon}\int_{0}^{1}|H(\tau_{1},z_{1},\mathfrak{g})(\omega_{1}-i\epsilon+\mu)|\,d\mu.

As mentioned above, we have Ω1(ω1iϵ+μ)=(τ1z1z1ω1iϵ+μ)K2\Omega_{1}\left(\omega_{1}-i\epsilon+\mu\right)=\left(\begin{smallmatrix}{\tau_{1}}&{z_{1}}\\ {z_{1}}&{\omega_{1}-i\epsilon+\mu}\end{smallmatrix}\right)\in K_{2} for 0μ10\leq\mu\leq 1, so that |H(τ1,z1,𝔤)(ω1iϵ+μ)|A0|H(\tau_{1},z_{1},\mathfrak{g})(\omega_{1}-i\epsilon+\mu)|\leq A_{0} by equation (12). Thus

|\displaystyle| m=1Mϕm(τ1,z1)e(Nmω1)|\displaystyle\sum_{m=1}^{M}\phi_{m}(\tau_{1},z_{1})e\left(Nm\omega_{1}\right)|
A0\displaystyle\leq A_{0} m=1Me2πNmϵ<A0e2πNϵ1e2πNϵ=A0e2πNϵ1=A1.\displaystyle\sum_{m=1}^{M}e^{-2\pi Nm\epsilon}<\dfrac{A_{0}e^{-2\pi N\epsilon}}{1-e^{-2\pi N\epsilon}}=\dfrac{A_{0}}{e^{2\pi N\epsilon}-1}=A_{1}.

This completes the main step and, as explained, consequently shows that |m=1Mϕm(τ,z)e(Nmω)|A1|\sum_{m=1}^{M}\phi_{m}(\tau,z)e\left(Nm\omega\right)|\leq A_{1} on the neighborhood B1B_{1} of Ωo\Omega_{o} for all MM\in{\mathbb{N}}. Since Ωo2\Omega_{o}\in\mathcal{H}_{2} was arbitrary, the sequence of partial sums m=1Mϕm(τ,z)e(Nmω)\sum_{m=1}^{M}\phi_{m}(\tau,z)e\left(Nm\omega\right) is locally bounded on 2\mathcal{H}_{2}. ∎

8. Holomorphicity.

In Theorem 8.2 we show that the divisibility of paramodular Fricke eigenforms is implied by the cuspidality of the quotient of their formal series. Later, after proving our main result, Theorem 10.1 improves Theorem 8.2 by dropping the cuspidality assumption.

We review divisor theory following Gunning [21]. Let GnG\subseteq{\mathbb{C}}^{n} be a domain. For open UGU\subseteq G, let 𝒪(U)\mathcal{O}(U) be the ring of holomorphic functions on UU. For pGp\in G, let 𝒪p\mathcal{O}_{p} be the ring of germs of holomorphic functions, and p\mathcal{M}_{p} the field of germs of meromorphic functions at pp. A holomorphic subvariety is a subset VGV\subseteq G such that each point pVp\in V has a neighborhood UU where UVU\cap V is the zero set of finitely many functions holomorphic on UU. A holomorphic subvariety VV is irreducible if V=V1V2V=V_{1}\cup V_{2} for holomorphic subvarieties V1V_{1}, V2V_{2}, implies V1=VV_{1}=V or V2=VV_{2}=V. The regular points of VV, (V)\mathcal{R}(V), are the points pVp\in V where UVU\cap V is a complex manifold for some neighborhood UU of pp. The regular points are an open dense subset of VV, locally finitely connected, and (V)\mathcal{R}(V) is connected precisely when VV is irreducible. The dimension, dimV\dim V, of an irreducible VV is the dimension of the complex manifold (V)\mathcal{R}(V), and the codimension of VV is ndimVn-\dim V.

We let (G)\mathcal{E}(G) be the set of irreducible, codimension one holomorphic subvarieties of GG. For a function Δ:(G)\Delta:\mathcal{E}(G)\to{\mathbb{Z}}, the support of Δ\Delta is supp(Δ)={V(G):Δ(V)0}\operatorname{supp}(\Delta)=\{V\in\mathcal{E}(G):\Delta(V)\neq 0\}. We say Δ\Delta is locally finite if, for all open UGU\subseteq G with U¯G{\bar{U}}\subseteq G, the set {Vsupp(Δ):UV}\{V\in\operatorname{supp}(\Delta):U\cap V\neq\emptyset\} is finite. The group of divisors in GG, 𝒟G\mathcal{D}_{G}, is the abelian group of locally finite functions Δ:(G)\Delta:\mathcal{E}(G)\to{\mathbb{Z}}. Since supp(Δ)\operatorname{supp}(\Delta) is countable, we often write a global divisor as Δ=jνjVj\Delta=\sum_{j\in{\mathbb{N}}}\nu_{j}V_{j}, for νj=Δ(Vj)\nu_{j}=\Delta(V_{j})\in{\mathbb{Z}} and distinct Vjsupp(Δ)V_{j}\in\operatorname{supp}(\Delta). The semigroup of effective divisors, 𝒟G+𝒟G\mathcal{D}_{G}^{+}\subseteq\mathcal{D}_{G}, is defined by νj0\nu_{j}\geq 0.

For the germ of a holomorphic subvariety at pp, VpV_{p}, the dimension, dimVp\dim V_{p}, is the maximal dimension of the finitely many connected components of (V)\mathcal{R}(V) whose closure contains pp. We say VpV_{p} is pure dimensionsal when all these connected components have the same dimension. The definition of irreducibility for VpV_{p} is as before. Let (p)\mathcal{E}(p) be the set of germs of holomorphic subvarieties at pp that are irreducible and have codimension one. A function δ:(p)\delta:\mathcal{E}(p)\to{\mathbb{Z}} has support supp(δ)={Vp(p):δ(Vp)0}\operatorname{supp}(\delta)=\{V_{p}\in\mathcal{E}(p):\delta(V_{p})\neq 0\}. The group of local divisors at pp, 𝒟p\mathcal{D}_{p}, is the free abelian group on (p)\mathcal{E}(p). We write local divisors δ𝒟p\delta\in\mathcal{D}_{p} as δ=j=1νjVj\delta=\sum_{j=1}^{\ell}\nu_{j}V_{j} with νj=δ(Vj)\nu_{j}=\delta(V_{j})\in{\mathbb{Z}} and distinct Vj(p)V_{j}\in\mathcal{E}(p) but, fundamentally, a local divisor at pp is a function δ:(p)\delta:\mathcal{E}(p)\to{\mathbb{Z}} with finite support. The semigroup of effective local divisors, 𝒟p+𝒟p\mathcal{D}_{p}^{+}\subseteq\mathcal{D}_{p}, is defined by νj0\nu_{j}\geq 0. If p(V)p\in\mathcal{R}(V) then VpV_{p} is irreducible. For V(G)V\in\mathcal{E}(G) and general pVp\in V, Vp=VpVp′′V_{p}=V_{p}^{\prime}\cup V_{p}^{\prime\prime}\cup\cdots decomposes into a finite union of distinct Vp,Vp′′,(p)V_{p}^{\prime},V_{p}^{\prime\prime},\ldots\in\mathcal{E}(p).

The ring 𝒪p\mathcal{O}_{p} is a unique factorization domain and noetherian. The fundamental correspondence of algebraic geometry holds between germs of holomorphic subvarieties and germs of holomorphic functions at pp. The ideal id(Vp)𝒪p\operatorname{id}(V_{p})\subseteq\mathcal{O}_{p} consists of the germs that vanish on VpV_{p} and, for an ideal 𝔞𝒪p\mathfrak{a}\subseteq\mathcal{O}_{p}, loc(𝔞)\operatorname{loc}(\mathfrak{a}) is the germ of the holomorphic subvariety defined by the simultaneous vanishing of the elements of 𝔞\mathfrak{a}. The germ VpV_{p} is irreducible precisely when id(Vp)\operatorname{id}(V_{p}) is prime, and VpV_{p} has pure codimension one precisely when id(Vp)\operatorname{id}(V_{p}) is principal. Therefore, for V(G)V\in\mathcal{E}(G) and pVp\in V, there is a neigborhood UU of pp and a ϖ𝒪(U)\varpi\in\mathcal{O}(U), with ϖp\varpi_{p} prime in 𝒪p\mathcal{O}_{p}, such that VU={ZU:ϖ(Z)=0}V\cap U=\{Z\in U:\varpi(Z)=0\}. We call this a local equation at pp and refer to ϖ\varpi as a uniformizer. As a consequence of Cartan’s Theorem, see Theorem F6 of [21], there is a neighborhood U′′UU^{\prime\prime}\subseteq U of pp such that ϖq\varpi_{q} generates idVq\operatorname{id}V_{q} at all points qU′′Vq\in U^{\prime\prime}\cap V. Cartan’s Theorem asserts the existence of a neighborhood UUU^{\prime}\subseteq U of pp and of functions h1,,h𝒪(U)h_{1},\ldots,h_{\ell}\in\mathcal{O}(U^{\prime}) such that idVq=𝒪q(h1q,,hq)\operatorname{id}V_{q}=\mathcal{O}_{q}(h_{1q},\ldots,h_{\ell q}) at all points qUVq\in U^{\prime}\cap V. Since ϖp\varpi_{p} divides each hjph_{jp} there is a neighborhood U′′UU^{\prime\prime}\subseteq U^{\prime} of pp where idVq=𝒪q(h1q,,hq)𝒪qϖq\operatorname{id}V_{q}=\mathcal{O}_{q}(h_{1q},\ldots,h_{\ell q})\subseteq\mathcal{O}_{q}\varpi_{q} at all points qU′′Vq\in U^{\prime\prime}\cap V. Hence idVq=𝒪qϖq\operatorname{id}V_{q}=\mathcal{O}_{q}\varpi_{q} and ϖq\varpi_{q} is prime for all qU′′Vq\in U^{\prime\prime}\cap V.

For a ring RR, let R×R^{\times} be the group of units and R=R{0}R^{*}=R\,{\setminus}\{0\}. The sequence of semigroups 0(𝒪p×,)(𝒪p,)divp(𝒟p+,+)00{\to}\left(\mathcal{O}_{p}^{\times},\cdot\right){\to}\left(\mathcal{O}_{p}^{*},\cdot\right)\overset{\operatorname{div}_{p}}{\to}\left(\mathcal{D}_{p}^{+},+\right){\to}0 is exact, where divp(fp)=j=1νjloc(𝒪pfj)\operatorname{div}_{p}(f_{p})=\sum_{j=1}^{\ell}\nu_{j}\operatorname{loc}(\mathcal{O}_{p}f_{j}) is the semigroup homomorphism determined by the factorization fp=(unit)f1ν1fνf_{p}=\text{\rm(unit)}f_{1}^{\nu_{1}}\cdots f_{\ell}^{\nu_{\ell}} in 𝒪p\mathcal{O}_{p} into powers of nonassociate irreducibles fjf_{j}, unique up to order and units. The rule divp(fp/gp)=divp(fp)divp(gp)\operatorname{div}_{p}\left(f_{p}/g_{p}\right)=\operatorname{div}_{p}(f_{p})-\operatorname{div}_{p}(g_{p}) defines a group homomorphism divp:(p,)(𝒟p,+)\operatorname{div}_{p}:\left(\mathcal{M}_{p}^{*},\cdot\right)\to\left(\mathcal{D}_{p},+\right). The local divisor divp(fp)\operatorname{div}_{p}(f_{p}) is effective precisely when fp𝒪ppf_{p}\in\mathcal{O}_{p}^{*}\subseteq\mathcal{M}_{p}^{*}. To each global divisor Δ𝒟G\Delta\in\mathcal{D}_{G} and point pGp\in G, we associate a local divisor, germp(Δ)=Vsupp(Δ):pVΔ(V)(Vp+Vp′′+)\operatorname{germ}_{p}(\Delta)=\sum_{V\in\operatorname{supp}(\Delta):\hskip 1.4457ptp\in V}\Delta(V)\left(V_{p}^{\prime}+V_{p}^{\prime\prime}+\cdots\right), where Vp=VpVp′′V_{p}=V_{p}^{\prime}\cup V_{p}^{\prime\prime}\cup\cdots is the finite decomposition of VpV_{p} into irreducibles. For each f𝒪(G)f\in\mathcal{O}(G)^{*}, the global divisor div(f)\operatorname{div}(f) is the unique Δ𝒟G+\Delta\in\mathcal{D}_{G}^{+} such that germp(Δ)=divp(fp)\operatorname{germ}_{p}(\Delta)=\operatorname{div}_{p}(f_{p}) for all pGp\in G. The rule div(f/g)=div(f)div(g)\operatorname{div}\left(f/g\right)=\operatorname{div}(f)-\operatorname{div}(g) extends div\operatorname{div} to a group homomorphism div:((G),)(𝒟G,+)\operatorname{div}:\left(\mathcal{M}(G)^{*},\cdot\right)\to\left(\mathcal{D}_{G},+\right), where the field of meromorphic functions (G)\mathcal{M}(G) is the quotient field of 𝒪(G)\mathcal{O}(G). The global assertion that div(f)\operatorname{div}(f) is effective precisely when f𝒪(G)(G)f\in\mathcal{O}(G)^{*}\subseteq\mathcal{M}(G)^{*} follows from the corresponding local assertion.

For V(G)V\in\mathcal{E}(G) and f𝒪(G)f\in\mathcal{O}(G)^{*}, a direct way to compute ν=ordV(f)=(div(f))(V)\nu={\operatorname{ord}}_{V}(f)=\left(\operatorname{div}(f)\right)(V) is to take a regular point p(V)p\in\mathcal{R}(V) and a local equation at pp, VU={ZU:ϖ(Z)=0}V\cap U=\{Z\in U:\varpi(Z)=0\}, so that VpV_{p} is irreducible and ϖp\varpi_{p} is prime with 𝒪pϖp=id(Vp)\mathcal{O}_{p}\varpi_{p}=\operatorname{id}(V_{p}). The factor ϖpν\varpi_{p}^{\nu} in the unique factorization of fpf_{p} in 𝒪p\mathcal{O}_{p} defines ν0\nu\in{\mathbb{N}}_{0}. This ν\nu is independent of the choice of the regular point pp and of the uniformizer ϖ\varpi.

We use cN=(1N01)GL(2,)\operatorname{c_{N}}=\left(\begin{smallmatrix}{-1}&{-N}\\ {0}&{1}\end{smallmatrix}\right)\in{\operatorname{GL}}(2,{\mathbb{Z}}) and CN=(cN00cN)K(N)\operatorname{C_{N}}=\left(\begin{smallmatrix}{\operatorname{c_{N}}}&{0}\\ {0}&{\operatorname{c_{N}}^{*}}\end{smallmatrix}\right)\in K(N) in the following lemma. The usefulness of the following lemma is that a point on the K(N)+K(N)^{+}-orbit of a codimension one irreducible holomorphic subvariety of 2\mathcal{H}_{2} can be found where the uniformizer ϖ\varpi is regular in ω\omega at that point. This allows us to use the Weierstrauss preparation theorem on ϖ\varpi in the proof of Theorem 8.2. In Lemma 8.1, the term “regular” regettably has two distinct meanings. A holomorphic ϖ\varpi is regular in ω\omega at (τo,zo,ωo)(\tau_{o},z_{o},\omega_{o}) means ϖ(τo,zo,ω)0\varpi(\tau_{o},z_{o},\omega)\not\equiv 0 in all neighborhoods of ωo\omega_{o}, whereas a point pp of a holomorphic subvariety VV is regular if VUV\cap U is a complex manifold for some neighborhood UU of pp.

Lemma 8.1.

Let k,Nk,N\in{\mathbb{N}}, and ϵ{±1}\epsilon\in\{\pm 1\}. Let fMk(K(N))ϵf\in M_{k}\left(K(N)\right)^{\epsilon}. Let V2V\subseteq\mathcal{H}_{2} be an irreducible component of div(f)\operatorname{div}(f). For any gK(N)+g\in K(N)^{+}, Vg=gVV_{g}=g\langle V\rangle is also an irreducible component of div(f)\operatorname{div}(f), and we have ordVf=ordVgf{\operatorname{ord}}_{V}f={\operatorname{ord}}_{V_{g}}f.

There is a g{I,μN,CN}K(N)+g\in\{I,\mu_{N},\operatorname{C_{N}}\}\subseteq K(N)^{+}, a regular point pVgp\in V_{g}, a neighborhood U2U\subseteq\mathcal{H}_{2} of pp, and a function ϖ𝒪(U)\varpi\in{\mathcal{O}}(U) with an irreducible germ ϖp𝒪p\varpi_{p}\in\mathcal{O}_{p}, and with ϖ(τzzω)\varpi\left(\begin{smallmatrix}{\tau}&{z}\\ {z}&{\omega}\end{smallmatrix}\right) regular in ω\omega at pp, such that VgU={ZU:ϖ(Z)=0}V_{g}\cap U=\{Z\in U:\varpi(Z)=0\}.

Proof.

Each gK(N)+g\in K(N)^{+} is biholomorphic on 2\mathcal{H}_{2} so VV is an irreducible holomorphic subvariety of codimension one if and only if VgV_{g} is. The automorphy of ff shows that VV is supported in div(f)\operatorname{div}(f) if and only if VgV_{g} is, although we remark that VV and VgV_{g} might be equal. Since gg is biholomorphic, we have ordVf=ordgVfg1{\operatorname{ord}}_{V}f={\operatorname{ord}}_{g\langle V\rangle}f\circ g^{-1}. The automorphy of ff gives us (fg1)(Ω)=j(g1,Ω)kf(Ω)(f\circ g^{-1})(\Omega)=j(g^{-1},\Omega)^{k}f(\Omega) so that ff and fg1f\circ g^{-1} differ by a multiplicative unit and ordgVfg1=ordVgf{\operatorname{ord}}_{g\langle V\rangle}f\circ g^{-1}={\operatorname{ord}}_{V_{g}}f.

Every regular point pVp\in V has a local defining equation for VV given by VU={ZU:ϖ(Z)=0}V\cap U=\{Z\in U:\varpi(Z)=0\} for some neighborhood UU of pp, and some uniformizer ϖ𝒪(U)\varpi\in{\mathcal{O}}(U) with an irreducible germ ϖp\varpi_{p} represented by (ϖ,U)(\varpi,U). We may select UU so that every point of UVU\cap V is also a regular point of VV because VV is a complex manifold at pp. By Cartan’s Theorem, we may also assume that a single function element (ϖ,U)(\varpi,U) satisfies idVq=𝒪qϖq\operatorname{id}V_{q}=\mathcal{O}_{q}\varpi_{q} for all qUVq\in U\cap V. For the proof of the second part, we break the discussion into two mutually exclusive cases. Case II. Some regular point pVp\in V has ϖ(τzzω)\varpi\left(\begin{smallmatrix}{\tau}&{z}\\ {z}&{\omega}\end{smallmatrix}\right) regular in ω\omega or τ\tau at pp. If ϖ(τozozoω)0\varpi\left(\begin{smallmatrix}{\tau_{o}}&{z_{o}}\\ {z_{o}}&{\omega}\end{smallmatrix}\right)\not\equiv 0 in ω\omega for some regular p=(τozozoωo)Vp=\left(\begin{smallmatrix}{\tau_{o}}&{z_{o}}\\ {z_{o}}&{\omega_{o}}\end{smallmatrix}\right)\in V then we take g=Ig=I to satisfy our conclusion. Whereas if ϖ(τzozoωo)0\varpi\left(\begin{smallmatrix}{\tau}&{z_{o}}\\ {z_{o}}&{\omega_{o}}\end{smallmatrix}\right)\not\equiv 0 in τ\tau we take g=μNg=\mu_{N} and consider q=(τ1z1z1ω1)=μNp=(ωo/NzozoNτo)Vgq=\left(\begin{smallmatrix}{\tau_{1}}&{z_{1}}\\ {z_{1}}&{\omega_{1}}\end{smallmatrix}\right)=\mu_{N}\langle p\rangle=\left(\begin{smallmatrix}{\omega_{o}/N}&{-z_{o}}\\ {-z_{o}}&{N\tau_{o}}\end{smallmatrix}\right)\in V_{g}. The point qgUq\in g\langle U\rangle is a regular point of VgV_{g} because pp is a regular point of VV and gg is biholomorphic. The local uniformizer at qVgq\in V_{g} may be represented by (ϖg1,gU)(\varpi\circ g^{-1},g\langle U\rangle). In this case

(ϖg1)(τ1z1z1ω)=ϖ(ω/Nz1z1Nτ1)=ϖ(ω/Nzozoωo)(\varpi\circ g^{-1})\left(\begin{smallmatrix}{\tau_{1}}&{z_{1}}\\ {z_{1}}&{\omega}\end{smallmatrix}\right)=\varpi\left(\begin{smallmatrix}{\omega/N}&{-z_{1}}\\ {-z_{1}}&{N\tau_{1}}\end{smallmatrix}\right)=\varpi\left(\begin{smallmatrix}{\omega/N}&{z_{o}}\\ {z_{o}}&{\omega_{o}}\end{smallmatrix}\right)

is regular in ω\omega at qq. From VU={ZU:ϖ(Z)=0}V\cap U=\{Z\in U:\varpi(Z)=0\} we deduce VggU={ZgU:(ϖg1)(Z)=0}V_{g}\cap g\langle U\rangle=\{Z\in g\langle U\rangle:(\varpi\circ g^{-1})(Z)=0\} so that qq and gUg\langle U\rangle play the role of pp and UU in the statement of the lemma. Case II. Every regular point pVp\in V has ϖ(τzzω)\varpi\left(\begin{smallmatrix}{\tau}&{z}\\ {z}&{\omega}\end{smallmatrix}\right) not regular in ω\omega and not regular in τ\tau at pp. Consider any regular point po=(τozozoωo)Vp_{o}=\left(\begin{smallmatrix}{\tau_{o}}&{z_{o}}\\ {z_{o}}&{\omega_{o}}\end{smallmatrix}\right)\in V. Take a polydisk Δ=Δ11×Δ12×Δ22U\Delta=\Delta_{11}\times\Delta_{12}\times\Delta_{22}\subseteq U about pop_{o} and use the function element (ϖ,Δ)(\varpi,\Delta). We will ultimately show that Case II is very special and that VΔ={ZΔ:zzo=0}V\cap\Delta=\{Z\in\Delta:z-z_{o}=0\}. We claim that ϖ(τzozoω)=0\varpi\left(\begin{smallmatrix}{\tau}&{z_{o}}\\ {z_{o}}&{\omega}\end{smallmatrix}\right)=0 for all (τzozoω)Δ\left(\begin{smallmatrix}{\tau}&{z_{o}}\\ {z_{o}}&{\omega}\end{smallmatrix}\right)\in\Delta. If not there is some ro=(τ1zozoω1)Δr_{o}=\left(\begin{smallmatrix}{\tau_{1}}&{z_{o}}\\ {z_{o}}&{\omega_{1}}\end{smallmatrix}\right)\in\Delta with ϖ(ro)0\varpi(r_{o})\neq 0. Since ϖ\varpi is not regular in ω\omega at pop_{o} we have ϖ(τozozoω)0\varpi\left(\begin{smallmatrix}{\tau_{o}}&{z_{o}}\\ {z_{o}}&{\omega}\end{smallmatrix}\right)\equiv 0 for ω\omega in some neighborhood of ωo\omega_{o}, hence necessarily for all ωΔ22\omega\in\Delta_{22}. Therefore ϖ(τozozoω1)=0\varpi\left(\begin{smallmatrix}{\tau_{o}}&{z_{o}}\\ {z_{o}}&{\omega_{1}}\end{smallmatrix}\right)=0. The setting of Case II applies equally well to r=(τozozoω1)VΔr=\left(\begin{smallmatrix}{\tau_{o}}&{z_{o}}\\ {z_{o}}&{\omega_{1}}\end{smallmatrix}\right)\in V\cap\Delta, which is still a regular point of VV, so that ϖ\varpi is not regular in τ\tau at rr, and ϖ(τzozoω1)0\varpi\left(\begin{smallmatrix}{\tau}&{z_{o}}\\ {z_{o}}&{\omega_{1}}\end{smallmatrix}\right)\equiv 0 for τ\tau in a neighborhood of τo\tau_{o}, and hence for all τΔ11\tau\in\Delta_{11}. Thus ϖ(ro)=ϖ(τ1zozoω1)=0\varpi(r_{o})=\varpi\left(\begin{smallmatrix}{\tau_{1}}&{z_{o}}\\ {z_{o}}&{\omega_{1}}\end{smallmatrix}\right)=0, contrary to our supposition. Thus ϖ(τzozoω)=0\varpi\left(\begin{smallmatrix}{\tau}&{z_{o}}\\ {z_{o}}&{\omega}\end{smallmatrix}\right)=0 for all (τzozoω)Δ\left(\begin{smallmatrix}{\tau}&{z_{o}}\\ {z_{o}}&{\omega}\end{smallmatrix}\right)\in\Delta. Therefore ϖ\varpi vanishes on {ZΔ:zzo=0}\{Z\in\Delta:z-z_{o}=0\} and so the irreducible zzoz-z_{o} divides ϖ\varpi in Δ\Delta; however, the germ ϖpo\varpi_{p_{o}} of ϖ\varpi at pop_{o} is irreducible and so ϖ\varpi and zzoz-z_{o} differ by a multiplicative unit in Δ\Delta, after possibly shrinking Δ\Delta further. Therefore in Case II we have the possible but special circumstance VΔ={ZΔ:zzo=0}V\cap\Delta=\{Z\in\Delta:z-z_{o}=0\}. Without loss of generality we may adjust the selection of ϖ\varpi by a unit and assume that ϖ(τzzω)=zzo\varpi\left(\begin{smallmatrix}{\tau}&{z}\\ {z}&{\omega}\end{smallmatrix}\right)=z-z_{o} in Δ\Delta. Now set g=CNg=\operatorname{C_{N}} and consider q=(τ1z1z1ω1)=CNpo=(τo+2Nzo+N2ωozo+Nωozo+Nωoωo)Vgq=\left(\begin{smallmatrix}{\tau_{1}}&{z_{1}}\\ {z_{1}}&{\omega_{1}}\end{smallmatrix}\right)=\operatorname{C_{N}}\langle p_{o}\rangle=\left(\begin{smallmatrix}{\tau_{o}+2Nz_{o}+N^{2}\omega_{o}}&{z_{o}+N\omega_{o}}\\ {z_{o}+N\omega_{o}}&{\omega_{o}}\end{smallmatrix}\right)\in V_{g}. The local uniformizer at qVgq\in V_{g} may be represented by (ϖg1,gΔ)(\varpi\circ g^{-1},g\langle\Delta\rangle). For this choice we have

(ϖg1)(τ1z1z1ω)\displaystyle(\varpi\circ g^{-1})\left(\begin{smallmatrix}{\tau_{1}}&{z_{1}}\\ {z_{1}}&{\omega}\end{smallmatrix}\right) =ϖ(τ12Nz1+N2ωz1Nωz1Nωω)\displaystyle=\varpi\left(\begin{smallmatrix}{\tau_{1}-2Nz_{1}+N^{2}\omega}&{z_{1}-N\omega}\\ {z_{1}-N\omega}&{\omega}\end{smallmatrix}\right)
=ϖ(τo+N2(ωωo)zo+N(ωoω)zo+N(ωoω)ω)=N(ωoω),\displaystyle=\varpi\left(\begin{smallmatrix}{\tau_{o}+N^{2}(\omega-\omega_{o})}&{z_{o}+N(\omega_{o}-\omega)}\\ {z_{o}+N(\omega_{o}-\omega)}&{\omega}\end{smallmatrix}\right)=N(\omega_{o}-\omega),

which is regular in ω\omega of order 11 at qq. Thus the conclusion of the lemma holds for qq and gΔg\langle\Delta\rangle in the role of pp and UU. ∎

Theorem 8.2.

Let k,k0,Nk,k_{0},N\in{\mathbb{N}} and ϵ,ϵ0{±1}\epsilon,\epsilon_{0}\in\{\pm 1\}. Let ΓΓ0(N)\Gamma\subseteq\Gamma^{0}(N) be a subgroup of finite index. Let 𝔣𝕊(k,N,ϵ;Γ)\mathfrak{f}\in\mathbb{S}(k,N,\epsilon;\Gamma). Let f0Mk0(K(N))ϵ0f_{0}\in M_{k_{0}}\left(K(N)\right)^{\epsilon_{0}} be nontrivial and GMk0+k(K(N))ϵ0ϵG\in M_{k_{0}+k}\left(K(N)\right)^{\epsilon_{0}\epsilon}. Let V2V\subseteq\mathcal{H}_{2} be an irreducible component of div(f0)\operatorname{div}(f_{0}).

If FJ(G)=𝔣FJ(f0){\operatorname{FJ}}(G)=\mathfrak{f}\,{\operatorname{FJ}}(f_{0}) in 𝕊(k0+k,N,ϵ0ϵ;Γ)\mathbb{S}(k_{0}+k,N,\epsilon_{0}\epsilon;\Gamma) then ordVf0ordVG{\operatorname{ord}}_{V}f_{0}\leq{\operatorname{ord}}_{V}G. The meromorphic G/f0G/f_{0} is holomorphic and G/f0Mk(K(N))ϵG/f_{0}\in M_{k}\left(K(N)\right)^{\epsilon}.

Proof.

By Lemma 8.1 we can find a point pVgp\in V_{g}, for some gK(N)+g\in K(N)^{+}, where the local defining equation VgU={ZU:ϖ(Z)=0}V_{g}\cap U=\{Z\in U:\varpi(Z)=0\} has the following nice properties. The point pp is a regular point of VgV_{g}, the uniformizer ϖ\varpi is regular in ω\omega at pp, and the germ ϖp\varpi_{p} is prime in 𝒪p\mathcal{O}_{p}. By Lemma 8.1, in order to prove ordVf0ordVG{\operatorname{ord}}_{V}f_{0}\leq{\operatorname{ord}}_{V}G it suffices to prove ordVgf0ordVgG{\operatorname{ord}}_{V_{g}}f_{0}\leq{\operatorname{ord}}_{V_{g}}G, so we may rename VgV_{g} as VV without loss of generality.

In this paragraph we outline the remainder of the proof. We will find a neighborhood U1U_{1} of pp such that every point qU1Vq\in U_{1}\cap V has the same nice properties that pp does. We use this to twice move to nearby points. First we select p1U1Vp_{1}\in U_{1}\cap V so that ϖp1\varpi_{p_{1}} is the only irreducible factor of Gp1G_{p_{1}} that vanishes at p1p_{1}. Then we use the Weierstrauss preparation theorem on ϖp1\varpi_{p_{1}} to find a p2𝒲(Γ)VU1p_{2}\in\mathcal{W}(\Gamma)\cap V\cap U_{1}, and a holomorphic curve Ω2\Omega_{2} inside 𝒲(Γ)V\mathcal{W}(\Gamma)\cap V and passing through p2p_{2}, which guarantees that the formal series 𝔣\mathfrak{f} converges on Ω2(ω)\Omega_{2}(\omega). The conclusion about the orders of vanishing readily follows from the convergence of 𝔣\mathfrak{f} on the holomorphic curve.

The set of regular points (V)\mathcal{R}(V) is open in VV. Regularity in ω\omega at pp is a local condition in pp because regularity of order less than or equal to ν\nu at pp is implied by the nonvanishing of a partial derivative ωνϖ(p){\partial^{\nu}_{\omega}\varpi}(p). By Cartan’s Theorem, there is a neighborhood U′′UU^{\prime\prime}\subseteq U of pp such that ϖq\varpi_{q} generates idVq\operatorname{id}V_{q} at all points qU′′Vq\in U^{\prime\prime}\cap V. Accordingly there is a smaller neighborhood U1U_{1} of pp where every qU1Vq\in U_{1}\cap V has the nice properties q(V)q\in\mathcal{R}(V), ϖ\varpi regular in ω\omega at qq, and ϖq\varpi_{q} prime in 𝒪q\mathcal{O}_{q}.

In the local ring 𝒪p{\mathcal{O}}_{p}, factor (f0)p=(Δ1)pϖpν(f0)(f_{0})_{p}=(\Delta_{1})_{p}\varpi_{p}^{\nu(f_{0})} and Gp=(Δ2)pϖpν(G)G_{p}=(\Delta_{2})_{p}\varpi_{p}^{\nu(G)}, where (Δ1)p,(Δ2)p(\Delta_{1})_{p},(\Delta_{2})_{p} are finite products of irreducibles that are not associate to ϖp\varpi_{p}, and where ν(f0)=ordVf0\nu(f_{0})={\operatorname{ord}}_{V}f_{0} and ν(G)=ordVG\nu(G)={\operatorname{ord}}_{V}G by the direct way of computing the vanishing order on VV. We select a neighborhood N1U1N_{1}\subseteq U_{1} where the germs of the above factors all have representative function elements. If Δ2(q)=0\Delta_{2}(q)=0 for all qN1Vq\in N_{1}\cap V then (Δ2)p(\Delta_{2})_{p} vanishes on VpV_{p} and ϖp\varpi_{p} divides (Δ2)p(\Delta_{2})_{p} in 𝒪p\mathcal{O}_{p} because ϖp\varpi_{p} is prime, contradicting the fact that (Δ2)p(\Delta_{2})_{p} is a finite product of irreducibles that are not associate to ϖp\varpi_{p} in 𝒪p\mathcal{O}_{p}. Hence there is a point p1VN1p_{1}\in V\cap N_{1} such that Δ2(p1)0\Delta_{2}(p_{1})\neq 0. By shrinking N1N_{1}, we may assume that Δ2\Delta_{2} is nonzero on N1N_{1}.

We apply the Weierstrauss preparation theorem to ϖ\varpi, which is regular in ω\omega at p1=(τ1z1z1ω1)p_{1}=\left(\begin{smallmatrix}{\tau_{1}}&{z_{1}}\\ {z_{1}}&{\omega_{1}}\end{smallmatrix}\right). There exist neighborhoods U2U_{2} of (τ1,z1)(\tau_{1},z_{1}) and W2W_{2} of ω1\omega_{1} such that, N2=U2×W2N1N_{2}=U_{2}\times W_{2}\subseteq N_{1} and

{Z=(τzzω)N2:ϖ(Z)=0}=\displaystyle\{Z=\left(\begin{smallmatrix}{\tau}&{z}\\ {z}&{\omega}\end{smallmatrix}\right)\in N_{2}:\varpi(Z)=0\}=
{Z=(τzzω)N2:W(τ,z,ω):=j=0thj(τ,z)(ωω1)tj=0},\displaystyle\{Z=\left(\begin{smallmatrix}{\tau}&{z}\\ {z}&{\omega}\end{smallmatrix}\right)\in N_{2}:W(\tau,z,\omega):=\sum_{j=0}^{t}h_{j}(\tau,z)(\omega-\omega_{1})^{t-j}=0\},

for some tt\in{\mathbb{N}}, hj𝒪(U2)h_{j}\in{\mathcal{O}}(U_{2}) for j=0,,tj=0,\ldots,t, and h0=1h_{0}=1. The set 𝒲1(Γ)\mathcal{W}_{1}(\Gamma) is dense in U2U_{2} by Lemma 6.6, so by taking (τ2,z2)𝒲1(Γ)(\tau_{2},z_{2})\in\mathcal{W}_{1}(\Gamma) close enough to (τ1,z1)U2(\tau_{1},z_{1})\in U_{2} so that we may choose a root ω2\omega_{2} of W(τ2,z2,ω)W(\tau_{2},z_{2},\omega) close enough to ω1\omega_{1}, we have

p2=(τ2z2z2ω2)VN2𝒲(Γ).p_{2}=\left(\begin{smallmatrix}{\tau_{2}}&{z_{2}}\\ {z_{2}}&{\omega_{2}}\end{smallmatrix}\right)\in V\cap N_{2}\cap\mathcal{W}(\Gamma).

From N2N1N_{2}\subseteq N_{1} we inherit the representative function elements f0f_{0}, GG, Δ1\Delta_{1}, Δ2\Delta_{2}, and ϖ\varpi, so that f0=Δ1ϖν(f0)f_{0}=\Delta_{1}\varpi^{\nu(f_{0})} and G=Δ2ϖν(G)G=\Delta_{2}\varpi^{\nu(G)} in N2N_{2}, and Δ2\Delta_{2} is nonzero on N2N_{2}.

The formal series of GG is given by FJ(G)=𝔣FJ(f0){\operatorname{FJ}}(G)=\mathfrak{f}\,{\operatorname{FJ}}(f_{0}). Recalling that (τ2,z2)𝒲1(Γ)(\tau_{2},z_{2})\in\mathcal{W}_{1}(\Gamma), and using Lemma 7.2 to apply A(τ2,z2)A(\tau_{2},z_{2}), we have the equality A(τ2,z2)FJ(G)=(A(τ2,z2)𝔣)(A(τ2,z2)FJ(f0))A(\tau_{2},z_{2}){\operatorname{FJ}}(G)=(A(\tau_{2},z_{2})\mathfrak{f})(A(\tau_{2},z_{2}){\operatorname{FJ}}(f_{0})) of convergent power series on D2={ξ:|ξ|<r2}D_{2}=\{\xi\in{\mathbb{C}}:|\xi|<r_{2}\} for r2=e2πη2r_{2}=e^{-2\pi\eta_{2}} and η2=(Im(z2))2/Im(τ2)\eta_{2}=({\rm Im}(z_{2}))^{2}/{{\rm Im}(\tau_{2})}. Also set Ω2(ω)=(τ2z2z2ω)2\Omega_{2}(\omega)=\left(\begin{smallmatrix}{\tau_{2}}&{z_{2}}\\ {z_{2}}&{\omega}\end{smallmatrix}\right)\in\mathcal{H}_{2} for ωN(η2)\omega\in N_{\infty}(\eta_{2}). For ωN(η2)\omega\in N_{\infty}(\eta_{2}) we have e(ω)D2e(\omega)\in D_{2} and may apply the evaluation homomorphism [e(ω)]:𝒮(D2)[e(\omega)]:{\mathcal{S}}(D_{2})\to{\mathbb{C}} to obtain

(A(τ2,z2)FJ(G))[e(ω)]=(A(τ2,z2)𝔣)[e(ω)](A(τ2,z2)FJ(f0))[e(ω)].\left(A(\tau_{2},z_{2}){\operatorname{FJ}}(G)\right)[e(\omega)]=\left(A(\tau_{2},z_{2})\mathfrak{f}\right)[e(\omega)]\,\left(A(\tau_{2},z_{2}){\operatorname{FJ}}(f_{0})\right)[e(\omega)].

Making use of Lemma 7.2, this may be written on W2N(η2)W_{2}\cap N_{\infty}(\eta_{2}), which contains a neighborhood of ω2\omega_{2}, as

G(Ω2(ω))=H(τ2,z2,𝔣)(ω)f0(Ω2(ω)).G\left(\Omega_{2}(\omega)\right)=H(\tau_{2},z_{2},\mathfrak{f})(\omega)f_{0}\left(\Omega_{2}(\omega)\right).

By specializing the factorizations of GG and f0f_{0} in 𝒪(N2){\mathcal{O}}(N_{2}) to the holomorphic curve Ω2(ω)=(τ2z2z2ω)\Omega_{2}(\omega)=\left(\begin{smallmatrix}{\tau_{2}}&{z_{2}}\\ {z_{2}}&{\omega}\end{smallmatrix}\right) we have

Δ2(Ω2(ω))ϖ(Ω2(ω))ν(G)=H(τ2,z2,𝔣)(ω)Δ1(Ω2(ω))ϖ(Ω2(ω))ν(f0).\Delta_{2}\left(\Omega_{2}(\omega)\right)\varpi\left(\Omega_{2}(\omega)\right)^{\nu(G)}=H(\tau_{2},z_{2},\mathfrak{f})(\omega)\,\Delta_{1}\left(\Omega_{2}(\omega)\right)\varpi\left(\Omega_{2}(\omega)\right)^{\nu(f_{0})}.

Note that ϖ(Ω2(ω))\varpi\left(\Omega_{2}(\omega)\right) has at most tt zeros on W2W_{2} and is hence nontrivial. Of the two cases ν(G)ν(f0)\nu(G)\geq\nu(f_{0}), and ν(G)<ν(f0)\nu(G)<\nu(f_{0}), the first is our conclusion, so we will conclude the proof by showing that the second does not occur. In the second case, by cancelling powers of the nontrivial ϖ(Ω2(ω))\varpi\left(\Omega_{2}(\omega)\right), we have

Δ2(Ω2(ω))=H(τ2,z2,𝔣)(ω)Δ1(Ω2(ω))ϖ(Ω2(ω))ν(f0)ν(G).\Delta_{2}\left(\Omega_{2}(\omega)\right)=H(\tau_{2},z_{2},\mathfrak{f})(\omega)\,\Delta_{1}\left(\Omega_{2}(\omega)\right)\varpi\left(\Omega_{2}(\omega)\right)^{\nu(f_{0})-\nu(G)}.

We evaluate these at ω2W2N(η2)\omega_{2}\in W_{2}\cap N_{\infty}(\eta_{2}) to obtain

Δ2(p2)=H(τ2,z2,𝔣)(ω2)Δ1(p2)(0ν(f0)ν(G))=0,\Delta_{2}(p_{2})=H(\tau_{2},z_{2},\mathfrak{f})(\omega_{2})\,\Delta_{1}(p_{2})\left(0^{\,\nu(f_{0})-\nu(G)}\right)=0,

which contradicts Δ2(p2)0\Delta_{2}(p_{2})\neq 0. Thus we have ν(G)ν(f0)\nu(G)\geq\nu(f_{0}).

A meromorphic function with an effective divisor is holomorphic. Therefore G/f0G/f_{0} is holomorphic and continuous on 2\mathcal{H}_{2}. There is an open dense subset of 2\mathcal{H}_{2} where G/f0G/f_{0} transforms like an element of Mk(K(N))ϵM_{k}\left(K(N)\right)^{\epsilon}. By the continuity of G/f0G/f_{0} and of the factor of automorphy, G/f0G/f_{0} transforms like an element of Mk(K(N))ϵM_{k}\left(K(N)\right)^{\epsilon} on 2\mathcal{H}_{2} and hence G/f0Mk(K(N))ϵG/f_{0}\in M_{k}\left(K(N)\right)^{\epsilon}. ∎

9. Main Theorem.

Theorem 9.1.

Let UdU\subseteq{\mathbb{C}}^{d} be open. Let {fj}\{f_{j}\} be a locally bounded sequence of holomorphic functions on UU that converges on a dense subset of UU. Then the sequence {fj}\{f_{j}\} converges on UU and uniformly on compact subsets of UU.

Proof.

This is Exercise 44a of section 44 in Chapter 11 of [14].

The proof given here imitates the proof of Lemma IV.4.8 in [13] for the one-dimensional case. It suffices to show that {fj}\{f_{j}\} is uniformly Cauchy on compact subsets of Ω\Omega. Pick a compact KΩK\subseteq\Omega and an ϵ>0\epsilon>0. There is no loss of generality in assuming that KK is a closed ball. We will construct an MM\in{\mathbb{N}} from the given data {fj}\{f_{j}\}, KK, and ϵ\epsilon.

The family {fj}\{f_{j}\} is equicontinuous on KK because it is locally bounded, see pages 1111-1212 of [31]. Select δ>0\delta>0 to enforce this equicontinuity on KK for 13ϵ\frac{1}{3}\epsilon.

(13) z,wK,n,|zw|<δ|fn(z)fn(w)|<13ϵ.\forall z,w\in K,\,\forall n\in{\mathbb{N}},\,|z-w|<\delta\implies|f_{n}(z)-f_{n}(w)|<\tfrac{1}{3}\epsilon.

Cover KK with open balls inside Ω\Omega at each point of KK with radii that are less than 12δ\frac{1}{2}\delta. By the compactness of KK we have Ki=1B(zi,δi)ΩK\subseteq\cup_{i=1}^{\ell}B(z_{i},\delta_{i})\subseteq\Omega for some ziKz_{i}\in K, and 0<δi<12δ0<\delta_{i}<\frac{1}{2}\delta. Since SS is dense in KK, we can pick an siKSB(zi,δi)s_{i}\in K\cap S\cap B(z_{i},\delta_{i}); this uses that KK is a closed ball. Note that for every point zB(zi,δi)z\in B(z_{i},\delta_{i}) we have |zsi|<δ|z-s_{i}|<\delta and therefore |fn(z)fn(si)|<13ϵ|f_{n}(z)-f_{n}(s_{i})|<\frac{1}{3}\epsilon by (13). By hypothesis each sequence fj(si)f_{j}(s_{i}) converges and so is Cauchy. Select Mi>0M_{i}>0 so that

(14) m,n,m,n>Mi|fm(si)fn(si)|<13ϵ.\forall m,n\in{\mathbb{N}},\,m,n>M_{i}\implies|f_{m}(s_{i})-f_{n}(s_{i})|<\tfrac{1}{3}\epsilon.

Let M=max1iMiM=\max_{1\leq i\leq\ell}M_{i} be the promised natural number. We have

|fm(z)fn(z)||fm(z)fm(si)|+|fm(si)fn(si)|+|fn(si)fn(z)||f_{m}(z){-}f_{n}(z)|\leq|f_{m}(z)-f_{m}(s_{i})|+|f_{m}(s_{i})-f_{n}(s_{i})|+|f_{n}(s_{i})-f_{n}(z)|

for any m,nm,n\in{\mathbb{N}}, zKz\in K, and 1i1\leq i\leq\ell. If m,n>Mm,n>M, then m,n>Mim,n>M_{i} and so |fm(si)fn(si)|<13ϵ|f_{m}(s_{i})-f_{n}(s_{i})|<\frac{1}{3}\epsilon by (14). Choose ii so that zB(zi,δi)z\in B(z_{i},\delta_{i}); then |fj(z)fj(si)|<13ϵ|f_{j}(z)-f_{j}(s_{i})|<\frac{1}{3}\epsilon for all jj\in{\mathbb{N}} by (13). Therefore, for all m,nm,n\in{\mathbb{N}} with m,n>Mm,n>M, and all zKz\in K, we have |fm(z)fn(z)|<ϵ|f_{m}(z)-f_{n}(z)|<\epsilon, and the sequence {fj}\{f_{j}\} is uniformly Cauchy on KK. ∎

The essence of the proof of the main result, Theorem 9.6, lies in the argument for the following special case of Fricke plus cusp forms. The general case will be reduced to Theorem 9.2.

Theorem 9.2.

The map FJ:Sk(K(N)+)𝕊(k,N,+){\operatorname{FJ}}:S_{k}\left(K(N)^{+}\right)\to\mathbb{S}(k,N,+) is an isomorphism for N,kN,k\in{\mathbb{N}}.

Proof.

Since the map FJ{\operatorname{FJ}} is injective, the only issue is surjectivity. Take a nontrivial 𝔣𝕊(k,N,+)\mathfrak{f}\in\mathbb{S}(k,N,+). By Proposition 5.4, there is a subgroup Γ\Gamma of finite index in Γ0(N)\Gamma^{0}(N) such that 𝔣𝕄(k,N,+;Γ)\mathfrak{f}\in\mathbb{M}(k,N,+\,;\Gamma), which implies 𝔣𝕄(k,N,+;Γ)𝕊(k,N,+)=𝕊(k,N,+;Γ)\mathfrak{f}\in\mathbb{M}(k,N,+\,;\Gamma)\cap\mathbb{S}(k,N,+)=\mathbb{S}(k,N,+\,;\Gamma). Completely separately, by Proposition 4.7, 𝔣\mathfrak{f} satisfies a polynomial relation in 𝕄(k0+dk,N,+)\mathbb{M}(k_{0}+dk,N,+),

FJ(f0)𝔣d+FJ(f1)𝔣d1++FJ(fj)𝔣dj++FJ(fd)=0,{\operatorname{FJ}}(f_{0})\mathfrak{f}^{d}+{\operatorname{FJ}}(f_{1})\mathfrak{f}^{d-1}+\cdots+{\operatorname{FJ}}(f_{j})\mathfrak{f}^{d-j}+\cdots+{\operatorname{FJ}}(f_{d})=0,

for some dd\in{\mathbb{N}}, k00k_{0}\in{\mathbb{N}}_{0}, and some fjMk0+jk(K(N)+)f_{j}\in M_{k_{0}+jk}(K(N)^{+}) with f0f_{0} not identically zero. Since FJ(fj){\operatorname{FJ}}(f_{j}), 𝔣𝕄(N,+;Γ)\mathfrak{f}\in\mathbb{M}(N,+\,;\Gamma), this polynomial relation is actually in 𝕄(k0+dk,N,+;Γ)\mathbb{M}(k_{0}+dk,N,+\,;\Gamma). Since 𝕊(N,+;Γ)\mathbb{S}(N,+\,;\Gamma) is a graded ideal in 𝕄(N,+;Γ)\mathbb{M}(N,+\,;\Gamma), set

𝔤=FJ(f0)𝔣=m=1ϕmξNm𝕊(k+k0,N,+;Γ).\mathfrak{g}={\operatorname{FJ}}(f_{0})\mathfrak{f}=\sum_{m=1}^{\infty}\phi_{m}\xi^{Nm}\in\mathbb{S}(k+k_{0},N,+\,;\Gamma).

Then 𝔤\mathfrak{g} is integral over FJ(M(K(N)+)){\operatorname{FJ}}\left(M(K(N)^{+})\right) and satisfies a monic polynomial relation in 𝕄((k+k0)d,N,+;Γ)\mathbb{M}((k+k_{0})d,N,+\,;\Gamma) given by

𝔤d+FJ(g1)𝔤d1++FJ(gj)𝔤dj++FJ(gd)=0\mathfrak{g}^{d}+{\operatorname{FJ}}(g_{1})\mathfrak{g}^{d-1}+\cdots+{\operatorname{FJ}}(g_{j})\mathfrak{g}^{d-j}+\cdots+{\operatorname{FJ}}(g_{d})=0

with gj=f0j1fjM(k+k0)j(K(N)+)g_{j}=f_{0}^{j-1}f_{j}\in M_{(k+k_{0})j}\left(K(N)^{+}\right) for j=1,,dj=1,\ldots,d.

By Theorem 7.4, the sequence of partial sums m=1Mϕm(τ,z)e(Nmω)\sum_{m=1}^{M}\phi_{m}(\tau,z)e\left(Nm\omega\right) is locally bounded for (τzzω)2\left(\begin{smallmatrix}{\tau}&{z}\\ {z}&{\omega}\end{smallmatrix}\right)\in\mathcal{H}_{2}. By Corollary 6.9, we also know that the formal series 𝔤\mathfrak{g} converges on the dense subset 𝒲(Γ)\mathcal{W}(\Gamma) of 2\mathcal{H}_{2}, noting 𝔤𝕊(k+k0,N,+;Γ)𝕊(k+k0,N;Γ)\mathfrak{g}\in\mathbb{S}(k+k_{0},N,+\,;\Gamma)\subseteq\mathbb{S}(k+k_{0},N;\Gamma). Therefore, by Theorem 9.1, the sequence of partial sums converges uniformly on compact sets of 2\mathcal{H}_{2}. The limiting function G𝒪(2)G\in{\mathcal{O}}(\mathcal{H}_{2}) is given by G(Ω)=m=1ϕm(τ,z)e(Nmω)G(\Omega)=\sum_{m=1}^{\infty}\phi_{m}(\tau,z)e\left(Nm\omega\right) and satisfies G|k+k0g=GG|_{k+k_{0}}g=G for all gP2,1()g\in\operatorname{P}_{2,1}({\mathbb{Z}}) because each term ϕm(τ,z)e(Nmω)\phi_{m}(\tau,z)e\left(Nm\omega\right) is so invariant. The holomorphic function GG is hence periodic with respect to the translation lattice {(Is0I)P2,1():sM2×2sym()}\{\left(\begin{smallmatrix}{I}&{s}\\ {0}&{I}\end{smallmatrix}\right)\in\operatorname{P}_{2,1}({\mathbb{Z}}):s\in M_{2\times 2}^{\text{\rm sym}}({\mathbb{Z}})\} and has a Fourier series that converges absolutley and uniformly on compact sets of 2\mathcal{H}_{2}. This absolute convergence of the Fourier series of GG is the crux of the matter because it will allow us to rearrange the order of summation and thereby deduce the crucial invariance property G|k+k0μN=GG|_{k+k_{0}}\mu_{N}=G from the involution condition c(n,r;ϕm)=(1)kϵc(m,r;ϕn)c(n,r;\phi_{m})=(-1)^{k}\epsilon\,c(m,r;\phi_{n}) on 𝔤\mathfrak{g}. Since the convergence of the sequence of partial sums is uniform on compact subsets, the Fourier coefficients of GG are supported on 𝒳¯(N)M2×2sym(){\bar{\mathcal{X}}}(N)\subseteq M_{2\times 2}^{\text{\rm sym}}({\mathbb{Q}}) and, for t=(nr/2r/2Nm)𝒳¯(N)t=\left(\begin{smallmatrix}{n}&{r/2}\\ {r/2}&{Nm}\end{smallmatrix}\right)\in{\bar{\mathcal{X}}}(N), satisfy a(t;G)=c(n,r;ϕm)a(t;G)=c(n,r;\phi_{m}). For 𝔤\mathfrak{g} in this proof we have ϵ=+1\epsilon=+1. Recall that tFN1tFNt\mapsto F_{N}^{-1}tF_{N}^{*} is an involution of 𝒳¯(N){\bar{\mathcal{X}}}(N) satisfying FN1(nr/2r/2Nm)FN=(mr/2r/2Nn)F_{N}^{-1}\left(\begin{smallmatrix}{n}&{r/2}\\ {r/2}&{Nm}\end{smallmatrix}\right)F_{N}^{*}=\left(\begin{smallmatrix}{m}&{-r/2}\\ {-r/2}&{Nn}\end{smallmatrix}\right), and that absolutely convergent summations over t𝒳¯(N)t\in{\bar{\mathcal{X}}}(N) may instead be rearranged to be taken over FN1tFNF_{N}^{-1}tF_{N}^{*}.

(G|μN)(Ω)=G(FNΩFN1)=t𝒳¯(N)a(t;G)e(FNΩFN1,t)\displaystyle\left(G|\mu_{N}\right)(\Omega)=G(F_{N}^{*}\Omega F_{N}^{-1})=\sum_{t\in{\bar{\mathcal{X}}}(N)}a(t;G)e\left(\langle F_{N}^{*}\Omega F_{N}^{-1},t\rangle\right)
=\displaystyle= ta(t;G)e(Ω,FN1tFN)=ta(FN1tFN;G)e(Ω,t)\displaystyle\sum_{t}a(t;G)e\left(\langle\Omega,F_{N}^{-1}tF_{N}^{*}\rangle\right)=\sum_{t}a(F_{N}^{-1}tF_{N}^{*};G)e\left(\langle\Omega,t\rangle\right)
=\displaystyle= t=(nr/2r/2Nm)𝒳¯(N)c(m,r;ϕn)e(Ω,t)=tc(n,r;ϕm)e(Ω,t)\displaystyle\sum_{t={\tiny\left(\begin{smallmatrix}{n}&{r/2}\\ {r/2}&{Nm}\end{smallmatrix}\right)}\in{\bar{\mathcal{X}}}(N)}c(m,-r;\phi_{n})e\left(\langle\Omega,t\rangle\right)=\sum_{t}c(n,r;\phi_{m})e\left(\langle\Omega,t\rangle\right)
=\displaystyle= ta(t;G)e(Ω,t)=G(Ω).\displaystyle\sum_{t}a(t;G)e\left(\langle\Omega,t\rangle\right)=G(\Omega).

Combining these automorphic properties of GG we obtain G|σ=GG|\sigma=G for all σP2,1(),μN=K(N),μN=K(N)+\sigma\in\langle\operatorname{P}_{2,1}({\mathbb{Z}}),\mu_{N}\rangle=\langle K(N),\mu_{N}\rangle=K(N)^{+}, see Gritsenko [17, 18]. Therefore GMk+k0(K(N)+)G\in M_{k+k_{0}}\left(K(N)^{+}\right) and FJ(G)=𝔤𝕊(k+k0,N,+;Γ){\operatorname{FJ}}(G)=\mathfrak{g}\in\mathbb{S}(k+k_{0},N,+\,;\Gamma). Together these imply that GSk+k0(K(N)+)G\in S_{k+k_{0}}\left(K(N)^{+}\right) by Reefschläger’s decomposition [34], as in Lemma 3.1.

We have FJ(G)=𝔤=FJ(f0)𝔣{\operatorname{FJ}}(G)=\mathfrak{g}={\operatorname{FJ}}(f_{0})\mathfrak{f} for the formal series 𝔣𝕊(k,N,+;Γ)\mathfrak{f}\in\mathbb{S}(k,N,+\,;\Gamma), and the paramodular forms f0Mk0(K(N)+)f_{0}\in M_{k_{0}}\left(K(N)^{+}\right), GSk+k0(K(N)+)G\in S_{k+k_{0}}\left(K(N)^{+}\right). By Theorem 8.2, f:=G/f0Mk(K(N)+)f{:=}\,{G}/{f_{0}}\in M_{k}\left(K(N)^{+}\right). From f0f=Gf_{0}f=G we see that FJ(f0)FJ(f)=FJ(G){\operatorname{FJ}}(f_{0}){\operatorname{FJ}}(f)={\operatorname{FJ}}(G). Combining this with FJ(f0)𝔣=FJ(G){\operatorname{FJ}}(f_{0})\mathfrak{f}={\operatorname{FJ}}(G), we have FJ(f0)(FJ(f)𝔣)=0{\operatorname{FJ}}(f_{0})\left({\operatorname{FJ}}(f)-\mathfrak{f}\right)=0, which implies FJ(f)=𝔣{\operatorname{FJ}}(f)=\mathfrak{f} since 𝕄(N)\mathbb{M}(N) is an integral domain and f0f_{0} is nontrivial. By Reefschläger’s result, FJ(f)=𝔣𝕊(k,N,+;Γ){\operatorname{FJ}}(f)=\mathfrak{f}\in\mathbb{S}(k,N,+\,;\Gamma) and fMk(K(N)+)f\in M_{k}\left(K(N)^{+}\right) imply fSk(K(N)+)f\in S_{k}\left(K(N)^{+}\right), which proves that FJ{\operatorname{FJ}} is surjective. ∎

Lemma 9.3.

There is a nontrivial cusp form χNS11(K(N))\chi_{N}\in S_{11}\left(K(N)\right)^{-} for N>1N>1.

Proof.

There are a number of possible approaches to proving the existence of nontrivial minus cusp forms. We might use an asymptotic trace formula or the oldform theory of Roberts and Schmidt [35, 36]. Perhaps the briefest is to use the injectivity of the Gritsenko lift [18], Grit:Jk,mcuspSk(K(m))ϵ\operatorname{Grit}:J_{k,m}^{\text{\rm cusp}}\to S_{k}\left(K(m)\right)^{\epsilon} for ϵ=(1)k\epsilon=(-1)^{k}. An estimate from the dimension formula [11] is dimJ11,mcusp74floor(m36)+712\dim J_{11,m}^{\text{\rm cusp}}\geq\frac{7}{4}\operatorname{floor}\left(\frac{m-3}{6}\right)+\frac{7}{12} for mm\in{\mathbb{N}}. Considering dimJ11,2cusp=1\dim J_{11,2}^{\text{\rm cusp}}=1, we see dimJ11,mcusp1\dim J_{11,m}^{\text{\rm cusp}}\geq 1 for m2m\geq 2. Thus there is a nontrivial Gritsenko lift χN\chi_{N} in S11(K(N))S_{11}\left(K(N)\right)^{-} for N>1N>1. ∎

Theorem 9.4.

The map FJ:Sk(K(N))ϵ𝕊(k,N,ϵ){\operatorname{FJ}}:S_{k}\left(K(N)\right)^{\epsilon}\to\mathbb{S}(k,N,\epsilon) is an isomorphism for N,kN,k\in{\mathbb{N}}, and ϵ{±1}\epsilon\in\{\pm 1\}.

Proof.

The case ϵ=+1\epsilon=+1 is Theorem 9.2, so assume ϵ=1\epsilon=-1 and N>1N>1. Take 𝔣𝕊(k,N,)\mathfrak{f}\in\mathbb{S}(k,N,-). There is a nontrivial cusp form χNS11(K(N))\chi_{N}\in S_{11}\left(K(N)\right)^{-} by Lemma 9.3. Multiply 𝔣\mathfrak{f} by FJ(χN){\operatorname{FJ}}(\chi_{N}) to obtain a new formal series 𝔤=FJ(χN)𝔣𝕊(k+11,N,+)\mathfrak{g}={\operatorname{FJ}}(\chi_{N})\mathfrak{f}\in\mathbb{S}(k+11,N,+). By Theorem 9.2, there exists a GSk+11(K(N))+G\in S_{k+11}\left(K(N)\right)^{+} such that FJ(G)=𝔤=FJ(χN)𝔣{\operatorname{FJ}}(G)=\mathfrak{g}={\operatorname{FJ}}(\chi_{N})\mathfrak{f}. Both FJ(G){\operatorname{FJ}}(G) and FJ(χN){\operatorname{FJ}}(\chi_{N}) have Γ0(N)\Gamma^{0}(N)-symmetries and we will show that 𝔣\mathfrak{f} does too. Apply the ring homomorphism AFS:𝕄(N)𝒳¯(N){\operatorname{AFS}}:\mathbb{M}(N)\to{\mathbb{C}}^{{\bar{\mathcal{X}}}(N)} of section 5 to FJ(G)=FJ(χN)𝔣{\operatorname{FJ}}(G)={\operatorname{FJ}}(\chi_{N})\mathfrak{f} to obtain FS(G)=FS(χN)AFS(𝔣){\operatorname{FS}}(G)={\operatorname{FS}}(\chi_{N}){\operatorname{AFS}}(\mathfrak{f}). Now, for σΓ0(N)\sigma\in\Gamma^{0}(N), apply the automorphism j(σ)\operatorname{j}(\sigma) of Lemma 5.3, which fixes FS(G){\operatorname{FS}}(G) and FS(χN){\operatorname{FS}}(\chi_{N}) to obtain 0=FS(χN)(AFS(𝔣)j(σ)AFS(𝔣))0={\operatorname{FS}}(\chi_{N})\left({\operatorname{AFS}}(\mathfrak{f})-\operatorname{j}(\sigma){\operatorname{AFS}}(\mathfrak{f})\right). Since 𝒳¯(N){\mathbb{C}}^{{\bar{\mathcal{X}}}(N)} is an integral domain and FS(χN){\operatorname{FS}}(\chi_{N}) is nontrivial, we have j(σ)AFS(𝔣)=AFS(𝔣)\operatorname{j}(\sigma){\operatorname{AFS}}(\mathfrak{f})={\operatorname{AFS}}(\mathfrak{f}) for all σΓ0(N)\sigma\in\Gamma^{0}(N). By Lemma 5.3, we conclude that 𝔣𝕊(k,N,;Γ0(N))\mathfrak{f}\in\mathbb{S}(k,N,-\,;\Gamma^{0}(N)). By Theorem 8.2 with Γ=Γ0(N)\Gamma=\Gamma^{0}(N), the equation FJ(G)=FJ(χN)𝔣{\operatorname{FJ}}(G)={\operatorname{FJ}}(\chi_{N})\mathfrak{f} implies that f:=G/χNf{:=}\,G/\chi_{N} is holomorphic, and, consequently, that fMk(K(N))f\in M_{k}\left(K(N)\right)^{-}. From FJ(G)=FJ(χN)𝔣{\operatorname{FJ}}(G)={\operatorname{FJ}}(\chi_{N})\mathfrak{f} and FJ(G)=FJ(χN)FJ(f){\operatorname{FJ}}(G)={\operatorname{FJ}}(\chi_{N}){\operatorname{FJ}}(f) we obtain 0=FJ(χN)(𝔣FJ(f))0={\operatorname{FJ}}(\chi_{N})(\mathfrak{f}-{\operatorname{FJ}}(f)). Since 𝕄(N)\mathbb{M}(N) is an integral domain and FJ(χN){\operatorname{FJ}}(\chi_{N}) is nontrivial, we have FJ(f)=𝔣{\operatorname{FJ}}(f)=\mathfrak{f}. By Reefschläger’s result and 𝔣𝕊(k,N,;Γ)\mathfrak{f}\in\mathbb{S}(k,N,-\,;\Gamma) we have fSk(K(N))f\in S_{k}\left(K(N)\right)^{-}. Thus FJ{\operatorname{FJ}} is surjective. ∎

Lemma 9.5.

Let NN\in{\mathbb{N}} and I=[K(N)+:K(N)+K(1)]I=[K(N)^{+}:K(N)^{+}\cap K(1)]. Let p2p\in\mathcal{H}_{2}. There is a cusp form χpS60I(K(N)+)\chi_{p}\in S_{60I}\left(K(N)^{+}\right) such that χp(p)0\chi_{p}(p)\neq 0.

Proof.

The divisor of χ10S10(K(1))\chi_{10}\in S_{10}(K(1)) is the reducible locus RR,

R=σK(1){σ(τ00ω)2:τ,ω1}.R=\cup_{\sigma\in K(1)}\{\sigma\langle\left(\begin{smallmatrix}{\tau}&{0}\\ {0}&{\omega}\end{smallmatrix}\right)\rangle\in\mathcal{H}_{2}:\tau,\omega\in\mathcal{H}_{1}\}.

For ψ12S12(K(1))\psi_{12}\in S_{12}(K(1)), we have ψ12(τ00ω)=Δ(τ)Δ(ω)\psi_{12}\left(\begin{smallmatrix}{\tau}&{0}\\ {0}&{\omega}\end{smallmatrix}\right)=\Delta(\tau)\Delta(\omega) for the nonvanishing ΔS12(SL(2,))\Delta\in S_{12}\left({\operatorname{SL}}(2,{\mathbb{Z}})\right). Thus χ10\chi_{10} and ψ12\psi_{12} have no common zeros on 2\mathcal{H}_{2}. Pick II coset representatives σ\sigma for K(N)+=σ(K(N)+K(1))σK(N)^{+}=\amalg_{\sigma}(K(N)^{+}\cap K(1))\sigma. For any (α,β)2{(0,0)}(\alpha,\beta)\in{\mathbb{C}}^{2}\setminus\{(0,0)\}, consider the nontrivial norm

Hα,β=[σ](K(N)+K(1))\K(N)+(αχ106+βψ125)|60σS60I(K(N)+).H^{\alpha,\beta}=\prod_{[\sigma]\in(K(N)^{+}\cap K(1))\backslash K(N)^{+}}\left(\alpha\chi_{10}^{6}+\beta\psi_{12}^{5}\right)|_{60}\sigma\in S_{60I}\left(K(N)^{+}\right).

We have Hα,β(p)=(σj(σ,p)60)σ(αχ106(σp)+βψ125(σp))H^{\alpha,\beta}(p)=\left(\prod_{\sigma}j(\sigma,p)^{-60}\right)\prod_{\sigma}\left(\alpha\chi_{10}^{6}(\sigma\langle p\rangle)+\beta\psi_{12}^{5}(\sigma\langle p\rangle)\right). Pick any (αo,βo)2(\alpha_{o},\beta_{o})\in{\mathbb{C}}^{2} that is not on any of the finitely many codimension one complex lines through the origin {(z1,z2)2:z1χ106(σp)+z2ψ125(σp)=0}\{(z_{1},z_{2})\in{\mathbb{C}}^{2}:z_{1}\chi_{10}^{6}(\sigma\langle p\rangle)+z_{2}\psi_{12}^{5}(\sigma\langle p\rangle)=0\}. The cusp form χp=Hαo,βo\chi_{p}=H^{\alpha_{o},\beta_{o}} is as claimed. ∎

Theorem 9.6 (Main Theorem).

Let NN\in{\mathbb{N}}, k0k\in{\mathbb{N}}_{0}, and ϵ{±1}\epsilon\in\{\pm 1\}. The map FJ:Mk(K(N))ϵ𝕄(k,N,ϵ){\operatorname{FJ}}:M_{k}\left(K(N)\right)^{\epsilon}\to\mathbb{M}(k,N,\epsilon) is an isomorphism. The map FJ:Sk(K(N))ϵ𝕊(k,N,ϵ){\operatorname{FJ}}:S_{k}\left(K(N)\right)^{\epsilon}\to\mathbb{S}(k,N,\epsilon) is also an isomorphism.

Proof.

The case k=0k=0 is easy so assume kk\in{\mathbb{N}}. Take 𝔣𝕄(k,N,ϵ)\mathfrak{f}\in\mathbb{M}(k,N,\epsilon). For any point p2p\in\mathcal{H}_{2} there is a cusp form χpS60I(K(N)+)\chi_{p}\in S_{60I}\left(K(N)^{+}\right) such that χp(p)0\chi_{p}(p)\neq 0, as in Lemma 9.5. Define a corrsponding product of formal series 𝔤p=FJ(χp)𝔣𝕊(k+60I,N,ϵ)\mathfrak{g}_{p}={\operatorname{FJ}}(\chi_{p})\mathfrak{f}\in\mathbb{S}(k+60I,N,\epsilon). By Theorem 9.4, there exists a GpSk+60I(K(N))ϵG_{p}\in S_{k+60I}(K(N))^{\epsilon} such that FJ(Gp)=𝔤p=FJ(χp)𝔣{\operatorname{FJ}}(G_{p})=\mathfrak{g}_{p}={\operatorname{FJ}}(\chi_{p})\mathfrak{f}. Note that Sk+60I(K(N))ϵ=Sk+60I(K(N)+,λϵ)S_{k+60I}(K(N))^{\epsilon}=S_{k+60I}(K(N)^{+},\lambda_{\epsilon}) where λϵ:K(N)+{±1}\lambda_{\epsilon}:K(N)^{+}\to\{\pm 1\} is the character defined by λϵ(μN)=ϵ\lambda_{\epsilon}(\mu_{N})=\epsilon and λϵ=1\lambda_{\epsilon}=1 on K(N)K(N).

In the same way, for any point q2q\in\mathcal{H}_{2} distinct from pp we also have FJ(Gq)=FJ(χq)𝔣{\operatorname{FJ}}(G_{q})={\operatorname{FJ}}(\chi_{q})\mathfrak{f}. Since 𝕄(N)\mathbb{M}(N) is an integral domain we necessarily have FJ(Gp)FJ(χq)FJ(Gq)FJ(χp)=0{\operatorname{FJ}}(G_{p}){\operatorname{FJ}}(\chi_{q})-{\operatorname{FJ}}(G_{q}){\operatorname{FJ}}(\chi_{p})=0, or Gpχq=GqχpG_{p}\chi_{q}=G_{q}\chi_{p}. Therefore we have Gpχp=Gqχq\frac{G_{p}}{\chi_{p}}=\frac{G_{q}}{\chi_{q}} whenever both quotients are defined. We may define f𝒪(2)f\in{\mathcal{O}}(\mathcal{H}_{2}) by setting f=Gpχpf=\frac{G_{p}}{\chi_{p}} in neighborhoods where the denominator does not vanish, and by noting this is independent of the choice of pp. Therefore we have χpf=Gp\chi_{p}f=G_{p}, χp|σf|σ=Gp|σ\chi_{p}|\sigma\,f|\sigma=G_{p}|\sigma, and χpf|σ=λϵ(σ)Gp\chi_{p}f|\sigma=\lambda_{\epsilon}(\sigma)G_{p}, for any point pp and any σK(N)+\sigma\in K(N)^{+}. Thus χp(f|σλϵ(σ)f)=0\chi_{p}\left(f|\sigma-\lambda_{\epsilon}(\sigma)f\right)=0. Since M(K(N))M(K(N)) is an integral domain and χp\chi_{p} is nontrivial, we have f|σ=λϵ(σ)ff|\sigma=\lambda_{\epsilon}(\sigma)f and fMk(K(N)+,λϵ)=Mk(K(N))ϵf\in M_{k}(K(N)^{+},\lambda_{\epsilon})=M_{k}(K(N))^{\epsilon}. It remains to show FJ(f)=𝔣{\operatorname{FJ}}(f)=\mathfrak{f}. This follows from FJ(Gp)=FJ(χp)𝔣{\operatorname{FJ}}(G_{p})={\operatorname{FJ}}(\chi_{p})\mathfrak{f}, FJ(Gp)=FJ(χp)FJ(f){\operatorname{FJ}}(G_{p})={\operatorname{FJ}}(\chi_{p}){\operatorname{FJ}}(f), and the fact that 𝕄(N)\mathbb{M}(N) is an integral domain and χp\chi_{p} is nontrivial. The result for cusp forms follows from the general result because, by Lemma 3.2, the inverse image of 𝕊(k,N,ϵ)\mathbb{S}(k,N,\epsilon) under FJ{\operatorname{FJ}} is Sk(K(N))ϵS_{k}\left(K(N)\right)^{\epsilon}. ∎

To help prove Corollary 1.2 we use a lemma from linear algebra.

Lemma 9.7.

Let UU, VV, and WW be finite dimensional vector spaces. Let L:UWL^{\prime}:U\to W and L′′:VWL^{\prime\prime}:V\to W be linear maps. Define a subspace XU×VX\subseteq U\times V by X={(u,v)U×V:Lu=L′′v}X=\{(u,v)\in U\times V:L^{\prime}u=L^{\prime\prime}v\}. Define the projection πU:XU\pi_{U}:X\to U by πU(u,v)=u\pi_{U}(u,v)=u. If L′′L^{\prime\prime} is injective then πU\pi_{U} is injective, and conversely.

Proof.

We have ker(πU)={(0,v)U×V:0=L′′v}={0}×ker(L′′)\ker(\pi_{U})=\{(0,v)\in U\times V:0=L^{\prime\prime}v\}=\{0\}\times\ker(L^{\prime\prime}). Therefore ker(πU)={(0,0)}\ker(\pi_{U})=\{(0,0)\} if and only if ker(L′′)={0}\ker(L^{\prime\prime})=\{0\}. ∎

We now prove Corollary 1.2 from the Introduction.

Proof of Corollary 1.2. To prove that the sequence dim𝕄(k,N,ϵ)[d]\dim\mathbb{M}(k,N,\epsilon)[d] is monotonically decreasing for d16Nkd\geq\frac{1}{6}Nk, it suffices to prove that the projection πd:𝕄(k,N,ϵ)[d+1]𝕄(k,N,ϵ)[d]\pi_{d}:\mathbb{M}(k,N,\epsilon)[d+1]\to\mathbb{M}(k,N,\epsilon)[d], sending (ϕm)m=0d+1\left(\phi_{m}\right)_{m=0}^{d+1} to (ϕm)m=0d\left(\phi_{m}\right)_{m=0}^{d}, injects. To employ the notation of Lemma 9.7, we set U=𝕄(k,N,ϵ)[d]U=\mathbb{M}(k,N,\epsilon)[d] and V=Jk,N(d+1)V=J_{k,N(d+1)}. To define WW, first define the finite set I={(m,r)2:0md, 4N(d+1)mr2}I=\{(m,r)\in{\mathbb{Z}}^{2}:0\leq m\leq d,\,4N(d+1)m\geq r^{2}\} and set W=|I|W={\mathbb{C}}^{|I|}. Define a linear map L:UWL^{\prime}:U\to W by L((ϕm)m=0d)=(c(d+1,r;ϕm))(m,r)IL^{\prime}\left(\left(\phi_{m}\right)_{m=0}^{d}\right)=\left(c(d+1,r;\phi_{m})\right)_{(m,r)\in I}, and a linear map L′′:VWL^{\prime\prime}:V\to W by L′′(ϕd+1)=((1)kϵc(m,r;ϕd+1))(m,r)IL^{\prime\prime}\left(\phi_{d+1}\right)=\left((-1)^{k}\epsilon\,c(m,r;\phi_{d+1})\right)_{(m,r)\in I}. From the definition of L′′L^{\prime\prime} we see that the kernel of L′′L^{\prime\prime} consists of the Jacobi forms vanishing to order d+1d+1, so that ker(L′′)=Jk,N(d+1)(d+1)\ker(L^{\prime\prime})=J_{k,N(d+1)}(d+1). From equation (1) we see that 𝕄(k,N,ϵ)[d]X\mathbb{M}(k,N,\epsilon)[d]\subseteq X for

X={((ϕm)m=0d,ϕd+1)U×V:L((ϕm)m=0d)=L′′(ϕd+1)}.X=\{(\left(\phi_{m}\right)_{m=0}^{d},\phi_{d+1})\in U\times V:L^{\prime}\left(\left(\phi_{m}\right)_{m=0}^{d}\right)=L^{\prime\prime}(\phi_{d+1})\}.

Since d+1>16Nkd+1>\frac{1}{6}Nk, Proposition 4.5 implies ker(L′′)=Jk,N(d+1)(d+1)={0}\ker(L^{\prime\prime})=J_{k,N(d+1)}(d+1)=\{0\}, so that πU:XU\pi_{U}:X\to U injects by Lemma 9.7. Since πd\pi_{d} is the restriction of πU\pi_{U} to 𝕄(k,N,ϵ)[d+1]\mathbb{M}(k,N,\epsilon)[d+1], we have shown that the projection πd:𝕄(k,N,ϵ)[d+1]𝕄(k,N,ϵ)[d]\pi_{d}:\mathbb{M}(k,N,\epsilon)[d+1]\to\mathbb{M}(k,N,\epsilon)[d] injects for d16Nkd\geq\frac{1}{6}Nk.

For the inverse system πd:𝕄(k,N,ϵ)[d+1]𝕄(k,N,ϵ)[d]\pi_{d}:\mathbb{M}(k,N,\epsilon)[d+1]\to\mathbb{M}(k,N,\epsilon)[d] we have 𝕄(k,N,ϵ)=limd𝕄(k,N,ϵ)[d]\mathbb{M}(k,N,\epsilon)=\varprojlim_{d}\mathbb{M}(k,N,\epsilon)[d]. The reason for this is that any equation c(n,r;ϕm)=(1)kϵc(m,r;ϕm)c(n,r;\phi_{m})=(-1)^{k}\epsilon\,c(m,r;\phi_{m}) imposed by equation (1) on 𝔣=(ϕm)m=0𝕄(k,N,ϵ)\mathfrak{f}=\left(\phi_{m}\right)_{m=0}^{\infty}\in\mathbb{M}(k,N,\epsilon) is also imposed on πd𝔣=(ϕm)m=0d𝕄(k,N,ϵ)[d]\pi_{d}\mathfrak{f}=\left(\phi_{m}\right)_{m=0}^{d}\in\mathbb{M}(k,N,\epsilon)[d] for all dmax(m,n)d\geq\max(m,n).

Since the sequence dim𝕄(k,N,ϵ)[d]0\dim\mathbb{M}(k,N,\epsilon)[d]\in{\mathbb{N}}_{0} is eventually monotonically decreasing, it is eventually constant, and the injective πd\pi_{d} are eventually isomorphisms. Thus dim𝕄(k,N,ϵ)=limddim𝕄(k,N,ϵ)[d]\dim\mathbb{M}(k,N,\epsilon)=\lim_{d}\dim\mathbb{M}(k,N,\epsilon)[d]. From Theorem 9.6 we have dimMk(K(N))ϵ=dim𝕄(k,N,ϵ)\dim M_{k}\left(K(N)\right)^{\epsilon}=\dim\mathbb{M}(k,N,\epsilon) and this completes the proof.

10. Concluding Remarks.

The Main Theorem 9.6 gives an quick proof, without needing to check convergence, of the existence of the Gritsenko lift [18], Grit:Jk,NMk(K(N))ϵ\operatorname{Grit}:J_{k,N}\to M_{k}\left(K(N)\right)^{\epsilon}, for ϵ=(1)k\epsilon=(-1)^{k}. For \ell\in{\mathbb{N}}, let V:Jk,mJk,mV_{\ell}:J_{k,m}\to J_{k,m\ell} be the index raising operator from [11]. For ϕJk,N\phi\in J_{k,N}, the formal series 𝔣=c(0,0;ϕ)12ζ(1k)Ek+m(ϕ|Vm)ξNm\mathfrak{f}=c(0,0;\phi)\frac{1}{2}\zeta(1-k)E_{k}+\sum_{m\in{\mathbb{N}}}\left(\phi|V_{m}\right)\,\xi^{Nm} is directly checked to be in 𝕄(k,N,ϵ)\mathbb{M}(k,N,\epsilon). Here, the first term with the elliptic Eisenstein series EkE_{k} is zero unless k4k\geq 4 is even. By the Main Theorem there is a Grit(ϕ)Mk(K(N))ϵ\operatorname{Grit}(\phi)\in M_{k}\left(K(N)\right)^{\epsilon} with FJ(Grit(ϕ))=𝔣{\operatorname{FJ}}\left(\operatorname{Grit}(\phi)\right)=\mathfrak{f}.

The Main Theorem 9.6 can be used to construct the global level raising operators in the paramodular newform theory of Roberts and Schmidt [35, 36]. For a prime \ell, the three level raising operators θ,θ:Sk(K(N))Sk(K(N))\theta_{\ell},\theta^{\prime}_{\ell}:S_{k}\left(K(N)\right)\to S_{k}\left(K(N\ell)\right), and η:Sk(K(N))Sk(K(N2))\eta_{\ell}:S_{k}\left(K(N)\right)\to S_{k}\left(K(N\ell^{2})\right), are used to create the oldforms. We can directly define the global level raising operators by giving their action on Fourier-Jacobi expansions. For \ell\in{\mathbb{N}}, let U:Jk,mJk,m2U_{\ell}:J_{k,m}\to J_{k,m\ell^{2}} be the index raising operator from [11]. The simplest to define is η\eta_{\ell}. For fSk(K(N))ϵf\in S_{k}\left(K(N)\right)^{\epsilon} with FJ(f)=m=1ϕmξNm{\operatorname{FJ}}(f)=\sum_{m=1}^{\infty}\phi_{m}\,\xi^{Nm}, define A(f)=m=1(ϕm|U)ξN2m𝕊(k,N2)A(f)=\sum_{m=1}^{\infty}(\phi_{m}|U_{\ell})\xi^{N\ell^{2}m}\in\mathbb{S}(k,N\ell^{2}). One directly checks A(f)𝕊(k,N2,ϵ)A(f)\in\mathbb{S}(k,N\ell^{2},\epsilon) so that by the main result there is an ηfSk(K(N2))ϵ\eta_{\ell}f\in S_{k}\left(K(N\ell^{2})\right)^{\epsilon} with FJ(ηf)=A(f){\operatorname{FJ}}(\eta_{\ell}f)=A(f). This defines η:Sk(K(N))Sk(K(N2))\eta_{\ell}:S_{k}\left(K(N)\right)\to S_{k}\left(K(N\ell^{2})\right) since the entire space is the direct sum of the plus and minus forms.

In order to define θ\theta_{\ell} and θ\theta^{\prime}_{\ell}, set Ξf=m=1(ϕm|V)ξNm𝕊(k,N)\Xi_{\ell}f=\sum_{m=1}^{\infty}(\phi_{m}|V_{\ell})\xi^{N\ell m}\in\mathbb{S}(k,N\ell), and Ξf=m=1(δgcd(,m)δk1ϕmδ2|Uδ)ξNm𝕊(k,N)\Xi^{\prime}_{\ell}f=\sum_{m=1}^{\infty}\left(\sum_{\delta\mid\gcd(\ell,m)}\delta^{k-1}\phi_{\frac{\ell m}{\delta^{2}}}|U_{\delta}\right)\xi^{N\ell m}\in\mathbb{S}(k,N\ell). One directly checks B±(f)=Ξ(f)±Ξ(f|μN)𝕊(k,N,±1)B^{\pm}(f)=\Xi_{\ell}(f)\pm\Xi^{\prime}_{\ell}(f|\mu_{N})\in\mathbb{S}(k,N\ell,\pm 1) so that by the main result there are b±(f)Sk(K(N))±b^{\pm}(f)\in S_{k}\left(K(N\ell)\right)^{\pm} with FJ(b±(f))=B±(f){\operatorname{FJ}}\left(b^{\pm}(f)\right)=B^{\pm}(f). Defining θf=12(b+(f)+b(f))Sk(K(N))\theta_{\ell}f=\frac{1}{2}(b^{+}(f)+b^{-}(f))\in S_{k}\left(K(N\ell)\right) and correspondingly θf=12(b+(f|μN)b(f|μN))Sk(K(N))\theta^{\prime}_{\ell}f=\frac{1}{2}(b^{+}(f|\mu_{N})-b^{-}(f|\mu_{N}))\in S_{k}\left(K(N\ell)\right), we have θ,θ:Sk(K(N))Sk(K(N))\theta_{\ell},\theta^{\prime}_{\ell}:S_{k}\left(K(N)\right)\to S_{k}\left(K(N\ell)\right) with FJ(θf)=Ξf{\operatorname{FJ}}\left(\theta_{\ell}f\right)=\Xi_{\ell}f, FJ(θf)=Ξf{\operatorname{FJ}}\left(\theta^{\prime}_{\ell}f\right)=\Xi^{\prime}_{\ell}f, and (θf)|μN=θ(f|μN)(\theta_{\ell}f)|\mu_{N\ell}=\theta^{\prime}_{\ell}(f|\mu_{N}). Since these constructions work for \ell\in{\mathbb{N}}, this gives a generalization of the global level raising operators.

The theory of formal series for arithmetic subgroups developed in [5] frames different hypotheses than we do in the case of the paramodular groups. Still, in their Theorem 4.8, Bruinier and Raum considered the implications of their general theory for paramodular groups when a single formal series at the standard 11-cusp is given. In our notation, they proved the following. For squarefree NN, let Γ0(N)\Gamma_{0}(N)^{*} be the extension of Γ0(N)\Gamma_{0}(N) by all the Atkin-Lehner involutions. Let χ:Γ0(N){±1}\chi:\Gamma_{0}(N)^{*}\to\{\pm 1\} be a character trivial on Γ0(N)\Gamma_{0}(N). If a formal series 𝔣𝕄(k,N)\mathfrak{f}\in\mathbb{M}(k,N) satisfies j(σ)AFS(𝔣)=χ(σ)AFS(𝔣)\operatorname{j}(\sigma^{*}){\operatorname{AFS}}(\mathfrak{f})=\chi(\sigma){\operatorname{AFS}}(\mathfrak{f}) for all σΓ0(N)\sigma\in\Gamma_{0}(N)^{*}, then 𝔣\mathfrak{f} converges on 2\mathcal{H}_{2}. Using simple generators for Γ0(N)\Gamma_{0}(N)^{*}, this reproves the cases N=2,3N=2,3 from [26]. In comparison, our main result allows general NN and assumes a single symmetry under the paramodular Fricke involution.

Theorem 8.2 only needed to be proven for cuspidal quotients 𝔣\mathfrak{f} because the Main Theorem 9.6 was reduced to the case of formal series of Jacobi cusp forms. With the Main Theorem in hand, however, we may improve Theorem 8.2 to obtain a general criterion for the divisibility of paramodular Fricke eigenforms that is interesting in its own right because it avoids all discussion of divisors.

Theorem 10.1.

Let k,k0,Nk,k_{0},N\in{\mathbb{N}} and ϵ,ϵ0{±1}\epsilon,\epsilon_{0}\in\{\pm 1\}. Let 𝔣𝕄(k,N)\mathfrak{f}\in\mathbb{M}(k,N), f0Mk0(K(N))ϵ0f_{0}\in M_{k_{0}}\left(K(N)\right)^{\epsilon_{0}} be nontrivial, and GMk0+k(K(N))ϵ0ϵG\in M_{k_{0}+k}\left(K(N)\right)^{\epsilon_{0}\epsilon}. If we have FJ(G)=FJ(f0)𝔣{\operatorname{FJ}}(G)={\operatorname{FJ}}(f_{0})\mathfrak{f} in 𝕄(k0+k,N)\mathbb{M}(k_{0}+k,N) then the meromorphic G/f0G/f_{0} is holomorphic and G/f0Mk(K(N))ϵG/f_{0}\in M_{k}\left(K(N)\right)^{\epsilon}.

Proof.

We have 𝔣𝕄(k,N,ϵ)\mathfrak{f}\in\mathbb{M}(k,N,\epsilon) if and only if j(FN)AFS(𝔣)=ϵAFS(𝔣)\operatorname{j}(F_{N}^{*}){\operatorname{AFS}}(\mathfrak{f})=\epsilon{\operatorname{AFS}}(\mathfrak{f}). Applying AFS{\operatorname{AFS}} to FJ(G)=FJ(f0)𝔣{\operatorname{FJ}}(G)={\operatorname{FJ}}(f_{0})\mathfrak{f} we obtain FS(G)=FS(f0)AFS(𝔣){\operatorname{FS}}(G)={\operatorname{FS}}(f_{0}){\operatorname{AFS}}(\mathfrak{f}). Applying j(FN)\operatorname{j}(F_{N}^{*}) we have ϵ0ϵFS(G)=ϵ0FS(f0)j(FN)AFS(𝔣)\epsilon_{0}\epsilon\,{\operatorname{FS}}(G)=\epsilon_{0}{\operatorname{FS}}(f_{0})\,\operatorname{j}(F_{N}^{*}){\operatorname{AFS}}(\mathfrak{f}). Since 𝒳¯(N){\mathbb{C}}^{{\bar{\mathcal{X}}}(N)} is an integral domain and the element FS(f0){\operatorname{FS}}(f_{0}) is nontrivial, we have j(FN)AFS(𝔣)=ϵAFS(𝔣)\operatorname{j}(F_{N}^{*}){\operatorname{AFS}}(\mathfrak{f})=\epsilon{\operatorname{AFS}}(\mathfrak{f}), and thus 𝔣𝕄(k,N,ϵ)\mathfrak{f}\in\mathbb{M}(k,N,\epsilon). By the Main Theorem 9.6 there is an fMk(K(N))ϵf\in M_{k}\left(K(N)\right)^{\epsilon} such that FJ(f)=𝔣{\operatorname{FJ}}(f)=\mathfrak{f}. Thus the quotient G/f0=fMk(K(N))ϵG/f_{0}=f\in M_{k}\left(K(N)\right)^{\epsilon} is holomorphic. ∎

References

  • [1] Hiroki Aoki. Estimating Siegel modular forms of genus 2 using Jacobi forms. J. Math. Kyoto Univ., 40(3):581–588, 2000.
  • [2] Hiroki Aoki. On the upper bound of the orders of Jacobi forms. Internat. J. Math., 33(8):Paper No. 2250056, 14, 2022.
  • [3] Eran Assaf, Watson Ladd, Gustavo Rama, Gonzalo Tornaría, and John Voight. A database of paramodular forms from quinary orthogonal modular forms. In LuCaNT: LMFDB, computation, and number theory, volume 796 of Contemp. Math., pages 243–259. Amer. Math. Soc., [Providence], RI, [2024] ©2024.
  • [4] Jeffery Breeding, II, Cris Poor, and David S. Yuen. Computations of spaces of paramodular forms of general level. J. Korean Math. Soc., 53(3):645–689, 2016.
  • [5] Jan Bruinier and Martin Raum. Formal Siegel modular forms for arithmetic subgroups. Trans. Amer. Math. Soc. Ser. B, 11:1394–1434, 2024.
  • [6] Jan Hendrik Bruinier. Vector valued formal Fourier-Jacobi series. Proc. Amer. Math. Soc., 143(2):505–512, 2015.
  • [7] Jan Hendrik Bruinier and Martin Westerholt-Raum. Kudla’s modularity conjecture and formal Fourier-Jacobi series. Forum Math. Pi, 3:7, 30, 2015.
  • [8] Armand Brumer and Kenneth Kramer. Paramodular abelian varieties of odd conductor. Trans. Amer. Math. Soc., 366(5):2463–2516, 2014.
  • [9] Armand Brumer and Kenneth Kramer. Paramodular abelian varieties of odd conductor. arXiv:1004.4699, 2018.
  • [10] N. Dummigan, A. Pacetti, G. Rama, and G. Tornaría. Quinary forms and paramodular forms. Math. Comp., 93(348):1805–1858, 2024.
  • [11] Martin Eichler and Don Zagier. The theory of Jacobi forms, volume 55 of Progress in Mathematics. Birkhäuser Boston, Inc., Boston, MA, 1985.
  • [12] E. Freitag. Siegelsche Modulfunktionen, volume 254 of Grundlehren der Mathematischen Wissenschaften [Fundamental Principles of Mathematical Sciences]. Springer-Verlag, Berlin, 1983.
  • [13] Eberhard Freitag and Rolf Busam. Complex analysis. Universitext. Springer-Verlag, Berlin, second edition, 2009.
  • [14] Klaus Fritzsche and Hans Grauert. From holomorphic functions to complex manifolds, volume 213 of Graduate Texts in Mathematics. Springer-Verlag, New York, 2002.
  • [15] Vasily Golyshev and Duco van Straten. Congruences via fibered motives. Pure Appl. Math. Q., 19(1):233–265, 2023.
  • [16] V. Gritsenko and K. Hulek. Commutator coverings of Siegel threefolds. Duke Math. J., 94(3):509–542, 1998.
  • [17] Valeri Gritsenko. Arithmetical lifting and its applications. In Number theory (Paris, 1992–1993), volume 215 of London Math. Soc. Lecture Note Ser., pages 103–126. Cambridge Univ. Press, Cambridge, 1995.
  • [18] Valeri Gritsenko. Irrationality of the moduli spaces of polarized abelian surfaces. In Abelian varieties (Egloffstein, 1993), pages 63–84. de Gruyter, Berlin, 1995. With an appendix by the author and K. Hulek.
  • [19] Valeri A. Gritsenko and Viacheslav V. Nikulin. Automorphic forms and Lorentzian Kac-Moody algebras. II. Internat. J. Math., 9(2):201–275, 1998.
  • [20] Valery Gritsenko and Haowu Wang. Powers of Jacobi triple product, Cohen’s numbers and the Ramanujan Δ\Delta-function. Eur. J. Math., 4(2):561–584, 2018.
  • [21] Robert C. Gunning. Introduction to holomorphic functions of several variables, Vol. II: Local theory. The Wadsworth & Brooks/Cole Mathematics Series. Wadsworth & Brooks/Cole Advanced Books & Software, Monterey, CA, 1990.
  • [22] Tomoyoshi Ibukiyama. On relations of dimensions of automorphic forms of Sp(2,𝐑){\rm Sp}(2,{\bf R}) and its compact twist Sp(2){\rm Sp}(2). I. In Automorphic forms and number theory (Sendai, 1983), volume 7 of Adv. Stud. Pure Math., pages 7–30. North-Holland, Amsterdam, 1985.
  • [23] Tomoyoshi Ibukiyama. Paramodular forms and compact twist. In Automorphic forms on GSp(4): Proceedings of the 9th Autumn Workshop on Number Theory, pages 37–48, 2007.
  • [24] Tomoyoshi Ibukiyama. Dimensions of paramodular forms and compact twist modular forms with involutions. arXiv:2208.13578v2, 2024.
  • [25] Tomoyoshi Ibukiyama and Hidetaka Kitayama. Dimension formulas of paramodular forms of squarefree level and comparison with inner twist. J. Math. Soc. Japan, 69(2):597–671, 2017.
  • [26] Tomoyoshi Ibukiyama, Cris Poor, and David S. Yuen. Jacobi forms that characterize paramodular forms. Abh. Math. Semin. Univ. Hambg., 83(1):111–128, 2013.
  • [27] Stephen S. Kudla. Algebraic cycles on Shimura varieties of orthogonal type. Duke Math. J., 86(1):39–78, 1997.
  • [28] Stephen S. Kudla. Integrals of Borcherds forms. Compositio Math., 137(3):293–349, 2003.
  • [29] Stephen S. Kudla. Special cycles and derivatives of Eisenstein series. In Heegner points and Rankin LL-series, volume 49 of Math. Sci. Res. Inst. Publ., pages 243–270. Cambridge Univ. Press, Cambridge, 2004.
  • [30] Joseph Lehner. A short course in automorphic functions. Holt, Rinehart and Winston, New York-Toronto-London, 1966.
  • [31] Junjiro Noguchi. Analytic function theory of several variables. Springer, Singapore, 2016. Elements of Oka’s coherence.
  • [32] Aaron Pollack. Automatic convergence for Siegel modular forms. arXiv:2408.16392v1, 2024.
  • [33] Cris Poor, Jerry Shurman, and David S. Yuen. Siegel paramodular forms of weight 2 and squarefree level. Int. J. Number Theory, 13(10):2627–2652, 2017.
  • [34] Helmut Reefschläger. Berechnung der Anzahl der 1-Spitzen der Paramodularen Gruppen 2-ten Grades. PhD thesis, Georg-August-Universität zu Göttingen, 1973.
  • [35] Brooks Roberts and Ralf Schmidt. On modular forms for the paramodular groups. In Automorphic forms and zeta functions, pages 334–364. World Sci. Publ., Hackensack, NJ, 2006.
  • [36] Brooks Roberts and Ralf Schmidt. Local Newforms for GSp(4){\rm GSp}(4), volume 1918 of Lecture Notes in Mathematics. Springer, Berlin, 2007.
  • [37] Mirko Rösner and Rainer Weissauer. Global liftings between inner forms of GSp(4){\rm GSp}(4). J. Number Theory, 263:79–138, 2024.
  • [38] Nils-Peter Skoruppa. Über den Zusammenhang zwischen Jacobiformen und Modulformen halbganzen Gewichts, volume 159 of Bonner Mathematische Schriften [Bonn Mathematical Publications]. Universität Bonn, Mathematisches Institut, Bonn, 1985. Dissertation, Rheinische Friedrich-Wilhelms-Universität, Bonn, 1984.
  • [39] Pol van Hoften. A geometric Jacquet-Langlands correspondence for paramodular Siegel threefolds. Math. Z., 299(3-4):2029–2061, 2021.
  • [40] Martin Westerholt-Raum. Formal Fourier Jacobi expansions and special cycles of codimension two. Compos. Math., 151(12):2187–2211, 2015.