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

On the non-existence of singular Borcherds products

Haowu Wang School of Mathematics and Statistics, Wuhan University, Wuhan 430072, Hubei, China [email protected]  and  Brandon Williams Lehrstuhl A für Mathematik, RWTH Aachen, 52056 Aachen, Germany Institute of Mathematics, University of Heidelberg, 69120 Heidelberg, Germany [email protected]
Abstract.

Let l3l\geq 3 and FF be a modular form of weight l/21l/2-1 on O(l,2)\mathrm{O}(l,2) which vanishes only on rational quadratic divisors. We prove that FF has only simple zeros and that FF is anti-invariant under every reflection fixing a quadratic divisor in the zeros of FF. In particular, FF is a reflective modular form. As a corollary, the existence of FF leads to l20l\leq 20 or l=26l=26, in which case FF equals the Borcherds form on II26,2\mathop{\mathrm{II}}\nolimits_{26,2}. This answers a question posed by Borcherds in 1995.

Key words and phrases:
Automorphic products, rational quadratic divisors, orthogonal modular forms, singular weight, Borcherds–Kac–Moody algebras
2020 Mathematics Subject Classification:
11F22, 11F27, 11F55

1. Introduction

In this paper we prove some nice properties of holomorphic automorphic products of singular weight on orthogonal groups O(l,2)\mathop{\hbox{}\mathrm{O}}\nolimits(l,2), and resolve an open problem posed by Borcherds in 1995.

Let MM be an even integral lattice of signature (l,2)(l,2) with l3l\geq 3. Let (,)(-,-) denote the bilinear form on MM and let MM^{\prime} be the dual lattice of MM. Either of the two connected components of the space

{[𝒵](M):(𝒵,𝒵)=0,(𝒵,𝒵¯)<0}\{[\mathcal{Z}]\in\mathbb{P}(M\otimes\mathbb{C}):(\mathcal{Z},\mathcal{Z})=0,(\mathcal{Z},\bar{\mathcal{Z}})<0\}

defines the attached Hermitian symmetric domain 𝒟(M)\mathcal{D}(M); we fix a component once and for all. We denote by O+(M)\mathop{\hbox{}\mathrm{O}}\nolimits^{+}(M) the subgroup of the orthogonal group of MM that preserves 𝒟(M)\mathcal{D}(M). Let kk be an integer, Γ\Gamma be a finite-index subgroup of O+(M)\mathop{\hbox{}\mathrm{O}}\nolimits^{+}(M) and χ\chi be a character of Γ\Gamma. A modular form of weight kk and character χ\chi for Γ\Gamma is a holomorphic function FF on the affine cone

𝒜(M)={𝒵(M):[𝒵]𝒟(M)}\mathcal{A}(M)=\{\mathcal{Z}\in\mathbb{P}(M\otimes\mathbb{C}):[\mathcal{Z}]\in\mathcal{D}(M)\}

over 𝒟(M)\mathcal{D}(M) which satisfies

F(t𝒵)\displaystyle F(t\mathcal{Z}) =tkF(𝒵),t×,\displaystyle=t^{-k}F(\mathcal{Z}),\quad\forall t\in\mathbb{C}^{\times},
F(g𝒵)\displaystyle F(g\mathcal{Z}) =χ(g)F(𝒵),gΓ.\displaystyle=\chi(g)F(\mathcal{Z}),\quad\forall g\in\Gamma.

The symmetric domain 𝒟(M)\mathcal{D}(M) can be realized as a tube domain around any cusp. The action of O+(M)\mathop{\hbox{}\mathrm{O}}\nolimits^{+}(M) on the tube domain induces an automorphy factor, with respect to which one can also define modular forms of half-integral weight. If FF is nonzero, then either k=0k=0 (in which case FF is constant), or kl/21k\geq l/2-1. The smallest possible weight l/21l/2-1 of a non-constant modular form is called the singular weight. On any tube domain realization, modular forms can be expanded into Fourier series, and singular-weight modular forms are characterized by the fact that their Fourier series is supported only on norm-zero vectors.

In 1995 and 1998 Borcherds [2, 3] established a remarkable lift to construct orthogonal modular forms with nice properties. The Borcherds lift is a multiplicative map and its input ff is a weakly holomorphic modular form of weight 1l/21-l/2 for the Weil representation of SL2()\mathop{\mathrm{SL}}\nolimits_{2}(\mathbb{Z}) attached to M/MM^{\prime}/M. The image 𝐁(f)\mathbf{B}(f) is a meromorphic modular form for the discriminant kernel

O~+(M)={gO+(M):g(v)vM,for all vM}\widetilde{\mathop{\hbox{}\mathrm{O}}\nolimits}^{+}(M)=\{g\in\mathop{\hbox{}\mathrm{O}}\nolimits^{+}(M):g(v)-v\in M,\quad\text{for all $v\in M^{\prime}$}\}

whose divisor is a linear combination of rational quadratic divisors

λ={[𝒵]𝒟(M):(𝒵,λ)=0},for λM with λ2>0.\lambda^{\perp}=\{[\mathcal{Z}]\in\mathcal{D}(M):(\mathcal{Z},\lambda)=0\},\quad\text{for $\lambda\in M$ with $\lambda^{2}>0$.}

The function 𝐁(f)\mathbf{B}(f) has an infinite product expansion around any cusp, so we call it a Borcherds product or an automorphic product.

Holomorphic Borcherds products of singular weight are very exceptional. They are usually the denominators of generalized Kac–Moody algebras that have a natural construction as the BRST cohomology related to a superconformal field theory (e.g. [4, 5, 19]). For example, the fake monster Lie algebra, which was constructed by Borcherds [4] in 1990, is the BRST cohomology related to the Leech lattice vertex operator algebra. This infinite dimensional Lie algebra describes the physical states of a bosonic string moving on a 26-dimensional torus, and its denominator is given by the singular-weight Borcherds product Φ12\Phi_{12} on II26,2\mathop{\mathrm{II}}\nolimits_{26,2}, the even unimodular lattice of signature (26,2)(26,2). Motivated by this connection, in 1995 Borcherds [2, Open problem 3 in Section 17] posed the following problems:

  1. (1)

    Are there a finite or infinite number of singular-weight modular forms which can be written as automorphic products?

  2. (2)

    Are there holomorphic automorphic products of singular weight on O(l,2)\mathop{\hbox{}\mathrm{O}}\nolimits(l,2) when l>26l>26?

These problems have remained open, although there are some partial results in the literature. Dittmann, Hagemeier and Schwagenscheidt [11] and Opitz and Schwagenscheidt [18] classified holomorphic Borcherds products of singular weight on simple lattices (i.e. lattices on which there is no obstruction to constructing Borcherds products). Note that there are only finitely many simple lattices and a full classification was given in [9]. Scheithauer [21] classified singular Borcherds products on unimodular lattices, and derived a conditional bound on the signature of lattices of prime level with singular Borcherds products.

This paper gives a complete solution to problem (2) and a weak solution to problem (1). To state the main results more conveniently, we introduce the following definition.

Definition 1.1.

A non-constant modular form for Γ<O+(M)\Gamma<\mathop{\hbox{}\mathrm{O}}\nolimits^{+}(M) is called special if its zero divisor is a linear combination of quadratic divisors λ\lambda^{\perp} for λM\lambda\in M\otimes\mathbb{Q} with λ2>0\lambda^{2}>0.

Bruinier’s converse theorem [7, 8] shows that every special modular form for the discriminant kernel of MM is a Borcherds product on MM if MM splits as UU(m)LU\oplus U(m)\oplus L, where UU is the unique even unimodular lattice of signature (1,1)(1,1). It is not known whether this holds for arbitrary MM (of signature (l,2)(l,2) with l3l\geq 3).

In this paper, we first describe zeros of special modular forms of singular weight. The proof follows from studying the Laplace operator on the tube domain around any cusp.

Theorem 1.2.

Let MM be an even lattice of signature (l,2)(l,2) with l3l\geq 3 and FF be a special modular form of singular weight for a finite-index subgroup of O+(M)\mathop{\hbox{}\mathrm{O}}\nolimits^{+}(M). Then FF has only simple zeros. Moreover, if FF vanishes on the quadratic divisor λ\lambda^{\perp}, then we have

F(σλ(𝒵))=F(𝒵),F(\sigma_{\lambda}(\mathcal{Z}))=-F(\mathcal{Z}),

where σλ\sigma_{\lambda} is the reflection fixing λ\lambda^{\perp} defined as

σλ(v)=v2(λ,v)(λ,λ)λ,vM.\sigma_{\lambda}(v)=v-\frac{2(\lambda,v)}{(\lambda,\lambda)}\lambda,\quad v\in M\otimes\mathbb{Q}.

Following Borcherds [3] and Gritsenko–Nikulin [13], we define reflective modular forms.

Definition 1.3.

Let FF be a special modular form for a finite-index subgroup of O+(M)\mathop{\hbox{}\mathrm{O}}\nolimits^{+}(M). If σλ\sigma_{\lambda} lies in O+(M)\mathop{\hbox{}\mathrm{O}}\nolimits^{+}(M) whenever FF vanishes on λ\lambda^{\perp}, then FF is called reflective.

As a corollary of Theorem 1.2, we prove the following result, which is a conjecture formulated by the first named author in 2019 (see [26, Conjecture 9.8]).

Theorem 1.4.

Let MM be an even lattice of signature (l,2)(l,2) with l3l\geq 3 and FF be a special modular form of singular weight for a finite-index subgroup Γ\Gamma of O+(M)\mathop{\hbox{}\mathrm{O}}\nolimits^{+}(M). Then there exists an even lattice 𝕄\mathbb{M} on MM\otimes\mathbb{Q} such that Γ<O+(𝕄)\Gamma<\mathop{\hbox{}\mathrm{O}}\nolimits^{+}(\mathbb{M}) and FF is a reflective modular form on 𝕄\mathbb{M}.

Theorem 1.4 reduces the difficult problem of classifying singular automorphic products to the easier problem of classifying reflective modular forms. A full classification of reflective modular forms is not yet available, but many partial results have been obtained over the past two decades [13, 12, 1, 20, 21, 15, 16, 10, 22, 14, 26, 23, 24, 25]. In particular, the first named author [22, 26, 25] classified lattices of large rank which has a reflective modular form. This yields a complete solution to Borcherds’ problem (2).

Theorem 1.5.

There is no special modular form of singular weight on O(l,2)\mathop{\hbox{}\mathrm{O}}\nolimits(l,2) when l21l\geq 21 and l26l\neq 26. The Borcherds form Φ12\Phi_{12} is the unique special modular form of singular weight on O(26,2)\mathop{\hbox{}\mathrm{O}}\nolimits(26,2) up to a constant factor.

Special modular forms of singular weight on O(l,2)\mathop{\hbox{}\mathrm{O}}\nolimits(l,2) do exist for l=3l=3, 44, 66, 88, 1010, 1212, 1414, 1818. They do not appear to exist on O(l,2)\mathop{\hbox{}\mathrm{O}}\nolimits(l,2) for any other l20l\leq 20.

The finiteness of lattices with reflective modular forms is known in certain specific cases [15, 16, 25]. However, the finiteness of reflective modular forms is unknown in general. Therefore, we cannot give a full solution to Borcherds’ problem (1). In 2018 Ma [16, Corollary 1.10] proved that up to scaling there are only finitely many lattices of signature (l,2)(l,2) with l4l\geq 4 which has a reflective modular form with simple zeros. Theorem 1.2, Theorem 1.4 and Ma’s result together yield the following weak solution to Borcherds’ problem (1).

Corollary 1.6.

The set of lattices 𝕄\mathbb{M} in Theorem 1.4 is finite up to scaling when l4l\geq 4.

This paper is organized as follows. In Section 2 we introduce the Laplace operator and prove Theorem 1.2. In Section 3 we prove Theorem 1.4 and give a corollary. Section 4 is devoted to the proof of Theorem 1.5.

2. Proof of Theorem 1.2

In this section we use the Laplace operator on a tube domain to prove Theorem 1.2.

Let MM be an even lattice of signature (l,2)(l,2) with l3l\geq 3. Let cMc\in M be a primitive norm-zero vector and cMc^{\prime}\in M^{\prime} with (c,c)=1(c,c^{\prime})=1. Let e1,,ele_{1},...,e_{l} be any \mathbb{R}-basis of the signature (l1,1)(l-1,1) lattice

Mc,c={xM:(x,c)=(x,c)=0}M_{c,c^{\prime}}=\{x\in M:(x,c)=(x,c^{\prime})=0\}

and define the Gram matrix S=(sij)i,j=1lS=(s_{ij})_{i,j=1}^{l} with inverse S1=(sij)i,jS^{-1}=(s^{ij})_{i,j}, where sij:=(ei,ej)s_{ij}:=(e_{i},e_{j}). The holomorphic Laplace operator is defined as

𝚫=𝚫c,c=12i,j=1lsij2eiej.\mathbf{\Delta}=\mathbf{\Delta}_{c,c^{\prime}}=\frac{1}{2}\sum_{i,j=1}^{l}s^{ij}\frac{\partial^{2}}{\partial e_{i}\partial e_{j}}.

This operator acts on functions defined on the tube domain c,c\mathbb{H}_{c,c^{\prime}} which is a connected component of the space

{Z=X+iY:X,YMc,c,(Y,Y)<0}.\{Z=X+iY:\;X,Y\in M_{c,c^{\prime}}\otimes\mathbb{R},\;(Y,Y)<0\}.

It is clear that 𝚫c,c\mathbf{\Delta}_{c,c^{\prime}} is independent of the basis eie_{i}.

For any λMc,c\lambda\in M_{c,c^{\prime}}\otimes\mathbb{Q}, we have

𝚫e2πi(λ,Z)=2π2(λ,λ)e2πi(λ,Z).\mathbf{\Delta}e^{2\pi i(\lambda,Z)}=-2\pi^{2}(\lambda,\lambda)e^{2\pi i(\lambda,Z)}.

Therefore, a function FF that is represented near cc by a Fourier series

F(Z)=λMc,cc(λ)e2πi(λ,Z)F(Z)=\sum_{\lambda\in M_{c,c^{\prime}}\otimes\mathbb{Q}}c(\lambda)e^{2\pi i(\lambda,Z)}

satisfies 𝚫F=0\mathbf{\Delta}F=0 if and only if it is singular at cc, that is, c(λ)=0c(\lambda)=0 for every λ\lambda of nonzero norm.

For any meromorphic function F:c,cF:\mathbb{H}_{c,c^{\prime}}\rightarrow\mathbb{C}, matrix elements gO+(M)g\in\mathrm{O}^{+}(M\otimes\mathbb{R}) and positive integer mm, we have (see e.g. [27, Lemma 2.4])

(𝚫mF)|l/2+mg=𝚫m(F|l/2mg).(\mathbf{\Delta}^{m}F)\Big{|}_{l/2+m}g=\mathbf{\Delta}^{m}\Big{(}F\Big{|}_{l/2-m}g\Big{)}.

In particular, if FF transforms like a modular form of weight l/2ml/2-m then 𝚫mF\mathbf{\Delta}^{m}F transforms like a modular form of weight l/2+ml/2+m. If FF is in fact a non-constant (holomorphic) modular form, then mm must be 1, i.e. FF is of singular weight, and thus 𝚫F=0\mathbf{\Delta}F=0.

Theorem 1.2 is a particular case of the following result.

Theorem 2.1.

Let FF be a modular form of singular weight on Γ<O+(M)\Gamma<\mathop{\hbox{}\mathrm{O}}\nolimits^{+}(M). Let vv be a positive-norm vector of MM. If FF vanishes on the quadratic divisor vv^{\perp}, then vv^{\perp} has multiplicity one in the divisor of FF and FF is anti-invariant under the associated reflection, i.e.

F(σv(𝒵))=F(𝒵).F(\sigma_{v}(\mathcal{Z}))=-F(\mathcal{Z}).
Proof.

First, note that the existence of a singular-weight modular form FF implies that MM contains an isotropic plane, i.e. 𝒟(M)\mathcal{D}(M) contains a one-dimensional cusp. (Indeed, MM has rank at least 55 and therefore contains a primitive isotropic vector cc. The Fourier expansion of FF about cc is supported only on norm-zero vectors orthogonal to cc, and as FF is non-constant, such vectors must exist.)

Fix a primitive isotropic vector cMc\in M and a vector cMc^{\prime}\in M^{\prime} with (c,c)=1(c,c^{\prime})=1 such that vMc,cv\in M_{c,c^{\prime}}. We view FF as a modular form on the associated tube domain c,c\mathbb{H}_{c,c^{\prime}}. Let KK be the orthogonal complement of vv in Mc,cM_{c,c^{\prime}}. We decompose ZMc,cZ\in M_{c,c^{\prime}}\otimes\mathbb{C} as Z=z1v+z2Z=z_{1}v+z_{2} for z1z_{1}\in\mathbb{C} and z2Kz_{2}\in K\otimes\mathbb{C}.

If dd is the multiplicity of vv^{\perp} in the divisor of FF, then the Taylor series of FF at z1=0z_{1}=0 has the form

F(Z)=fd(z2)z1d+O(z1d+1),fd(z2)0.F(Z)=f_{d}(z_{2})z_{1}^{d}+O(z_{1}^{d+1}),\quad f_{d}(z_{2})\neq 0.

By applying the Laplace operator on the tube domain c,c\mathbb{H}_{c,c^{\prime}} to FF, we have

0=𝚫c,c(F)=d(d1)2fd(z2)z1d2+O(z1d1),0=\mathbf{\Delta}_{c,c^{\prime}}(F)=\frac{d(d-1)}{2}f_{d}(z_{2})z_{1}^{d-2}+O(z_{1}^{d-1}),

which shows that d=1d=1. Similarly, applying the Laplace operator to the function

G(Z):=F(Z)+F|σv(Z)=F(z1v+z2)+F(z1v+z2)=O(z12)G(Z):=F(Z)+F|\sigma_{v}(Z)=F(z_{1}v+z_{2})+F(-z_{1}v+z_{2})=O(z_{1}^{2})

we find that 𝚫(G)=0\mathbf{\Delta}(G)=0, which forces F(Z)+F|σv(Z)=0F(Z)+F|\sigma_{v}(Z)=0. ∎

3. Proof of Theorem 1.4

We divide the proof of Theorem 1.4 into two lemmas. The notation is the same as Theorem 1.4.

Lemma 3.1.

Let ΓF\Gamma_{F} be the group generated by Γ\Gamma and the reflections σr\sigma_{r} for which FF vanishes on rr^{\perp}. Then ΓF\Gamma_{F} is arithmetic in O+(M)\mathop{\hbox{}\mathrm{O}}\nolimits^{+}(M\otimes\mathbb{Q}). In particular, ΓFO+(M)\Gamma_{F}\cap\mathop{\hbox{}\mathrm{O}}\nolimits^{+}(M) has finite index in ΓF\Gamma_{F}.

Proof.

Let mm be the order of the character of F2F^{2} on Γ\Gamma. Then F2m|g=F2mF^{2m}|g=F^{2m} for all gΓFg\in\Gamma_{F} by Theorem 2.1. It follows that ΓF\Gamma_{F} is a discrete subgroup of O+(M)\mathop{\hbox{}\mathrm{O}}\nolimits^{+}(M\otimes\mathbb{R}). Since 𝒟(M)/Γ\mathcal{D}(M)/\Gamma has finite volume, the smaller quotient space 𝒟(M)/ΓF\mathcal{D}(M)/\Gamma_{F} also has finite volume. Therefore, ΓF\Gamma_{F} is a lattice in O+(M)\mathop{\hbox{}\mathrm{O}}\nolimits^{+}(M\otimes\mathbb{Q}) and thus an arithmetic subgroup by the Margulis arithmeticity theorem [17]. ∎

Lemma 3.2.

The lattice 𝕄\mathbb{M} in Theorem 1.4 can be obtained by rescaling the rational lattice generated by g(M)g(M), where gg runs over representatives of ΓF/(ΓFO+(M))\Gamma_{F}/(\Gamma_{F}\cap\mathop{\hbox{}\mathrm{O}}\nolimits^{+}(M)).

Proof.

Let MFM_{F} be the lattice generated by g(M)g(M) over \mathbb{Z}. Since gO+(M)g\in\mathop{\hbox{}\mathrm{O}}\nolimits^{+}(M\otimes\mathbb{Q}), the lattice MFM_{F} is rational. By construction, ΓF\Gamma_{F} fixes MFM_{F}. Lemma 3.1 shows that the set of gg is finite. Therefore, there exists dd\in\mathbb{N} such that the rescaled lattice MF(d)M_{F}(d) is integral. Since O+(MF)=O+(MF(d))\mathop{\hbox{}\mathrm{O}}\nolimits^{+}(M_{F})=\mathop{\hbox{}\mathrm{O}}\nolimits^{+}(M_{F}(d)), we can conclude that MF(d)M_{F}(d) is the desired 𝕄\mathbb{M}. ∎

For lattices that split a unimodular plane, Theorem 1.4 sharply restricts the principal parts of vector-valued modular forms that lift to holomorphic Borcherds products of singular weight:

Corollary 3.3.

Let FF be a holomorphic Borcherds product of singular weight on a lattice of type M=UKM=U\oplus K. We denote by mK\mathrm{m}_{K} the positive generator of the integral ideal generated by v2/2v^{2}/2 for vKv\in K. We write the principal part of the input of FF as

n>0xK/Kc(x,n)qn𝐞x.\sum_{n>0}\sum_{x\in K^{\prime}/K}c(x,-n)q^{-n}\mathbf{e}_{x}.

If c(x,n)0c(x,-n)\neq 0, then mK/n\mathrm{m}_{K}/n\in\mathbb{Z}.

Proof.

Write vectors λUK\lambda\in U\oplus K^{\prime} in the form (a,r,b)(a,r,b) with a,ba,b\in\mathbb{Z} and vKv\in K^{\prime}, such that λ2=r22ab\lambda^{2}=r^{2}-2ab. Suppose c(x,n)0c(x,-n)\neq 0, and let dd be the largest integer such that c(dx,d2n)0c(dx,-d^{2}n)\neq 0. For simplicity we set y=dxK/Ky=dx\in K^{\prime}/K, m=d2nm=d^{2}n, and vr:=(1,r,r2/2m)v_{r}:=(1,r,r^{2}/2-m) for any ry+Kr\in y+K. Then vrv_{r}^{\perp} has multiplicity c(y,m)c(y,-m) in the divisor of FF, so c(y,m)c(y,-m) is a positive integer.

Let tt be the minimal positive integer such that tvrtv_{r} lies in the lattice MFM_{F} generated by g(M)g(M), where gg runs through the representatives of ΓF/(ΓFO+(M))\Gamma_{F}/(\Gamma_{F}\cap\mathop{\hbox{}\mathrm{O}}\nolimits^{+}(M)) for Γ=O~+(M)\Gamma=\widetilde{\mathop{\hbox{}\mathrm{O}}\nolimits}^{+}(M) (as in Lemmas 3.1 and 3.2). Since all vectors of type vrv_{r} lie in a single O~+(M)\widetilde{\mathop{\hbox{}\mathrm{O}}\nolimits}^{+}(M)-orbit, the integer tt depends only on the class yK/Ky\in K^{\prime}/K and the number mm. For any r,sy+Kr,s\in y+K, we have σvr(tvs)MF\sigma_{v_{r}}(tv_{s})\in M_{F}, which is equivalent to

(tvs,vr)mvr=(vs,vr)mtvrMF.\frac{(tv_{s},v_{r})}{m}v_{r}=\frac{(v_{s},v_{r})}{m}\cdot tv_{r}\in M_{F}.

It follows that (vs,vr)/m(v_{s},v_{r})/m\in\mathbb{Z} for all r,sy+Kr,s\in y+K. Therefore, mK/m\mathrm{m}_{K}/m\in\mathbb{Z} and mK/(nd2)\mathrm{m}_{K}/(nd^{2})\in\mathbb{Z}. ∎

4. A proof of Theorem 1.5

In this section we prove Theorem 1.5 as a corollary of Theorem 1.4 using the following results:

Theorem 4.1 (Theorem 1.2 of [25]).
  1. (1)

    There is no even lattice of signature (l,2)(l,2) with a reflective modular form when l=21l=21 or l23l\geq 23 and l26l\neq 26.

  2. (2)

    Let M=UKM=U\oplus K be an even lattice of signature (l,2)(l,2) which has a reflective Borcherds product. If l=26l=26, then MII26,2M\cong\mathop{\mathrm{II}}\nolimits_{26,2} and if l=22l=22, then M2UD20M\cong 2U\oplus D_{20}.

Let MM be an even lattice of signature (l,2)(l,2) with l21l\geq 21 and suppose FF is a special modular form of singular weight for some finite-index subgroup Γ\Gamma of O+(M)\mathop{\hbox{}\mathrm{O}}\nolimits^{+}(M). By Theorem 1.4, Lemma 3.1 and Lemma 3.2, FF is a reflective modular form for ΓF<O+(𝕄)\Gamma_{F}<\mathop{\hbox{}\mathrm{O}}\nolimits^{+}(\mathbb{M}). Theorem 4.1 implies that l=22l=22 or 2626. By [16, Corollory 3.2], there exists a lattice 𝕄1\mathbb{M}_{1} in 𝕄\mathbb{M}\otimes\mathbb{Q} such that O+(𝕄)O+(𝕄1)\mathop{\hbox{}\mathrm{O}}\nolimits^{+}(\mathbb{M})\subset\mathop{\hbox{}\mathrm{O}}\nolimits^{+}(\mathbb{M}_{1}) and such that 𝕄1\mathbb{M}_{1} is a scaling of an even lattice of type 𝕄2=2UL\mathbb{M}_{2}=2U\oplus L. The subgroup ΓF\Gamma_{F} has finite index in O+(𝕄2)\mathop{\hbox{}\mathrm{O}}\nolimits^{+}(\mathbb{M}_{2}), and we denote the index by aa. Then the function

F^=gO+(𝕄2)/ΓFF|g\hat{F}=\prod_{g\in\mathop{\hbox{}\mathrm{O}}\nolimits^{+}(\mathbb{M}_{2})/\Gamma_{F}}F|g

defines a reflective modular form of weight a(l/21)a(l/2-1) for O+(𝕄2)\mathop{\hbox{}\mathrm{O}}\nolimits^{+}(\mathbb{M}_{2}).

When l=26l=26, we conclude from Theorem 4.1 that 𝕄2II26,2\mathbb{M}_{2}\cong\mathop{\mathrm{II}}\nolimits_{26,2} and F^=Φ12a\hat{F}=\Phi_{12}^{a} up to a constant factor. By Theorem 1.2, FF has only simple zeros. Therefore, Φ12/F\Phi_{12}/F is a (holomorphic) modular form of weight 0 for ΓF\Gamma_{F} and thus constant.

When l=22l=22, Theorem 4.1 implies that 𝕄22UD20\mathbb{M}_{2}\cong 2U\oplus D_{20}. Borcherds [6] constructed a reflective modular form Ψ24\Psi_{24} of weight 2424 for O+(2UD20)\mathop{\hbox{}\mathrm{O}}\nolimits^{+}(2U\oplus D_{20}) which vanishes on vv^{\perp} with multiplicity 11 and vanishes on uu^{\perp} with multiplicity 88, where v2UD20v\in 2U\oplus D_{20} with v2=2v^{2}=2 and u2UD20u\in 2U\oplus D_{20}^{\prime} with u2=1u^{2}=1. We know from [25, Lemma 3.2] that Ψ24\Psi_{24} is the unique reflective modular form for O+(2UD20)\mathop{\hbox{}\mathrm{O}}\nolimits^{+}(2U\oplus D_{20}) up to a power. It follows that F^=Ψ24d\hat{F}=\Psi_{24}^{d} for some positive integer dd. Recall that FF has only simple zeros. By comparing the weights and zero orders of FF and Ψ24\Psi_{24}, we find

10a=24dand8da,10a=24d\quad\text{and}\quad 8d\leq a,

which leads to 10a/3a10a/3\leq a, a contradiction. This finishes the proof of Theorem 1.5.

\bibliofont

References

  • Barnard [2003] Alexander Graham Barnard. The singular theta correspondence, Lorentzian lattices and Borcherds-Kac-Moody algebras. ProQuest LLC, Ann Arbor, MI, 2003. Thesis (Ph.D.)–University of California, Berkeley.
  • Borcherds [1995] Richard Borcherds. Automorphic forms on Os+2,2(𝐑){\rm O}_{s+2,2}({\bf R}) and infinite products. Invent. Math., 120(1):161–213, 1995.
  • Borcherds [1998] Richard Borcherds. Automorphic forms with singularities on Grassmannians. Invent. Math., 132(3):491–562, 1998.
  • Borcherds [1990] Richard E. Borcherds. The monster Lie algebra. Adv. Math., 83(1):30–47, 1990.
  • Borcherds [1992] Richard E. Borcherds. Monstrous moonshine and monstrous Lie superalgebras. Invent. Math., 109(2):405–444, 1992.
  • Borcherds [2000] Richard E. Borcherds. Reflection groups of Lorentzian lattices. Duke Math. J., 104(2):319–366, 2000.
  • Bruinier [2002] Jan H. Bruinier. Borcherds products on O(2, ll) and Chern classes of Heegner divisors, volume 1780 of Lecture Notes in Mathematics. Springer-Verlag, Berlin, 2002.
  • Bruinier [2014] Jan Hendrik Bruinier. On the converse theorem for Borcherds products. J. Algebra, 397:315–342, 2014.
  • Bruinier et al. [2016] Jan Hendrik Bruinier, Stephan Ehlen, and Eberhard Freitag. Lattices with many Borcherds products. Math. Comp., 85(300):1953–1981, 2016.
  • Dittmann [2019] Moritz Dittmann. Reflective automorphic forms on lattices of squarefree level. Trans. Amer. Math. Soc., 372(2):1333–1362, 2019.
  • Dittmann et al. [2015] Moritz Dittmann, Heike Hagemeier, and Markus Schwagenscheidt. Automorphic products of singular weight for simple lattices. Math. Z., 279(1-2):585–603, 2015.
  • Gritsenko and Nikulin [2002] V. A. Gritsenko and V. V. Nikulin. On the classification of Lorentzian Kac-Moody algebras. Uspekhi Mat. Nauk, 57(5(347)):79–138, 2002.
  • Gritsenko and Nikulin [1998] Valeri A. Gritsenko and Viacheslav V. Nikulin. Automorphic forms and Lorentzian Kac-Moody algebras. II. Internat. J. Math., 9(2):201–275, 1998.
  • Gritsenko and Nikulin [2018] Valery Gritsenko and Viacheslav V. Nikulin. Lorentzian Kac-Moody algebras with Weyl groups of 2-reflections. Proc. Lond. Math. Soc. (3), 116(3):485–533, 2018.
  • Ma [2017] Shouhei Ma. Finiteness of 2-reflective lattices of signature (2,n)(2,n). Amer. J. Math., 139(2):513–524, 2017.
  • Ma [2018] Shouhei Ma. On the Kodaira dimension of orthogonal modular varieties. Invent. Math., 212(3):859–911, 2018.
  • Margulis [1991] G. A. Margulis. Discrete subgroups of semisimple Lie groups, volume 17 of Ergebnisse der Mathematik und ihrer Grenzgebiete (3) [Results in Mathematics and Related Areas (3)]. Springer-Verlag, Berlin, 1991.
  • Opitz and Schwagenscheidt [2019] Sebastian Opitz and Markus Schwagenscheidt. Holomorphic Borcherds products of singular weight for simple lattices of arbitrary level. Proc. Amer. Math. Soc., 147(11):4639–4653, 2019.
  • Scheithauer [2000] Nils R. Scheithauer. The fake monster superalgebra. Adv. Math., 151(2):226–269, 2000.
  • Scheithauer [2006] Nils R. Scheithauer. On the classification of automorphic products and generalized Kac-Moody algebras. Invent. Math., 164(3):641–678, 2006.
  • Scheithauer [2017] Nils R. Scheithauer. Automorphic products of singular weight. Compos. Math., 153(9):1855–1892, 2017.
  • Wang [2021] Haowu Wang. Reflective modular forms: a Jacobi forms approach. Int. Math. Res. Not. IMRN, (3):2081–2107, 2021.
  • Wang [2022] Haowu Wang. Reflective modular forms on lattices of prime level. Trans. Amer. Math. Soc., 375(5):3451–3468, 2022.
  • Wang [2023a] Haowu Wang. 2-reflective lattices of signature (n,2)(n,2) with n8n\geq 8. Int. Math. Res. Not. IMRN, (20):17953–17971, 2023a.
  • Wang [2023b] Haowu Wang. On the classification of reflective modular forms. preprint, 2023b. URL arXiv:2301.12606.
  • Wang [2024] Haowu Wang. The classification of 2-reflective modular forms. J. Eur. Math. Soc. (JEMS), 26(1):111–151, 2024.
  • Williams [2021] Brandon Williams. Higher pullbacks of modular forms on orthogonal groups. Forum Math., 33(3):631–652, 2021.