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

Exceptional theta functions and arithmeticity of modular forms on G2G_{2}

Aaron Pollack Department of Mathematics
The University of California San Diego
La Jolla, CA USA
[email protected]
Abstract.

Quaternionic modular forms on the split exceptional group G2=G2sG_{2}=G_{2}^{s} were defined by Gan-Gross-Savin. A remarkable property of these automorphic functions is that they have a robust notion of Fourier expansion and Fourier coefficients, similar to the classical holomorphic modular forms on Shimura varieties. In this paper we prove that in even weight \ell at least 66, there is a basis of the space of cuspidal modular forms of weight \ell such that all the Fourier coefficients of elements of this basis are in the cyclotomic extension of 𝐐{\mathbf{Q}}.

Our main tool for proving this is to develop a notion of “exceptional theta functions” on G2G_{2}. We also develop a notion of exceptional theta functions on Sp6\operatorname{Sp}_{6}. In the case of Sp6\operatorname{Sp}_{6}, these are level one, holomorphic vector-valued Siegel modular forms, with explicit Fourier expansions, that are the theta lifts from algebraic modular forms on anisotropic G2G_{2} for the dual pair Sp6×G2aE7,3\operatorname{Sp}_{6}\times G_{2}^{a}\subseteq E_{7,3}. In the case of split G2G_{2}, our exceptional theta functions are level one quaternionic modular forms, with explicit Fourier expansions, that are the theta lifts from algebraic modular forms on anistropic F4F_{4} for the dual pair G2×F4IE8,4G_{2}\times F_{4}^{I}\subseteq E_{8,4}.

As further consequences of this theory of exceptional theta functions, we also obtain the following corollaries: 1) there is an algorithm to determine if any cuspidal, level one Siegel modular form on Sp6\operatorname{Sp}_{6} of most weights is a lift from G2aG_{2}^{a}; 2) the level one theta lifts from F4IF_{4}^{I} possess an integral structure, in the sense of Fourier coefficients; 3) in every even weight k6k\geq 6, there is a nonzero, level one Hecke eigen quaternionic cusp form on split G2G_{2} with nonzero 𝐙×𝐙×𝐙{\mathbf{Z}}\times{\mathbf{Z}}\times{\mathbf{Z}} Fourier coefficient. Finally, we obtain evidence for a conjecture of Gross relating Fourier coefficients of certain G2G_{2} quaternionic modular forms to LL-values.

Funding information: AP has been supported by the Simons Foundation via Collaboration Grant number 585147, by the NSF via grant numbers 2101888 and 2144021, and by an AMS Centennial Research Fellowship. Part of this work was done while the author visited the Erwin Schrodinger Institute in Vienna, and the author thanks them for their support and hospitality. The author would also like to thank the Isaac Newton Institute for Mathematical Sciences, Cambridge, for support and hospitality during the programme New Connections in Number Theory and Physics where some work on this paper was undertaken. The work at the INI was supported by EPSRC grant no EP/R014604/1 and 34.

1. Introduction

Holomorphic modular forms on Shimura varieties have a good notion of Fourier coefficients. It is theorem, going back to Shimura, that the Fourier coefficients of holomorphic modular forms on Hermitian tube domains give these modular forms an algebraic structure: there is a basis of the space of holomorphic modular forms so that all the Fourier coefficients of elements of this basis are algebraic.

Outside the realm of holomorphic modular forms on Shimura varieties, there is little one can say111One exception might be the case of globally generic cohomological automorphic forms, for which the Whittaker coefficients can be directly related to Satake paremeters. about the arithmeticity of Fourier coefficients of automorphic functions. In fact, at present, it is not clear in general how one might define a good notion of Fourier coefficients for spaces of non-holomorphic automorphic forms.

Nevertheless, the quaternionic exceptional groups possess a special class of automorphic functions called the quaternionic modular forms, which do have a good notion of Fourier expansion and Fourier coefficients. These automorphic forms go back to Gross-Wallach [GW94, GW96] and Gan-Gross-Savin [GGS02]. The precise shape of their Fourier expansion was determined in [Pol20a], extending and refining earlier work of Wallach [Wal03]. In particular, it is possible to define what it means for a quaternionic modular form to have Fourier coefficients in some ring R𝐂R\subseteq{\mathbf{C}}. In [Pol22], we conjectured that the space of quaternionic modular forms of some fixed weight \ell on a quaternionic exceptional group GG has a basis consisting of elements all of whose Fourier coefficients are algebraic numbers.222It seems useful to point out that there has been other recent work, specifically [PV21], that conjectures the existence of surprising algebraic structures on spaces of non-holomorphic automorphic forms.

In this paper, we provide substantial evidence toward this conjecture in the case of GG the split exceptional group G2G_{2}. We setup the result now.

Suppose φ\varphi is a cuspidal quaternionic modular form on G2G_{2} of weight 1\ell\geq 1. Let NN denote the unipotent radical of the Heisenberg parabolic of G2G_{2} and ZZ its center. Then

φZ(gfg)=χaχ(gf)Wχ(g)\varphi_{Z}(g_{f}g_{\infty})=\sum_{\chi}{a_{\chi}(g_{f})W_{\chi}(g_{\infty})}

is the Fourier expansion of φ\varphi. Here the χ\chi range over non-degenerate characters of N(𝐐)\N(𝐀)N({\mathbf{Q}})\backslash N({\mathbf{A}}) and the WχW_{\chi} are the generalized Whittaker functions of [Pol20a]. The functions aχ:G2(𝐀f)𝐂a_{\chi}:G_{2}({\mathbf{A}}_{f})\rightarrow{\mathbf{C}} are locally constant and called the Fourier coefficients of φ\varphi. We say that φ\varphi has Fourier coefficients in a ring RR if aχ(gf)Ra_{\chi}(g_{f})\in R for all characters χ\chi of N(𝐐)\N(𝐀)N({\mathbf{Q}})\backslash N({\mathbf{A}}) and all gfG2(𝐀f)g_{f}\in G_{2}({\mathbf{A}}_{f}). We write S(G2;R)S_{\ell}(G_{2};R) for the space cuspidal quaternionic modular forms on G2G_{2} of weight \ell with Fourier coefficients in RR.

Let 𝐐cyc=𝐐(μ){\mathbf{Q}}_{cyc}={\mathbf{Q}}(\mu_{\infty}) be the cyclotomic extension of 𝐐{\mathbf{Q}}.

Theorem 1.0.1.

Suppose 6\ell\geq 6 is even. Then there is a basis of the cuspidal quaternionic modular forms of weight \ell with all Fourier coefficients in 𝐐cyc{\mathbf{Q}}_{cyc}. In other words, S(G2,𝐂)=S(G2,𝐐cyc)𝐐cyc𝐂S_{\ell}(G_{2},{\mathbf{C}})=S_{\ell}(G_{2},{\mathbf{Q}}_{cyc})\otimes_{{\mathbf{Q}}_{cyc}}{\mathbf{C}}.

Our main tool for proving Theorem 1.0.1 is a notion of “exceptional” theta functions, that mirrors the classical theory of Siegel modular theta functions associated to pluriharmonic polynomials. Recall that these classical pluriharmonic theta functions are (often cuspidal) Siegel modular forms, with completely explicit Fourier expansions, that can be considered as arising from the Weil representation restricted to Sp2g×O(V)\operatorname{Sp}_{2g}\times O(V) where VV is a rational quadratic space whose quadratic form is positive definite. In other words, they are the theta lifts from algebraic modular forms on O(V)O(V) with nontrivial archimedean weight to holomorphic Siegel modular forms. From the perspective of theta lifts and algebraic modular forms, the remarkable fact is that one can give the Fourier expansions of these lifts completely explicitly.

To prove Theorem 1.0.1, we develop a notion of exceptional theta functions on G2G_{2}. These are quaternionic modular forms on G2G_{2} whose Fourier coefficients we can tightly control. They arise as theta lifts from an anisotropic group of type F4F_{4}. Using the Siegel-Weil theorem of [Pol23], we can prove that every cuspidal quaternionic modular form of even weight at least 66 on G2G_{2} is one of our exceptional theta functions. This establishes the theorem.

Our theory of exceptional theta functions on G2G_{2} has a parallel–but easier–development on Sp6\operatorname{Sp}_{6}. As this theory is a bit easier, and might be more familiar to the reader, we begin with the Sp6\operatorname{Sp}_{6}-case.

1.1. Holomorphic theta functions

Let G2aG_{2}^{a} denote the the algebraic 𝐐{\mathbf{Q}} group of type G2G_{2} that is split at every finite place and compact at the archimedean place. Let HJ1H_{J}^{1} denote the simply connected group of type E7E_{7} that is split at every finite place and the group E7,3E_{7,3} at the archimedean place. There is a dual pair Sp6×G2aHJ1\operatorname{Sp}_{6}\times G_{2}^{a}\subseteq H_{J}^{1}, and a corresponding theta lift from G2aG_{2}^{a} to Sp6\operatorname{Sp}_{6} using the automorphic minimal representation on HJ1H_{J}^{1} [Kim93] that was studied by Gross-Savin [GS98]. This lift produces (in general) vector-valued holomorphic Siegel modular forms on Sp6\operatorname{Sp}_{6}. It makes sense to ask if, given an algebraic modular form on G2aG_{2}^{a}, one can give an explicit Fourier expansion of its theta lift to Sp6\operatorname{Sp}_{6}. This is easy if the weight of the algebraic modular form is trivial, but is not immediate (at least to us) if the weight is nontrivial. Theorem 1.1.1 computes this Fourier expansion in the level one case.

We now setup this theorem. Let Θ\Theta denote the octonions over 𝐐{\mathbf{Q}} with positive definite norm. Denote by J=H3(Θ)J=H_{3}(\Theta) the 2727-dimensional exceptional cubic norm structure consisting of the 3×33\times 3 Hermitian matrices with coefficients in Θ\Theta. Let V7V_{7} denote the 77-dimensional space of trace 0 octonions, let W3W_{3} denote the standard representation of GL3\operatorname{GL}_{3} and set V3=2W3V_{3}=\wedge^{2}W_{3}. There is a projection 𝒫:JJV3V7\mathcal{P}:J^{\vee}\simeq J\rightarrow V_{3}\otimes V_{7} given by taking the trace 0 projections of the off-diagonal entries of an element of JJ. Consider the natural map

(V3V7)k1+2k2\displaystyle(V_{3}\otimes V_{7})^{\otimes k_{1}+2k_{2}} Sk1(V3)V7k1Sk2(2V3)(2V7)k2\displaystyle\rightarrow S^{k_{1}}(V_{3})\otimes V_{7}^{\otimes k_{1}}\otimes S^{k_{2}}(\wedge^{2}V_{3})\otimes(\wedge^{2}V_{7})^{\otimes k_{2}}
Sk1(2W3)V7k1Sk2(W3)(2V7)k2det(W3)k2,\displaystyle\simeq S^{k_{1}}(\wedge^{2}W_{3})\otimes V_{7}^{\otimes k_{1}}\otimes S^{k_{2}}(W_{3})\otimes(\wedge^{2}V_{7})^{\otimes k_{2}}\otimes\det(W_{3})^{k_{2}},

and denote by 𝒫k1,k2(T)\mathcal{P}_{k_{1},k_{2}}(T) the image of (T)(k1+2k2)(T)^{\otimes(k_{1}+2k_{2})} under this map.

Let ω1\omega_{1} denote the highest weight of the representation V7V_{7} of G2G_{2}, and let ω2\omega_{2} denote the highest weight of the 1414-dimensional adjoint representation 𝔤22V7{\mathfrak{g}}_{2}\subseteq\wedge^{2}V_{7}. For non-negative integers k1,k2k_{1},k_{2}, let W(k1,k2)W(k_{1},k_{2}) denote the representation of G2(𝐂)G_{2}({\mathbf{C}}) with highest weight k1ω1+k2ω2k_{1}\omega_{1}+k_{2}\omega_{2}, embedded in V7k1(2V7)k2𝐂V_{7}^{\otimes k_{1}}\otimes(\wedge^{2}V_{7})^{\otimes k_{2}}\otimes{\mathbf{C}}. This representation is the one generated by vω1k1vω2k2v_{\omega_{1}}^{\otimes k_{1}}\otimes v_{\omega_{2}}^{\otimes k_{2}} where vω1v_{\omega_{1}} and vω2v_{\omega_{2}} are highest weight vectors for V7V_{7} and 𝔤2{\mathfrak{g}}_{2} for the same Borel subgroup of G2(𝐂)G_{2}({\mathbf{C}}). Given βW(k1,k2)\beta\in W(k_{1},k_{2}), and TJT\in J, we can form the pairing

{Pk1,k2(T),β}Sk1(2W3)Sk2(W3)det(W3)k2+4,\{P_{k_{1},k_{2}}(T),\beta\}\in S^{k_{1}}(\wedge^{2}W_{3})\otimes S^{k_{2}}(W_{3})\otimes\det(W_{3})^{k_{2}+4},

where we have shifted the k2k2+4k_{2}\mapsto k_{2}+4 in the exponent of det(W3)\det(W_{3}).

Denote by RΘΘR_{\Theta}\subseteq\Theta Coxeter’s order of integral octonions, and set JRJJ_{R}\subseteq J the elements whose diagonal entries are in 𝐙{\mathbf{Z}} and off-diagonal entries are in RΘR_{\Theta}. Recall that Kim’s modular form on HJ1H_{J}^{1} has Fourier expansion

ΘKim(Z)=T0,TJR,rk(T)1a(T)e2πi(T,Z)\Theta_{Kim}(Z)=\sum_{T\geq 0,T\in J_{R},rk(T)\leq 1}{a(T)e^{2\pi i(T,Z)}}

where a(0)=1a(0)=1 and if TT is rank one then a(T)=240σ3(dT)a(T)=240\sigma_{3}(d_{T}) where dTd_{T} is the largest integer with dT1TJRd_{T}^{-1}T\in J_{R}.

Finally, set ΓG2=G2a(𝐙)\Gamma_{G_{2}}=G_{2}^{a}({\mathbf{Z}}).

Theorem 1.1.1.

Suppose βW(k1,k2)\beta\in W(k_{1},k_{2}) and α\alpha is the level one algebraic modular form on G2aG_{2}^{a} with α(1)=1|ΓG2|γΓG2γβ\alpha(1)=\frac{1}{|\Gamma_{G_{2}}|}\sum_{\gamma\in\Gamma_{G_{2}}}{\gamma\cdot\beta}. Then the theta lift Θ(α)\Theta(\alpha) of α\alpha is a vector-valued Siegel modular form of weight (k1+2k2+4,k1+k2+4,k2+4)(k_{1}+2k_{2}+4,k_{1}+k_{2}+4,k_{2}+4) with Fourier expansion

Θ(α)(Z)=T0,TJR,rk(T)1a(T){Pk1,k2(T),β}e2πi(T,Z).\Theta(\alpha)(Z)=\sum_{T\geq 0,T\in J_{R},rk(T)\leq 1}{a(T)\{P_{k_{1},k_{2}}(T),\beta\}e^{2\pi i(T,Z)}}.

When k2>0k_{2}>0 [GS98] proves that Θ(α)\Theta(\alpha) is a cusp form.

A simple restatement of Theorem 1.1.1 is as follows. Consider the projection map JH3(𝐐)J\rightarrow H_{3}({\mathbf{Q}}) given by sending

(c1a3a2a3c2a1a2a1c3)12(2c1tr(a3)tr(a2)tr(a3)2c2tr(a1)tr(a2)tr(a1)2c3).\left(\begin{array}[]{ccc}c_{1}&a_{3}&a_{2}^{*}\\ a_{3}^{*}&c_{2}&a_{1}\\ a_{2}&a_{1}^{*}&c_{3}\end{array}\right)\mapsto\frac{1}{2}\left(\begin{array}[]{ccc}2c_{1}&\operatorname{tr}(a_{3})&\operatorname{tr}(a_{2})\\ \operatorname{tr}(a_{3})&2c_{2}&\operatorname{tr}(a_{1})\\ \operatorname{tr}(a_{2})&\operatorname{tr}(a_{1})&2c_{3}\end{array}\right).

Then if T0H3(𝐐)T_{0}\in H_{3}({\mathbf{Q}}) (Hermitian 3×33\times 3 matrices with 𝐐{\mathbf{Q}} coefficients) is half-integral, the T0T_{0} Fourier coefficient of Θ(α)(Z)\Theta(\alpha)(Z) is

aΘ(α)(T0)=T0,TJR,rk(T)1,TT0a(T){Pk1,k2(T),β}.a_{\Theta(\alpha)}(T_{0})=\sum_{T\geq 0,T\in J_{R},rk(T)\leq 1,T\mapsto T_{0}}{a(T)\{P_{k_{1},k_{2}}(T),\beta\}}. (1)

(The sum is finite and explicitly determinable.) We verify that the aΘ(α)(T0)a_{\Theta(\alpha)}(T_{0}) live in the highest weight submodule

V[k1,k2]:=V(k1+2k2+4,k1+k2+4,k2+4)Sk1(2W3)Sk2(W3)det(W3)k2+4V_{[k_{1},k_{2}]}:=V_{(k_{1}+2k_{2}+4,k_{1}+k_{2}+4,k_{2}+4)}\subseteq S^{k_{1}}(\wedge^{2}W_{3})\otimes S^{k_{2}}(W_{3})\otimes\det(W_{3})^{k_{2}+4}

so that Θ(α)\Theta(\alpha) really is a vector-valued Siegel modular form for the representation V[k1,k2]V_{[k_{1},k_{2}]}.

An important computational aspect of Theorem 1.1.1 is that one can use β\beta in the pairing {Pk1,k2(T),β}\{P_{k_{1},k_{2}}(T),\beta\} instead of α(1)=1|ΓG2|γΓG2γβ\alpha(1)=\frac{1}{|\Gamma_{G_{2}}|}\sum_{\gamma\in\Gamma_{G_{2}}}{\gamma\cdot\beta}. This enables one to compute theta lifts much more quickly than if one had to use the algebraic modular form α(1)W(k1,k2)ΓG2\alpha(1)\in W(k_{1},k_{2})^{\Gamma_{G_{2}}}. Moreover, even if one does not know a priori that W(k1,k2)ΓG20W(k_{1},k_{2})^{\Gamma_{G_{2}}}\neq 0, the theorem still holds as stated. In particular, verifying that the right hand side of equation (1) is nonzero for a single T0T_{0} shows that 1|ΓG2|γΓG2γβ0\frac{1}{|\Gamma_{G_{2}}|}\sum_{\gamma\in\Gamma_{G_{2}}}{\gamma\cdot\beta}\neq 0. The reader can, of course, easily check this claim directly.

The Fourier expansion in Theorem 1.1.1 can be seen as completely analogous to the Fourier expansion of classical pluriharmonic theta functions. Indeed, the βW(k1,k2)\beta\in W(k_{1},k_{2}) becomes the pluriharmonic polynomial, and the rank one TT’s become the lattice vectors over which one sums.

When the Siegel modular form Θ(α)\Theta(\alpha) is a Hecke eigenform, it has Satake parameters cpc_{p}, one for each prime number pp, which are semisimple elements in Spin7(𝐂)\operatorname{Spin}_{7}({\mathbf{C}}). (Because we work with level one forms, we blur the distinction between Sp6\operatorname{Sp}_{6} and PGSp6\operatorname{PGSp}_{6}.) It is proved by Gross-Savin [GS98], Maagard-Savin [MS97], and Gan-Savin [GS21] that the theta lift is functorial for spherical representations; in fact, it is functorial for all representations, see Gan-Savin [GS22]. In particular, these conjugacy classes cpc_{p} are in G2(𝐂)Spin7(𝐂)G_{2}({\mathbf{C}})\subseteq\operatorname{Spin}_{7}({\mathbf{C}}). Thus Theorem 1.1.1 can produce numerous explicit examples of level one vector-valued Siegel modular forms all of whose Satake parameters are in G2(𝐂)G_{2}({\mathbf{C}}). We have taken the liberty of providing some small illustration of this, as follows. Let λ1=(12,8,8)\lambda_{1}=(12,8,8) and λ2=(14,10,8)\lambda_{2}=(14,10,8). Then it is known from Chenevier-Taibi [CT20] that the space of vector-valued, level one cuspidal Siegel modular forms of these weights are one dimensional.

Corollary 1.1.2.

For both λ1=(12,8,8)\lambda_{1}=(12,8,8) and λ2=(14,10,8)\lambda_{2}=(14,10,8), the cuspidal Siegel modular forms of these weights are lifts from G2aG_{2}^{a}. In particular, their Satake parameters all lie in G2(𝐂)G_{2}({\mathbf{C}}).

Indeed, the proof of Corollary 1.1.2 is to produce a single β1W(0,4)\beta_{1}\in W(0,4) and β2W(2,4)\beta_{2}\in W(2,4), and a single T0T_{0} so that the right hand side of (1) is nonzero for these βi\beta_{i}. It then follows that the Θ(α)\Theta(\alpha) are nonzero level one Siegel modular forms, and thus by the dimension computation of [CT20] are the unique cuspidal level one eigenforms of weights λ1\lambda_{1} and λ2\lambda_{2}. Producing many more such explicit examples would be possible. We remark that one can find conjectures here [BCFvdG17] of which small weight level one Siegel modular forms have their Satake parameters in G2(𝐂)G_{2}({\mathbf{C}}).

More than just a couple of examples, however, Theorem 1.1.1 produces an algorithm to determine if any fixed level one Siegel modular cusp form of most weights is a lift from G2aG_{2}^{a}. To setup the result, note that for a weight λ=(λ1,λ2,λ3)\lambda=(\lambda_{1},\lambda_{2},\lambda_{3}), it is known [Ibu02] that there exists explicitly determinable finite sets 𝒞λ\mathcal{C}_{\lambda} of half-integral symmetric matrices TT so that if FF is a cuspidal level one Siegel modular form of weight λ\lambda, and if aF(T)=0a_{F}(T)=0 for all T𝒞λT\in\mathcal{C}_{\lambda}, then F=0F=0.

Corollary 1.1.3.

Suppose λ=(k1+2k2+4,k1+k2+4,k2+4)\lambda=(k_{1}+2k_{2}+4,k_{1}+k_{2}+4,k_{2}+4) with k2>0k_{2}>0, and FF is a level one Siegel modular cusp form of weight λ\lambda on Sp6\operatorname{Sp}_{6}, whose Fourier coefficients aF(T)a_{F}(T) are given for all T𝒞λT\in\mathcal{C}_{\lambda}. Then there is an algorithm to determine if F=Θ(α)F=\Theta(\alpha) for some algebraic modular form α\alpha on G2aG_{2}^{a}.

Let us indicate now some of the ingredients that go into the proof of Corollary 1.1.3. First, let us clarify that the theta lifts Θ(α)\Theta(\alpha) of level one algebraic modular forms α(1)W(k1,k2)ΓG2\alpha(1)\in W(k_{1},k_{2})^{\Gamma_{G_{2}}} that appear in Theorem 1.1.1 are defined as integrals

Θ(α)(g)=G2a(𝐐)\G2a(𝐀){Θk1,k2(g,h),α(h)}𝑑h\Theta(\alpha)(g)=\int_{G_{2}^{a}({\mathbf{Q}})\backslash G_{2}^{a}({\mathbf{A}})}{\{\Theta_{k_{1},k_{2}}(g,h),\alpha(h)\}\,dh}

for a certain specific vector-valued element of Θk1,k2\Theta_{k_{1},k_{2}} in the minimal representation Πmin\Pi_{min} on HJ1H_{J}^{1}. Here, in the integral, {,}\{\,,\,\} is a G2a(𝐑)G_{2}^{a}({\mathbf{R}})-equivariant pairing valued in Sk1(V3)Sk2(2V3)S^{k_{1}}(V_{3})\otimes S^{k_{2}}(\wedge^{2}V_{3}). Let {β1,,βN}\{\beta_{1},\ldots,\beta_{N}\} be a spanning set of W(k1,k2)W(k_{1},k_{2}), and αj=1|ΓG2|γΓG2γβj\alpha_{j}=\frac{1}{|\Gamma_{G_{2}}|}\sum_{\gamma\in\Gamma_{G_{2}}}{\gamma\beta_{j}}. One can use Theorem 1.1.1 to explicitly compute the Fourier coefficients of Θ(αj)\Theta(\alpha_{j}) associated to the various TsT^{\prime}s in 𝒞λ\mathcal{C}_{\lambda}, where λ=(k1+2k2+4,k1+k2+4,k2+4)\lambda=(k_{1}+2k_{2}+4,k_{1}+k_{2}+4,k_{2}+4). Using linear algebra, one can check algorithmically if FF is some linear combination of the Θ(α)\Theta(\alpha)’s. Thus Theorem 1.1.1 gives an algorithm to determine if the Siegel modular form FF is a Θ(α)\Theta(\alpha) for some algebraic modular form α\alpha on G2aG_{2}^{a}. Thus Corollary 1.1.3 will be proved if one knew the following claim:

Claim 1.1.4.

Suppose FF is a level one cuspidal Siegel modular form of weight λ=(k1+2k2+4,k1+k2+4,k2+4)\lambda=(k_{1}+2k_{2}+4,k_{1}+k_{2}+4,k_{2}+4) with k2>0k_{2}>0. Suppose FF is in the image of the theta correspondence for G2a×Sp6HJ1G_{2}^{a}\times\operatorname{Sp}_{6}\subseteq H_{J}^{1}. Then F=Θ(α)F=\Theta(\alpha) for some algebraic modular form α(1)W(k1,k2)ΓG2\alpha(1)\in W(k_{1},k_{2})^{\Gamma_{G_{2}}}.

We prove this claim. The proof uses some powerful ingredients: The Howe Duality theorem of Gan-Savin [GS21]; an analysis of the minimal representation, as provided by Gross-Savin [GS98], Magaard-Savin [MS97] and Gan-Savin [GS21]; the existence of enough nonvanishing Fourier coefficients of FF of a certain form, as proved by Böcherer-Das [BD21]; and another argument from Gross-Savin [GS98].

Our proof of Theorem 1.1.1 is very simple. Let 𝔭J{\mathfrak{p}}_{J} be the complexification of the 1-1 eigenspace for the Cartan involution on the Lie algebra of E7,3=HJ1(𝐑)E_{7,3}=H_{J}^{1}({\mathbf{R}}). Write 𝔭J=𝔭J+𝔭J{\mathfrak{p}}_{J}={\mathfrak{p}}_{J}^{+}\oplus{\mathfrak{p}}_{J}^{-}, the natural decomposition, so that 𝔭J{\mathfrak{p}}_{J}^{-} annihilates the automorphic form on E7,3E_{7,3} associated to any holomorphic modular form for this group. Let {Xα}α\{X_{\alpha}\}_{\alpha} be a basis of 𝔭J+{\mathfrak{p}}_{J}^{+} and {Xα}α\{X_{\alpha}^{\vee}\}_{\alpha} be the dual basis of 𝔭J+,𝔭J{\mathfrak{p}}_{J}^{+,\vee}\simeq{\mathfrak{p}}_{J}^{-}. For an automorphic form φ\varphi on HJ1H_{J}^{1}, set Dφ(g)=αXαφXα.D\varphi(g)=\sum_{\alpha}{X_{\alpha}\varphi\otimes X_{\alpha}^{\vee}}. For an integer m0m\geq 0, let Dmφ=DDDφD^{m}\varphi=D\circ D\circ\cdots\circ D\varphi, so that

Dmφ=α1,,αmXαmXα1φXα1Xαm.D^{m}\varphi=\sum_{\alpha_{1},\ldots,\alpha_{m}}{X_{\alpha_{m}}\cdots X_{\alpha_{1}}\varphi\otimes X_{\alpha_{1}}^{\vee}\otimes\cdots\otimes X_{\alpha_{m}}^{\vee}}.

If (abusing notation) ΘKim\Theta_{Kim} denotes the automorphic form on HJ1H_{J}^{1} associated to Kim’s holomorphic modular form, we set Θk1,k2(g)=Dk1+2k2ΘKim(g)\Theta_{k_{1},k_{2}}(g)=D^{k_{1}+2k_{2}}\Theta_{Kim}(g).

With this definition, using Kim’s expansion of ΘKim\Theta_{Kim}, we compute the Fourier expansion of Θk1,k2(g)\Theta_{k_{1},k_{2}}(g). This is the main step in the proof of Theorem 1.1.1. We then use this to compute the Fourier expansion of Θ(α)\Theta(\alpha). One obtains (the automorphic form associated to) a holomorphic function with the Fourier expansion as given in Theorem 1.1.1. We remark that the use of differential operators as we do has some overlap with the works [Ibu99, Cle22] of Ibukiyama and Clerc.

1.2. Quaternionic theta functions

Denote by GJG_{J} the group of type E8E_{8} which is split at every finite place and E8,4E_{8,4} at the archimedean place and by G2G_{2} the split exceptional group of this Dynkin type. Analogous to the dual pair Sp6×G2aHJ1\operatorname{Sp}_{6}\times G_{2}^{a}\subseteq H_{J}^{1} is the dual pair G2×F4IGJG_{2}\times F_{4}^{I}\subseteq G_{J}, where F4IF_{4}^{I} is a specific form of F4F_{4} that is split at all finite places and compact at the archimedean place, defined as the stabilizer of the identity matrix IJI\in J. We compute the theta lifts of certain algebraic modular forms on F4IF_{4}^{I} to G2G_{2}, and obtain cuspidal quaternionic modular forms on G2G_{2} together with their exact Fourier expansions.

More precisely, let J0J^{0} denote the trace 0 elements of JJ. In other words, J0J^{0} consists of the XJX\in J with (I,X)I=0(I,X)_{I}=0, where (,)I(\,,\,)_{I} is the symmetric non-degenerate pairing on JJ determined by II. There is a surjective F4IF_{4}^{I}-equivariant map 2J0𝔣4\wedge^{2}J^{0}\rightarrow\mathfrak{f}_{4} from 2J0\wedge^{2}J^{0} to the Lie algebra 𝔨4\mathfrak{k}_{4} of F4IF_{4}^{I}. Denote by Vλ3V_{\lambda_{3}} the kernel of this map. It is an irreducible representation of F4F_{4} of dimension 273273. For an integer m>0m>0, let Vmλ3(2J0)mV_{m\lambda_{3}}\subseteq(\wedge^{2}J^{0})^{\otimes m} denote the irreducible representation of F4F_{4} with highest weight mλ3m\lambda_{3}, generated by the tensor product of a highest weight vector of Vλ3V_{\lambda_{3}}. It follows from the archimedean theta correspondence calculated in [HPS96] that algebraic modular forms on F4IF_{4}^{I} for the representation Vmλ3V_{m\lambda_{3}} should lift to quaternionic modular forms on G2G_{2} of integer weight 4+m4+m. We explicitly compute the Fourier expansion of this lift, and as a result, obtain exceptional “pluriharmonic” cuspidal quaternionic theta functions on G2G_{2}.

We setup the statement of the result. To do so, recall that the group F4IF_{4}^{I} has not one but two integral structures [Gro96], [EG96]. More precisely, if F4I(𝐙^)F_{4}^{I}(\widehat{{\mathbf{Z}}}) denotes one of these integral structures, then the double coset space F4I(𝐐)\F4I(𝐀f)/F4I(𝐙^)F_{4}^{I}({\mathbf{Q}})\backslash F_{4}^{I}({\mathbf{A}}_{f})/F_{4}^{I}(\widehat{{\mathbf{Z}}}) has size two. Because of this, algebraic modular forms for F4IF_{4}^{I} can be described as follows. Denote by MJ1M_{J}^{1} the subgroup of GL(J)\operatorname{GL}(J) fixing the cubic norm. Set ΓI\Gamma_{I} to be the subgroup of MJ1M_{J}^{1} preserving the lattice JRJ_{R} and fixing the element II. Recall the element EJRE\in J_{R} of norm one from [EG96] or [Gro96]. Let ΓE\Gamma_{E} denote the subgroup of MJ1M_{J}^{1} preserving the lattice JRJ_{R} and fixing the element EE. Fix an element δE𝐐MJ1(𝐐)\delta_{E}^{{\mathbf{Q}}}\in M_{J}^{1}({\mathbf{Q}}) with δE𝐐E=I\delta_{E}^{\mathbf{Q}}E=I. If VV is a representation of F4I(𝐑)F_{4}^{I}({\mathbf{R}}), let ΓE\Gamma_{E} act on VV via γEv=(δE𝐐vδE𝐐,1)v\gamma\cdot_{E}v=(\delta_{E}^{\mathbf{Q}}v\delta_{E}^{{\mathbf{Q}},-1})v, where the conjugation takes place in MJ1M_{J}^{1}. Then a level one algebraic modular form for F4IF_{4}^{I} can be considered as a pair

(αI,αE)VΓIVΓE.(\alpha_{I},\alpha^{\prime}_{E})\in V^{\Gamma_{I}}\oplus V^{\Gamma_{E}}.

In the case of V=Vmλ3V=V_{m\lambda_{3}}, we rephrase this as follows. Let (,)E(\,,\,)_{E} be the symmetric non-degenerate pairing on JJ determined by EE; one has

(u,v)E=14(E,E,u)(E,E,v)(E,u,v)(u,v)_{E}=\frac{1}{4}(E,E,u)(E,E,v)-(E,u,v)

where (,,)(\cdot,\cdot,\cdot) is the symmetric trilinear form on JJ satisfying (x,x,x)=6n(x)(x,x,x)=6n(x), 66 times the cubic norm on JJ. Set JE0J_{E}^{0} to be the perpendicular space to EE under the pairing (,)E(\,,\,)_{E}. One easily verifies that δE𝐐,1J0=JE0\delta_{E}^{{\mathbf{Q}},-1}J^{0}=J_{E}^{0}. Thus

αE:=δE𝐐,1αE(2JE0)m\alpha_{E}:=\delta_{E}^{{\mathbf{Q}},-1}\alpha^{\prime}_{E}\subseteq(\wedge^{2}J_{E}^{0})^{\otimes m}

and is ΓE\Gamma_{E} invariant for the natural action of ΓE\Gamma_{E}. It will be convenient for us to consider the algebraic modular form to be the pair (αI,αE)(\alpha_{I},\alpha_{E}).

Now, for x,y,bJx,y,b\in J and cJc\in J^{\vee}, write

xy,bcI=(x,b)I(y,c)(x,c)(y,b)I.\langle x\wedge y,b\wedge c\rangle_{I}=(x,b)_{I}(y,c)-(x,c)(y,b)_{I}.

We use the same notation for the pairing (2J)m(2J)m𝐂(\wedge^{2}J)^{\otimes m}\otimes(\wedge^{2}J)^{\otimes m}\rightarrow{\mathbf{C}} that extends this one via z1zm,z1zmI=j=1mzj,zjI\langle z_{1}\otimes\cdots\otimes z_{m},z_{1}^{\prime}\otimes\cdots\otimes z_{m}^{\prime}\rangle_{I}=\prod_{j=1}^{m}\langle z_{j},z_{j}^{\prime}\rangle_{I}. Thus if βVmλ3\beta\in V_{m\lambda_{3}}, bJb\in J and cJc\in J^{\vee}, we can compute the quantity β,(bc)m𝐂\langle\beta,(b\wedge c)^{\otimes m}\rangle\in{\mathbf{C}}. We similarly define ,E\langle\,,\,\rangle_{E}, by replacing the pairing (,)I(\,,\,)_{I} with (,)E(\,,\,)_{E}.

Set WJR=𝐙JRJR𝐙W_{J_{R}}={\mathbf{Z}}\oplus J_{R}\oplus J_{R}^{\vee}\oplus{\mathbf{Z}}. For wWJRw\in W_{J_{R}} of rank one, set a(w)=σ4(dw)a(w)=\sigma_{4}(d_{w}) where dwd_{w} is the largest integer with dw1wWJRd_{w}^{-1}w\in W_{J_{R}}. In [Pol20b], we proved that the minimal modular form ΘGan\Theta_{Gan} on GJG_{J} has Fourier expansion

ΘGan,Z(g)=ΘGan,N(g)+wWJR,rk(w)=1a(w)W2πw(g).\Theta_{Gan,Z}(g)=\Theta_{Gan,N}(g)+\sum_{w\in W_{J_{R}},rk(w)=1}{a(w)W_{2\pi w}(g)}.

Here NZ1N\supseteq Z\supseteq 1 is the unipotent radical of the Heisenberg parabolic of E8,4E_{8,4}, ΘGan,Z\Theta_{Gan,Z}, ΘGan,N\Theta_{Gan,N} denote constant terms, and W2πwW_{2\pi w} is the generalized Whittaker function of [Pol20a].

Now, if wWJw\in W_{J} and m>0m>0, set Pm(w)=(bc)m(2J)mP_{m}(w)=(b\wedge c)^{\otimes m}\in(\wedge^{2}J)^{\otimes m} if w=(a,b,c,d)w=(a,b,c,d). Moreover, with this notation, let pI(w)p_{I}(w) and pE(w)p_{E}(w) be the binary cubic forms given as

pI(w)(u,v)=au3+(b,I#)u2v+(c,I)uv2+dv3;pE(w)(u,v)=au3+(b,E#)u2v+(c,E)uv2+dv3.p_{I}(w)(u,v)=au^{3}+(b,I^{\#})u^{2}v+(c,I)uv^{2}+dv^{3};\,\,\,p_{E}(w)(u,v)=au^{3}+(b,E^{\#})u^{2}v+(c,E)uv^{2}+dv^{3}.

We prove the following.

Theorem 1.2.1.

Suppose α\alpha is a level one algebraic modular form on F4IF_{4}^{I} for the representation Vmλ3V_{m\lambda_{3}} with m>0m>0. Represent α\alpha as a pair (αI,αE)(\alpha_{I},\alpha_{E}) with αI(2J)m\alpha_{I}\in(\wedge^{2}J)^{\otimes m} being ΓI\Gamma_{I} invariant and αE(2J)m\alpha_{E}\in(\wedge^{2}J)^{\otimes m} being ΓE\Gamma_{E} invariant. Let βI,βE\beta_{I},\beta_{E} be in Vmλ3V_{m\lambda_{3}}, respectively δE𝐐,1Vmλ3\delta_{E}^{{\mathbf{Q}},-1}V_{m\lambda_{3}} so that αI=1|ΓI|γΓIγβI\alpha_{I}=\frac{1}{|\Gamma_{I}|}\sum_{\gamma\in\Gamma_{I}}{\gamma\beta_{I}} and αE=1|ΓE|γΓEγβE.\alpha_{E}=\frac{1}{|\Gamma_{E}|}\sum_{\gamma\in\Gamma_{E}}{\gamma\beta_{E}}. Then the theta lift Θ(α)\Theta(\alpha) is a cuspidal, level one, quaternionic modular form on G2G_{2} of weight 4+m4+m with Fourier expansion

Θ(α)Z(g)\displaystyle\Theta(\alpha)_{Z}(g) =1|ΓI|wWJR,rk(w)=1a(w)Pm(w),βIIW2πprI(w)(g)\displaystyle=\frac{1}{|\Gamma_{I}|}\sum_{w\in W_{J_{R}},rk(w)=1}{a(w)\langle P_{m}(w),\beta_{I}\rangle_{I}W_{2\pi pr_{I}(w)}(g)}
+1|ΓE|wWJR,rk(w)=1a(w)Pm(w),βEEW2πprE(w)(g).\displaystyle\,\,+\frac{1}{|\Gamma_{E}|}\sum_{w\in W_{J_{R}},rk(w)=1}{a(w)\langle P_{m}(w),\beta_{E}\rangle_{E}W_{2\pi pr_{E}(w)}(g)}.

A simple restatement of Theorem 1.2.1 is as follows. If w0w_{0} is an integral binary cubic form, then the w0w_{0} Fourier coefficient of Θ(α)\Theta(\alpha) is

aΘ(α)(w0)=1|ΓI|wWJR,prw,I(w)=w0a(w)Pm(w),βII+1|ΓE|wWJR,prw,E(w)=w0a(w)Pm(w),βEE.a_{\Theta(\alpha)}(w_{0})=\frac{1}{|\Gamma_{I}|}\sum_{w\in W_{J_{R}},pr_{w,I}(w)=w_{0}}{a(w)\langle P_{m}(w),\beta_{I}\rangle_{I}}+\frac{1}{|\Gamma_{E}|}\sum_{w\in W_{J_{R}},pr_{w,E}(w)=w_{0}}{a(w)\langle P_{m}(w),\beta_{E}\rangle_{E}}.

These sums are finite.

The reason we do not state Theorem 1.2.1 in the case m=0m=0 is because then the theta lifts will be non-cuspidal. In fact, the theta lifts obtained for m=0m=0 are exactly the automorphic forms obtained in [GGS02, Section 10]. Thus Theorem 1.2.1 may be considered a generalization of [GGS02, Section 10].

Again, just like Theorem 1.1.1, the Fourier expansion given by Theorem 1.2.1 is completely parallel to the classical pluriharmonic theta functions: The βs\beta^{\prime}s in Vmλ3V_{m\lambda_{3}} are the pluriharmonic polynomial, and the sum over wWJRw\in W_{J_{R}} of rank one is the sum over lattice vectors.

We now state a few corollaries of Theorem 1.2.1. For the first corollary, we can partially refine Theorem 1.0.1 in the case of level one.

Corollary 1.2.2.

There is a lattice LmIVmλ3L_{m}^{I}\subseteq V_{m\lambda_{3}} and a lattice LmEδE𝐐,1Vmλ3L_{m}^{E}\subseteq\delta_{E}^{{\mathbf{Q}},-1}V_{m\lambda_{3}} so that the level one theta lifts of elements of these lattices to G2G_{2} have Fourier coefficients that are integers when evaluated at gf=1g_{f}=1.

For the second corollary, recall that it was proved in [cDD+22] that if π\pi is a cuspidal automorphic representation on G2(𝐀)G_{2}({\mathbf{A}}) that corresponds to a level one quaternionic modular form φπ\varphi_{\pi} of even weight \ell, then the completed standard LL-function Λ(π,Std,s)\Lambda(\pi,Std,s) satisfies the exact functional equation Λ(π,Std,s)=Λ(π,Std,1s)\Lambda(\pi,Std,s)=\Lambda(\pi,Std,1-s), so long as the w0=u2vuv2w_{0}=u^{2}v-uv^{2} Fourier coefficient of φπ\varphi_{\pi} is nonzero. At the time of the writing of [cDD+22], it was not known whether such π\pi exist. Using Theorem 1.2.1, one easily obtains the following.

Corollary 1.2.3.

Suppose 6\ell\geq 6 is even. Then there is a cuspidal automorphic representation π\pi on G2(𝐀)G_{2}({\mathbf{A}}) that corresponds to a level one quaternionic modular form φπ\varphi_{\pi} of weight \ell with nonzero w0=u2vuv2w_{0}=u^{2}v-uv^{2} Fourier coefficient.

This corollary is, in fact, an ingredient in the proof of Theorem 1.0.1.

The third corollary we state has to do with the Fourier expansion of a particular cuspidal quaternionic modular form. To setup this corollary, recall that Dalal [Dal21] has recently given an explicit formula for the dimension of the level one quaternionic cuspidal modular forms of weights at least 33. From his dimension formula, one has that the first such nonzero cusp form appears in weight 66, and the space of weight 66 cuspidal quaternionic modular forms is one-dimensional, spanned by a cusp form ΔG2\Delta_{G_{2}}. Combining the (proof of) Corollary 1.2.3 with Corollary 1.2.2, one obtains that ΔG2\Delta_{G_{2}} can be normalized to have integer Fourier coefficients, and that the Fourier coefficient associated to the cubic ring 𝐙×𝐙×𝐙{\mathbf{Z}}\times{\mathbf{Z}}\times{\mathbf{Z}} is nonzero.

Now, Benedict Gross has suggested that the Fourier coefficients of certain non-tempered cuspidal quaternionic modular forms on G2G_{2} of weight \ell should be related to square-roots of twists of LL-values of classical modular forms of weight 22\ell by Artin motives associated to totally real cubic fields. As pointed out to the author by Mundy, the quaternionic modular form ΔG2\Delta_{G_{2}} should be one of these non-tempered lifts, to which Gross’s conjecture applies. Now, for DD congruent to 0 or 11 modulo 44, denote by 𝐙D{\mathbf{Z}}_{D} the quadratic ring of discriminant DD. Then on the one hand, in the case of ΔG2\Delta_{G_{2}}, Gross’s conjecture implies that the square Fourier coefficients aΔG2(𝐙×𝐙D)2a_{\Delta_{G_{2}}}({\mathbf{Z}}\times{\mathbf{Z}}_{D})^{2} associated to the cubic ring 𝐙×𝐙D{\mathbf{Z}}\times{\mathbf{Z}}_{D} should be related to the central333We here use the classical normalization of LL-functions, instead of the automorphic normalization. LL-value L(Δ,D,6)L(\Delta,D,6) of the twist of Ramanujan’s Δ\Delta function by the quadratic character associated to DD. Denote by δ(z)S13/2(Γ0(4))+\delta(z)\in S_{13/2}(\Gamma_{0}(4))^{+}, δ(z)=D0,1(mod4)α(D)qD\delta(z)=\sum_{D\equiv 0,1\pmod{4}}{\alpha(D)q^{D}} the Shimura lift of Δ(z)\Delta(z). On the other hand, following Waldspurger [Wal81], Kohnen-Zagier [KZ81] relate the squares of Fourier coefficients α(D)\alpha(D) to the same LL-value, L(Δ,D,6)L(\Delta,D,6). It thus makes sense to ask, in light of Gross’s conjecture, if there is some relationship between aΔG2(𝐙×𝐙D)a_{\Delta_{G_{2}}}({\mathbf{Z}}\times{\mathbf{Z}}_{D}) and α(D)\alpha(D). It turns out, the numbers are equal:

Corollary 1.2.4.

Normalize ΔG2\Delta_{G_{2}} so that aΔG2(𝐙×𝐙×𝐙)=1a_{\Delta_{G_{2}}}({\mathbf{Z}}\times{\mathbf{Z}}\times{\mathbf{Z}})=1, and normalize δ(z)\delta(z) so that α(1)=1\alpha(1)=1. Then aΔG2(𝐙×𝐙D)=α(D)a_{\Delta_{G_{2}}}({\mathbf{Z}}\times{\mathbf{Z}}_{D})=\alpha(D) for all DD.

1.3. Acknowledgements

It is pleasure to thank Wee Teck Gan, Nadya Gurevich, and Gordan Savin for engaging with the author in a “Research in Teams” project in Spring 2022 at the Erwin Schrodinger Institute, which helped to stimulate thinking about exceptional theta correspondences. We thank them for fruitful discussions. We thank Tomoyoshi Ibukiyama for sending us his note [Ibu02], and we also thank Gaetan Chenevier, Chao Li, Finley McGlade, Sam Mundy, Cris Poor and David Yuen for helpful discussions.

2. Groups and embeddings

In this section, we explain various group theoretic facts regarding the groups with which we work.

2.1. Some exceptional groups

We begin by defining the group HJ1H_{J}^{1}. Thus let JJ be our exceptional cubic norm structure, WJ=𝐐JJ𝐐W_{J}={\mathbf{Q}}\oplus J\oplus J^{\vee}\oplus{\mathbf{Q}}. A typical element of WJW_{J} we write as an ordered four-tuple (a,b,c,d)(a,b,c,d), so that a,d𝐐a,d\in{\mathbf{Q}}, bJb\in J and cJc\in J^{\vee}. We put on WJW_{J} Freudenthal’s symplectic form, and quartic form. The group HJ1H_{J}^{1} is defined as the algebraic 𝐐{\mathbf{Q}}-group preserving these two forms. We let HJH_{J} be the group that preserves the forms on WJW_{J} up to similitude, and let ν:HJGL1\nu:H_{J}\rightarrow\operatorname{GL}_{1} be this similitude.

The Siegel parabolic PJP_{J} of HJ1H_{J}^{1} is defined as the stabilizer of the line 𝐐(0,0,0,1){\mathbf{Q}}(0,0,0,1). Write MJM_{J} for the group of linear automorphisms of JJ that preserve the norm on JJ, up to scaling. Let λ:MJGL1\lambda:M_{J}\rightarrow\operatorname{GL}_{1} be this scaling factor, and MJ1M_{J}^{1} the kernel of λ\lambda. A Levi subgroup of PJP_{J} can be defined as the subgroup that also stabilizes the line 𝐐(1,0,0,0){\mathbf{Q}}(1,0,0,0). This is isomorphic to the group of pairs (δ,m)GL1×MJ(\delta,m)\in\operatorname{GL}_{1}\times M_{J} with δ2=λ(m)\delta^{2}=\lambda(m). Such a pair acts on WJW_{J} as (a,b,c,d)(δ1a,δ1m(b),δm~(c),δd)(a,b,c,d)\mapsto(\delta^{-1}a,\delta^{-1}m(b),\delta\widetilde{m}(c),\delta d). We write M(δ,m)M(\delta,m) for this group element of HJ1H_{J}^{1}.

2.2. The first dual pair

We now explain how GSp6×G2a\operatorname{GSp}_{6}\times G_{2}^{a} embeds in HJH_{J}. To accomplish this, we describe a linear isomorphism between 03W6ν1W6Θ0\wedge^{3}_{0}W_{6}\otimes\nu^{-1}\oplus W_{6}\otimes\Theta^{0} and WJW_{J}. Here Θ0=V7\Theta^{0}=V_{7} is the trace 0 octonions, W6W_{6} is the 66-dimensional defining representation of GSp6\operatorname{GSp}_{6}, and 03W6\wedge^{3}_{0}W_{6} is the kernel of the contraction map 3W6W6ν\wedge^{3}W_{6}\rightarrow W_{6}\otimes\nu. We let e1,e2,e3,f1,f2,f3e_{1},e_{2},e_{3},f_{1},f_{2},f_{3} be the standard symplectic basis of W6W_{6}.

  • If v1,v2,v3,v1,v2,v3Θ0v_{1},v_{2},v_{3},v_{1}^{\prime},v_{2}^{\prime},v_{3}^{\prime}\in\Theta^{0}, we map v1e1+v2e2+v3e3+v1f1+v2f2+v3f3v_{1}^{\prime}e_{1}+v_{2}^{\prime}e_{2}+v_{3}^{\prime}e_{3}+v_{1}f_{1}+v_{2}f_{2}+v_{3}f_{3} to (0,Y,Y,0)(0,Y,Y^{\prime},0) where Y=(0v3v2v30v1v2v10)JY=\left(\begin{array}[]{ccc}0&v_{3}&-v_{2}\\ -v_{3}&0&v_{1}\\ v_{2}&-v_{1}&0\end{array}\right)\in J and Y=(0v3v2v30v1v2v10)Y^{\prime}=-\left(\begin{array}[]{ccc}0&v_{3}^{\prime}&-v_{2}^{\prime}\\ -v_{3}^{\prime}&0&v_{1}^{\prime}\\ v_{2}^{\prime}&-v_{1}^{\prime}&0\end{array}\right).

  • We map f1f2f3f_{1}\wedge f_{2}\wedge f_{3} to (1,0,0,0)(1,0,0,0) and e1e2e3e_{1}\wedge e_{2}\wedge e_{3} to (0,0,0,1)(0,0,0,1).

  • Set fi=fi+1fi1f_{i}^{*}=f_{i+1}\wedge f_{i-1} and ei=ei+1ei1e_{i}^{*}=e_{i+1}\wedge e_{i-1} (with indices taken modulo 33). We map i,jbijfiej\sum_{i,j}{b_{ij}f_{i}^{*}\wedge e_{j}} to (0,(bij),0,0)(0,(b_{ij}),0,0) and i,jcijeifj\sum_{i,j}{c_{ij}e_{i}^{*}\wedge f_{j}} to (0,0,(cij),0)(0,0,(c_{ij}),0).

Via the natural action of GSp6×G2a\operatorname{GSp}_{6}\times G_{2}^{a} on 03W6ν1W6Θ0\wedge^{3}_{0}W_{6}\otimes\nu^{-1}\oplus W_{6}\otimes\Theta^{0}, we obtain an action of this group on WJW_{J}. It is clear that the action is faithful. To obtain the embedding into HJH_{J}, we must check that GSp6×G2a\operatorname{GSp}_{6}\times G_{2}^{a} preserves the symplectic and quartic form on WJW_{J}, up to scaling:

Proposition 2.2.1.

The defined action of GSp6×G2a\operatorname{GSp}_{6}\times G_{2}^{a} on WJW_{J} gives an embedding GSp6×G2aHJ\operatorname{GSp}_{6}\times G_{2}^{a}\subseteq H_{J}.

To prove the proposition, we first make a few lemmas. We work over a general field FF of characteristic 0.

Lemma 2.2.2.

Suppose X=(s1u1u2u3s2u1u2u1s3)X=\left(\begin{array}[]{ccc}s_{1}&u_{1}&u_{2}\\ u_{3}&s_{2}&u_{1}\\ u_{2}&u_{1}&s_{3}\end{array}\right) is in H3(F)H_{3}(F) and vjΘ0v_{j}\in\Theta^{0} so that the element Y=(0v3v2v30v1v2v10)Y=\left(\begin{array}[]{ccc}0&v_{3}&-v_{2}\\ -v_{3}&0&v_{1}\\ v_{2}&-v_{1}&0\end{array}\right) is in JJ. Then X×Y=(0v3v2v30v1v2v10)X\times Y=-\left(\begin{array}[]{ccc}0&v_{3}^{\prime}&-v_{2}^{\prime}\\ -v_{3}^{\prime}&0&v_{1}^{\prime}\\ v_{2}^{\prime}&-v_{1}^{\prime}&0\end{array}\right) with

  1. (1)

    v1=s1v1+u3v2+u2v3v_{1}^{\prime}=s_{1}v_{1}+u_{3}v_{2}+u_{2}v_{3}

  2. (2)

    v2=u3v1+s2v2+u1v3v_{2}^{\prime}=u_{3}v_{1}+s_{2}v_{2}+u_{1}v_{3}

  3. (3)

    v3=u2v1+u1v2+s3v3v_{3}^{\prime}=u_{2}v_{1}+u_{1}v_{2}+s_{3}v_{3}.

Proof.

This is direct computation.∎

Let now X=(Xij)H3(F)X=(X_{ij})\in H_{3}(F) and consider the element nL,Sp6(X)=(0X00)n_{L,\operatorname{Sp}_{6}}(X)=\left(\begin{smallmatrix}0&X\\ 0&0\end{smallmatrix}\right) in the Lie algebra of Sp6\operatorname{Sp}_{6}. Recall the element nL(X)n_{L}(X), in the Lie algebra of HJ1H_{J}^{1}, that acts on (a,b,c,d)WJ(a,b,c,d)\in W_{J} as nL(X)(a,b,c,d)=(0,aX,b×X,(c,X))n_{L}(X)(a,b,c,d)=(0,aX,b\times X,(c,X)).

Lemma 2.2.3.

Under the above identification 03W6ν1W6Θ0WJ\wedge^{3}_{0}W_{6}\otimes\nu^{-1}\oplus W_{6}\otimes\Theta^{0}\simeq W_{J}, the operator nL,Sp6(X)n_{L,\operatorname{Sp}_{6}}(X) acts as nL(X)n_{L}(X).

Proof.

We first compute how nL,Sp6(X)n_{L,\operatorname{Sp}_{6}}(X) acts on (0,(bij),0,0)=ijbijfiej(0,(b_{ij}),0,0)=\sum_{ij}{b_{ij}f_{i}^{*}\wedge e_{j}}. Writing out the action of nL,Sp6(X)n_{L,\operatorname{Sp}_{6}}(X) on this element, we obtain

ijbij(mXm,i1fi+1emej+X,i+1efi1ej).\sum_{ij}{b_{ij}(\sum_{m}X_{m,i-1}f_{i+1}\wedge e_{m}\wedge e_{j}+\sum_{\ell}X_{\ell,i+1}e_{\ell}\wedge f_{i-1}\wedge e_{j})}.

The coefficient of e1f1=e2e3f1e_{1}^{*}\wedge f_{1}=e_{2}\wedge e_{3}\wedge f_{1} comes from 44 terms:

  • i=3,m=2,j=3i=3,m=2,j=3: b33X22b_{33}X_{22}

  • i=3,m=3,j=2i=3,m=3,j=2: b32X32-b_{32}X_{32}

  • i=2,=3,j=2i=2,\ell=3,j=2: b22X33b_{22}X_{33}

  • i=2,=2,j=3i=2,\ell=2,j=3: b23X23-b_{23}X_{23}

The coefficient of e2f3=e3e1f3e_{2}^{*}\wedge f_{3}=e_{3}\wedge e_{1}\wedge f_{3} again comes from 44 terms:

  • i=2,m=3,j=1i=2,m=3,j=1: b21X31b_{21}X_{31}

  • i=2,m=1,j=3i=2,m=1,j=3: b23X11-b_{23}X_{11}

  • i=1,=3,j=1i=1,\ell=3,j=1: b11X32-b_{11}X_{32}

  • i=1,=1,j=3i=1,\ell=1,j=3: b13X12b_{13}X_{12}.

Putting the above computations together, one obtains nL,Sp6(X)(0,b,0,0)=(0,0,c,0)n_{L,\operatorname{Sp}_{6}}(X)(0,b,0,0)=(0,0,c,0) where c=b×Xc=b\times X.

Finally, one immediately obtains that nL,Sp6(X)(0,0,c,0)=(0,0,0,(c,X))n_{L,\operatorname{Sp}_{6}}(X)(0,0,c,0)=(0,0,0,(c,X)) and nL,Sp6(X)(1,0,0,0)=(0,X,0,0)n_{L,\operatorname{Sp}_{6}}(X)(1,0,0,0)=(0,X,0,0). The lemma follows. ∎

The following lemma is immediate.

Lemma 2.2.4.

The element (0110)\left(\begin{smallmatrix}0&-1\\ 1&0\end{smallmatrix}\right) of Sp6\operatorname{Sp}_{6} acts on WJW_{J} as (a,b,c,d)(d,c,b,a)(a,b,c,d)\mapsto(d,-c,b,-a). The element (ν1)\left(\begin{smallmatrix}\nu&\\ &1\end{smallmatrix}\right) acts on (a,b,c,d)(a,b,c,d) as (ν1a,b,νc,ν2d)(\nu^{-1}a,b,\nu c,\nu^{2}d).

Lemma 2.2.5.

Suppose (mn)Sp6\left(\begin{smallmatrix}m&\\ &n\end{smallmatrix}\right)\in\operatorname{Sp}_{6}, so that n=tm1n=\,^{t}m^{-1}. The action of this element on WJW_{J} is in MJ1M_{J}^{1}; in particular, it preserves the symplectic and quartic form on WJW_{J}.

Proof.

We compute the action of (mn)\left(\begin{smallmatrix}m&\\ &n\end{smallmatrix}\right) on (a,b,c,d)(a,b,c,d) when n=tm1n=\,^{t}m^{-1}. First, one computes that nn acts on fif_{i}^{*} taking fif_{i}^{*} to k(nk+1,i+1nk1,i1nk1,i+1nk+1,i1)fk=kc(n)kifk\sum_{k}{(n_{k+1,i+1}n_{k-1,i-1}-n_{k-1,i+1}n_{k+1,i-1})f_{k}^{*}}=\sum_{k}{c(n)_{ki}f_{k}^{*}}, where c(n)=det(n)tn1c(n)=\det(n)\,^{t}n^{-1} is the cofactor matrix of nn. This gives that the 03W6ν1\wedge^{3}_{0}W_{6}\otimes\nu^{-1} part of bb maps to c(n)btm=det(n)tn1bn1c(n)b\,^{t}m=\det(n)\,^{t}n^{-1}bn^{-1}. The vector (i.e., Θ0\Theta^{0}) part of bb moves as jvjfjk(jnkjvj)fk\sum_{j}{v_{j}f_{j}}\mapsto\sum_{k}{(\sum_{j}{n_{kj}v_{j}})f_{k}}.

Completely similarly (it is the same calculation), the 03W6ν1\wedge^{3}_{0}W_{6}\otimes\nu^{-1} part of cc maps to c(m)ctn=det(m)tm1cm1c(m)c\,^{t}n=\det(m)\,^{t}m^{-1}cm^{-1}. The vector part of cc moves as jvjejk(jmkjvj)ek\sum_{j}{v_{j}^{\prime}e_{j}}\mapsto\sum_{k}{(\sum_{j}{m_{kj}v_{j}^{\prime}})e_{k}}.

To finish the proof of the lemma, one must verify that the map JH3(F)Θ0V3JJ\simeq H_{3}(F)\oplus\Theta^{0}\otimes V_{3}\rightarrow J given by

(X,v)(tn1Xn1,det(n)1nv)(X,v)\mapsto(\,^{t}n^{-1}Xn^{-1},\det(n)^{-1}nv)

scales the norm on JJ by det(n)2\det(n)^{-2}. This can be done, for example, by repeatedly applying the Cayley-Dickson construction and using the formulas of [Pol18, Section 8.1]. ∎

We can now prove Proposition 2.2.1.

Proof of Proposition 2.2.1.

Exponentiating the action of nL,Sp6(X)n_{L,\operatorname{Sp}_{6}}(X), we see that the action of (1X1)Sp6\left(\begin{smallmatrix}1&X\\ &1\end{smallmatrix}\right)\in\operatorname{Sp}_{6} lands in HJ1H_{J}^{1}. The group GSp6\operatorname{GSp}_{6} is generated by these elements, together with the (ν1)\left(\begin{smallmatrix}\nu&\\ &1\end{smallmatrix}\right) and with (0110)\left(\begin{smallmatrix}0&-1\\ 1&0\end{smallmatrix}\right). The proposition thus follows from the above lemmas. ∎

2.3. Action of the maximal compact

Let KSp6K_{\operatorname{Sp}_{6}} denote the standard maximal compact subgroup of Sp6(𝐑)\operatorname{Sp}_{6}({\mathbf{R}}), so that

KSp6={(ABBA):A+iBU(3)}.K_{\operatorname{Sp}_{6}}=\left\{\left(\begin{array}[]{cc}A&-B\\ B&A\end{array}\right):A+iB\in U(3)\right\}.

Let KE7K_{E_{7}} denote the subgroup of HJ1(𝐑)H_{J}^{1}({\mathbf{R}}) that fixes the line spanned by (1,i,1,i)WJ(𝐂)(1,i,-1,-i)\in W_{J}({\mathbf{C}}). This is a maximal compact subgoup of HJ1(𝐑)H_{J}^{1}({\mathbf{R}}). Let 𝔭J{\mathfrak{p}}_{J} be the complexification of the 1-1 part for the Cartan involution on the Lie algebra of HJ1(𝐑)H_{J}^{1}({\mathbf{R}}) for this choice of maximal compact. Write 𝔭J±{\mathfrak{p}}_{J}^{\pm} for its two KE7K_{E_{7}} factors. We now work out how KSp6K_{\operatorname{Sp}_{6}} acts on 𝔭J±{\mathfrak{p}}_{J}^{\pm}.

We begin by discussing the Cayley transform ChHJ1(𝐂)C_{h}\in H_{J}^{1}({\mathbf{C}}) [Pol20a, section 5]. To do, we review some notation from [Pol20a]:

  • nG(X):=exp(nL(X))n_{G}(X):=\exp(n_{L}(X));

  • nG(γ):=exp(nL(γ))n_{G}^{\vee}(\gamma):=\exp(n_{L}^{\vee}(\gamma)), with, for γJ\gamma\in J^{\vee}, nL(γ)(a,b,c,d)=((b,γ),c×γ,dγ,0)n_{L}^{\vee}(\gamma)(a,b,c,d)=((b,\gamma),c\times\gamma,d\gamma,0);

  • for λGL1\lambda\in\operatorname{GL}_{1}, η(λ)(a,b,c,d)=(λ3a,λb,λ1c,λ3d)\eta(\lambda)(a,b,c,d)=(\lambda^{3}a,\lambda b,\lambda^{-1}c,\lambda^{-3}d);

  • 𝔪J{\mathfrak{m}}_{J} is the Lie algebra of MJM_{J}, which acts on WJW_{J} as given in [Pol20a, section 3.4].

The Cayley transform is defined as Ch=nG(i)nG(i/2)η(21/2)C_{h}=n_{G}(-i)n_{G}^{\vee}(-i/2)\eta(2^{-1/2}), so that Ch1=η(21/2)nG(i/2)nG(i)C_{h}^{-1}=\eta(2^{1/2})n_{G}^{\vee}(i/2)n_{G}(i). It satisfies:

  1. (1)

    Ch1nL(J𝐂)Ch=𝔭J+C_{h}^{-1}n_{L}(J\otimes{\mathbf{C}})C_{h}={\mathfrak{p}}_{J}^{+}

  2. (2)

    Ch1nL(J𝐂)Ch=𝔭JC_{h}^{-1}n_{L}^{\vee}(J\otimes{\mathbf{C}})C_{h}={\mathfrak{p}}_{J}^{-}

  3. (3)

    Ch1(𝔪J𝐂)Ch=𝔨E7C_{h}^{-1}({\mathfrak{m}}_{J}\otimes{\mathbf{C}})C_{h}={\mathfrak{k}}_{E_{7}}.

For ZJ𝐂Z\in J\otimes{\mathbf{C}}, we set r1(Z)WJ𝐂r_{1}(Z)\in W_{J}\otimes{\mathbf{C}} as

r1(Z)=(1,Z,Z#,n(Z))=r0(Z).r_{1}(Z)=(1,Z,Z^{\#},n(Z))=r_{0}(-Z).

For gHJ(𝐑)g\in H_{J}({\mathbf{R}}) and ZJZ\in\mathcal{H}_{J}, the factor of automorphy j(g,Z)𝐂×j(g,Z)\in{\mathbf{C}}^{\times} and the action of gg on J\mathcal{H}_{J} is defined as gr1(Z)=j(g,Z)r1(gZ)gr_{1}(Z)=j(g,Z)r_{1}(g\cdot Z).

Recall that we let M(δ,m)M(\delta,m), for mMJm\in M_{J} and δGL1\delta\in\operatorname{GL}_{1} such that δ2=λ(m)\delta^{2}=\lambda(m), act on WJW_{J} as

M(δ,m)(a,b,c,d)=(δ1a,δ1m(b),δm~(c),δ(d)).M(\delta,m)(a,b,c,d)=(\delta^{-1}a,\delta^{-1}m(b),\delta\widetilde{m}(c),\delta(d)).

For YJ>0Y\in J_{>0} set MY=M(n(Y)1/2,UY1/2)M_{Y}=M(n(Y)^{1/2},U_{Y^{1/2}}). Here Y1/2Y^{1/2} is the positive definite square root of YY, and for xJx\in J, Ux:JJU_{x}:J\rightarrow J is the map defined as Ux(z)=x#×z+(x,z)xU_{x}(z)=-x^{\#}\times z+(x,z)x. For Z=X+iYJZ=X+iY\in\mathcal{H}_{J} set gZ=nG(X)MYg_{Z}=n_{G}(X)M_{Y}. Then gZr1(i)=n(Y)1/2r1(Z)g_{Z}r_{1}(i)=n(Y)^{-1/2}r_{1}(Z).

Lemma 2.3.1.

One has

  1. (1)

    Ch1(1,0,0,0)=122r1(i)C_{h}^{-1}(1,0,0,0)=\frac{1}{2\sqrt{2}}r_{1}(i)

  2. (2)

    Ch1(0,0,0,1)=122r1(i)C_{h}^{-1}(0,0,0,1)=\frac{1}{2\sqrt{2}}r_{1}(-i).

Consequently, 𝔨=Ch1𝔪JCh{\mathfrak{k}}=C_{h}^{-1}{\mathfrak{m}}_{J}C_{h}. Moreover,

Ch1(0,z,0,0)=122(i(1,z),2z(1,z)1,i(2z+(1,z)1),(1,z))C_{h}^{-1}(0,z,0,0)=\frac{1}{2\sqrt{2}}(i(1,z),2z-(1,z)1,i(-2z+(1,z)1),(-1,z))

and

Ch1(0,0,E,0)=122((1,E),i((1,E)2E),2E(1,E),i(1,E)).C_{h}^{-1}(0,0,E,0)=\frac{1}{2\sqrt{2}}(-(1,E),i((1,E)-2E),2E-(1,E),i(1,E)).
Proof.

The first parts are direct verifications. For the second, recall that KK is the subgroup of HJ1(𝐑)H_{J}^{1}({\mathbf{R}}) that stabilizes the lines 𝐂r1(i){\mathbf{C}}r_{1}(i) and 𝐂r1(i){\mathbf{C}}r_{1}(-i), while MJM_{J} is the subgroup of HJ1(𝐑)H_{J}^{1}({\mathbf{R}}) that stabilizes the lines spanned by (1,0,0,0)(1,0,0,0) and (0,0,0,1)(0,0,0,1). The last part is again a direct verification. ∎

Here is the statement of the result.

Proposition 2.3.2.

Suppose k=(ABBA)KSp6k=\left(\begin{smallmatrix}A&-B\\ B&A\end{smallmatrix}\right)\in K_{\operatorname{Sp}_{6}}, so that A+iBU(3)A+iB\in U(3). Let EJE\in J, with EE perpendicular to H3(𝐑)H_{3}({\mathbf{R}}), so that E=v1x1+v2x2+v3x3E=v_{1}\otimes x_{1}+v_{2}\otimes x_{2}+v_{3}\otimes x_{3}. Then Ad(k)Ch1nL(E)Ch=Ch1nL(E)ChAd(k)C_{h}^{-1}n_{L}^{\vee}(E)C_{h}=C_{h}^{-1}n_{L}^{\vee}(E^{\prime})C_{h} with E=det(A+iB)t(A+iB)1EE^{\prime}=\det(A+iB)\,^{t}(A+iB)^{-1}E.

Proof.

We know from lemmas above that Ad(k)Ch1nL(E)Ch=Ch1nL(E)ChAd(k)C_{h}^{-1}n_{L}^{\vee}(E)C_{h}=C_{h}^{-1}n_{L}^{\vee}(E^{\prime})C_{h} for some EJE^{\prime}\in J. To compute it, we apply both sides to r0(i)=r1(i)r_{0}(i)=r_{1}(-i). The right-hand side gives 22Ch1(0,0,E,0)2\sqrt{2}C_{h}^{-1}(0,0,E^{\prime},0), which is (0,2iE,2E,0)(0,-2iE^{\prime},2E^{\prime},0). The left-hand side gives j(k1,i)k(0,2iE,2E,0)j(k^{-1},i)^{*}k\cdot(0,-2iE,2E,0).

To further compute this left-hand side, note that we have (0,iE,E,0)(0,iE,-E,0) is identified with (e1+if1)x1+(e2+if2)x2+(e3+if3)x3(e_{1}+if_{1})\otimes x_{1}+(e_{2}+if_{2})\otimes x_{2}+(e_{3}+if_{3})\otimes x_{3}. Applying kk, one gets the action of AiB=t(A+iB)1A-iB=\,^{t}(A+iB)^{-1} on the ej+ifje_{j}+if_{j}. ∎

2.4. More exceptional groups

We define the group F4IF_{4}^{I} to be the stabilizer of II inside of MJ1M_{J}^{1}. For the group GJG_{J}, recall the Lie algebra 𝔤(J){\mathfrak{g}}(J) of [Pol20a], so that GJG_{J} is the identity component of the group of automorphisms of 𝔤(J){\mathfrak{g}}(J). Let us remark that an integral model of the group GJG_{J} is described in [Gan00]. We use this integral model.

2.5. The second and third dual pairs

We will now define two more dual pairs, GL2×F4IHJ\operatorname{GL}_{2}\times F_{4}^{I}\subseteq H_{J} and G2×F4IGJG_{2}\times F_{4}^{I}\subseteq G_{J}.

For the second dual pair, GL2×F4IHJ\operatorname{GL}_{2}\times F_{4}^{I}\subseteq H_{J}, we explicate an identification W𝐐V2J0WJW_{\mathbf{Q}}\oplus V_{2}\otimes J^{0}\simeq W_{J}, as follows. Here V2V_{2} is the standard representation of GL2\operatorname{GL}_{2}, with standard basis e,fe,f. Identify (a,b,c,d)+fB+eC(a,b,c,d)+f\otimes B+e\otimes C with (a,bI+B,cIC,d)(a,bI+B,cI-C,d). We claim that under this identification, and the natural action of GL2×F4I\operatorname{GL}_{2}\times F_{4}^{I} on the left, this group action preserves the symplectic and quartic form on WJW_{J}.

It is clear that F4IF_{4}^{I} preserves the symplectic and quartic form. For the action of GL2\operatorname{GL}_{2}, the proof is similar to (but much easier than) that given above for GSp6×G2\operatorname{GSp}_{6}\times G_{2}. One simply needs to observe that nL(b)(0,B,0,0)=(0,0,bI×B,0)=(0,0,bB,0)n_{L}(b)(0,B,0,0)=(0,0,bI\times B,0)=(0,0,-bB,0) and nL(b)(fB)=beBn_{L}(b)(f\otimes B)=be\otimes B.

For the third dual pair, G2×F4IGJG_{2}\times F_{4}^{I}\subseteq G_{J}, we proceed as follows. Observe that F4IGJF_{4}^{I}\subseteq G_{J} from the action of F4IF_{4}^{I} on JJ and JJ^{\vee}. Now notice that 𝔤(J)F4I=𝔤2{\mathfrak{g}}(J)^{F_{4}^{I}}={\mathfrak{g}}_{2}. Because G2G_{2} is simply connected, we thus obtain a map from G2G_{2} to the centralizer of the group F4IF_{4}^{I} in GJG_{J}. Thus we have a unique map G2×F4IGJG_{2}\times F_{4}^{I}\rightarrow G_{J} so that F4IF_{4}^{I} acts on J,JJ,J^{\vee} naturally, and the differential of the map G2GJG_{2}\rightarrow G_{J} is the Lie algebra embedding 𝔤2𝔤(J){\mathfrak{g}}_{2}\rightarrow{\mathfrak{g}}(J).

We now relate these two dual pairs. Thus let GL2s\operatorname{GL}_{2}^{s} be the Levi of the Heisenberg parabolic in G2G_{2}. Observe that GL2s\operatorname{GL}_{2}^{s} fixes the line spanned by E13E_{13} in 𝔤(J){\mathfrak{g}}(J), because it does so in 𝔤2{\mathfrak{g}}_{2}. Thus because HJH_{J} is the Levi of the Heisenberg parabolic in GJG_{J}, we obtain a map GL2s×F4IHJGL(WJ)\operatorname{GL}_{2}^{s}\times F_{4}^{I}\rightarrow H_{J}\rightarrow\operatorname{GL}(W_{J}).

Proposition 2.5.1.

The above-described two maps GL2s×F4IHJ\operatorname{GL}_{2}^{s}\times F_{4}^{I}\rightarrow H_{J} are identical.

Proof.

Indeed, it is clear that the F4IF_{4}^{I}’s act exactly the same way. As for GL2s\operatorname{GL}_{2}^{s}, we have two algebraic representations of GL2s\operatorname{GL}_{2}^{s} on WJW_{J}. By the formulas for the Lie bracket on 𝔤(J){\mathfrak{g}}(J), it is easy to see that the differential of these representations of GL2s\operatorname{GL}_{2}^{s} are identical. Thus, they agree on the level of algebraic groups, as desired.∎

3. Facts about F4IF_{4}^{I}

We set down some notations and results we will need concerning the representations Vmλ3V_{m\lambda_{3}} and algebraic modular forms on F4IF_{4}^{I}.

3.1. Special elements of representations of F4F_{4}

Let J0JJ^{0}\subseteq J be the trace 0 subspace. Let ι:JJ\iota:J\rightarrow J^{\vee} be the identification of the JJ with its dual given by the trace pairing. Recall that if γJ\gamma\in J^{\vee} and xJx\in J then Φγ,xEnd(J)\Phi_{\gamma,x}\in End(J) is defined as

Φγ,x(z)=γ×(x×z)+(γ,z)x+(γ,x)z.\Phi_{\gamma,x}(z)=-\gamma\times(x\times z)+(\gamma,z)x+(\gamma,x)z.

If X,YJX,Y\in J then one defines ΦXY=Φι(X),YΦι(Y),X\Phi_{X\wedge Y}=\Phi_{\iota(X),Y}-\Phi_{\iota(Y),X}. This defines a map 2J𝔞(J)=𝔣4\wedge^{2}J\rightarrow{\mathfrak{a}}(J)={\mathfrak{f}}_{4}. (Here 𝔞(J)=f4{\mathfrak{a}}(J)=f_{4} is the Lie algebra of the subgroup of MJ1M_{J}^{1} that preserves the trace pairing, or equivalently, fixes the element IJI\in J.) As Φι(1),x=Φι(x),1\Phi_{\iota(1),x}=\Phi_{\iota(x),1}, this map factors through the projection 2J2J0\wedge^{2}J\rightarrow\wedge^{2}J^{0}.

Set

Vλ3=ker{2J0𝔣4}.V_{\lambda_{3}}=\ker\{\wedge^{2}J^{0}\rightarrow{\mathfrak{f}}_{4}\}.

One can construct special elements of Vλ3V_{\lambda_{3}} using the following two lemmas.

Lemma 3.1.1.

Suppose xJx\in J is rank one. Suppose zJz\in J and γ=z×x\gamma=z\times x. Then Φγ,x=0\Phi_{\gamma,x}=0.

Proof.

Observe that (γ,x)=(z×x,x)=2(z,x#=0)(\gamma,x)=(z\times x,x)=2(z,x^{\#}=0) so that

Φγ,x(z)=(x×z)×(x×z)+(x,z,z)x.\Phi_{\gamma,x}(z^{\prime})=-(x\times z)\times(x\times z^{\prime})+(x,z,z^{\prime})x.

Keeping this expression in mind, we now recall the identity

(u×v)#+u#×v×=(u,v#)u+(v,u#)v(u\times v)^{\#}+u^{\#}\times v^{\times}=(u,v^{\#})u+(v,u^{\#})v

valid for all u,vu,v in JJ. If u#=0u^{\#}=0, then symmetrizing this identity in vv gives

(u×v1)×(u×v2)=(u,v1,v2)u.(u\times v_{1})\times(u\times v_{2})=(u,v_{1},v_{2})u.

Consequently, taking u=xu=x, v1=zv_{1}=z and v2=zv_{2}=z^{\prime} we obtain

(x×z)×(x×z)=(x,z,z)x.(x\times z)\times(x\times z^{\prime})=(x,z,z^{\prime})x.

This gives the lemma. ∎

Lemma 3.1.2.

Suppose xJx\in J and γJ\gamma\in J^{\vee} are such that Φγ,x=0\Phi_{\gamma,x}=0. Then Φι(x),ι(γ)=0\Phi_{\iota(x),\iota(\gamma)}=0.

Proof.

The point is that one has (x,γ)=0(x,\gamma)=0 and Φγ,x=0\Phi_{\gamma,x}=0 (in fact, Φγ,x=0\Phi_{\gamma,x}=0 implies (γ,x)=0(\gamma,x)=0) if and only if (0,x,γ,0)(0,x,\gamma,0) is rank at most one in WJW_{J}. But J2(0,x,γ,0)=(0,ι(γ),ι(x),0)J_{2}(0,x,\gamma,0)=(0,-\iota(\gamma),\iota(x),0) so the lemma follows.

Of course, one can also give a more direct proof: One has

(ι(y),Φι(x),ι(γ)(ι(μ)))\displaystyle(\iota(y),\Phi_{\iota(x),\iota(\gamma)}(\iota(\mu))) =(ι(y),ι(x)×(ι(γ)×ι(μ))+(x,μ)ι(γ)+(x,γ)ι(μ))\displaystyle=(\iota(y),-\iota(x)\times(\iota(\gamma)\times\iota(\mu))+(x,\mu)\iota(\gamma)+(x,\gamma)\iota(\mu))
=(γ×(x×y),μ)+((y,γ)x,μ)+((x,γ)y,μ)\displaystyle=(-\gamma\times(x\times y),\mu)+((y,\gamma)x,\mu)+((x,\gamma)y,\mu)
=(Φγ,x(y),μ).\displaystyle=(\Phi_{\gamma,x}(y),\mu).

Thus if Φγ,x=0\Phi_{\gamma,x}=0, then Φι(x),ι(γ)=0\Phi_{\iota(x),\iota(\gamma)}=0. ∎

We now write down some specific elements x,yJ0x,y\in J^{0} with the following property:

  • The four dimensional space

    Ex,y:={(0,α1x+α2y,α3ι(x)+α4ι(y),0):αiF}E_{x,y}:=\{(0,\alpha_{1}x+\alpha_{2}y,\alpha_{3}\iota(x)+\alpha_{4}\iota(y),0):\alpha_{i}\in F\}

    is isotropic and singular in the sense that symplectic form on WJW_{J} restricted to Ex,yE_{x,y} is 0 and every element of Ex,yE_{x,y} has rank at most one.

Note that Ex,yE_{x,y} is isotropic and singular is equivalent to:

  • (x,x)=(x,y)=(y,y)=0(x,x)=(x,y)=(y,y)=0

  • x,yx,y are rank at most one, and x×y=0x\times y=0

  • Φι(x),x=0\Phi_{\iota(x),x}=0, Φι(x),y=0\Phi_{\iota(x),y}=0, Φι(y),x=0\Phi_{\iota(y),x}=0 and Φι(y),y=0\Phi_{\iota(y),y}=0.

If x,yx,y satisfy the above properties, then xyVλ3x\wedge y\in V_{\lambda_{3}} is a highest weight vector for some Borel. Indeed, one can show that the span of x,yx,y is “amber”, in the sense of Aschbacher [Asc87, 9.3-9.5], see also [MS97, Definition 7.2, Proposition 7.3, Proposition 7.4(1)]. Consequently, such x,yx,y will allow us to construct explicit elements of Vmλ3.V_{m\lambda_{3}}.

Lemma 3.1.3.

Suppose xJ0x\in J^{0} is rank one, and that zJz\in J is such that (z,x)=0(z,x)=0 and (x,z#)=0(x,z^{\#})=0. Set y=ι(z×x)y=\iota(z\times x). Then x,yJ0x,y\in J^{0} and Ex,yE_{x,y} is isotropic and singular.

Proof.

First suppose xJ0x\in J^{0} is rank one and zJz\in J is arbitrary. Then if y=ι(z×x)y=\iota(z\times x), then Φι(y),x=0\Phi_{\iota(y),x}=0 and Φι(x),y=0\Phi_{\iota(x),y}=0.

Moreover, observe that if vJ0v\in J^{0} is rank one then 1×v=(1,v)1v=v1\times v=(1,v)1-v=-v, so that Φι(v),v=0\Phi_{\iota(v),v}=0. Thus Φι(x),x=0\Phi_{\iota(x),x}=0, and if we arrange that yJ0y\in J^{0} is rank one, then Φι(y),y=0\Phi_{\iota(y),y}=0.

To ensure that yJ0y\in J^{0} we use (z,x)=0(z,x)=0. Indeed,

(y,1)=(z×x,1)=(z,x×1)=(z,(1,x)1x)=(z,x).(y,1)=(z\times x,1)=(z,x\times 1)=(z,(1,x)1-x)=-(z,x).

To ensure yy has rank at most one, we use (x,z#)=0(x,z^{\#})=0. Indeed,

y#=(x×z)#=(x,z#)x.y^{\#}=(x\times z)^{\#}=(x,z^{\#})x.

Because Φι(x),x=0\Phi_{\iota(x),x}=0 and Φι(y),x=0\Phi_{\iota(y),x}=0 and Φι(y),y=0\Phi_{\iota(y),y}=0, we get for free that (x,x)=(x,y)=(y,y)=0(x,x)=(x,y)=(y,y)=0. (Of course, one can also check this directly.)

To complete the proof, we must verify that x×y=0x\times y=0. For this, we compute

(z,x×y)\displaystyle(z^{\prime},x\times y) =(z×x,y)\displaystyle=(z^{\prime}\times x,y)
=(z×x,1×y)\displaystyle=(z^{\prime}\times x,-1\times y)
=((z×x)×(z×x),1)\displaystyle=((z\times x)\times(z^{\prime}\times x),-1)
=((z,z,x)x,1)=0.\displaystyle=((z,z^{\prime},x)x,-1)=0.

Example 3.1.4.

As an example of such x,y,zJ𝐂x,y,z\in J\otimes{\mathbf{C}} we can take:

  • x=(1a3a2a31a3a2a2a2a30)x=\left(\begin{array}[]{ccc}1&a_{3}&a_{2}^{*}\\ a_{3}^{*}&-1&a_{3}^{*}a_{2}^{*}\\ a_{2}&a_{2}a_{3}&0\end{array}\right) with n(a2)=0n(a_{2})=0, a20a_{2}\neq 0, and n(a3)=1n(a_{3})=-1.

  • z=(00(a2)010a200)z=\left(\begin{array}[]{ccc}0&0&(a_{2}^{\prime})^{*}\\ 0&1&0\\ a_{2}^{\prime}&0&0\end{array}\right) with n(a2)=1n(a_{2}^{\prime})=1 and (a2,a2)=1(a_{2}^{\prime},a_{2})=1. Note that z#=zz^{\#}=-z and (z,x)=0(z,x)=0, so that (z#,x)=0(z^{\#},x)=0 as well.

  • Then y=z×x=(0(a2)(a2a3)1a3(a2)a2a21)y=z\times x=\left(\begin{array}[]{ccc}0&(a_{2}^{\prime})^{*}(a_{2}a_{3})&*\\ &-1&a_{3}^{*}(a_{2}^{\prime})^{*}\\ a_{2}^{\prime}-a_{2}&*&1\end{array}\right).

We will use this example to prove Corollaries 1.2.3 and 1.2.4.

3.2. Algebraic modular forms

Suppose VV is a representation of F4I(𝐑)F_{4}^{I}({\mathbf{R}}). By an algebraic modular form for F4IF_{4}^{I}, we mean an automorphic form α:F4I(𝐐)\F4I(𝐀)V\alpha:F_{4}^{I}({\mathbf{Q}})\backslash F_{4}^{I}({\mathbf{A}})\rightarrow V satisfying α(gk)=k1α(g)\alpha(gk)=k^{-1}\cdot\alpha(g) for all gF4I(𝐀)g\in F_{4}^{I}({\mathbf{A}}) and kF4I(𝐑)k\in F_{4}^{I}({\mathbf{R}}). If α\alpha has level one, then because the double coset F4I(𝐐)\F4I(𝐀f)/F4I(𝐙^)F_{4}^{I}({\mathbf{Q}})\backslash F_{4}^{I}({\mathbf{A}}_{f})/F_{4}^{I}(\widehat{{\mathbf{Z}}}) has size two, such α\alpha can be described by two elements of VV. In this subsection, we make this identification explicit.

Recall the elements I,EJRI,E\in J_{R} of norm 11, see [EG96]. Define F4IF_{4}^{I} to be the stabilizer of II in MJ1E6M_{J}^{1}\simeq E_{6} and F4EF_{4}^{E} to be the stabilizer of EE in MJ1M_{J}^{1}. From the point of view of double cosets, the element EJRE\in J_{R} arises as follows. Let {1,γE}\{1,\gamma_{E}\} be representatives for F4I(𝐐)\F4I(𝐀f)/F4I(𝐙^)F_{4}^{I}({\mathbf{Q}})\backslash F_{4}^{I}({\mathbf{A}}_{f})/F_{4}^{I}(\widehat{{\mathbf{Z}}}). Using strong approximation on MJ1M_{J}^{1}, we can write γE=δE𝐐(δE𝐑)1δE𝐙^\gamma_{E}=\delta_{E}^{\mathbf{Q}}(\delta_{E}^{\mathbf{R}})^{-1}\delta_{E}^{\widehat{{\mathbf{Z}}}} with δE𝐐MJ1(𝐐)\delta_{E}^{\mathbf{Q}}\in M_{J}^{1}({\mathbf{Q}}) etc. We can choose γE\gamma_{E} and δE?\delta_{E}^{?} so that E=(δE𝐐)1IE=(\delta_{E}^{{\mathbf{Q}}})^{-1}\cdot I. Indeed, observe that

  1. (1)

    E=(δE𝐐)1IJR𝐐E=(\delta_{E}^{\mathbf{Q}})^{-1}I\in J_{R}\otimes{\mathbf{Q}}

  2. (2)

    (δE𝐐)1=(δE𝐑)1δE𝐙^γE1(\delta_{E}^{{\mathbf{Q}}})^{-1}=(\delta_{E}^{{\mathbf{R}}})^{-1}\delta_{E}^{\widehat{{\mathbf{Z}}}}\gamma_{E}^{-1} so that (δE𝐐)1I(\delta_{E}^{\mathbf{Q}})^{-1}I has finite part in JR𝐙^J_{R}\otimes\widehat{{\mathbf{Z}}}.

We have F4E=(δE𝐐)1F4IδE𝐐F_{4}^{E}=(\delta_{E}^{{\mathbf{Q}}})^{-1}F_{4}^{I}\delta_{E}^{\mathbf{Q}}. If VV is a representation of F4I(𝐑)F_{4}^{I}({\mathbf{R}}), we let F4E(𝐑)F_{4}^{E}({\mathbf{R}}) act on VV via gEv=(δE𝐑g(δE𝐑)1)vg\cdot_{E}v=(\delta_{E}^{{\mathbf{R}}}g(\delta_{E}^{{\mathbf{R}}})^{-1})v.

We make the following notations:

  • Let ΓI\Gamma_{I} be the image of F4I(𝐐)(F4I(𝐙^)F4I(𝐑))F_{4}^{I}({\mathbf{Q}})\cap(F_{4}^{I}(\widehat{{\mathbf{Z}}})F_{4}^{I}({\mathbf{R}})) in F4I(𝐑)F_{4}^{I}({\mathbf{R}}). Thus

    ΓI=(F4I(𝐐)F4I(𝐙^))F4I(𝐑).\Gamma_{I}=(F_{4}^{I}({\mathbf{Q}})F_{4}^{I}(\widehat{{\mathbf{Z}}}))\cap F_{4}^{I}({\mathbf{R}}).
  • Let ΓE\Gamma_{E} be the image of F4E(𝐐)(F4E(𝐙^)F4E(𝐑))F_{4}^{E}({\mathbf{Q}})\cap(F_{4}^{E}(\widehat{{\mathbf{Z}}})F_{4}^{E}({\mathbf{R}})) in F4E(𝐑)F_{4}^{E}({\mathbf{R}}). Thus

    ΓE=(F4E(𝐐)F4E(𝐙^))F4E(𝐑).\Gamma_{E}=(F_{4}^{E}({\mathbf{Q}})F_{4}^{E}(\widehat{{\mathbf{Z}}}))\cap F_{4}^{E}({\mathbf{R}}).
  • Let ΓγE\Gamma_{\gamma_{E}} be the image in F4I(𝐑)F_{4}^{I}({\mathbf{R}}) of γE1F4I(𝐐)γE(F4I(𝐙^)F4I(𝐑))\gamma_{E}^{-1}F_{4}^{I}({\mathbf{Q}})\gamma_{E}\cap(F_{4}^{I}(\widehat{{\mathbf{Z}}})F_{4}^{I}({\mathbf{R}})). Thus

    ΓγE=((γE1F4I(𝐐)γE)F4I(𝐙^))F4I(𝐑).\Gamma_{\gamma_{E}}=((\gamma_{E}^{-1}F_{4}^{I}({\mathbf{Q}})\gamma_{E})F_{4}^{I}(\widehat{{\mathbf{Z}}}))\cap F_{4}^{I}({\mathbf{R}}).
Lemma 3.2.1.

One has δE𝐑ΓEδE𝐑,1=ΓγE.\delta_{E}^{\mathbf{R}}\Gamma_{E}\delta_{E}^{{\mathbf{R}},-1}=\Gamma_{\gamma_{E}}.

Proof.

Observe

ΓE=(F4E(𝐐)MJ1(𝐙^))F4E(𝐑)\Gamma_{E}=(F_{4}^{E}({\mathbf{Q}})M_{J}^{1}(\widehat{{\mathbf{Z}}}))\cap F_{4}^{E}({\mathbf{R}})

and

ΓγE=((γE1F4I(𝐐)γE)MJ1(𝐙^))F4I(𝐑).\Gamma_{\gamma_{E}}=((\gamma_{E}^{-1}F_{4}^{I}({\mathbf{Q}})\gamma_{E})M_{J}^{1}(\widehat{{\mathbf{Z}}}))\cap F_{4}^{I}({\mathbf{R}}).

Using that δE𝐑=δE𝐙^γE1δE𝐐\delta_{E}^{\mathbf{R}}=\delta_{E}^{\widehat{{\mathbf{Z}}}}\gamma_{E}^{-1}\delta_{E}^{{\mathbf{Q}}}, the lemma follows. ∎

Suppose VV is a representation of F4I(𝐑)F_{4}^{I}({\mathbf{R}}) and

α:F4I(𝐐)\F4I(𝐀)V\alpha:F_{4}^{I}({\mathbf{Q}})\backslash F_{4}^{I}({\mathbf{A}})\rightarrow V

satisfies α(gkfk𝐑)=k𝐑1α(g)\alpha(gk_{f}k_{\mathbf{R}})=k_{\mathbf{R}}^{-1}\alpha(g) for all gF4I(𝐀)g\in F_{4}^{I}({\mathbf{A}}), kfF4I(𝐙^)k_{f}\in F_{4}^{I}(\widehat{{\mathbf{Z}}}) and k𝐑F4I(𝐑)k_{\mathbf{R}}\in F_{4}^{I}({\mathbf{R}}). Then α\alpha is determined by its values at g=1g=1 and g=γEg=\gamma_{E}. Moreover, because F4(𝐙^)F_{4}(\widehat{{\mathbf{Z}}}) acts freely on F4I(𝐐)\F4I(𝐀)F_{4}^{I}({\mathbf{Q}})\backslash F_{4}^{I}({\mathbf{A}}), we have

F4I(𝐐)\F4I(𝐀)=(ΓI\F4I(𝐑))F4I(𝐙^)(ΓγE\F4I(𝐑))F4I(𝐙^).F_{4}^{I}({\mathbf{Q}})\backslash F_{4}^{I}({\mathbf{A}})=\left(\Gamma_{I}\backslash F_{4}^{I}({\mathbf{R}})\right)\cdot F_{4}^{I}(\widehat{{\mathbf{Z}}})\bigsqcup\left(\Gamma_{\gamma_{E}}\backslash F_{4}^{I}({\mathbf{R}})\right)\cdot F_{4}^{I}(\widehat{{\mathbf{Z}}}).

Note that the measures of the two open sets are in the proportion 1|ΓI|:1|ΓE|\frac{1}{|\Gamma_{I}|}:\frac{1}{|\Gamma_{E}|}. Consequently, if V=Vmλ3V=V_{m\lambda_{3}}, one has

[F4I]{D2mΘ(g,h),α(h)}𝑑h=1|ΓI|{D2mΘ(g,1),α(1)}+1|ΓE|{D2mΘ(g,γE),α(γE)}.\int_{[F_{4}^{I}]}\{D^{2m}\Theta(g,h),\alpha(h)\}\,dh=\frac{1}{|\Gamma_{I}|}\{D^{2m}\Theta(g,1),\alpha(1)\}+\frac{1}{|\Gamma_{E}|}\{D^{2m}\Theta(g,\gamma_{E}),\alpha(\gamma_{E})\}.

Set αI=α(1)\alpha_{I}=\alpha(1). Consider Vmλ3(2J0)m(2J)mV_{m\lambda_{3}}\subseteq(\wedge^{2}J^{0})^{\otimes m}\subseteq(\wedge^{2}J)^{\otimes m}, and let αE=δE𝐐,1α(γE)\alpha_{E}=\delta_{E}^{{\mathbf{Q}},-1}\alpha(\gamma_{E}). Observe the following lemma:

Lemma 3.2.2.

If b,xJb,x\in J then (δE𝐐b,x)I=(b,δE𝐐,1x)E(\delta_{E}^{{\mathbf{Q}}}b,x)_{I}=(b,\delta_{E}^{{\mathbf{Q}},-1}x)_{E}.

Proof.

One has

(δE𝐐b,x)I\displaystyle(\delta_{E}^{{\mathbf{Q}}}b,x)_{I} =14(I,I,δE𝐐b)(I,I,x)(I,δE𝐐b,x)\displaystyle=\frac{1}{4}(I,I,\delta_{E}^{{\mathbf{Q}}}b)(I,I,x)-(I,\delta_{E}^{{\mathbf{Q}}}b,x)
=14(E,E,b)(E,E,δE𝐐,1x)(E,b,δE𝐐,1x)\displaystyle=\frac{1}{4}(E,E,b)(E,E,\delta_{E}^{{\mathbf{Q}},-1}x)-(E,b,\delta_{E}^{{\mathbf{Q}},-1}x)
=(b,δE𝐐,1x)E.\displaystyle=(b,\delta_{E}^{{\mathbf{Q}},-1}x)_{E}.

If one writes JE0J_{E}^{0} to be the perpendicular space to EE under the pairing (,)E(\,,\,)_{E}, then αE(2JE0)m\alpha_{E}\in(\wedge^{2}J_{E}^{0})^{\otimes m}. This follows from the lemma with b=Eb=E. Moreover, observe that, for the action of MJ1M_{J}^{1} on (2J)m(\wedge^{2}J)^{\otimes m}, αE\alpha_{E} is stabilized by the action of ΓE\Gamma_{E}. Thus, we can think of our algebraic modular form as being the pair

(αI,αE)[(2J)m]ΓI[(2J)m]ΓE.(\alpha_{I},\alpha_{E})\in[(\wedge^{2}J)^{\otimes m}]^{\Gamma_{I}}\oplus[(\wedge^{2}J)^{\otimes m}]^{\Gamma_{E}}.

4. Theorems on Siegel modular forms

In this section, we give the proof of Theorem 1.1.1 and its corollaries, Corollary 1.1.2 and Corollary 1.1.3. We do this assuming Theorem 4.0.1 stated below, which is the main technical ingredient in the proof of Theorem 1.1.1. Theorem 4.0.1 will be proved in section 6.

To setup the statement of Theorem 4.0.1, suppose VV is a KE7K_{E_{7}} representation, and φ:HJ1(𝐑)V\varphi:H_{J}^{1}({\mathbf{R}})\rightarrow V is a function satisfying φ(gk)=k1φ(g)\varphi(gk)=k^{-1}\cdot\varphi(g). In this scenario, let DφD\varphi be the V𝔭J+,V\otimes{\mathfrak{p}}_{J}^{+,\vee}-valued function defined as

Dφ=αXαφXαD\varphi=\sum_{\alpha}{X_{\alpha}\varphi\otimes X_{\alpha}^{\vee}}

where {Xα}α\{X_{\alpha}\}_{\alpha} is a 𝐂{\mathbf{C}}-basis of 𝔭J+{\mathfrak{p}}_{J}^{+}. It is easily checked that DφD\varphi is again KE7K_{E_{7}}-equivariant. Recall also that j(g,Z):HJ1(𝐑)×J𝐂×j(g,Z):H_{J}^{1}({\mathbf{R}})\times\mathcal{H}_{J}\rightarrow{\mathbf{C}}^{\times} is the factor of automorphy defined in section 2. Let ρ[k1,k2]\rho_{[k_{1},k_{2}]} be the representation of GL3(𝐂)\operatorname{GL}_{3}({\mathbf{C}}) on Sk1(V3)Sk2(2V3)S^{k_{1}}(V_{3})\otimes S^{k_{2}}(\wedge^{2}V_{3}).

Theorem 4.0.1.

Suppose gSp6(𝐑)g\in\operatorname{Sp}_{6}({\mathbf{R}}) and βW(k1,k2)\beta\in W(k_{1},k_{2}). Let 0\ell\geq 0 be an integer. There is a nonzero constant Bk1,k2B_{k_{1},k_{2}}, independent of gg and TT, so that

j(g,i)ρ[k1,k2](J(g,i)){Dk1+2k2(j(g,i)e2πi(T,gi)),β}=Bk1,k2{Pk1,k2(T),β}e2πi(pr(T),gi).j(g,i)^{\ell}\rho_{[k_{1},k_{2}]}(J(g,i))\{D^{k_{1}+2k_{2}}(j(g,i)^{-\ell}e^{2\pi i(T,g\cdot i)}),\beta\}=B_{k_{1},k_{2}}\{P_{k_{1},k_{2}}(T),\beta\}e^{2\pi i(pr(T),g\cdot i)}.

Moreover, {Pk1,k2(T),β}\{P_{k_{1},k_{2}}(T),\beta\} lies in the highest weight submodule S[k1,k2]S^{[k_{1},k_{2}]} of Sk1(V3)Sk2(2V3)S^{k_{1}}(V_{3})\otimes S^{k_{2}}(\wedge^{2}V_{3}).

Recall that if α𝒜(G2a)W(k1,k2)\alpha\in\mathcal{A}(G_{2}^{a})\otimes W(k_{1},k_{2}) is a level one algebraic modular form for the representation W(k1,k2)W(k_{1},k_{2}), then we defined the theta lift of α\alpha as

Θ(α)(g):=G2a(𝐐)\G2a(𝐀){Dk1+2k2ΘKim((g,h)),α(h)}𝑑h=1|ΓG2|{Dk1+2k2ΘKim((g,1)),α(1)}\Theta(\alpha)(g):=\int_{G_{2}^{a}({\mathbf{Q}})\backslash G_{2}^{a}({\mathbf{A}})}{\{D^{k_{1}+2k_{2}}\Theta_{Kim}((g,h)),\alpha(h)\}\,dh}=\frac{1}{|\Gamma_{G_{2}}|}\{D^{k_{1}+2k_{2}}\Theta_{Kim}((g,1)),\alpha(1)\}

where we have normalized the measure so that G2a(𝐙^)G2a(𝐑)G_{2}^{a}(\widehat{{\mathbf{Z}}})G_{2}^{a}({\mathbf{R}}) has measure 11. By rescaling α\alpha or this measure, we can (and will) ignore the term 1|ΓG2|\frac{1}{|\Gamma_{G_{2}}|}.

We first state the fact that Θ(α)\Theta(\alpha) is the automorphic form corresponding to a Siegel modular form of weight [k1,k2][k_{1},k_{2}].

Proposition 4.0.2.

For gSp6(𝐑)g\in\operatorname{Sp}_{6}({\mathbf{R}}) and Z3Z\in\mathcal{H}_{3} the Siegel upper half-space of degree three, define

fΘ(α)(Z)=ρ[k1,k2](J(g,i))Θ(α)(g)f_{\Theta(\alpha)}(Z)=\rho_{[k_{1},k_{2}]}(J(g,i))\Theta(\alpha)(g)

for any gg with g(i13)=Zg\cdot(i1_{3})=Z in 3\mathcal{H}_{3}. Then fΘ(α)(Z)f_{\Theta(\alpha)}(Z) is well-defined, and is a level one Siegel modular form of weight [k1,k2][k_{1},k_{2}]. If k2>0k_{2}>0, it is a cusp form.

Proof.

The fact that fΘ(α)(Z)f_{\Theta(\alpha)}(Z) is well-defined comes from the KSp6U(3)K_{\operatorname{Sp}_{6}}\simeq U(3) action on 𝔭J+,{\mathfrak{p}}_{J}^{+,\vee}, which was determined in section 2. To see that it’s a holomorphic modular form of the correct weight, we use [GS98, Theorem 3.5]. The cuspidality when k2>0k_{2}>0 is proved in [GS98, Corollary 4.9].

In more detail, if S[k1,k2],\ell\in S^{[k_{1},k_{2}],\vee} is a linear form on S[k1,k2]S^{[k_{1},k_{2}]}, then one can write (Θ(α)(g))\ell(\Theta(\alpha)(g)) as a sum of terms of the form Θ(vj,αj)\Theta(v_{j},\alpha_{j}), where

Θ(vj,αj)=[G2a]Θvj((g,h))αj(h)𝑑h\Theta(v_{j},\alpha_{j})=\int_{[G_{2}^{a}]}{\Theta_{v_{j}}((g,h))\alpha_{j}(h)\,dh}

is a usual scalar-valued theta lift. By [GS98, Theorem 3.5], as functions of gg, these lifts all lie in the holomorphic discrete series representation π(k1,k2)\pi(k_{1},k_{2}) with minimal KSp6K_{\operatorname{Sp}_{6}}-type S[k1,k2]det4S^{[k_{1},k_{2}]}\det^{4}. Moreover, by the KSp6K_{\operatorname{Sp}_{6}}-equivariance that proves that fΘ(α)(Z)f_{\Theta(\alpha)}(Z) is well-defined, the vector-valued function Θ(α)\Theta(\alpha) exactly encompasses the minimal KSp6K_{\operatorname{Sp}_{6}}-type in π(k1,k2)\pi(k_{1},k_{2}). Here we use that this minimal KSp6K_{\operatorname{Sp}_{6}}-type appears in π(k1,k2)\pi(k_{1},k_{2}) with multiplicity one. Consequently, fΘ(α)(Z)f_{\Theta(\alpha)}(Z) is a holomorphic Siegel modular form. It is clearly level one. Finally, [GS98, Corollary 4.9] shows that all the Θ(vj,αj)\Theta(v_{j},\alpha_{j}) are cusp forms if k2>0k_{2}>0, thus so is Θ(α)\Theta(\alpha).

This completes the proof. ∎

Proof of Theorem 1.1.1.

Theorem 1.1.1 follows directly from Proposition 4.0.2 and Theorem 4.0.1 by plugging in the Fourier expansion of Kim’s modular form ΘKim(Z)\Theta_{Kim}(Z).∎

We now explain the proofs of Corollary 1.1.2 and Corollary 1.1.3, especially that of Claim 1.1.4.

Proof of Corollary 1.1.2.

Let u,vu,v span a null, two-dimensional subspace of the trace zero elements Θ0𝐂\Theta^{0}\otimes{\mathbf{C}}. That they are null means that u2=uv=vu=v2=0u^{2}=uv=vu=v^{2}=0. We set β=uk1(uv)k2\beta=u^{\otimes k_{1}}\otimes(u\wedge v)^{\otimes k_{2}}. It is easy to see that βW(k1,k2)\beta\in W(k_{1},k_{2}). Indeed, we can choose a Borel subgroup of G2(𝐂)G_{2}({\mathbf{C}}) to be the one that stabilizes the flag 𝐂u𝐂u𝐂v{\mathbf{C}}u\subseteq{\mathbf{C}}u\oplus{\mathbf{C}}v, and then it is clear that β\beta is a highest weight vector in W(k1,k2)W(k_{1},k_{2}). A computer calculation shows that, if β1\beta_{1} is defined as above with k1=0k_{1}=0 and k2=4k_{2}=4, then the

T0:=12(211121112)T_{0}:=\frac{1}{2}\left(\begin{array}[]{ccc}2&1&1\\ 1&2&1\\ 1&1&2\end{array}\right)

Fourier coefficient of Θ(β1)\Theta(\beta_{1}) is nonzero. Similarly, if β2\beta_{2} is defined as above with k1=2k_{1}=2 and k2=4k_{2}=4, then the T0T_{0} Fourier coefficient of Θ(β2)\Theta(\beta_{2}) is nonzero.

To actually do the computation on a computer, we proceed as follows. First, we set HH to be the quaternion algebra over 𝐐{\mathbf{Q}} ramified at 22 and the archimedean place. Let 1,i,j,k1,i,j,k be its usual basis. We obtain the octonion algebra Θ=HH\Theta=H\oplus H via the Cayley-Dickson construction using γ=1\gamma=-1. This means that the addition in Θ\Theta is component-wise and the multiplication is

(x1,y1)(x2,y2)=(x1x2+γy2y1,y2x1+y1x2).(x_{1},y_{1})(x_{2},y_{2})=(x_{1}x_{2}+\gamma y_{2}^{*}y_{1},y_{2}x_{1}+y_{1}x_{2}^{*}).

Set e=(0,1)e=(0,1) and h=12(i+j+k+e)h=\frac{1}{2}(i+j+k+e). Then, the following are a 𝐙{\mathbf{Z}} basis of RΘR_{\Theta}, Coxeter’s ring [Cox46]: jh,e,h,j,ih,1,eh,kejh,e,-h,j,ih,1,eh,ke. These are the simple roots of the E8E_{8} root lattice, with jhjh the extended node, 11 the branch vertex, and e,h,j,ih,1,eh,kee,-h,j,ih,1,eh,ke going along longways.

For u,vu,v we take elements inside of Θ𝐐(1)\Theta\otimes{\mathbf{Q}}(\sqrt{-1}) as

u=12((0,1)1(0,i));v=12((0,j)1(0,k)).u=\frac{1}{2}((0,1)-\sqrt{-1}(0,i));\,\,\,v=\frac{1}{2}((0,j)-\sqrt{-1}(0,k)).

Finally, to compute the T0T_{0} Fourier coefficient, where

T0=12(2afef2bded2c),T_{0}=\frac{1}{2}\left(\begin{array}[]{ccc}2a&f&e\\ f&2b&d\\ e&d&2c\end{array}\right),

we must explain how to enumerate the rank one TJRT\in J_{R} with pr(T)=T0pr(T)=T_{0}. The point is that

T=(c1x3x2x3c2x1x2x1c3)T=\left(\begin{array}[]{ccc}c_{1}&x_{3}&x_{2}^{*}\\ x_{3}^{*}&c_{2}&x_{1}\\ x_{2}&x_{1}^{*}&c_{3}\end{array}\right)

being rank one with pr(T)=T0pr(T)=T_{0} implies n(x1)=c2c3=bc,n(x_{1})=c_{2}c_{3}=bc, n(x2)=c3c1=can(x_{2})=c_{3}c_{1}=ca, n(x3)=c1c2=abn(x_{3})=c_{1}c_{2}=ab. Thus one only must search through a finite list of xix_{i} (namely, those with these norms) to find all such TT. ∎

In order to prove Corollary 1.1.3, we require Claim 1.1.4, which we prove now.

Proof of Claim 1.1.4.

Recall that we assume k2>0k_{2}>0, that FF is a level one Siegel modular form of weight [k1,k2][k_{1},k_{2}] and we wish to prove that if FF is in the image of the theta correspondence from G2aG_{2}^{a}, then F=Θ(α)F=\Theta(\alpha) for a level one algebraic modular form α(1)W(k1,k2)ΓG2\alpha(1)\in W(k_{1},k_{2})^{\Gamma_{G_{2}}}.

We first use the Howe Duality theorem of Gan-Savin [GS21] to reduce to the case of eigenforms. Thus write F=jFjF=\sum_{j}{F_{j}} as am orthogonal sum of eigenforms with each FjF_{j} nonzero. Then the Petersson inner product (Fj,F)0(F_{j},F)\neq 0 for each jj. Let πFj\pi_{F_{j}} be the automorphic cuspidal representation of PGSp6\operatorname{PGSp}_{6} generated by FjF_{j}. Then, by changing the order of the theta integral, one sees that the big theta lift Θ(πFj)\Theta(\pi_{F_{j}}) of πFj\pi_{F_{j}} to G2aG_{2}^{a} is nonzero. By Howe Duality [GS21] and the argument in [Gan22, Proposition 3.1], τj:=Θ(πFj)\tau_{j}:=\Theta(\pi_{F_{j}}) is an irreducible representation of G2aG_{2}^{a}.

We now argue that Θ(τj)=πFj\Theta(\tau_{j})=\pi_{F_{j}}. Let W(k1,k2)W(k_{1}^{\prime},k_{2}^{\prime}) the archimedean component of τj\tau_{j}, which a priori could depend upon jj. For any fixed vector yjy_{j} in the finite part of τjτj,fW(k1,k2)\tau_{j}\simeq\tau_{j,f}\otimes W(k_{1}^{\prime},k_{2}^{\prime}), and any fixed vector vjv_{j} in the finite part of ΠminΠmin,fΠmin,\Pi_{min}\simeq\Pi_{min,f}\otimes\Pi_{min,\infty} one can consider the map

Πmin,W(k1,k2)𝒜(Sp6)\Pi_{min,\infty}\otimes W(k_{1}^{\prime},k_{2}^{\prime})\rightarrow\mathcal{A}(\operatorname{Sp}_{6})

given by the theta lift. By the archimedean correspondence proved in Gross-Savin [GS98], the image is a holomorphic discrete series representation π(k1,k2)\pi(k_{1}^{\prime},k_{2}^{\prime}) with lowest KSp6K_{\operatorname{Sp}_{6}}-type S[k1,k2]S^{[k_{1}^{\prime},k_{2}^{\prime}]}. Because for some choice of yjy_{j} and vjv_{j} we must obtain a theta lift that is not orthogonal to FjF_{j}, we must have k1=k1k_{1}^{\prime}=k_{1} and k2=k2k_{2}^{\prime}=k_{2}.

It now follows from [GS98], because k2=k2>0k_{2}^{\prime}=k_{2}>0, that Θ(τj)\Theta(\tau_{j}) consists of cusp forms. Thus again by Howe Duality [GS21] and the argument in [Gan22, Proposition 3.1], Θ(τj)\Theta(\tau_{j}) is irreducible, so Θ(τj)=πFj\Theta(\tau_{j})=\pi_{F_{j}}. By the results of [GS98, MS97, GS21] that apply to spherical representations, it must be that τj\tau_{j} is unramified at every finite place. Thus we have a level one algebraic modular form αj(1)W(k1,k2)ΓG2\alpha_{j}(1)\in W(k_{1},k_{2})^{\Gamma_{G_{2}}}, and we must show that Θ(αj)\Theta(\alpha_{j}) is a nonzero multiple of FjF_{j}. We have now reduced ourselves to the case of eigenforms, so will drop the jj from αj\alpha_{j}, FjF_{j} and τj\tau_{j}.

Let b1,,bNb_{1},\ldots,b_{N} be a basis of W(k1,k2)W(k_{1},k_{2}), φj\varphi_{j} the level one automorphic form on G2a(𝐀)G_{2}^{a}({\mathbf{A}}) corresponding to bjb_{j}, and let w1,,wMw_{1},\ldots,w_{M} be a basis of S[k1,k2]S^{[k_{1},k_{2}]}. By changing variables in the integral defining the theta lift from G2aG_{2}^{a} to Sp6\operatorname{Sp}_{6}, one sees that there exists vi,jv_{i,j} in Πmin\Pi_{min} so that

F=i,jΘ(vi,j,φj)wi.F=\sum_{i,j}{\Theta(v_{i,j},\varphi_{j})\otimes w_{i}}.

Set v=i,jvi,jwibjΠminS[k1,k2]W(k1,k2)v^{\prime}=\sum_{i,j}{v_{i,j}^{\prime}\otimes w_{i}\otimes b_{j}}\in\Pi_{min}\otimes S^{[k_{1},k_{2}]}\otimes W(k_{1},k_{2}) and α=jφjbj\alpha=\sum_{j}\varphi_{j}\otimes b_{j}^{\vee}. Then we have

F=[G2a]{Θv((g,h)),α(h)}𝑑h.F=\int_{[G_{2}^{a}]}{\{\Theta_{v^{\prime}}((g,h)),\alpha(h)\}\,dh}.

Let

v=KSp6×G2a(𝐑)(k1,k2)v𝑑k1𝑑k2v=\int_{K_{\operatorname{Sp}_{6}}\times G_{2}^{a}({\mathbf{R}})}{(k_{1},k_{2})\cdot v^{\prime}\,dk_{1}\,dk_{2}}

where the action is diagonal on the minimal representation and the vector space S[k1,k2]W(k1,k2)S^{[k_{1},k_{2}]}\otimes W(k_{1},k_{2}). Because α\alpha is G2a(𝐑)G_{2}^{a}({\mathbf{R}})-equivariant and FF is KSp6K_{\operatorname{Sp}_{6}}-equivariant, one has

F=[G2a]{Θv((g,h)),α(h)}𝑑h.F=\int_{[G_{2}^{a}]}{\{\Theta_{v}((g,h)),\alpha(h)\}\,dh}.

Because the KSp6×G2a(𝐑)K_{\operatorname{Sp}_{6}}\times G_{2}^{a}({\mathbf{R}})-type (S[k1,k2]W(k1,k2))(S^{[k_{1},k_{2}]}\otimes W(k_{1},k_{2}))^{\vee} appears in Πmin,\Pi_{min,\infty} with multiplicity one [GS98], we can write

v=i,jvi,j,fvi,j,wibjv=\sum_{i,j}{v_{i,j,f}\otimes v_{i,j,\infty}\otimes w_{i}\otimes b_{j}}

where vi,j,Πmin,v_{i,j,\infty}\in\Pi_{min,\infty} is the basis of (S[k1,k2]W(k1,k2))Πmin,(S^{[k_{1},k_{2}]}\otimes W(k_{1},k_{2}))^{\vee}\subseteq\Pi_{min,\infty} dual to the basis wibjw_{i}\otimes b_{j}. We wish to show that we can set vi,j,fv_{i,j,f} to be the spherical vector in Πmin,f\Pi_{min,f} for all i,ji,j in this equality.

Fixing the vectors vi,j,v_{i,j,\infty} and letting the vi,j,fv_{i,j,f} vary, the theta integral gives an equivariant map

Πmin,fτfπF,f.\Pi_{min,f}\otimes\tau_{f}\rightarrow\pi_{F,f}.

This map is nonzero by assumption, so by pp-adic Howe Duality again [GS21] the map is uniquely determined up to scalar multiple. Our goal is to show that the image of the spherical vector on the left-hand side is nonzero on the right-hand side. We will do this by a global argument.

It suffices to show that, for each prime pp, in the unique up to scalar map Πmin,pτpπF,p\Pi_{min,p}\otimes\tau_{p}\rightarrow\pi_{F,p}, the image of the spherical vector is nonzero. Because this map is unique up to scalar multiple, we must only find some vΠminv\in\Pi_{min} which is spherical at pp, some φτ\varphi\in\tau which is spherical at pp, and so that Θ(v,φ)0\Theta(v,\varphi)\neq 0. By [GS98, Proposition 4.5], it suffices to check that α\alpha has an appropriate period. More precisely, let qq be an odd prime with qpq\neq p, and let BqB_{q} be the quaternion algebra over 𝐐{\mathbf{Q}} ramified at qq and infinity. Then it suffices to show that α\alpha has a BqB_{q} period. But finally, by Bocherer-Das [BD21], for some odd prime qpq\neq p, FF has a nonzero Fourier coefficient associated to the maximal order in BqB_{q}. Consequently, α\alpha does have a BqB_{q} period and we have shown that the unique map Πmin,pτpπF,p\Pi_{min,p}\otimes\tau_{p}\rightarrow\pi_{F,p} is nonzero on the spherical vector.

Let

v0=i,jv0,fvi,j,wibjv_{0}=\sum_{i,j}{v_{0,f}\otimes v_{i,j,\infty}\otimes w_{i}\otimes b_{j}}

where v0,fv_{0,f} is the spherical vector in Πmin,f\Pi_{min,f}. We have shown

[G2a]{Θv0((g,h)),α(h)}𝑑h.\int_{[G_{2}^{a}]}{\{\Theta_{v_{0}}((g,h)),\alpha(h)\}\,dh}.

is nonzero, and thus a nonzero multiple of FF. Because

Θ(α)(g):=[G2a]{Dk1+2k2ΘKim((g,h)),α(h)}𝑑h=[G2a]{Θv0((g,h)),α(h)}𝑑h,\Theta(\alpha)(g):=\int_{[G_{2}^{a}]}{\{D^{k_{1}+2k_{2}}\Theta_{Kim}((g,h)),\alpha(h)\}\,dh}=\int_{[G_{2}^{a}]}{\{\Theta_{v_{0}}((g,h)),\alpha(h)\}\,dh},

we obtain Θ(α)\Theta(\alpha) is nonzero, and thus is nonzero multiple of FF. The claim is proved. ∎

Proof of Corollary 1.1.3.

We have explained, given βW(k1,k2)V7k1(2V7)k2\beta\in W(k_{1},k_{2})\subseteq V_{7}^{\otimes k_{1}}\otimes(\wedge^{2}V_{7})^{\otimes k_{2}}, how to compute individual Fourier coefficients of Θ(β)\Theta(\beta). It remains to explain how to enumerate a spanning set of W(k1,k2)W(k_{1},k_{2}). To do this, we define elements e1,e2,e3,e1,e2,e3,u0=ϵ1ϵ2e_{1},e_{2},e_{3},e_{1}^{*},e_{2}^{*},e_{3}^{*},u_{0}=\epsilon_{1}-\epsilon_{2} which are a basis of V7𝐐(1)V_{7}\otimes{\mathbf{Q}}(\sqrt{-1}), as follows.

  • e2=12((0,1)1(0,i))e_{2}=\frac{1}{2}((0,1)-\sqrt{-1}(0,i))

  • e3=12((0,j)1(0,k))e_{3}^{*}=\frac{1}{2}((0,j)-\sqrt{-1}(0,k))

  • e3=12((0,j)1(0,k))e_{3}=\frac{1}{2}((0,-j)-\sqrt{-1}(0,k)).

  • e2=12((0,1)1(0,i))e_{2}^{*}=\frac{1}{2}((0,-1)-\sqrt{-1}(0,i))

  • ϵ1=12((1,0)1(i,0))\epsilon_{1}=\frac{1}{2}((1,0)-\sqrt{-1}(i,0))

  • ϵ2=12((1,0)+1(i,0))\epsilon_{2}=\frac{1}{2}((1,0)+\sqrt{-1}(i,0))

  • e1=12((j,0)1(k,0))e_{1}=\frac{1}{2}((j,0)-\sqrt{-1}(k,0))

  • e1=12((j,0)1(k,0))e_{1}^{*}=\frac{1}{2}((-j,0)-\sqrt{-1}(k,0))

Now, in terms of these elements, a basis for root spaces of 𝔤2{\mathfrak{g}}_{2} can be found in [Pol19]. In particular, for the standard Borel chosen in [Pol19], v(k1,k2):=e1k1(e1e3)k2v(k_{1},k_{2}):=e_{1}^{\otimes k_{1}}\otimes(e_{1}\wedge e_{3}^{*})^{\otimes k_{2}} will be a highest weight vector for W(k1,k2)W(k_{1},k_{2}). By the Poincare-Birkoff-Witt theorem, the representation W(k1,k2)W(k_{1},k_{2}) is spanned by {y1m1y2m2v(k1,k2)|m1,m2N(k1,k2)}\{y_{1}^{m_{1}}y_{2}^{m_{2}}v(k_{1},k_{2})|m_{1},m_{2}\leq N(k_{1},k_{2})\} where here

  • y1,y2y_{1},y_{2} span the negative root spaces of the simple roots of 𝔤2{\mathfrak{g}}_{2};

  • the element y1m1y2m2v(k1,k2)y_{1}^{m_{1}}y_{2}^{m_{2}}v(k_{1},k_{2}) can be explicitly computed using the formulas of [Pol19];

  • one can come up with easy bounds for the integer N(k1,k2)N(k_{1},k_{2}).

The corollary follows. ∎

5. Theorems on quaternionic modular forms

In this section we prove Theorem 1.2.1, assuming a crucial technical result, Theorem 5.0.1, which is proved in section 7. We also prove Corollaries 1.2.2, 1.2.3, and 1.2.4.

For wWJ(𝐑)w\in W_{J}({\mathbf{R}}) that is positive semi-definite, and 1\ell\geq 1 an integer, let Ww,W_{w,\ell} be the generalized Whittaker function [Pol20a] on GJ(𝐑)=E8,4G_{J}({\mathbf{R}})=E_{8,4} associated to ww and \ell. Similarly, for w0w_{0} a real binary cubic form that is positive semi-definite, let Ww0,W_{w_{0},\ell} be the associated generalized Whittaker function on G2(𝐑)G_{2}({\mathbf{R}}).

Let now 𝔭{\mathfrak{p}} denote the complexification of the 1-1 part for the Cartan involution on 𝔤(J)𝐑{\mathfrak{g}}(J)\otimes{\mathbf{R}} for the Cartan involution defined in [Pol20a]. Recall that if φ:GJ(𝐑)V\varphi:G_{J}({\mathbf{R}})\rightarrow V is a smooth function, then DφC(GJ(𝐑);V𝔭)D\varphi\in C^{\infty}(G_{J}({\mathbf{R}});V\otimes{\mathfrak{p}}^{\vee}) is defined as Dφ=αXαφXαD\varphi=\sum_{\alpha}{X_{\alpha}\varphi\otimes X_{\alpha}^{\vee}}, where {Xα}α\{X_{\alpha}\}_{\alpha} is a basis of 𝔭{\mathfrak{p}}.

Theorem 5.0.1.

Suppose =4\ell=4 and βVmλ3\beta\in V_{m\lambda_{3}} with m0m\geq 0. Then there is a nonzero constant BmB_{m} so that for all wWJw\in W_{J} rank one and gG2(𝐑)g\in G_{2}({\mathbf{R}}), one has

{D2mWw,(g),β}=BmPm(w),βIWprI(w),+m(g).\{D^{2m}W_{w,\ell}(g),\beta\}=B_{m}\langle P_{m}(w),\beta\rangle_{I}W_{pr_{I}(w),\ell+m}(g).

Here, the F4F_{4}-equivariant pairing {,}\{\,,\,\} is defined as follows. By virtue of the exceptional Cayley transform of [Pol20a] and the explanations of subsection 2.5, one has 𝔭V2WJ{\mathfrak{p}}\simeq V_{2}^{\ell}\otimes W_{J}, and WJ=W𝐐V2sJ0W_{J}=W_{\mathbf{Q}}\oplus V_{2}^{s}\otimes J^{0}. Thus

𝔭2m(V2)2m(V2sJ0)2mS2m(V2)det(V2s)m(2J0)m.{\mathfrak{p}}^{\otimes 2m}\rightarrow(V_{2}^{\ell})^{\otimes 2m}\otimes(V_{2}^{s}\otimes J^{0})^{\otimes 2m}\rightarrow S^{2m}(V_{2}^{\ell})\otimes\det(V_{2}^{s})^{\otimes m}\otimes(\wedge^{2}J^{0})^{\otimes m}.

Thus if βVmλ3(2J0)m\beta\in V_{m\lambda_{3}}\subseteq(\wedge^{2}J^{0})^{\otimes m}, and rS2(V2)𝔭2mr\in S^{2\ell}(V_{2}^{\ell})\otimes{\mathfrak{p}}^{\otimes 2m} we obtain an element {r,β}\{r,\beta\} in S2m+2(V2)S^{2m+2\ell}(V_{2}^{\ell}).

We can use the theorem to compute the Fourier expansion of the theta lift Θ(α)\Theta(\alpha) of a level one algebraic modular form α\alpha on F4IF_{4}^{I}. We begin with the statement that these lifts are quaternionic modular forms of weight 4+m4+m on G2G_{2}.

Proposition 5.0.2.

Suppose m0m\geq 0, and βVmλ3\beta\in V_{m\lambda_{3}}. Then

Θ(β):={D2mΘGan(g,1),β}\Theta(\beta):=\{D^{2m}\Theta_{Gan}(g,1),\beta\}

is a quaternionic modular form on G2G_{2} of weight 4+m4+m. If m>0m>0, it is a cusp form.

Proof.

First note that, if γΓI\gamma\in\Gamma_{I}, then

{D2mΘGan(g,1),γβ}={γ1D2mΘGan(g,1),β}={D2mΘGan(g,γ),β}={D2mΘGan(g,1),β}.\{D^{2m}\Theta_{Gan}(g,1),\gamma\cdot\beta\}=\{\gamma^{-1}\cdot D^{2m}\Theta_{Gan}(g,1),\beta\}=\{D^{2m}\Theta_{Gan}(g,\gamma),\beta\}=\{D^{2m}\Theta_{Gan}(g,1),\beta\}.

Thus if α\alpha is the level one algebraic modular form on F4IF_{4}^{I} with α(1)=γΓIγβ\alpha(1)=\sum_{\gamma\in\Gamma_{I}}{\gamma\cdot\beta} and α(γE)=0\alpha(\gamma_{E})=0, then

Θ(α)=[F4I]{D2mΘGan(g,h),α(h)}𝑑h=1|ΓI|{D2mΘGan(g,1),α(1)}=Θ(β).\Theta(\alpha)=\int_{[F_{4}^{I}]}{\{D^{2m}\Theta_{Gan}(g,h),\alpha(h)\}\,dh}=\frac{1}{|\Gamma_{I}|}\{D^{2m}\Theta_{Gan}(g,1),\alpha(1)\}=\Theta(\beta).

Now, for vΠmin,v\in\Pi_{min,\infty} and wVmλ3w\in V_{m\lambda_{3}}, consider the theta lift

Θ(v,φw)(g)=[F4I]Θv(g,h)φw(h)𝑑h\Theta(v,\varphi_{w})(g)=\int_{[F_{4}^{I}]}{\Theta_{v}(g,h)\varphi_{w}(h)\,dh}

where φw(h)={α(h),w}\varphi_{w}(h)=\{\alpha(h),w\} is the level one automorphic function on F4I(𝐀)F_{4}^{I}({\mathbf{A}}) associated to wVmλ3w\in V_{m\lambda_{3}}. This lift gives an equivariant pairing Πmin,Vmλ3𝒜(G2)\Pi_{min,\infty}\otimes V_{m\lambda_{3}}\rightarrow\mathcal{A}(G_{2}). By [HPS96], it thus gives a map π4+m𝒜(G2)\pi_{4+m}\rightarrow\mathcal{A}(G_{2}). Finally, by the KG2K_{G_{2}}-equivariance of Θ(α)\Theta(\alpha), we see that Θ(α)\Theta(\alpha) is the minimal KK-type of this copy of π4+m\pi_{4+m}, so it is a quaternionic modular form of weight 4+m4+m.

We now show the cuspidality of the theta lifts if m>0m>0. Let PP and QQ, respectively, be the two standard maximal parabolic subgroups of G2G_{2}, so that PP is the Heisenberg parabolic. Let PJP_{J} be the Heisenberg parabolic of GJG_{J}, and QJQ_{J} the standard maximal parabolic with QJG2=QQ_{J}\cap G_{2}=Q. In terms of the F4F_{4} root system underlying the group GJG_{J}, with long roots α1,α2\alpha_{1},\alpha_{2} and short roots α3,α4\alpha_{3},\alpha_{4}, QJQ_{J} is the parabolic for which α2\alpha_{2} is in its unipotent radical. The Levi subgroup of QJQ_{J} is of absolute Dynkin type A1×E6A_{1}\times E_{6}. Write P=MNP=MN, Q=LVQ=LV, PJ=MJNJP_{J}=M_{J}N_{J}, and QJ=LJVJQ_{J}=L_{J}V_{J} for the Levi decompositions.

We must check that the constant terms Θ(v,φ)N\Theta(v,\varphi)_{N} and Θ(v,φ)V\Theta(v,\varphi)_{V} are 0, if φ\varphi is an automorphic form in a representation τ\tau with τ=Vmλ3\tau_{\infty}=V_{m\lambda_{3}} with m>0m>0. We first observe the following claim:

Claim 5.0.3.

One has an equality of constant terms Θv,N=Θv,NJ\Theta_{v,N}=\Theta_{v,N_{J}} and Θv,V=Θv,VJ\Theta_{v,V}=\Theta_{v,V_{J}}.

Granting the claim for the moment, we obtain that Θ(v,φ)N(g)=[F4I]Θv,NJ(g,h)φ(h)𝑑h\Theta(v,\varphi)_{N}(g)=\int_{[F_{4}^{I}]}{\Theta_{v,N_{J}}(g,h)\varphi(h)\,dh} and Θ(v,φ)V(g)=[F4I]Θv,VJ(g,h)φ(h)𝑑h\Theta(v,\varphi)_{V}(g)=\int_{[F_{4}^{I}]}{\Theta_{v,V_{J}}(g,h)\varphi(h)\,dh}. The constant terms Θv,NJ\Theta_{v,N_{J}} and Θv,VJ\Theta_{v,V_{J}}, restricted to their Levi subgroups, were determined in [Gan00]. For the first one, see page 174 of [Gan00], it is a sum of terms from a one-dimensional representation of MJM_{J} and the minimal representation of MJM_{J}. Both of these have integral 0 against φ\varphi, because by [HPS96], the representation Vmλ3V_{m\lambda_{3}} with m>0m>0 does not participate in the theta correspondence for the dual pair SL2×F4IHJ1\operatorname{SL}_{2}\times F_{4}^{I}\subseteq H_{J}^{1}. For the second one, see page 176 of [Gan00], the constant term restricted to F4IF_{4}^{I} is the trivial representation. Thus this too has integral 0 against φ\varphi.

It remains to explain the proof of Claim 5.0.3. For the equality Θv,N=Θv,NJ\Theta_{v,N}=\Theta_{v,N_{J}}, note that Θv,N\Theta_{v,N} is a sum of terms of the form wWJ:rk(w)1,prI(w)=0Θw(g)\sum_{w\in W_{J}:rk(w)\leq 1,pr_{I}(w)=0}{\Theta_{w}(g)}. But, using that if TJT\in J is rank one with tr(T)=0\operatorname{tr}(T)=0 then T=0T=0, we find that the only wWJw\in W_{J} with rk(w)1rk(w)\leq 1 and prI(w)=0pr_{I}(w)=0 is w=0w=0. Thus Θv,N(g)=Θv,0(g)=Θv,NJ(g)\Theta_{v,N}(g)=\Theta_{v,0}(g)=\Theta_{v,N_{J}}(g). One makes a completely similar argument for the constant term Θv,V(g)\Theta_{v,V}(g). The proposition is proved. ∎

Proof of Theorem 1.2.1.

Suppose αIVmλ3ΓI\alpha_{I}\in V_{m\lambda_{3}}^{\Gamma_{I}}, and α\alpha is the level one algebraic modular form on F4IF_{4}^{I} with α(1)=αI\alpha(1)=\alpha_{I} and αE=0\alpha_{E}=0. Then it follows from Proposition 5.0.2 that the theta lift Θ(α)\Theta(\alpha) of α\alpha to G2G_{2} is a quaternionic modular form of weight 4+m4+m, and cuspidal if m>0m>0. Up to the constant BmB_{m}, its Fourier expansion is given exactly as in the statement of Theorem 1.2.1. The general case, where αE0\alpha_{E}\neq 0, is explained in subsection 5.1 below.∎

We explain the proof of Corollary 1.2.2.

Proof of Corollary 1.2.2.

Over 𝐂{\mathbf{C}}, we have a decomposition (2J0)m𝐂=Vmλ3V(\wedge^{2}J^{0})^{\otimes m}\otimes{\mathbf{C}}=V_{m\lambda_{3}}\oplus V^{\prime}, where VV^{\prime} is F4IF_{4}^{I}-stable. Because F4IF_{4}^{I} is a pure inner form of split F4F_{4}, one can use the results of [BGW15, sections 2.1, 2.2] to give such a decomposition over 𝐐{\mathbf{Q}}: (2J0)m=Vmλ3,𝐐V𝐐(\wedge^{2}J^{0})^{\otimes m}=V_{m\lambda_{3},{\mathbf{Q}}}\oplus V_{{\mathbf{Q}}}^{\prime}, where Vmλ3,𝐐V_{m\lambda_{3},{\mathbf{Q}}} is a rational representation of the algebraic group F4IF_{4}^{I} whose complexification is Vmλ3V_{m\lambda_{3}}. Now let LmL_{m} be the intersection of Vmλ3,𝐐V_{m\lambda_{3},{\mathbf{Q}}} with (2JR)m(\wedge^{2}J_{R})^{\otimes m}; it is immediately seen to be an integral lattice in Vmλ3V_{m\lambda_{3}}, so that Lm𝐂=Vmλ3L_{m}\otimes{\mathbf{C}}=V_{m\lambda_{3}}. But now, from the explicit formula from Theorem 1.2.1, it is clear that if βI|ΓI|Lm\beta_{I}\in|\Gamma_{I}|L_{m}, then Θ(βI)\Theta(\beta_{I}) has integral Fourier coefficients. One makes a similar argument for Θ(βE)\Theta(\beta_{E}). This proves the corollary. ∎

We now explain the proof of Corollary 1.2.3. To do so, we first construct a special βmVmλ3\beta_{m}\in V_{m\lambda_{3}}. Thus let

  • KK be an imaginary quadratic field, so that HKH\otimes K is split, such as K=𝐐(1)K={\mathbf{Q}}(\sqrt{-1}).

  • a2ΘKa_{2}\in\Theta\otimes K, with a20a_{2}\neq 0 but n(a2)=0n(a_{2})=0

  • a3ΘKa_{3}\in\Theta\otimes K with n(a3)=1n(a_{3})=-1

  • a2ΘKa_{2}^{\prime}\in\Theta\otimes K with n(a2)=1n(a_{2}^{\prime})=1 and (a2,a2)=1(a_{2}^{\prime},a_{2})=1.

Then, as in section 3, we set

  • x=(1a3a2a31a3a2a2a2a30)x=\left(\begin{array}[]{ccc}1&a_{3}&a_{2}^{*}\\ a_{3}^{*}&-1&a_{3}^{*}a_{2}^{*}\\ a_{2}&a_{2}a_{3}&0\end{array}\right)

  • z=(00(a2)010a200)z=\left(\begin{array}[]{ccc}0&0&(a_{2}^{\prime})^{*}\\ 0&1&0\\ a_{2}^{\prime}&0&0\end{array}\right)

  • and y=z×x=(0(a2)(a2a3)1a3(a2)a2a21)y=z\times x=\left(\begin{array}[]{ccc}0&(a_{2}^{\prime})^{*}(a_{2}a_{3})&*\\ &-1&a_{3}^{*}(a_{2}^{\prime})^{*}\\ a_{2}^{\prime}-a_{2}&*&1\end{array}\right).

We set βK,m=(xy)m\beta_{K,m}=(x\wedge y)^{\otimes m}. It is proved in section 3 that βK,mVmλ3\beta_{K,m}\in V_{m\lambda_{3}}.

We require the following lemma.

Lemma 5.0.4.

Let the notation be as above, with m>0m>0 even. Set w0=u2vuv2w_{0}=u^{2}v-uv^{2}. Then

(wWJR:rk(w)=1,prI(w)=w0Pm(w),βK,mI)=6.\left(\sum_{w\in W_{J_{R}}:rk(w)=1,pr_{I}(w)=w_{0}}{\langle P_{m}(w),\beta_{K,m}\rangle_{I}}\right)=6.
Proof.

As explained in [Pol20b, proof of Corollary 2.5.2], it follows from [EG96, Proposition 5.5] that the set of wWJRw\in W_{J_{R}}, with rk(w)=1,prI(w)=w0rk(w)=1,pr_{I}(w)=w_{0} consists of the six elements (0,eii,ejj,0)(0,e_{ii},-e_{jj},0), iji\neq j, where ekke_{kk} is the diagonal matrix in JJ with a 11 in the kthk^{th} place and 0’s elsewhere. Now observe that

  1. (1)

    e22e33,xyI=1\langle e_{22}\wedge e_{33},x\wedge y\rangle_{I}=-1

  2. (2)

    e33e11,xyI=1\langle e_{33}\wedge e_{11},x\wedge y\rangle_{I}=-1

  3. (3)

    e11e22,xyI=1\langle e_{11}\wedge e_{22},x\wedge y\rangle_{I}=-1

The lemma follows directly. ∎

Proof of Corollary 1.2.3.

The corollaries now follow immediately from Theorem 1.2.1 and Lemma 5.0.4.∎

We now explain the proof of Corollary 1.2.4.

Proof of Corollary 1.2.4.

First note that the cubic ring 𝐙×𝐙[t]/(t2ptq){\mathbf{Z}}\times{\mathbf{Z}}[t]/(t^{2}-pt-q) is associated to the binary cubic form y(x2+pxy+qy2)-y(x^{2}+pxy+qy^{2}). Indeed, setting ω=(1,0)\omega=(1,0) and θ=(0,t)\theta=(0,t) in 𝐙×𝐙[t]/(t2ptq){\mathbf{Z}}\times{\mathbf{Z}}[t]/(t^{2}-pt-q), this is a good basis, and computing its multiplication table gives rise to the binary cubic form y(x2+pxy+qy2)-y(x^{2}+pxy+qy^{2}). The ring 𝐙×𝐙D{\mathbf{Z}}\times{\mathbf{Z}}_{D} is of this form with p=Dp=D and q=DD24q=\frac{D-D^{2}}{4}.

To explicitly compute the Fourier coefficients of ΔG2\Delta_{G_{2}}, we now make a specific choice of x,z,yx,z,y as in the proof of Corollary 1.2.3. Namely, we take K=𝐐(1)K={\mathbf{Q}}(\sqrt{-1}), a2=e2a_{2}=e_{2}, a3=u0a_{3}=u_{0}, and a2=e2e2a_{2}^{\prime}=e_{2}-e_{2}^{*}. We obtain

x=(1x3x2x31x1x2x10),y=(0y3y2y31y1y2y11)x=\left(\begin{array}[]{ccc}1&x_{3}&x_{2}^{*}\\ x_{3}^{*}&-1&x_{1}\\ x_{2}&x_{1}^{*}&0\end{array}\right),\,\,\,\,y=\left(\begin{array}[]{ccc}0&y_{3}&y_{2}^{*}\\ y_{3}^{*}&-1&y_{1}\\ y_{2}&y_{1}^{*}&1\end{array}\right)

with

  • x1=12(0,11i)x_{1}=\frac{1}{2}(0,1-\sqrt{-1}i)

  • x2=12(0,11i)x_{2}=\frac{1}{2}(0,1-\sqrt{-1}i)

  • x3=1(i,0)x_{3}=-\sqrt{-1}(i,0)

  • y1=1(0,i)y_{1}=-\sqrt{-1}(0,i)

  • y2=12(0,1+1i)y_{2}=\frac{1}{2}(0,1+\sqrt{-1}i)

  • y3=12(1+1i,0)y_{3}=-\frac{1}{2}(1+\sqrt{-1}i,0).

Now, given the binary cubic y(x2+pxy+qy2)-y(x^{2}+pxy+qy^{2}), in order to compute the associated Fourier coefficient of ΔG2\Delta_{G_{2}}, we must compute the set of (0,T1,T2,q)WJR(0,T_{1},T_{2},q)\in W_{J_{R}} so that tr(T2)=p\operatorname{tr}(T_{2})=p, tr(T1)=1\operatorname{tr}(T_{1})=-1, and (0,T1,T2,q)(0,T_{1},T_{2},q) rank one. We assume q0q\neq 0. Then (0,T1,T2,q)(0,T_{1},T_{2},q) is rank one if and only if n(T2)=0n(T_{2})=0 and T2#=qT1T_{2}^{\#}=qT_{1}, which implies that T1T_{1} is rank one. As mentioned above, the set of rank one T1T_{1} in JRJ_{R} with tr(T1)=1\operatorname{tr}(T_{1})=-1 consists just of the three elements e11,e22,e33-e_{11},-e_{22},-e_{33}.

Suppose that T1=e11T_{1}=-e_{11}. Then 0=(T1,T2)0=(T_{1},T_{2}) because (T2,T2#)=3n(T2)=0(T_{2},T_{2}^{\#})=3n(T_{2})=0. So, the (1,1)(1,1) entry of T2=0T_{2}=0. It now follows easily, using that T2#=qe11T_{2}^{\#}=-qe_{11}, that T2T_{2} is of the form

T2=(0000c2x0xc3)T_{2}=\left(\begin{array}[]{ccc}0&0&0\\ 0&c_{2}&x\\ 0&x^{*}&c_{3}\end{array}\right)

with c2c3n(x)=qc_{2}c_{3}-n(x)=-q and c2+c3=pc_{2}+c_{3}=p. Substituting p=Dp=D, q=DD24q=\frac{D-D^{2}}{4}, c2=12(v+D)c_{2}=\frac{1}{2}(v+D), c3=12(v+D)c_{3}=\frac{1}{2}(-v+D), one finds that vv is an integer and v2+4n(x)=Dv^{2}+4n(x)=D. The contribution to the Fourier coefficient aΔG2(𝐙×𝐙D)a_{\Delta_{G_{2}}}({\mathbf{Z}}\times{\mathbf{Z}}_{D}) for such a term is computed to be (v+1(x,(0,i)))2(v+\sqrt{-1}(x,(0,i)))^{2}.

One makes a completely similar calculation if T1=e22T_{1}=-e_{22} or T1=e33T_{1}=-e_{33}. One obtains

aΔG2(𝐙×𝐙D)=[v+1(x,(0,i))]2+[v1(x,(0,i))]2+[v+1(x,(i,0))]2a_{\Delta_{G_{2}}}({\mathbf{Z}}\times{\mathbf{Z}}_{D})=\sum{[v+\sqrt{-1}(x,(0,i))]^{2}+[v-\sqrt{-1}(x,(0,i))]^{2}+[v+\sqrt{-1}(x,(-i,0))]^{2}}

where the sum is over pairs (v,x)(v,x) in 𝐙RΘ{\mathbf{Z}}\oplus R_{\Theta} such that v2+4n(x)=Dv^{2}+4n(x)=D.

But, this is clearly the Fourier coefficient of a harmonic theta function in S13/2(Γ0(4))+S_{13/2}(\Gamma_{0}(4))^{+}, associated to the lattice 22E8\langle 2\rangle\perp 2E_{8}. This space has dimension 11, spanned by δ(z)\delta(z). By looking at the coefficient of D=1D=1, one obtains the corollary. ∎

5.1. Theta lifts

In this subsection, we complete the proof of Theorem 1.2.1. More specifically, we will now use our knowledge of the computation of {D2mΘ(g,1),β}\{D^{2m}\Theta(g,1),\beta\} for arbitrary β\beta to compute {D2mΘ(g,γE),β}\{D^{2m}\Theta(g,\gamma_{E}),\beta^{\prime}\}.

Let Θ=ΘGan\Theta=\Theta_{Gan} and denote by Θw\Theta_{w} the ww-Fourier coefficient of this automorphic form. Suppose gGJ(𝐑)g_{\infty}\in G_{J}({\mathbf{R}}). To begin, observe that

a(w)(γE)W2πw(g)\displaystyle a(w)(\gamma_{E})W_{2\pi w}(g_{\infty}) =Θw(gγE)\displaystyle=\Theta_{w}(g_{\infty}\gamma_{E})
=Θw(γEg)\displaystyle=\Theta_{w}(\gamma_{E}g_{\infty})
=Θw(δE𝐐δE𝐑,1δE𝐙^g)\displaystyle=\Theta_{w}(\delta_{E}^{\mathbf{Q}}\delta_{E}^{{\mathbf{R}},-1}\delta_{E}^{\widehat{{\mathbf{Z}}}}g_{\infty})
=ΘδE𝐐,1w(δE𝐑,1g)\displaystyle=\Theta_{\delta_{E}^{{\mathbf{Q}},-1}w}(\delta_{E}^{{\mathbf{R}},-1}g_{\infty})
=a(δE𝐐,1w)W2πwδE𝐐(δE𝐑,1g)\displaystyle=a(\delta_{E}^{{\mathbf{Q}},-1}w)W_{2\pi w\cdot\delta_{E}^{{\mathbf{Q}}}}(\delta_{E}^{{\mathbf{R}},-1}g_{\infty})
=a(δE𝐐,1w)W2πw(g).\displaystyle=a(\delta_{E}^{{\mathbf{Q}},-1}w)W_{2\pi w}(g_{\infty}).

Now, if gG2(𝐑)g\in G_{2}({\mathbf{R}}), we prove

{D2mWw(g),β}=Pm(w),βWw0,4+m(g)\{D^{2m}W_{w}(g),\beta\}=\langle P_{m}(w),\beta\rangle W_{w_{0},4+m}(g)

where w0w_{0} is the binary cubic prI(w)=au3+(b,I#)u2v+(c,I)uv2+dv3pr_{I}(w)=au^{3}+(b,I^{\#})u^{2}v+(c,I)uv^{2}+dv^{3} if w=(a,b,c,d)w=(a,b,c,d). Observe that prI(δQw)=prE(w)pr_{I}(\delta_{Q}w)=pr_{E}(w), where prE(w)=au3+(b,E#)u2v+(c,E)uv2+dv3pr_{E}(w)=au^{3}+(b,E^{\#})u^{2}v+(c,E)uv^{2}+dv^{3}. Combining these two facts, one obtains

{D2mΘ(g,γE),β}\displaystyle\{D^{2m}\Theta(g,\gamma_{E}),\beta\} =rk(w)=1a(δE𝐐,1w)Pm(w),βWprI(w)(g)\displaystyle=\sum_{rk(w)=1}a(\delta_{E}^{{\mathbf{Q}},-1}w)\langle P_{m}(w),\beta\rangle W_{pr_{I}(w)}(g)
=rk(w)=1a(w)Pm(δE𝐐w),βWprE(w)(g)\displaystyle=\sum_{rk(w)=1}a(w)\langle P_{m}(\delta_{E}^{\mathbf{Q}}w),\beta\rangle W_{pr_{E}(w)}(g)
=rk(w)=1a(w)Pm(w),δE𝐐,1βEWprE(w)(g).\displaystyle=\sum_{rk(w)=1}a(w)\langle P_{m}(w),\delta_{E}^{{\mathbf{Q}},-1}\beta\rangle_{E}W_{pr_{E}(w)}(g).

Here, if bJ,cJJb\in J,c\in J^{\vee}\simeq J, x,yJx,y\in J then

bc,xyE=(b,x)E(c,y)(b,y)E(c,x)\langle b\wedge c,x\wedge y\rangle_{E}=(b,x)_{E}(c,y)-(b,y)_{E}(c,x)

and one extends to (2J)m(\wedge^{2}J)^{\otimes m}. We have used Lemma 3.2.2 to get

Pw(δE𝐐w),βI=Pm(w),δE𝐐,1βE.\langle P_{w}(\delta_{E}^{\mathbf{Q}}w),\beta\rangle_{I}=\langle P_{m}(w),\delta_{E}^{{\mathbf{Q}},-1}\beta\rangle_{E}.

Thus, putting everything together,

Θ(α)(g)\displaystyle\Theta(\alpha)(g) :=[F4I]{D2mΘ(g,h),α(h)}𝑑h\displaystyle:=\int_{[F_{4}^{I}]}\{D^{2m}\Theta(g,h),\alpha(h)\}\,dh
=1|ΓI|{D2mΘ(g,1),α(1)}+1|ΓE|{D2mΘ(g,γE),α(γE)}\displaystyle=\frac{1}{|\Gamma_{I}|}\{D^{2m}\Theta(g,1),\alpha(1)\}+\frac{1}{|\Gamma_{E}|}\{D^{2m}\Theta(g,\gamma_{E}),\alpha(\gamma_{E})\}
=1|ΓI|rk(w)=1a(w)Pm(w),αIIWprI(w)(g)+1|ΓE|rk(w)=1a(w)Pm(w),αEEWprE(w)(g).\displaystyle=\frac{1}{|\Gamma_{I}|}\sum_{rk(w)=1}{a(w)\langle P_{m}(w),\alpha_{I}\rangle_{I}W_{pr_{I}(w)}(g)}+\frac{1}{|\Gamma_{E}|}\sum_{rk(w)=1}{a(w)\langle P_{m}(w),\alpha_{E}\rangle_{E}W_{pr_{E}(w)}(g)}.

This completes the proof of Theorem 1.2.1.

6. The exponential derivative

The purpose of this section is to prove Theorem 4.0.1, which we restate here:

Theorem 6.0.1.

Suppose gSp6(𝐑)g\in\operatorname{Sp}_{6}({\mathbf{R}}) and βW(k1,k2)\beta\in W(k_{1},k_{2}). Let 0\ell\geq 0 be an integer. There is a nonzero constant Bk1,k2B_{k_{1},k_{2}}, independent of gg and TT, so that

j(g,i)ρ[k1,k2](J(g,i)){Dk1+2k2(j(g,i)e2πi(T,gi)),β}=Bk1,k2{Pk1,k2(T),β}e2πi(pr(T),gi).j(g,i)^{\ell}\rho_{[k_{1},k_{2}]}(J(g,i))\{D^{k_{1}+2k_{2}}(j(g,i)^{-\ell}e^{2\pi i(T,g\cdot i)}),\beta\}=B_{k_{1},k_{2}}\{P_{k_{1},k_{2}}(T),\beta\}e^{2\pi i(pr(T),g\cdot i)}.

Moreover, {Pk1,k2(T),β}\{P_{k_{1},k_{2}}(T),\beta\} lies in the highest weight submodule S[k1,k2]S^{[k_{1},k_{2}]} of Sk1(V3)Sk2(2V3)S^{k_{1}}(V_{3})\otimes S^{k_{2}}(\wedge^{2}V_{3}).

Our proof of this theorem is to, essentially completely explicitly, calculate the derivatives

Xα1Xαn(j(g,i)e2πi(T,gi)).X_{\alpha_{1}}\cdots X_{\alpha_{n}}(j(g,i)^{-\ell}e^{2\pi i(T,g\cdot i)}). (2)

To do this, we use the Iwasawa decomposition 𝔥J=n(J)+𝔪J+𝔨E7{\mathfrak{h}}_{J}=n(J)+{\mathfrak{m}}_{J}+{\mathfrak{k}}_{E_{7}}, write each XαX_{\alpha} as a sum in terms of this decomposition, and calculate the derivatives for each piece.

6.1. Preliminaries

Recall from above the Cayley transform ChHJ(𝐂)C_{h}\in H_{J}({\mathbf{C}}), which satisfies

  1. (1)

    Ch1nL(J𝐂)Ch=𝔭J+C_{h}^{-1}n_{L}(J\otimes{\mathbf{C}})C_{h}={\mathfrak{p}}_{J}^{+}

  2. (2)

    Ch1nL(J𝐂)Ch=𝔭JC_{h}^{-1}n_{L}^{\vee}(J\otimes{\mathbf{C}})C_{h}={\mathfrak{p}}_{J}^{-}

  3. (3)

    Ch1(𝔪J𝐂)Ch=𝔨E7C_{h}^{-1}({\mathfrak{m}}_{J}\otimes{\mathbf{C}})C_{h}={\mathfrak{k}}_{E_{7}}.

Given ϕ𝔪J\phi\in{\mathfrak{m}}_{J}, let M(ϕ)M(\phi) denote its action on WJW_{J} [Pol20a, section 3.4].

Proposition 6.1.1.

One has the following identities:

  1. (1)

    Ch1nL(x)Ch=nL(x)i2M(Φ1,x)+12(nL(x)nL(x))C_{h}^{-1}n_{L}(x)C_{h}=n_{L}(x)-\frac{i}{2}M(\Phi_{1,x})+\frac{1}{2}(n_{L}^{\vee}(x)-n_{L}(x))

  2. (2)

    Ch1nL(γ)Ch=nL(γ)+i2M(Φ1,γ)+12(nL(γ)nL(γ))C_{h}^{-1}n_{L}^{\vee}(\gamma)C_{h}=n_{L}(\gamma)+\frac{i}{2}M(\Phi_{1,\gamma})+\frac{1}{2}(n_{L}^{\vee}(\gamma)-n_{L}(\gamma))

  3. (3)

    If ϕ(1)=0\phi(1)=0, then Ch1M(ϕ)Ch=M(ϕ)C_{h}^{-1}M(\phi)C_{h}=M(\phi).

  4. (4)

    If ϕ=Φ1,z\phi=\Phi_{1,z}, then Ch1M(ϕ)Ch=inL(z)+inL(z)C_{h}^{-1}M(\phi)C_{h}=-in_{L}(z)+in_{L}^{\vee}(z).

Proof.

Some of the facts needed to prove this are in Lemma 3.4.1 of [Pol20a]. Also useful is the identity ΦY,1=Φ1,Y\Phi_{Y,1}=\Phi_{1,Y} (see Theorem 4.0.10 in [Pol21]). To help with the proof of the second statement, one computes that

nG(δ)M(ϕ)nG(δ)=M(ϕ)nL(ϕ~(δ)).n_{G}^{\vee}(\delta)M(\phi)n_{G}^{\vee}(-\delta)=M(\phi)-n_{L}^{\vee}(\widetilde{\phi}(\delta)).

We omit the rest of the proof. ∎

6.2. Some computations

For EJ𝐂E\in J\otimes{\mathbf{C}}, define X+(E)=iCh1nL(E)ChX^{+}(E)=iC_{h}^{-1}n_{L}(E)C_{h}. To warm up, we will compute X+(E)(j(g,i)e2πi(T,gi))X^{+}(E)(j(g,i)^{-\ell}e^{2\pi i(T,g\cdot i)}). Let hh denote the function h(g)=j(g,i)e2πi(T,gi)h(g)=j(g,i)^{-\ell}e^{2\pi i(T,g\cdot i)}. Let mYm_{Y} denote any element of MJM_{J} for which mYi=YiJm_{Y}\cdot i=Yi\in\mathcal{H}_{J}.

Lemma 6.2.1.

One has the following computations:

  1. (1)

    nL(E)h(M(δ,m))=2πi(T,m(E))h(M(δ,m))n_{L}(E)h(M(\delta,m))=2\pi i(T,m(E))h(M(\delta,m));

  2. (2)

    M(Φ1,E)e2π(T,m(1))=4π(T,m(E))e2π(T,m(1))M(\Phi_{1,E})e^{-2\pi(T,m(1))}=-4\pi(T,m(E))e^{-2\pi(T,m(1))};

  3. (3)

    M(Φ1,E)j(mY,i)=(1,E)n(Y)/2=(1,E)j(mY,i)M(\Phi_{1,E})j(m_{Y},i)^{-\ell}=\ell(1,E)n(Y)^{\ell/2}=\ell(1,E)j(m_{Y},i)^{-\ell};

  4. (4)

    if k=nL(E)nL(E)k=n_{L}^{\vee}(E)-n_{L}(E), then kj(g,i)=i(1,E)j(g,i)kj(g,i)^{-\ell}=-\ell i(1,E)j(g,i)^{-\ell}.

Proof.

The first statement follows from the identity M(δ,m)nL(E)M(δ,m)1=nL(m(E))M(\delta,m)n_{L}(E)M(\delta,m)^{-1}=n_{L}(m(E)). The second statement follows from the fact that Φ1,E(E)={E,E}\Phi_{1,E}(E^{\prime})=\{E,E^{\prime}\}, so that Φ1,E(1)=2E\Phi_{1,E}(1)=2E. For the third statement, we have

M(Φ1,E)j(mY,i)=ddt|t=0(mYexp(tΦ1,E)r0(i),f).M(\Phi_{1,E})j(m_{Y},i)^{-\ell}=\frac{d}{dt}|_{t=0}(\langle m_{Y}\exp(t\Phi_{1,E})r_{0}(i),f\rangle)^{-\ell}.

Now M(Φ1,E)r1(i)=((1,E),)M(\Phi_{1,E})r_{1}(i)=(-(1,E),\ldots) so M(Φ1,E)j(mY,i)M(\Phi_{1,E})j(m_{Y},i)^{-\ell} is the coefficient in tt of (n(Y)1/2(1(1,E)t))(n(Y)^{-1/2}(1-(1,E)t))^{-\ell}, which is n(Y)/2(1,E)n(Y)^{\ell/2}\cdot\ell(1,E). Thus M(Φ1,E)j(mY,i)=(1,E)n(Y)/2M(\Phi_{1,E})j(m_{Y},i)^{-\ell}=\ell(1,E)n(Y)^{\ell/2} as claimed. For the fourth statement, observe that nL(E)r1(i)=(i(1,E),)n_{L}^{\vee}(E)r_{1}(i)=(i(1,E),\ldots), nL(E)r1(i)=(0,)n_{L}(E)r_{1}(i)=(0,\ldots), so etkr1(i)=(1+i(1,E)t,)+O(t2)e^{tk}r_{1}(i)=(1+i(1,E)t,\ldots)+O(t^{2}). As j(getk,i)=j(g,i)j(etk,i)j(ge^{tk},i)=j(g,i)j(e^{tk},i), the statement follows. ∎

Putting everything together, from Lemma 6.2.1 and Proposition 6.1.1 one obtains

(Ch1nL(E)Ch)h(m)\displaystyle(C_{h}^{-1}n_{L}(E)C_{h})h(m) =2πi(T,m(E))h(m)+(i/2)(4π(T,m(E))+(1,E))h(m)\displaystyle=2\pi i(T,m(E))h(m)+(-i/2)(-4\pi(T,m(E))+\ell(1,E))h(m)
+12(i(1,E))h(m)\displaystyle\,+\frac{1}{2}(-\ell i(1,E))h(m)
=(4πi(T,m(E))i(1,E))h(m)\displaystyle=(4\pi i(T,m(E))-\ell i(1,E))h(m)
(Ch1nL(E)Ch)h(m)\displaystyle(C_{h}^{-1}n_{L}^{\vee}(E)C_{h})h(m) =2πi(T,m(E))h(m)+(i/2)(4π(T,m(E))+(1,E))h(m)\displaystyle=2\pi i(T,m(E))h(m)+(i/2)(-4\pi(T,m(E))+\ell(1,E))h(m)
+12(i(1,E))h(m)\displaystyle\,+\frac{1}{2}(-\ell i(1,E))h(m)
=0.\displaystyle=0.

Thus

XE+h(m)=(4π(T,m(E))+(1,E))h(m).X^{+}_{E}h(m)=(-4\pi(T,m(E))+\ell(1,E))h(m).

We now explain the computation of (2). To setup the result, define Pk:JkT(J)P_{k}:J^{\otimes k}\rightarrow T(J) inductively as follows: P0=1P_{0}=1, and for k0k\geq 0,

Pk+1(E1,,Ek,Ek+1)\displaystyle P_{k+1}(E_{1},\ldots,E_{k},E_{k+1}) =Pk(E1,Ek)Ek+1+(1,Ek+1)Pk(E1,,Ek)+12{Ek+1,Pk(E1,Ek)}\displaystyle=P_{k}(E_{1},\ldots E_{k})\otimes E_{k+1}+\ell(1,E_{k+1})P_{k}(E_{1},\ldots,E_{k})+\frac{1}{2}\{E_{k+1},P_{k}(E_{1},\ldots E_{k})\}
+12Pk({Ek+1,E1,,Ek}).\displaystyle\,+\frac{1}{2}P_{k}(\{E_{k+1},E_{1},\ldots,E_{k}\}).

Here we write

{E,V1Vr}:=j=1rV1{E,Vj}Vr\{E,V_{1}\otimes\cdots\otimes V_{r}\}:=\sum_{j=1}^{r}V_{1}\otimes\cdots\otimes\{E,V_{j}\}\otimes\cdots\otimes V_{r}

and we interpret {E,C}=0\{E,C\}=0 if CT0(J)C\in T^{0}(J) is constant. Thus

  1. (1)

    P1(E1)=E1+(1,E1)P_{1}(E_{1})=E_{1}+\ell(1,E_{1}).

  2. (2)

    P2(E1,E2)=E1E2+(1,E2)E1+(1,E1)E2+{E1,E2}+2(1,E1)(1,E2)+2(1,{E1,E2}).P_{2}(E_{1},E_{2})=E_{1}\otimes E_{2}+\ell(1,E_{2})E_{1}+\ell(1,E_{1})E_{2}+\{E_{1},E_{2}\}+\ell^{2}(1,E_{1})(1,E_{2})+\frac{\ell}{2}(1,\{E_{1},E_{2}\}).

Define now wT,m:T(J)𝐂w_{T,m}:T(J)\rightarrow{\mathbf{C}} as

wT,m(V1Vr)=(4π)r(T,m(V1))(T,m(Vr))w_{T,m}(V_{1}\otimes\cdots\otimes V_{r})=(-4\pi)^{r}(T,m(V_{1}))\cdots(T,m(V_{r}))

and extending to T(J)T(J) by linearity.

Proposition 6.2.2.

Let the notation be as above. Then XEkXE1h(m)=wT,m(P(E1,,Ek))h(m)X_{E_{k}}\cdots X_{E_{1}}h(m)=w_{T,m}(P(E_{1},\ldots,E_{k}))h(m).

Proof.

We proceed by induction. Observe that

  1. (1)

    nL(Ek+1)XEkXE1h(m)=2πi(T,mEk+1)XEkXE1h(m)n_{L}(E_{k+1})X_{E_{k}}\cdots X_{E_{1}}h(m)=2\pi i(T,mE_{k+1})X_{E_{k}}\cdots X_{E_{1}}h(m).

  2. (2)

    M(Φ1,E)wT,m(V1Vr)=wT,m({E,V1Vr})M(\Phi_{1,E})w_{T,m}(V_{1}\otimes\cdots\otimes V_{r})=w_{T,m}(\{E,V_{1}\otimes\cdots\otimes V_{r}\}).

  3. (3)

    M(Φ1,E)h(m)=((1,E)4π(T,m(E)))h(m)M(\Phi_{1,E})h(m)=(\ell(1,E)-4\pi(T,m(E)))h(m).

  4. (4)

    If μK\mu\in K, then μXEkXE1h(m)=j(μ,i)XμEkXμE1h(m)\mu X_{E_{k}}\cdots X_{E_{1}}h(m)=j(\mu,i)^{-\ell}X_{\mu\cdot E_{k}}\cdots X_{\mu\cdot E_{1}}h(m).

Now, if k=nL(E)nL(E)k=n_{L}^{\vee}(E)-n_{L}(E), then

[k,Ch1nL(E)Ch]\displaystyle[k,C_{h}^{-1}n_{L}(E^{\prime})C_{h}] =[iCh1M(Φ1,E)Ch,Ch1nL(E)Ch]\displaystyle=[-iC_{h}^{-1}M(\Phi_{1,E})C_{h},C_{h}^{-1}n_{L}(E^{\prime})C_{h}]
=iCh1[M(Φ1,E),nL(E)]Ch\displaystyle=-iC_{h}^{-1}[M(\Phi_{1,E}),n_{L}(E^{\prime})]C_{h}
=iCh1nL({E,E})Ch.\displaystyle=-iC_{h}^{-1}n_{L}(\{E,E^{\prime}\})C_{h}.

Thus [ik/2,XE]=12X{E,E}[ik/2,X_{E^{\prime}}]=\frac{1}{2}X_{\{E,E^{\prime}\}}. Consequently,

(ik/2)(XEkXE1)h(m)=2(1,E)(XEkXE1)h(m)+12(j=1kXEkX{E,Ej}XE1)h(m).(ik/2)(X_{E_{k}}\cdots X_{E_{1}})h(m)=\frac{\ell}{2}(1,E)(X_{E_{k}}\cdots X_{E_{1}})h(m)+\frac{1}{2}(\sum_{j=1}^{k}X_{E_{k}}\cdots X_{\{E,E_{j}\}}\cdots X_{E_{1}})h(m).

As

XE=iCh1nL(E)Ch=inL(E)+12M(Φ1,E)+i2(nL(E)nL(E)),X_{E}=iC_{h}^{-1}n_{L}(E)C_{h}=in_{L}(E)+\frac{1}{2}M(\Phi_{1,E})+\frac{i}{2}(n_{L}^{\vee}(E)-n_{L}(E)),

one can now easily verify the proposition. ∎

We now prove:

Lemma 6.2.3.

Let Pn:JkT(J)P_{n}:J^{\otimes k}\rightarrow T(J) be as above and let βW(k1,k2)V7k1(2V7)k2\beta\in W(k_{1},k_{2})\subseteq V_{7}^{\otimes k_{1}}\otimes(\wedge^{2}V_{7})^{\otimes k_{2}}. Set n=k1+2k2n=k_{1}+2k_{2}. Then

αiPn(Eα1,,Eαn){Eα1Eαn,β}=αiEα1Eαn{Eα1Eαn,β}.\sum_{\alpha_{i}}{P_{n}(E_{\alpha_{1}},\ldots,E_{\alpha_{n}})\{E_{\alpha_{1}}^{\vee}\otimes\cdots\otimes E_{\alpha_{n}}^{\vee},\beta\}}=\sum_{\alpha_{i}}{E_{\alpha_{1}}\otimes\cdots\otimes E_{\alpha_{n}}\{E_{\alpha_{1}}^{\vee}\otimes\cdots\otimes E_{\alpha_{n}}^{\vee},\beta\}}. (3)

In other words, only the leading term contributes.

Proof.

Suppose hG2(𝐑)h\in G_{2}({\mathbf{R}}). Observe

αiPn(Eα1,,Eαn){Eα1Eαn,hβ}\displaystyle\sum_{\alpha_{i}}P_{n}(E_{\alpha_{1}},\ldots,E_{\alpha_{n}})\{E_{\alpha_{1}}^{\vee}\otimes\cdots\otimes E_{\alpha_{n}}^{\vee},h\cdot\beta\} =αiPn(Eα1,,Eαn){h1Eα1h1Eαn,β}\displaystyle=\sum_{\alpha_{i}}P_{n}(E_{\alpha_{1}},\ldots,E_{\alpha_{n}})\{h^{-1}E_{\alpha_{1}}^{\vee}\otimes\cdots\otimes h^{-1}E_{\alpha_{n}}^{\vee},\beta\}
=αiPn(hEα1,,hEαn){Eα1Eαn,β}\displaystyle=\sum_{\alpha_{i}}P_{n}(hE_{\alpha_{1}},\ldots,hE_{\alpha_{n}})\{E_{\alpha_{1}}^{\vee}\otimes\cdots\otimes E_{\alpha_{n}}^{\vee},\beta\}
=αihPn(Eα1,,Eαn){Eα1Eαn,β}\displaystyle=\sum_{\alpha_{i}}h\cdot P_{n}(E_{\alpha_{1}},\ldots,E_{\alpha_{n}})\{E_{\alpha_{1}}^{\vee}\otimes\cdots\otimes E_{\alpha_{n}}^{\vee},\beta\}

where we have used that PnP_{n} is equivariant for the action of F4G2F_{4}\supseteq G_{2}. (This follows, for example, from the recursive formula.)

Now, because of the above equivariance, we can assume β=e2k1(e2e3)k2\beta=e_{2}^{k_{1}}\otimes(e_{2}\wedge e_{3}^{*})^{k_{2}}. Indeed, the two sides of the equality to be proved are linear in β\beta and e2k1(e2e3)k2W(k1,k2)e_{2}^{k_{1}}\otimes(e_{2}\wedge e_{3}^{*})^{k_{2}}\in W(k_{1},k_{2}).

We now compute in the basis {u0,e1,e2,e3,e1,e2,e3}\{u_{0},e_{1},e_{2},e_{3},e_{1}^{*},e_{2}^{*},e_{3}^{*}\} of Θ0\Theta^{0}. More precisely, we take a basis of JJ that is a basis of H3(F)H_{3}(F) union a basis viujv_{i}\otimes u_{j} of V3Θ0V_{3}\otimes\Theta^{0}, where uj{u0,e1,e2,e3,e1,e2,e3}u_{j}\in\{u_{0},e_{1},e_{2},e_{3},e_{1}^{*},e_{2}^{*},e_{3}^{*}\} and v1,v2,v3v_{1},v_{2},v_{3} is a basis of V3V_{3}. We compute in this basis. Then the only terms that contribute to the left-hand side of (6.2) are those with Eαi=vαiuαiE_{\alpha_{i}}=v_{\alpha_{i}}\otimes u_{\alpha_{i}} for all ii. Then, still in order for these terms to contribute in a nonzero way to (6.2), we must have uk=e2u_{k}=e_{2} for 1kk11\leq k\leq k_{1} and {uα2j1,uα2j}={e2,e3}\{u_{\alpha_{2j-1}},u_{\alpha_{2j}}\}=\{e_{2},e_{3}^{*}\} for 2j1>k12j-1>k_{1}. Consequently, all the EαiE_{\alpha_{i}} that contribute to the sum in our basis satisfy (1,Eαi)=0(1,E_{\alpha_{i}})=0, Eαi#=0E_{\alpha_{i}}^{\#}=0, and {Eαi,Eαj}=0\{E_{\alpha_{i}},E_{\alpha_{j}}\}=0. It now follows from our recursive formula for PnP_{n} that only the leading term contributes. ∎

It follows immediately from Lemma 6.2.3 that

αiwT,m(Pn(Eα1,,Eαn)){Eα1Eαn,β}\displaystyle\sum_{\alpha_{i}}{w_{T,m}(P_{n}(E_{\alpha_{1}},\ldots,E_{\alpha_{n}}))\{E_{\alpha_{1}}^{\vee}\otimes\cdots\otimes E_{\alpha_{n}}^{\vee},\beta\}} =αiwT,m(Eα1Eαn){Eα1Eαn,β}\displaystyle=\sum_{\alpha_{i}}{w_{T,m}(E_{\alpha_{1}}\otimes\cdots\otimes E_{\alpha_{n}})\{E_{\alpha_{1}}^{\vee}\otimes\cdots\otimes E_{\alpha_{n}}^{\vee},\beta\}}
=(4π)n{m~1(Tn),β}.\displaystyle=(-4\pi)^{n}\{\widetilde{m}^{-1}(T^{\otimes n}),\beta\}.

Combining this with Proposition 6.2.2, we obtain

Proposition 6.2.4.

Suppose βW(k1,k2)\beta\in W(k_{1},k_{2}), n=k1+2k2n=k_{1}+2k_{2} and mMJm\in M_{J}. Then

{Dn(j(m,i)e2π(T,m(1))),β}=Bk1,k2j(m,i){m~1(Tn),β}e2π(T,m(1))\{D^{n}(j(m,i)^{-\ell}e^{-2\pi(T,m(1))}),\beta\}=B_{k_{1},k_{2}}j(m,i)^{-\ell}\{\widetilde{m}^{-1}(T^{\otimes n}),\beta\}e^{-2\pi(T,m(1))}

for a nonzero constant Bk1,k2.B_{k_{1},k_{2}}.

We can now prove Theorem 4.0.1.

Proof of Theorem 4.0.1.

Observe that by appropriate KSp6K_{\operatorname{Sp}_{6}}-equivariance, and by equivariance for the unipotent radical of the Siegel parabolic, it suffices to check the equality of the statement of Theorem 4.0.1 when gMg\in M, the Levi of the Siegel parabolic. But this follows from Proposition 6.2.4 and the definition of ρ[k1,k2](J(m,i))\rho_{[k_{1},k_{2}]}(J(m,i)); see the proof of Lemma 2.2.5 for the action of m~1\widetilde{m}^{-1} on TT. The proof of the theorem now follows from Lemma 6.2.5 below.∎

Recall that S[k1,k2]S^{[k_{1},k_{2}]} is the kernel of the contraction

Sk1(V3)Sk2(2V3)Sk11(V3)Sk21(2V3)det(V3).S^{k_{1}}(V_{3})\otimes S^{k_{2}}(\wedge^{2}V_{3})\rightarrow S^{k_{1}-1}(V_{3})\otimes S^{k_{2}-1}(\wedge^{2}V_{3})\otimes\det(V_{3}).
Lemma 6.2.5.

If βW(k1,k2)\beta\in W(k_{1},k_{2}), then {Pk1,k2(T),β}S[k1,k2]\{P_{k_{1},k_{2}}(T),\beta\}\in S^{[k_{1},k_{2}]}.

Proof.

By equivariance and linearity, it suffices to verify the claim of the lemma for β=e2k1(e2e3)k2\beta=e_{2}^{\otimes k_{1}}\otimes(e_{2}\wedge e_{3}^{*})^{\otimes k_{2}}. Now suppose T=T0+x1v1+x2v2+x3v3T=T_{0}+x_{1}\otimes v_{1}+x_{2}\otimes v_{2}+x_{3}\otimes v_{3}. Then

(T,e2)=(x1,e2)v1+(x2,e2)v2+(x3,e2)v3(T,e_{2})=(x_{1},e_{2})v_{1}+(x_{2},e_{2})v_{2}+(x_{3},e_{2})v_{3}

and

(TT,e2e3)=(x2x3,e2e3)v2v3+(x3x1,e2e3)v3v1+(x1x2,e2e3)v1v2.(T\otimes T,e_{2}\wedge e_{3}^{*})=(x_{2}\wedge x_{3},e_{2}\wedge e_{3}^{*})v_{2}\wedge v_{3}+(x_{3}\wedge x_{1},e_{2}\wedge e_{3}^{*})v_{3}\wedge v_{1}+(x_{1}\wedge x_{2},e_{2}\wedge e_{3}^{*})v_{1}\wedge v_{2}.

Thus contracting yields the term

(x1,e2)(x2x3,e2e3)+(x2,e2)(x3x1,e2e3)+(x3,e2)(x1x2,e2e3).(x_{1},e_{2})(x_{2}\wedge x_{3},e_{2}\wedge e_{3}^{*})+(x_{2},e_{2})(x_{3}\wedge x_{1},e_{2}\wedge e_{3}^{*})+(x_{3},e_{2})(x_{1}\wedge x_{2},e_{2}\wedge e_{3}^{*}).

This is (x1x2x3,e2e2e3)=0(x_{1}\wedge x_{2}\wedge x_{3},e_{2}\wedge e_{2}\wedge e_{3}^{*})=0. The lemma follows. ∎

7. The quaternionic Whittaker derivative

The goal of this section is to prove Theorem 5.0.1, which we restate here:

Theorem 7.0.1.

Suppose =4\ell=4 and βVmλ3\beta\in V_{m\lambda_{3}} with m0m\geq 0. Then there is a nonzero constant BmB_{m} so that for all wWJw\in W_{J} rank one and gG2(𝐑)g\in G_{2}({\mathbf{R}}), one has

{D2mWw,(g),β}=BmPm(w),βIWprI(w),+m(g).\{D^{2m}W_{w,\ell}(g),\beta\}=B_{m}\langle P_{m}(w),\beta\rangle_{I}W_{pr_{I}(w),\ell+m}(g).

We begin with the following proposition.

Proposition 7.0.2.

Let βVmλ3(2J0)m\beta\in V_{m\lambda 3}\subseteq(\wedge^{2}J^{0})^{\otimes m}. Suppose =4\ell=4 and ww is rank one. Then the function Fm,β:G2(𝐑)S2m+2(V2)F_{m,\beta}:G_{2}({\mathbf{R}})\rightarrow S^{2m+2\ell}(V_{2}) defined as {D2mWw,(g),β}\{D^{2m}W_{w,\ell}(g),\beta\} is quaternionic.

Proof.

Fix a rank one wWJw\in W_{J}. Let χ=χw\chi=\chi_{w} be the character of NJN_{J} given as χ(n)=eiw,n¯\chi(n)=e^{i\langle w,\overline{n}\rangle}. Here n¯\overline{n} is the image of nn in WJNJabW_{J}\simeq N_{J}^{ab}, the abelianization of NJN_{J}. Let L:Πmin𝐂L:\Pi_{min}\rightarrow{\mathbf{C}} be the unique (up to scalar multiple) moderate growth linear functional satisfying L(nv)=χ(n)L(v)L(nv)=\chi(n)L(v) for all nN(𝐑)n\in N({\mathbf{R}}) and vΠminv\in\Pi_{min}. (Such an LL exists by a global argument: The global minimal representation has nonzero Fourier coefficients, so there is an L0L_{0} for some w0w_{0}. Now L(v):=L0(mwv)L(v):=L_{0}(m_{w}v) for an appropriate mwm_{w} is the desired functional. The uniqueness of LL follows from [Pol20a].)

Now let xjx_{j} be a basis of the minimal K=(SU(2)×E7)/μ2K=(\operatorname{SU}(2)\times E_{7})/\mu_{2}-type of Πmin\Pi_{min} and xjx_{j}^{\vee} in S8(V2)S^{8}(V_{2}) the dual basis. Note that Ww(g)=jL(gxj)xjW_{w}(g)=\sum_{j}L(gx_{j})\otimes x_{j}^{\vee}. Then E=jxjxjE=\sum_{j}x_{j}\otimes x_{j}^{\vee} is in (VminS8(V2))K(V_{min}\otimes S^{8}(V_{2}))^{K}. One obtains that D2mE(VminS8(V2)𝔭2m)KD^{2m}E\in(V_{min}\otimes S^{8}(V_{2})\otimes{\mathfrak{p}}^{\otimes 2m})^{K}. This latter space maps K:=SU(2)×SU(2)s×F4I(𝐑)K^{\prime}:=\operatorname{SU}(2)\times\operatorname{SU}(2)_{s}\times F_{4}^{I}({\mathbf{R}})-equivariantly to

S8+2m(V2)det(V2s)m(2J0)m.S^{8+2m}(V_{2})\otimes\det(V_{2}^{s})^{\otimes m}\otimes(\wedge^{2}J_{0})^{\otimes m}.

Finally, mapping (2J0)m(\wedge^{2}J^{0})^{\otimes m} to Vmλ3V_{m\lambda_{3}}, we obtain a KK^{\prime} invariant element EE^{\prime} in Vmin(S8+2m𝟏Vmλ3)V_{min}\otimes(S^{8+2m}\otimes\mathbf{1}\otimes V_{m\lambda_{3}}). By Huang-Pandzic-Savin [HPS96], this EE^{\prime} is either 0 or the minimal type of πm+4Vmλ3\pi_{m+4}\otimes V_{m\lambda_{3}}. Contracting now against some βVmλ3\beta\in V_{m\lambda_{3}}, we obtain some (possibly 0) multiple of S2m+8(V2)πm+4S^{2m+8}(V_{2})\subseteq\pi_{m+4}. Applying L(g)L(g\cdot), it follows that Fm,βF_{m,\beta} is quaternionic. ∎

7.1. General strategy

As before, let Ww,(g)W_{w,\ell}(g) be the generalized Whittaker function of weight \ell associated to wWJw\in W_{J}, which is positive semi-definite. Set Ww,(g)=(x2,Ww(g))W_{w,\ell}^{-\ell}(g)=(x^{2\ell},W_{w}(g)) be the component multiplying y2/(2)!y^{2\ell}/(2\ell)!. (See [Pol20a] for the definition of x,yx,y.) Here (,)(\,,\,) is an SL2(𝐂)\operatorname{SL}_{2}({\mathbf{C}})-equivariant pairing on S(V2)S^{\cdot}(V_{2}).

Consider the quantity {D2mWw,(m),β}\{D^{2m}W_{w,\ell}(m),\beta\}. Then

{D2mWw(m),kβ}={k1D2mWw(m),β}={D2mWw(mk),β}.\{D^{2m}W_{w}(m),k\cdot\beta\}=\{k^{-1}D^{2m}W_{w}(m),\beta\}=\{D^{2m}W_{w}(mk),\beta\}.

Thus if we can compute {D2mWw(m),β}\{D^{2m}W_{w}(m),\beta\} for β=(xy)m\beta=(x\wedge y)^{\otimes m} where Ex,yE_{x,y} is assumed singular and isotropic, then we can compute this quantity for general β\beta.

What we actually do is compute (x2m+2,{D2mWw(m),β})(x^{2m+2\ell},\{D^{2m}W_{w}(m),\beta\}) for β=(xy)m\beta=(x\wedge y)^{\otimes m} where Ex,yE_{x,y} is assumed singular and isotropic. To setup the computation, we fix a basis x1=x,x2=y,x_{1}=x,x_{2}=y,\ldots of J0J^{0}, fix a basis of WJ=WQV2sJ0W_{J}=W_{Q}\oplus V_{2}^{s}\otimes J^{0} that is the union of bases of W𝐐W_{\mathbf{Q}} and of the tensor product basis {xs,ys}{x1,x2,}\{x_{s},y_{s}\}\otimes\{x_{1},x_{2},\ldots\} of V2sJ0V_{2}^{s}\otimes J^{0}. Then we fix our basis of 𝔭V2WJ{\mathfrak{p}}\simeq V_{2}^{\ell}\otimes W_{J} to be the tensor product basis of {x,y}\{x,y\} with the above fixed basis of WJW_{J}.

Now, it is clear that (x2m+2,{D2mWw(m),β})(x^{2m+2\ell},\{D^{2m}W_{w}(m),\beta\}) only contains the terms in D2mD^{2m} where the XαiX_{\alpha_{i}} equal one of

X1=y(xsx),X2=y(xsy),X3=y(ysx),X4=y(ysy).X_{1}=y\otimes(x_{s}\otimes x),X_{2}=y\otimes(x_{s}\otimes y),X_{3}=y\otimes(y_{s}\otimes x),X_{4}=y\otimes(y_{s}\otimes y).

Note also that because Ex,yE_{x,y} is isotropic, the above Lie algebra elements all commute. We obtain that

(x2m+,{D2mWw(m),(xy)m})\displaystyle(x^{2m+\ell},\{D^{2m}W_{w}(m),(x\wedge y)^{\otimes m}\}) =2mj=0m(1)j(mj)(X1X4)mj(X2X3)jWw,(m)\displaystyle=2^{m}\sum_{j=0}^{m}{(-1)^{j}\binom{m}{j}(X_{1}X_{4})^{m-j}(X_{2}X_{3})^{j}W_{w,\ell}^{-\ell}(m)}
=2m(X1X4X2X3)mWw,(m).\displaystyle=2^{m}(X_{1}X_{4}-X_{2}X_{3})^{m}W_{w,\ell}^{-\ell}(m).

We now observe the following fact: if kF4I(𝐑)k\in F_{4}^{I}({\mathbf{R}}), then

(x2m+2,{D2mWw,(m),kβ})=(x2m+2,{k1D2mWw,(m),β})=(x2m+2,{D2mWw,(mk),β}).(x^{2m+2\ell},\{D^{2m}W_{w,\ell}(m),k\cdot\beta\})=(x^{2m+2\ell},\{k^{-1}D^{2m}\cdot W_{w,\ell}(m),\beta\})=(x^{2m+2\ell},\{D^{2m}W_{w,\ell}(mk),\beta\}).

Thus, if we can compute the right hand side, then we can compute the left hand side.

To compute the quantity (x2m+2,{D2mWw,(mk),β})(x^{2m+2\ell},\{D^{2m}W_{w,\ell}(mk),\beta\}), we will represent Ww,(g)W_{w,\ell}^{-\ell}(g) as an integral, and differentiate under the integral sign. This is inspired by the work of McGlade-Pollack [MP22]. More exactly, set

a,v(g)=(e,g1v)(pr(g1v),pr(g1v))+12.a_{\ell,v}(g)=\frac{(e_{\ell},g^{-1}v)^{\ell}}{(pr(g^{-1}v),pr(g^{-1}v))^{\ell+\frac{1}{2}}}.

Here pr:𝔤(J)𝐑𝔰𝔲2pr:{\mathfrak{g}}(J)\otimes{\mathbf{R}}\rightarrow{\mathfrak{su}}_{2} is the projection onto the Lie algebra of the long root SU2\operatorname{SU}_{2} and we write (X,Y)=B(X,Y)(X,Y)=B(X,Y) for short.

We prove the following theorem. To setup the theorem, recall from [Pol20a] that if zJz\in J with tr(z)=0\operatorname{tr}(z)=0, then

V(z)=(0,iz,z,0),V(z)=(0,iz,z,0).V(z)=(0,iz,-z,0),V^{*}(z)=(0,-iz,-z,0).

Moreover, let ν:MJGL1\nu:M_{J}\rightarrow\operatorname{GL}_{1} be the similitude character on the Levi of the Heisenberg parabolic of GJG_{J}. There is an identification MJHJM_{J}\simeq H_{J} (see [Pol20a, Lemma 4.3.1]), and ν\nu is the similitude character of HJH_{J} via this identification. I.e., for hHJh\in H_{J}, ν\nu satisfies hv,hv=ν(h)v,v\langle hv,hv^{\prime}\rangle=\nu(h)\langle v,v^{\prime}\rangle for all v,vWJv,v^{\prime}\in W_{J} and ,\langle\,,\,\rangle Freudenthal’s symplectic form on WJW_{J}.

Theorem 7.1.1.

Let the notation be as above. Let NwNN_{w}\subseteq N consist of the nn with w,n¯=0\langle w,\overline{n}\rangle=0. Suppose gMJ(𝐑)g\in M_{J}({\mathbf{R}}), the Levi of the Heisenberg parabolic of GJ(𝐑)G_{J}({\mathbf{R}}) and wWJw\in W_{J} is positive semidefinite. Set w1=g1ww_{1}=g^{-1}w. There is a nonzero constant B,mB^{\prime}_{\ell,m}, independent of w0w\geq 0 and independent of gg so that the integral

𝐑=Nw\Neiw,n¯(X1X4X2X3)ma,v(ng)𝑑n\int_{{\mathbf{R}}=N_{w}\backslash N}e^{-i\langle w,\overline{n}\rangle}(X_{1}X_{4}-X_{2}X_{3})^{m}a_{\ell,v}(ng)\,dn

is equal to

B,mν(g)mWw,+m(+m)(g)(V(x),w1V(y),w1V(y),w1V(x),w1)m.B^{\prime}_{\ell,m}\nu(g)^{m}W_{w,\ell+m}^{-(\ell+m)}(g)(\langle V(x),w_{1}\rangle\langle V^{*}(y),w_{1}\rangle-\langle V(y),w_{1}\rangle\langle V^{*}(x),w_{1}\rangle)^{m}.

Note that the m=0m=0 case of Theorem 7.1.1 represents the function Ww,(g)W_{w,\ell}^{-\ell}(g) as a integral, and the cases m>0m>0 of this theorem compute (by exchaning the order of integration and differentiation) the derivatives (X1X4X2X3)mWw,(g)(X_{1}X_{4}-X_{2}X_{3})^{m}W_{w,\ell}^{-\ell}(g). We justify this exchange of integration and differentiation in subsection 7.6.

Corollary 7.1.2.

Suppose gg is in the Levi of the Heisenberg parabolic of G2(𝐑)G_{2}({\mathbf{R}}), and βVmλ3\beta^{\prime}\in V_{m\lambda_{3}}. Then

(x2m+2,{D2mWw,(g),β})=Bm,′′Ww,+m(+m)(g)Pm(w),βI.(x^{2m+2\ell},\{D^{2m}W_{w,\ell}(g),\beta^{\prime}\})=B_{m,\ell}^{\prime\prime}W_{w,\ell+m}^{-(\ell+m)}(g)\langle P_{m}(w),\beta^{\prime}\rangle_{I}.

for a nonzero constant Bm,′′B_{m,\ell}^{\prime\prime}, independent of ww and gg.

Proof.

Suppose w=(a,b,c,d)w=(a,b,c,d). First consider the case β=β\beta^{\prime}=\beta. We must simplify the quantity

V(x),w1V(y),w1V(y),w1V(x),w1.\langle V(x),w_{1}\rangle\langle V^{*}(y),w_{1}\rangle-\langle V(y),w_{1}\rangle\langle V^{*}(x),w_{1}\rangle.

If w1=w1+(0,b1,c1,0)w_{1}=w_{1}^{\prime}+(0,b_{1},-c_{1},0) with tr(b1)=tr(c1)=0\operatorname{tr}(b_{1})=\operatorname{tr}(c_{1})=0 and w1W𝐐w_{1}^{\prime}\in W_{\mathbf{Q}}, then

V(x),w1V(y),w1V(y),w1V(x),w1\displaystyle\langle V(x),w_{1}\rangle\langle V^{*}(y),w_{1}\rangle-\langle V(y),w_{1}\rangle\langle V^{*}(x),w_{1}\rangle =(i(x,c1)(x,b1))(i(y,c1)(y,b1))\displaystyle=(i(x,c_{1})-(x,b_{1}))(-i(y,c_{1})-(y,b_{1}))
(i(x,c1)(x,b1))(i(y,c1)(y,b1))\displaystyle\,\,-(-i(x,c_{1})-(x,b_{1}))(i(y,c_{1})-(y,b_{1}))
=2i((x,b1)(y,c1)(x,c1)(y,b1))\displaystyle=2i((x,b_{1})(y,c_{1})-(x,c_{1})(y,b_{1}))
=2i(xy,b1c1).\displaystyle=2i(x\wedge y,b_{1}\wedge c_{1}).

Consequently, if gGL2sg\in\operatorname{GL}_{2}^{s}, the Heisenberg Levi on G2G_{2}, and w=w+(0,b,c,0)w=w^{\prime}+(0,b^{\prime},-c^{\prime},0), then b1c1=det(g)1bcb_{1}\wedge c_{1}=\det(g)^{-1}b^{\prime}\wedge c^{\prime}, so

V(x),w1V(y),w1V(y),w1V(x),w1=(2i)det(g)1(xy,bc).\langle V(x),w_{1}\rangle\langle V^{*}(y),w_{1}\rangle-\langle V(y),w_{1}\rangle\langle V^{*}(x),w_{1}\rangle=(2i)\det(g)^{-1}(x\wedge y,b^{\prime}\wedge c^{\prime}).

In general, if β=jαjkj(xy)m\beta^{\prime}=\sum_{j}{\alpha_{j}k_{j}(x\wedge y)^{\otimes m}}, with αj𝐂\alpha_{j}\in{\mathbf{C}} and kjF4I(𝐑)k_{j}\in F_{4}^{I}({\mathbf{R}}), one finds that

jαj(V(x),kj1g1wV(y),kj1g1wV(y),kj1g1wV(x),kj1g1w)m\sum_{j}\alpha_{j}(\langle V(x),k_{j}^{-1}g^{-1}w\rangle\langle V^{*}(y),k_{j}^{-1}g^{-1}w\rangle-\langle V(y),k_{j}^{-1}g^{-1}w\rangle\langle V^{*}(x),k_{j}^{-1}g^{-1}w\rangle)^{m}

is equal to

(2i)mdet(g)mjαj(kj(xy),bc)m\displaystyle(2i)^{m}\det(g)^{-m}\sum_{j}\alpha_{j}(k_{j}(x\wedge y),b^{\prime}\wedge c^{\prime})^{m} =(2i)mdet(g)mjαj(kj(xy)m,(bc)m)\displaystyle=(2i)^{m}\det(g)^{-m}\sum_{j}\alpha_{j}(k_{j}(x\wedge y)^{\otimes m},(b^{\prime}\wedge c^{\prime})^{\otimes m})
=(2i)mdet(g)m(β,(bc)m)\displaystyle=(2i)^{m}\det(g)^{-m}(\beta,(b^{\prime}\wedge c^{\prime})^{\otimes m})
=(2i)mdet(g)m(β,(bc)m)\displaystyle=(-2i)^{m}\det(g)^{-m}(\beta,(b\wedge c)^{\otimes m})

if w=(a,b,c,d)w=(a,b,c,d).

The det(g)m\det(g)^{-m} cancels the ν(g)m\nu(g)^{m} from Theorem 7.1.1, giving the corollary. ∎

Theorem 5.0.1 follows easily from Corollary 7.1.2:

Proof of Theorem 5.0.1.

Both sides of the desired equality transform on the left under N(𝐑)N({\mathbf{R}}) in the same way, and on the right under KG2K_{G_{2}} in the same way. Moreover, for =4\ell=4, they are both known to be quaternionic functions. Thus to prove their equality, it suffices to pair against x2x^{2\ell}, and evaluate on gg in the Levi of the Heisenberg parabolic. But this is precisely what is done in Corollary 7.1.2, so the theorem is proved. ∎

7.2. Some derivatives

We now focus on proving Theorem 7.1.1. We must consider some derivatives of the function

a,v(g)=(e,g1v)(pr(g1v),pr(g1v))+12.a_{\ell,v}(g)=\frac{(e_{\ell},g^{-1}v)^{\ell}}{(pr(g^{-1}v),pr(g^{-1}v))^{\ell+\frac{1}{2}}}.

Suppose X=xw𝔭X=x\otimes w\in{\mathfrak{p}}. Then Xe=0X\cdot e_{\ell}=0, and thus X{g(e,g1v)}=0X\{g\mapsto(e_{\ell},g^{-1}v)\}=0. Consequently, the function (e,g1v)(e_{\ell},g^{-1}v)^{\ell} can be considered constant for the purposes of differentiating with respect to xW𝔭x\otimes W\subseteq{\mathfrak{p}}. We therefore must just differentiate the function

b,v(g)=(pr(g1v),pr(g1v))(+12).b_{\ell,v}(g)=(pr(g^{-1}v),pr(g^{-1}v))^{-(\ell+\frac{1}{2})}.

We obtain

Xb,v(g)\displaystyle Xb_{\ell,v}(g) =((+12))(pr(g1v),pr(g1v))(+32)(2)(pr([X,g1v]),pr(g1v))\displaystyle=(-(\ell+\frac{1}{2}))\cdot(pr(g^{-1}v),pr(g^{-1}v))^{-(\ell+\frac{3}{2})}\cdot(2)\cdot(-pr([X,g^{-1}v]),pr(g^{-1}v))
=(2+1)(pr([X,g1v]),pr(g1v))(pr(g1v),pr(g1v))(+32).\displaystyle=(2\ell+1)(pr([X,g^{-1}v]),pr(g^{-1}v))(pr(g^{-1}v),pr(g^{-1}v))^{-(\ell+\frac{3}{2})}.

We write 𝒞v(X,)=(pr([X,g1v]),pr(g1v))\mathcal{C}_{v}(X,\cdot)=(pr([X,g^{-1}v]),pr(g^{-1}v)) and 𝒞v(X1,X2)=(pr([X1,g1v]),pr([X2,g1v]))\mathcal{C}_{v}(X_{1},X_{2})=(pr([X_{1},g^{-1}v]),pr([X_{2},g^{-1}v])) and 𝒞v=(pr(g1v),pr(g1v)).{\mathcal{C}_{v}}=(pr(g^{-1}v),pr(g^{-1}v)). Thus,

Xb,v(g)=(2+1)𝒞v(X,)𝒞v(+32).Xb_{\ell,v}(g)=(2\ell+1)\mathcal{C}_{v}(X,\cdot){\mathcal{C}_{v}}^{-(\ell+\frac{3}{2})}.

Now suppose X1=X=xw1X_{1}=X=x\otimes w_{1}, and X2=xw2X_{2}=x\otimes w_{2} is such that Span{w1,w2}\operatorname{Span}\{w_{1},w_{2}\} is isotropic and singular. Because it is isotropic, [X1,X2]=0[X_{1},X_{2}]=0. Because it is singular, SpanX1,X2\operatorname{Span}{X_{1},X_{2}} consists of rank one elements of 𝔤(J){\mathfrak{g}}(J). Recall that X𝔤(J)X\in{\mathfrak{g}}(J) is rank one means [X,[X,y]]+2B(X,y)X=0[X,[X,y]]+2B(X,y)X=0 for all y𝔤(J)y\in{\mathfrak{g}}(J), where B(,)B(\,,\,) is the bilinear form proportional to the Killing form defined in [Pol20a, section 4.2.2]. Thus, by symmetrizing and using that X1,X2X_{1},X_{2} commute, one arrives at [X1,[X2,y]]=B(X1,y)X2+B(X2,y)X1-[X_{1},[X_{2},y]]=B(X_{1},y)X_{2}+B(X_{2},y)X_{1}. Thus, pr([X1,[X2,y]])=0pr([X_{1},[X_{2},y]])=0. Using this, we differentiate b,v(g)b_{\ell,v}(g) again to obtain

X2X1b,v(g)=(2+1)𝒞v(X1,X2)𝒞v(+32)+(2+1)(2+3)𝒞v(X1,)𝒞v(X2,)𝒞v(+52).X_{2}X_{1}b_{\ell,v}(g)=-(2\ell+1)\mathcal{C}_{v}(X_{1},X_{2}){\mathcal{C}_{v}}^{-(\ell+\frac{3}{2})}+(2\ell+1)(2\ell+3)\mathcal{C}_{v}(X_{1},\cdot)\mathcal{C}_{v}(X_{2},\cdot){\mathcal{C}_{v}}^{-(\ell+\frac{5}{2})}.

For 0kn/20\leq k\leq\lfloor n/2\rfloor, define 𝒞v,n,k\mathcal{C}_{v,n,k} to be the symmetric sum of terms of the form

𝒞v(X1,X2)𝒞v(X2k1,X2k)𝒞v(X2k+1,)𝒞v(Xn,).\mathcal{C}_{v}(X_{1},X_{2})\cdots\mathcal{C}_{v}(X_{2k-1},X_{2k})\mathcal{C}_{v}(X_{2k+1},\cdot)\cdots\mathcal{C}_{v}(X_{n},\cdot).

Then we have the following proposition.

Proposition 7.2.1.

Suppose X1,,XnX_{1},\ldots,X_{n} are such that

  1. (1)

    Xj=xwjX_{j}=x\otimes w_{j} for all jj with

  2. (2)

    Span(w1,,wn)\operatorname{Span}(w_{1},\ldots,w_{n}) singular and isotropic.

Then

XnX1b,v(g)=k=0n/2(1)k2nk(+12)nk𝒞v,n,k𝒞v(+nk+12).X_{n}\cdots X_{1}b_{\ell,v}(g)=\sum_{k=0}^{\lfloor n/2\rfloor}{(-1)^{k}2^{n-k}(\ell+\frac{1}{2})_{n-k}\mathcal{C}_{v,n,k}{\mathcal{C}_{v}}^{-(\ell+n-k+\frac{1}{2})}}.
Proof.

We proceed by induction, noting that the proposition is true for n=1n=1 and n=2n=2 as checked above. Note that

Xn+1𝒞v(+nk+12)=2(+nk+12)𝒞v(Xn+1,)𝒞v(+n+1k+12).X_{n+1}{\mathcal{C}_{v}}^{-(\ell+n-k+\frac{1}{2})}=2(\ell+n-k+\frac{1}{2})\mathcal{C}_{v}(X_{n+1},\cdot){\mathcal{C}_{v}}^{-(\ell+n+1-k+\frac{1}{2})}.

Thus, making the induction assumption, Xn+1X1b,v(g)X_{n+1}\cdots X_{1}b_{\ell,v}(g) is equal to

k=0n/2(1)k2n+1k(+12)n+1k𝒞v,n,k𝒞v(Xn+1,)𝒞v(+n+1k+12)\sum_{k=0}^{\lfloor n/2\rfloor}{(-1)^{k}2^{n+1-k}(\ell+\frac{1}{2})_{n+1-k}\mathcal{C}_{v,n,k}\mathcal{C}_{v}(X_{n+1},\cdot){\mathcal{C}_{v}}^{-(\ell+n+1-k+\frac{1}{2})}}

plus

k=0n/2(1)k12nk(+12)nk(Xn+1𝒞v,n,k)𝒞v(+nk+12).\sum_{k=0}^{\lfloor n/2\rfloor}{(-1)^{k-1}2^{n-k}(\ell+\frac{1}{2})_{n-k}(-X_{n+1}\mathcal{C}_{v,n,k}){\mathcal{C}_{v}}^{-(\ell+n-k+\frac{1}{2})}}.

But now note that

𝒞v,n,j𝒞v(Xn+1,)+(Xn+1𝒞v,n,j1)=𝒞v,n+1,j.\mathcal{C}_{v,n,j}\mathcal{C}_{v}(X_{n+1},\cdot)+(-X_{n+1}\mathcal{C}_{v,n,j-1})=\mathcal{C}_{v,n+1,j}.

The proposition follows. ∎

The next step is to calculate the 𝒞v(Xj,)\mathcal{C}_{v}(X_{j},\cdot) and 𝒞v(Xi,Xj)\mathcal{C}_{v}(X_{i},X_{j}), then integrate.

7.3. Some Lie algebra calculations

We set v=ewv=e\otimes w and x=ngNJ(𝐑)MJ(𝐑)x=ng\in N_{J}({\mathbf{R}})M_{J}({\mathbf{R}}).

Then if n=exp(en¯)n=\exp(e\otimes\overline{n}), n1v=ew+w,n¯(0100)n^{-1}v=e\otimes w+\langle w,\overline{n}\rangle\left(\begin{smallmatrix}0&1\\ 0&0\end{smallmatrix}\right), so x1v=ew1+z1(0100)x^{-1}v=e\otimes w_{1}+z_{1}\left(\begin{smallmatrix}0&1\\ 0&0\end{smallmatrix}\right) where w1=g1ww_{1}=g^{-1}w and z1=ν(g)1w,n¯z_{1}=\nu(g)^{-1}\langle w,\overline{n}\rangle.

Now recall the long root 𝔰𝔩2{\mathfrak{sl}}_{2} triple of 𝔤(J){\mathfrak{g}}(J) given by

  • e=14(ie+f)r0(i)e_{\ell}=\frac{1}{4}(ie+f)\otimes r_{0}(i),

  • f=14(ief)r0(i)f_{\ell}=\frac{1}{4}(ie-f)\otimes r_{0}(-i) and

  • h=i2((0110)+nL(1)+nL(1))h_{\ell}=\frac{i}{2}(\left(\begin{smallmatrix}0&1\\ -1&0\end{smallmatrix}\right)+n_{L}(-1)+n_{L}^{\vee}(1)).

For X𝔤X\in{\mathfrak{g}} we have pr(X)=B(X,f)e+12B(X,h)h+B(X,e)fpr(X)=B(X,f_{\ell})e_{\ell}+\frac{1}{2}B(X,h_{\ell})h_{\ell}+B(X,e_{\ell})f_{\ell}. Note that the exceptional Cayley transform CC [Pol20a, section 5] takes e,h,fe_{\ell},h_{\ell},f_{\ell} to (0100),(11),(0010)\left(\begin{smallmatrix}0&1\\ 0&0\end{smallmatrix}\right),\left(\begin{smallmatrix}1&\\ &-1\end{smallmatrix}\right),\left(\begin{smallmatrix}0&0\\ 1&0\end{smallmatrix}\right). We obtain

pr(x1v)=14(αeiz1hαf)pr(x^{-1}v)=\frac{1}{4}(\alpha^{*}e_{\ell}-iz_{1}h_{\ell}-\alpha f_{\ell})

where444This is slightly different than how α\alpha is defined in [Pol19]. α=w1,r0(i)\alpha=\langle w_{1},r_{0}(i)\rangle.

We now compute 𝒞v(Xj,)=(pr([Xj,x1v]),pr(x1v))\mathcal{C}_{v}(X_{j},\cdot)=(pr([X_{j},x^{-1}v]),pr(x^{-1}v)). We have

  1. (1)

    h3=12(1ii1)h_{3}=\frac{1}{2}\left(\begin{smallmatrix}-1&i\\ i&1\end{smallmatrix}\right), which under CC goes to ie(1,0,0,0)-ie\otimes(1,0,0,0).

  2. (2)

    h1(X)=12(ie+f)V(X)h_{1}(X)=\frac{1}{2}(ie+f)\otimes V(X), which goes to ie(0,X,0,0)-ie\otimes(0,X,0,0)

  3. (3)

    h1(Z)=i2(nL(Z)+nL(Z))+12M(Φ1,Z)h_{-1}(Z)=\frac{i}{2}(n_{L}(Z)+n_{L}^{\vee}(Z))+\frac{1}{2}M(\Phi_{1,Z}), which goes to ie(0,0,Z,0)-ie\otimes(0,0,Z,0)

  4. (4)

    h3=14(ief)r0(i)h_{-3}=\frac{1}{4}(ie-f)\otimes r_{0}(i), which goes to ie(0,0,0,1)-ie\otimes(0,0,0,1).

Denote by pr𝔭pr_{{\mathfrak{p}}} the projection from 𝔤(J)𝐑𝔭{\mathfrak{g}}(J)\otimes{\mathbf{R}}\rightarrow{\mathfrak{p}}. Let’s suppose C1(pr𝔭(x1v))=ewe+fwfC^{-1}(pr_{{\mathfrak{p}}}(x^{-1}v))=e\otimes w_{e}+f\otimes w_{f}. Write also Xj=C(ewj)X_{j}=C(e\otimes w_{j}) with wj=(0,bj,cj,0)w_{j}=(0,b_{j},c_{j},0). Now

𝒞v(Xj,)\displaystyle\mathcal{C}_{v}(X_{j},\cdot) =([C1Xj,C1x1v],C1(pr(x1v)))\displaystyle=([C^{-1}X_{j},C^{-1}x^{-1}v],C^{-1}(pr(x^{-1}v)))
=14([ewj,ewe+fwf],αEiz1HαF)\displaystyle=\frac{1}{4}([e\otimes w_{j},e\otimes w_{e}+f\otimes w_{f}],\alpha^{*}E-iz_{1}H-\alpha F)
=14(wj,weE12wj,wfH,αEiz1HαF)\displaystyle=\frac{1}{4}(\langle w_{j},w_{e}\rangle E-\frac{1}{2}\langle w_{j},w_{f}\rangle H,\alpha^{*}E-iz_{1}H-\alpha F)
=α4wj,we+iz14wj,wf\displaystyle=-\frac{\alpha}{4}\langle w_{j},w_{e}\rangle+i\frac{z_{1}}{4}\langle w_{j},w_{f}\rangle

Here E=(0100),F=(0010)E=\left(\begin{smallmatrix}0&1\\ 0&0\end{smallmatrix}\right),F=\left(\begin{smallmatrix}0&0\\ 1&0\end{smallmatrix}\right) and H=(11)H=\left(\begin{smallmatrix}1&\\ &-1\end{smallmatrix}\right).

We now compute

wj,wf\displaystyle\langle w_{j},w_{f}\rangle =(ewj,fwf)\displaystyle=(-e\otimes w_{j},f\otimes w_{f})
=(ewj,C1(pr𝔭(x1v)))\displaystyle=(-e\otimes w_{j},C^{-1}(pr_{\mathfrak{p}}(x^{-1}v)))
=(C(ewj),x1v=ew1+z1E)\displaystyle=(-C(e\otimes w_{j}),x^{-1}v=e\otimes w_{1}+z_{1}E)
=(ih1(bj)ih1(cj),ew1+z1E)\displaystyle=(-ih_{1}(b_{j})-ih_{-1}(c_{j}),e\otimes w_{1}+z_{1}E)
=i(h1(bj),ew1)\displaystyle=-i(h_{1}(b_{j}),e\otimes w_{1})
=12((eif)V(bj),ew1)\displaystyle=\frac{1}{2}((e-if)\otimes V(b_{j}),e\otimes w_{1})
=i2V(bj),w1.\displaystyle=-\frac{i}{2}\langle V(b_{j}),w_{1}\rangle.

Similarly we compute

wj,we=(fwj,ewe)\displaystyle\langle w_{j},w_{e}\rangle=(f\otimes w_{j},e\otimes w_{e})
=(fwj,C1(pr𝔭(x1v)))\displaystyle=(f\otimes w_{j},C^{-1}(pr_{\mathfrak{p}}(x^{-1}v)))
=(C(fwj),x1v=ew1+z1E)\displaystyle=(C(f\otimes w_{j}),x^{-1}v=e\otimes w_{1}+z_{1}E)
=(ih1¯(bj)ih1¯(cj),ew1+z1E)\displaystyle=(i\overline{h_{-1}}(b_{j})-i\overline{h_{1}}(c_{j}),e\otimes w_{1}+z_{1}E)
=(ih1¯(cj),ew1)\displaystyle=(-i\overline{h_{1}}(c_{j}),e\otimes w_{1})
=i(12(ie+f)V(cj),ew1)\displaystyle=-i(\frac{1}{2}(-ie+f)\otimes V^{*}(c_{j}),e\otimes w_{1})
=i2V(cj),w1.\displaystyle=-\frac{i}{2}\langle V^{*}(c_{j}),w_{1}\rangle.

We therefore obtain

𝒞v(Xj,)=iα8V(cj),w1+z18V(bj),w1.\mathcal{C}_{v}(X_{j},\cdot)=\frac{i\alpha}{8}\langle V^{*}(c_{j}),w_{1}\rangle+\frac{z_{1}}{8}\langle V(b_{j}),w_{1}\rangle.

We now compute 𝒞v(Xi,Xj)\mathcal{C}_{v}(X_{i},X_{j}). We assume Xi=C(ewi=e(0,bi,ci,0))X_{i}=C(e\otimes w_{i}=e\otimes(0,b_{i},c_{i},0)), Xj=C(ewj=e(0,bj,cj,0))X_{j}=C(e\otimes w_{j}=e\otimes(0,b_{j},c_{j},0)). Then for k=i,jk=i,j, we have

pr([Xk,x1v])\displaystyle pr([X_{k},x^{-1}v]) =pr([C(ewk),pr𝔭(x1v)])\displaystyle=pr([C(e\otimes w_{k}),pr_{\mathfrak{p}}(x^{-1}v)])
=pr([C(ewk),C(ewe+fwf)])\displaystyle=pr([C(e\otimes w_{k}),C(e\otimes w_{e}+f\otimes w_{f})])
=(prC)([ewk,ewe+fwf])\displaystyle=(pr\circ C)([e\otimes w_{k},e\otimes w_{e}+f\otimes w_{f}])
=(prC)(wk,weEwk,wf2H)\displaystyle=(pr\circ C)(\langle w_{k},w_{e}\rangle E-\frac{\langle w_{k},w_{f}\rangle}{2}H)
=wk,weewk,wf2h.\displaystyle=\langle w_{k},w_{e}\rangle e_{\ell}-\frac{\langle w_{k},w_{f}\rangle}{2}h_{\ell}.

Thus

𝒞v(Xi,Xj)=12wi,wfwj,wf=18V(bi),w1V(bj),w1.\mathcal{C}_{v}(X_{i},X_{j})=\frac{1}{2}\langle w_{i},w_{f}\rangle\langle w_{j},w_{f}\rangle=-\frac{1}{8}\langle V(b_{i}),w_{1}\rangle\langle V(b_{j}),w_{1}\rangle.

We summarize what we’ve proved in a proposition.

Proposition 7.3.1.

Suppose Xi=C(ewi=e(0,bi,ci,0))X_{i}=C(e\otimes w_{i}=e\otimes(0,b_{i},c_{i},0)), Xj=C(ewj=e(0,bj,cj,0))X_{j}=C(e\otimes w_{j}=e\otimes(0,b_{j},c_{j},0)). Then one has

  1. (1)

    (e,x1v)=α4(e_{\ell},x^{-1}v)=-\frac{\alpha}{4};

  2. (2)

    𝒞v=18(|α|2+z12){\mathcal{C}_{v}}=-\frac{1}{8}(|\alpha|^{2}+z_{1}^{2});

  3. (3)

    𝒞v(Xj,)=iα8V(cj),w1+z18V(bj),w1\mathcal{C}_{v}(X_{j},\cdot)=\frac{i\alpha}{8}\langle V^{*}(c_{j}),w_{1}\rangle+\frac{z_{1}}{8}\langle V(b_{j}),w_{1}\rangle;

  4. (4)

    𝒞v(Xi,Xj)=18V(bi),w1V(bj),w1\mathcal{C}_{v}(X_{i},X_{j})=-\frac{1}{8}\langle V(b_{i}),w_{1}\rangle\langle V(b_{j}),w_{1}\rangle.

Here x=ngNJ(𝐑)MJ(𝐑)x=ng\in N_{J}({\mathbf{R}})M_{J}({\mathbf{R}}), α=w1,r0(i)\alpha=\langle w_{1},r_{0}(i)\rangle, w1=g1ww_{1}=g^{-1}w and z1=ν(g)1w,n¯z_{1}=\nu(g)^{-1}\langle w,\overline{n}\rangle.

Rewriting the above, we have

  1. (1)

    𝒞v(X1,)=z18V(x),w1-\mathcal{C}_{v}(X_{1},\cdot)=-\frac{z_{1}}{8}\langle V(x),w_{1}\rangle

  2. (2)

    𝒞v(X2,)=z18V(y),w1-\mathcal{C}_{v}(X_{2},\cdot)=-\frac{z_{1}}{8}\langle V(y),w_{1}\rangle

  3. (3)

    𝒞v(X3,)=iα8V(x),w1-\mathcal{C}_{v}(X_{3},\cdot)=i\frac{\alpha}{8}\langle V^{*}(x),w_{1}\rangle

  4. (4)

    𝒞v(X4,)=iα8V(y),w1-\mathcal{C}_{v}(X_{4},\cdot)=i\frac{\alpha}{8}\langle V^{*}(y),w_{1}\rangle

  5. (5)

    𝒞v(X1,X1)=18V(x),w1V(x),w1-\mathcal{C}_{v}(X_{1},X_{1})=\frac{1}{8}\langle V(x),w_{1}\rangle\langle V(x),w_{1}\rangle

  6. (6)

    𝒞v(X1,X2)=18V(x),w1V(y),w1-\mathcal{C}_{v}(X_{1},X_{2})=\frac{1}{8}\langle V(x),w_{1}\rangle\langle V(y),w_{1}\rangle

  7. (7)

    𝒞v(X2,X2)=18V(y),w1V(y),w1-\mathcal{C}_{v}(X_{2},X_{2})=\frac{1}{8}\langle V(y),w_{1}\rangle\langle V(y),w_{1}\rangle

7.4. More calculations

Putting together the work we have done above, we obtain the following lemma. Define 𝒞v=𝒞v{\mathcal{C}_{v}}^{*}=-{\mathcal{C}_{v}}, 𝒞v(Xj,)=𝒞v(Xj,)\mathcal{C}_{v}(X_{j},\cdot)^{*}=-\mathcal{C}_{v}(X_{j},\cdot), 𝒞v(Xi,Xj)=𝒞v(Xi,Xj)\mathcal{C}_{v}(X_{i},X_{j})^{*}=-\mathcal{C}_{v}(X_{i},X_{j}) and 𝒞v,n,k(X1Xn)\mathcal{C}_{v,n,k}^{*}(X_{1}\cdots X_{n}) to be the quantity one obtains by replacing every 𝒞v(Xj,)\mathcal{C}_{v}(X_{j},\cdot) with a 𝒞v(Xj,)\mathcal{C}_{v}(X_{j},\cdot)^{*} and every 𝒞v(Xi,Xj)\mathcal{C}_{v}(X_{i},X_{j}) with a 𝒞v(Xi,Xj)\mathcal{C}_{v}(X_{i},X_{j})^{*}.

Lemma 7.4.1.

Suppose n=2mn=2m. Consider

r=0m(1)r(mr)𝒞v,2m,k((X1X4)mr(X2X3)r).\sum_{r=0}^{m}(-1)^{r}\binom{m}{r}\mathcal{C}_{v,2m,k}^{*}((X_{1}X_{4})^{m-r}(X_{2}X_{3})^{r}).

This quantity is

(18)k(iα8)m(z18)m2k(m2k)(2k)!k!2k(V(x),w1V(y),w1V(y),w1V(x),w1)m.\left(\frac{1}{8}\right)^{k}\left(\frac{i\alpha}{8}\right)^{m}\left(\frac{-z_{1}}{8}\right)^{m-2k}\binom{m}{2k}\frac{(2k)!}{k!2^{k}}(\langle V(x),w_{1}\rangle\langle V^{*}(y),w_{1}\rangle-\langle V(y),w_{1}\rangle\langle V^{*}(x),w_{1}\rangle)^{m}.

(Interpret this formula to mean that it is 0 if 2k>m2k>m.)

Proof.

We evaluate 𝒞v,2m,k((X1X4)mr(X2X3)r)\mathcal{C}_{v,2m,k}^{*}((X_{1}X_{4})^{m-r}(X_{2}X_{3})^{r}). The point is that every nonzero term that goes into the definition of this sum is the same. One obtains that any such term is equal to

(18)k(iα8)m(z18)m2k(V(x),w1V(y),w1)mr(V(y),w1V(x),w1)r.\left(\frac{1}{8}\right)^{k}\left(\frac{i\alpha}{8}\right)^{m}\left(\frac{-z_{1}}{8}\right)^{m-2k}(\langle V(x),w_{1}\rangle\langle V^{*}(y),w_{1}\rangle)^{m-r}(\langle V(y),w_{1}\rangle\langle V^{*}(x),w_{1}\rangle)^{r}.

There are

1k!(m2)(m22)(m42)(m2k+22)=(m2k)(2k)!k!2k\frac{1}{k!}\binom{m}{2}\binom{m-2}{2}\binom{m-4}{2}\cdots\binom{m-2k+2}{2}=\binom{m}{2k}\frac{(2k)!}{k!2^{k}}

such nonzero terms. Thus

𝒞v,2m,k((X1X4)mr(X2X3)r)\displaystyle\mathcal{C}_{v,2m,k}^{*}((X_{1}X_{4})^{m-r}(X_{2}X_{3})^{r}) =(m2k)(2k)!k!2k(18)k(iα8)m(z18)m2k\displaystyle=\binom{m}{2k}\frac{(2k)!}{k!2^{k}}\left(\frac{1}{8}\right)^{k}\left(\frac{i\alpha}{8}\right)^{m}\left(\frac{-z_{1}}{8}\right)^{m-2k}
×(V(x),w1V(y),w1)mr(V(y),w1V(x),w1)r\displaystyle\,\times(\langle V(x),w_{1}\rangle\langle V^{*}(y),w_{1}\rangle)^{m-r}(\langle V(y),w_{1}\rangle\langle V^{*}(x),w_{1}\rangle)^{r}

and the lemma follows. ∎

Putting everything together, we obtain the following proposition. Set

b,v(g)=((pr(g1v),pr(g1v)))(+12).b_{\ell,v}^{*}(g)=(-(pr(g^{-1}v),pr(g^{-1}v)))^{-(\ell+\frac{1}{2})}.
Proposition 7.4.2.

One has

(X1X4X2X3)mb,v(g)\displaystyle(X_{1}X_{4}-X_{2}X_{3})^{m}b_{\ell,v}^{*}(g) =8+12(iα)m(V(x),w1V(y),w1V(y),w1V(x),w1)m\displaystyle=8^{\ell+\frac{1}{2}}(-i\alpha)^{m}(\langle V(x),w_{1}\rangle\langle V^{*}(y),w_{1}\rangle-\langle V(y),w_{1}\rangle\langle V^{*}(x),w_{1}\rangle)^{m}
×(k=0m/2(1)k22mkz1m2k(m2k)(2k)!k!2k(+12)2mk(|α|2+z12)(+2mk+12)).\displaystyle\,\times\left(\sum_{k=0}^{\lfloor m/2\rfloor}(-1)^{k}2^{2m-k}z_{1}^{m-2k}\binom{m}{2k}\frac{(2k)!}{k!2^{k}}(\ell+\frac{1}{2})_{2m-k}(|\alpha|^{2}+z_{1}^{2})^{-(\ell+2m-k+\frac{1}{2})}\right).

For a positive real number β\beta set

I,m(β)=𝐑eiz(k=0m/2Cm,kzm2k(+12)2mk(β2+z2)+2mk+12)𝑑zI_{\ell,m}(\beta)=\int_{{\mathbf{R}}}{e^{-iz}\left(\sum_{k=0}^{\lfloor m/2\rfloor}\frac{C_{m,k}z^{m-2k}(\ell+\frac{1}{2})_{2m-k}}{(\beta^{2}+z^{2})^{\ell+2m-k+\frac{1}{2}}}\right)\,dz}

where

Cm,k=(1)k22mk(m2k)(2k)!k!2k.C_{m,k}=(-1)^{k}2^{2m-k}\binom{m}{2k}\frac{(2k)!}{k!2^{k}}.

Set α,v(g)=(e,g1v)b,v(g)\alpha_{\ell,v}^{*}(g)=-(e_{\ell},g^{-1}v)b_{\ell,v}^{*}(g). We have proved:

Proposition 7.4.3.

Set z=ν(g)z1z=\nu(g)z_{1}. One has

𝐑=Nw\Neiw,n¯(X1X4X2X3)ma,v(ng)𝑑n\displaystyle\int_{{\mathbf{R}}=N_{w}\backslash N}e^{-i\langle w,\overline{n}\rangle}(X_{1}X_{4}-X_{2}X_{3})^{m}a_{\ell,v}^{*}(ng)\,dn =(i)m2+32(ν(g)α)+m|ν(g)|ν(g)+2mI,m(|ν(g)α|)\displaystyle=(-i)^{m}2^{\ell+\frac{3}{2}}(\nu(g)\alpha)^{\ell+m}|\nu(g)|\nu(g)^{\ell+2m}I_{\ell,m}(|\nu(g)\alpha|)
×(V(x),w1V(y),w1V(y),w1V(x),w1)m.\displaystyle\,\times(\langle V(x),w_{1}\rangle\langle V^{*}(y),w_{1}\rangle-\langle V(y),w_{1}\rangle\langle V^{*}(x),w_{1}\rangle)^{m}.

The following proposition now implies Theorem 7.1.1.

Proposition 7.4.4.

One has

I,m(β)=21im(12)β(+m)K+m(β).I_{\ell,m}(\beta)=\frac{2^{1-\ell}i^{m}}{(\frac{1}{2})_{\ell}}\beta^{-(\ell+m)}K_{\ell+m}(\beta).

7.5. Evaluation of an integral

In this subsection, we prove Proposition 7.4.4.

For m=0m=0, from equation (5) of [Wei], one obtains

I,0(β)=2(21)(23)(3)(1)βK(β).I_{\ell,0}(\beta)=\frac{2}{(2\ell-1)(2\ell-3)\cdots(3)(1)}\beta^{-\ell}K_{\ell}(\beta).

Set, for r>0r>0, v𝐙>0v\in{\mathbf{Z}}_{>0},

Iv,0(r,β)=𝐑eirz(β2+z2)(v+12)𝑑z.I_{v,0}(r,\beta)=\int_{{\mathbf{R}}}{e^{-irz}(\beta^{2}+z^{2})^{-(v+\frac{1}{2})}\,dz}.

Changing variables, one obtains

Iv,0(r,β)=Dv(r/β)vKv(rβ).I_{v,0}(r,\beta)=D_{v}(r/\beta)^{v}K_{v}(r\beta).

where Dv=2v+11(12)vD_{v}=2^{-v+1}\frac{1}{(\frac{1}{2})_{v}}. Differentiating under the integral sign, and using that

Cm,k(+12)2mk(i)m2kD+2mk=21(i)m(12),C_{m,k}(\ell+\frac{1}{2})_{2m-k}(-i)^{m-2k}D_{\ell+2m-k}=\frac{2^{1-\ell}(-i)^{m}}{(\frac{1}{2})_{\ell}},

we get

I,m(β)=21(i)m(12)(k=0m/2(m2k)(2k)!k!2krm2k((r/β)+2mkK+2mk(rβ))|r=1).I_{\ell,m}(\beta)=\frac{2^{1-\ell}(-i)^{m}}{(\frac{1}{2})_{\ell}}\left(\sum_{k=0}^{\lfloor m/2\rfloor}\binom{m}{2k}\frac{(2k)!}{k!2^{k}}\partial_{r}^{m-2k}((r/\beta)^{\ell+2m-k}K_{\ell+2m-k}(r\beta))|_{r=1}\right).

Finally, making the variable change u=rβu=r\beta, one gets

I,m(β)=21(i)m(12)(k=0m/2(m2k)(2k)!k!2kβ23mum2k(u+2mkK+2mk(u))|u=β).I_{\ell,m}(\beta)=\frac{2^{1-\ell}(-i)^{m}}{(\frac{1}{2})_{\ell}}\left(\sum_{k=0}^{\lfloor m/2\rfloor}\binom{m}{2k}\frac{(2k)!}{k!2^{k}}\beta^{-2\ell-3m}\partial_{u}^{m-2k}(u^{\ell+2m-k}K_{\ell+2m-k}(u))|_{u=\beta}\right).

Set

cnj=n!j!(n2j)!2j=(n2j)(2j)!j!2j.c_{n}^{j}=\frac{n!}{j!(n-2j)!2^{j}}=\binom{n}{2j}\frac{(2j)!}{j!2^{j}}.
Lemma 7.5.1.

Suppose bnb\geq n. One has

un(ubKb(u))=j=0n/2(1)njcnjubjKbn+j(u).\partial_{u}^{n}(u^{b}K_{b}(u))=\sum_{j=0}^{\lfloor n/2\rfloor}{(-1)^{n-j}c_{n}^{j}u^{b-j}K_{b-n+j}(u)}.
Proof.

The proof is by induction, using the recurrence cn+1j=cnj+(n2j+2)cnj1c_{n+1}^{j}=c_{n}^{j}+(n-2j+2)c_{n}^{j-1}. See also [oIS]. ∎

From the immediately verified identity

0j+k= fixedm/2(1)jcmkcm2kj=0\sum_{0\leq j+k=\text{ fixed}\leq\lfloor m/2\rfloor}(-1)^{j}c_{m}^{k}c_{m-2k}^{j}=0

if j+k>0j+k>0, we obtain Proposition 7.4.4.

7.6. Technical justification

We still must justify our differentiation under the integral. In other words, we must justify the identity

Nw\N𝐑Z1Zr(eiw,n¯(a,v(ng))dn=Z1Zr(Nw\N𝐑eiw,n¯a,v(ng)dn).\int_{N_{w}\backslash N\simeq{\mathbf{R}}}{Z_{1}\cdots Z_{r}\left(e^{-i\langle w,\overline{n}\rangle}(a_{\ell,v}(ng)\right)\,dn}=Z_{1}\cdots Z_{r}\left(\int_{N_{w}\backslash N\simeq{\mathbf{R}}}{e^{-i\langle w,\overline{n}\rangle}a_{\ell,v}(ng)\,dn}\right).

Here the ZiZ_{i} are in 𝔭{\mathfrak{p}} and differentiate with respect to the gg variable.

Write n=n(z)n=n(z) if nzn\mapsto z under the identification Nw\N𝐑N_{w}\backslash N\simeq{\mathbf{R}}. To do this justification, it suffices to show (for all non-negative integers rr) that there is a small neighborhood UU of gg, and a positive function Ar(z)A_{r}(z) on 𝐑{\mathbf{R}}, so that for all xUx\in U,

|Z1Zra,v(n(z)x)|Ar(z) and 𝐑Ar(z)𝑑z<.|Z_{1}\cdots Z_{r}a_{\ell,v}(n(z)x)|\leq A_{r}(z)\text{ and }\int_{{\mathbf{R}}}A_{r}(z)\,dz<\infty.

The function Ar(z)A_{r}(z) can depend upon gg and Z1,,ZrZ_{1},\ldots,Z_{r}, but we drop them from the notation.

It is easy to see (e.g., by induction) that the derivative Z1Zr(a,v(nx))Z_{1}\cdots Z_{r}(a_{\ell,v}(nx)) is of the form

(e,x1n1v)(pr(x1n1v),pr(x1n1v))+r+12PZ1,,Zr(x,n)\frac{(e_{\ell},x^{-1}n^{-1}v)^{\ell}}{(pr(x^{-1}n^{-1}v),pr(x^{-1}n^{-1}v))^{\ell+r+\frac{1}{2}}}P_{Z_{1},\ldots,Z_{r}}(x,n)

where PZ1,,Zr(x,n)P_{Z_{1},\ldots,Z_{r}}(x,n) consists of sums of products of terms of the form (pr([Zj,x1n1v],pr(x1n1v))(pr([Z_{j},x^{-1}n^{-1}v],pr(x^{-1}n^{-1}v)), (pr([Zj,x1n1v],pr([Zk,x1n1v]))(pr([Z_{j},x^{-1}n^{-1}v],pr([Z_{k},x^{-1}n^{-1}v])) etc. The key point is that, if n=n(z)n=n(z), then when written as a polynomial in zz, PZ1,,Zr(n,z)P_{Z_{1},\ldots,Z_{r}}(n,z) has degree at most 2r2r. The coefficients of this polynomial depend on xx and Z1,,ZrZ_{1},\ldots,Z_{r}, but are easily seen to be bounded for xx in a small compact set around gg.

To finish the proof, we now must bound

|(e,x1n1v)(pr(x1n1v),pr(x1n1v))+r+12|pr(x1n1v)(+2r+1).\left|\frac{(e_{\ell},x^{-1}n^{-1}v)^{\ell}}{(pr(x^{-1}n^{-1}v),pr(x^{-1}n^{-1}v))^{\ell+r+\frac{1}{2}}}\right|\leq||pr(x^{-1}n^{-1}v)||^{-(\ell+2r+1)}.

Here v=|B(v,θ(v))|1/2||v||=|B(v,\theta(v))|^{1/2} is the KK-equivariant norm on 𝔤(J)𝐑{\mathfrak{g}}(J)\otimes{\mathbf{R}}, where θ\theta is the Cartan involution. Write g=ngmgkgg=n_{g}m_{g}k_{g} in the Iwasawa decomposition. We take small open neighborhoods around ng,mgn_{g},m_{g} and kgk_{g}, and let UU be the product of these neighborhoods. Then, if n=n(z)n=n(z), pr(x1n1v)||pr(x^{-1}n^{-1}v)|| is bounded away from 0 for zz small, independent of xx, and is bounded below by (|z|2+|α0|2)1/2(|z|^{2}+|\alpha_{0}|^{2})^{1/2} for zz large, with α0\alpha_{0} independent of xUx\in U. The existence of Ar(z)A_{r}(z) with the desired properties follows.

8. Arithmeticity of modular forms on G2G_{2}

The purpose of this section is to prove Theorem 1.0.1, which we recall here. Suppose φ\varphi is a cuspidal quaternionic modular form on G2G_{2} of weight 1\ell\geq 1. Then

φZ(gfg)=χaχ(gf)Wχ(g)\varphi_{Z}(g_{f}g_{\infty})=\sum_{\chi}{a_{\chi}(g_{f})W_{\chi}(g_{\infty})}

is its Fourier expansion. The locally constant functions aχ:G2(𝐀f)𝐂a_{\chi}:G_{2}({\mathbf{A}}_{f})\rightarrow{\mathbf{C}} are its Fourier coefficients. We say that φ\varphi has Fourier coefficients in a ring RR if aχ(gf)Ra_{\chi}(g_{f})\in R for all characters χ\chi of N(𝐐)\N(𝐀)N({\mathbf{Q}})\backslash N({\mathbf{A}}) and all gfG2(𝐀f)g_{f}\in G_{2}({\mathbf{A}}_{f}). We write S(G2;R)S_{\ell}(G_{2};R) for the space cuspidal quaternionic modular forms on G2G_{2} of weight \ell with Fourier coefficients in RR.

Let 𝐐cyc=𝐐(μ){\mathbf{Q}}_{cyc}={\mathbf{Q}}(\mu_{\infty}) be the cyclotomic extension of 𝐐{\mathbf{Q}}.

Theorem 8.0.1.

Suppose 6\ell\geq 6 is even. Then there is a basis of the cuspidal quaternionic modular forms of weight \ell with all Fourier coefficients in 𝐐cyc{\mathbf{Q}}_{cyc}. In other words, S(G2,𝐂)=S(G2,𝐐cyc)𝐐cyc𝐂S_{\ell}(G_{2},{\mathbf{C}})=S_{\ell}(G_{2},{\mathbf{Q}}_{cyc})\otimes_{{\mathbf{Q}}_{cyc}}{\mathbf{C}}.

The proof of this theorem has the following steps:

  1. (1)

    Set S(G2)ΘS_{\ell}(G_{2})_{\Theta} the subspace of S(G2,𝐂)S_{\ell}(G_{2},{\mathbf{C}}) consisting of theta lifts from algebraic modular forms on F4IF_{4}^{I}. It is clear that it is a G2(𝐀f)G_{2}({\mathbf{A}}_{f}) submodule. Moreover, as an application of Theorem 5.0.1, it is easy to see that S(G2)ΘS_{\ell}(G_{2})_{\Theta} is defined over 𝐐cyc{\mathbf{Q}}_{cyc}, i.e., that it has a basis consisting of elements with Fourier coefficients in 𝐐cyc{\mathbf{Q}}_{cyc}.

  2. (2)

    To show that every element of S(G2;𝐂)S_{\ell}(G_{2};{\mathbf{C}}) is a theta lift, one uses the Siegel-Weil theorem of [Pol23] and the Rankin-Selberg integral of [GS15, Seg17].

  3. (3)

    For the previous step to go through, a certain archimedean Zeta integral must be shown to be non-vanishing. One shows the non-vanishing of this integral by a global method, using Corollary 1.2.3.

We will break the proof into various lemmas.

Lemma 8.0.2.

Suppose ϕVmin\phi\in V_{min} is ϕ=jμjgjϕ0\phi=\sum_{j}{\mu_{j}g_{j}\cdot\phi_{0}} with μj𝐐\mu_{j}\in{\mathbf{Q}} and ϕ0\phi_{0} the spherical vector. Let 𝒜(F4I,U;Vmλ3)\mathcal{A}(F_{4}^{I},U;V_{m\lambda_{3}}) be the algebraic modular forms on F4IF_{4}^{I} of level UF4I(𝐀f)U\subseteq F_{4}^{I}({\mathbf{A}}_{f}) and for the representation Vmλ3V_{m\lambda_{3}}. Then there is a lattice Λ𝒜(F4I,U;Vmλ3)\Lambda\subseteq\mathcal{A}(F_{4}^{I},U;V_{m\lambda_{3}}) so that if αΛ\alpha\in\Lambda, then Θϕ(α)S+m(G2;𝐂)\Theta_{\phi}(\alpha)\in S_{\ell+m}(G_{2};{\mathbf{C}}) has Fourier coefficients in 𝐐cyc{\mathbf{Q}}_{cyc}.

Proof.

Write

φ(g)=[F4I]{D𝔭2mΘϕ(g,h),α(h)}𝑑h.\varphi(g)=\int_{[F_{4}^{I}]}{\{D_{{\mathfrak{p}}}^{2m}\Theta_{\phi}(g,h),\alpha(h)\}\,dh}. (4)

Now

F4I(𝐐)\F4I(𝐑)F4I(𝐀f)/U=j=1NΓj\F4I(𝐑)γjF_{4}^{I}({\mathbf{Q}})\backslash F_{4}^{I}({\mathbf{R}})F_{4}^{I}({\mathbf{A}}_{f})/U=\cup_{j=1}^{N}\Gamma_{j}\backslash F_{4}^{I}({\mathbf{R}})\gamma_{j}

where γjF4I(𝐀f)\gamma_{j}\in F_{4}^{I}({\mathbf{A}}_{f}). In other words,

F4I(𝐐)\F4I(𝐀)=j=1NΓj\F4I(𝐑)γjU.F_{4}^{I}({\mathbf{Q}})\backslash F_{4}^{I}({\mathbf{A}})=\cup_{j=1}^{N}\Gamma_{j}\backslash F_{4}^{I}({\mathbf{R}})\gamma_{j}U.

Thus the integral over [F4I][F_{4}^{I}] is a finite sum of terms cj{D𝔭2mΘϕ(g,1),α(γj)}c_{j}\{D_{{\mathfrak{p}}}^{2m}\Theta_{\phi}(g,1),\alpha(\gamma_{j})\} where cj=meas(U)|Γj|c_{j}=\frac{meas(U)}{|\Gamma_{j}|} is rational. But now Theorem 5.0.1 implies that if the Fourier coefficients of Θϕ(g)\Theta_{\phi}(g) are in some ring RR, and α(γj)\alpha(\gamma_{j}) is in an appropriate lattice, then the Fourier coefficients of {D𝔭2mΘϕ(g,1),α(γj)}\{D_{{\mathfrak{p}}}^{2m}\Theta_{\phi}(g,1),\alpha(\gamma_{j})\} are in RR. Thus we obtain the fact that the Fourier coefficients of theta lifts can all be made in 𝐐cyc=𝐐(μ){\mathbf{Q}}_{cyc}={\mathbf{Q}}(\mu_{\infty}), as soon as we prove the same result for the Fourier coefficients of Θϕ(g)\Theta_{\phi}(g).

For the latter, simply observe the following identities: suppose gfGJ(𝐀f)g_{f}\in G_{J}({\mathbf{A}}_{f}). Then we can write gf=ufmfkfg_{f}=u_{f}m_{f}k_{f} with ufu_{f} in the NJN_{J}, kfGJ(𝐙^)=Kfk_{f}\in G_{J}(\widehat{{\mathbf{Z}}})=K_{f}, and mfHJ(𝐀f)m_{f}\in H_{J}({\mathbf{A}}_{f}). Then we can further write mf=(m𝐐m𝐑1)km_{f}=(m_{\mathbf{Q}}m_{\mathbf{R}}^{-1})k^{\prime} with m𝐐HJ(𝐐)m_{\mathbf{Q}}\in H_{J}({\mathbf{Q}}) and kKfk^{\prime}\in K_{f}. This follows from strong approximation on the simply connected group HJ1H_{J}^{1}. Thus, if a(ω)(gf)a(\omega)(g_{f}) denotes a Fourier coefficient of ΘGan(g)\Theta_{Gan}(g) (level one), then

a(ω)(gf)=ψ(ω,uf)a(ω)(m𝐐m𝐑1).a(\omega)(g_{f})=\psi(\langle\omega,u_{f}\rangle)a(\omega)(m_{\mathbf{Q}}m_{\mathbf{R}}^{-1}).

Additionally,

a(ω)(m𝐐m𝐑1)Wω(g)=Θω(m𝐐m𝐑1g)=Θωm𝐐(m𝐑1g)=aωm𝐐(1)Wωm𝐐(m𝐑1g).a(\omega)(m_{\mathbf{Q}}m_{\mathbf{R}}^{-1})W_{\omega}(g_{\infty})=\Theta_{\omega}(m_{\mathbf{Q}}m_{\mathbf{R}}^{-1}g_{\infty})=\Theta_{\omega\cdot m_{\mathbf{Q}}}(m_{\mathbf{R}}^{-1}g_{\infty})=a_{\omega\cdot m_{\mathbf{Q}}}(1)W_{\omega\cdot m_{\mathbf{Q}}}(m_{\mathbf{R}}^{-1}g_{\infty}).

This last term is det(m𝐑)|det(m𝐑)|1aωm𝐐(1)Wω(g).\det(m_{\mathbf{R}})^{-\ell}|\det(m_{\mathbf{R}})|^{-1}a_{\omega\cdot m_{\mathbf{Q}}}(1)W_{\omega}(g_{\infty}). Thus since all the aω(1)a_{\omega^{\prime}}(1) are integral, all a(ω)(gf)𝐐cyca(\omega)(g_{f})\in{\mathbf{Q}}_{cyc}. It thus follows that if ϕ=jμjgjϕ0\phi=\sum_{j}{\mu_{j}g_{j}\cdot\phi_{0}} with μj𝐐\mu_{j}\in{\mathbf{Q}}, then all the Fourier coefficients of Θϕ\Theta_{\phi} are in 𝐐cyc{\mathbf{Q}}_{cyc}. This completes the argument. ∎

Write S(G2)ΘS_{\ell}(G_{2})_{\Theta} for the space of all lifts φ\varphi as in equation (4).

Lemma 8.0.3.

The subspace S(G2)ΘS_{\ell}(G_{2})_{\Theta} is a G2(𝐀f)G_{2}({\mathbf{A}}_{f})-submodule.

Proof.

This is clear. ∎

Recall the projection 𝔭mS2m(V)(2J0)m.{\mathfrak{p}}^{\otimes m}\rightarrow S^{2m}(V_{\ell})\otimes(\wedge^{2}J^{0})^{\otimes m}. Let W(2J0)mW\subseteq(\wedge^{2}J^{0})^{\otimes m} be the set of all w(2J0)mw\in(\wedge^{2}J^{0})^{\otimes m} for which {w,v}=0\{w,v\}=0 for all vVmλ3v\in V_{m\lambda_{3}}. We set Vm=(2J0)m/WV_{m}^{*}=(\wedge^{2}J^{0})^{\otimes m}/W, and let P:𝔭mS2m(V)VmP:{\mathfrak{p}}^{\otimes m}\rightarrow S^{2m}(V_{\ell})\otimes V_{m}^{*} be the composite projection.

To prove that S(G2)Θ=S(G2;𝐂)S_{\ell}(G_{2})_{\Theta}=S_{\ell}(G_{2};{\mathbf{C}}) it suffices to show that if φS(G2;𝐂)\varphi\in S_{\ell}(G_{2};{\mathbf{C}}) generates an irreducible representation π\pi, then

[G2]{φ(g),P(D𝔭2mΘϕ(g,h))}𝑑g0\int_{[G_{2}]}{\{\varphi(g),P(D_{\mathfrak{p}}^{2m}\Theta_{\phi}(g,h))\}\,dg}\neq 0 (5)

for some φVmin\varphi\in V_{min}. Indeed, in this case, the submodule S(G2)ΘS_{\ell}(G_{2})_{\Theta} has orthocomplement equal to 0. Moreover, we can assume that φ\varphi is a pure tensor in π\pi.

Suppose EE is a totally real cubic étale extension of 𝐐{\mathbf{Q}}, and let SES_{E} be the group of type Spin8\operatorname{Spin}_{8} defined in terms of EE that has SE(𝐑)S_{E}({\mathbf{R}}) compact. See [Pol23] for a precise definition. To prove (5), it then further suffices to show that

[G2]×[SE]{φ(g),P(D𝔭2mΘϕ(g,h))}𝑑g0\int_{[G_{2}]\times[S_{E}]}{\{\varphi(g),P(D_{\mathfrak{p}}^{2m}\Theta_{\phi}(g,h))\}\,dg}\neq 0 (6)

for some such EE.

The integral in (6) can be evaluated using the main theorem of [Pol23] and the Rankin-Selberg integral studied in [GS15, Seg17] (see also [Pol19]). To set up the result, following [Pol23], write GEG_{E} for a certain simply connected group of absolute Dynkin type D4D_{4}, defined in terms of EE and split over 𝐑{\mathbf{R}}.

We have

D𝔭2mΘ(g)=D𝔭2m(vΘvj(g)[x4+j][y4j])=α,jΘuαvj(g)vjuαD_{{\mathfrak{p}}}^{2m}\otimes\Theta(g)=D_{{\mathfrak{p}}}^{2m}\left(\sum_{v}\Theta_{v_{j}}(g)\otimes[x^{4+j}][y^{4-j}]\right)=\sum_{\alpha,j}\Theta_{u_{\alpha}v_{j}}(g)\otimes v_{j}^{\vee}\otimes u_{\alpha}^{\vee}

for elements uαU(𝔤(J)𝐂)u_{\alpha}\in U({\mathfrak{g}}(J)\otimes{\mathbf{C}}). Thus by Corollary 9.4.8 of [Pol23],

[G2]×[SE]{φ(g),P(D𝔭2mΘϕ(g,h))}𝑑g𝑑h=[G2]{φ(g),α,jE1(ϕ,uαvj)(g,s=5)P(vjuα)}𝑑g.\int_{[G_{2}]\times[S_{E}]}{\{\varphi(g),P(D_{{\mathfrak{p}}}^{2m}\Theta_{\phi}(g,h))\}\,dg\,dh}=\int_{[G_{2}]}{\{\varphi(g),\sum_{\alpha,j}E_{1}(\phi,u_{\alpha}v_{j})(g,s=5)\otimes P(v_{j}^{\vee}\otimes u_{\alpha}^{\vee})\}\,dg}. (7)

Here E1(ϕ,uαvj,s=5)E_{1}(\phi,u_{\alpha}v_{j},s=5) is the Siegel-Weil Eisenstein series on the group GEG_{E}.

The integral of (7) can now be written as a partial LL-function times some local Zeta integrals at bad finite places (including the archimedean place). Specifically, we have the following proposition. Moreover, these local zeta integrals at the finite places can be trivialized with Siegel-Weil inducing data for the Eisenstein series E1E_{1} on GEG_{E}. Specifically, we have the following proposition.

Suppose χ:N(𝐐)\N(𝐀)𝐂×\chi:N({\mathbf{Q}})\backslash N({\mathbf{A}})\rightarrow{\mathbf{C}}^{\times} is a unitary character. To setup the proposition, we define

I,χ(s)=N0,χ\G2(𝐑){Wχ(g),α,jfuαvj(γ0,χg,s)P(vjuα)}𝑑g.I_{\infty,\chi}(s)=\int_{N_{0,\chi}\backslash G_{2}({\mathbf{R}})}\{W_{\chi}(g),\sum_{\alpha,j}f_{u_{\alpha}v_{j}}(\gamma_{0,\chi}g,s)\otimes P(v_{j}^{\vee}\otimes u_{\alpha}^{\vee})\}\,dg.

This is the local archimedean Zeta integral that comes from the Rankin-Selberg integral (7). The notation is from [Pol19, Theorem 5.2], which is a restatement of a Theorem of [GS15, Seg17]. Here also fuαvj(g,s)f_{u_{\alpha}v_{j}}(g,s) is the Siegel-Weil inducing section from [Pol23]. It follows from Proposition 8.0.8 below that I,χ(s)I_{\infty,\chi}(s) converges absolutely for Re(s)>1Re(s)>1.

Proposition 8.0.4.

Let χ\chi be a unitary character of N(𝐐)\N(𝐀)N({\mathbf{Q}})\backslash N({\mathbf{A}}) for which aχ(φ)(1)0a_{\chi}(\varphi)(1)\neq 0. Recall that from the theory of binary cubic forms one associates to φ\varphi is a rank three 𝐙{\mathbf{Z}}-module RχR_{\chi} in a cubic étale algebra EE over 𝐐{\mathbf{Q}}. Suppose SS is a set of finite places of 𝐐{\mathbf{Q}} that satisfies the following conditions:

  1. (1)

    If pSp\notin S, then πp\pi_{p} is unramified and φ\varphi is spherical at pp;

  2. (2)

    If pSp\notin S, then Rχ𝐙pR_{\chi}\otimes{\mathbf{Z}}_{p} is a ring, and in fact the maximal order of E𝐙pE\otimes{\mathbf{Z}}_{p};

  3. (3)

    S{2,3}S\supseteq\{2,3\}.

Then the finite vector ϕ\phi can be chosen so that ϕ\phi is spherical outside SS, and the integral (7) is equal to LS(π,Std,s=3)I,E(s=3)L^{S}(\pi,Std,s=3)I_{\infty,E}(s=3). Moreover, if φ\varphi is unramified at all finite primes, and Rχ=𝐙×𝐙×𝐙R_{\chi}={\mathbf{Z}}\times{\mathbf{Z}}\times{\mathbf{Z}} in 𝐐×𝐐×𝐐{\mathbf{Q}}\times{\mathbf{Q}}\times{\mathbf{Q}}, then SS may be chosen to be empty.

Observe that if ϕ\phi is level one and has 𝐙×𝐙×𝐙{\mathbf{Z}}\times{\mathbf{Z}}\times{\mathbf{Z}} Fourier coefficient nonzero, then we may take SS to be empty.

Proof.

The fact that the global integral represents the partial LL-function is from [GS15, Seg17], for a slightly larger set SS. In [Pol19] the set SS is shrunk to that in the statement of the proposition, except that [Pol19] includes 2,3S2,3\in S in all cases. Then, in [cDD+22], the case where φ\varphi is level one and RχR_{\chi} is 𝐙×𝐙×𝐙{\mathbf{Z}}\times{\mathbf{Z}}\times{\mathbf{Z}} is handled.

That the bad local integrals may be trivialized with some data is in [Seg17, Section 7]. What we state and use is slightly stronger. Specifically, we must verify that the bad local integrals can be trivialized for Siegel-Weil inducing sections. This follows simply because the Siegel-Weil inducing sections make up all of the induced representation IndPE(𝐐p)GE(𝐐p)(δPE)Ind_{P_{E}({\mathbf{Q}}_{p})}^{G_{E}({\mathbf{Q}}_{p})}(\delta_{P_{E}}), where PEP_{E} is the Heisenberg parabolic of GEG_{E}. To see this, recall that IndPE(𝐐p)GE(𝐐p)(δPE)Ind_{P_{E}({\mathbf{Q}}_{p})}^{G_{E}({\mathbf{Q}}_{p})}(\delta_{P_{E}}) is generated by any vector which is not annihilated by the long intertwining operator, which turns out to be given by an absolutely convergent integral. The restriction of the spherical vector from GJ(𝐐p)G_{J}({\mathbf{Q}}_{p}) is positively valued on GE(𝐐p)G_{E}({\mathbf{Q}}_{p}), so it cannot be annihilated by the long intertwining operator.

Finally, what we have stated is that ϕ\phi may be chosen to trivialize the integral, without changing the cusp form φ\varphi. This is slightly stronger from what is stated in [Seg17]. This claim follows from Lemma 8.0.7 below. ∎

It turns out the LL-value LS(π,Std,s=3)L^{S}(\pi,Std,s=3) is always nonzero. This is a direct consequence of the main theorem of [Mui97].

Theorem 8.0.5.

Let π\pi be a cuspidal automorphic representation of G2G_{2} over 𝐐{\mathbf{Q}}. Then the Euler product defining the partial standard LL-function LS(π,Std,s)L^{S}(\pi,Std,s) converges absolutely for Re(s)>2Re(s)>2.

Proof.

The local factors πp\pi_{p} are unitarizable. In [Mui97], the unitary dual of pp-adic G2G_{2} is completely and explicitly determined. In particular, when πp\pi_{p} is spherical, one has tight bounds on the Satake parameters of πp\pi_{p}. These bounds imply the absolute convergence statement of the theorem. ∎

Finally, it turns out that if 6\ell\geq 6 is even, the archimedean Zeta integral I,χ(s=3)I_{\infty,\chi}(s=3) is nonzero for all non-degenerate χ\chi.

Proposition 8.0.6.

Suppose 6\ell\geq 6 is even. Then I,χ(s=3)I_{\infty,\chi}(s=3) is nonzero for all non-degenerate χ\chi.

Proof.

A change of variables in the integral I,χ(s)I_{\infty,\chi}(s) shows that the non-vanishing of I,χ(s=3)I_{\infty,\chi}(s=3) is equivalent for all χ\chi. So, we take χ\chi with Rχ=𝐙×𝐙×𝐙R_{\chi}={\mathbf{Z}}\times{\mathbf{Z}}\times{\mathbf{Z}}. We will prove that this integral is nonvanishing by a global argument, using Proposition 8.0.4, Corollary 1.2.3, and Theorem 8.0.5. Fix now E=𝐐×𝐐×𝐐E={\mathbf{Q}}\times{\mathbf{Q}}\times{\mathbf{Q}}.

Let βK,m\beta_{K,m} be as in the proof of Corollary 1.2.3. Set β0=SE(𝐑)kβK,m𝑑k\beta_{0}=\int_{S_{E}({\mathbf{R}})}{k\cdot\beta_{K,m}\,dk}. Then one sees that Θ(kβK,m)\Theta(k\cdot\beta_{K,m}) has 𝐙×𝐙×𝐙{\mathbf{Z}}\times{\mathbf{Z}}\times{\mathbf{Z}} Fourier coefficient equal to 66, so Θ(β0)\Theta(\beta_{0}) does as well.

We can write Θ(β0)\Theta(\beta_{0}) as a finite sum of level one cuspidal eigenforms forms φj\varphi_{j}, with φj,Θ(β0)0\langle\varphi_{j},\Theta(\beta_{0})\rangle\neq 0. Thus there is some such φ\varphi with 𝐙×𝐙×𝐙{\mathbf{Z}}\times{\mathbf{Z}}\times{\mathbf{Z}} Fourier coefficient nonzero; fix this φ\varphi.

Now

[G2]×[SE]{φ(g),P(D𝔭2mΘ(g,h))}𝑑g𝑑h=L(πφ,Std,s=3)I,χ(s=3)\int_{[G_{2}]\times[S_{E}]}{\{\varphi(g),P(D_{{\mathfrak{p}}}^{2m}\Theta(g,h))\}\,dg\,dh}=L(\pi_{\varphi},Std,s=3)I_{\infty,\chi}(s=3)

from Proposition 8.0.4.

We have {φ,Θ(β0)}0\{\varphi,\Theta(\beta_{0})\}\neq 0. So

[G2]{φ(g)β0,P(D𝔭2mΘ(g,1))}𝑑g0.\int_{[G_{2}]}\{\varphi(g)\otimes\beta_{0},P(D_{{\mathfrak{p}}}^{2m}\Theta(g,1))\}\,dg\neq 0.

Let w(h)=[G2]{φ(g),P(D𝔭2mΘ(g,h))}𝑑gw(h)=\int_{[G_{2}]}{\{\varphi(g),P(D_{{\mathfrak{p}}}^{2m}\Theta(g,h))\}\,dg}. Then

[SE]w(h)𝑑h=|ΓSE|1SE(𝐑)w(k)𝑑k\int_{[S_{E}]}{w(h)\,dh}=|\Gamma_{S_{E}}|^{-1}\int_{S_{E}({\mathbf{R}})}{w(k)\,dk}

where ΓSE\Gamma_{S_{E}} is some finite group. Here we are using that SE(𝐀)=SE(𝐐)SE(𝐑)SE(𝐙^)S_{E}({\mathbf{A}})=S_{E}({\mathbf{Q}})S_{E}({\mathbf{R}})S_{E}(\widehat{{\mathbf{Z}}}).

But this latter integral is nonzero, because it is nonzero after pairing with β0\beta_{0}. Consequently, we have deduced that I,χ(s=3)I_{\infty,\chi}(s=3) is nonzero. ∎

Proof of Theorem 8.0.1.

We have proved that the space of lifts S(G2)ΘS_{\ell}(G_{2})_{\Theta} has a 𝐐cyc{\mathbf{Q}}_{cyc} structure, from the Fourier coefficients. We have also prove that if 6\ell\geq 6 is even, then S(G2)Θ=S(G2;𝐂)S_{\ell}(G_{2})_{\Theta}=S_{\ell}(G_{2};{\mathbf{C}}). This proves the theorem.∎

We end with some of the technical details that were used in the proofs above.

Lemma 8.0.7.

Let VpV_{p} denote the space of the representation πp\pi_{p}, and suppose L:Vp𝐂L:V_{p}\rightarrow{\mathbf{C}} is an (N,χ)(N,\chi) functional. Given vVpv\in V_{p}, there is a Schwartz-Bruhat function Φ\Phi on 𝔤E𝐐p{\mathfrak{g}}_{E}\otimes{\mathbf{Q}}_{p} so that

Ip(Φ,v,s)=N0,E\G2(𝐐p)L(gv)f(γ0,Eg,Φ,s)𝑑gI_{p}(\Phi,v,s)=\int_{N_{0,E}\backslash G_{2}({\mathbf{Q}}_{p})}{L(gv)f(\gamma_{0,E}g,\Phi,s)\,dg}

is equal to L(v)L(v), independent of ss.

Proof.

Write γ0,E1E13=eω\gamma_{0,E}^{-1}E_{13}=e\otimes\omega in the notation of [Pol19]. We have

Ip(Φ,v,s)=GL1×N0,E\G2|t|sΦ(tg1eω)L(gv)𝑑g𝑑t.I_{p}(\Phi,v,s)=\int_{\operatorname{GL}_{1}\times N_{0,E}\backslash G_{2}}{|t|^{s}\Phi(tg^{-1}e\otimes\omega)L(gv)\,dg\,dt}.

The function Φ\Phi is on 𝔤E\mathfrak{g}_{E}, and we have 𝔤E=𝔤2E0V7\mathfrak{g}_{E}={\mathfrak{g}}_{2}\oplus E^{0}\otimes V_{7}. In this decomposition, we can write eω=(eω)+(α0e1+α1e3)e\otimes\omega=(e\otimes\omega^{\prime})+(\alpha_{0}e_{1}+\alpha_{1}e_{3}^{*}), where α0,α1\alpha_{0},\alpha_{1} are a basis of E0E^{0}, the trace 0 elements of EE.

We take Φ\Phi a pure tensor, Φ=Φ𝔤2ΦEV7\Phi=\Phi_{{\mathfrak{g}}_{2}}\otimes\Phi_{E^{\otimes}V_{7}}. We make ΦE0V7\Phi_{E^{0}\otimes V_{7}} be the characteristic function of a set very close to α0e1+α1e3\alpha_{0}e_{1}+\alpha_{1}e_{3}^{*}. Let ZGL2Z_{\operatorname{GL}_{2}} be the center of the GL2\operatorname{GL}_{2} Levi of the Heisenberg parabolic on G2G_{2}. Then ΦE0V7(tg1)0\Phi_{E^{0}\otimes V_{7}}(tg^{-1})\neq 0 implies gZGL2(t)NHeisKG2(pM)g\in Z_{\operatorname{GL}_{2}}(t)N_{Heis}K_{G_{2}}(p^{M}) for some M>>0M>>0. Here KG2(pM)K_{G_{2}}(p^{M}) is the elements of G2(𝐐p)G_{2}({\mathbf{Q}}_{p}) that are 11 modulo pMp^{M} in the 7×77\times 7 matrix representation of G2G_{2}. Here we are using that if hG2h\in G_{2}, with h1e1=e1+δ1h^{-1}e_{1}=e_{1}+\delta_{1} and h1e3=e3+δ2h^{-1}e_{3}^{*}=e_{3}^{*}+\delta_{2}, with δjpMV7(𝐙p)\delta_{j}\in p^{M}V_{7}({\mathbf{Z}}_{p}), then there is kKG2(pM)k\in K_{G_{2}}(p^{M}) so that (hk)1e1=e1(hk)^{-1}e_{1}=e_{1} and (hk)1e3=e3(hk)^{-1}e_{3}^{*}=e_{3}^{*}. (Indeed, this latter fact can be proved by using KGL2(pM)K_{\operatorname{GL}_{2}}(p^{M}) and also unipotent elements in N(pM𝐙p)N(p^{M}{\mathbf{Z}}_{p}).)

Thus we must evaluate

GL1×GaL(z(t)v)Φ𝔤2(t(eω+zE13))|t|sψ(z)𝑑t𝑑z.\int_{\operatorname{GL}_{1}\times G_{a}}{L(z(t)v)\Phi_{{\mathfrak{g}}_{2}}(t(e\otimes\omega^{\prime}+zE_{13}))|t|^{s}\psi(z)\,dt\,dz}.

We choose Φ𝔤2\Phi_{{\mathfrak{g}}_{2}} to be a pure tensor in our root basis of 𝔤2{\mathfrak{g}}_{2}, so that Φ𝔤2(t(eω+zE13))=Φ𝔤2(t(eω))ΦE13(tz)\Phi_{{\mathfrak{g}}_{2}}(t(e\otimes\omega^{\prime}+zE_{13}))=\Phi_{{\mathfrak{g}}_{2}}^{\prime}(t(e\otimes\omega^{\prime}))\Phi_{E_{13}}(tz). But GaΦE13(tz)ψ(z)𝑑z=ΦE13^(t1)|t|1\int_{G_{a}}{\Phi_{E_{13}}(tz)\psi(z)\,dz}=\widehat{\Phi_{E_{13}}}(t^{-1})|t|^{-1}. By choosing ΦE13\Phi_{E_{13}} so that its Fourier transform is supported near t=1t=1, we see that we can trivialize the integral to a constant multiple of L(v)L(v). This proves the lemma. ∎

We now prove the absolute convergence of the archimedean Zeta integral. The integral in question is

N0,E\G2(𝐑){Wχ(g),f(γ0g,s)}𝑑g.\int_{N_{0,E}\backslash G_{2}({\mathbf{R}})}{\{W_{\chi}(g),f(\gamma_{0}g,s)\}\,dg}. (8)
Proposition 8.0.8.

The integral (8) converges absolutely for Re(s)>1Re(s)>1.

Proof.

Let Φ\Phi be a Sschwartz function on 𝔤𝐑{\mathfrak{g}}_{\mathbf{R}}. We obtain an inducing section from Φ\Phi as

f(g,Φ,s)=GL1(𝐑)|t|sΦ(tg1E13)𝑑t.f(g,\Phi,s)=\int_{\operatorname{GL}_{1}({\mathbf{R}})}{|t|^{s}\Phi(tg^{-1}E_{13})\,dt}.

We will check that every inducing section in I(s)I(s) is of this form, and we will prove the proposition for these inducing sections.

For the first part, observe that if fI(s)f\in I(s), then restricting to KK_{\infty} we obtain f(k1k,s)=f(k,s)f(k_{1}k,s)=f(k,s) for all k1KM(𝐑)k_{1}\in K_{\infty}\cap M({\mathbf{R}}). These are the k1Kk_{1}\in K_{\infty} for which k1E13=±E13k_{1}E_{13}=\pm E_{13}, and note that the negative sign does indeed occur. Now let β\beta be an arbitrary even smooth function on 𝔤E𝐑{\mathfrak{g}}_{E}\otimes{\mathbf{R}} and α\alpha a smooth compactly supported function on 𝐑>0{\mathbf{R}}_{>0}. We set Φ(v)=α(v)β(vv)\Phi(v)=\alpha(||v||)\beta\left(\frac{v}{||v||}\right); this is a Scwhartz function.

We have now

f(k,Φ,s)=𝐑×|t|sα(|t|)β(k1E13)𝑑t=2β(k1E13)𝐑>0tsα(t)𝑑t.f(k,\Phi,s)=\int_{{\mathbf{R}}^{\times}}{|t|^{s}\alpha(|t|)\beta(k^{-1}E_{13})\,dt}=2\beta(k^{-1}E_{13})\int_{{\mathbf{R}}_{>0}}{t^{s}\alpha(t)\,dt}.

We have used the evenness of β\beta. But because β(k1E13)\beta(k^{-1}E_{13}) gives an arbitrary function on (KM(𝐑))\K(K_{\infty}\cap M({\mathbf{R}}))\backslash K_{\infty}, we see that every inducing section in I(s)I(s) is an f(g,Φ,s)f(g,\Phi,s).

We thus now proceed to prove the absolute convergence of the double integral

GL1(𝐑)N0,E\G2(𝐑)Wχ(g)|t|sΦ(tg1eω)𝑑g𝑑t\int_{\operatorname{GL}_{1}({\mathbf{R}})}\int_{N_{0,E}\backslash G_{2}({\mathbf{R}})}W_{\chi}(g)|t|^{s}\Phi(tg^{-1}e\otimes\omega)\,dg\,dt

when Re(s)>1Re(s)>1.

The bound we use below on Wχ(mk)W_{\chi}(mk) is independent of kk, so it suffices to integrate over GL1×(Ga×GL2)\operatorname{GL}_{1}\times(\mathrm{G}_{a}\times\operatorname{GL}_{2}). Without loss of generality, we can assume that Φ\Phi is a pure tensor of the appropriate sort. We thus must bound the integral

GL1×Ga×GL2|t|sΦ13(tdet(m)1z)Φ𝔤2(tm1ω)|det(m)|3ΦM2(tm1)Wχ(m)𝑑t𝑑z𝑑m.\int_{\operatorname{GL}_{1}\times\mathrm{G}_{a}\times\operatorname{GL}_{2}}|t|^{s}\Phi_{13}(t\det(m)^{-1}z)\Phi_{{\mathfrak{g}}_{2}}(tm^{-1}\omega^{\prime})|\det(m)|^{-3}\Phi_{M_{2}}(tm^{-1})W_{\chi}(m)\,dt\,dz\,dm.

Here ΦM2\Phi_{M_{2}} is a Schwartz function on the 2×22\times 2 matrices M2(𝐑)M_{2}({\mathbf{R}}), and the rest of the notation is as in Lemma 8.0.7.

One has 𝐑|Φ13(tdet(m)1z)|𝑑z<C|det(m)||t|1\int_{{\mathbf{R}}}{|\Phi_{13}(t\det(m)^{-1}z)|\,dz}<C|det(m)||t|^{-1} and Φ𝔤2\Phi_{{\mathfrak{g}}_{2}} is bounded above, so we must bound

GL1×GL2|t|s1|det(m)|2ΦM2(tm1)Wχ(m)𝑑t𝑑m.\int_{\operatorname{GL}_{1}\times\operatorname{GL}_{2}}|t|^{s-1}|\det(m)|^{-2}\Phi_{M_{2}}(tm^{-1})W_{\chi}(m)\,dt\,dm.

Make the variable change tdet(m)tt\mapsto\det(m)t, and set m=det(m)m1m^{\prime}=\det(m)m^{-1}. Then we must bound

GL1×GL2|t|s1|det(m)|s3ΦM2(tm)Wχ(m)𝑑t𝑑m.\int_{\operatorname{GL}_{1}\times\operatorname{GL}_{2}}|t|^{s-1}|\det(m)|^{s-3}\Phi_{M_{2}}(tm^{\prime})W_{\chi}(m)\,dt\,dm.

Now we claim GL1|t|s1Φ(tm)𝑑t\int_{\operatorname{GL}_{1}}{|t|^{s-1}\Phi(tm^{\prime})\,dt} is, for s>1s>1 (assumed real now) bounded by Cms1s1C\frac{||m||^{s-1}}{s-1}. Indeed, Φ(tm)\Phi(tm^{\prime}) is rapidly decreasing, so |Φ(tm)|<Cmax{1,|t|NmN}|\Phi(tm^{\prime})|<C\max\{1,|t|^{-N}||m||^{-N}\} for an NN sufficiently large of our choosing. Thus

GL1|t|s1|Φ(tm)|𝑑t<C(0m|t|s1dtt+m|t|s1NmNdtt).\int_{\operatorname{GL}_{1}}{|t|^{s-1}|\Phi(tm^{\prime})|\,dt}<C\left(\int_{0}^{||m||}{|t|^{s-1}\,\frac{dt}{t}}+\int_{||m||}^{\infty}{|t|^{s-1-N}||m||^{-N}\,\frac{dt}{t}}\right).

Dropping terms of the form 1s1\frac{1}{s-1} (since we are fixing ss), both integrals above are bounded by Cms1C||m||^{s-1}. Thus we must bound

GL2ms1|det(m)|s3Wχ(m)𝑑m.\int_{\operatorname{GL}_{2}}{||m||^{s-1}|\det(m)|^{s-3}W_{\chi}(m)\,dm}.

We break this integral into two pieces, one where m1||m||\leq 1 and the other with m>1||m||>1. The first integral has a compact domain, so can be ignored. To show the convergence of the second integral, it suffices to show that |Wχ(m)|<ϕ(m)|W_{\chi}(m)|<\phi(||m||), where ϕ\phi is a rapidly decreasing function. And since the KK-Bessel function is rapidly decreasing, it suffices to show that |ω,mr0(i)|Cm|\langle\omega^{\prime},mr_{0}(i)\rangle|\geq C||m|| for all mGL2(𝐑)m\in\operatorname{GL}_{2}({\mathbf{R}}).

Both sides of the desired inequality scale linearly with the center of GL2(𝐑)\operatorname{GL}_{2}({\mathbf{R}}), so it suffices to check that

|ω,mr0(i)|2m2\frac{|\langle\omega^{\prime},mr_{0}(i)\rangle|^{2}}{||m||^{2}}

is bounded away from 0 for mSL2(𝐑)m\in\operatorname{SL}_{2}({\mathbf{R}}). Furthermore, by a change of variables again (now or in the initial integral), we can assume ω=(0,1,1,0)\omega^{\prime}=(0,1,-1,0). Then if

m=(1x01)(y1/2y1/2)k,m=\left(\begin{array}[]{cc}1&x\\ 0&1\end{array}\right)\left(\begin{array}[]{cc}y^{1/2}&\\ &y^{-1/2}\end{array}\right)k,

we wish to bound below the quantity

y3|h(z)|((y2+x2+1)/y)1=((x1)2+y2)(x2+y2)(x2+y2+1)y2y^{-3}|h(z)|((y^{2}+x^{2}+1)/y)^{-1}=\frac{((x-1)^{2}+y^{2})(x^{2}+y^{2})}{(x^{2}+y^{2}+1)y^{2}}

where h(z)=z2z=z(z1)h(z)=z^{2}-z=z(z-1).

Finally, to see that this rational function in x,yx,y is bounded below for y>0y>0, we work in polar coordinates. We have

((x1)2+y2)(x2+y2)(x2+y2+1)y2=(12xr2+1)(1+x2y2)=(1+(x1)2y2)(11r2+1).\frac{((x-1)^{2}+y^{2})(x^{2}+y^{2})}{(x^{2}+y^{2}+1)y^{2}}=\left(1-\frac{2x}{r^{2}+1}\right)\left(1+\frac{x^{2}}{y^{2}}\right)=\left(1+\frac{(x-1)^{2}}{y^{2}}\right)\left(1-\frac{1}{r^{2}+1}\right).

If |r|1/2|r|\geq 1/2 then r21/4r^{2}\geq 1/4 so 1+r25/41+r^{2}\geq 5/4 so 1/(1+r2)4/51/(1+r^{2})\leq 4/5 and then 1(1/(r2+1))1/51-(1/(r^{2}+1))\geq 1/5, so the quantity is at least 1/51/5. If |r|1/2|r|\leq 1/2, then

2x/(r2+1)=2rcos(θ)/(r2+1)2r/(r2+1)2(1/2)/((1/2)2+1)=4/52x/(r^{2}+1)=2r\cos(\theta)/(r^{2}+1)\leq 2r/(r^{2}+1)\leq 2(1/2)/((1/2)^{2}+1)=4/5

because f(r)=2r/(r2+1)f(r)=2r/(r^{2}+1) is increasing on (0,1)(0,1). This proves the claim, and thus the proposition. ∎

References

  • [Asc87] Michael Aschbacher, The 2727-dimensional module for E6E_{6}. I, Invent. Math. 89 (1987), no. 1, 159–195. MR 892190
  • [BCFvdG17] Jonas Bergström, Fabien Cléry, Carel Faber, and Gerard van der Geer, Siegel modular forms of degree two and three, http://smf.compositio.nl, Retrieved October 6, 2022.
  • [BD21] Siegfried Böcherer and Soumya Das, On fundamental fourier coefficients of siegel modular forms, Journal of the Institute of Mathematics of Jussieu (2021), 1–41.
  • [BGW15] Manjul Bhargava, Benedict H. Gross, and Xiaoheng Wang, Arithmetic invariant theory II: Pure inner forms and obstructions to the existence of orbits, Representations of reductive groups, Progr. Math., vol. 312, Birkhäuser/Springer, Cham, 2015, pp. 139–171. MR 3495795
  • [cDD+22] Fatma Çiçek, Giuliana Davidoff, Sarah Dijols, Trajan Hammonds, Aaron Pollack, and Manami Roy, The completed standard LL-function of modular forms on G2G_{2}, Math. Z. 302 (2022), no. 1, 483–517. MR 4462682
  • [Cle22] Jean-Louis Clerc, Construction à la Ibukiyama of symmetry breaking differential operators, J. Math. Sci. Univ. Tokyo 29 (2022), no. 1, 51–88. MR 4414247
  • [Cox46] H. S. M. Coxeter, Integral Cayley numbers, Duke Math. J. 13 (1946), 561–578. MR 19111
  • [CT20] Gaëtan Chenevier and Olivier Taïbi, Discrete series multiplicities for classical groups over 𝐙\bf Z and level 1 algebraic cusp forms, Publ. Math. Inst. Hautes Études Sci. 131 (2020), 261–323, https://otaibi.perso.math.cnrs.fr/levelone/siegel/siegel_genus3.txt. MR 4106796
  • [Dal21] Rahul Dalal, Counting Discrete, Level-11, Quaternionic Automorphic Representations on G2G_{2}, arXiv e-prints (2021), arXiv:2106.09313.
  • [EG96] Noam D. Elkies and Benedict H. Gross, The exceptional cone and the Leech lattice, Internat. Math. Res. Notices (1996), no. 14, 665–698. MR 1411589
  • [Gan00] Wee Teck Gan, A Siegel-Weil formula for exceptional groups, J. Reine Angew. Math. 528 (2000), 149–181. MR 1801660
  • [Gan22] by same author, Automorphic forms and the theta correspondence, Arizona Winter School notes (2022), https://swc-math.github.io/aws/2022/index.html.
  • [GGS02] Wee Teck Gan, Benedict Gross, and Gordan Savin, Fourier coefficients of modular forms on G2G_{2}, Duke Math. J. 115 (2002), no. 1, 105–169. MR 1932327
  • [Gro96] Benedict H. Gross, Groups over 𝐙{\bf Z}, Invent. Math. 124 (1996), no. 1-3, 263–279. MR 1369418
  • [GS98] Benedict H. Gross and Gordan Savin, Motives with Galois group of type G2G_{2}: an exceptional theta-correspondence, Compositio Math. 114 (1998), no. 2, 153–217. MR 1661756
  • [GS15] Nadya Gurevich and Avner Segal, The Rankin-Selberg integral with a non-unique model for the standard LL-function of G2G_{2}, J. Inst. Math. Jussieu 14 (2015), no. 1, 149–184. MR 3284482
  • [GS21] Wee Teck Gan and Gordan Savin, Howe duality and dichotomy for exceptional theta correspondences, 2021.
  • [GS22] by same author, The local langlands conjecture for g2g_{2}, 2022.
  • [GW94] Benedict H. Gross and Nolan R. Wallach, A distinguished family of unitary representations for the exceptional groups of real rank =4=4, Lie theory and geometry, Progr. Math., vol. 123, Birkhäuser Boston, Boston, MA, 1994, pp. 289–304. MR 1327538
  • [GW96] by same author, On quaternionic discrete series representations, and their continuations, J. Reine Angew. Math. 481 (1996), 73–123. MR 1421947
  • [HPS96] Jing-Song Huang, Pavle Pandžić, and Gordan Savin, New dual pair correspondences, Duke Math. J. 82 (1996), no. 2, 447–471. MR 1387237
  • [Ibu99] Tomoyoshi Ibukiyama, On differential operators on automorphic forms and invariant pluri-harmonic polynomials, Comment. Math. Univ. St. Paul. 48 (1999), no. 1, 103–118. MR 1684769
  • [Ibu02] by same author, Vanishing of vector valued siegel modular forms, Unpublished note (2002).
  • [Kim93] Henry H. Kim, Exceptional modular form of weight 44 on an exceptional domain contained in 𝐂27{\bf C}^{27}, Rev. Mat. Iberoamericana 9 (1993), no. 1, 139–200. MR 1216126
  • [KZ81] W. Kohnen and D. Zagier, Values of LL-series of modular forms at the center of the critical strip, Invent. Math. 64 (1981), no. 2, 175–198. MR 629468
  • [MP22] Finley McGlade and Aaron Pollack, Quaternionic functions and applications, Work in progress (2022).
  • [MS97] K. Magaard and G. Savin, Exceptional Θ\Theta-correspondences. I, Compositio Math. 107 (1997), no. 1, 89–123. MR 1457344
  • [Mui97] Goran Muić, The unitary dual of pp-adic G2G_{2}, Duke Math. J. 90 (1997), no. 3, 465–493. MR 1480543
  • [oIS] The On-Line Encyclopedia of Integer Sequences, A100861: Triangle of bessel numbers read by rows, https://oeis.org/A100861.
  • [Pol18] Aaron Pollack, Lifting laws and arithmetic invariant theory, Camb. J. Math. 6 (2018), no. 4, 347–449. MR 3870360
  • [Pol19] by same author, Modular forms on G2{G}_{2} and their standard L{L}-function, Proceedings of the Simons Symposium ”Relative Trace Formulas” (accepted) (2019).
  • [Pol20a] by same author, The Fourier expansion of modular forms on quaternionic exceptional groups, Duke Math. J. 169 (2020), no. 7, 1209–1280. MR 4094735
  • [Pol20b] by same author, The minimal modular form on quaternionic E8{E}_{8}, Jour. Inst. Math. Juss. (accepted) (2020).
  • [Pol21] by same author, Exceptional algebraic structures and applications, Notes from a topics course at UCSD (2021), https://www.math.ucsd.edu/~apollack/course_notes.pdf.
  • [Pol22] by same author, Modular forms on exceptional groups, Arizona Winter School notes (2022), https://swc-math.github.io/aws/2022/index.html.
  • [Pol23] by same author, Exceptional siegel weil theorems for compact Spin8\mathrm{{S}pin}_{8}, Preprint (2023).
  • [PV21] Kartik Prasanna and Akshay Venkatesh, Automorphic cohomology, motivic cohomology, and the adjoint LL-function, Astérisque (2021), no. 428, viii+132. MR 4372499
  • [Seg17] Avner Segal, A family of new-way integrals for the standard \mathcal{L}-function of cuspidal representations of the exceptional group of type G2G_{2}, Int. Math. Res. Not. IMRN (2017), no. 7, 2014–2099.
  • [Wal81] J.-L. Waldspurger, Sur les coefficients de Fourier des formes modulaires de poids demi-entier, J. Math. Pures Appl. (9) 60 (1981), no. 4, 375–484. MR 646366
  • [Wal03] Nolan R. Wallach, Generalized Whittaker vectors for holomorphic and quaternionic representations, Comment. Math. Helv. 78 (2003), no. 2, 266–307. MR 1988198
  • [Wei] Eric W. Weisstein, Modified bessel function of the second kind, From MathWorld–A Wolfram Web Resource. https://mathworld.wolfram.com/ModifiedBesselFunctionoftheSecondKind.html.