1 Introduction

Let F be a global field with the ring of adeles \({\mathbb {A}}\). We assume that the characteristics of F is not 2. We present in this paper a Shimura type integral on the exceptional group \(G_2\) which represents the L-function

$$\begin{aligned} L(s,\widetilde{\pi }\times (\chi \otimes \tau ))L(s,\widetilde{\pi }\otimes (\chi \otimes \omega _\tau )), \end{aligned}$$

where \(\widetilde{\pi }\) is an irreducible genuine cuspidal representation of \(\widetilde{\mathrm {SL}}_2({\mathbb {A}}) \), \(\tau \) is an irreducible generic cuspidal representation of \({\mathrm {GL}}_2({\mathbb {A}})\) and \(\chi \) is the quadratic character of \(F^\times \backslash {\mathbb {A}}^\times \) defined by \(\chi (a)=\prod _v (a_v,-1)_{F_v}\), where \(a=(a_v)_v \in {\mathbb {A}}^\times \) and \((~,~)_{F_v}\) is the Hilbert symbol on \(F_v\).

To give more details about the integral, we introduce some notations. The group \(G_2\) has two simple roots and we label the short root by \(\alpha \) and the long root by \(\beta \). Let \(P=MV\) (resp. \(P'=M'V'\)) be the maximal parabolic subgroup of \(G_2\) such that the root space of \(\beta \) is in the Levi M (resp. the root space of \(\alpha \) is in the Levi \(M'\)). The Levi subgroups M and \(M'\) are isomorphic to \({\mathrm {GL}}_2\). Let J be the subgroup of P which is isomorphic to \({\mathrm {SL}}_2 < imes V\). Let \(\widetilde{\mathrm {SL}}_2({\mathbb {A}})\) be the metaplectic double cover of \( {\mathrm {SL}}_2({\mathbb {A}})\). There is a Weil representation \(\omega _\psi \) of \(\widetilde{\mathrm {SL}}_2({\mathbb {A}})\) for a nontrivial additive character \(\psi \) of \(F\backslash {\mathbb {A}}\). Let \(\widetilde{\theta }_\phi \) be a corresponding theta series associated with a function \(\phi \in {\mathcal {S}}({\mathbb {A}})\). Let \(\tau \) be an irreducible cuspidal automorphic representations of \({\mathrm {GL}}_2({\mathbb {A}})\). For \(f_s\in {\mathrm {Ind}}_{P'({\mathbb {A}})}^{G_2({\mathbb {A}})}(\tau \otimes \delta _{P'}^s)\), we can form an Eisenstein series \(E(g,f_s)\) on \(G_2({\mathbb {A}})\). Let \(\widetilde{\pi }\) be an irreducible genuine cuspidal automorphic forms of \(\widetilde{\mathrm {SL}}_2({\mathbb {A}})\). For a cusp form \(\widetilde{\varphi }\in \widetilde{\pi }\), we consider the integral

$$\begin{aligned} I(\widetilde{\varphi }, \phi ,f_s)=\int _{{\mathrm {SL}}_2(F)\backslash {\mathrm {SL}}_2({\mathbb {A}})}\int _{V(F)\backslash V({\mathbb {A}})}\widetilde{\varphi }(g)\widetilde{\theta }_\phi (vg)E(vg,f_s)dvdg. \end{aligned}$$

Our main result is the following

Theorem 1.1

The above integral is absolutely convergent for \({\mathrm {Re}}(s)\gg 0\) and can be meromorphically continued to all \(s\in {\mathbb {C}}\). When \({\mathrm {Re}}(s)\gg 0\), the integral \(I(\widetilde{\varphi },\phi ,f_s)\) is Eulerian. Moreover, at an unramified place v, the local integral represents the L-function

$$\begin{aligned} \frac{L(3s-1,\widetilde{\pi }_v\times (\chi _v\otimes \tau _v))L(6s-5/2,\widetilde{\pi }_v\otimes (\chi _v\otimes \omega _{\tau _v}))}{L(3s-1/2,\tau _v)L(6s-2,\omega _{\tau _v})L(9s-7/2,\tau _v\otimes \omega _{\tau _v})}. \end{aligned}$$

This is Theorem 3.1 and Proposition 4.6. We remark that Ginzburg–Rallis–Soudry gave integral representations for L-functions of generic cuspidal representations of \( \widetilde{{\mathrm {Sp}}}_{2n}\times {\mathrm {GL}}_m\) in [8] using symplectic groups. It is still interesting to have different integral representations. As an application of Theorem 1.1, we show that if \(Wd_\psi (\widetilde{\pi })=\chi \otimes \tau \), then a Shimura type period with respect to \(\widetilde{\pi }\) and the residue of Eisenstein series on \(G_2\) is non-vanishing, where \(Wd_\psi \) is the Shimura–Waldspurger lift. It is an interesting theme in number theory to investigate the relations between poles of L-functions and non-vanishing of automorphic periods. There are many examples of this kind relations. See [5, 7, 9] for some examples. The non-vanishing results of automorphic periods have many interesting applications in automorphic forms. We expect the non-vanishing period in our case would be useful on problems related to the residue spectrum of \(G_2\).

There are several known Rankin–Selberg integrals on \(G_2\) which represents different L-functions and have many applications, see [4,5,6] for example. The integral \(I(\widetilde{\varphi },\phi ,f_s)\) can be viewed as a dual integral of the standard \(G_2\) L-function integral in [5] in the following sense. The integral \(I(\widetilde{\varphi },\phi ,f_s)\) is an integral of a triple product of a cusp form on \(\widetilde{\mathrm {SL}}_2({\mathbb {A}})\), a theta series and an Eisenstein series on \(G_2({\mathbb {A}})\), while the integral in [5] is an integral of a triple product of a cusp form on \(G_2({\mathbb {A}})\), a theta series and an Eisenstein series on \(\widetilde{\mathrm {SL}}_2({\mathbb {A}})\). The integral in [6] is also in a similar pattern, which is an integral of a triple product of a cusp form on \({\mathrm {SL}}_2({\mathbb {A}})\), a theta series and an Eisenstein series on a cover of \(G_2({\mathbb {A}})\). The results presented here were known for D. Ginzburg. But we still think that it might be useful to write up the details.

2 The group \({ {G}}_2\)

2.1 Roots and Weyl group for \({ {G}}_2\)

Let \({{G}}_2\) be the split algebraic reductive group of type \({{G}}_2\) (defined over \({\mathbb {Z}}\)). The group \({{G}}_2\) has two simple roots, the short root \(\alpha \) and the long root \(\beta \). The set of the positive roots is \(\Sigma ^+=\left\{ {\alpha , \beta , \alpha +\beta , 2\alpha +\beta , 3\alpha +\beta , 3\alpha +2\beta }\right\} \). Let ( ,  ) be the inner product in the root system and \(\langle {~,~} \rangle \) be the pair defined by \(\langle {\gamma _1,\gamma _2} \rangle =\frac{2(\gamma _1,\gamma _2)}{(\gamma _2,\gamma _2)}\). For the root space \({{G}}_2\), we have the relations:

$$\begin{aligned} \langle {\alpha ,\beta } \rangle =-1, \langle {\beta , \alpha } \rangle =-3. \end{aligned}$$

For a root \(\gamma \), let \(s_\gamma \) be the reflection defined by \(\gamma \), i.e., \(s_\gamma (\gamma ')=\gamma '-\langle {\gamma ',\gamma } \rangle \gamma \). We have the relation

$$\begin{aligned} s_\alpha (\beta )=3\alpha +\beta , s_\beta (\alpha )=\alpha +\beta . \end{aligned}$$

The Weyl group \(\mathbf{{W}}=\mathbf{{W}}({{G}}_2)\) of \({{G}}_2\) has 12 elements, which is explicitly given by

$$\begin{aligned} \mathbf{{W}}= & {} \left\{ 1, s_\alpha , s_\beta , s_\alpha s_\beta , s_\beta s_\alpha , s_\alpha s_\beta s_\alpha , s_\beta s_\alpha s_\beta ,\right. \\&\quad \left. (s_\alpha s_\beta )^2, (s_\beta s_\alpha )^2, s_\beta (s_\alpha s_\beta )^2, s_\alpha (s_\beta s_\alpha )^2, (s_\alpha s_\beta )^3\right\} . \end{aligned}$$

For a root \(\gamma \), let \(U_\gamma \subset G\) be the root space of \(\gamma \), and let \(\mathbf{{x}}_\gamma : F\rightarrow U_\gamma \) be a fixed isomorphism which satisfies various Chevalley relations, see Chapter 3 of [14]. Among other things, \(\mathbf{{x}}_{\gamma }\) satisfies the following commutator relations:

$$\begin{aligned} \begin{aligned} {[}\mathbf{{x}}_\alpha (x), \mathbf{{x}}_\beta (y)]&= \mathbf{{x}}_{\alpha +\beta }(-xy)\mathbf{{x}}_{2\alpha +\beta }(-x^2 y) \mathbf{{x}}_{3\alpha +\beta }( x^3 y)\mathbf{{x}}_{3\alpha +2\beta }(-2x^3 y^2)\\ {[}\mathbf{{x}}_\alpha (x), \mathbf{{x}}_{\alpha +\beta }(y)]&= \mathbf{{x}}_{2\alpha +\beta }( -2xy)\mathbf{{x}}_{3\alpha +\beta }(3x^2 y) \mathbf{{x}}_{3\alpha +2\beta }(3xy^2) \\ {[}\mathbf{{x}}_\alpha (x), \mathbf{{x}}_{2\alpha +\beta }(y)]&=\mathbf{{x}}_{3\alpha +\beta }( 3xy) \\ {[}\mathbf{{x}}_\beta (x), \mathbf{{x}}_{3\alpha +\beta }(y)]&=\mathbf{{x}}_{3\alpha +2\beta }(xy) \\ {[}\mathbf{{x}}_{\alpha +\beta }(x), \mathbf{{x}}_{2\alpha +\beta }(y)]&=\mathbf{{x}}_{3\alpha +2\beta }(3xy). \end{aligned} \end{aligned}$$
(2.1)

For all the other pairs of positive roots \(\gamma _1, \gamma _2\), we have \([\mathbf{{x}}_{\gamma _1}(x), \mathbf{{x}}_{\gamma _2}(y)]=1\). Here \([g_1,g_2]=g_1^{-1}g_2^{-1}g_1g_2\) for \(g_1,g_2\in G_2\). For these commutator relationships, see [12].

Following [14], we denote \(w_\gamma (t)= \mathbf{{x}}_{\gamma }(t)\mathbf{{x}}_{-\gamma }(-t^{-1})\mathbf{{x}}_{\gamma }(t)\) and \(w_\gamma =w_\gamma (1)\). Note that \(w_\gamma \) is a representative of \(s_\gamma \). Let \(h_\gamma (t)=w_\gamma (t) w_\gamma ^{-1}\). Let T be the subgroup of G which consists of elements of the form \(h_\alpha (t_1)h_\beta (t_2), t_1,t_2\in T\) and U be the subgroup of \(G_2\) generated by \(U_\gamma \) for all \(\gamma \in \Sigma ^+\). Let \(B=TU\), which is a Borel subgroup of \(G_2\).

For \(t_1,t_2\in \mathbb {G}_m,\) denote \(h(t_1,t_2)=h_\alpha (t_1t_2)h_\beta (t_1^2t_2)\). From the Chevalley relation \(h_{\gamma _1}(t)\mathbf{{x}}_{\gamma _2}(r) h_{\gamma _1}(t)^{-1}=\mathbf{{x}}_{\gamma _2}(t^{\langle {\gamma _2,\gamma _1} \rangle } r)\) (see [14, Lemma 20, (c)]), we can check the following relations

$$\begin{aligned} \begin{aligned} h^{-1}(t_1,t_2)\mathbf{{x}}_\alpha (r)h(t_1,t_2)&=\mathbf{{x}}_\alpha (t_2^{-1} r), \\ h^{-1}(t_1,t_2)\mathbf{{x}}_\beta (r)h(t_1,t_2)&=\mathbf{{x}}_\beta (t_1^{-1}t_2 r)\\ h^{-1}(t_1,t_2)\mathbf{{x}}_{\alpha +\beta }(r) h(t_1,t_2)&=\mathbf{{x}}_{\alpha +\beta }(t_1^{-1}r),\\ h^{-1}(t_1,t_2)\mathbf{{x}}_{2\alpha +\beta }(r) h(t_1,t_2)&= \mathbf{{x}}_{2\alpha +\beta }(t_1^{-1}t_2^{-1}r) \\ h^{-1}(t_1,t_2)\mathbf{{x}}_{3\alpha +\beta }(r)h(t_1,t_2)&=\mathbf{{x}}_{3\alpha +\beta }(t_1^{-1}t_2^{-2}r),\\ h^{-1}(t_1,t_2) \mathbf{{x}}_{3\alpha +2\beta }(r) h(t_1,t_2)&=\mathbf{{x}}_{3\alpha +2\beta }(t_1^{-2} t_2^{-1} r). \end{aligned} \end{aligned}$$
(2.2)

Thus the notation h(ab) agrees with that of [5].

One can also check that

$$\begin{aligned} w_\alpha h(t_1,t_2) w_\alpha ^{-1}= h(t_1 t_2, t_2^{-1}), \quad w_\beta h(t_1,t_2) w_\beta ^{-1}= h(t_2,t_1). \end{aligned}$$

2.2 Subgroups

Let F be a field and denote \(G=G_2(F)\). The group G has two proper parabolic subgroups. Let \(P=M < imes V\) be the parabolic subgroup of G such that \(U_\beta \subset M\cong {\mathrm {GL}}_2\). Thus the unipotent subgroup V is consisting of root spaces of \(\alpha , \alpha +\beta , 2\alpha +\beta , 3\alpha +\beta ,3\alpha +2\beta \), and a typical element of V is of the form

$$\begin{aligned} \mathbf{{x}}_{\alpha }(r_1)\mathbf{{x}}_{\alpha +\beta }(r_2)\mathbf{{x}}_{2\alpha +\beta }(r_3)\mathbf{{x}}_{3\alpha +\beta }(r_4)\mathbf{{x}}_{3\alpha +2\beta }(r_5), r_i\in F. \end{aligned}$$

To ease the notation, we will write the above element as \([r_1,r_2,r_3,r_4,r_5]\). Denote by J the following subgroup of P

$$\begin{aligned} J={\mathrm {SL}}_2(F) < imes V. \end{aligned}$$

Let \( V_1\) (resp. Z) be the subgroup of V which consists root spaces of \(3\alpha +\beta \) and \(3\alpha +2\beta \) (resp. \(2\alpha +\beta , 3\alpha +\beta \) and \(3\alpha +2\beta \)). Note that P and hence J normalizes \(V_1\) and Z. We will always view \({\mathrm {SL}}_2(F)\) as a subgroup of G via the inclusion \({\mathrm {SL}}_2(F)\subset M\). Denote by \(A_{{\mathrm {SL}}_2}\), \(N_{{\mathrm {SL}}_2}\) and \(B_{{\mathrm {SL}}_2}\) the standard torus, the upper triangular unipotent subgroup and the upper triangular Borel subgroup of \({\mathrm {SL}}_2(F)\). Note that the torus element h(ab) can be identified with

$$\begin{aligned} \begin{pmatrix}a&{}\\ &{}b \end{pmatrix}\in {\mathrm {GL}}_2(F)\cong M, \end{aligned}$$

and thus \(A_{{\mathrm {SL}}_2}=\left\{ {h(a,a^{-1})|a\in F^\times }\right\} \) and \(B_{{\mathrm {SL}}_2}=A_{{\mathrm {SL}}_2} < imes U_\beta \).

Let \(P'=M'V'\) be the other maximal parabolic subgroup G with \(U_\alpha \) in the Levi subgroup \(M'\). The Levi \(M'\) is isomorphic to \({\mathrm {GL}}_2(F)\), and from relations in (2.2), one can check that one isomorphism \(M'\cong {\mathrm {GL}}_2(F)\) can be determined by

$$\begin{aligned} \begin{aligned}\mathbf{{x}}_\alpha (r)&\mapsto \begin{pmatrix}1&{} r\\ &{}1 \end{pmatrix},\\ h(a,b)&\mapsto \begin{pmatrix}ab&{}\\ &{}a \end{pmatrix}. \end{aligned} \end{aligned}$$

In particular, we see that \(h(a,1)\in T\subset M'\) can be identified with \({\mathrm {diag}}(a,a).\) Let \(\delta _{P'}\) be the modulus character of \(P'\). One can check that \(\delta _{P'}(m')=|\det (m')|^3\) for \(m'\in M'\), where \(\det (m')\) can be computed using the above isomorphism \(M'\cong {\mathrm {GL}}_2(F)\).

2.3 Weil representation of \(\widetilde{\mathrm {SL}}_2({\mathbb {A}}) < imes V({\mathbb {A}})\)

In this subsection, we assume that F is a global field and \({\mathbb {A}}\) is its ring of adeles. In \({\mathrm {SL}}_2(F)\), we denote \(t(a)={\mathrm {diag}}(a,a^{-1}),a\in F^\times \) and

$$\begin{aligned} n(b)=\begin{pmatrix}1&{} b\\ &{}1 \end{pmatrix}, b\in F. \end{aligned}$$

Denote \(w^1=\begin{pmatrix}&{}1\\ -1&{} \end{pmatrix}\), which represents the unique nontrivial Weyl element of \({\mathrm {SL}}_2(F)\). Under the embedding \({\mathrm {SL}}_2(F)\subset M\subset G\), the element \(w^1\) can be identified with \(w_\beta \).

Let \(\widetilde{\mathrm {SL}}_2({\mathbb {A}})\) be the metaplectic double cover of \({\mathrm {SL}}_2({\mathbb {A}})\). Then we have an exact sequence

$$\begin{aligned} 0\rightarrow \mu _2\rightarrow \widetilde{\mathrm {SL}}_2({\mathbb {A}}) \rightarrow {\mathrm {SL}}_2({\mathbb {A}})\rightarrow 0, \end{aligned}$$

where \(\mu _2=\left\{ {\pm 1}\right\} \).

We will identify \({\mathrm {SL}}_2({\mathbb {A}})\) with the symplectic group of \({\mathbb {A}}^2\) with symplectic structure defined by

$$\begin{aligned} \langle {(x_1,y_1),(x_2,y_2)} \rangle =-2x_1y_2+2x_2y_1. \end{aligned}$$

Let \({\mathscr {H}}({\mathbb {A}})\) be the Heisenberg group of the symplectic space \(({\mathbb {A}}^2, \langle {~,~} \rangle )\), i.e., \({\mathscr {H}}({\mathbb {A}})={\mathbb {A}}^3\) with group law

$$\begin{aligned} (x_1,y_1,z_1)(x_2,y_2,z_2)=(x_1+x_2,y_1+y_2, z_1+z_2-x_1y_2+y_1 x_2). \end{aligned}$$

Let \({\mathrm {SL}}_2({\mathbb {A}})\) act on \({\mathscr {H}}({\mathbb {A}})\) from the right side by

$$\begin{aligned} (x_1, y_1, z_1).g=((x_1,y_1)g, z_1), g\in {\mathrm {SL}}_2({\mathbb {A}}), \end{aligned}$$

where \((x_1,y_1)g\) is the usual matrix multiplication.

We then can form the semi-direct product \({\mathrm {SL}}_2({\mathbb {A}}) < imes {\mathscr {H}}({\mathbb {A}}),\) where the product is defined by

$$\begin{aligned} (g_1, h_1)(g_2,h_2)=(g_1g_2, (h_1.g_2 ) h_2), g_i\in {\mathrm {SL}}_2({\mathbb {A}}), h_i\in {\mathscr {H}}({\mathbb {A}}), i=1,2. \end{aligned}$$

Let \(\psi \) be a nontrivial additive character of \(F\backslash {\mathbb {A}}\). Then there is a Weil representation \(\omega _\psi \) of \(\widetilde{\mathrm {SL}}_2({\mathbb {A}}) < imes {\mathscr {H}}({\mathbb {A}})\). The space of \(\omega _\psi \) is \({\mathcal {S}}({\mathbb {A}})\), the Bruhat–Schwartz functions on \({\mathbb {A}}\).

For \(\phi \in {\mathcal {S}}({\mathbb {A}})\), we have the well-know formulas:

$$\begin{aligned} (\omega _\psi (n(b))\phi )(x)&=\psi (bx^2)\phi (x), b\in {\mathbb {A}}\\ (\omega _{\psi }((r_1,r_2,r_3))\phi )(x)&=\psi \left( r_3-2xr_2-r_1r_2\right) \phi (x+r_1), (r_1,r_2,r_3)\in {\mathscr {H}}({\mathbb {A}}), \end{aligned}$$

The above formulas could be found in [11].

Recall that for \(r_1,r_2,r_3,r_4,r_5\in {\mathbb {A}}\), the notation \([r_1,r_2,r_3,r_4,r_5]\) is an abbreviation of

$$\begin{aligned} \mathbf{{x}}_\alpha (r_1)\mathbf{{x}}_{\alpha +\beta }(r_2)\mathbf{{x}}_{2\alpha +\beta }(r_3)\mathbf{{x}}_{3\alpha +\beta }(r_4) \mathbf{{x}}_{3\alpha +2\beta }(r_5)\in V({\mathbb {A}}). \end{aligned}$$

Define a map \({\mathrm {pr}}: V({\mathbb {A}})\rightarrow {\mathscr {H}}({\mathbb {A}})\)

$$\begin{aligned} {\mathrm {pr}}([r_1,r_2,r_3,r_4,r_5])=(r_1,r_2,r_3-r_1 r_2) . \end{aligned}$$

From the commutator relation (2.1), we can check that \({\mathrm {pr}}\) is a group homomorphism and defines an exact sequence

$$\begin{aligned} 0\rightarrow V_1({\mathbb {A}}) \rightarrow V({\mathbb {A}})\rightarrow {\mathscr {H}}({\mathbb {A}})\rightarrow 0. \end{aligned}$$

Recall that \(V_1\) is the subgroup of V which is generated by the root space of \(3\alpha +\beta , 3\alpha +2\beta \). Note that there is a typo in the formula of the projection map \({\mathrm {pr}}\) in [5, p.316].

For \(g=\begin{pmatrix}a &{} b\\ c&{} d \end{pmatrix}\in {\mathrm {SL}}_2(F) \subset M\), we can check that

$$\begin{aligned} g^{-1}[r_1,r_2,r_3,0,0] g=[r_1', r_2', r_3', r_4',r_5'], \end{aligned}$$

where \(r_1'=a r_1- cr_2, r_2'=-b r_1 + dr_2, r_3'-r_1'r_2'=r_3-r_1r_2\).

Consider the map \(\overline{{\mathrm {pr}}}: J({\mathbb {A}})={\mathrm {SL}}_2({\mathbb {A}}) < imes V({\mathbb {A}})\rightarrow {\mathrm {SL}}_2({\mathbb {A}}) < imes {\mathscr {H}}({\mathbb {A}})\),

$$\begin{aligned} (g,v)\mapsto (g^*, {\mathrm {pr}}(v)), g\in {\mathrm {SL}}_2({\mathbb {A}}), v\in V({\mathbb {A}}). \end{aligned}$$

where \(g^*=\begin{pmatrix} a&{} -b \\ -c &{} d \end{pmatrix}=d_1 g d_1^{-1}\), where \(d_1={\mathrm {diag}}(1,-1)\in {\mathrm {GL}}_2(F)\). From the above discussion, the map \(\overline{{\mathrm {pr}}}\) is a group homomorphism and its kernel is also \( V_1({\mathbb {A}})\). We will also view \(\overline{\text {pr}}\) as a homomorphism \(\widetilde{{\mathrm {SL}}}_2({\mathbb {A}}) < imes V({\mathbb {A}})\rightarrow \widetilde{{\mathrm {SL}}}_2({\mathbb {A}}) < imes {\mathscr {H}}({\mathbb {A}})\).

In the following, we will also view \(\omega _\psi \) as a representation of \(\widetilde{{\mathrm {SL}}_2}({\mathbb {A}}) < imes V({\mathbb {A}})\) via the projection map \(\overline{{\mathrm {pr}}}\). For \(\phi \in {\mathcal {S}}({\mathbb {A}})\), we form the theta series

$$\begin{aligned} \widetilde{\theta }_\phi (vg)=\sum _{\xi \in F}\omega _\psi (vh)\phi (\xi ),v\in V({\mathbb {A}}),g\in \widetilde{\mathrm {SL}}_2({\mathbb {A}}). \end{aligned}$$

Note that given a genuine cusp form \(\widetilde{\varphi }\) on \(\widetilde{\mathrm {SL}}_2({\mathbb {A}})\), the product

$$\begin{aligned} \widetilde{\varphi }(g)\widetilde{\theta }_\phi (vg),v\in V({\mathbb {A}}),g\in \widetilde{\mathrm {SL}}_2({\mathbb {A}}) \end{aligned}$$

can be viewed as a function on \(J({\mathbb {A}})={\mathrm {SL}}_2({\mathbb {A}}) < imes V({\mathbb {A}})\).

2.4 An Eisenstein series on \({ {G}}_2\)

In this subsection and in the rest of the paper, every representation appeared is assumed to be irreducible. Let \(\tau \) be a cuspidal automorphic representation on \({\mathrm {GL}}_2({\mathbb {A}})\). We will view \(\tau \) as a representation of \(M'({\mathbb {A}})\) via the identification \(M'\cong {\mathrm {GL}}_2\). We then consider the induced representation \(I(s,\tau )={\mathrm {Ind}}_{P'({\mathbb {A}})}^{{{G}}_2({\mathbb {A}})}(\tau \otimes \delta _{P'}^s)\). A section \(f_s\in I(s,\tau )\) is a smooth function satisfying

$$\begin{aligned} f_s(v'm'g)=\delta _{P'}(m')^sf_s(g), \forall v'\in V'({\mathbb {A}}), m'\in M'({\mathbb {A}}),g\in { {G}}_2({\mathbb {A}}). \end{aligned}$$

For \(f_s\in I(s,\tau )\), we consider the Eisenstein series

$$\begin{aligned} E(g,f_s)=\sum _{\delta \in P'(F){\setminus }{{G}}_2(F)}f_s(\delta g),g\in {{G}}_2({\mathbb {A}}). \end{aligned}$$

3 A global integral

Let \(\widetilde{\pi }\) be a genuine cuspidal automorphic representation on \(\widetilde{\mathrm {SL}}_2({\mathbb {A}}),\) and \(\tau \) be a cuspidal automorphic representation of \({\mathrm {GL}}_2({\mathbb {A}})\). For \(\widetilde{\varphi }\in V_\pi , \phi \in {\mathcal {S}}({\mathbb {A}})\) and \(f_s\in I(s,\tau )\), we consider the integral

$$\begin{aligned} I(\widetilde{\varphi },\phi ,f_s)=\int _{{\mathrm {SL}}_2(F){\setminus }{\mathrm {SL}}_2({\mathbb {A}})}\int _{V(F)\setminus V({\mathbb {A}})}\widetilde{\varphi }(g) \widetilde{\theta }_\phi (vg) E(vg,f_s)dvdg. \end{aligned}$$

Let \(\gamma =w_\beta w_\alpha w_\beta w_\alpha \in {{G}}_2(F)\).

Theorem 3.1

The integral \(I(\widetilde{\varphi },\phi ,f_s)\) is absolutely convergent when \({\mathrm {Re}}(s)\gg 0\) and can be meromorphically continued to all \(s\in {\mathbb {C}}\). Moreover, when \({\mathrm {Re}}(s)\gg 0\), we have

$$\begin{aligned} I(\tilde{\varphi },\phi ,f_s)=\int _{N_{{\mathrm {SL}}_2}({\mathbb {A}}){\setminus }{\mathrm {SL}}_2({\mathbb {A}})}\int _{U_{\alpha +\beta }({\mathbb {A}}){\setminus }V({\mathbb {A}})} W_{\widetilde{\varphi }}(g)\omega _\psi (vg)\phi (1)W_{f_s}(\gamma vg)dvdg, \end{aligned}$$

where

$$\begin{aligned} W_{\widetilde{\varphi }}(g)=\int _{F{\setminus }{\mathbb {A}}}\widetilde{\varphi }(\mathbf{{x}}_\beta (r)g)\psi (r)dr, \end{aligned}$$

and

$$\begin{aligned} W_{f_s}(\gamma vg)=\int _{F{\setminus }{\mathbb {A}}}f_s(\mathbf{{x}}_\alpha (r)\gamma vg)\psi (-2r)dr. \end{aligned}$$

Proof

The first assertion is standard. We only show that the above integral is Eulerian when \({\mathrm {Re}}(s)\gg 0\). Unfolding the Eisenstein series, we can get

$$\begin{aligned} I(\widetilde{\varphi },\phi ,f_s)=\sum _{\delta \in P'(F){\setminus }{{G}}_2(F)/P(F)}\int _{{\mathrm {SL}}_2^\delta (F){\setminus }{\mathrm {SL}}_2({\mathbb {A}})}\int _{V^\delta (F){\setminus }V({\mathbb {A}})}\widetilde{\varphi }(g)\widetilde{\theta }_\phi (vg)f_s(\delta vg)dvdg, \end{aligned}$$

where \(X^\delta =\delta ^{-1} P'\delta \cap X\) for \(X\subset G_2(F)\). We can check that a set of representatives of the double coset \(P'(F){\setminus }{{G}}_2(F)/P(F)\) can be taken as \(\left\{ {1,w_\beta w_\alpha , \gamma = w_\beta w_\alpha w_\beta w_\alpha }\right\} .\) For \(\delta =1, w_\beta w_\alpha ,\) or \( \gamma =w_\beta w_\alpha w_\beta w_\alpha \), denote

$$\begin{aligned} I_\delta =\int _{{\mathrm {SL}}_2^\delta (F){\setminus }{\mathrm {SL}}_2({\mathbb {A}})}\int _{V^\delta (F)\setminus V({\mathbb {A}})}\widetilde{\varphi }(g)\widetilde{\theta }_\phi (vg)f_s(\delta vg)dvdg. \end{aligned}$$

If \(\delta =1\), the above integral \(I_\delta \) has an inner integral

$$\begin{aligned} \int _{U_{2\alpha +\beta }(F){\setminus }U_{2\alpha +\beta }({\mathbb {A}})}\widetilde{\theta }_\phi (\mathbf{{x}}_{2\alpha +\beta }(r)vg)f_s(\mathbf{{x}}_{2\alpha +\beta }(r)vg)dr, \end{aligned}$$

which is zero because \(f_s(\mathbf{{x}}_{2\alpha +\beta }(r)vg)=f_s(vg) , \widetilde{\theta }_\phi (\mathbf{{x}}_{2\alpha +\beta }(r)vg)=\psi (r)\tilde{\theta }_\phi (vg)\) and \(\int _{F{\setminus }{\mathbb {A}}}\psi (r)dr=0.\) The last equation follows from the fact that \(\psi \) is non-trivial.

We next consider the term when \(\delta =w_\beta w_\alpha \). We write

$$\begin{aligned} \widetilde{\theta }_\phi (vg)=\omega _\psi (vg)\phi (0)+\sum _{\xi \in F^\times }\omega _\psi (vg)\phi (\xi ). \end{aligned}$$

The contribution of the first term to the integral \(I_{\delta }\) is

$$\begin{aligned} \int _{{\mathrm {SL}}_2^\delta (F){\setminus }{\mathrm {SL}}_2({\mathbb {A}})}\int _{V^\delta (F){\setminus } V({\mathbb {A}})}\widetilde{\varphi }(g)\omega _\psi (vg)\phi (0)f_s(\delta vg)dvdg. \end{aligned}$$

Note that \(\delta \mathbf{{x}}_\beta (r)\delta ^{-1}\subset U_{2\alpha +\beta }\subset V'\), we have \(f_s(\delta v \mathbf{{x}}_\beta (r)g)=f_s(\delta \mathbf{{x}}_\beta (-r)v\mathbf{{x}}_\beta (r)g)\). On the other hand, we have \(\omega _\psi (\mathbf{{x}}_\beta (r)vg)\phi (0)=\omega _\psi (vg)\phi (0)\) . After a changing variable on v, we can see that the above integral contains an inner integral

$$\begin{aligned} \int _{F{\setminus }{\mathbb {A}}}\widetilde{\varphi }(\mathbf{{x}}_\beta (r)vg)dr, \end{aligned}$$

which is zero since \(\widetilde{\varphi }\) is cuspidal. Thus the contribution of the term \(\omega _\psi (vg)\phi (0)\) is zero when \(\delta =w_\beta w_\alpha \). The contribution of \(\sum _{\xi \in F^\times }\omega _\psi (vg)\phi (\xi ) \) is

$$\begin{aligned} \int _{{\mathrm {SL}}_2^\delta (F){\setminus }{\mathrm {SL}}_2({\mathbb {A}})}\int _{V^\delta (F){\setminus } V({\mathbb {A}})}\widetilde{\varphi }(g) \sum _{\xi \in F^\times }\omega _\psi (vg)\phi (\xi )f_s(\delta vg)dvdg. \end{aligned}$$

We consider the inner integral on \(U_{\alpha +\beta }(F){\setminus }U_{\alpha +\beta }({\mathbb {A}})\). Note that \(U_{\alpha +\beta }\subset V\) and \(\delta U_{\alpha +\beta } \delta ^{-1}=U_{2\alpha +\beta }\subset V'\), we get \(f_s(\delta \mathbf{{x}}_{\alpha +\beta }(r)vg)=f_s(\delta vg)\). On the other hand, we have \(\omega _{\psi }(\mathbf{{x}}_{\alpha +\beta }(r)vg)\phi (\xi )=\psi (-2r\xi )\omega _\psi (vg)\phi (\xi )\). Thus the above integral has an inner integral

$$\begin{aligned} \int _{F{\setminus }{\mathbb {A}}}\sum _{\xi \in F^\times }\psi (-2r\xi )\omega _\psi (vg)\phi (\xi )dr=\sum _{\xi \in F^\times }\omega _\psi (vg)\phi (\xi )\int _{F{\setminus }{\mathbb {A}}}\psi (-2r\xi )dr=0. \end{aligned}$$

Thus when \(\delta =w_\beta w_\alpha \), the corresponding term is zero. Thus we get

$$\begin{aligned} I(\widetilde{\varphi },\phi ,f_s)=\int _{{\mathrm {SL}}_2^\gamma (F){\setminus }{\mathrm {SL}}_2({\mathbb {A}})}\int _{V^\gamma (F){\setminus }V({\mathbb {A}})}\widetilde{\varphi }(g)\widetilde{\theta }_\phi (vg)f_s(\gamma vg)dvdg. \end{aligned}$$

We have \({\mathrm {SL}}_2^\gamma =B_{{\mathrm {SL}}_2}\) and \(V^\gamma =U_{\alpha +\beta }\). We decompose \(\widetilde{\theta }_\phi \) as

$$\begin{aligned} \widetilde{\theta }_\phi (vg)=\omega _\psi (vg)\phi (0)+\sum _{\xi \in F^\times }\omega _\psi (vg)\phi (\xi )=\omega _\psi (vg)\phi (0)+\sum _{a\in F^\times }\omega _\psi (t(a)vg)\phi (1). \end{aligned}$$

Recall that \(t(a)={\mathrm {diag}}(a,a^{-1})\). Since \(\gamma U_\beta \gamma ^{-1}\subset U_{3\alpha +\beta }\subset V'\), we have

$$\begin{aligned} f_s(\gamma v\mathbf{{x}}_\beta (r)g)=f_s(\gamma \mathbf{{x}}_\beta (-r)v\mathbf{{x}}_\beta (r)g). \end{aligned}$$

On the other hand we have \(\omega _\psi (v\mathbf{{x}}_\beta (r)g)\phi (0)=\omega _\psi (\mathbf{{x}}_\beta (-r)v\mathbf{{x}}_\beta (r)g)\phi (0) \). Thus after a changing variable on v, we can get that the contribution of \(\omega _\psi (vg)\phi (0)\) to \(I(\widetilde{\varphi },\phi ,f_s)\) has an inner integral

$$\begin{aligned} \int _{F{\setminus }{\mathbb {A}}}\widetilde{\varphi }(\mathbf{{x}}_\beta (r)g)dr, \end{aligned}$$

which is zero by the cuspidality of \(\widetilde{\varphi }\). Thus we get

$$\begin{aligned} I(\widetilde{\varphi },\phi ,f_s)=\int _{B_{{\mathrm {SL}}_2}(F){\setminus }{\mathrm {SL}}_2({\mathbb {A}})}\int _{U_{\alpha +\beta }(F){\setminus }V({\mathbb {A}})}\widetilde{\varphi }(g)\sum _{a\in F^\times }\omega _\psi (t(a)vg)\phi (1)f_s(\gamma vg)dvdg. \end{aligned}$$

Collapsing the summation with the integration, we then get

$$\begin{aligned}&\quad I(\widetilde{\varphi },\phi ,f_s)\\&=\int _{N_{{\mathrm {SL}}_2}(F){\setminus }{\mathrm {SL}}_2({\mathbb {A}})}\int _{U_{\alpha +\beta }(F){\setminus }V({\mathbb {A}})}\widetilde{\varphi }(g)\omega _\psi (vg)\phi (1)f_s(\gamma vg)dvdg\\&=\int _{N_{{\mathrm {SL}}_2}({\mathbb {A}}){\setminus }{\mathrm {SL}}_2({\mathbb {A}})}\int _{U_{\alpha +\beta }(F){\setminus }V({\mathbb {A}})}\int _{F{\setminus }{\mathbb {A}}}\widetilde{\varphi }(\mathbf{{x}}_\beta (r)g)\omega _\psi (v\mathbf{{x}}_\beta (r)g)\phi (1)f_s(\gamma v\mathbf{{x}}_\beta (r)g)drdvdg. \end{aligned}$$

Note that we have \(\omega _\psi (v\mathbf{{x}}_\beta (r)g)\phi (1)=\omega _\psi (\mathbf{{x}}_\beta (r)\mathbf{{x}}_\beta (-r)v\mathbf{{x}}_\beta (r)g)\phi (1)=\psi (r) \omega _\psi (\mathbf{{x}}_\beta (-r)v\mathbf{{x}}_\beta (r)g)\phi (1)\). On the other hand, we have \(\gamma \mathbf{{x}}_\beta (r)\gamma ^{-1}\subset U_{3\alpha +\beta }\subset V'\). Thus \(f_s(\gamma v\mathbf{{x}}_\beta (r)g)=f_s(\gamma \mathbf{{x}}_\beta (-r)v\mathbf{{x}}_\beta (r)g) .\) After a changing of variable on v, we get

$$\begin{aligned} I(\widetilde{\varphi },\phi ,f_s)=\int _{N_{{\mathrm {SL}}_2}({\mathbb {A}}){\setminus }{\mathrm {SL}}_2({\mathbb {A}})}\int _{U_{\alpha +\beta }(F){\setminus }V({\mathbb {A}})}W_{\widetilde{\varphi }}(g)\omega _\psi (vg)\phi (1)f_s(\gamma vg)dvdg, \end{aligned}$$

where

$$\begin{aligned} W_{\widetilde{\varphi }}(g)=\int _{F{\setminus }{\mathbb {A}}}\tilde{\varphi }(\mathbf{{x}}_\beta (r)g)\psi (r)dr. \end{aligned}$$

We can further decompose the above integral as

$$\begin{aligned}&I(\widetilde{\varphi },\phi ,f_s)=\int _{N_{{\mathrm {SL}}_2}({\mathbb {A}}){\setminus }{\mathrm {SL}}_2({\mathbb {A}})}\int _{U_{\alpha +\beta }({\mathbb {A}}){\setminus }V({\mathbb {A}})}\\&\quad \int _{F{\setminus }{\mathbb {A}}}W_{\widetilde{\varphi }}(g)\omega _\psi (\mathbf{{x}}_{\alpha +\beta }(r)vg)\phi (1)f_s(\gamma \mathbf{{x}}_{\alpha +\beta }(r)vg)dr dvdg. \end{aligned}$$

Note that \(\omega _\psi (\mathbf{{x}}_{\alpha +\beta }(r)vg)\phi (1)=\psi (-2r)\omega _\psi (vg)\phi (1) \) and \( f_s(\gamma \mathbf{{x}}_{\alpha +\beta }(r)vg)=f_s(\mathbf{{x}}_\alpha (r)\gamma vg)\) since \(\gamma \mathbf{{x}}_{\alpha +\beta }(r)\gamma ^{-1}=\mathbf{{x}}_\alpha (r)\). We then get

$$\begin{aligned} I(\widetilde{\varphi },\phi ,f_s)=\int _{N_{{\mathrm {SL}}_2}({\mathbb {A}}){\setminus }{\mathrm {SL}}_2({\mathbb {A}})}\int _{U_{\alpha +\beta }({\mathbb {A}}){\setminus }V({\mathbb {A}})} W_{\widetilde{\varphi }}(g)\omega _\psi (vg)\phi (1)W_{f_s}(\gamma vg)dvdg, \end{aligned}$$

where

$$\begin{aligned} W_{f_s}(\gamma vg)=\int _{F{\setminus }{\mathbb {A}}}f_s(\mathbf{{x}}_\alpha (r)\gamma vg)\psi (-2r)dr. \end{aligned}$$

This concludes the proof.\(\square \)

4 Unramified calculation

In this section, let F be a p-adic field with \(p\ne 2\). Let \({\mathfrak {o}}\) be the ring of integers of F, and let p be a uniformizer of \({\mathfrak {o}}\) by abuse of notation. Let q be the cardinality of the residue field \({\mathfrak {o}}/(p)\).

4.1 Local Weil representations

Let \(\psi \) be an additive character of F and let \(\gamma (\psi )\) be the Weil index and let \(\mu _\psi (a)=\frac{\gamma (\psi )}{\gamma (\psi _a)}\). Let \(\omega _\psi \) be the Weil representation of \(\widetilde{\mathrm {SL}}_2(F) < imes V\) on \({\mathcal {S}}(F)\) via the projection \(\widetilde{\mathrm {SL}}_2(F) < imes V\rightarrow \widetilde{\mathrm {SL}}_2(F) < imes {\mathscr {H}}\). For \(\phi \in {\mathcal {S}}(F)\), we have the well-know formulas:

$$\begin{aligned} (\omega _{\psi }(w^1)\phi )(x)&=\gamma (\psi )\hat{\phi }(x),\\ (\omega _\psi (n(b))\phi )(x)&=\psi (bx^2)\phi (x), b\in F\\ (\omega _\psi (t(a))\phi )(x)&=|a|^{1/2} \mu _\psi (a)\phi (ax), a\in F^\times \\ (\omega _{\psi }((r_1,r_2,r_3))\phi )(x)&=\psi \left( r_3-2xr_2-r_1r_2\right) \phi (x+r_1), (r_1,r_2,r_3)\in {\mathscr {H}}(F). \end{aligned}$$

where \(\hat{\phi }(x)=\int _F \phi (y)\psi (2xy)dy\) is the Fourier transform of \(\phi \) with respect to \(\psi \). Note that under the embedding \({\mathrm {SL}}_2(F)\hookrightarrow {{G}}_2(F)\), we have \(w^1=w_\beta ,n(b)=\mathbf{{x}}_\beta (b)\) and \(t(a)=h(a,a^{-1})\).

4.2 Unramified calculation

In this subsection, we compute the local integral in last section. The strategy is similar to the unramified calculation in [6].

Let \(\widetilde{\pi }\) be an unramified genuine representation of \(\widetilde{\mathrm {SL}}_2(F)\) with Satake parameter a, and let \(\tau \) be an unramified irreducible representation of \({\mathrm {GL}}_2(F)\) with Satake parameters \(b_1,b_2\). Let \(\widetilde{W}\in {\mathcal {W}}(\tilde{\pi }, \psi )\) with \(\widetilde{W}(1)=1\). Let \(v_0\in V_\tau \) be an unramified vector and \(\lambda \in {\mathrm {Hom}}_{N}(V_\tau ,\psi )\) such that \(\lambda (v_0)=1\). Let \(f_s:G_2\rightarrow V_\tau \) be the unramified section in \(I(s,\tau )\) with \(f_s(e)=v_0\). Let

$$\begin{aligned} W_{f_s}:G_2\times {\mathrm {GL}}_2(F)\rightarrow {\mathbb {C}}\end{aligned}$$

be the function \(W_{f_s}(g,a)=\lambda (\tau (a)f_s(g)).\) We will write \(W_{f_s}(g)\) for \(W_{f_s}(g,1)\) in the following. By assumption and Shintani formula, we have

$$\begin{aligned} \begin{aligned} W_{f_s}(h(p^k,p^l))&=q^{-3s(2k+l)}\lambda (\tau ({\mathrm {diag}}(p^{k+l},p^k))v_0)\\&=q^{-3s(2k+l)}W_{v_0}({\mathrm {diag}}(p^{k+l},p^k))\\&=\left\{ \begin{array}{lll}q^{-3s(2k+l)}\frac{(b_1b_2)^kq^{-l/2}}{b_1-b_2}(b_1^{l+1}-b_2^{l+1}), &{} \text { if } l\ge 0,\\ 0, &{} \text { if } l<0. \end{array} \right. \end{aligned} \end{aligned}$$
(4.1)

Let \(\phi \in {\mathcal {S}}(F)\) be the characteristic function of \({\mathfrak {o}}\). We need to compute the integral

$$\begin{aligned} I(\widetilde{W}, W_{f_s},\phi )=\int _{N_2{\setminus } {\mathrm {SL}}_2(F)}\int _{U_{\alpha +\beta }{\setminus }V}\widetilde{W}(g)\omega _\psi (vg)\phi (1)W_{f_s}(\gamma vg)dvdg. \end{aligned}$$

In the following, we fix the Haar measure such that \({\mathrm {vol}}(dr,{\mathfrak {o}})=1\). Thus \({\mathrm {vol}}(d^*r,{\mathfrak {o}}^\times )=1-q^{-1}\).

Using the Iwasawa decomposition \({\mathrm {SL}}_2(F)=N_2(F)A_2(F){\mathrm {SL}}_2({\mathfrak {o}})\), we have

$$\begin{aligned}&I(\widetilde{W}, W_{f_s},\phi )\\ =&\int _{F^\times }\int _{F^4}\widetilde{W}(t(a))\omega _\psi ([r_1,0,r_3]t(a))\phi (1)W_{f_s}\\&(\gamma (r_1,0,r_3,r_4,r_5)t(a))|a|^{-2}dr_1dr_3dr_4dr_5 d^\times a\\ =&\int _{F^\times }\int _{F^4}\widetilde{W}(t(a))\omega _\psi (t(a)[r_1,0,r_3])\phi (1)W_{f_s}\\&(\gamma t(a) (r_1,0,r_3,r_4,r_5))|a|^{-3}dr_1dr_3dr_4dr_5 d^\times a \end{aligned}$$

If \(\widetilde{W}(t(a))\ne 0\), then \(|a|\le 1\). On the other hand, we have

$$\begin{aligned} \omega _\psi (t(a)[r_1,0,r_3])\phi (1)=\mu _\psi (a)|a|^{1/2}\psi (r_3)\phi (a+r_1). \end{aligned}$$

If \(\phi (a+r_1)\ne 0\) and \(a\in {\mathfrak {o}}\), then \(r_1\in {\mathfrak {o}}\). Thus the domain for a and \(r_1\) in the above integral is \(\left\{ {a\in F^\times \cap {\mathfrak {o}},r_1\in {\mathfrak {o}}}\right\} \). Note that \(\gamma t(a)=h(1,a)\gamma =h(1,a)w_\beta w_\alpha w_\beta w_\alpha \). Thus, if we conjugate \(w_\alpha \mathbf{{x}}_\alpha (r_1)\) to the right side, we can get

$$\begin{aligned} h(1,a)\gamma [r_1,0,r_3,r_4,r_5]=h(1,a)w_\beta w_\alpha w_\beta \mathbf{{x}}_{\alpha +\beta }(-r_3)\mathbf{{x}}_\beta (-r_4-3r_1r_3)\mathbf{{x}}_{3\alpha +2\beta }(r_5)w_\alpha \mathbf{{x}}_\alpha (r_1). \end{aligned}$$

Since \(w_\alpha \mathbf{{x}}_\alpha (r_1)\in K\) for \(r_1\in {\mathfrak {o}}\), by changing of variables, we get

$$\begin{aligned}&I(\widetilde{W}, W_{f_s},\phi )\\ =&\int _{|a|\le 1}\widetilde{W}(t(a))|a|^{-5/2}\mu _\psi (a) \\&\quad \cdot \int _{F^3} W_{f_s}(h(1,a)w_\beta w_\alpha w_\beta \mathbf{{x}}_{\alpha +\beta }(r_3)\mathbf{{x}}_\beta (r_4)\mathbf{{x}}_{3\alpha +2\beta }(r_5))\psi (-r_3)dr_3dr_4dr_5 d^*a\\ =&\sum _{n\ge 0} \widetilde{W}(t(p^n)) q^{5n/2}\mu _\psi (p^n) J(n), \end{aligned}$$

where

$$\begin{aligned} J(n)= \int _{F^3} W_{f_s}(h(1,p^n)w_\beta w_\alpha w_\beta \mathbf{{x}}_{\alpha +\beta }(r_3)\mathbf{{x}}_\beta (r_4)\mathbf{{x}}_{3\alpha +2\beta }(r_5))\psi (-r_3)dr_3dr_4dr_5. \end{aligned}$$

By dividing the domain of \(r_3\) into two parts, we can write \(J(n)=J_1(n)+J_2(n)\), where

$$\begin{aligned} J_1(n)&=\int _{|r_3|\le 1}\int _{F^2} W_{f_s}(h(1,p^n)w_\beta w_\alpha w_\beta \mathbf{{x}}_{\alpha +\beta }(r_3)\mathbf{{x}}_\beta (r_4)\mathbf{{x}}_{3\alpha +2\beta }(r_5))\psi (-r_3)dr_3dr_4dr_5\\&=\int _{F^2} W_{f_s}(h(1,p^n)w_\beta w_\alpha w_\beta \mathbf{{x}}_\beta (r_4)\mathbf{{x}}_{3\alpha +2\beta }(r_5))dr_4dr_5, \end{aligned}$$

and

$$\begin{aligned} J_2(n)=\int _{|r_3|> 1}\int _{F^2} W_{f_s}(h(1,p^n)w_\beta w_\alpha w_\beta \mathbf{{x}}_{\alpha +\beta }(r_3)\mathbf{{x}}_\beta (r_4)\mathbf{{x}}_{3\alpha +2\beta }(r_5))\psi (-r_3)dr_3dr_4dr_5. \end{aligned}$$

Lemma 4.1

Set

$$\begin{aligned} I(n)=\int _F W_{f_s}(h(1,p^n)w_\beta \mathbf{{x}}_\beta (r))dr. \end{aligned}$$

Then

$$\begin{aligned} I(n)&=\frac{q^{-(3s+1/2)n}}{b_1-b_2}\left[ (b_1^{n+1}-b_2^{n+1}) \right. \\&\quad +(1-q^{-1})\frac{b_1b_2X}{(1-b_1X) (1-b_2X)},\\&\quad \left. (b_1^n-b_2^n-b_1^{n+1}X+b_2^{n+1}X+b_1X(b_1b_2X)^n-b_2X(b_1b_2X)^n)\right] , \end{aligned}$$

where \(X=q^{-(3s-3/2)}\).

Proof

We have

$$\begin{aligned} I(n)&=\int _F W_{f_s}(h(1,p^n)w_\beta \mathbf{{x}}_\beta (r))dr\\&=\int _{|r|\le 1}W_{f_s}(h(1,p^n)w_\beta \mathbf{{x}}_\beta (r))dr\\&\quad +\int _{|r|> 1}W_{f_s}(h(1,p^n)w_\beta \mathbf{{x}}_\beta (r))dr\\&=W_{f_s}(h(1,p^n))+\int _{|r|> 1}W_{f_s}(h(1,p^n)w_\beta \mathbf{{x}}_\beta (r))dr. \end{aligned}$$

To deal with the integral when \(|r|>1\), we consider the following Iwasawa decomposition of \(w_\beta \mathbf{{x}}_\beta (r)\):

$$\begin{aligned} w_\beta \mathbf{{x}}_\beta (r)=\mathbf{{x}}_\beta (-r^{-1})h(-r^{-1},-r)\mathbf{{x}}_{-\beta }(r^{-1}). \end{aligned}$$

Since \(\mathbf{{x}}_{-\beta }(r^{-1})\) is in the maximal compact subgroup for \(|r|>1\), we have

$$\begin{aligned} W_{f_s}(h(1,p^n)w_\beta \mathbf{{x}}_\beta (r))=W_{f_s}(h(1,p^n)\mathbf{{x}}_{\beta }(-r^{-1})h(-r^{-1},-r))=W_{f_s}(h(1,p^n)h(r^{-1},r)), \end{aligned}$$

where we used \(U_\beta \subset V'\). For \(|r|>1\), we can write \(r= p^{-m}u\) for some \(m\ge 1\) and \(u\in {\mathfrak {o}}^\times \). We then have \(dr=q^{m}du\). Note that \({\mathrm {vol}}({\mathfrak {o}}^\times )=1-q^{-1}\). Thus we have

$$\begin{aligned} I(n)=W_{f_s}(h(1,p^n))+\sum _{m\ge 1}(1-q^{-1})q^mW_{f_s}(h(p^m,p^{n-m})). \end{aligned}$$

Note that \(h(p^m,1)\mapsto {\mathrm {diag}}(p^m,p^m)\) under the isomorphism \(M'\cong {\mathrm {GL}}_2\). Thus we have

$$\begin{aligned} W_{f_s}(h(p^m,1)h(1,p^{n-m}))=q^{-6sm}\omega _\tau (p)^mW_{f_s}(h(1,p^{n-m})). \end{aligned}$$

Thus we get

$$\begin{aligned} I(n)=W_{f_s}(h(1,p^n))+\sum _{m\ge 1}(1-q^{-1})q^{(-6s+1)m}\omega _\tau (p)^mW_{f_s}(h(1,p^{n-m})). \end{aligned}$$

By (4.1), we have

$$\begin{aligned} W_{f_s}(h(1,p^{n-m}))&=\left\{ \begin{array}{lll} \frac{q^{-3s(n-m)-(n-m)/2}}{b_1-b_2}(b_1^{n-m+1}-b_2^{n-m+1}), &{} \text { if } n\ge m,\\ 0, &{} \text { if } n<m. \end{array} \right. \end{aligned}$$

Thus for \(n\ge 1\), we have

$$\begin{aligned} I(n)&=\frac{q^{-(3s+1/2)n}}{b_1-b_2}\left( (b_1^{n+1}-b_2^{n+1})+\sum _{m= 1}^n(1-q^{-1})q^{-(3s-3/2)m}(b_1^{n+1}b_2^m-b_2^{n+1}b_1^m)\right) . \end{aligned}$$

Thus result can be computed using the geometric summation formula. One can check that the given formula also satisfies \(I(0)=1\). \(\square \)

Lemma 4.2

We have

$$\begin{aligned} J_1(n)=\frac{1-q^{-6s+1}b_1b_2}{1-q^{-6s+2}b_1b_2}I(n). \end{aligned}$$

Proof

To compute \(J_1(n)\), we break up the domain of integration in \(r_4\) and get

$$\begin{aligned} J_1(n)&=\int _{F}\int _{|r_4|\le 1} W_{f_s}(h(1,p^n)w_\beta w_\alpha w_\beta \mathbf{{x}}_\beta (r_4)\mathbf{{x}}_{3\alpha +2\beta }(r_5))dr_4dr_5\\&\quad +\int _{F}\int _{|r_4|> 1} W_{f_s}(h(1,p^n)w_\beta w_\alpha w_\beta \mathbf{{x}}_\beta (r_4)\mathbf{{x}}_{3\alpha +2\beta }(r_5))dr_4dr_5\\&:=J_{11}(n)+J_{12}(n), \end{aligned}$$

where

$$\begin{aligned} J_{11}(n)=&\int _{F}\int _{|r_4|\le 1}W_{f_s}(h(1,p^n)w_\beta w_\alpha w_\beta \mathbf{{x}}_\beta (r_4)\mathbf{{x}}_{3\alpha +2\beta }(r_5))dr_4 dr_5 \\ =&\int _{F}\int _{|r_4|\le 1}W_{f_s}(h(1,p^n) w_\beta w_\alpha w_\beta \mathbf{{x}}_{3\alpha +2\beta }(r_5)w_\beta ^{-1}w_\alpha ^{-1} w_\alpha w_\beta \mathbf{{x}}_\beta (r_4))dr_4dr_5\\ =&\int _{F}W_{f_s}(h(1,p^n)w_\beta \mathbf{{x}}_\beta (r_5))dr_5\\ =&I(n), \end{aligned}$$

and

$$\begin{aligned} J_{12}(n)&=\int _{F}\int _{|r_4|> 1}W_{f_s}(h(1,p^n)w_\beta w_\alpha w_\beta \mathbf{{x}}_\beta (r_4)\mathbf{{x}}_{3\alpha +2\beta }(r_5))dr_4 dr_5 \\&=\int _{F}\int _{|r_4|> 1}W_{f_s}(h(1,p^n) w_\beta w_\alpha w_\beta \mathbf{{x}}_{3\alpha +2\beta }(r_5)w_\beta ^{-1}w_\alpha ^{-1} w_\alpha w_\beta \mathbf{{x}}_\beta (r_4))dr_4dr_5\\&=\int _{F}\int _{|r_4|> 1}W_{f_s}(h(1,p^n) w_\beta \mathbf{{x}}_{\beta }(r_5) w_\alpha w_\beta \mathbf{{x}}_\beta (r_4))dr_4dr_5. \end{aligned}$$

We have the Iwasawa decomposition of \(w_\beta \mathbf{{x}}_\beta (r_4)\):

$$\begin{aligned} w_\beta \mathbf{{x}}_\beta (r_4)=\mathbf{{x}}_\beta (-r_4^{-1})h(-r_4^{-1},-r_4)\mathbf{{x}}_{-\beta }(r_4^{-1}). \end{aligned}$$

Since \(\mathbf{{x}}_{-\beta }(r_4^{-1})\) is in the maximal compact subgroup for \(|r_4|>1\), we then get

$$\begin{aligned} J_{12}(n)&=\int _{F}\int _{|r_4|> 1}W_{f_s}(h(1,p^n) w_\beta \mathbf{{x}}_{\beta }(r_5) w_\alpha \mathbf{{x}}_\beta (-r_4^{-1})h(r_4^{-1},r_4))dr_4dr_5\\&=\int _F \int _{|r_4|>1}W_{f_s}(h(1,p^n) h(r_4^{-1},1)w_\beta \mathbf{{x}}_{\beta }(r_4^{-1}r_5))dr_4dr_5\\&=\int _F \int _{|r_4|>1}|r_4|W_{f_s}(h(1,p^n) h(r_4^{-1},1)w_\beta \mathbf{{x}}_{\beta }(r_5))dr_4dr_5\\&=\sum _{m\ge 1}(1-q^{-1})q^{2m} \int _F W_{f_s}(h(p^m,1)h(1,p^n)w_\beta \mathbf{{x}}_\beta (r_5))dr_5, \end{aligned}$$

where in the second equality, we conjugated \(\mathbf{{x}}_\beta (-r_4^{-1})h(r_4^{-1},r_4)\) to the left, and in the third equality, we wrote \(r_4=p^{-m}u\) for \(m\ge 1, u\in {\mathfrak {o}}^\times \) and used \(dr_4=q^mdu, {\mathrm {vol}}({\mathfrak {o}}^\times )=1-q^{-1}\). Note that \(h(p^m,1)\) is in the center of \(M'\), and thus

$$\begin{aligned} W_{f_s}(h(p^m,1)g)=q^{-6sm}\omega _\tau (p)^mW_{f_s}(g), \end{aligned}$$

we get

$$\begin{aligned} J_{12}(n)=(1-q^{-1})\sum _{m\ge 1}q^{-6sm+2m}\omega _\tau (p)^m\int _F W_{f_s}(h(1,p^n)w_\beta \mathbf{{x}}_\beta (r_5))dr_5. \end{aligned}$$

Thus we get

$$\begin{aligned} J_1(n)=I(n)+\sum _{m\ge 1}(1-q^{-1})q^{(-6s+2)m}(b_1b_2)^m I(n). \end{aligned}$$

A simple calculation gives the formula of \(J_1(n)\). \(\square \)

We next consider the term

$$\begin{aligned} J_2(n)=\int _{|r_3|> 1}\int _{F^2} W_{f_s}(h(1,p^n)w_\beta w_\alpha w_\beta \mathbf{{x}}_{\alpha +\beta }(r_3)\mathbf{{x}}_\beta (r_4)\mathbf{{x}}_{3\alpha +2\beta }(r_5))\psi (-r_3)dr_3dr_4dr_5. \end{aligned}$$

For \(|r_3|>1\), we can write \(r_3\in p^{-m}u\) with \(m\ge 1, u\in {\mathfrak {o}}^\times \). We then have,

$$\begin{aligned}&J_2(n)\\= & {} \int _{F^2}\sum _{m\ge 1}q^m W_{f_s}(h(1,p^n)w_\beta w_\alpha w_\beta \mathbf{{x}}_{\alpha +\beta }(p^{-m}u)\mathbf{{x}}_\beta (r_4)\mathbf{{x}}_{3\alpha +2\beta }(r_5))\psi (-p^{-m}u)dudr_4dr_5. \end{aligned}$$

Write \(\mathbf{{x}}_{\alpha +\beta }(p^{-m}u)=h(u,u^{-1})\mathbf{{x}}_{\alpha +\beta }(p^{-m})h(u^{-1},u)\), and by conjugation and changing of variables, we get

$$\begin{aligned}&J_2(n)\\= & {} \int _{F^2}\sum _{m\ge 1}q^m W_{f_s}(h(u^{-1},p^n)w_\beta w_\alpha w_\beta \mathbf{{x}}_{\alpha +\beta }(p^{-m})\mathbf{{x}}_\beta (r_4)\mathbf{{x}}_{3\alpha +2\beta }(r_5))\psi (-p^{-m}u)dudr_4dr_5, \end{aligned}$$

where we used \(h(u,u^{-1})\) is in the maximal compact subgroup of \({ {G}}_2(F)\). Since \(h(u^{-1},1)\) maps to the center of \(M'\) and \( |\omega _\tau (u)|=1 \), we have

$$\begin{aligned}&W_{f_s}(h(u^{-1},p^n)w_\beta w_\alpha w_\beta \mathbf{{x}}_{\alpha +\beta }(p^{-m})\mathbf{{x}}_\beta (r_4)\mathbf{{x}}_{3\alpha +2\beta }(r_5))\\= & {} W_{f_s}(1,p^n)w_\beta w_\alpha w_\beta \mathbf{{x}}_{\alpha +\beta }(p^{-m})\mathbf{{x}}_\beta (r_4)\mathbf{{x}}_{3\alpha +2\beta }(r_5)). \end{aligned}$$

Thus we get

$$\begin{aligned}&J_2(n)\\= & {} \int _{F^2}\sum _{m\ge 1}q^m W_{f_s}(h(1,p^n)w_\beta w_\alpha w_\beta \mathbf{{x}}_{\alpha +\beta }(p^{-m})\mathbf{{x}}_\beta (r_4)\mathbf{{x}}_{3\alpha +2\beta }(r_5))\psi (-p^{-m}u)dudr_4dr_5. \end{aligned}$$

Since

$$\begin{aligned} \int _{{\mathfrak {o}}^\times }\psi (p^ku)du=\left\{ \begin{array}{lll}1-q^{-1}, &{} \text { if } k\ge 0,\\ -q^{-1}, &{} \text { if } k=-1, \\ 0, &{} \text { if } k\le -2, \end{array}\right. \end{aligned}$$

we get \(J_2(n)=-R(n)\), where

$$\begin{aligned} R(n)=\int _{F^2} W_{f_s}(h(1,p^n)w_\beta w_\alpha w_\beta \mathbf{{x}}_{\alpha +\beta }(p^{-1})\mathbf{{x}}_\beta (r_4)\mathbf{{x}}_{3\alpha +2\beta }(r_5))dr_4dr_5. \end{aligned}$$

To evaluate R(n), we split the domain of \(r_4\), and write \(R(n)=R_1(n)+R_{2}(n)\), where

$$\begin{aligned} R_1(n)&=\int _{|r_4|\le 1} \int _{F} W_{f_s}(h(1,p^n)w_\beta w_\alpha w_\beta \mathbf{{x}}_{\alpha +\beta }(p^{-1})\mathbf{{x}}_\beta (r_4)\mathbf{{x}}_{3\alpha +2\beta }(r_5))dr_4dr_5,\\&= \int _{F} W_{f_s}(h(1,p^n)w_\beta w_\alpha w_\beta \mathbf{{x}}_{\alpha +\beta }(p^{-1})\mathbf{{x}}_{3\alpha +2\beta }(r_5))dr_5, \end{aligned}$$

and

$$\begin{aligned} R_2(n)=\int _{|r_4|>1} \int _{F} W_{f_s}(h(1,p^n)w_\beta w_\alpha w_\beta \mathbf{{x}}_{\alpha +\beta }(p^{-1})\mathbf{{x}}_\beta (r_4)\mathbf{{x}}_{3\alpha +2\beta }(r_5))dr_4dr_5. \end{aligned}$$

We now compute \(R_1(n)\). We conjugate \(w_\alpha w_\beta \mathbf{{x}}_{\alpha +\beta }(p^{-1})\) to the right and then get

$$\begin{aligned} R_1(n)&=\int _{F} W_{f_s}(h(1,p^n)w_\beta \mathbf{{x}}_{\beta }(r_5)w_\alpha w_\beta \mathbf{{x}}_{\alpha +\beta }(p^{-1}))dr_5\\&=\int _{F} W_{f_s}(h(1,p^n)w_\beta \mathbf{{x}}_{\beta }(r_5)w_\alpha \mathbf{{x}}_{\alpha }(-p^{-1}))dr_5 \end{aligned}$$

Next, we use the Iwasawa decomposition of \(w_\alpha \mathbf{{x}}_\alpha (p^{-1})\):

$$\begin{aligned} w_\alpha \mathbf{{x}}_\alpha (-p^{-1})=\mathbf{{x}}_\alpha (p)h(p^{-1},p^2)\mathbf{{x}}_{-\alpha }(-p) \end{aligned}$$

to get

$$\begin{aligned} R_1(n)=\int _F W_{f_s}(h(1,p^n)w_\beta \mathbf{{x}}_\beta (r_5)\mathbf{{x}}_\alpha (p)h(p^{-1},p^2))dr_5. \end{aligned}$$

Next, we use the commutator relation

$$\begin{aligned} \mathbf{{x}}_\beta (r_5)\mathbf{{x}}_{\alpha }(p)=\mathbf{{x}}_{\alpha +\beta }(pr_5)u\mathbf{{x}}_\alpha (p)\mathbf{{x}}_\beta (r_5), \end{aligned}$$

where u is in the root space of \(2\alpha +\beta ,3\alpha +\beta ,3\alpha +2\beta \). Then we get

$$\begin{aligned} R_1(n)=\int _F W_{f_s}(h(1,p^n)w_\beta \mathbf{{x}}_{\alpha +\beta }(pr_5)u\mathbf{{x}}_\alpha (p)\mathbf{{x}}_\beta (r_5)h(p^{-1},p^2))dr_5. \end{aligned}$$

Note that \(w_\beta u \mathbf{{x}}_\alpha (r) w_\beta (1)\in V'\), and \(h(1,p^n)w_\beta \mathbf{{x}}_{\alpha +\beta }(pr_5)(h(1,p^n)w_\beta )^{-1}=\mathbf{{x}}_\alpha (-p^{n+1}r_5),\) and \(W_{f_s}(\mathbf{{x}}_\alpha (r)g)=\psi (2r)W_{f_s}(g)\), we get

$$\begin{aligned} R_1(n)&=\int _F W_{f_s}(h(1,p^n)w_\beta \mathbf{{x}}_\beta (r_5)h(p^{-1},p^2))\psi (-2p^{n+1}r_5)dr_5\\&=\int _F W_{f_s}(h(p^2,1)h(1,p^{n-1})w_\beta \mathbf{{x}}_\beta (p^3r_5))\psi (-2p^{n+1}r_5)dr_5\\&=q^{-12s+3}\omega _\tau (p^2)\int _{F}W_{f_s}(h(1,p^{n-1})w_\beta \mathbf{{x}}_\beta (r_5))\psi (-2 p^{n-2}r_5)dr_5, \end{aligned}$$

where the last equality comes from a changing of variable on \(r_5\) and the fact that \(h(p^2,1)\mapsto {\mathrm {diag}}(p^2,p^2)\) under the isomorphism \(M'\cong {\mathrm {GL}}_2\). We next break up the integral on \(r_5\) and get

$$\begin{aligned} R_1(n)&=q^{-12s+3}\omega _\tau (p^2)W_{f_s}(h(1,p^{n-1}))\int _{|r_5|\le 1}\psi (-2p^{n-2}r_5)dr_5\\&\quad +q^{-12s+3}\omega _\tau (p^2)\int _{|r_5|>1}W_{f_s}(h(1,p^{n-1})w_\beta \mathbf{{x}}_\beta (r_5))\psi (-2p^{n-2}r_5)dr_5. \end{aligned}$$

Using the Iwasawa decomposition of \(w_\beta \mathbf{{x}}_\beta (r_5)\), we have

$$\begin{aligned} R_1(n)&=q^{-12s+3}\omega _\tau (p^2)\\&\quad \left( W_{f_s}(h(1,p^{n-1}))\int _{|r_5|\le 1}\psi (-2p^{n-2}r_5)dr_5 \right. \\&\quad \left. +\sum _{m=1}^\infty W_{f_s}(h(p^{m},p^{n-m-1}))q^m\int _{{\mathfrak {o}}^\times }\psi (-2p^{n-m-2}u)du \right) . \end{aligned}$$

Lemma 4.3

We have \(R_1(n)=0\) if \(n\le 1\), and

$$\begin{aligned} R_1(n)=q^{-12s+3}\omega _\tau (p)^2I(n-1)-q^{-6s(n+1)+n+2}\omega _\tau (p)^{n+1}, \end{aligned}$$

for \(n\ge 2\).

Proof

Note that \(\int _{|r|\le 1}\psi (p^kr)dr=0\) if \(k<0\) and \(\int _{|r|\le 1}\psi (p^kr)dr=1 \) if \(k\ge 0\). Moreover, we have

$$\begin{aligned} \int _{{\mathfrak {o}}^\times }\psi (p^ku)du=\left\{ \begin{array}{lll}1-q^{-1}, &{} \text { if } k\ge 0,\\ -q^{-1}, &{} \text { if } k=-1, \\ 0, &{} \text { if } k\le -2. \end{array}\right. \end{aligned}$$

Thus we get \(R_1(n)=0\) for \(n\le 1\). For \(n\ge 2\), we have

$$\begin{aligned} R_1(n)&= q^{-12s+3}\omega _\tau (p^2)\\&\quad \cdot \left( W_{f_s}(h(1,p^{n-1}))+\sum _{m=1}^{n-2}(1-q^{-1})q^mW_{f_s}( h(p^{m},p^{n-m-1}))\right. \\&\quad \left. -q^{-1}q^{n-1}W_{f_s}(h(p^{(n-1)},1))\right) =q^{-12s+3}\omega _\tau (p^2)\\&\quad \cdot \left( W_{f_s}(h(1,p^{n-1}))+\sum _{m=1}^{n-1}(1-q^{-1})q^mW_{f_s}( h(p^{m},p^{n-m-1}))\right. \\&\quad \left. -q^{n-1}W_{f_s}(h(p^{(n-1)},1)) \right) =q^{-12s+3}\\&\quad \omega _\tau (p)^2I(n-1)-q^{-12s+3+n-1}\omega _\tau (p)^2W_{f_s}(h(p^{n-1},1)), \end{aligned}$$

where in the last equation, we used the formula in the computation of I(n). Since \(h(p^{n-1},1)\) is in the center of \(M'\), we have \(W_{f_s}(h(p^{n-1},1))=q^{-6s(n-1)}\omega _\tau (p)^{n-1}\). The result follows. \(\square \)

We next consider

$$\begin{aligned} R_2(n)&=\int _{|r_4|>1} \int _{F} W_{f_s}(h(1,p^n)w_\beta w_\alpha w_\beta \mathbf{{x}}_{\alpha +\beta }(p^{-1})\mathbf{{x}}_\beta (r_4)\mathbf{{x}}_{3\alpha +2\beta }(r_5))dr_4dr_5. \end{aligned}$$

Conjugating \(w_\beta \) to the right side and using the Iwasawa decomposition of \(w_\beta \mathbf{{x}}_\beta (r_4)\), we can get

$$\begin{aligned} R_2(n)=\int _F \int _{|r_4|>1}W_{f_s}(h(1,p^n)w_\beta w_\alpha \mathbf{{x}}_\alpha (p^{-1})\mathbf{{x}}_{3\alpha +\beta }(r_5)\mathbf{{x}}_\beta (r_4^{-1})h(r_4^{-1},r_4))dr_4dr_5. \end{aligned}$$

From the commutator relation, we have

$$\begin{aligned} \mathbf{{x}}_\alpha (p^{-1})\mathbf{{x}}_\beta (r_4^{-1})=\mathbf{{x}}_\beta (r_4^{-1})\mathbf{{x}}_\alpha (p^{-1})\mathbf{{x}}_{2\alpha +\beta }(p^{-2}r_4^{-1})u, \end{aligned}$$

for some u in the group generated by roots subgroups of \(\alpha +\beta ,3\alpha +\beta ,3\alpha +2\beta \). Like in the computation of \(R_1(n)\), we have

$$\begin{aligned} R_2(n)&=\int _F\int _{|r_4|>1}W_{f_s}(h(1,p^n)w_\beta w_\alpha \mathbf{{x}}_\alpha (p^{-1})\mathbf{{x}}_{3\alpha +\beta }(r_5)h(r_4^{-1},r_4))\psi \\&\quad (-2p^{n-2}r_4^{-1})dr_4dr_5\\&=\int _F \int _{|r_4|>1}W_{f_s}(h(1,p^n)h(r_4^{-1},1)w_\beta \mathbf{{x}}_\beta (r_5r_4^{-1})w_\alpha \mathbf{{x}}_\alpha (p^{-1}r_4^{-1}))\psi (-2p^{n-2}r_4^{-1})\\&\quad dr_4dr_5\\&=\int _F \int _{|r_4|>1}|r_4|W_{f_s}(h(1,p^n)h(r_4^{-1},1)w_\beta \mathbf{{x}}_\beta (r))\psi (-2p^{n-2}r_4^{-1})dr_4dr\\&=I(n)\int _{|r_4|>1}|r_4|^{-6s+1}\omega _\tau (r_4^{-1})\psi (-2p^{n-2}r_4^{-1})dr_4\\&=I(n)\sum _{m= 1}^\infty q^{(-6s+2)m}\omega _\tau (p)^{m}\int _{{\mathfrak {o}}^\times }\psi (-2p^{m+n-2}u)du. \end{aligned}$$

Lemma 4.4

We have

$$\begin{aligned} R_2(n)=\left\{ \begin{array}{lll}I(0)q^{-6s+2}\omega _\tau (p)\left( -q^{-1}+(1-q^{-1})\frac{q^{-6s+2}\omega _\tau (p)}{1-q^{-6s+2}\omega _\tau (p)} \right) ,&{} n=0, \\ I(n)(1-q^{-1})\frac{q^{-6s+2}\omega _\tau (p)}{1-q^{-6s+2}\omega _\tau (p)},&{} n\ge 1 \end{array}\right. \end{aligned}$$

Proof

If \(n\ge 1\), then \(\int _{{\mathfrak {o}}^\times }\psi (p^{m+n-2}u)du=(1-q^{-1}) \) for \(m\ge 1\). Thus, we have

$$\begin{aligned} R_2(n)&=I(n)\sum _{m= 1}^\infty q^{(-6s+2)m}\omega _\tau (p)^{m}(1-q^{-1})\\&=I(n)(1-q^{-1})\frac{q^{-6s+2}\omega _\tau (p)}{1-q^{-6s+2}\omega _\tau (p)}. \end{aligned}$$

If \(n=0\), then \(\int _{{\mathfrak {o}}^\times }\psi (p^{m+n-2}u)du=(1-q^{-1}) \) for \(m\ge 2\), and \(\int _{{\mathfrak {o}}^\times }\psi (p^{m+n-2}u)du=-q^{-1}\) for \(m=1\). Thus, we have

$$\begin{aligned} R_2(0)&=I(0)(-q^{-1}q^{-6s+2}\omega _\tau (p)+(1-q^{-1})\sum _{m=2}^\infty q^{(-6s+2)m}\omega _\tau (p)^{m})\\&=I(0)q^{-6s+2}\omega _\tau (p)\left( -q^{-1}+(1-q^{-1})\frac{q^{-6s+2}\omega _\tau (p)}{1-q^{-6s+2}\omega _\tau (p)} \right) . \end{aligned}$$

The completes the proof of the lemma. \(\square \)

Combining the above results, we get the following

Lemma 4.5

We have

$$\begin{aligned} R(n)=\left\{ \begin{array}{lll}-I(0)q^{-6s+1}\omega _\tau (p)\frac{1-q^{-6s+3}\omega _\tau (p)}{1-q^{-6s+2}\omega _\tau (p)}, &{} n=0,\\ I(1)(1-q^{-1})\frac{q^{-6s+2}\omega _\tau (p)}{1-q^{-6s+2}\omega _\tau (p)}, &{} n=1,\\ q^{-12s+3}\omega _\tau (p)^2I(n-1)-q^{-6s(n+1)+n+2}\omega _\tau (p)^{n+1} &{}\\ \qquad \qquad \qquad +I(n)(1-q^{-1})\frac{q^{-6s+2}\omega _\tau (p)}{1-q^{-6s+2}\omega _\tau (p)}, &{} n\ge 2, \end{array}\right. \end{aligned}$$

and

$$\begin{aligned} J(n)&=J_1(n)-R(n)\\&=\left\{ \begin{array}{lll}1+Y, &{} n=0\\ I(1), &{} n=1,\\ I(n)-q^{-1}Y^2 I(n-1)+q^{-n}Y^{n+1}, &{} n\ge 2. \end{array} \right. \end{aligned}$$

where \(Y=q^{-6s+2}\omega _\tau (p)\)

By the main result of [1], we have

$$\begin{aligned} \widetilde{W}(t(p^n))=\frac{\mu _\psi (p^{n})q^{-n}}{a-a^{-1}}\left( (1-\chi (p)q^{-1/2}a^{-1})a^{n+1}-(1-\chi (p)q^{-1/2}a)a^{-(n+1)}\right) , \end{aligned}$$

where \(\chi (p)=(p,p)_F=(p,-1)_F.\) Note that the notation \(\gamma (a)\) in [1] is our \(\mu _\psi (a)^{-1}\). Note that \(\mu _\psi (p^n)\mu _\psi (p^n)=(p^n,p^n)_F=\chi (p)^n\). Thus

$$\begin{aligned}&I(\widetilde{W},W_{f_s},\phi )=\sum _{n\ge 0}\frac{q^{3n/2}\chi (p)^n}{a-a^{-1}}\\&\quad \left( (1-\chi (p)q^{-1/2}a^{-1})a^{n+1}-(1-\chi (p)q^{-1/2}a)a^{-(n+1)}\right) J(n). \end{aligned}$$

Plugging the formula J(n) into the above equation, we can get that

$$\begin{aligned}&I(\widetilde{W},W_f,\phi )\\&\quad =\frac{(1-b_1q^{-1}X)(1-b_2q^{-1}X)(1-b_1b_2q^{-1}X^2)(1-b_1^2b_2q^{-1}X^3)(1-b_1b_2^2q^{-1}X^3)}{(1-\chi (p)a^{-1}b_1b_2q^{-1/2}X^2)(1-\chi (p)ab_1b_2q^{-1/2}X^2)}\\&\quad \cdot \frac{1}{\prod _{i=1}^2(1-\chi (p)a^{-1}b_iq^{-1/2}X)\prod _{i=1}^2(1-\chi (p)ab_iq^{-1/2}X)}\\&\quad =\frac{L(3s-1,\widetilde{\pi }\times (\chi \otimes \tau ))L(6s-5/2,\widetilde{\pi }\otimes (\chi \otimes \omega _\tau ))}{L(3s-1/2,\tau )L(6s-2,\omega _\tau )L(9s-7/2,\tau \otimes \omega _\tau )}. \end{aligned}$$

Here

$$\begin{aligned} \begin{aligned} L(s,\widetilde{\pi }\otimes (\chi \otimes \omega _{\tau } ))=\frac{1}{(1-a\chi (p)b_1b_2q^{-s})((1-a^{-1}\chi (p)b_1b_2q^{-s}))} \end{aligned} \end{aligned}$$

is the L function of \(\widetilde{\pi }\) twisted by the character \(\chi \otimes \omega _\tau \), and

$$\begin{aligned} L(s,\widetilde{\pi }\times (\chi \otimes \tau ))=\frac{1}{\prod _{i=1}^2(1-\chi (p)a^{-1}b_iq^{-s})\prod _{i=1}^2(1-\chi (p)ab_iq^{-s})} \end{aligned}$$

is the Rankin–Selberg L-function of \(\widetilde{\pi }\) twisted by \(\chi \otimes \tau \). We record the above calculation in the following

Proposition 4.6

Let \(\widetilde{W}\in {\mathcal {W}}(\widetilde{\pi },\psi )\) be the normalized unramified Whittaker function, \(f_s\) be the normalized unramified section in \(I(s,\tau )\) and \(\phi \in {\mathcal {S}}(F)\) is the characteristic function of \({\mathfrak {o}}\), we have

$$\begin{aligned} I(\widetilde{W},W_{f_s},\phi )=\frac{L(3s-1,\widetilde{\pi }\times (\chi \otimes \tau ))L(6s-5/2,\widetilde{\pi }\otimes (\chi \otimes \omega _\tau ))}{L(3s-1/2,\tau )L(6s-2,\omega _\tau )L(9s-7/2,\tau \otimes \omega _\tau )}. \end{aligned}$$

5 Some local theory

In this section, let F be a local field, which can be archimedean or non-archimedean. If F is non-archimedean, let \({\mathfrak {o}}\) be the ring of integers of F, p be a uniformizer of \({\mathfrak {o}}\) and \(q={\mathfrak {o}}/(p)\). Let \(\widetilde{\pi }\) be an irreducible genuine generic representation of \(\widetilde{\mathrm {SL}}_2(F)\), \(\tau \) be an irreducible generic representation of \({\mathrm {GL}}_2(F)\). Let \(\psi \) be a nontrivial additive character of F.

Lemma 5.1

Let \(\widetilde{W}\in {\mathcal {W}}(\widetilde{\pi },\psi ), f_s\in I(s,\tau ),\phi \in {\mathcal {S}}(F)\), then the integral \(I(\widetilde{W},W_{f_s},\phi )\) converges absolutely for \({\mathrm {Re}}(s)\) large and has a meromorphic continuation to the whole s-plane. Moreover, if F is a p-adic field, then \(I(\widetilde{W},W_{f_s},\phi )\) is a rational function in \(q^{-s}\).

The proof is similar to [5, Lemma 4.2–4.7] and [6, Lemma 3.10, Lemma 3.3]. We omit the details.

Lemma 5.2

Let \(s_0\in {\mathbb {C}}\). Then there exists \(\widetilde{W}\in {\mathcal {W}}(\widetilde{\pi },\psi ),f_{s_0}\in I(s_0,\tau ),\phi \in {\mathcal {S}}(F)\) such that \(I(\widetilde{W},W_{f_{s_0}},\phi )\ne 0\).

Proof

The proof is similar to the proof of [5, Lemma 4.4,4.7], [6, Proposition 3.4]. We omit the details. \(\square \)

6 Nonvanishing of certain periods on \(G_2\)

6.1 Poles of Eisenstein series on \({ {G}}_2\)

Let \(\tau \) be a cuspidal unitary representation of \({\mathrm {GL}}_2({\mathbb {A}})\cong M'({\mathbb {A}})\). Let K be a maximal compact subgroup of \({ {G}}_2({\mathbb {A}})\). Given a \(K\cap {\mathrm {GL}}_2({\mathbb {A}})\)-finite cusp form f in \(\tau \), we can extend f to a function \(\widetilde{f}:{ {G}}_2({\mathbb {A}})\rightarrow {\mathbb {C}}\) as in [13, §2]. We then define

$$\begin{aligned} \Phi _{\widetilde{f},s}(g)=\widetilde{f}(g) \delta _{P'}(m')^{s/3+1/2}, \end{aligned}$$

for \(g=v'm'k\) with \(v'\in V'({\mathbb {A}}),m'\in M'({\mathbb {A}}),k\in K\). Then \(\Phi _{\widetilde{f},s}\) is well-defined and \(\Phi _{\widetilde{f},s}\in I(\frac{s}{3}+\frac{1}{2},\tau )\). Then we can consider the Eisenstein series

$$\begin{aligned} E(s,\widetilde{f},g)=\sum _{P'(F)\backslash { {G}}_2(F)}\Phi _{\widetilde{f},s}(\gamma g). \end{aligned}$$

Proposition 6.1

The Eisenstein series \(E(s,\widetilde{f}, g)\) has a pole on the half plane \({\mathrm {Re}}(s)>0\) if and only if \(s=\frac{1}{2}, \omega _\tau =1\) and \(L(\frac{1}{2},\tau )\ne 0\).

For a proof of the above proposition, see [16, §1] or [10, §5]. If \(\omega _\tau =1\) and \(L(\frac{1}{2},\tau )\ne 0\), denote by \({\mathcal {R}}(\frac{1}{2},\tau )\) the space generated by the residues of Eisenstein series \(E(s,\widetilde{f},g)\) defined as above. Note that an element \( R\in {\mathcal {R}}(\frac{1}{2},\tau )\) is an automorphic form on \(G_2({\mathbb {A}})\).

6.2 On the Shimura–Waldspurger lift

Let \(\widetilde{\pi }\) be a genuine cuspidal automorphic representation of \(\widetilde{{\mathrm {SL}}}_2({\mathbb {A}})\). Let \(Wd_\psi (\widetilde{\pi })\) be the Shimura–Waldspurger lift of \(\widetilde{\pi }\). Then \(Wd_\psi (\widetilde{\pi })\) is a cuspidal representation of \({\mathrm {P}}{\mathrm {GL}}_2({\mathbb {A}})\). A cuspidal automorphic representation \( \tau \) is in the image of \(Wd_\psi \) if and only if \(L(\frac{1}{2},\tau )\ne 0\). Moreover, the correspondence \(\widetilde{\pi }\mapsto Wd_\psi (\widetilde{\pi })\) respects the Rankin-Selberg L-functions. For these assertions, see [15] or [2].

6.3 A period on \(G_2\)

Theorem 6.2

Let \(\widetilde{\pi }\) be a genuine cuspidal automorphic representation of \(\widetilde{\mathrm {SL}}_2({\mathbb {A}})\) and \(\tau \) be a unitary cuspidal automorphic representation of \({\mathrm {GL}}_2({\mathbb {A}})\). Assume that \(\omega _\tau =1\) and \(L(\frac{1}{2},\tau )\ne 0\). In particular, \(\tau \) can be viewed as a cuspidal automorphic representation of \({\mathrm {P}}{\mathrm {GL}}_2({\mathbb {A}})\). If \(Wd_\psi (\widetilde{\pi })=\chi \otimes \tau ,\) then there exists \(\widetilde{\varphi }\in V_{\widetilde{\pi }}, \phi \in {\mathcal {S}}({\mathbb {A}}), R\in {\mathcal {S}}(\frac{1}{2},\tau )\) such that the period

$$\begin{aligned} \begin{aligned} {\mathcal {P}}(\widetilde{\varphi }, \widetilde{\theta }_\phi ,R)=\int _{{\mathrm {SL}}_2(F)\backslash {\mathrm {SL}}_2({\mathbb {A}})}\int _{V(F)\backslash V({\mathbb {A}})}\widetilde{\varphi }(g)\widetilde{\theta }_\phi (vg)R(vg)dvdg \end{aligned} \end{aligned}$$

is non-vanishing.

Proof

For \(\widetilde{\varphi }\in V_\pi , \phi \in {\mathcal {S}}({\mathbb {A}})\) and a good section \(\Phi _{\widetilde{f}, s}\) as in Sect. 6.1, by Theorem 3.1 and Proposition 4.6, we have

$$\begin{aligned} I(\widetilde{\varphi },\phi ,\widetilde{f},s)&=\int _{{\mathrm {SL}}_2(F)\backslash {\mathrm {SL}}_2({\mathbb {A}})}\int _{V(F)\backslash V({\mathbb {A}})} \widetilde{\varphi }(g)\widetilde{\theta }_{\phi }(vg)E(vg,\Phi _{\widetilde{f},s})dvdg\\&=\int _{N_{{\mathrm {SL}}_2}({\mathbb {A}})\backslash {\mathrm {SL}}_2({\mathbb {A}})}\int _{U_{\alpha +\beta }({\mathbb {A}})\backslash V({\mathbb {A}})}W_{\widetilde{\varphi }}(g)\omega _\psi (vg)\phi (1)W_{\Phi _{\widetilde{f},s}}(\gamma vg)dvdg \nonumber \\&=I_S \cdot \frac{L^S(s+\frac{1}{2},\widetilde{\pi }\times (\chi \otimes \tau ))L^S(2s+\frac{1}{2},\widetilde{\pi }\otimes (\chi \otimes \omega _\tau ))}{L^S(s+1,\tau )L^S(2s+1,\omega _\tau )L^S(3s+1,\tau \otimes \omega _\tau )}. \nonumber \end{aligned}$$

Here S is a finite set of places of F such that for \(v\notin S\), \( \pi _v,\tau _v\) are unramified, and \( I_S\) is the product of the local zeta integrals over all places \(v\in S\) and \(L^S\) denotes the partial L-function which is the product of all local L-function as the place v runs over \(v\notin S\). Note that \(\tau \cong \tau ^\vee \) since \(\omega _\tau =1\). Suppose that \(Wd_\psi (\widetilde{\pi })=\chi \otimes \tau =\chi \otimes \tau ^\vee \), then \(L^S(s+1/2, \widetilde{\pi }\times (\chi \otimes \tau )) \) has a pole at \(s=1/2\). Note that at \(s=\frac{1}{2}\), \(L^{S}(2s+1/2,\widetilde{\pi }\otimes (\chi \otimes \omega _\tau ))\) is holomorphic and nonzero, while \(L^S(s+1,\tau )L^S(2s+1,\omega _\tau )L^S(3s+1,\tau \otimes \omega _\tau )\) is holomorphic. Moreover, \(I_S\) can be chosen to be nonzero. Thus we get that \(I(\widetilde{\varphi },\phi , \widetilde{f},s)\) has a pole at \(s=1/2\), which means that there exists a residue R(g) of \(E(s,\widetilde{f},g)\) such that

$$\begin{aligned} \begin{aligned} {\mathcal {P}}(\widetilde{\varphi },\theta _\phi ,R)=\int _{{\mathrm {SL}}_2(F)\backslash {\mathrm {SL}}_2({\mathbb {A}})}\int _{V(F)\backslash V({\mathbb {A}})}\widetilde{\varphi }(g)\widetilde{\theta }_\phi (vg)R(vg)dvdg\ne 0. \end{aligned} \end{aligned}$$

This completes the proof. \(\square \)

Remark 6.3

For an \(L^2\)-automorphic form \(\eta \in L^2(G_2(F)\backslash G_2({\mathbb {A}}))\), one can form the period

$$\begin{aligned} \eta _{\widetilde{\phi },\widetilde{\theta }_\phi }(g)=\int _{{\mathrm {SL}}_2(F)\backslash {\mathrm {SL}}_2({\mathbb {A}})}\int _{V(F)\backslash V({\mathbb {A}})}\widetilde{\varphi }(h)\widetilde{\theta }_\phi (vh)\eta (vhg)dvdh. \end{aligned}$$

Theorem 6.2 says that if \(\eta \in {\mathcal {S}}(\frac{1}{2},\tau )\), then under the condition \(Wd_\psi (\widetilde{\pi })=\chi \otimes \tau \), the period \(\eta _{\widetilde{\varphi },\widetilde{\theta }_\phi } \) is non-vanishing for certain \(\widetilde{\varphi }\) and \(\phi \). For general \(\eta \), one can ask under what conditions \(\eta _{\widetilde{\varphi },\widetilde{\theta }_\phi }\) is not identically zero as \(\widetilde{\varphi }\) varies in \(\widetilde{\pi }\) and \(\phi \in {\mathcal {S}}({\mathbb {A}})\). In the classical group case, this is the global Gan–Gross–Prasad conjecture for Fourier–Jacobi case, see [3]. It is natural to ask if it is possible to extend the GGP-conjecture to the \(G_2\)-case.