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

On a Generalization of Motohashi’s Formula: Non-archimedean Weight Functions

Han Wu School of Mathematical Sciences, University of Scinece and Technology of China, 230026 Hefei, P. R. China wuhan1121@yahoo.com
Abstract.

This is a continuation of the adelic version of Kwan’s formula. At non-archimedean places we give a bound of the weight function on the mixed moment side, when the weight function on the PGL3×PGL2\displaystyle{\rm PGL}_{3}\times{\rm PGL}_{2} side is nearly the characteristic function of a short family. Our method works for any tempered representation Π\displaystyle\Pi of PGL3\displaystyle{\rm PGL}_{3}, and reveals the structural reason for the appearance of Katz’s hypergeometric sums in a previous joint work with P.Xi.

1. Introduction

1.1. Main Results

This paper is the first follow-up of our previous [19]. Let 𝐅\displaystyle\mathbf{F} be a number field with adele ring 𝔸\displaystyle\mathbb{A}. Let ψ\displaystyle\psi be the additive character of 𝐅\𝔸\displaystyle\mathbf{F}\backslash\mathbb{A} à la Tate. Let S𝐅\displaystyle S_{\mathbf{F}} be the set of places of 𝐅\displaystyle\mathbf{F}, and S\displaystyle S_{\infty} be the subset of archimedean places. Fix a cuspidal automorphic representation Π\displaystyle\Pi of PGL3(𝔸)\displaystyle{\rm PGL}_{3}(\mathbb{A}). We established an adelic version of Kwan’s spectral reciprocity identity (see [19, Theorem 1.1 & (5.19)-(5.21)]), which we recall in the special case of trivial central characters as follows.

Theorem 1.1.

Let SSS𝐅\displaystyle S_{\infty}\subset S\subset S_{\mathbf{F}} be any finite subset. At every place vS\displaystyle v\in S there is a pair of weight functions hv\displaystyle h_{v} and h~v\displaystyle\widetilde{h}_{v} with the auxiliary normalized ones at finite places 𝔭<\displaystyle\mathfrak{p}<\infty

(1.1) H𝔭(π𝔭):=h𝔭(π𝔭)L(1,π𝔭×π𝔭)L(1/2,Π𝔭×π𝔭),H~𝔭(χ𝔭):=h~𝔭(χ𝔭)L(1/2,Π𝔭×χ𝔭1)L(1/2,χ𝔭),H_{\mathfrak{p}}(\pi_{\mathfrak{p}}):=h_{\mathfrak{p}}(\pi_{\mathfrak{p}})\frac{L(1,\pi_{\mathfrak{p}}\times\pi_{\mathfrak{p}})}{L(1/2,\Pi_{\mathfrak{p}}\times\pi_{\mathfrak{p}})},\quad\widetilde{H}_{\mathfrak{p}}(\chi_{\mathfrak{p}}):=\frac{\widetilde{h}_{\mathfrak{p}}(\chi_{\mathfrak{p}})}{L(1/2,\Pi_{\mathfrak{p}}\times\chi_{\mathfrak{p}}^{-1})L(1/2,\chi_{\mathfrak{p}})},

so that the following equation holds (where π\displaystyle\pi runs through cuspidal automorphic forms of PGL2(𝔸)\displaystyle{\rm PGL}_{2}(\mathbb{A}))

πL(1/2,Π×π)2Λ𝐅(2)L(1,π,Ad)vhv(πv)𝔭SH𝔭(π𝔭)+χ+𝐅×\𝔸×^|L(1/2+iτ,Π×χ)|22Λ𝐅(2)|L(1+2iτ,χ2)|2vhv(π(χv,iτ))𝔭SH𝔭(π(χ𝔭,iτ))dτ2π=1ζ𝐅χ+𝐅×\𝔸×^L(1/2iτ,Π×χ1)L(1/2+iτ,χ)vh~v(χv||viτ)𝔭SH~𝔭(χ𝔭||𝔭iτ)dτ2π+1ζ𝐅±Ress1=±12L(1/2s1,Π)ζ𝐅(1/2+s1)vh~v(||vs1)𝔭SH~𝔭(||𝔭s1),\sum_{\pi}\frac{L(1/2,\Pi\times\pi)}{2\Lambda_{\mathbf{F}}(2)L(1,\pi,\mathrm{Ad})}\cdot\prod_{v\mid\infty}h_{v}(\pi_{v})\cdot\prod_{\mathfrak{p}\in S}H_{\mathfrak{p}}(\pi_{\mathfrak{p}})+\\ \sum_{\chi\in\widehat{\mathbb{R}_{+}\mathbf{F}^{\times}\backslash\mathbb{A}^{\times}}}\int_{-\infty}^{\infty}\frac{\left\lvert L(1/2+i\tau,\Pi\times\chi)\right\rvert^{2}}{2\Lambda_{\mathbf{F}}(2)\left\lvert L(1+2i\tau,\chi^{2})\right\rvert^{2}}\cdot\prod_{v\mid\infty}h_{v}(\pi(\chi_{v},i\tau))\cdot\prod_{\mathfrak{p}\in S}H_{\mathfrak{p}}(\pi(\chi_{\mathfrak{p}},i\tau))\frac{\mathrm{d}\tau}{2\pi}\\ =\frac{1}{\zeta_{\mathbf{F}}^{*}}\sum_{\chi\in\widehat{\mathbb{R}_{+}\mathbf{F}^{\times}\backslash\mathbb{A}^{\times}}}\int_{-\infty}^{\infty}L(1/2-i\tau,\Pi\times\chi^{-1})L(1/2+i\tau,\chi)\cdot\prod_{v\mid\infty}\widetilde{h}_{v}(\chi_{v}\lvert\cdot\rvert_{v}^{i\tau})\cdot\prod_{\mathfrak{p}\in S}\widetilde{H}_{\mathfrak{p}}(\chi_{\mathfrak{p}}\lvert\cdot\rvert_{\mathfrak{p}}^{i\tau})\frac{\mathrm{d}\tau}{2\pi}+\\ \frac{1}{\zeta_{\mathbf{F}}^{*}}\sum_{\pm}{\rm Res}_{s_{1}=\pm\frac{1}{2}}L(1/2-s_{1},\Pi)\zeta_{\mathbf{F}}(1/2+s_{1})\cdot\prod_{v\mid\infty}\widetilde{h}_{v}(\lvert\cdot\rvert_{v}^{s_{1}})\cdot\prod_{\mathfrak{p}\in S}\widetilde{H}_{\mathfrak{p}}(\lvert\cdot\rvert_{\mathfrak{p}}^{s_{1}}),

where we have used the abbreviation π(χv,s):=π(χv||vs,χv1||vs)\displaystyle\pi(\chi_{v},s):=\pi(\chi_{v}\lvert\cdot\rvert_{v}^{s},\chi_{v}^{-1}\lvert\cdot\rvert_{v}^{-s}).

We only mention that the weight functions hv\displaystyle h_{v} and h~v\displaystyle\widetilde{h}_{v} are integrals of a smooth Whittaker function Wv𝒲(Πv,ψv)\displaystyle W_{v}\in\mathcal{W}(\Pi_{v}^{\infty},\psi_{v}). More details from [19] will be recalled when we need them.

In this paper we focus on the weight functions at a non-archimedean place 𝔭<\displaystyle\mathfrak{p}<\infty. We omit the subscript 𝔭\displaystyle\mathfrak{p} for simplicity. Let 𝐅\displaystyle\mathbf{F} be a non-archimedean local field of characteristic 0\displaystyle 0 with valuation ring 𝒪𝐅\displaystyle\mathcal{O}_{\mathbf{F}}, and write q=Nr(𝔭)\displaystyle q={\rm Nr}(\mathfrak{p}) from now on. In the case χ=||𝐅s\displaystyle\chi=\lvert\cdot\rvert_{\mathbf{F}}^{s}, the function H~(||𝐅s)\displaystyle\widetilde{H}(\lvert\cdot\rvert_{\mathbf{F}}^{s}) is a polynomial in [qs,qs]\displaystyle\mathbb{C}[q^{s},q^{-s}], hence is entire. We introduce its Taylor expansion at any point s0\displaystyle s_{0} as

(1.2) H~(||𝐅s)=k0H~(k;s0)(ss0)k.\widetilde{H}(\lvert\cdot\rvert_{\mathbf{F}}^{s})=\sideset{}{{}_{k\geqslant 0}}{\sum}\widetilde{H}(k;s_{0})\left(s-s_{0}\right)^{k}.

We summarize our main results (namely Proposition 5.13, 5.16, 6.10 & 6.12) as follows.

Theorem 1.2.

Suppose the residual characteristic of 𝐅\displaystyle\mathbf{F} is not 2\displaystyle 2. Let ΠPGL3(𝐅)^\displaystyle\Pi\in\widehat{{\rm PGL}_{3}(\mathbf{F})} be generic and tempered. Let π0PGL2(𝐅)^\displaystyle\pi_{0}\in\widehat{{\rm PGL}_{2}(\mathbf{F})} with 𝔠(π0)>1\displaystyle\mathfrak{c}(\pi_{0})>1. We can choose W𝒲(Π,ψ)\displaystyle W\in\mathcal{W}(\Pi^{\infty},\psi) so that:

  • (1)

    The weight function satisfies h(π)0\displaystyle h(\pi)\geqslant 0 for any πPGL2(𝐅)^\displaystyle\pi\in\widehat{{\rm PGL}_{2}(\mathbf{F})} and h(π0)>0\displaystyle h(\pi_{0})>0;

  • (2)

    For any unitary χ𝐅×^\displaystyle\chi\in\widehat{\mathbf{F}^{\times}} and ϵ>0\displaystyle\epsilon>0 the dual weight function satisfies

    h(π0)1h~(χ)ϵ𝐂(Π)2+ϵ{𝟙max(𝔠(π0)2,6𝔠(Π))(𝔠(χ))+q12𝟙(1)q12=ε0𝟙2𝔠(χ)=𝔠(π0)23𝟙𝓔(π0)(χ2)},\displaystyle h(\pi_{0})^{-1}\widetilde{h}(\chi)\ll_{\epsilon}\mathbf{C}(\Pi)^{2+\epsilon}\cdot\left\{\mathbbm{1}_{\leqslant\max(\lfloor\frac{\mathfrak{c}(\pi_{0})}{2}\rfloor,6\mathfrak{c}(\Pi))}(\mathfrak{c}(\chi))+q^{\frac{1}{2}}\mathbbm{1}_{(-1)^{\frac{q-1}{2}}=\varepsilon_{0}}\mathbbm{1}_{2\nmid\mathfrak{c}(\chi)=\frac{\mathfrak{c}(\pi_{0})}{2}\geqslant 3}\mathbbm{1}_{\boldsymbol{\mathcal{E}}(\pi_{0})}(\chi^{2})\right\},

    where ε0=1\displaystyle\varepsilon_{0}=-1 if π0\displaystyle\pi_{0} is dihedral supercuspidal, and ε0=1\displaystyle\varepsilon_{0}=1 otherwise; and the exceptional set of characters 𝓔(π0)\displaystyle\boldsymbol{\mathcal{E}}(\pi_{0}) of 𝒪𝐅×\displaystyle\mathcal{O}_{\mathbf{F}}^{\times} has size O(q1𝐂(π0)12)\displaystyle O(q^{-1}\mathbf{C}(\pi_{0})^{\frac{1}{2}}) and is given below in Lemma 5.7 (2);

  • (3)

    For any k0\displaystyle k\in\mathbb{Z}_{\geqslant 0} and ϵ>0\displaystyle\epsilon>0 the normalized unramified dual weight function satisfies the bounds

    h(π0)1H~(k;±1/2)ϵ𝐂(Π)4+ϵq𝔠(π0)2+ϵ.\displaystyle h(\pi_{0})^{-1}\widetilde{H}(k;\pm 1/2)\ll_{\epsilon}\mathbf{C}(\Pi)^{4+\epsilon}q^{\lceil\frac{\mathfrak{c}(\pi_{0})}{2}\rceil+\epsilon}.
Remark 1.3.

Our primary goal is to reveal the structural reason for the bounds of the dual weight functions. Our main discovery is the quadratic elementary functions given in (1.6), which are “building blocks” of the Bessel functions of the relevant representations. See §1.3 for more details. Further extension of our method to the non-dihedral supercuspidal and Steinberg π0\displaystyle\pi_{0} requires only plugging in the integral representation of the Bessel functions of such π0\displaystyle\pi_{0}, analogous to [19, Theorem 1.6].

Remark 1.4.

Specializing to 𝐅=p\displaystyle\mathbf{F}=\mathbb{Q}_{p} and Π=𝟙𝟙𝟙\displaystyle\Pi=\mathbbm{1}\boxplus\mathbbm{1}\boxplus\mathbbm{1}, Theorem 1.2 corresponds to the main local non-archimedean computation of the recent work of Hu–Petrow–Young [8] in the case p2\displaystyle p\neq 2.

1.2. Notation and Convention

For a locally compact group G\displaystyle G, let G^\displaystyle\widehat{G} be the topological dual of unitary irreducible representations. For πG^\displaystyle\pi\in\widehat{G}, we write Vπ\displaystyle V_{\pi} for the underlying Hilbert space, and write VπVπ\displaystyle V_{\pi}^{\infty}\subset V_{\pi} for the subspace of smooth vectors if G\displaystyle G carries extra structure to make sense of the notion.

Throughout the paper 𝐅\displaystyle\mathbf{F} is a local field of characteristic 0\displaystyle 0, with residual characteristic 2\displaystyle\neq 2. Let ||𝐅\displaystyle\lvert\cdot\rvert_{\mathbf{F}} (resp. v𝐅\displaystyle v_{\mathbf{F}}) be the valuation (resp. normalized additive valuation) of 𝐅\displaystyle\mathbf{F}. Fix ψ\displaystyle\psi an additive character of conductor exponent 0\displaystyle 0 and normalize the measures accordingly. The valuation ring of 𝐅\displaystyle\mathbf{F} is 𝒪𝐅\displaystyle\mathcal{O}_{\mathbf{F}}, while the valuation ideal is 𝒫𝐅\displaystyle\mathcal{P}_{\mathbf{F}}. We choose a uniformizer ϖ𝐅𝒫𝐅𝒫𝐅2\displaystyle\varpi_{\mathbf{F}}\in\mathcal{P}_{\mathbf{F}}-\mathcal{P}_{\mathbf{F}}^{2}. Different choices of ϖ𝐅\displaystyle\varpi_{\mathbf{F}} give different ramified quadratic extensions of 𝐅\displaystyle\mathbf{F}. Write 𝐆d:=GLd\displaystyle\mathbf{G}_{d}:={\rm GL}_{d} for simplicity, introduce some compact open subgroups of 𝐆2(𝐅)\displaystyle\mathbf{G}_{2}(\mathbf{F}) as

𝐊:=GL2(𝒪𝐅);𝐊0[𝒫𝐅n]:={(abcd)𝐊|c𝒫𝐅n},n0;\displaystyle\mathbf{K}:={\rm GL}_{2}(\mathcal{O}_{\mathbf{F}});\quad\mathbf{K}_{0}[\mathcal{P}_{\mathbf{F}}^{n}]:=\left\{\begin{pmatrix}a&b\\ c&d\end{pmatrix}\in\mathbf{K}\ \middle|\ c\in\mathcal{P}_{\mathbf{F}}^{n}\right\},\ \forall n\in\mathbb{Z}_{\geqslant 0};

and some algebraic subgroups of 𝐆2(𝐅)\displaystyle\mathbf{G}_{2}(\mathbf{F}) as

𝐙=𝐙2(𝐅):={z𝟙2|z𝐅×},𝐍2(𝐅):={n(x):=(1x1)|x𝐅},\displaystyle\mathbf{Z}=\mathbf{Z}_{2}(\mathbf{F}):=\left\{z\mathbbm{1}_{2}\ \middle|\ z\in\mathbf{F}^{\times}\right\},\quad\mathbf{N}_{2}(\mathbf{F}):=\left\{n(x):=\begin{pmatrix}1&x\\ &1\end{pmatrix}\ \middle|\ x\in\mathbf{F}\right\},
𝐀2(𝐅):={(t1t2)|t1,t2𝐅×},𝐁2(𝐅):=𝐀2(𝐅)𝐍2(𝐅).\displaystyle\mathbf{A}_{2}(\mathbf{F}):=\left\{\begin{pmatrix}t_{1}&\\ &t_{2}\end{pmatrix}\ \middle|\ t_{1},t_{2}\in\mathbf{F}^{\times}\right\},\quad\mathbf{B}_{2}(\mathbf{F}):=\mathbf{A}_{2}(\mathbf{F})\mathbf{N}_{2}(\mathbf{F}).

For integers n,m1\displaystyle n,m\geqslant 1 we write 𝒮(n×m,𝐅)\displaystyle\mathcal{S}(n\times m,\mathbf{F}) for the space of Schwartz–Bruhat functions on M(n×m,𝐅)\displaystyle{\rm M}(n\times m,\mathbf{F}) the n×m\displaystyle n\times m matrices with entries in 𝐅\displaystyle\mathbf{F}. The (inverse) ψ\displaystyle\psi-Fourier transform is denoted and defined by

Ψ^(X)=𝔉ψ(Ψ)(X)=M(n×m,𝐅)Ψ(Y)ψ(Tr(XYT))dY.\displaystyle\widehat{\Psi}(X)=\mathfrak{F}_{\psi}(\Psi)(-X)=\int_{{\rm M}(n\times m,\mathbf{F})}\Psi(Y)\psi\left({\rm Tr}(XY^{T})\right)\mathrm{d}Y.

If no confusion occurs, we may omit ψ\displaystyle\psi from the notation. If this is the case, then the inverse Fourier transform is denoted by 𝔉¯=𝔉ψ¯=𝔉ψ1\displaystyle\overline{\mathfrak{F}}=\mathfrak{F}_{\overline{\psi}}=\mathfrak{F}_{\psi^{-1}}. We introduce some elementary operators on the space of functions on 𝐅×\displaystyle\mathbf{F}^{\times}:

  • For functions ϕ\displaystyle\phi on 𝐅×\displaystyle\mathbf{F}^{\times}, its extension by 0\displaystyle 0 to 𝐅\displaystyle\mathbf{F} is denoted by e(ϕ)\displaystyle\mathrm{e}(\phi), and its inverse is Inv(ϕ)(t):=ϕ(t1)\displaystyle\mathrm{Inv}(\phi)(t):=\phi(t^{-1}); for functions ϕ\displaystyle\phi on 𝐅\displaystyle\mathbf{F}, its restriction to 𝐅×\displaystyle\mathbf{F}^{\times} is denoted by r(ϕ)\displaystyle\mathrm{r}(\phi), and the operator i\displaystyle\mathrm{i} is

    i=eInvr.\displaystyle\mathrm{i}=\mathrm{e}\circ\mathrm{Inv}\circ\mathrm{r}.
  • For s\displaystyle s\in\mathbb{C}, μ𝐅×^\displaystyle\mu\in\widehat{\mathbf{F}^{\times}} and functions ϕ\displaystyle\phi on 𝐅\displaystyle\mathbf{F}, we introduce the operator 𝔪s(μ)\displaystyle\mathfrak{m}_{s}(\mu) by

    𝔪s(μ)(ϕ)(t)=ϕ(t)μ(t)|t|𝐅s.\displaystyle\mathfrak{m}_{s}(\mu)(\phi)(t)=\phi(t)\mu(t)\lvert t\rvert_{\mathbf{F}}^{s}.
  • For δ𝐅×\displaystyle\delta\in\mathbf{F}^{\times} we introduce the operator 𝔱(δ)\displaystyle\mathfrak{t}(\delta) by

    𝔱(δ)(ϕ)(y)=ϕ(yδ).\displaystyle\mathfrak{t}(\delta)(\phi)(y)=\phi(y\delta).
  • Let In=In,𝐅:C(𝐅×)C(GLn(𝐅))\displaystyle I_{n}=I_{n,\mathbf{F}}:{\rm C}(\mathbf{F}^{\times})\to{\rm C}({\rm GL}_{n}(\mathbf{F})) be given by In(h)(g):=h(detg)\displaystyle I_{n}(h)(g):=h(\det g).

These notation apply to finite (field) extensions of 𝐅\displaystyle\mathbf{F}.

We introduce the standard involution of inverse-transpose on GLn(R)\displaystyle{\rm GL}_{n}(R) as gι:=g1t\displaystyle g^{\iota}:={}^{t}g^{-1}.

If (U,du)\displaystyle(U,\mathrm{d}u) is a measured space with finite total mass, we introduce the normalized integral as

(1.3) Uf(u)du:=1Vol(U,du)Uf(u)du.\oint_{U}f(u)\mathrm{d}u:=\frac{1}{{\rm Vol}(U,\mathrm{d}u)}\int_{U}f(u)\mathrm{d}u.

For n1\displaystyle n\in\mathbb{Z}_{\geqslant 1} we will frequently perform the following process of regularization to an integral

(1.4) 𝐅f(y)dy=𝐅(𝒪𝐅f(y(1+ϖ𝐅nx))dx)dy,\int_{\mathbf{F}}f(y)\mathrm{d}y=\int_{\mathbf{F}}\left(\oint_{\mathcal{O}_{\mathbf{F}}}f(y(1+\varpi_{\mathbf{F}}^{n}x))\mathrm{d}x\right)\mathrm{d}y,

which we shall refer to as the level n\displaystyle n regularization (with respect) to dy\displaystyle\mathrm{d}y.

1.3. Outline of Proof

From the local main result [19, Theorem 1.4] the weight h(π)\displaystyle h(\pi) and the dual weight h~(χ)\displaystyle\widetilde{h}(\chi) functions are related to each other by a (hidden) relative orbital integral H(y)\displaystyle H(y) via

(1.5) h(π)=𝐅×H(y)jπ~,ψ1(y1)d×y|y|𝐅,h~(χ)=𝐅×ψ(y)χ1(y)|y|𝐅12𝒱~Π(H)(y)d×yh(\pi)=\int_{\mathbf{F}^{\times}}H(y)\cdot\mathrm{j}_{\widetilde{\pi},\psi^{-1}}\begin{pmatrix}&-y\\ 1&\end{pmatrix}\frac{\mathrm{d}^{\times}y}{\lvert y\rvert_{\mathbf{F}}},\quad\widetilde{h}(\chi)=\int_{\mathbf{F}^{\times}}\psi(-y)\chi^{-1}(y)\lvert y\rvert_{\mathbf{F}}^{-\frac{1}{2}}\cdot\widetilde{\mathcal{V}}_{\Pi}(H)(y)\mathrm{d}^{\times}y

where jπ~,ψ1\displaystyle\mathrm{j}_{\widetilde{\pi},\psi^{-1}} is the Bessel function of the contragredient representation π~\displaystyle\widetilde{\pi} in the sense of [2, §3.5], and 𝒱~Π\displaystyle\widetilde{\mathcal{V}}_{\Pi} is the extended Voronoi transform characterized by the equations

𝒱~Π:=𝒱~Π𝔪1,𝔪2(β)I3𝒱~Π=𝔉¯𝔪1(βι)I3\displaystyle\widetilde{\mathcal{VH}}_{\Pi}:=\widetilde{\mathcal{V}}_{\Pi}\circ\mathfrak{m}_{1},\quad\mathfrak{m}_{-2}(\beta)\circ I_{3}\circ\widetilde{\mathcal{VH}}_{\Pi}=\overline{\mathfrak{F}}\circ\mathfrak{m}_{-1}(\beta^{\iota})\circ I_{3}

for all smooth matrix coefficients β\displaystyle\beta of Π\displaystyle\Pi.

We need to choose H(y)\displaystyle H(y) so that h(π)0\displaystyle h(\pi)\geqslant 0 selects a short family containing π0\displaystyle\pi_{0}. In other words the weight function h(π)\displaystyle h(\pi) should be an approximation of the characteristic functions of {π0}\displaystyle\{\pi_{0}\}. From the first formula of (1.5) and the orthogonality of Bessel functions, one should expect that any such H(y)\displaystyle H(y) is an approximation of jπ0,ψ(y1)\displaystyle\mathrm{j}_{\pi_{0},\psi}\begin{pmatrix}&-y\\ 1&\end{pmatrix}. If π0\displaystyle\pi_{0} is supercuspidal, the asymptotic analysis of both functions (in §2.1 and Lemma 3.10) show that the two functions can be equal. In general, we construct H(y)\displaystyle H(y) from some test function ϕ0ιϕ0\displaystyle\phi_{0}^{\iota}*\phi_{0} on PGL2(𝐅)\displaystyle{\rm PGL}_{2}(\mathbf{F}) of positive type in order to ensure h(π)0\displaystyle h(\pi)\geqslant 0; we also require ϕ0\displaystyle\phi_{0} to be left covariant with respect to some character of an open compact subgroup, which determines the (minimal 𝐊\displaystyle\mathbf{K}-)type of π0\displaystyle\pi_{0} in the sense of [12]. We use ϕ0\displaystyle\phi_{0} to deduce some integral representation (3.5) of H(y)\displaystyle H(y) which approximates the one of jπ0,ψ(y1)\displaystyle\mathrm{j}_{\pi_{0},\psi}\begin{pmatrix}&-y\\ 1&\end{pmatrix} obtained in [19, Theorem 1.6]. The integral representation of H(y)\displaystyle H(y) in all cases is summarized in Corollary 3.12, and is the departure point of our refined analysis of the dual weight function h~(χ)\displaystyle\widetilde{h}(\chi) described below.

Our key observation on the dual weight function is the following decomposition into three parts. The first part is the contribution from H(y)\displaystyle H_{\infty}(y), essentially the restriction of H(y)\displaystyle H(y) to v𝐅(y)𝔠(π0)\displaystyle v_{\mathbf{F}}(y)\leqslant-\mathfrak{c}(\pi_{0}) (see (4.1)). In this region H(y)\displaystyle H_{\infty}(y) has a stable behavior regardless π0\displaystyle\pi_{0}, i.e., the germ of any Bessel relative orbital integral at infinity. To treat the corresponding h~(χ)\displaystyle\widetilde{h}_{\infty}(\chi) we crucially rely on the extension of the Voronoi–Hankel transform developed in our previous paper [19, Theorem 1.3]. The remaining part h~c(χ)\displaystyle\widetilde{h}_{c}(\chi) corresponding to Hc:=HH\displaystyle H_{c}:=H-H_{\infty} can be written as h~c(χ)=h~c+(χ)+h~c(χ)\displaystyle\widetilde{h}_{c}(\chi)=\widetilde{h}_{c}^{+}(\chi)+\widetilde{h}_{c}^{-}(\chi), where

h~c+(χ)=𝒪𝐅χ1(y)|y|𝐅12𝒱~Π(H)(y)d×y,h~c(χ)=𝐅𝒪𝐅ψ(y)χ1(y)|y|𝐅12𝒱~Π(H)(y)d×y.\displaystyle\widetilde{h}_{c}^{+}(\chi)=\int_{\mathcal{O}_{\mathbf{F}}}\chi^{-1}(y)\lvert y\rvert_{\mathbf{F}}^{-\frac{1}{2}}\cdot\widetilde{\mathcal{V}}_{\Pi}(H)(y)\mathrm{d}^{\times}y,\quad\widetilde{h}_{c}^{-}(\chi)=\int_{\mathbf{F}-\mathcal{O}_{\mathbf{F}}}\psi(-y)\chi^{-1}(y)\lvert y\rvert_{\mathbf{F}}^{-\frac{1}{2}}\cdot\widetilde{\mathcal{V}}_{\Pi}(H)(y)\mathrm{d}^{\times}y.

The bound of h~c+\displaystyle\widetilde{h}_{c}^{+} is offered by the local functional equations via Lemma 4.2. The method via the local functional equations can also offer a crude bound in Lemma 4.8 of h~c\displaystyle\widetilde{h}_{c}^{-} when 𝔠(π0)𝔠(Π)\displaystyle\mathfrak{c}(\pi_{0})\ll\mathfrak{c}(\Pi).

To refine the bound of h~c\displaystyle\widetilde{h}_{c}^{-} in the case 𝔠(π0)𝔠(Π)\displaystyle\mathfrak{c}(\pi_{0})\gg\mathfrak{c}(\Pi) we observe that Hc(y)\displaystyle H_{c}(y) is a linear combination of translations of the following quadratic elementary functions (see Definition 5.1 & 6.1, (5.1)-(5.4) & (6.1)-(6.3) for more details)

(1.6) Fn,Gn(y2)=𝟙v(y)=n±η𝐋/𝐅(±y)ψ(±y),F_{n},G_{n}(y^{2})=\mathbbm{1}_{v(y)=-n}\cdot\sideset{}{{}_{\pm}}{\sum}\eta_{\mathbf{L}/\mathbf{F}}(\pm y)\psi(\pm y),

where 𝐋/𝐅\displaystyle\mathbf{L}/\mathbf{F} is the quadratic algebra extension associated with the type of π0\displaystyle\pi_{0}, and η𝐋/𝐅\displaystyle\eta_{\mathbf{L}/\mathbf{F}} is the corresponding quadratic character of 𝐅×\displaystyle\mathbf{F}^{\times}. We essentially change the order of integrations by first computing the dual weight of (the translates of) the quadratic elementary functions. In particular, the translation pattern of the above quadratic elementary functions is responsible for the appearance of Katz’s hypergeometric sums in our previous joint work with Xi [21].

Remark 1.5.

Xi [22] has yet another transformation of the algebraic exponential sums in Petrow–Young’s work [14]. This is mysterious and seems to lie beyond the framework of GL2\displaystyle{\rm GL}_{2} or GL3\displaystyle{\rm GL}_{3}.

Remark 1.6.

In the case of principal series π0\displaystyle\pi_{0}, our ϕ0\displaystyle\phi_{0} coincides with Nelson’s test function in [13]. But we do not view it within the theory of micro-localized vectors, nor does our method of bounding the dual weight function rely on anything in that theory. Our choice of test function follows the idea of a previous work [1] in the real case by analogy in terms of “minimal 𝐊\displaystyle\mathbf{K}-type” [12].

1.4. Acknowledgement

We thank Zhi Qi and Ping Xi for discussions related to the topics of the paper.

2. Local Weight Transforms Revisited

We have expressed the local weight transforms in terms of the extended Voronoi transforms in [19, Theorem 1.4]. In that version, a test function H(y)\displaystyle H(y) is some integral of a Kirillov function of a generic unitary irreducible Π\displaystyle\Pi (namely the restriction of a W𝒲(Π,ψ)\displaystyle W\in\mathcal{W}(\Pi^{\infty},\psi) to the left-upper embedding of 𝐆2(𝐅)\displaystyle\mathbf{G}_{2}(\mathbf{F})). As explained in [19, (6.16)], the space of test functions H(y)\displaystyle H(y) contains the Bessel orbital integrals of Cc(𝐆2(𝐅))\displaystyle{\rm C}_{c}^{\infty}(\mathbf{G}_{2}(\mathbf{F})). We shall restrict to the latter subspace of test functions and get more refined information on the local weight transforms in the case of trivial central characters.

2.1. Relative Orbital Integrals

For any m0\displaystyle m\in\mathbb{Z}_{\geqslant 0} we introduce a function EmCc(𝐅×)\displaystyle E_{m}\in{\rm C}_{c}^{\infty}(\mathbf{F}^{\times}) supported in the subset of square elements of 𝐅×\displaystyle\mathbf{F}^{\times} given by

(2.1) Em(y2):=𝟙v𝐅(y)=m|y|𝐅±±1+𝒫𝐅m2ψ(y(u+u1))du.E_{m}(y^{2}):=\mathbbm{1}_{v_{\mathbf{F}}(y)=-m}\cdot\lvert y\rvert_{\mathbf{F}}\sum_{\pm}\int_{\pm 1+\mathcal{P}_{\mathbf{F}}^{\lfloor\frac{m}{2}\rfloor}}\psi\left(y(u+u^{-1})\right)\mathrm{d}u.

Recall that for any fCc(GL2(𝐅))\displaystyle f\in{\rm C}_{c}^{\infty}({\rm GL}_{2}(\mathbf{F})), the Bessel orbital integral [19, (5.17)] is given by

(2.2) h(y)=𝐅×𝐅2f((1x11)(y1)(1x21)(zz))ψ(x1x2)dx1dx2d×z.h(y)=\int_{\mathbf{F}^{\times}}\int_{\mathbf{F}^{2}}f\left(\begin{pmatrix}1&x_{1}\\ &1\end{pmatrix}\begin{pmatrix}&-y\\ 1&\end{pmatrix}\begin{pmatrix}1&x_{2}\\ &1\end{pmatrix}\begin{pmatrix}z&\\ &z\end{pmatrix}\right)\psi(-x_{1}-x_{2})\mathrm{d}x_{1}\mathrm{d}x_{2}\mathrm{d}^{\times}z.
Proposition 2.1.

As f\displaystyle f traverses Cc(𝐆2(𝐅))\displaystyle{\rm C}_{c}^{\infty}(\mathbf{G}_{2}(\mathbf{F})), the Bessel orbital integrals h\displaystyle h traverses

Cc(𝐅×)En\displaystyle{\rm C}_{c}^{\infty}(\mathbf{F}^{\times})\bigoplus\mathbb{C}E_{\geqslant n}

where n0\displaystyle n\in\mathbb{Z}_{\geqslant 0} can be chosen arbitrarily and we have written

(2.3) En(y2):=m=nEm(y2)=𝟙v𝐅(y)n|y|𝐅±±1+𝒫𝐅v𝐅(y)2ψ(y(u+u1))du.E_{\geqslant n}(y^{2}):=\sideset{}{{}_{m=n}^{\infty}}{\sum}E_{m}(y^{2})=\mathbbm{1}_{v_{\mathbf{F}}(y)\leqslant-n}\cdot\lvert y\rvert_{\mathbf{F}}\sideset{}{{}_{\pm}}{\sum}\int_{\pm 1+\mathcal{P}_{\mathbf{F}}^{\lfloor-\frac{v_{\mathbf{F}}(y)}{2}\rfloor}}\psi\left(y(u+u^{-1})\right)\mathrm{d}u.
Proof.

If f\displaystyle f traverses Cc(𝐁2(𝐅)w𝐍2(𝐅))\displaystyle{\rm C}_{c}^{\infty}(\mathbf{B}_{2}(\mathbf{F})w\mathbf{N}_{2}(\mathbf{F})), then clearly h\displaystyle h traverses Cc(𝐅×)\displaystyle{\rm C}_{c}^{\infty}(\mathbf{F}^{\times}). Consider a function fCc(𝐁2(𝐅)𝐍2(𝐅)t)\displaystyle f\in{\rm C}_{c}^{\infty}(\mathbf{B}_{2}(\mathbf{F})\mathbf{N}_{2}(\mathbf{F})^{t}) and let ϕCc(𝐅××𝐅)\displaystyle\phi\in{\rm C}_{c}^{\infty}(\mathbf{F}^{\times}\times\mathbf{F}) be defined by

ϕ(y,x):=𝐅×𝐅2f((1x11)(y1)(1x1)(zz))ψ(x1)dx1d×z.\displaystyle\phi(y,x):=\int_{\mathbf{F}^{\times}}\int_{\mathbf{F}^{2}}f\left(\begin{pmatrix}1&x_{1}\\ &1\end{pmatrix}\begin{pmatrix}y&\\ &1\end{pmatrix}\begin{pmatrix}1&\\ x&1\end{pmatrix}\begin{pmatrix}z&\\ &z\end{pmatrix}\right)\psi(-x_{1})\mathrm{d}x_{1}\mathrm{d}^{\times}z.

From the equation of matrices

(y1)(1x21)=(1y/x21)(y/x2x2)(11/x21)\displaystyle\begin{pmatrix}&-y\\ 1&\end{pmatrix}\begin{pmatrix}1&x_{2}\\ &1\end{pmatrix}=\begin{pmatrix}1&-y/x_{2}\\ &1\end{pmatrix}\begin{pmatrix}y/x_{2}&\\ &x_{2}\end{pmatrix}\begin{pmatrix}1&\\ 1/x_{2}&1\end{pmatrix}

we easily deduce that the relative orbital integral (2.2) is given by

h(y)=𝐅ϕ(yx22,1x2)ψ(yx2x2)dx2=𝐅ϕ(yu2,u)ψ(yuu1)|u|𝐅2du.\displaystyle h(y)=\int_{\mathbf{F}}\phi\left(\frac{y}{x_{2}^{2}},\frac{1}{x_{2}}\right)\psi\left(-\frac{y}{x_{2}}-x_{2}\right)\mathrm{d}x_{2}=\int_{\mathbf{F}}\phi(yu^{2},u)\psi(-yu-u^{-1})\lvert u\rvert_{\mathbf{F}}^{-2}\mathrm{d}u.

Let τ{1,ε}\displaystyle\tau\in\{1,\varepsilon\}, we obtain by an obvious change of variables

h(τy2)=|y|𝐅𝐅ϕ(τu2,y1u)ψ(y(τu+u1))|u|𝐅2du.\displaystyle h(\tau y^{2})=\lvert y\rvert_{\mathbf{F}}\int_{\mathbf{F}}\phi(\tau u^{2},y^{-1}u)\psi(-y(\tau u+u^{-1}))\lvert u\rvert_{\mathbf{F}}^{-2}\mathrm{d}u.

Choose i01\displaystyle i_{0}\in\mathbb{Z}_{\geqslant 1} and k0,m0\displaystyle k_{0},m_{0}\in\mathbb{Z} such that:

  • ϕ(y(1+δ1),x(1+δ2))=ϕ(y,x),δ1,δ2𝔭i0\displaystyle\phi(y(1+\delta_{1}),x(1+\delta_{2}))=\phi(y,x),\quad\forall\ \delta_{1},\delta_{2}\in\mathfrak{p}^{i_{0}};

  • for any y𝔭2+2k0\displaystyle y\in\mathfrak{p}^{2+2k_{0}} and any x𝐅\displaystyle x\in\mathbf{F}, we have ϕ(y,x)=0\displaystyle\phi(y,x)=0;

  • for any x𝔭m0\displaystyle x\in\mathfrak{p}^{m_{0}} and any y𝐅×\displaystyle y\in\mathbf{F}^{\times}, we have ϕ(y,x)=ϕ(y,0)\displaystyle\phi(y,x)=\phi(y,0).

Let v𝐅(y)=m\displaystyle v_{\mathbf{F}}(y)=-m for some mmax(m0,2i0+k0,2i0)\displaystyle m\geqslant\max(m_{0},2i_{0}+k_{0},2i_{0}), and take any n1\displaystyle n\in\mathbb{Z}_{\geqslant 1} satisfying

2nm+k0,ni0,mni0.\displaystyle 2n\geqslant m+k_{0},\quad n\geqslant i_{0},\quad m-n\geqslant i_{0}.

Performing the level n\displaystyle n regularization to dy\displaystyle\mathrm{d}y we get

h(τy2)\displaystyle\displaystyle h(\tau y^{2}) =|y|𝐅𝐅𝒫𝐅1+k0ϕ(τu2,y1u)[𝒪𝐅ψ(y(τu(1+ϖ𝐅nx)+u1(1+ϖ𝐅nx)1))dx]|u|𝐅2du\displaystyle\displaystyle=\lvert y\rvert_{\mathbf{F}}\int_{\mathbf{F}-\mathcal{P}_{\mathbf{F}}^{1+k_{0}}}\phi(\tau u^{2},y^{-1}u)\left[\oint_{\mathcal{O}_{\mathbf{F}}}\psi\left(-y\left(\tau u(1+\varpi_{\mathbf{F}}^{n}x)+u^{-1}(1+\varpi_{\mathbf{F}}^{n}x)^{-1}\right)\right)\mathrm{d}x\right]\lvert u\rvert_{\mathbf{F}}^{-2}\mathrm{d}u
=|y|𝐅𝐅𝒫𝐅1+k0ϕ(τu2,y1u)ψ(y(τu+u1))[𝒪𝐅ψ(yϖ𝐅n(τuu1)x)dx]|u|𝐅2du.\displaystyle\displaystyle=\lvert y\rvert_{\mathbf{F}}\int_{\mathbf{F}-\mathcal{P}_{\mathbf{F}}^{1+k_{0}}}\phi(\tau u^{2},y^{-1}u)\psi(-y(\tau u+u^{-1}))\left[\oint_{\mathcal{O}_{\mathbf{F}}}\psi\left(-y\varpi_{\mathbf{F}}^{n}(\tau u-u^{-1})x\right)\mathrm{d}x\right]\lvert u\rvert_{\mathbf{F}}^{-2}\mathrm{d}u.

The non-vanishing of the inner integral implies

v𝐅(τuu1)mni0,\displaystyle v_{\mathbf{F}}(\tau u-u^{-1})\geqslant m-n\geqslant i_{0},

which can be satisfied only if τ\displaystyle\tau is a square modulo 𝒫𝐅\displaystyle\mathcal{P}_{\mathbf{F}}, i.e., τ=1\displaystyle\tau=1. Moreover, we have v𝐅(uu1)i0u±1+𝒫𝐅i0\displaystyle v_{\mathbf{F}}(u-u^{-1})\geqslant i_{0}\Leftrightarrow u\in\pm 1+\mathcal{P}_{\mathbf{F}}^{i_{0}}. We therefore get h(εy2)=0\displaystyle h(\varepsilon y^{2})=0 and

h(y2)\displaystyle\displaystyle h(y^{2}) =𝟙k00ϕ(1,0)|y|𝐅±±1+𝒫𝐅i0ψ(y(u+u1))du\displaystyle\displaystyle=\mathbbm{1}_{k_{0}\geqslant 0}\phi(1,0)\cdot\lvert y\rvert_{\mathbf{F}}\sum_{\pm}\int_{\pm 1+\mathcal{P}_{\mathbf{F}}^{i_{0}}}\psi(-y(u+u^{-1}))\mathrm{d}u
=𝟙k00ϕ(1,0)|y|𝐅±±1+𝒫𝐅i0ψ(y(u+u1))[𝒪𝐅ψ(yϖ𝐅n(uu1)x)dx]du,\displaystyle\displaystyle=\mathbbm{1}_{k_{0}\geqslant 0}\phi(1,0)\cdot\lvert y\rvert_{\mathbf{F}}\sum_{\pm}\int_{\pm 1+\mathcal{P}_{\mathbf{F}}^{i_{0}}}\psi(-y(u+u^{-1}))\left[\oint_{\mathcal{O}_{\mathbf{F}}}\psi\left(-y\varpi_{\mathbf{F}}^{n}(u-u^{-1})x\right)\mathrm{d}x\right]\mathrm{d}u,

where we have performed the level n=m/2i0\displaystyle n=\lceil m/2\rceil\geqslant i_{0} regularization to du\displaystyle\mathrm{d}u. Again the inner integral is non-vanishing only if

v𝐅(uu1)mn=m/2u±1+𝒫𝐅m/2.\displaystyle v_{\mathbf{F}}(u-u^{-1})\geqslant m-n=\lfloor m/2\rfloor\quad\Leftrightarrow\quad u\in\pm 1+\mathcal{P}_{\mathbf{F}}^{\lfloor m/2\rfloor}.

We have obtained

h(y2)=𝟙k00ϕ(1,0)|y|𝐅±±1+𝒫𝐅v𝐅(y)2ψ(y(u+u1))du,\displaystyle h(y^{2})=\mathbbm{1}_{k_{0}\geqslant 0}\phi(1,0)\cdot\lvert y\rvert_{\mathbf{F}}\sum_{\pm}\int_{\pm 1+\mathcal{P}_{\mathbf{F}}^{\lfloor-\frac{v_{\mathbf{F}}(y)}{2}\rfloor}}\psi(-y(u+u^{-1}))\mathrm{d}u,

hence h\displaystyle h lies in the desired space of functions. Applying a smooth partition of unity to the open covering

𝐆2(𝐅)=𝐁2(𝐅)w𝐍2(𝐅)𝐁2(𝐅)𝐍2(𝐅)t\displaystyle\mathbf{G}_{2}(\mathbf{F})=\mathbf{B}_{2}(\mathbf{F})w\mathbf{N}_{2}(\mathbf{F})\bigcup\mathbf{B}_{2}(\mathbf{F})\mathbf{N}_{2}(\mathbf{F})^{t}

we conclude the proof. ∎

Definition 2.2.

We call En\displaystyle E_{\geqslant n} the elementary Bessel orbital integrals, abbreviated as EBOIs.

2.2. Voronoi–Hankel Transforms of EBOIs

We notice that the Mellin transform of Em\displaystyle E_{m} is simple.

Lemma 2.3.

Let m2\displaystyle m\geqslant 2 and χ\displaystyle\chi be a quasi-character. We have

𝐅×Em(y)χ(y)d×y=𝟙𝔠(χ)=mζ𝐅(1)γ(1/2,χ1,ψ)2.\displaystyle\int_{\mathbf{F}^{\times}}E_{m}(y)\chi(y)\mathrm{d}^{\times}y=\mathbbm{1}_{\mathfrak{c}(\chi)=m}\cdot\zeta_{\mathbf{F}}(1)\gamma(1/2,\chi^{-1},\psi)^{2}.
Proof.

Applying the change of variables yy2\displaystyle y\mapsto y^{2} and the level m/2\displaystyle\lceil m/2\rceil regularization to du\displaystyle\mathrm{d}u we get

𝐅×Em(y)χ(y)d×y=12ϖ𝐅m𝒪𝐅×Em(y2)χ2(y)d×y=qm1+𝒫𝐅m2(ϖ𝐅m𝒪𝐅×ψ(y(u+u1))χ2(y)d×y)du=qm𝒪𝐅×(ϖ𝐅m𝒪𝐅×ψ(y(u+u1)χ2(y)d×y)du=ζ𝐅(1)qm𝒪𝐅×𝒪𝐅×ψ(y2u+u1ϖ𝐅m)χ2(yϖ𝐅m)dydu.\int_{\mathbf{F}^{\times}}E_{m}(y)\chi(y)\mathrm{d}^{\times}y=\frac{1}{2}\int_{\varpi_{\mathbf{F}}^{-m}\mathcal{O}_{\mathbf{F}}^{\times}}E_{m}(y^{2})\chi^{2}(y)\mathrm{d}^{\times}y\\ =q^{m}\int_{1+\mathcal{P}_{\mathbf{F}}^{\lfloor\frac{m}{2}\rfloor}}\left(\int_{\varpi_{\mathbf{F}}^{-m}\mathcal{O}_{\mathbf{F}}^{\times}}\psi\left(y(u+u^{-1})\right)\chi^{2}(y)\mathrm{d}^{\times}y\right)\mathrm{d}u\\ =q^{m}\int_{\mathcal{O}_{\mathbf{F}}^{\times}}\left(\int_{\varpi_{\mathbf{F}}^{-m}\mathcal{O}_{\mathbf{F}}^{\times}}\psi\left(y(u+u^{-1}\right)\chi^{2}(y)\mathrm{d}^{\times}y\right)\mathrm{d}u\\ =\zeta_{\mathbf{F}}(1)q^{m}\int_{\mathcal{O}_{\mathbf{F}}^{\times}}\int_{\mathcal{O}_{\mathbf{F}}^{\times}}\psi\left(\frac{y^{2}u+u^{-1}}{\varpi_{\mathbf{F}}^{m}}\right)\chi^{2}\left(\frac{y}{\varpi_{\mathbf{F}}^{m}}\right)\mathrm{d}y\mathrm{d}u.

While the level m/2\displaystyle\lceil m/2\rceil regularization to du\displaystyle\mathrm{d}u also implies

𝒪𝐅×𝒪𝐅×ψ(εy2u+u1ϖ𝐅m)χ(εy2ϖ𝐅2m)dydu=0,\displaystyle\int_{\mathcal{O}_{\mathbf{F}}^{\times}}\int_{\mathcal{O}_{\mathbf{F}}^{\times}}\psi\left(\frac{\varepsilon y^{2}u+u^{-1}}{\varpi_{\mathbf{F}}^{m}}\right)\chi\left(\frac{\varepsilon y^{2}}{\varpi_{\mathbf{F}}^{2m}}\right)\mathrm{d}y\mathrm{d}u=0,

we can sum the above two equations to get

𝐅×Em(y)χ(y)d×y=ζ𝐅(1)qm𝒪𝐅×𝒪𝐅×ψ(yu+u1ϖ𝐅m)χ(yϖ𝐅2m)dydu=ζ𝐅(1)qm(𝒪𝐅×ψ(tϖ𝐅m)χ(tϖ𝐅m)dt)2.\int_{\mathbf{F}^{\times}}E_{m}(y)\chi(y)\mathrm{d}^{\times}y=\zeta_{\mathbf{F}}(1)q^{m}\int_{\mathcal{O}_{\mathbf{F}}^{\times}}\int_{\mathcal{O}_{\mathbf{F}}^{\times}}\psi\left(\frac{yu+u^{-1}}{\varpi_{\mathbf{F}}^{m}}\right)\chi\left(\frac{y}{\varpi_{\mathbf{F}}^{2m}}\right)\mathrm{d}y\mathrm{d}u\\ =\zeta_{\mathbf{F}}(1)q^{m}\left(\int_{\mathcal{O}_{\mathbf{F}}^{\times}}\psi\left(\frac{t}{\varpi_{\mathbf{F}}^{m}}\right)\chi\left(\frac{t}{\varpi_{\mathbf{F}}^{m}}\right)\mathrm{d}t\right)^{2}.

The last integral was studied in [20, Proposition 4.6], which is non-vanishing if and only if 𝔠(χ)=m\displaystyle\mathfrak{c}(\chi)=m. Its relation with the local gamma factor

𝒪𝐅×ψ(tϖ𝐅m)χ(tϖ𝐅m)dt=γ(1,χ1,ψ)\displaystyle\int_{\mathcal{O}_{\mathbf{F}}^{\times}}\psi\left(\frac{t}{\varpi_{\mathbf{F}}^{m}}\right)\chi\left(\frac{t}{\varpi_{\mathbf{F}}^{m}}\right)\mathrm{d}t=\gamma(1,\chi^{-1},\psi)

is the content of [5, Exercise 23.5]. ∎

Proposition 2.4.

There is a=a(Π)2\displaystyle a=a(\Pi)\in\mathbb{Z}_{\geqslant 2}, called the stability barrier of Π\displaystyle\Pi, such that for any ma\displaystyle m\geqslant a

𝒱Π,ψ𝔪1(Em)(t)=ψ(t)𝟙ϖm𝒪𝐅×(t),𝒱~Π,ψ𝔪1(Em)(t)=ψ(t)𝟙𝒫𝐅m(t1).\displaystyle\mathcal{VH}_{\Pi,\psi}\circ\mathfrak{m}_{-1}(E_{m})(t)=\psi(t)\cdot\mathbbm{1}_{\varpi^{-m}\mathcal{O}_{\mathbf{F}}^{\times}}(t),\quad\widetilde{\mathcal{VH}}_{\Pi,\psi}\circ\mathfrak{m}_{-1}(E_{\geqslant m})(t)=\psi(t)\cdot\mathbbm{1}_{\mathcal{P}_{\mathbf{F}}^{m}}(t^{-1}).

Moreover, we have amax(2𝔠(Π),1)\displaystyle a\leqslant\max(2\mathfrak{c}(\Pi),1). More precisely, we have:

  • (1)

    If Π\displaystyle\Pi is equal to or is included in μ1μ2μ3\displaystyle\mu_{1}\boxplus\mu_{2}\boxplus\mu_{3} we can take a=max(2𝔠(μ1),2𝔠(μ2),2𝔠(μ3),1)\displaystyle a=\max(2\mathfrak{c}(\mu_{1}),2\mathfrak{c}(\mu_{2}),2\mathfrak{c}(\mu_{3}),1);

  • (2)

    If Π=πμ\displaystyle\Pi=\pi\boxplus\mu for a supercuspidal π\displaystyle\pi of 𝐆2(𝐅)\displaystyle\mathbf{G}_{2}(\mathbf{F}) we can take a=max(𝔠(π),2𝔠(μ))\displaystyle a=\max(\mathfrak{c}(\pi),2\mathfrak{c}(\mu));

  • (3)

    If Π\displaystyle\Pi is supercuspidal we have a2𝔠(Π)\displaystyle a\leqslant 2\mathfrak{c}(\Pi).

Proof.

By Lemma 2.3 and the local functional equation, the integral

𝐅×𝒱Π,ψ𝔪1(Em)(t)χ1(t)|t|𝐅sd×t=γ(s,Π×χ,ψ)𝐅×Em(y)χ(y)|y|𝐅s1d×y\displaystyle\int_{\mathbf{F}^{\times}}\mathcal{VH}_{\Pi,\psi}\circ\mathfrak{m}_{-1}(E_{m})(t)\chi^{-1}(t)\lvert t\rvert_{\mathbf{F}}^{-s}\mathrm{d}^{\times}t=\gamma(s,\Pi\times\chi,\psi)\int_{\mathbf{F}^{\times}}E_{m}(y)\chi(y)\lvert y\rvert_{\mathbf{F}}^{s-1}\mathrm{d}^{\times}y

is vanishing for any χ\displaystyle\chi with 𝔠(χ)m\displaystyle\mathfrak{c}(\chi)\neq m. By the stability of the local gamma factors (see [11, Proposition (2.2)], [5, Theorem 23.8] and [5, Exercise 23.5]), there is a2\displaystyle a\in\mathbb{Z}_{\geqslant 2} depending only on Π\displaystyle\Pi so that the factors

γ(s,Π×χ,ψ)=γ(s,χ,ψ)3\displaystyle\gamma(s,\Pi\times\chi,\psi)=\gamma(s,\chi,\psi)^{3}

depend only on ωΠ=𝟙\displaystyle\omega_{\Pi}=\mathbbm{1} if 𝔠(χ)a\displaystyle\mathfrak{c}(\chi)\geqslant a. Hence for 𝔠(χ)=ma\displaystyle\mathfrak{c}(\chi)=m\geqslant a we have by Lemma 2.3

γ(s,Π×χ,ψ)𝐅×Em(y)χ(y)|y|𝐅s1d×y=γ(s,χ,ψ)3ζ𝐅(1)γ(3/2s,χ1,ψ)2=ζ𝐅(1)γ(s+1,χ,ψ).\displaystyle\gamma(s,\Pi\times\chi,\psi)\int_{\mathbf{F}^{\times}}E_{m}(y)\chi(y)\lvert y\rvert_{\mathbf{F}}^{s-1}\mathrm{d}^{\times}y=\gamma(s,\chi,\psi)^{3}\cdot\zeta_{\mathbf{F}}(1)\gamma(3/2-s,\chi^{-1},\psi)^{2}=\zeta_{\mathbf{F}}(1)\gamma(s+1,\chi,\psi).

One verifies easily the desired formula for 𝒱Π,ψ(Em)\displaystyle\mathcal{VH}_{\Pi,\psi}(E_{m}) by comparing their Mellin transforms. The desired formula for Em\displaystyle E_{\geqslant m} follows easily by taking limits in the sense of tempered distributions on M3(𝐅)\displaystyle{\rm M}_{3}(\mathbf{F}) via I3:C(𝐅×)C(GL3(𝐅))\displaystyle I_{3}:{\rm C}(\mathbf{F}^{\times})\to{\rm C}({\rm GL}_{3}(\mathbf{F})) by [19, Theorem 1.3]. In the “moreover” part, (1) and (2) follow from the effective versions of the stability theorems for 𝐆1\displaystyle\mathbf{G}_{1} and 𝐆2\displaystyle\mathbf{G}_{2} [5, Theorem 23.8 & 25.7] together with the multiplicativity of the local gamma factors. For (3) we can take a=max(2𝔠(Π),6)\displaystyle a=\max(2\mathfrak{c}(\Pi),6) by examining the proof in [11, §2]. Now that a supercuspidal Π\displaystyle\Pi of 𝐆3\displaystyle\mathbf{G}_{3} has 𝔠(Π)3\displaystyle\mathfrak{c}(\Pi)\geqslant 3, we conclude the last assertion. ∎

Remark 2.5.

It would be interesting to know a sharp bound for a\displaystyle a, say in general for stability theorems for 𝐆n×𝐆t\displaystyle\mathbf{G}_{n}\times\mathbf{G}_{t}. It could follow from the work of Bushnell–Henniart–Kutzko [6].

3. Local Weight Functions

3.1. Choice of Test Functions

The target representation π0PGL2(𝐅)^\displaystyle\pi_{0}\in\widehat{{\rm PGL}_{2}(\mathbf{F})} are:

  • (1)

    (Split) π(χ0,χ01)\displaystyle\pi(\chi_{0},\chi_{0}^{-1}) with a ramified and ϑ3\displaystyle\vartheta_{3}-tempered quasi-character χ0\displaystyle\chi_{0} of 𝐅×\displaystyle\mathbf{F}^{\times};

  • (2)

    (Special) the quadratic twists Stη\displaystyle\mathrm{St}_{\eta} of the Steinberg representation St\displaystyle\mathrm{St};

  • (3)

    (Dihedral) πβ\displaystyle\pi_{\beta} with a unitary regular character β\displaystyle\beta of 𝐄×\displaystyle\mathbf{E}^{\times}, 𝐄/𝐅\displaystyle\mathbf{E}/\mathbf{F} being a quadratic field extension.

Note that Stη\displaystyle\mathrm{St}_{\eta} is a sub-representation of π(η||𝐅1/2,η||𝐅1/2)\displaystyle\pi(\eta\lvert\cdot\rvert_{\mathbf{F}}^{1/2},\eta\lvert\cdot\rvert_{\mathbf{F}}^{-1/2}).

Let 𝐋\displaystyle\mathbf{L} be a separable quadratic algebra extension of 𝐅\displaystyle\mathbf{F}. The non-trivial element of the group Aut𝐅(𝐋)\displaystyle\mathrm{Aut}_{\mathbf{F}}(\mathbf{L}) is denoted by 𝐋𝐋,xx¯\displaystyle\mathbf{L}\to\mathbf{L},x\mapsto\bar{x}. Define

Tr=Tr𝐋/𝐅:𝐋𝐅,xx+x¯;Nr=Nr𝐋/𝐅:𝐋×𝐅×,xxx¯;|x|𝐋:=|Nr(x)|.\displaystyle{\rm Tr}={\rm Tr}_{\mathbf{L}/\mathbf{F}}:\mathbf{L}\to\mathbf{F},\ x\mapsto x+\bar{x};\quad{\rm Nr}={\rm Nr}_{\mathbf{L}/\mathbf{F}}:\mathbf{L}^{\times}\to\mathbf{F}^{\times},\ x\mapsto x\bar{x};\quad\lvert x\rvert_{\mathbf{L}}:=\lvert{\rm Nr}(x)\rvert.

We associate to each target π0\displaystyle\pi_{0} a parameter (𝐋,β)\displaystyle(\mathbf{L},\beta) by:

  • (1)

    π(χ0,χ01)\displaystyle\pi(\chi_{0},\chi_{0}^{-1}): Let 𝐋𝐅𝐅\displaystyle\mathbf{L}\simeq\mathbf{F}\oplus\mathbf{F} and β:𝐋×𝐅××𝐅×S1,(t1,t2)χ0(t1t21)\displaystyle\beta:\mathbf{L}^{\times}\simeq\mathbf{F}^{\times}\times\mathbf{F}^{\times}\to\mathrm{S}^{1},(t_{1},t_{2})\mapsto\chi_{0}(t_{1}t_{2}^{-1});

  • (2)

    Stη\displaystyle\mathrm{St}_{\eta}: Let 𝐋𝐅𝐅\displaystyle\mathbf{L}\simeq\mathbf{F}\oplus\mathbf{F} and β:𝐋×𝐅××𝐅××,(t1,t2)η(t1t21)|t1t21|𝐅1/2\displaystyle\beta:\mathbf{L}^{\times}\simeq\mathbf{F}^{\times}\times\mathbf{F}^{\times}\to\mathbb{C}^{\times},(t_{1},t_{2})\mapsto\eta(t_{1}t_{2}^{-1})\lvert t_{1}t_{2}^{-1}\rvert_{\mathbf{F}}^{1/2};

  • (3)

    πβ\displaystyle\pi_{\beta}: Let 𝐋=𝐄\displaystyle\mathbf{L}=\mathbf{E} and β\displaystyle\beta be the obvious one.

We equip 𝐋\displaystyle\mathbf{L} with the self-dual Haar measure dz\displaystyle\mathrm{d}z with respect to ψ𝐋:=ψTr\displaystyle\psi_{\mathbf{L}}:=\psi\circ{\rm Tr}. Write

𝒪𝐋:={x𝐋|Tr(x),Nr(x)𝒪𝐅}.\displaystyle\mathcal{O}_{\mathbf{L}}:=\left\{x\in\mathbf{L}\ \middle|\ {\rm Tr}(x),{\rm Nr}(x)\in\mathcal{O}_{\mathbf{F}}\right\}.
Definition 3.1.

Define 𝒫𝐋:=ϖ𝐋𝒪𝐋\displaystyle\mathcal{P}_{\mathbf{L}}:=\varpi_{\mathbf{L}}\mathcal{O}_{\mathbf{L}} and ϖ𝐋:=ϖ𝐅\displaystyle\varpi_{\mathbf{L}}:=\varpi_{\mathbf{F}} if 𝐋\displaystyle\mathbf{L} is split, otherwise ϖ𝐋\displaystyle\varpi_{\mathbf{L}} is a uniformizer of 𝐋\displaystyle\mathbf{L}. For any (quasi-)character β\displaystyle\beta of 𝐋×\displaystyle\mathbf{L}^{\times} we define its 𝐅\displaystyle\mathbf{F}-norm-1\displaystyle 1 conductor as

𝔠1(β):=min{n|β(𝐋1(1+𝒫𝐋n))={1}}.\displaystyle\mathfrak{c}_{1}(\beta):=\min\left\{n\ \middle|\ \beta\left(\mathbf{L}^{1}\cap(1+\mathcal{P}_{\mathbf{L}}^{n})\right)=\{1\}\right\}.

If 𝔠1(β)>0\displaystyle\mathfrak{c}_{1}(\beta)>0 we say that β\displaystyle\beta is a regular character.

Remark 3.2.

Directly from the definition we have for any (quasi-)character β\displaystyle\beta of 𝐋×\displaystyle\mathbf{L}^{\times}

  • 𝔠1(β)𝔠(β)\displaystyle\mathfrak{c}_{1}(\beta)\leqslant\mathfrak{c}(\beta);

  • 𝔠1(β)=𝔠1(β(χNr))\displaystyle\mathfrak{c}_{1}(\beta)=\mathfrak{c}_{1}(\beta\cdot(\chi\circ{\rm Nr})) for any (quasi-)character χ\displaystyle\chi of 𝐅×\displaystyle\mathbf{F}^{\times}.

Lemma 3.3.

If β\displaystyle\beta is a regular character of 𝐋×\displaystyle\mathbf{L}^{\times} so that the restriction β𝐅×=η𝐋/𝐅\displaystyle\beta\mid_{\mathbf{F}^{\times}}=\eta_{\mathbf{L}/\mathbf{F}} coincides with the quadratic character associated with the quadratic extension 𝐋/𝐅\displaystyle\mathbf{L}/\mathbf{F}, then we have 𝔠1(β)=𝔠(β)\displaystyle\mathfrak{c}_{1}(\beta)=\mathfrak{c}(\beta).

Proof.

Write n0:=𝔠(β)(𝔠1(β)1)\displaystyle n_{0}:=\mathfrak{c}(\beta)(\geqslant\mathfrak{c}_{1}(\beta)\geqslant 1). We shall prove that 𝔠1(β)<n0\displaystyle\mathfrak{c}_{1}(\beta)<n_{0} is impossible.

(1) If 𝐋𝐅𝐅\displaystyle\mathbf{L}\simeq\mathbf{F}\oplus\mathbf{F} is split then η𝐋/𝐅=𝟙\displaystyle\eta_{\mathbf{L}/\mathbf{F}}=\mathbbm{1}. The assertion follows from 𝔠(χ0)=𝔠(χ02)\displaystyle\mathfrak{c}(\chi_{0})=\mathfrak{c}(\chi_{0}^{2}), because taking square is a group automorphism of 1+𝒫𝐅\displaystyle 1+\mathcal{P}_{\mathbf{F}} and the regularity of β\displaystyle\beta is equivalent with 𝔠(χ02)>0\displaystyle\mathfrak{c}(\chi_{0}^{2})>0.

(2) Assume 𝐋/𝐅\displaystyle\mathbf{L}/\mathbf{F} is non-split. Then β\displaystyle\beta coincides with η𝐋/𝐅\displaystyle\eta_{\mathbf{L}/\mathbf{F}} on 𝐅×\displaystyle\mathbf{F}^{\times}. If n0=1\displaystyle n_{0}=1 then the assertion follows from the condition β𝐄1𝟙\displaystyle\beta\mid_{\mathbf{E}^{1}}\neq\mathbbm{1} (equivalent to β\displaystyle\beta being regular). Assume n02\displaystyle n_{0}\geqslant 2 from now on. We choose the uniformizers ϖ𝐅\displaystyle\varpi_{\mathbf{F}} and ϖ𝐋\displaystyle\varpi_{\mathbf{L}} of 𝒪𝐅\displaystyle\mathcal{O}_{\mathbf{F}} and 𝒪𝐋\displaystyle\mathcal{O}_{\mathbf{L}} respectively so that

(3.1) {ϖ𝐋=ϖ𝐅if e=e(𝐋/𝐅)=1ϖ𝐋¯=ϖ𝐋,ϖ𝐅=Nr(ϖ𝐋)=ϖ𝐋2if e=e(𝐄/𝐅)=2.\begin{cases}\varpi_{\mathbf{L}}=\varpi_{\mathbf{F}}&\text{if }e=e(\mathbf{L}/\mathbf{F})=1\\ \overline{\varpi_{\mathbf{L}}}=-\varpi_{\mathbf{L}},\varpi_{\mathbf{F}}={\rm Nr}(\varpi_{\mathbf{L}})=-\varpi_{\mathbf{L}}^{2}&\text{if }e=e(\mathbf{E}/\mathbf{F})=2\end{cases}.

Claim: We have 2n0\displaystyle 2\mid n_{0} in the ramified case.

Proof of Claim:.

In fact, if β\displaystyle\beta is trivial on 1+𝒫𝐋2n+1\displaystyle 1+\mathcal{P}_{\mathbf{L}}^{2n+1} with n1\displaystyle n\geqslant 1, then for any element α1+𝒫𝐋2n\displaystyle\alpha\in 1+\mathcal{P}_{\mathbf{L}}^{2n} we can find u0𝒪𝐅\displaystyle u_{0}\in\mathcal{O}_{\mathbf{F}} and u1𝒪𝐋\displaystyle u_{1}\in\mathcal{O}_{\mathbf{L}} such that α=1+ϖ𝐋2nu0+ϖ𝐋2n+1u1\displaystyle\alpha=1+\varpi_{\mathbf{L}}^{2n}u_{0}+\varpi_{\mathbf{L}}^{2n+1}u_{1}, since the residue class fields of 𝐋\displaystyle\mathbf{L} and 𝐅\displaystyle\mathbf{F} are isomorphic. Then β(α)=η𝐋/𝐅(1+(ϖ𝐅)nu0)=1\displaystyle\beta(\alpha)=\eta_{\mathbf{L}/\mathbf{F}}(1+(-\varpi_{\mathbf{F}})^{n}u_{0})=1. Hence β\displaystyle\beta is also trivial on 1+𝒫𝐋2n\displaystyle 1+\mathcal{P}_{\mathbf{L}}^{2n}. ∎

With these reductions, we find α1+𝒫𝐋n01\displaystyle\alpha\in 1+\mathcal{P}_{\mathbf{L}}^{n_{0}-1} such that β(α)1\displaystyle\beta(\alpha)\neq 1. But Nr(α)𝐅(1+𝒫𝐋n01)=1+𝒫𝐅(n01)/e\displaystyle{\rm Nr}(\alpha)\in\mathbf{F}\cap(1+\mathcal{P}_{\mathbf{L}}^{n_{0}-1})=1+\mathcal{P}_{\mathbf{F}}^{\lceil(n_{0}-1)/e\rceil} is a square. Hence Nr(α)=k2\displaystyle{\rm Nr}(\alpha)=k^{2} for some k1+𝒫𝐅(n01)/e1+𝒫𝐋n01\displaystyle k\in 1+\mathcal{P}_{\mathbf{F}}^{\lceil(n_{0}-1)/e\rceil}\subset 1+\mathcal{P}_{\mathbf{L}}^{n_{0}-1}, and β(k1α)=β(α)1\displaystyle\beta(k^{-1}\alpha)=\beta(\alpha)\neq 1 with k1α𝐋1(1+𝒫𝐋n01)\displaystyle k^{-1}\alpha\in\mathbf{L}^{1}\cap(1+\mathcal{P}_{\mathbf{L}}^{n_{0}-1}), proving the assertion. ∎

Corollary 3.4.

Any dihedral supercuspidal representations πβ\displaystyle\pi_{\beta} with trivial central character (the case (3) in the beginning of this subsection) is twisted minimal.

Proof.

For any (quasi-)character χ\displaystyle\chi of 𝐅×\displaystyle\mathbf{F}^{\times} we have 𝔠(β)=𝔠1(β)=𝔠1(β(χNr))𝔠(β(χNr))\displaystyle\mathfrak{c}(\beta)=\mathfrak{c}_{1}(\beta)=\mathfrak{c}_{1}\left(\beta\cdot(\chi\circ{\rm Nr})\right)\leqslant\mathfrak{c}\left(\beta\cdot(\chi\circ{\rm Nr})\right). For any regular (quasi-)character β\displaystyle\beta of 𝐄×\displaystyle\mathbf{E}^{\times} we have by [10, Theorem 4.7] with e=e(𝐄/𝐅)\displaystyle e=e(\mathbf{E}/\mathbf{F})

(3.2) 𝔠(πβ)=𝔠(β)f(𝐄/𝐅)+𝔠(ψ𝐄)=2n0e+e1,\mathfrak{c}(\pi_{\beta})=\mathfrak{c}(\beta)f(\mathbf{E}/\mathbf{F})+\mathfrak{c}(\psi_{\mathbf{E}})=\tfrac{2n_{0}}{e}+e-1,

where f(𝐄/𝐅)\displaystyle f(\mathbf{E}/\mathbf{F}) is the residual field index and ψ𝐄=ψ𝐅Tr𝐄/𝐅\displaystyle\psi_{\mathbf{E}}=\psi_{\mathbf{F}}\circ{\rm Tr}_{\mathbf{E}/\mathbf{F}}. The desired assertion follows readily. ∎

Proposition 3.5.

Let 𝐋1\displaystyle\mathbf{L}^{1} be the kernel of the norm map Nr\displaystyle{\rm Nr}.

  • (1)

    For any x𝐋×\displaystyle x\in\mathbf{L}^{\times}, the following map

    ιx:𝐅××𝐋1𝐋×,(r,α)xrα\displaystyle\iota_{x}:\mathbf{F}^{\times}\times\mathbf{L}^{1}\to\mathbf{L}^{\times},\quad(r,\alpha)\mapsto xr\alpha

    is a 2\displaystyle 2-to-1\displaystyle 1 covering map onto an open subset of 𝐋×\displaystyle\mathbf{L}^{\times}.

  • (2)

    (Polar decomposition) The Haar measure dα\displaystyle\mathrm{d}\alpha on 𝐋1\displaystyle\mathbf{L}^{1} determined by

    (3.3) |z|𝐋1dz=|r|𝐅1drdα,z=ιx(r,α)=xrα,\lvert z\rvert_{\mathbf{L}}^{-1}\mathrm{d}z=\lvert r\rvert_{\mathbf{F}}^{-1}\mathrm{d}r\mathrm{d}\alpha,\quad z=\iota_{x}(r,\alpha)=xr\alpha,

    where dr\displaystyle\mathrm{d}r is the self-dual Haar measure of 𝐅\displaystyle\mathbf{F} with respect to ψ\displaystyle\psi, is independent of x\displaystyle x, and satisfies

    Vol(𝐋1𝒪𝐋,dα)=2e1Vol(𝒪𝐋×,dz)Vol(𝒪𝐅×,dr),\displaystyle{\rm Vol}(\mathbf{L}^{1}\cap\mathcal{O}_{\mathbf{L}},\mathrm{d}\alpha)=2^{e-1}\frac{{\rm Vol}(\mathcal{O}_{\mathbf{L}}^{\times},\mathrm{d}z)}{{\rm Vol}(\mathcal{O}_{\mathbf{F}}^{\times},\mathrm{d}r)},

    where the number e=e(𝐋/𝐅)\displaystyle e=e(\mathbf{L}/\mathbf{F}) is the generalized ramification index of 𝐋/𝐅\displaystyle\mathbf{L}/\mathbf{F} given by

    (3.4) e={1if 𝐋/𝐅 is not ramified2if 𝐋/𝐅 is ramified.e=\begin{cases}1&\text{if }\mathbf{L}/\mathbf{F}\text{ is not ramified}\\ 2&\text{if }\mathbf{L}/\mathbf{F}\text{ is ramified}\end{cases}.
  • (3)

    For any n1\displaystyle n\in\mathbb{Z}_{\geqslant 1} we have Vol(𝐋1(1+𝒫𝐋n),dα)=qnee12\displaystyle{\rm Vol}(\mathbf{L}^{1}\cap(1+\mathcal{P}_{\mathbf{L}}^{n}),\mathrm{d}\alpha)=q^{-\lfloor\frac{n}{e}\rfloor-\frac{e-1}{2}}.

Proof.

The proof of (1) is easy and omitted. The relation (ιx)=x(ι1)\displaystyle\Im(\iota_{x})=x\Im(\iota_{1}) implies the independence of x\displaystyle x in (3.3), since the measure |z|𝐋1dz\displaystyle\lvert z\rvert_{\mathbf{L}}^{-1}\mathrm{d}z is invariant by multiplication by elements in 𝐋×\displaystyle\mathbf{L}^{\times}. To show that (3.3) implies the stated formula for Vol(𝐋1𝒪𝐋,dα)\displaystyle{\rm Vol}(\mathbf{L}^{1}\cap\mathcal{O}_{\mathbf{L}},\mathrm{d}\alpha), it suffices to show

Vol(ι1(𝒪𝐅××(𝒪𝐋𝐋1)))=2e2Vol(𝒪𝐋×).\displaystyle{\rm Vol}(\iota_{1}(\mathcal{O}_{\mathbf{F}}^{\times}\times(\mathcal{O}_{\mathbf{L}}\cap\mathbf{L}^{1})))=2^{e-2}{\rm Vol}(\mathcal{O}_{\mathbf{L}}^{\times}).

Now that ι1(𝒪𝐅××(𝒪𝐋𝐋1))=𝒪𝐋×Nr1((𝒪𝐅×)2)\displaystyle\iota_{1}(\mathcal{O}_{\mathbf{F}}^{\times}\times(\mathcal{O}_{\mathbf{L}}\cap\mathbf{L}^{1}))=\mathcal{O}_{\mathbf{L}}^{\times}\cap{\rm Nr}^{-1}((\mathcal{O}_{\mathbf{F}}^{\times})^{2}), we get

[𝒪𝐋×:ι1(𝒪𝐅××(𝒪𝐋𝐋1))]=[Nr(𝒪𝐋×):(𝒪𝐅×)2]=[𝒪𝐅×:(𝒪𝐅×)2][𝒪𝐅×:Nr(𝒪𝐋×)]=22e1\displaystyle[\mathcal{O}_{\mathbf{L}}^{\times}:\iota_{1}(\mathcal{O}_{\mathbf{F}}^{\times}\times(\mathcal{O}_{\mathbf{L}}\cap\mathbf{L}^{1}))]=[{\rm Nr}(\mathcal{O}_{\mathbf{L}}^{\times}):(\mathcal{O}_{\mathbf{F}}^{\times})^{2}]=\frac{[\mathcal{O}_{\mathbf{F}}^{\times}:(\mathcal{O}_{\mathbf{F}}^{\times})^{2}]}{[\mathcal{O}_{\mathbf{F}}^{\times}:{\rm Nr}(\mathcal{O}_{\mathbf{L}}^{\times})]}=\frac{2}{2^{e-1}}

and conclude the proof of (2). To prove (3) we note that the map

σ:𝒪𝐋×𝐋1𝒪𝐋,xx/x¯\displaystyle\sigma:\mathcal{O}_{\mathbf{L}}^{\times}\to\mathbf{L}^{1}\cap\mathcal{O}_{\mathbf{L}},\quad x\mapsto x/\bar{x}

is surjective onto 𝐋1(1+𝒫𝐋n)\displaystyle\mathbf{L}^{1}\cap(1+\mathcal{P}_{\mathbf{L}}^{n}), whose pre-image is 𝒪𝐅×(1+𝒫𝐋n)\displaystyle\mathcal{O}_{\mathbf{F}}^{\times}(1+\mathcal{P}_{\mathbf{L}}^{n}). It is also easy to see that the image of σ\displaystyle\sigma is a subgroup of 𝐋1𝒪𝐋\displaystyle\mathbf{L}^{1}\cap\mathcal{O}_{\mathbf{L}} with index 2e1\displaystyle 2^{e-1}, by a refinement of Hilbert’s 90. Therefore we get

(σ)𝐋1(1+𝒫𝐋n)𝒪𝐋×𝒪𝐅×(1+𝒫𝐋n)(𝒪𝐋/𝒫𝐋n)×(𝒪𝐅/𝒫𝐅n/e)×\displaystyle\frac{\Im(\sigma)}{\mathbf{L}^{1}\cap(1+\mathcal{P}_{\mathbf{L}}^{n})}\simeq\frac{\mathcal{O}_{\mathbf{L}}^{\times}}{\mathcal{O}_{\mathbf{F}}^{\times}(1+\mathcal{P}_{\mathbf{L}}^{n})}\simeq\frac{(\mathcal{O}_{\mathbf{L}}/\mathcal{P}_{\mathbf{L}}^{n})^{\times}}{(\mathcal{O}_{\mathbf{F}}/\mathcal{P}_{\mathbf{F}}^{\lceil n/e\rceil})^{\times}}

and deduce from it the measure relation

Vol(𝐋1𝒪𝐋,dα)Vol(𝐋1(1+𝒫𝐋n),dα)=2e1q2n/eqn/eVol(𝒪𝐋×,dz)Vol(𝒪𝐅×,dr)Vol(𝒪𝐅,dr)Vol(𝒪𝐋,dz).\displaystyle\frac{{\rm Vol}(\mathbf{L}^{1}\cap\mathcal{O}_{\mathbf{L}},\mathrm{d}\alpha)}{{\rm Vol}(\mathbf{L}^{1}\cap(1+\mathcal{P}_{\mathbf{L}}^{n}),\mathrm{d}\alpha)}=2^{e-1}\frac{q^{2n/e}}{q^{\lceil n/e\rceil}}\cdot\frac{{\rm Vol}(\mathcal{O}_{\mathbf{L}}^{\times},\mathrm{d}z)}{{\rm Vol}(\mathcal{O}_{\mathbf{F}}^{\times},\mathrm{d}r)}\cdot\frac{{\rm Vol}(\mathcal{O}_{\mathbf{F}},\mathrm{d}r)}{{\rm Vol}(\mathcal{O}_{\mathbf{L}},\mathrm{d}z)}.

We conclude the proof of (3) by comparing with (2) and Vol(𝒪𝐋,dz)=qe12\displaystyle{\rm Vol}(\mathcal{O}_{\mathbf{L}},\mathrm{d}z)=q^{-\frac{e-1}{2}}. ∎

Let η𝐋/𝐅\displaystyle\eta_{\mathbf{L}/\mathbf{F}} be the quadratic character of 𝐅×\displaystyle\mathbf{F}^{\times} associated with 𝐋/𝐅\displaystyle\mathbf{L}/\mathbf{F}, which is trivial on Nr(𝐋×)\displaystyle{\rm Nr}(\mathbf{L}^{\times}). We have β𝐅×=η𝐋/𝐅\displaystyle\beta\mid_{\mathbf{F}^{\times}}=\eta_{\mathbf{L}/\mathbf{F}} since π0\displaystyle\pi_{0} has trivial central character. Write λ(𝐋/𝐅,ψ)\displaystyle\lambda(\mathbf{L}/\mathbf{F},\psi) for the Weil index, which is equal to 1\displaystyle 1 if 𝐋/𝐅\displaystyle\mathbf{L}/\mathbf{F} is not ramified (see [7, Corollary 3.2]). Let τ\displaystyle\tau run through a system of representatives of 𝒪𝐅×/(𝒪𝐅×)2\displaystyle\mathcal{O}_{\mathbf{F}}^{\times}/(\mathcal{O}_{\mathbf{F}}^{\times})^{2} (split), resp. Nr(𝐋×)/(𝐅×)2\displaystyle{\rm Nr}(\mathbf{L}^{\times})/(\mathbf{F}^{\times})^{2} (non-split), and choose any xτ𝐋×\displaystyle x_{\tau}\in\mathbf{L}^{\times} s.t. Nr(xτ)=τ\displaystyle{\rm Nr}(x_{\tau})=\tau; if 𝐋/𝐅\displaystyle\mathbf{L}/\mathbf{F} is split we also require Tr(xτ)=1+τ\displaystyle{\rm Tr}(x_{\tau})=1+\tau. Define a function H\displaystyle H on 𝐅×\displaystyle\mathbf{F}^{\times} with support contained in Nr(𝐋×)\displaystyle{\rm Nr}(\mathbf{L}^{\times}) by

(3.5) H(τy2)=λ(𝐋/𝐅,ψ)|τ|𝐅12|y|𝐅η𝐋/𝐅(y)𝐋1𝒪𝐋β(xτα)ψ(Tr(xτα)y)dα.H(\tau y^{2})=\lambda(\mathbf{L}/\mathbf{F},\psi)\lvert\tau\rvert_{\mathbf{F}}^{\frac{1}{2}}\cdot\lvert y\rvert_{\mathbf{F}}\cdot\eta_{\mathbf{L}/\mathbf{F}}(y)\cdot\int_{\mathbf{L}^{1}\cap\mathcal{O}_{\mathbf{L}}}\beta(x_{\tau}\alpha)\psi({\rm Tr}(x_{\tau}\alpha)y)\mathrm{d}\alpha.

It is clear that H\displaystyle H is independent of any choice made in the definition. For definiteness, we fix a ε𝒪𝐅×(𝒪𝐅×)2\displaystyle\varepsilon\in\mathcal{O}_{\mathbf{F}}^{\times}-(\mathcal{O}_{\mathbf{F}}^{\times})^{2} and choose a system {τ}\displaystyle\{\tau\} of representatives of 𝒪𝐅×/(𝒪𝐅×)2\displaystyle\mathcal{O}_{\mathbf{F}}^{\times}/(\mathcal{O}_{\mathbf{F}}^{\times})^{2}, resp. Nr(𝐋×)/(𝐅×)2\displaystyle{\rm Nr}(\mathbf{L}^{\times})/(\mathbf{F}^{\times})^{2} as

  • (1)

    {1,ε}\displaystyle\{1,\varepsilon\} if 𝐋/𝐅\displaystyle\mathbf{L}/\mathbf{F} is split or unramified (in the latter case 𝐋=𝐅[ε]\displaystyle\mathbf{L}=\mathbf{F}[\sqrt{\varepsilon}]),

  • (2)

    {1,ϖ𝐅}\displaystyle\{1,-\varpi_{\mathbf{F}}\} if 𝐋/𝐅\displaystyle\mathbf{L}/\mathbf{F} is ramified (with 𝐋=𝐅[ϖ𝐅]\displaystyle\mathbf{L}=\mathbf{F}[\sqrt{\varpi_{\mathbf{F}}}] and we fix a uniformizer ϖ𝐋=ϖ𝐅\displaystyle\varpi_{\mathbf{L}}=\sqrt{\varpi_{\mathbf{F}}} of 𝐋\displaystyle\mathbf{L}).

The (partial) Mellin transform of H\displaystyle H will be fundamental for our analysis. We record it as follows.

Lemma 3.6.

Let χ𝐅×^\displaystyle\chi\in\widehat{\mathbf{F}^{\times}} and n1\displaystyle n\in\mathbb{Z}_{\geqslant 1}. We have

εn(χ,H):=ϖ𝐅n𝒪𝐅×H(y)χ(y)d×y=ζ𝐅(1){𝟙2n,𝔠(χ0χ)=𝔠(χ01χ)=n/2ε(1/2,χ0χ1,ψ)ε(1/2,χ01χ1,ψ)if 𝐋/𝐅 split and 𝔠(χ0χ),𝔠(χ01χ)1𝟙n=2,𝔠(χ0)=1,𝔠(χ02)0q1χ0χ(ϖ𝐅)ε(1/2,χ01χ1,ψ)if 𝐋/𝐅 split and 𝔠(χ0χ)=0𝔠(χ01χ)𝟙n=2,𝔠(χ0)=1,𝔠(χ02)0q1χ01χ(ϖ𝐅)ε(1/2,χ0χ1,ψ)if 𝐋/𝐅 split and 𝔠(χ01χ)=0𝔠(χ0χ)𝟙n=2,𝔠(χ0)=1,𝔠(χ02)=0q2χ2(ϖ𝐅)if 𝐋/𝐅 split and 𝔠(χ0χ)=0=𝔠(χ01χ)𝟙2en,𝔠(β(χNr))=en/2+1eε(1/2,πβ1χ1,ψ)if 𝐋/𝐅 non-split.\varepsilon_{n}(\chi,H):=\int_{\varpi_{\mathbf{F}}^{-n}\mathcal{O}_{\mathbf{F}}^{\times}}H(y)\chi(y)\mathrm{d}^{\times}y=\zeta_{\mathbf{F}}(1)\cdot\\ \begin{cases}\mathbbm{1}_{2\mid n,\mathfrak{c}(\chi_{0}\chi)=\mathfrak{c}(\chi_{0}^{-1}\chi)=n/2}\cdot\varepsilon(1/2,\chi_{0}\chi^{-1},\psi)\varepsilon(1/2,\chi_{0}^{-1}\chi^{-1},\psi)&\text{if }\mathbf{L}/\mathbf{F}\text{ split and }\mathfrak{c}(\chi_{0}\chi),\mathfrak{c}(\chi_{0}^{-1}\chi)\geqslant 1\\ -\mathbbm{1}_{n=2,\mathfrak{c}(\chi_{0})=1,\mathfrak{c}(\chi_{0}^{2})\neq 0}\cdot q^{-1}\chi_{0}\chi(\varpi_{\mathbf{F}})\varepsilon(1/2,\chi_{0}^{-1}\chi^{-1},\psi)&\text{if }\mathbf{L}/\mathbf{F}\text{ split and }\mathfrak{c}(\chi_{0}\chi)=0\neq\mathfrak{c}(\chi_{0}^{-1}\chi)\\ -\mathbbm{1}_{n=2,\mathfrak{c}(\chi_{0})=1,\mathfrak{c}(\chi_{0}^{2})\neq 0}\cdot q^{-1}\chi_{0}^{-1}\chi(\varpi_{\mathbf{F}})\varepsilon(1/2,\chi_{0}\chi^{-1},\psi)&\text{if }\mathbf{L}/\mathbf{F}\text{ split and }\mathfrak{c}(\chi_{0}^{-1}\chi)=0\neq\mathfrak{c}(\chi_{0}\chi)\\ -\mathbbm{1}_{n=2,\mathfrak{c}(\chi_{0})=1,\mathfrak{c}(\chi_{0}^{2})=0}\cdot q^{-2}\chi^{2}(\varpi_{\mathbf{F}})&\text{if }\mathbf{L}/\mathbf{F}\text{ split and }\mathfrak{c}(\chi_{0}\chi)=0=\mathfrak{c}(\chi_{0}^{-1}\chi)\\ \mathbbm{1}_{2\mid en,\mathfrak{c}(\beta\cdot(\chi\circ{\rm Nr}))=en/2+1-e}\cdot\varepsilon(1/2,\pi_{\beta^{-1}}\otimes\chi^{-1},\psi)&\text{if }\mathbf{L}/\mathbf{F}\text{ non-split}\end{cases}.

In particular, we have |εn(χ,H)|ζ𝐅(1)\displaystyle\left\lvert\varepsilon_{n}(\chi,H)\right\rvert\leqslant\zeta_{\mathbf{F}}(1).

Proof.

We divide the domain of integration into cosets of square elements.

(A) In the split case, we have

ϖ𝐅n𝒪𝐅×H(y)χ(y)d×y\displaystyle\displaystyle\int_{\varpi_{\mathbf{F}}^{-n}\mathcal{O}_{\mathbf{F}}^{\times}}H(y)\chi(y)\mathrm{d}^{\times}y =𝟙2n12τ{1,ε}ϖ𝐅n2𝒪𝐅×χ(τy2)|y|(𝒪𝐅×χ0(τα2)ψ((τα+α1)y)dα)d×y\displaystyle\displaystyle=\mathbbm{1}_{2\mid n}\cdot\frac{1}{2}\sum_{\tau\in\{1,\varepsilon\}}\int_{\varpi_{\mathbf{F}}^{-\frac{n}{2}}\mathcal{O}_{\mathbf{F}}^{\times}}\chi(\tau y^{2})\lvert y\rvert\left(\int_{\mathcal{O}_{\mathbf{F}}^{\times}}\chi_{0}(\tau\alpha^{2})\psi\left((\tau\alpha+\alpha^{-1})y\right)\mathrm{d}\alpha\right)\mathrm{d}^{\times}y
=𝟙2nζ𝐅(1)1ϖ𝐅n2𝒪𝐅××ϖ𝐅n2𝒪𝐅×χ(t1t2)|t1t2|12χ0(t1t21)ψ(t1+t2)d×t1d×t2\displaystyle\displaystyle=\mathbbm{1}_{2\mid n}\cdot\zeta_{\mathbf{F}}(1)^{-1}\int_{\varpi_{\mathbf{F}}^{-\frac{n}{2}}\mathcal{O}_{\mathbf{F}}^{\times}\times\varpi_{\mathbf{F}}^{-\frac{n}{2}}\mathcal{O}_{\mathbf{F}}^{\times}}\chi(t_{1}t_{2})\lvert t_{1}t_{2}\rvert^{\frac{1}{2}}\cdot\chi_{0}(t_{1}t_{2}^{-1})\psi(t_{1}+t_{2})\mathrm{d}^{\times}t_{1}\mathrm{d}^{\times}t_{2}
=𝟙2nζ𝐅(1)qn2(ϖ𝐅n2𝒪𝐅×ψ(t)χχ0(t)dt|t|)(ϖ𝐅n2𝒪𝐅×ψ(t)χχ01(t)dt|t|),\displaystyle\displaystyle=\mathbbm{1}_{2\mid n}\cdot\zeta_{\mathbf{F}}(1)q^{\frac{n}{2}}\left(\int_{\varpi_{\mathbf{F}}^{-\frac{n}{2}}\mathcal{O}_{\mathbf{F}}^{\times}}\psi(t)\cdot\chi\chi_{0}(t)\frac{\mathrm{d}t}{\lvert t\rvert}\right)\left(\int_{\varpi_{\mathbf{F}}^{-\frac{n}{2}}\mathcal{O}_{\mathbf{F}}^{\times}}\psi(t)\cdot\chi\chi_{0}^{-1}(t)\frac{\mathrm{d}t}{\lvert t\rvert}\right),

where we have taken into account that for each τ\displaystyle\tau the map

ϖ𝐅n2𝒪𝐅××𝒪𝐅×ϖ𝐅n2𝒪𝐅××ϖ𝐅n2𝒪𝐅×,(y,α)(τyα,yα1)\displaystyle\varpi_{\mathbf{F}}^{-\frac{n}{2}}\mathcal{O}_{\mathbf{F}}^{\times}\times\mathcal{O}_{\mathbf{F}}^{\times}\to\varpi_{\mathbf{F}}^{-\frac{n}{2}}\mathcal{O}_{\mathbf{F}}^{\times}\times\varpi_{\mathbf{F}}^{-\frac{n}{2}}\mathcal{O}_{\mathbf{F}}^{\times},\quad(y,\alpha)\mapsto(\tau y\alpha,y\alpha^{-1})

is 2\displaystyle 2-to-1\displaystyle 1. The stated formulae then follow from the relation between the Gauss integrals and the local epsilon-factors given in [5, Exercise 23.5], and a direct computation in the unramified case.

(B) Similarly in the non-split case we have

ϖ𝐅n𝒪𝐅×H(y)χ(y)d×y\displaystyle\displaystyle\int_{\varpi_{\mathbf{F}}^{-n}\mathcal{O}_{\mathbf{F}}^{\times}}H(y)\chi(y)\mathrm{d}^{\times}y =𝟙2enζ𝐅(1)Vol(𝐋1)ϖ𝐋en2𝒪𝐋×H(Nr(z))χ(Nr(z))dz|z|𝐋\displaystyle\displaystyle=\mathbbm{1}_{2\mid en}\cdot\frac{\zeta_{\mathbf{F}}(1)}{{\rm Vol}(\mathbf{L}^{1})}\int_{\varpi_{\mathbf{L}}^{-\frac{en}{2}}\mathcal{O}_{\mathbf{L}}^{\times}}H({\rm Nr}(z))\chi({\rm Nr}(z))\frac{\mathrm{d}z}{\lvert z\rvert_{\mathbf{L}}}
=𝟙2enζ𝐅(1)qn2λ(𝐋/𝐅,ψ)ϖ𝐋en2𝒪𝐋×β(z)χ(Nr(z))ψ𝐋(z)dz|z|𝐋\displaystyle\displaystyle=\mathbbm{1}_{2\mid en}\cdot\zeta_{\mathbf{F}}(1)q^{\frac{n}{2}}\lambda(\mathbf{L}/\mathbf{F},\psi)\int_{\varpi_{\mathbf{L}}^{-\frac{en}{2}}\mathcal{O}_{\mathbf{L}}^{\times}}\beta(z)\chi({\rm Nr}(z))\cdot\psi_{\mathbf{L}}(z)\frac{\mathrm{d}z}{\lvert z\rvert_{\mathbf{L}}}
=𝟙2enζ𝐅(1)qn2λ(𝐋/𝐅,ψ)𝟙𝔠(β(χNr))+e1=en2ε(1,β1(χ1Nr),ψ𝐋),\displaystyle\displaystyle=\mathbbm{1}_{2\mid en}\cdot\zeta_{\mathbf{F}}(1)q^{\frac{n}{2}}\lambda(\mathbf{L}/\mathbf{F},\psi)\cdot\mathbbm{1}_{\mathfrak{c}\left(\beta\cdot(\chi\circ{\rm Nr})\right)+e-1=\frac{en}{2}}\varepsilon(1,\beta^{-1}\cdot(\chi^{-1}\circ{\rm Nr}),\psi_{\mathbf{L}}),

where we have applied [5, Exercise 23.5] in the last line. Now that [10, Theorem 4.7] implies

λ(𝐋/𝐅,ψ)ε(s,β(χNr),ψ𝐋)=ε(s,πβχ,ψ)=qn(12s)ε(12,πβχ,ψ),\displaystyle\lambda(\mathbf{L}/\mathbf{F},\psi)\varepsilon(s,\beta\cdot(\chi\circ{\rm Nr}),\psi_{\mathbf{L}})=\varepsilon(s,\pi_{\beta}\otimes\chi,\psi)=q^{n\cdot(\frac{1}{2}-s)}\varepsilon(\tfrac{1}{2},\pi_{\beta}\otimes\chi,\psi),

we conclude the desired formula. Note that in this case β(χNr)\displaystyle\beta\cdot(\chi\circ{\rm Nr}) is never unramified. ∎

We give a first asymptotic analysis of H\displaystyle H as follows.

Lemma 3.7.

Let α𝐋1\displaystyle\alpha\in\mathbf{L}^{1} and n11\displaystyle n_{1}\in\mathbb{Z}_{\geqslant 1}. Then αα¯𝒫𝐋n1\displaystyle\alpha-\overline{\alpha}\in\mathcal{P}_{\mathbf{L}}^{n_{1}} if and only if α±1+𝒫𝐋n1\displaystyle\alpha\in\pm 1+\mathcal{P}_{\mathbf{L}}^{n_{1}}.

Proof.

If α±1+𝒫𝐋n1\displaystyle\alpha\in\pm 1+\mathcal{P}_{\mathbf{L}}^{n_{1}}, then it clear that αα¯𝒫𝐋n1\displaystyle\alpha-\overline{\alpha}\in\mathcal{P}_{\mathbf{L}}^{n_{1}}. The converse is justified as follows.

(a) If 𝐋/𝐅\displaystyle\mathbf{L}/\mathbf{F} is split, we write α=(t,t1)\displaystyle\alpha=(t,t^{-1}) with t𝐅×\displaystyle t\in\mathbf{F}^{\times}. Then tt1𝒫𝐅n1\displaystyle t-t^{-1}\in\mathcal{P}_{\mathbf{F}}^{n_{1}}. We must have t𝒪𝐅×\displaystyle t\in\mathcal{O}_{\mathbf{F}}^{\times}, and tt1(mod𝒫𝐅)\displaystyle t\equiv t^{-1}\pmod{\mathcal{P}_{\mathbf{F}}}. Hence t±1+𝒫𝐅\displaystyle t\in\pm 1+\mathcal{P}_{\mathbf{F}}. Writing t=±1+ϖ𝐅ku\displaystyle t=\pm 1+\varpi_{\mathbf{F}}^{k}u for some k1\displaystyle k\in\mathbb{Z}_{\geqslant 1} and u𝒪𝐅×\displaystyle u\in\mathcal{O}_{\mathbf{F}}^{\times} we infer that tt12ϖ𝐅ku+𝒫𝐅k+1𝒫𝐅k𝒫𝐅k+1\displaystyle t-t^{-1}\in 2\varpi_{\mathbf{F}}^{k}u+\mathcal{P}_{\mathbf{F}}^{k+1}\subset\mathcal{P}_{\mathbf{F}}^{k}-\mathcal{P}_{\mathbf{F}}^{k+1}. Hence kn1\displaystyle k\geqslant n_{1}, and t±1+𝒫𝐅n1\displaystyle t\in\pm 1+\mathcal{P}_{\mathbf{F}}^{n-1}.

(b) If 𝐋/𝐅\displaystyle\mathbf{L}/\mathbf{F} is a field, then α¯=α1\displaystyle\overline{\alpha}=\alpha^{-1}. We repeat the above argument with (subscript) 𝐅\displaystyle\mathbf{F} replaced by 𝐋\displaystyle\mathbf{L}. ∎

Lemma 3.8.

There is cβ𝒪𝐅×\displaystyle c_{\beta}\in\mathcal{O}_{\mathbf{F}}^{\times}, called the additive parameter and is uniquely determined up to multiplication by elements in 1+𝒫𝐅n02e\displaystyle 1+\mathcal{P}_{\mathbf{F}}^{\lceil\frac{n_{0}}{2e}\rceil}, so that for all u𝒫𝐋n02\displaystyle u\in\mathcal{P}_{\mathbf{L}}^{\lfloor\frac{n_{0}}{2}\rfloor}

β(1+u)={ψ(ϖ𝐅n0cβ(uu¯))if 𝐋/𝐅 split and 2n0,ψ(ϖ𝐅n0cβ((uu¯)21(u2u¯2)))if 𝐋/𝐅 split and 2n0;ψ(ϖ𝐅n0εcβ(uu¯))if 𝐋/𝐅 unramified and 2n0,ψ(ϖ𝐅n0εcβ((uu¯)21(u2u¯2)))if 𝐋/𝐅 unramified and 2n0;ψ(ϖ𝐋n01cβ(uu¯))if 𝐋/𝐅 ramified.\displaystyle\beta(1+u)=\begin{cases}\psi\left(\varpi_{\mathbf{F}}^{-n_{0}}c_{\beta}(u-\bar{u})\right)&\text{if }\mathbf{L}/\mathbf{F}\text{ split and }2\mid n_{0},\\ \psi\left(\varpi_{\mathbf{F}}^{-n_{0}}c_{\beta}\left((u-\bar{u})-2^{-1}(u^{2}-\bar{u}^{2})\right)\right)&\text{if }\mathbf{L}/\mathbf{F}\text{ split and }2\nmid n_{0};\\ \psi\left(\varpi_{\mathbf{F}}^{-n_{0}}\sqrt{\varepsilon}c_{\beta}(u-\bar{u})\right)&\text{if }\mathbf{L}/\mathbf{F}\text{ unramified and }2\mid n_{0},\\ \psi\left(\varpi_{\mathbf{F}}^{-n_{0}}\sqrt{\varepsilon}c_{\beta}\left((u-\bar{u})-2^{-1}(u^{2}-\bar{u}^{2})\right)\right)&\text{if }\mathbf{L}/\mathbf{F}\text{ unramified and }2\nmid n_{0};\\ \psi\left(\varpi_{\mathbf{L}}^{-n_{0}-1}c_{\beta}(u-\bar{u})\right)&\text{if }\mathbf{L}/\mathbf{F}\text{ ramified}.\end{cases}
Proof.

We only treat the cases 2n0:=2m+1\displaystyle 2\nmid n_{0}:=2m+1 if 𝐋/𝐅\displaystyle\mathbf{L}/\mathbf{F} is not ramified, the even ones being simpler. The following map

log𝐅:(1+𝒫𝐅m)/(1+𝒫𝐅2m+1)𝒫𝐅m/𝒫𝐅2m+1,1+xx21x2\displaystyle\log_{\mathbf{F}}:(1+\mathcal{P}_{\mathbf{F}}^{m})/(1+\mathcal{P}_{\mathbf{F}}^{2m+1})\to\mathcal{P}_{\mathbf{F}}^{m}/\mathcal{P}_{\mathbf{F}}^{2m+1},\quad 1+x\mapsto x-2^{-1}x^{2}

is a group isomorphism. If 𝐋/𝐅\displaystyle\mathbf{L}/\mathbf{F} is split there is cβ𝒪𝐅×/(1+𝒫𝐅m+1)\displaystyle c_{\beta}^{\prime}\in\mathcal{O}_{\mathbf{F}}^{\times}/(1+\mathcal{P}_{\mathbf{F}}^{m+1}) such that

(3.6) β(1+ϖ𝐅mu)=ψ(cβ(ϖ𝐅(m+1)u21ϖ𝐅1u2)),u𝒪𝐋(𝒪𝐅×𝒪𝐅).\beta(1+\varpi_{\mathbf{F}}^{m}u)=\psi\left(c_{\beta}^{\prime}\left(\varpi_{\mathbf{F}}^{-(m+1)}u-2^{-1}\varpi_{\mathbf{F}}^{-1}u^{2}\right)\right),\quad\forall u\in\mathcal{O}_{\mathbf{L}}\left(\simeq\mathcal{O}_{\mathbf{F}}\times\mathcal{O}_{\mathbf{F}}\right).

One checks easily that cβ:=cβ\displaystyle c_{\beta}:=c_{\beta}^{\prime} is the required one. If 𝐋/𝐅\displaystyle\mathbf{L}/\mathbf{F} is unramified we get an analogue of (3.6)

(3.7) β(1+ϖ𝐅mu)=ψ(Tr(cβ(ϖ𝐅(m+1)u21ϖ𝐅1u2))),u𝒪𝐋.\beta(1+\varpi_{\mathbf{F}}^{m}u)=\psi\left({\rm Tr}\left(c_{\beta}^{\prime}\left(\varpi_{\mathbf{F}}^{-(m+1)}u-2^{-1}\varpi_{\mathbf{F}}^{-1}u^{2}\right)\right)\right),\quad\forall u\in\mathcal{O}_{\mathbf{L}}.

Now that β\displaystyle\beta is trivial on 𝒪𝐅×\displaystyle\mathcal{O}_{\mathbf{F}}^{\times} and log𝐅\displaystyle\log_{\mathbf{F}} is surjective, implying ψTr(cβ𝒫𝐅m1)=1\displaystyle\psi\circ{\rm Tr}(c_{\beta}^{\prime}\mathcal{P}_{\mathbf{F}}^{-m-1})=1, i.e., cβ+cβ¯𝒫𝐅m+1\displaystyle c_{\beta}^{\prime}+\overline{c_{\beta}^{\prime}}\in\mathcal{P}_{\mathbf{F}}^{m+1}. Therefore we may write cβcβε+𝒫𝐅m+1\displaystyle c_{\beta}^{\prime}\in c_{\beta}\sqrt{\varepsilon}+\mathcal{P}_{\mathbf{F}}^{m+1} for some cβ𝒪𝐅×/(1+𝒫𝐅m+1)\displaystyle c_{\beta}\in\mathcal{O}_{\mathbf{F}}^{\times}/(1+\mathcal{P}_{\mathbf{F}}^{m+1}). If 𝐋/𝐅\displaystyle\mathbf{L}/\mathbf{F} is ramified, then 2n0\displaystyle 2\mid n_{0} by the claim proved in Lemma 3.3 and ψ𝐋:=ψTr𝐋/𝐅\displaystyle\psi_{\mathbf{L}}:=\psi\circ{\rm Tr}_{\mathbf{L}/\mathbf{F}} has conductor exponent 1\displaystyle-1. We similarly get

(3.8) β(1+u)=ψ(ϖ𝐅n02ϖ𝐋1(cβucβ¯u¯)),u𝒫𝐋n02\beta(1+u)=\psi\left(\varpi_{\mathbf{F}}^{-\frac{n_{0}}{2}}\varpi_{\mathbf{L}}^{-1}(c_{\beta}^{\prime}u-\overline{c_{\beta}^{\prime}}\bar{u})\right),\quad\forall u\in\mathcal{P}_{\mathbf{L}}^{\frac{n_{0}}{2}}

for some cβ𝒪𝐋×/(1+𝒫𝐋n02)\displaystyle c_{\beta}^{\prime}\in\mathcal{O}_{\mathbf{L}}^{\times}/(1+\mathcal{P}_{\mathbf{L}}^{\frac{n_{0}}{2}}). Note that β\displaystyle\beta is trivial on 1+𝒫𝐅\displaystyle 1+\mathcal{P}_{\mathbf{F}}, implying ψ(ϖ𝐅n02ϖ𝐋1(cβcβ¯)𝒫𝐅n04)=1\displaystyle\psi\left(\varpi_{\mathbf{F}}^{-\frac{n_{0}}{2}}\varpi_{\mathbf{L}}^{-1}(c_{\beta}^{\prime}-\overline{c_{\beta}^{\prime}})\mathcal{P}_{\mathbf{F}}^{\lceil\frac{n_{0}}{4}\rceil}\right)=1, i.e., cβcβ¯𝒫𝐋2n04+1\displaystyle c_{\beta}^{\prime}-\overline{c_{\beta}^{\prime}}\in\mathcal{P}_{\mathbf{L}}^{2\lfloor\frac{n_{0}}{4}\rfloor+1}. Therefore we may write cβcβ+𝒫𝐋2n04+1\displaystyle c_{\beta}^{\prime}\in c_{\beta}+\mathcal{P}_{\mathbf{L}}^{2\lfloor\frac{n_{0}}{4}\rfloor+1}, hence may take cβ=cβ\displaystyle c_{\beta}^{\prime}=c_{\beta} in (3.8), for some cβ𝒪𝐅×/(1+𝒫𝐅n04)\displaystyle c_{\beta}\in\mathcal{O}_{\mathbf{F}}^{\times}/(1+\mathcal{P}_{\mathbf{F}}^{\lceil\frac{n_{0}}{4}\rceil}). ∎

Remark 3.9.

In the split case the character β\displaystyle\beta is constructed from a character χ0\displaystyle\chi_{0} of 𝐅×\displaystyle\mathbf{F}^{\times}. We also call cβ=cχ0\displaystyle c_{\beta}=c_{\chi_{0}} the additive parameter of χ0\displaystyle\chi_{0}.

Lemma 3.10.

Let e\displaystyle e be the generalized ramification index given in (3.4), and write n0:=𝔠(β)\displaystyle n_{0}:=\mathfrak{c}(\beta).

  • (1)

    In the domain v𝐅(y)4n0/e2(e1)\displaystyle v_{\mathbf{F}}(y)\leqslant-4n_{0}/e-2(e-1) we have H(y)=E2n0/e+e1(y)\displaystyle H(y)=E_{\geqslant 2n_{0}/e+e-1}(y).

  • (2)

    Assume β\displaystyle\beta is regular. For any τ\displaystyle\tau we have H(τy2)=0\displaystyle H(\tau y^{2})=0 if v𝐅(y)(2n0)e11\displaystyle v_{\mathbf{F}}(y)\geqslant(2-n_{0})e^{-1}-1.

  • (3)

    If τ1\displaystyle\tau\neq 1, then H(τy2)=0\displaystyle H(\tau y^{2})=0 unless v𝐅(y)=1ee1n0\displaystyle v_{\mathbf{F}}(y)=1-e-e^{-1}n_{0}.

  • (4)

    Suppose e=1,n02\displaystyle e=1,n_{0}\geqslant 2. If τ=1\displaystyle\tau=1 and v𝐅(y)=m\displaystyle v_{\mathbf{F}}(y)=-m with n0m2n01\displaystyle n_{0}\leqslant m\leqslant 2n_{0}-1, then we may put the following condition in the domain of integration in (3.5):

    Tr(α){±2(1+𝒫𝐅2(n01))if m=2n01±2(1+ϖ𝐅2(mn0)𝒪𝐅×)if n0<m<2n01𝒪𝐅±±2(1+𝒫𝐅)if m=n0.\displaystyle{\rm Tr}(\alpha)\in\begin{cases}\pm 2(1+\mathcal{P}_{\mathbf{F}}^{2(n_{0}-1)})&\text{if }m=2n_{0}-1\\ \pm 2(1+\varpi_{\mathbf{F}}^{2(m-n_{0})}\mathcal{O}_{\mathbf{F}}^{\times})&\text{if }n_{0}<m<2n_{0}-1\\ \mathcal{O}_{\mathbf{F}}-\sideset{}{{}_{\pm}}{\bigcup}\pm 2(1+\mathcal{P}_{\mathbf{F}})&\text{if }m=n_{0}\end{cases}.
  • (5)

    Suppose e=2\displaystyle e=2. We may put the following condition in the domain of integration in (3.5):

    {Tr(α)±2(1+𝒫𝐅n01)if τ=1, and v𝐅(y):=m=n0Tr(α)±2(1+ϖ𝐅2mn0+1𝒪𝐅×)if τ=1, and n02+1v𝐅(y):=mn01.\displaystyle\begin{cases}{\rm Tr}(\alpha)\in\pm 2(1+\mathcal{P}_{\mathbf{F}}^{n_{0}-1})&\text{if }\tau=1,\text{ and }-v_{\mathbf{F}}(y):=m=n_{0}\\ {\rm Tr}(\alpha)\in\pm 2(1+\varpi_{\mathbf{F}}^{2m-n_{0}+1}\mathcal{O}_{\mathbf{F}}^{\times})&\text{if }\tau=1,\text{ and }\tfrac{n_{0}}{2}+1\leqslant-v_{\mathbf{F}}(y):=m\leqslant n_{0}-1\end{cases}.
Proof.

(1) We only consider non-split 𝐋\displaystyle\mathbf{L}, leaving the simpler split case as an exercise. Let n=v𝐅(y)4n0/e+2(e1)\displaystyle n=-v_{\mathbf{F}}(y)\geqslant 4n_{0}/e+2(e-1). For χ𝒪𝐅×^\displaystyle\chi\in\widehat{\mathcal{O}_{\mathbf{F}}^{\times}}, by Lemma 3.6 the integral ϖ𝐅n𝒪𝐅×H(y)χ(y)d×y0\displaystyle\int_{\varpi_{\mathbf{F}}^{-n}\mathcal{O}_{\mathbf{F}}^{\times}}H(y)\chi(y)\mathrm{d}^{\times}y\neq 0 is non-zero only if

en2=𝔠(β(χNr))+𝔠(ψ𝐋)max(𝔠(β),𝔠(χNr))𝔠(β(χNr))={n1if e=2n/2if e=1.\displaystyle\tfrac{en}{2}=\mathfrak{c}(\beta(\chi\circ{\rm Nr}))+\mathfrak{c}(\psi_{\mathbf{L}})\quad\Leftrightarrow\quad\max(\mathfrak{c}(\beta),\mathfrak{c}(\chi\circ{\rm Nr}))\geqslant\mathfrak{c}(\beta(\chi\circ{\rm Nr}))=\begin{cases}n-1&\text{if }e=2\\ n/2&\text{if }e=1\end{cases}.

By [16, Proposition \@slowromancapv@.2.3 & Corollay \@slowromancapv@.3.3] we have

(3.9) 𝔠(χNr)={max(2𝔠(χ)1,0)if e=2𝔠(χ)if e=1.\mathfrak{c}(\chi\circ{\rm Nr})=\begin{cases}\max(2\mathfrak{c}(\chi)-1,0)&\text{if }e=2\\ \mathfrak{c}(\chi)&\text{if }e=1\end{cases}.

In view of the assumption on n\displaystyle n (implying 𝔠(β(χNr))>n0\displaystyle\mathfrak{c}(\beta(\chi\circ{\rm Nr}))>n_{0}), the non-vanishing condition is equivalent to 2n&𝔠(χ)=n/2\displaystyle 2\mid n\ \&\ \mathfrak{c}(\chi)=n/2. Now that by [10, Theorem 4.7] (χ)=𝔠(χ)1=n/21>2(πβ)=𝔠(πβ)2=2𝔠(β)/e+e3=2n0/e+e3\displaystyle\ell(\chi)=\mathfrak{c}(\chi)-1=n/2-1>2\ell(\pi_{\beta})=\mathfrak{c}(\pi_{\beta})-2=2\mathfrak{c}(\beta)/e+e-3=2n_{0}/e+e-3, we can apply the stability theorem [5, Theorem 25.7] and obtain

(3.10) ϖ𝐅n𝒪𝐅×H(y)χ(y)d×y=𝟙2n,𝔠(χ)=n2ζ𝐅(1)ε(1/2,χ1,ψ)2.\int_{\varpi_{\mathbf{F}}^{-n}\mathcal{O}_{\mathbf{F}}^{\times}}H(y)\chi(y)\mathrm{d}^{\times}y=\mathbbm{1}_{2\mid n,\mathfrak{c}(\chi)=\frac{n}{2}}\cdot\zeta_{\mathbf{F}}(1)\varepsilon(1/2,\chi^{-1},\psi)^{2}.

We conclude by comparing with Lemma 2.3 and applying the Mellin inversion on 𝒪𝐅×\displaystyle\mathcal{O}_{\mathbf{F}}^{\times}.

(2) We average over the change of variables ααδ\displaystyle\alpha\mapsto\alpha\delta for δ𝐋1(1+𝒫𝐋n01)=:U\displaystyle\delta\in\mathbf{L}^{1}\cap(1+\mathcal{P}_{\mathbf{L}}^{n_{0}-1})=:U. Note that

yxτα(δ1)𝒫𝐋1e,Tr(𝒫𝐋1e)𝒫𝐋1e𝐅=𝒪𝐅ψ(Tr(xταδ)y)=ψ(Tr(xτα)y)\displaystyle yx_{\tau}\alpha(\delta-1)\in\mathcal{P}_{\mathbf{L}}^{1-e},\quad{\rm Tr}(\mathcal{P}_{\mathbf{L}}^{1-e})\subset\mathcal{P}_{\mathbf{L}}^{1-e}\cap\mathbf{F}=\mathcal{O}_{\mathbf{F}}\quad\Rightarrow\quad\psi({\rm Tr}(x_{\tau}\alpha\delta)y)=\psi({\rm Tr}(x_{\tau}\alpha)y)

for y\displaystyle y satisfying the stated condition. By Lemma 3.3, the character β\displaystyle\beta is non-trivial on U\displaystyle U, hence

H(τy2)=λ(𝐋/𝐅,ψ)|τ|𝐅12|y|𝐅η𝐋/𝐅(y)𝐋1𝒪𝐋β(xτα)ψ(Tr(xτα)y)(Uβ(δ)dδ)dα=0.\displaystyle H(\tau y^{2})=\lambda(\mathbf{L}/\mathbf{F},\psi)\lvert\tau\rvert_{\mathbf{F}}^{\frac{1}{2}}\cdot\lvert y\rvert_{\mathbf{F}}\cdot\eta_{\mathbf{L}/\mathbf{F}}(y)\cdot\int_{\mathbf{L}^{1}\cap\mathcal{O}_{\mathbf{L}}}\beta(x_{\tau}\alpha)\psi({\rm Tr}(x_{\tau}\alpha)y)\cdot\left(\oint_{U}\beta(\delta)\mathrm{d}\delta\right)\mathrm{d}\alpha=0.

(3) The idea is to average over similar change of variables ααδ\displaystyle\alpha\mapsto\alpha\delta for δ𝐋1(1+𝒫𝐋n)=:Un\displaystyle\delta\in\mathbf{L}^{1}\cap(1+\mathcal{P}_{\mathbf{L}}^{n})=:U_{n} with

(3.11) nmax(n0,(1eev𝐅(y))/2)n\geqslant\max(n_{0},(1-e-ev_{\mathbf{F}}(y))/2)

and perform a further change of variables δ=δ(u)\displaystyle\delta=\delta(u)

δ=1+ϖ𝐋nu1+ϖ𝐋nu¯,u𝒪𝐋δ1ϖ𝐋nuϖ𝐋nu¯+y1𝒫𝐋1e.\displaystyle\delta=\frac{1+\varpi_{\mathbf{L}}^{n}u}{1+\overline{\varpi_{\mathbf{L}}^{n}u}},\ u\in\mathcal{O}_{\mathbf{L}}\quad\Rightarrow\quad\delta-1\in\varpi_{\mathbf{L}}^{n}u-\overline{\varpi_{\mathbf{L}}^{n}u}+y^{-1}\mathcal{P}_{\mathbf{L}}^{1-e}.

Then we have β(δ)=1\displaystyle\beta(\delta)=1, ψ(Tr(xτα(δ1))y)=ψ𝐋(yxτα(ϖ𝐋nuϖ𝐋nu¯))\displaystyle\psi({\rm Tr}(x_{\tau}\alpha(\delta-1))y)=\psi_{\mathbf{L}}(yx_{\tau}\alpha(\varpi_{\mathbf{L}}^{n}u-\overline{\varpi_{\mathbf{L}}^{n}u})) and get

(3.12) H(τy2)=λ(𝐋/𝐅,ψ)|τ|𝐅12|y|𝐅η𝐋/𝐅(y)𝐋1𝒪𝐋β(xτα)ψ(Tr(xτα)y)[𝒪𝐋ψ(y(xταxτα¯)(ϖ𝐋nuϖ𝐋nu¯))du]dα.H(\tau y^{2})=\lambda(\mathbf{L}/\mathbf{F},\psi)\lvert\tau\rvert_{\mathbf{F}}^{\frac{1}{2}}\cdot\lvert y\rvert_{\mathbf{F}}\cdot\eta_{\mathbf{L}/\mathbf{F}}(y)\cdot\int_{\mathbf{L}^{1}\cap\mathcal{O}_{\mathbf{L}}}\beta(x_{\tau}\alpha)\psi({\rm Tr}(x_{\tau}\alpha)y)\cdot\\ \left[\oint_{\mathcal{O}_{\mathbf{L}}}\psi\left(y(x_{\tau}\alpha-\overline{x_{\tau}\alpha})(\varpi_{\mathbf{L}}^{n}u-\overline{\varpi_{\mathbf{L}}^{n}u})\right)\mathrm{d}u\right]\mathrm{d}\alpha.

For any m\displaystyle m\in\mathbb{Z} we introduce 𝒫𝐋m,:={x𝒫𝐋m|x¯=x}\displaystyle\mathcal{P}_{\mathbf{L}}^{m,-}:=\left\{x\in\mathcal{P}_{\mathbf{L}}^{m}\ \middle|\ \bar{x}=-x\right\}. We see that

𝒪𝐋𝒫𝐋n,,uϖ𝐋nuϖ𝐋nu¯\displaystyle\mathcal{O}_{\mathbf{L}}\to\mathcal{P}_{\mathbf{L}}^{n,-},\quad u\mapsto\varpi_{\mathbf{L}}^{n}u-\overline{\varpi_{\mathbf{L}}^{n}u}

is a surjective group homomorphism. Therefore the inner integral in (3.12) is non-vanishing only if

(3.13) y(xταxτα¯)𝒫𝐋1en,v𝐋(xταxτα¯)1enev𝐅(y).y(x_{\tau}\alpha-\overline{x_{\tau}\alpha})\in\mathcal{P}_{\mathbf{L}}^{1-e-n,-}\quad\Leftrightarrow\quad v_{\mathbf{L}}(x_{\tau}\alpha-\overline{x_{\tau}\alpha})\geqslant 1-e-n-ev_{\mathbf{F}}(y).

We choose n=12eev𝐅(y)\displaystyle n=1-2e-ev_{\mathbf{F}}(y), which is consistent with the condition (3.11) if

(3.14) v𝐅(y)min(e12e1n0,e13)=ee1n0,v_{\mathbf{F}}(y)\leqslant\lfloor\min(e^{-1}-2-e^{-1}n_{0},e^{-1}-3)\rfloor=-e-e^{-1}n_{0},

so that (3.13) becomes v𝐋(xταxτα¯)e\displaystyle v_{\mathbf{L}}(x_{\tau}\alpha-\overline{x_{\tau}\alpha})\geqslant e or v𝐅((xταxτα¯)2)2\displaystyle v_{\mathbf{F}}((x_{\tau}\alpha-\overline{x_{\tau}\alpha})^{2})\geqslant 2. In view of the equality

(3.15) τ=xταxτα¯=14{(xτα+xτα¯)2(xταxτα¯)2}\tau=x_{\tau}\alpha\cdot\overline{x_{\tau}\alpha}=\tfrac{1}{4}\left\{(x_{\tau}\alpha+\overline{x_{\tau}\alpha})^{2}-(x_{\tau}\alpha-\overline{x_{\tau}\alpha})^{2}\right\}

and v𝐅(τ){0,1}\displaystyle v_{\mathbf{F}}(\tau)\in\{0,1\}, we deduce that v𝐅(τ)=v𝐅((xτα+xτα¯)2)2\displaystyle v_{\mathbf{F}}(\tau)=v_{\mathbf{F}}((x_{\tau}\alpha+\overline{x_{\tau}\alpha})^{2})\in 2\mathbb{Z}. Hence v𝐅(τ)=0\displaystyle v_{\mathbf{F}}(\tau)=0 and τ\displaystyle\tau is a square in 𝒪𝐅×\displaystyle\mathcal{O}_{\mathbf{F}}^{\times}. This is possible only if τ=1\displaystyle\tau=1. In other words, if τ1\displaystyle\tau\neq 1 and v𝐅(y)\displaystyle v_{\mathbf{F}}(y) satisfies (3.14) then H(τy2)=0\displaystyle H(\tau y^{2})=0. We conclude because the only integer not satisfying (3.14) nor the inequality in (1) is v𝐅(y)=1ee1n0\displaystyle v_{\mathbf{F}}(y)=1-e-e^{-1}n_{0}.

(4) Suppose 𝐋/𝐅\displaystyle\mathbf{L}/\mathbf{F} is unramified. Let n:=m2\displaystyle n:=\lceil\tfrac{m}{2}\rceil. Recall cβ\displaystyle c_{\beta} defined in Lemma 3.8 (1). We average over the change of variables αα(1+u)(1+u¯)1\displaystyle\alpha\mapsto\alpha(1+u)(1+\bar{u})^{-1} for u𝒫𝐋n\displaystyle u\in\mathcal{P}_{\mathbf{L}}^{n} and get

𝐋1𝒪𝐋β(α)ψ(yTr(α))dα=𝐋1𝒪𝐋β(α)ψ(yTr(α)){𝒫𝐋nψ(2cβε(uu¯)ϖ𝐅n0+yTr(αuu¯1+u¯))du}dα=𝐋1𝒪𝐋β(α)ψ(yTr(α)){𝒪𝐋ψ((2εcβϖ𝐅nn0+yϖ𝐅n(αα¯))(uu¯))du}dα.\int_{\mathbf{L}^{1}\cap\mathcal{O}_{\mathbf{L}}}\beta(\alpha)\psi(y{\rm Tr}(\alpha))\mathrm{d}\alpha=\int_{\mathbf{L}^{1}\cap\mathcal{O}_{\mathbf{L}}}\beta(\alpha)\psi(y{\rm Tr}(\alpha))\left\{\oint_{\mathcal{P}_{\mathbf{L}}^{n}}\psi\left(\tfrac{2c_{\beta}\sqrt{\varepsilon}(u-\bar{u})}{\varpi_{\mathbf{F}}^{n_{0}}}+y{\rm Tr}\left(\alpha\cdot\tfrac{u-\bar{u}}{1+\bar{u}}\right)\right)\mathrm{d}u\right\}\mathrm{d}\alpha\\ =\int_{\mathbf{L}^{1}\cap\mathcal{O}_{\mathbf{L}}}\beta(\alpha)\psi(y{\rm Tr}(\alpha))\left\{\oint_{\mathcal{O}_{\mathbf{L}}}\psi\left((2\sqrt{\varepsilon}c_{\beta}\varpi_{\mathbf{F}}^{n-n_{0}}+y\varpi_{\mathbf{F}}^{n}(\alpha-\bar{\alpha}))(u-\bar{u})\right)\mathrm{d}u\right\}\mathrm{d}\alpha.

The non-vanishing of the inner integral implies

(2εcβϖ𝐅nn0+yϖ𝐅n(αα¯))ε𝒪𝐅(αα¯)ε{ϖ𝐅mn0𝒪𝐅×if m<2n01𝒫𝐅n01if m=2n01(α+α¯)2{4+𝒫𝐅2(n01)if m=2n014+ϖ𝐅2(mn0)𝒪𝐅×if n0<m<2n01𝒪𝐅4(1+𝒫𝐅)if m=n0\left(2\sqrt{\varepsilon}c_{\beta}\varpi_{\mathbf{F}}^{n-n_{0}}+y\varpi_{\mathbf{F}}^{n}(\alpha-\bar{\alpha})\right)\sqrt{\varepsilon}\in\mathcal{O}_{\mathbf{F}}\ \Rightarrow\\ (\alpha-\bar{\alpha})\sqrt{\varepsilon}\in\begin{cases}\varpi_{\mathbf{F}}^{m-n_{0}}\mathcal{O}_{\mathbf{F}}^{\times}&\text{if }m<2n_{0}-1\\ \mathcal{P}_{\mathbf{F}}^{n_{0}-1}&\text{if }m=2n_{0}-1\end{cases}\Rightarrow\ (\alpha+\bar{\alpha})^{2}\in\begin{cases}4+\mathcal{P}_{\mathbf{F}}^{2(n_{0}-1)}&\text{if }m=2n_{0}-1\\ 4+\varpi_{\mathbf{F}}^{2(m-n_{0})}\mathcal{O}_{\mathbf{F}}^{\times}&\text{if }n_{0}<m<2n_{0}-1\\ \mathcal{O}_{\mathbf{F}}-4(1+\mathcal{P}_{\mathbf{F}})&\text{if }m=n_{0}\end{cases}

from which we conclude. Replacing ε\displaystyle\sqrt{\varepsilon} by 1\displaystyle 1 we obtain the proof in the split case.

(5) In the case τ=1\displaystyle\tau=1 and n02+1mn01\displaystyle\tfrac{n_{0}}{2}+1\leqslant m\leqslant n_{0}-1, we average over the change of variables αα(1+u)(1+u¯)1\displaystyle\alpha\mapsto\alpha(1+u)(1+\bar{u})^{-1} for u𝒫𝐋m\displaystyle u\in\mathcal{P}_{\mathbf{L}}^{m}, noting that αuu¯1+u¯α(uu¯)+y1𝒪𝐋\displaystyle\alpha\tfrac{u-\bar{u}}{1+\bar{u}}\in\alpha(u-\bar{u})+y^{-1}\mathcal{O}_{\mathbf{L}}, and get

𝐋1β(α)ψ(yTr(α))dα=𝐋1β(α)ψ(yTr(α)){𝒫𝐋mψ(2cβ(uu¯)ϖ𝐋n0+1+yTr(αuu¯1+u¯))du}dα=𝐋1β(α)ψ(yTr(α)){𝒫𝐋mψ((2cβϖ𝐋n0+1+(αα¯)y)(uu¯))du}dα=𝐋1β(α)ψ(yTr(α)){𝒫𝐅m2ψ((2cβϖ𝐅n0/2+ϖ𝐋(αα¯)y)u)du}dα.\int_{\mathbf{L}^{1}}\beta(\alpha)\psi(y{\rm Tr}(\alpha))\mathrm{d}\alpha=\int_{\mathbf{L}^{1}}\beta(\alpha)\psi(y{\rm Tr}(\alpha))\left\{\oint_{\mathcal{P}_{\mathbf{L}}^{m}}\psi\left(\tfrac{2c_{\beta}(u-\bar{u})}{\varpi_{\mathbf{L}}^{n_{0}+1}}+y{\rm Tr}\left(\alpha\cdot\tfrac{u-\bar{u}}{1+\bar{u}}\right)\right)\mathrm{d}u\right\}\mathrm{d}\alpha\\ =\int_{\mathbf{L}^{1}}\beta(\alpha)\psi(y{\rm Tr}(\alpha))\left\{\oint_{\mathcal{P}_{\mathbf{L}}^{m}}\psi\left(\left(\tfrac{2c_{\beta}}{\varpi_{\mathbf{L}}^{n_{0}+1}}+(\alpha-\bar{\alpha})y\right)(u-\bar{u})\right)\mathrm{d}u\right\}\mathrm{d}\alpha\\ =\int_{\mathbf{L}^{1}}\beta(\alpha)\psi(y{\rm Tr}(\alpha))\left\{\oint_{\mathcal{P}_{\mathbf{F}}^{\lfloor\frac{m}{2}\rfloor}}\psi\left(\left(\tfrac{2c_{\beta}}{\varpi_{\mathbf{F}}^{n_{0}/2}}+\varpi_{\mathbf{L}}(\alpha-\bar{\alpha})y\right)u\right)\mathrm{d}u\right\}\mathrm{d}\alpha.

The non-vanishing of the inner integral implies

2cβϖ𝐅n02+ϖ𝐋(αα¯)y𝒫𝐅m2{(αα¯)ϖ𝐋1ϖ𝐅mn021𝒪𝐅×if m<n0(αα¯)ϖ𝐋1ϖ𝐅n021𝒪𝐅if m=n0.\displaystyle 2c_{\beta}\varpi_{\mathbf{F}}^{-\frac{n_{0}}{2}}+\varpi_{\mathbf{L}}(\alpha-\bar{\alpha})y\in\mathcal{P}_{\mathbf{F}}^{-\lfloor\frac{m}{2}\rfloor}\ \Rightarrow\ \begin{cases}(\alpha-\bar{\alpha})\varpi_{\mathbf{L}}^{-1}\in\varpi_{\mathbf{F}}^{m-\frac{n_{0}}{2}-1}\mathcal{O}_{\mathbf{F}}^{\times}&\text{if }m<n_{0}\\ (\alpha-\bar{\alpha})\varpi_{\mathbf{L}}^{-1}\in\varpi_{\mathbf{F}}^{\frac{n_{0}}{2}-1}\mathcal{O}_{\mathbf{F}}&\text{if }m=n_{0}\end{cases}.

We conclude by noting that for any m0\displaystyle m\in\mathbb{Z}_{\geqslant 0} we have

(αα¯)ϖ𝐋1𝒫𝐅mTr(α)±2(1+𝒫𝐅2m+1),\displaystyle(\alpha-\bar{\alpha})\varpi_{\mathbf{L}}^{-1}\in\mathcal{P}_{\mathbf{F}}^{m}\ \Leftrightarrow\ {\rm Tr}(\alpha)\in\pm 2(1+\mathcal{P}_{\mathbf{F}}^{2m+1}),

and that Tr(α)24=(αα¯)2\displaystyle{\rm Tr}(\alpha)^{2}-4=(\alpha-\bar{\alpha})^{2} never has even valuation. ∎

Remark 3.11.

If β\displaystyle\beta is not regular, which can happen only if 𝐋/𝐅\displaystyle\mathbf{L}/\mathbf{F} is split with 𝔠(χ02)=0\displaystyle\mathfrak{c}(\chi_{0}^{2})=0, then (2) fails. In fact, it is easy to see that H(τy2)\displaystyle H(\tau y^{2}) is proportional to χ0(τ)|y|\displaystyle\chi_{0}(\tau)\lvert y\rvert as |y|0\displaystyle\lvert y\rvert\to 0. In this case we modify the definition of H\displaystyle H by truncating it so that (2) still holds.

Corollary 3.12.

The test function H\displaystyle H defined by (3.5) and Remark 3.11 is a Bessel orbital integral.

Proof.

This is a direct consequence of Lemma 3.10 (1) & (2) and Proposition 2.1. ∎

3.2. Weight Functions

To every target representation π0\displaystyle\pi_{0} we have associated a parameter (𝐋,β)\displaystyle(\mathbf{L},\beta) and a test function H\displaystyle H in (3.5).

Lemma 3.13.

(1) In the split case π0=π(χ0,χ01)\displaystyle\pi_{0}=\pi(\chi_{0},\chi_{0}^{-1}) consider the function ϕ0Cc(PGL2(𝐅))\displaystyle\phi_{0}\in{\rm C}_{c}^{\infty}({\rm PGL}_{2}(\mathbf{F})) given by

ϕ0(x1x2x3x4)=χ0(x4x1)𝟙𝐙𝐊0[𝒫𝐅n0](x1x2x3x4).\displaystyle\phi_{0}\begin{pmatrix}x_{1}&x_{2}\\ x_{3}&x_{4}\end{pmatrix}=\chi_{0}\left(\frac{x_{4}}{x_{1}}\right)\mathbbm{1}_{\mathbf{Z}\mathbf{K}_{0}[\mathcal{P}_{\mathbf{F}}^{n_{0}}]}\begin{pmatrix}x_{1}&x_{2}\\ x_{3}&x_{4}\end{pmatrix}.

The partial integral I0\displaystyle I_{0} defined by

I0(g):=𝐅ϕ0(g(1x1))ψ(x)dx\displaystyle I_{0}(g):=\int_{\mathbf{F}}\phi_{0}\left(g\begin{pmatrix}1&x\\ &1\end{pmatrix}\right)\psi(-x)\mathrm{d}x

has support contained in 𝐙𝐊0[𝒫𝐅n0]𝐍(𝐅)\displaystyle\mathbf{Z}\mathbf{K}_{0}[\mathcal{P}_{\mathbf{F}}^{n_{0}}]\mathbf{N}(\mathbf{F}), and satisfies

I0(κn(x))=ϕ0(κ)ψ(x),κ𝐙𝐊𝟎,x𝐅.\displaystyle I_{0}(\kappa n(x))=\phi_{0}(\kappa)\psi(x),\quad\forall\kappa\in\mathbf{Z}\mathbf{K_{0}},x\in\mathbf{F}.

(2) We can identify H\displaystyle H as a Bessel orbital integral

H(y)=𝐅2ϕ0((1x11)(y1)(1x21))ψ(x1x2)dx1dx2.\displaystyle H(y)=\int_{\mathbf{F}^{2}}\phi_{0}\left(\begin{pmatrix}1&x_{1}\\ &1\end{pmatrix}\begin{pmatrix}&-y\\ 1&\end{pmatrix}\begin{pmatrix}1&x_{2}\\ &1\end{pmatrix}\right)\psi(-x_{1}-x_{2})\mathrm{d}x_{1}\mathrm{d}x_{2}.

(3) The weight function h(π)\displaystyle h(\pi) is non-negative. We have h(π)0\displaystyle h(\pi)\neq 0 only if 𝔠(πχ01)n0\displaystyle\mathfrak{c}(\pi\otimes\chi_{0}^{-1})\leqslant n_{0}, in which case we have a lower bound h(π)qn0\displaystyle h(\pi)\gg q^{-n_{0}}.

Proof.

(1) Note that ϕ0\displaystyle\phi_{0} is a character upon restriction to 𝐙𝐊0[𝒫𝐅n0]\displaystyle\mathbf{Z}\mathbf{K}_{0}[\mathcal{P}_{\mathbf{F}}^{n_{0}}]. Hence I0(g)\displaystyle I_{0}(g) satisfies

I0(κgn(x))=ϕ0(κ)ψ(x)I0(g),κ𝐙𝐊𝟎,x𝐅.\displaystyle I_{0}(\kappa gn(x))=\phi_{0}(\kappa)\psi(x)I_{0}(g),\quad\forall\kappa\in\mathbf{Z}\mathbf{K_{0}},x\in\mathbf{F}.

From the Cartan decomposition n0𝐙𝐊(ϖ𝐅n1)𝐍(𝐅)\displaystyle\sideset{}{{}_{n\in\mathbb{Z}_{\geqslant 0}}}{\bigsqcup}\mathbf{Z}\mathbf{K}\begin{pmatrix}\varpi_{\mathbf{F}}^{n}&\\ &1\end{pmatrix}\mathbf{N}(\mathbf{F}) we deduce supp(I0)𝐙𝐊𝐍(𝐅)\displaystyle\mathrm{supp}(I_{0})\subset\mathbf{Z}\mathbf{K}\mathbf{N}(\mathbf{F}). We have

𝐙𝐊𝐍(𝐅)=𝐙𝐊0[𝒫𝐅n0](11)𝐍(𝐅)u𝒫𝐅/𝒫𝐅n0𝐙𝐊0[𝒫𝐅n0](1u1)𝐍(𝐅)\displaystyle\mathbf{Z}\mathbf{K}\mathbf{N}(\mathbf{F})=\mathbf{Z}\mathbf{K}_{0}[\mathcal{P}_{\mathbf{F}}^{n_{0}}]\begin{pmatrix}&1\\ 1&\end{pmatrix}\mathbf{N}(\mathbf{F})\sqcup\sideset{}{{}_{u\in\mathcal{P}_{\mathbf{F}}/\mathcal{P}_{\mathbf{F}}^{n_{0}}}}{\bigcup}\mathbf{Z}\mathbf{K}_{0}[\mathcal{P}_{\mathbf{F}}^{n_{0}}]\begin{pmatrix}1&\\ u&1\end{pmatrix}\mathbf{N}(\mathbf{F})

by the Bruhat decomposition over the residue field of 𝐅\displaystyle\mathbf{F}. We easily verify for u𝒫𝐅n0\displaystyle u\notin\mathcal{P}_{\mathbf{F}}^{n_{0}}

(11)𝐍(𝐅)𝐙𝐊0[𝒫𝐅n0]=,(1u1)𝐍(𝐅)𝐙𝐊0[𝒫𝐅n0]=.\displaystyle\begin{pmatrix}&1\\ 1&\end{pmatrix}\mathbf{N}(\mathbf{F})\cap\mathbf{Z}\mathbf{K}_{0}[\mathcal{P}_{\mathbf{F}}^{n_{0}}]=\emptyset,\quad\begin{pmatrix}1&\\ u&1\end{pmatrix}\mathbf{N}(\mathbf{F})\cap\mathbf{Z}\mathbf{K}_{0}[\mathcal{P}_{\mathbf{F}}^{n_{0}}]=\emptyset.

Hence we deduce supp(I0)𝐙𝐊0[𝒫𝐅n0]𝐍(𝐅)\displaystyle\mathrm{supp}(I_{0})\subset\mathbf{Z}\mathbf{K}_{0}[\mathcal{P}_{\mathbf{F}}^{n_{0}}]\mathbf{N}(\mathbf{F}) and conclude by I0(𝟙)=Vol(𝒪𝐅)=1\displaystyle I_{0}(\mathbbm{1})={\rm Vol}(\mathcal{O}_{\mathbf{F}})=1.

(2) Let ϕι(g):=ϕ(g1)¯\displaystyle\phi^{\iota}(g):=\overline{\phi(g^{-1})} for all ϕCc(PGL2(𝐅))\displaystyle\phi\in{\rm C}_{c}^{\infty}({\rm PGL}_{2}(\mathbf{F})). We have ϕ0ι=ϕ0\displaystyle\phi_{0}^{\iota}=\phi_{0} and ϕ0ιϕ0=Vol(𝐙𝐊0[𝒫𝐅n0])ϕ0\displaystyle\phi_{0}^{\iota}*\phi_{0}={\rm Vol}(\mathbf{Z}\mathbf{K}_{0}[\mathcal{P}_{\mathbf{F}}^{n_{0}}])\phi_{0} since ϕ0\displaystyle\phi_{0} is a unitary character upon restriction to 𝐙𝐊0[𝒫𝐅n0]\displaystyle\mathbf{Z}\mathbf{K}_{0}[\mathcal{P}_{\mathbf{F}}^{n_{0}}]. Here the convolution is taken over the group G:=PGL2(𝐅)\displaystyle G:={\rm PGL}_{2}(\mathbf{F}). We can rewrite the relevant Bessel orbital integral as

Vol(𝐙𝐊0[𝒫𝐅n0])1𝐅2Gϕ0(g)¯ϕ0(g(1x11)(y1)(1x21))ψ(x1x2)dgdx1dx2=Vol(𝐙𝐊0[𝒫𝐅n0])1G{𝐅ϕ0(g(1x11))¯ψ(x1)dx1}{𝐅ϕ0(g(y1)(1x21))ψ(x2)dx2}dg=Vol(𝐙𝐊0[𝒫𝐅n0])1GI0(g)¯I0(g(y1))dg=𝐅ψ(u)I0((1u1)(y1))du.{\rm Vol}(\mathbf{Z}\mathbf{K}_{0}[\mathcal{P}_{\mathbf{F}}^{n_{0}}])^{-1}\int_{\mathbf{F}^{2}}\int_{G}\overline{\phi_{0}(g)}\phi_{0}\left(g\begin{pmatrix}1&x_{1}\\ &1\end{pmatrix}\begin{pmatrix}&-y\\ 1&\end{pmatrix}\begin{pmatrix}1&x_{2}\\ &1\end{pmatrix}\right)\psi(-x_{1}-x_{2})\mathrm{d}g\mathrm{d}x_{1}\mathrm{d}x_{2}\\ ={\rm Vol}(\mathbf{Z}\mathbf{K}_{0}[\mathcal{P}_{\mathbf{F}}^{n_{0}}])^{-1}\int_{G}\\ \left\{\int_{\mathbf{F}}\overline{\phi_{0}\left(g\begin{pmatrix}1&-x_{1}\\ &1\end{pmatrix}\right)}\psi(-x_{1})\mathrm{d}x_{1}\right\}\cdot\left\{\int_{\mathbf{F}}\phi_{0}\left(g\begin{pmatrix}&-y\\ 1&\end{pmatrix}\begin{pmatrix}1&x_{2}\\ &1\end{pmatrix}\right)\psi(-x_{2})\mathrm{d}x_{2}\right\}\mathrm{d}g\\ ={\rm Vol}(\mathbf{Z}\mathbf{K}_{0}[\mathcal{P}_{\mathbf{F}}^{n_{0}}])^{-1}\int_{G}\overline{I_{0}(g)}I_{0}\left(g\begin{pmatrix}&-y\\ 1&\end{pmatrix}\right)\mathrm{d}g=\int_{\mathbf{F}}\psi(-u)I_{0}\left(\begin{pmatrix}1&u\\ &1\end{pmatrix}\begin{pmatrix}&-y\\ 1&\end{pmatrix}\right)\mathrm{d}u.

Considering the support of I0\displaystyle I_{0}, the above integral is non-vanishing only if for some u𝐅\displaystyle u^{\prime}\in\mathbf{F}

(1u1)(y1)(1u1)=(uuuy1u)𝐙𝐊0[𝒫𝐅n0],\displaystyle\begin{pmatrix}1&u\\ &1\end{pmatrix}\begin{pmatrix}&-y\\ 1&\end{pmatrix}\begin{pmatrix}1&u^{\prime}\\ &1\end{pmatrix}=\begin{pmatrix}u&uu^{\prime}-y\\ 1&u^{\prime}\end{pmatrix}\in\mathbf{Z}\mathbf{K}_{0}[\mathcal{P}_{\mathbf{F}}^{n_{0}}],

implying yϖ𝐅2n𝒪𝐅×\displaystyle y\in\varpi_{\mathbf{F}}^{-2n}\mathcal{O}_{\mathbf{F}}^{\times} for some nn0\displaystyle n\geqslant n_{0}, and u=ϖnu1\displaystyle u=\varpi^{-n}u_{1}, u=ϖnu2\displaystyle u^{\prime}=\varpi^{-n}u_{2} for some u1,u2𝒪𝐅×\displaystyle u_{1},u_{2}\in\mathcal{O}_{\mathbf{F}}^{\times} satisfying u1u2ϖ2ny𝒫𝐅n\displaystyle u_{1}u_{2}-\varpi^{2n}y\in\mathcal{P}_{\mathbf{F}}^{n}. Writing y=ϖ𝐅2ny0\displaystyle y=\varpi_{\mathbf{F}}^{-2n}y_{0} for some y0𝒪𝐅×\displaystyle y_{0}\in\mathcal{O}_{\mathbf{F}}^{\times} we may take u2=u11y0\displaystyle u_{2}=u_{1}^{-1}y_{0}. We recognize the Bessel orbital integral as H(y)\displaystyle H(y) by the following equation, which is easy to verify

qn𝒪𝐅×χ0(u12y0)ψ(u1+u11y0ϖ𝐅n)du1=H(y).\displaystyle q^{n}\int_{\mathcal{O}_{\mathbf{F}}^{\times}}\chi_{0}(u_{1}^{-2}y_{0})\psi\left(-\tfrac{u_{1}+u_{1}^{-1}y_{0}}{\varpi_{\mathbf{F}}^{n}}\right)\mathrm{d}u_{1}=H(y).

(3) Up to notation this is due to Nelson [13, Theorem 3.1]. Our former work [1, Lemma 4.1] contains more details in a similar situation. ∎

Lemma 3.14.

Let π\displaystyle\pi be a supercuspidal representation of GL2(𝐅)\displaystyle{\rm GL}_{2}(\mathbf{F}). Let C(g)\displaystyle C(g) be a matrix coefficient for smooth vectors in π\displaystyle\pi. Take a non-trivial additive character ψ\displaystyle\psi of 𝐅\displaystyle\mathbf{F}. Then the (relative) orbital integral

I(g):=𝐅2C(n(u1)gn(u2))ψ(u1u2)𝑑u1𝑑u2\displaystyle I(g):=\int_{\mathbf{F}^{2}}C(n(u_{1})gn(u_{2}))\psi(-u_{1}-u_{2})du_{1}du_{2}

is equal to the ψ\displaystyle\psi-Bessel function of π\displaystyle\pi up to a constant.

Proof.

Assume C(g)=π(g)v1,v2\displaystyle C(g)=\langle\pi(g)v_{1},v_{2}\rangle for smooth vectors v1\displaystyle v_{1} and v2\displaystyle v_{2}. One checks that the functional

Vπ,v𝐅π(n(u1)).v,v2ψ(u1)du1\displaystyle V_{\pi}^{\infty}\to\mathbb{C},\quad v\mapsto\int_{\mathbf{F}}\langle\pi(n(u_{1})).v,v_{2}\rangle\psi(-u_{1})du_{1}

is a ψ\displaystyle\psi-Whittaker functional. Hence the following function

Wv1(g):=𝐅C(n(u1)g)ψ(u1)𝑑u1\displaystyle W_{v_{1}}(g):=\int_{\mathbf{F}}C(n(u_{1})g)\psi(-u_{1})du_{1}

is the/a Whittaker function of v1\displaystyle v_{1}. We then apply [17, Lemma 4.1] to get

I(g)=jπ,ψ(g)Wv1(𝟙),g𝐁(𝐅)w2𝐁(𝐅).\displaystyle I(g)=\mathrm{j}_{\pi,\psi}(g)\cdot W_{v_{1}}(\mathbbm{1}),\quad\forall g\in\mathbf{B}(\mathbf{F})w_{2}\mathbf{B}(\mathbf{F}).

The integral converges absolutely because C\displaystyle C is smooth of compact support modulo the center. ∎

Lemma 3.15.

In the dihedral case π0=πβ\displaystyle\pi_{0}=\pi_{\beta} the weight function h(π)0\displaystyle h(\pi)\neq 0 is non-vanishing only if ππβ\displaystyle\pi\simeq\pi_{\beta}, in which case it is positive and satisfies a lower bound h(πβ)qe1n0+1e\displaystyle h(\pi_{\beta})\gg q^{-e^{-1}n_{0}+1-e}.

Proof.

Up to some constant approximately equal to 1\displaystyle 1 we recognize H(y)\displaystyle H(y) as jπβ,ψ(y1)\displaystyle\mathrm{j}_{\pi_{\beta},\psi}\begin{pmatrix}&-y\\ 1&\end{pmatrix} by [3, Theorem 1.1]. By Lemma 3.14 there exists some ϕ0Cc(GL2(𝐅))\displaystyle\phi_{0}\in{\rm C}_{c}^{\infty}({\rm GL}_{2}(\mathbf{F})), whose integral along the center ϕ1Cc(PGL2(𝐅))\displaystyle\phi_{1}\in{\rm C}_{c}^{\infty}({\rm PGL}_{2}(\mathbf{F})) is a matrix coefficient of πβ\displaystyle\pi_{\beta}, so that H\displaystyle H is the Bessel orbital integral of ϕ0\displaystyle\phi_{0} as given in Lemma 3.13 (1). Now that π(ϕ1)0\displaystyle\pi(\phi_{1})\neq 0 only if ππβ\displaystyle\pi\simeq\pi_{\beta} by Schur’s lemma, it remains to show that h(πβ)\displaystyle h(\pi_{\beta}) is positive with the stated lower bound. By (1.5) the value h(πβ)\displaystyle h(\pi_{\beta}) is the square of the L2\displaystyle{\rm L}^{2}-norm of H(y)\displaystyle H(y) up to a constant essentially equal to 1\displaystyle 1. Taking Lemma 3.10 (2) into account we have

𝐅×|H(y)|2d×y|y|𝐅\displaystyle\displaystyle\int_{\mathbf{F}^{\times}}\left\lvert H(y)\right\rvert^{2}\tfrac{\mathrm{d}^{\times}y}{\lvert y\rvert_{\mathbf{F}}} =12τ𝐅×|H(τy2)|2d×y|τy2|𝐅\displaystyle\displaystyle=\frac{1}{2}\sum_{\tau}\int_{\mathbf{F}^{\times}}\left\lvert H(\tau y^{2})\right\rvert^{2}\tfrac{\mathrm{d}^{\times}y}{\lvert\tau y^{2}\rvert_{\mathbf{F}}}
=12τ𝐅×(𝐋1×𝐋1β(xτα1)β(xτα2)¯ψ𝐋(xτ(α1α2)y)dα1dα2)d×y\displaystyle\displaystyle=\frac{1}{2}\sum_{\tau}\int_{\mathbf{F}^{\times}}\left(\int_{\mathbf{L}^{1}\times\mathbf{L}^{1}}\beta(x_{\tau}\alpha_{1})\overline{\beta(x_{\tau}\alpha_{2})}\psi_{\mathbf{L}}(x_{\tau}(\alpha_{1}-\alpha_{2})y)\mathrm{d}\alpha_{1}\mathrm{d}\alpha_{2}\right)\mathrm{d}^{\times}y
=12n(n02)e1+2𝐋1β(α)(τ𝐋1ϖ𝐅n𝒪𝐅×ψ𝐋(xτα2y(α1))d×ydα2)dα,\displaystyle\displaystyle=\frac{1}{2}\sum_{n\geqslant(n_{0}-2)e^{-1}+2}\int_{\mathbf{L}^{1}}\beta(\alpha)\left(\sum_{\tau}\int_{\mathbf{L}^{1}}\int_{\varpi_{\mathbf{F}}^{-n}\mathcal{O}_{\mathbf{F}}^{\times}}\psi_{\mathbf{L}}(x_{\tau}\alpha_{2}y(\alpha-1))\mathrm{d}^{\times}y\mathrm{d}\alpha_{2}\right)\mathrm{d}\alpha,

where we applied the change of variables α1αα2\displaystyle\alpha_{1}\mapsto\alpha\alpha_{2} in the last line. By measure relation (3.3) we get

12τ𝐋1ϖ𝐅n𝒪𝐅×ψ𝐋(xτα2y(α1))d×ydα2=ζ𝐅(1)env𝐋(z)<eenψ𝐋(z(α1))dz|z|𝐋.\displaystyle\frac{1}{2}\sum_{\tau}\int_{\mathbf{L}^{1}}\int_{\varpi_{\mathbf{F}}^{-n}\mathcal{O}_{\mathbf{F}}^{\times}}\psi_{\mathbf{L}}(x_{\tau}\alpha_{2}y(\alpha-1))\mathrm{d}^{\times}y\mathrm{d}\alpha_{2}=\zeta_{\mathbf{F}}(1)\int_{-en\leqslant v_{\mathbf{L}}(z)<e-en}\psi_{\mathbf{L}}(z(\alpha-1))\frac{\mathrm{d}z}{\lvert z\rvert_{\mathbf{L}}}.

Therefore we continue the calculation as

𝐅×|H(y)|2d×y|y|𝐅\displaystyle\displaystyle\int_{\mathbf{F}^{\times}}\left\lvert H(y)\right\rvert^{2}\tfrac{\mathrm{d}^{\times}y}{\lvert y\rvert_{\mathbf{F}}} =ζ𝐅(1)nn01+e𝐋1β(α)(ϖ𝐋n𝒪𝐋×ψ𝐋(z(α1))dz|z|𝐋)dα\displaystyle\displaystyle=\zeta_{\mathbf{F}}(1)\sum_{n\geqslant n_{0}-1+e}\int_{\mathbf{L}^{1}}\beta(\alpha)\left(\int_{\varpi_{\mathbf{L}}^{-n}\mathcal{O}_{\mathbf{L}}^{\times}}\psi_{\mathbf{L}}(z(\alpha-1))\tfrac{\mathrm{d}z}{\lvert z\rvert_{\mathbf{L}}}\right)\mathrm{d}\alpha
=ζ𝐅(1)ζ𝐋(1)q𝐋e12nn01+e𝐋1β(α)(𝟙𝒫𝐋n+1e(α1)q𝐋1𝟙𝒫𝐋ne(α1))dα\displaystyle\displaystyle=\tfrac{\zeta_{\mathbf{F}}(1)}{\zeta_{\mathbf{L}}(1)}q_{\mathbf{L}}^{-\frac{e-1}{2}}\sum_{n\geqslant n_{0}-1+e}\int_{\mathbf{L}^{1}}\beta(\alpha)\left(\mathbbm{1}_{\mathcal{P}_{\mathbf{L}}^{n+1-e}}(\alpha-1)-q_{\mathbf{L}}^{-1}\mathbbm{1}_{\mathcal{P}_{\mathbf{L}}^{n-e}}(\alpha-1)\right)\mathrm{d}\alpha
=ζ𝐅(1)ζ𝐋(1)2q𝐋e12nn0Vol(𝐋1(1+𝒫𝐋n)),\displaystyle\displaystyle=\tfrac{\zeta_{\mathbf{F}}(1)}{\zeta_{\mathbf{L}}(1)^{2}}q_{\mathbf{L}}^{-\frac{e-1}{2}}\sum_{n\geqslant n_{0}}{\rm Vol}(\mathbf{L}^{1}\cap(1+\mathcal{P}_{\mathbf{L}}^{n})),

and conclude the lower bound by Proposition 3.5 (3) and the fact en0\displaystyle e\mid n_{0} (see Claim in Lemma 3.3). ∎

Remark 3.16.

By (3.2), the bounds in Lemma 3.13 (1) & 3.15 can be rewritten as h(π0)q𝔠(π0)2\displaystyle h(\pi_{0})\gg q^{-\lceil\frac{\mathfrak{c}(\pi_{0})}{2}\rceil}.

4. Dual Weight Functions: Common Reductions

4.1. Infinite Part

Recall the constant a(Π)\displaystyle a(\Pi) for the stability range defined in Proposition 2.4. Take the “infinite part” of the test function as H=En1\displaystyle H_{\infty}=E_{\geqslant n_{1}} for some

(4.1) n1max(2n0/e+e1,a(Π))(2).n_{1}\geqslant\max(2n_{0}/e+e-1,a(\Pi))(\geqslant 2).

Recall the formula for the dual weight function

(4.2) h~(χ)=𝐅×𝒱~Π,ψ𝔪1(H)(t)ψ(t)χ1(t)|t|12d×t.\widetilde{h}(\chi)=\int_{\mathbf{F}^{\times}}\widetilde{\mathcal{VH}}_{\Pi,\psi}\circ\mathfrak{m}_{-1}(H)(t)\cdot\psi(-t)\chi^{-1}(t)\lvert t\rvert^{-\frac{1}{2}}\mathrm{d}^{\times}t.
Lemma 4.1.

The partial dual weight function

h~(χ):=𝐅×𝒱~Π,ψ𝔪1(H)(t)ψ(t)χ1(t)|t|12d×t\displaystyle\widetilde{h}_{\infty}(\chi):=\int_{\mathbf{F}^{\times}}\widetilde{\mathcal{VH}}_{\Pi,\psi}\circ\mathfrak{m}_{-1}(H_{\infty})(t)\cdot\psi(-t)\chi^{-1}(t)\lvert t\rvert^{-\frac{1}{2}}\mathrm{d}^{\times}t

is non-vanishing only if χ\displaystyle\chi is unramified, in which case we have

h~(χ)=χ(ϖ𝐅)n1qn121χ(ϖ𝐅)q12.\displaystyle\widetilde{h}_{\infty}(\chi)=\frac{\chi(\varpi_{\mathbf{F}})^{n_{1}}q^{-\frac{n_{1}}{2}}}{1-\chi(\varpi_{\mathbf{F}})q^{-\frac{1}{2}}}.
Proof.

By Proposition 2.4 we have

𝒱~Π,ψ𝔪1(En1)(t)=ψ(t)𝟙𝒫𝐅n1(t1).\displaystyle\widetilde{\mathcal{VH}}_{\Pi,\psi}\circ\mathfrak{m}_{-1}(E_{\geqslant n_{1}})(t)=\psi(t)\cdot\mathbbm{1}_{\mathcal{P}_{\mathbf{F}}^{n_{1}}}(t^{-1}).

The desired formula follows readily from the above one. ∎

Taking into account Lemma 3.10 (2) we introduce the “finite part” of the test function as

(4.3) Hc:=HH=n=2e(n02)+32en2n11Hn,Hn:=H𝟙ϖ𝐅n𝒪𝐅×.H_{c}:=H-H_{\infty}=\sideset{}{{}_{\begin{subarray}{c}n=\frac{2}{e}(n_{0}-2)+3\\ 2\mid en\end{subarray}}^{2n_{1}-1}}{\sum}H_{n},\qquad H_{n}:=H\cdot\mathbbm{1}_{\varpi_{\mathbf{F}}^{-n}\mathcal{O}_{\mathbf{F}}^{\times}}.

Since each function HnCc(𝐅×)\displaystyle H_{n}\in{\rm C}_{c}^{\infty}(\mathbf{F}^{\times}), we may replace the extended Voronoi–Hankel transform with its original version in the dual weight formula and turn to the study of

(4.4) h~n(χ):=𝐅×𝒱Π,ψ𝔪1(Hn)(t)ψ(t)χ1(t)|t|12d×t.\widetilde{h}_{n}(\chi):=\int_{\mathbf{F}^{\times}}\mathcal{VH}_{\Pi,\psi}\circ\mathfrak{m}_{-1}(H_{n})(t)\cdot\psi(-t)\chi^{-1}(t)\lvert t\rvert^{-\frac{1}{2}}\mathrm{d}^{\times}t.

We divide the domain of integration into two parts: 𝒪𝐅{0}\displaystyle\mathcal{O}_{\mathbf{F}}-\{0\} and 𝐅𝒪𝐅\displaystyle\mathbf{F}-\mathcal{O}_{\mathbf{F}}, giving the decomposition

(4.5) h~n(χ)=h~n+(χ)+h~n(χ).\widetilde{h}_{n}(\chi)=\widetilde{h}_{n}^{+}(\chi)+\widetilde{h}_{n}^{-}(\chi).

Accordingly the dual weight functions (see (1.1)) have the following decomposition

(4.6) h~(χ)=h~(χ)+h~c(χ),h~c(χ)=h~c+(χ)+h~c(χ),h~c±(χ)=n=2e(n02)+32en2n11h~n±(χ),\widetilde{h}(\chi)=\widetilde{h}_{\infty}(\chi)+\widetilde{h}_{c}(\chi),\quad\widetilde{h}_{c}(\chi)=\widetilde{h}_{c}^{+}(\chi)+\widetilde{h}_{c}^{-}(\chi),\quad\widetilde{h}_{c}^{\pm}(\chi)=\sideset{}{{}_{\begin{subarray}{c}n=\frac{2}{e}(n_{0}-2)+3\\ 2\mid en\end{subarray}}^{2n_{1}-1}}{\sum}\widetilde{h}_{n}^{\pm}(\chi),
(4.7) H~(χ)=H~(χ)+H~c(χ),H~c(χ)=H~c+(χ)+H~c(χ),H~c±(χ)=n=2e(n02)+32en2n11H~n±(χ).\widetilde{H}(\chi)=\widetilde{H}_{\infty}(\chi)+\widetilde{H}_{c}(\chi),\quad\widetilde{H}_{c}(\chi)=\widetilde{H}_{c}^{+}(\chi)+\widetilde{H}_{c}^{-}(\chi),\quad\widetilde{H}_{c}^{\pm}(\chi)=\sideset{}{{}_{\begin{subarray}{c}n=\frac{2}{e}(n_{0}-2)+3\\ 2\mid en\end{subarray}}^{2n_{1}-1}}{\sum}\widetilde{H}_{n}^{\pm}(\chi).

We call those (partial) dual weight functions with “+\displaystyle+” (resp. “\displaystyle-”) the postive (resp. negative) part.

4.2. Positive Part

We first study h~n+(χ)\displaystyle\widetilde{h}_{n}^{+}(\chi). Since ψ\displaystyle\psi is trivial on 𝒪𝐅\displaystyle\mathcal{O}_{\mathbf{F}}, we have h~n+(χ)=h~n(1/2,χ)\displaystyle\widetilde{h}_{n}^{+}(\chi)=\widetilde{h}_{n}(1/2,\chi) for

(4.8) h~n(s,χ):=𝒪𝐅{0}𝒱Π,ψ𝔪1(Hn)(t)χ1(t)|t|sd×t,\widetilde{h}_{n}(s,\chi):=\int_{\mathcal{O}_{\mathbf{F}}-\{0\}}\mathcal{VH}_{\Pi,\psi}\circ\mathfrak{m}_{-1}(H_{n})(t)\cdot\chi^{-1}(t)\lvert t\rvert^{-s}\mathrm{d}^{\times}t,

which is simply the partial sum of non-negative powers of X=qs\displaystyle X=q^{s} in the Laurent series expansion of

(4.9) fn(qs;χ,H)=𝐅×𝒱Π,ψ𝔪1(Hn)(t)χ1(t)|t|sd×t.f_{n}(q^{s};\chi,H)=\int_{\mathbf{F}^{\times}}\mathcal{VH}_{\Pi,\psi}\circ\mathfrak{m}_{-1}(H_{n})(t)\cdot\chi^{-1}(t)\lvert t\rvert^{-s}\mathrm{d}^{\times}t.

By the local functional equation we have (see Lemma 3.6)

(4.10) fn(qs;χ,H)=εn(χ,H)qn(s1)ε(s,Πχ,ψ)L(1s,Π~χ1)L(s,Πχ).f_{n}(q^{s};\chi,H)=\varepsilon_{n}(\chi,H)q^{n(s-1)}\cdot\varepsilon\left(s,\Pi\otimes\chi,\psi\right)\frac{L(1-s,\widetilde{\Pi}\otimes\chi^{-1})}{L(s,\Pi\otimes\chi)}.

To relate fn(qs;χ,H)\displaystyle f_{n}(q^{s};\chi,H) with h~n(s,χ)\displaystyle\widetilde{h}_{n}(s,\chi) we need the following crucial lemma.

Lemma 4.2.

Let f(X)=n>anXn\displaystyle f(X)=\sideset{}{{}_{n>-\infty}}{\sum}a_{n}X^{n} be a Laurent series converging in 0<|X|<ρ\displaystyle 0<\lvert X\rvert<\rho for some ρ>1\displaystyle\rho>1. Let f+(X)=n>anXn\displaystyle f_{+}(X)=\sideset{}{{}_{n>-\infty}}{\sum}a_{n}X^{n}. Let D=XddX\displaystyle D=X\tfrac{\mathrm{d}}{\mathrm{d}X}. Assume f(X)\displaystyle f(X) (hence f+(X)\displaystyle f_{+}(X)) has a meromorphic continuation to X\displaystyle X\in\mathbb{C}.

  • (0)

    For any X\displaystyle X and any ϵ<min(1,ρ|X|1)\displaystyle\epsilon<\min(1,\rho\lvert X\rvert^{-1}) we have the relation

    f+(X)=f(X)|z|=ϵf(Xz)1zdz2πi.\displaystyle f_{+}(X)=f(X)-\int_{\lvert z\rvert=\epsilon}\frac{f(Xz)}{1-z}\frac{\mathrm{d}z}{2\pi i}.
  • (1)

    For 0<|X|<ρ\displaystyle 0<\lvert X\rvert<\rho and any 1<r<ρ|X|1\displaystyle 1<r<\rho\lvert X\rvert^{-1} we have the relation

    f+(X)=|z|=rf(Xz)z1dz2πi=f(X)+|z|=1f(Xz)f(X)z1dz2πi.\displaystyle f_{+}(X)=\int_{\lvert z\rvert=r}\frac{f(Xz)}{z-1}\frac{\mathrm{d}z}{2\pi i}=f(X)+\int_{\lvert z\rvert=1}\frac{f(Xz)-f(X)}{z-1}\frac{\mathrm{d}z}{2\pi i}.

    Consequently, for 0<|X|<ρ\displaystyle 0<\lvert X\rvert<\rho and n0\displaystyle n\in\mathbb{Z}_{\geqslant 0} we have the bound

    |(Dnf+)(X)||(Dnf)(X)|+sup|z|=1|(Dn+1f)(Xz)|.\displaystyle\left\lvert(D^{n}f_{+})(X)\right\rvert\ll\left\lvert(D^{n}f)(X)\right\rvert+\sideset{}{{}_{\lvert z\rvert=1}}{\sup}\left\lvert(D^{n+1}f)(Xz)\right\rvert.
  • (2)

    Suppose f(X)=Q(X)P(X)1\displaystyle f(X)=Q(X)P(X)^{-1} for some Q[X,X1],P[X]\displaystyle Q\in\mathbb{C}[X,X^{-1}],P\in\mathbb{C}[X]. Let

    P(X)=j=1r(1bjX)mj\displaystyle P(X)=\sideset{}{{}_{j=1}^{r}}{\prod}(1-b_{j}X)^{m_{j}}

    with mj1\displaystyle m_{j}\in\mathbb{Z}_{\geqslant 1} and distinct bj0\displaystyle b_{j}\neq 0. Introduce

    Pj(X):=ij(1biX)mi,Cj,k:=(1)kk!bjk(QPj)(k)(bj1).\displaystyle P_{j}(X):=\sideset{}{{}_{i\neq j}}{\prod}(1-b_{i}X)^{m_{i}},\quad C_{j,k}:=\frac{(-1)^{k}}{k!\cdot b_{j}^{k}}\left(\frac{Q}{P_{j}}\right)^{(k)}(b_{j}^{-1}).

    Suppose the highest power Xm\displaystyle X^{m} in Q\displaystyle Q satisfies m<degP\displaystyle m<\deg P. Then we have

    P(X)f+(X)=j=1rk=0mj1Cj,k(1bjX)kPj(X).\displaystyle P(X)f_{+}(X)=\sideset{}{{}_{j=1}^{r}}{\sum}\sideset{}{{}_{k=0}^{m_{j}-1}}{\sum}C_{j,k}\cdot(1-b_{j}X)^{k}P_{j}(X).
Proof.

(0) The stated formula follows from the residue theorem via

|z|=ϵf(Xz)1zdz2πi=|z|=ϵ(n>anXnzn)(k=0zk)dz2πi=n<0anXn,\displaystyle\int_{\lvert z\rvert=\epsilon}\frac{f(Xz)}{1-z}\frac{\mathrm{d}z}{2\pi i}=\int_{\lvert z\rvert=\epsilon}\left(\sum_{n>-\infty}a_{n}X^{n}z^{n}\right)\left(\sum_{k=0}^{\infty}z^{k}\right)\frac{\mathrm{d}z}{2\pi i}=\sum_{n<0}a_{n}X^{n},

since only the terms for n+k=1\displaystyle n+k=-1 give non-zero contribution.

(1) Clearly the stated formula follows from the one in (0) by the residue theorem. To see the bound, we introduce g(z)=f(Xz)\displaystyle g(z)=f(Xz) and rewrite the integral as

|z|=1f(Xz)f(X)z1dz2πi=ππg(eiθ)g(1)eiθ1dθ2π=ππieiθ1(0θg(eit)eitdt)dθ2π.\displaystyle\int_{\lvert z\rvert=1}\frac{f(Xz)-f(X)}{z-1}\frac{\mathrm{d}z}{2\pi i}=\int_{-\pi}^{\pi}\frac{g(e^{i\theta})-g(1)}{e^{i\theta}-1}\frac{\mathrm{d}\theta}{2\pi}=\int_{-\pi}^{\pi}\frac{i}{e^{i\theta}-1}\left(\int_{0}^{\theta}g^{\prime}(e^{it})e^{it}\mathrm{d}t\right)\frac{\mathrm{d}\theta}{2\pi}.

Note that zg(z)=Df(Xz)\displaystyle zg^{\prime}(z)=Df(Xz). Therefore we obtain the formula

f+(X)=f(X)+ππieiθ1(0θ(Df)(Xeit)dt)dθ2π.\displaystyle f_{+}(X)=f(X)+\int_{-\pi}^{\pi}\frac{i}{e^{i\theta}-1}\left(\int_{0}^{\theta}(Df)(Xe^{it})\mathrm{d}t\right)\frac{\mathrm{d}\theta}{2\pi}.

Applying Dn\displaystyle D^{n} on both sides we get a formula and conclude the stated bound as

(Dnf+)(X)=(Dnf)(X)+ππieiθ1(0θ(Dn+1f)(Xeit)dt)dθ2π|(Dnf)(X)|+sup|z|=1|(Dn+1f)(Xz)|ππ|θeiθ1|dθ2π.(D^{n}f_{+})(X)=(D^{n}f)(X)+\int_{-\pi}^{\pi}\frac{i}{e^{i\theta}-1}\left(\int_{0}^{\theta}(D^{n+1}f)(Xe^{it})\mathrm{d}t\right)\frac{\mathrm{d}\theta}{2\pi}\\ \ll\left\lvert(D^{n}f)(X)\right\rvert+\sideset{}{{}_{\lvert z\rvert=1}}{\sup}\left\lvert(D^{n+1}f)(Xz)\right\rvert\cdot\int_{-\pi}^{\pi}\left\lvert\frac{\theta}{e^{i\theta}-1}\right\rvert\frac{\mathrm{d}\theta}{2\pi}.

(2) We may assume 0<|X|<ρ\displaystyle 0<\lvert X\rvert<\rho and depart from the first equation in (1), then move the contour to |z|=r\displaystyle\lvert z\rvert=r for r1\displaystyle r\gg 1, picking up the residues. The contour integral tends to 0\displaystyle 0 as r+\displaystyle r\to+\infty due to the assumption m<degP\displaystyle m<\deg P. Multiplying the resulted formula by P(X)\displaystyle P(X), we get the stated formula. ∎

Definition 4.3.

For any generic irreducible representation Π\displaystyle\Pi of GLr(𝐅)\displaystyle{\rm GL}_{r}(\mathbf{F}), let d(Π)\displaystyle d(\Pi) be the degree of the L\displaystyle L-function L(s,Π)\displaystyle L(s,\Pi), namely the degree of the polynomial L(s,Π)1\displaystyle L(s,\Pi)^{-1} in X:=qs\displaystyle X:=q^{-s}. Write

ρ(Π):=𝔠(Π)+d(Π).\displaystyle\mathfrak{\rho}(\Pi):=\mathfrak{c}(\Pi)+d(\Pi).

Introduce the set of exponents of Π\displaystyle\Pi by

E(Π):={ξ𝒪𝐅×^|d(Πξ)>0}.\displaystyle\mathrm{E}(\Pi):=\left\{\xi\in\widehat{\mathcal{O}_{\mathbf{F}}^{\times}}\ \middle|\ d(\Pi\otimes\xi)>0\right\}.
Lemma 4.4.

Let Π\displaystyle\Pi be a generic irreducible representation of GL3(𝐅)\displaystyle{\rm GL}_{3}(\mathbf{F}). We have ρ(Π)3\displaystyle\mathfrak{\rho}(\Pi)\geqslant 3, |E(Π)|3\displaystyle\left\lvert\mathrm{E}(\Pi)\right\rvert\leqslant 3. For any ξE(Π)\displaystyle\xi\in\mathrm{E}(\Pi) we have the bounds

𝔠(ξ)𝔠(Π),𝔠(Πξ)2𝔠(Π),ρ(Πξ)2𝔠(Π)+3.\displaystyle\mathfrak{c}(\xi)\leqslant\mathfrak{c}(\Pi),\quad\mathfrak{c}(\Pi\otimes\xi)\leqslant 2\mathfrak{c}(\Pi),\quad\mathfrak{\rho}(\Pi\otimes\xi)\leqslant 2\mathfrak{c}(\Pi)+3.
Proof.

The upper bound for ρ(Πξ)\displaystyle\mathfrak{\rho}(\Pi\otimes\xi) obviously follows from the one for 𝔠(Πξ)\displaystyle\mathfrak{c}(\Pi\otimes\xi) since d(Πξ)3\displaystyle d(\Pi\otimes\xi)\leqslant 3 in any case. For the other bounds we distinguish cases for Π\displaystyle\Pi.

(1) If 𝔠(Π)=0\displaystyle\mathfrak{c}(\Pi)=0, then Π\displaystyle\Pi is spherical, E(Π)={𝟙}\displaystyle\mathrm{E}(\Pi)=\{\mathbbm{1}\} and d(Π)=3\displaystyle d(\Pi)=3. The stated bounds clearly hold.

(2) If 𝔠(Π)>0\displaystyle\mathfrak{c}(\Pi)>0 and if Π=χ1χ2χ3\displaystyle\Pi=\chi_{1}\boxplus\chi_{2}\boxplus\chi_{3} is induced from the Borel subgroup, then we have

ρ(Π)=i=13ρ(χi)=i=13max(𝔠(χi),1)3.\displaystyle\mathfrak{\rho}(\Pi)=\sideset{}{{}_{i=1}^{3}}{\sum}\mathfrak{\rho}(\chi_{i})=\sideset{}{{}_{i=1}^{3}}{\sum}\max(\mathfrak{c}(\chi_{i}),1)\geqslant 3.

We also have E(Π){ξ11,ξ21,ξ31}\displaystyle\mathrm{E}(\Pi)\subset\{\xi_{1}^{-1},\xi_{2}^{-1},\xi_{3}^{-1}\} with χj𝒪𝐅×=ξj\displaystyle\chi_{j}\mid_{\mathcal{O}_{\mathbf{F}}^{\times}}=\xi_{j}, from which we deduce the bound for |E(Π)|\displaystyle\left\lvert\mathrm{E}(\Pi)\right\rvert. Obviously we have 𝔠(ξj)𝔠(Π)\displaystyle\mathfrak{c}(\xi_{j})\leqslant\mathfrak{c}(\Pi). Taking ξ=ξ11\displaystyle\xi=\xi_{1}^{-1} for example, we have the second stated bound as

𝔠(Πξ)=𝔠(χ2ξ11)+𝔠(χ3ξ11)2max{𝔠(χj):1j3}2𝔠(Π).\displaystyle\mathfrak{c}(\Pi\otimes\xi)=\mathfrak{c}(\chi_{2}\xi_{1}^{-1})+\mathfrak{c}(\chi_{3}\xi_{1}^{-1})\leqslant 2\max\{\mathfrak{c}(\chi_{j}):1\leqslant j\leqslant 3\}\leqslant 2\mathfrak{c}(\Pi).

(3) If 𝔠(Π)>0\displaystyle\mathfrak{c}(\Pi)>0 and if Π=πχ\displaystyle\Pi=\pi\boxplus\chi for some supercuspidal π\displaystyle\pi of GL2(𝐅)\displaystyle{\rm GL}_{2}(\mathbf{F}), then the central character of π\displaystyle\pi is χ1\displaystyle\chi^{-1}. We have ρ(Π)=𝔠(π)+max(𝔠(χ),1)2+1=3\displaystyle\mathfrak{\rho}(\Pi)=\mathfrak{c}(\pi)+\max(\mathfrak{c}(\chi),1)\geqslant 2+1=3. We also have E(Π)={χ1𝒪𝐅×}\displaystyle\mathrm{E}(\Pi)=\{\chi^{-1}\mid_{\mathcal{O}_{\mathbf{F}}^{\times}}\}. For ξ=χ1𝒪𝐅×\displaystyle\xi=\chi^{-1}\mid_{\mathcal{O}_{\mathbf{F}}^{\times}} we have 𝔠(ξ)=𝔠(χ)𝔠(Π)\displaystyle\mathfrak{c}(\xi)=\mathfrak{c}(\chi)\leqslant\mathfrak{c}(\Pi) and get the second stated bound via πχ1π~\displaystyle\pi\otimes\chi^{-1}\simeq\widetilde{\pi} as

𝔠(Πξ)=𝔠(πχ1)=𝔠(π~)=𝔠(π)𝔠(Π).\displaystyle\mathfrak{c}(\Pi\otimes\xi)=\mathfrak{c}(\pi\otimes\chi^{-1})=\mathfrak{c}(\widetilde{\pi})=\mathfrak{c}(\pi)\leqslant\mathfrak{c}(\Pi).

(4) If 𝔠(Π)>0\displaystyle\mathfrak{c}(\Pi)>0 and if Π=Stηχ\displaystyle\Pi=\mathrm{St}_{\eta}\boxplus\chi, then χη2=𝟙\displaystyle\chi\eta^{2}=\mathbbm{1}. By [15, §8 Proposition & §10 Proposition] we have

ρ(Π)={2𝔠(η)if 𝔠(η)11if 𝔠(η)=0+d(η)+max(𝔠(χ),1)3.\displaystyle\mathfrak{\rho}(\Pi)=\begin{cases}2\mathfrak{c}(\eta)&\text{if }\mathfrak{c}(\eta)\geqslant 1\\ 1&\text{if }\mathfrak{c}(\eta)=0\end{cases}+d(\eta)+\max(\mathfrak{c}(\chi),1)\geqslant 3.

We also have E(Π)={χ1𝒪𝐅×,η1𝒪𝐅×}\displaystyle\mathrm{E}(\Pi)=\{\chi^{-1}\mid_{\mathcal{O}_{\mathbf{F}}^{\times}},\eta^{-1}\mid_{\mathcal{O}_{\mathbf{F}}^{\times}}\}. For ξ=χ1𝒪𝐅×\displaystyle\xi=\chi^{-1}\mid_{\mathcal{O}_{\mathbf{F}}^{\times}} we argue the same as in the proof of (3). For ξ=η1𝒪𝐅×\displaystyle\xi=\eta^{-1}\mid_{\mathcal{O}_{\mathbf{F}}^{\times}} we have

𝔠(ξ)=𝔠(η)<1+𝔠(η)=𝔠(Πξ)max(2𝔠(η),1)+𝔠(χ)=𝔠(Π).\displaystyle\mathfrak{c}(\xi)=\mathfrak{c}(\eta)<1+\mathfrak{c}(\eta)=\mathfrak{c}(\Pi\otimes\xi)\leqslant\max(2\mathfrak{c}(\eta),1)+\mathfrak{c}(\chi)=\mathfrak{c}(\Pi).

(5) If 𝔠(Π)>0\displaystyle\mathfrak{c}(\Pi)>0 and if Π=Stη\displaystyle\Pi=\mathrm{St}_{\eta} is a twist of the Steinberg representation of PGL3(𝐅)\displaystyle{\rm PGL}_{3}(\mathbf{F}), then by [15, §8 Proposition & §10 Proposition] we have

ρ(Π)={3𝔠(η)if 𝔠(η)12if 𝔠(η)=0+d(η)3.\displaystyle\mathfrak{\rho}(\Pi)=\begin{cases}3\mathfrak{c}(\eta)&\text{if }\mathfrak{c}(\eta)\geqslant 1\\ 2&\text{if }\mathfrak{c}(\eta)=0\end{cases}+d(\eta)\geqslant 3.

We also have E(Π)={η1𝒪𝐅×}\displaystyle\mathrm{E}(\Pi)=\{\eta^{-1}\mid_{\mathcal{O}_{\mathbf{F}}^{\times}}\}. For ξ=η1𝒪𝐅×\displaystyle\xi=\eta^{-1}\mid_{\mathcal{O}_{\mathbf{F}}^{\times}} we have

𝔠(ξ)=𝔠(η)max(3𝔠(η),2)=𝔠(Π),𝔠(Πξ)=2𝔠(Π).\displaystyle\mathfrak{c}(\xi)=\mathfrak{c}(\eta)\leqslant\max(3\mathfrak{c}(\eta),2)=\mathfrak{c}(\Pi),\quad\mathfrak{c}(\Pi\otimes\xi)=2\leqslant\mathfrak{c}(\Pi).

(6) If Π\displaystyle\Pi is supercuspidal, then E(Π)=\displaystyle\mathrm{E}(\Pi)=\emptyset. We have ρ(Π)=𝔠(Π)3\displaystyle\mathfrak{\rho}(\Pi)=\mathfrak{c}(\Pi)\geqslant 3. ∎

The following subset of 𝒪𝐅×^\displaystyle\widehat{\mathcal{O}_{\mathbf{F}}^{\times}} will turn out to be important for the bound of h~n+(χ)\displaystyle\widetilde{h}_{n}^{+}(\chi):

(4.11) 𝒜n=𝒜n(β,Π):={ξ𝒪𝐅×^|ρ(Πξ)n,εn(ξ,H)0}.\mathcal{A}_{n}=\mathcal{A}_{n}(\beta,\Pi):=\left\{\xi\in\widehat{\mathcal{O}_{\mathbf{F}}^{\times}}\ \middle|\ \mathfrak{\rho}(\Pi\otimes\xi)\leqslant n,\ \varepsilon_{n}(\xi,H)\neq 0\right\}.
Lemma 4.5.

For any ξ𝒜n\displaystyle\xi\in\mathcal{A}_{n} we have 𝔠(ξ)max(n0e,2𝔠(Π))\displaystyle\mathfrak{c}(\xi)\leqslant\max\left(\tfrac{n_{0}}{e},2\mathfrak{c}(\Pi)\right).

Proof.

Suppose ξ𝒜n\displaystyle\xi\in\mathcal{A}_{n} and 𝔠(ξ)>max(n0e,2𝔠(Π))\displaystyle\mathfrak{c}(\xi)>\max\left(\tfrac{n_{0}}{e},2\mathfrak{c}(\Pi)\right). In particular, we have 𝔠(ξ)a(Π)\displaystyle\mathfrak{c}(\xi)\geqslant a(\Pi), the stability barrier of Π\displaystyle\Pi, by Proposition 2.4. Therefore we get 𝔠(Πξ)=3𝔠(ξ)\displaystyle\mathfrak{c}(\Pi\otimes\xi)=3\mathfrak{c}(\xi), d(Πξ)=0\displaystyle d(\Pi\otimes\xi)=0 and deduce

(4.12) 3𝔠(ξ)=ρ(Πξ)n.3\mathfrak{c}(\xi)=\mathfrak{\rho}(\Pi\otimes\xi)\leqslant n.

On the other hand, we claim that the condition 𝔠(ξ)>n0e\displaystyle\mathfrak{c}(\xi)>\tfrac{n_{0}}{e} and εn(ξ,H)0\displaystyle\varepsilon_{n}(\xi,H)\neq 0 imply

(4.13) 𝔠(ξ)=n2,\mathfrak{c}(\xi)=\tfrac{n}{2},

and conclude the proof by comparing (4.12) and (4.13) taking into account n2e(n02)+3>0\displaystyle n\geqslant\tfrac{2}{e}(n_{0}-2)+3>0. In fact, the proof of (4.13) is case-by-case check with Lemma 3.6: (1) If 𝐋/𝐅\displaystyle\mathbf{L}/\mathbf{F} is split, then e=1\displaystyle e=1 and 𝔠(ξ)>n0=𝔠(χ0±1)\displaystyle\mathfrak{c}(\xi)>n_{0}=\mathfrak{c}(\chi_{0}^{\pm 1}). Only the first case in Lemma 3.6 is possible, yielding (4.13). (2) If 𝐋/𝐅\displaystyle\mathbf{L}/\mathbf{F} is unramified, then e=1\displaystyle e=1 and by (3.9) we have 𝔠(ξNr)=𝔠(ξ)>n0=𝔠(β)\displaystyle\mathfrak{c}(\xi\circ{\rm Nr})=\mathfrak{c}(\xi)>n_{0}=\mathfrak{c}(\beta). The last case in Lemma 3.6 yields (4.13). (3) If 𝐋/𝐅\displaystyle\mathbf{L}/\mathbf{F} is ramified, then e=2\displaystyle e=2 and by (3.9) we have 𝔠(ξNr)2𝔠(ξ)12(n0/2+1)1=n0+1>𝔠(β)\displaystyle\mathfrak{c}(\xi\circ{\rm Nr})\geqslant 2\mathfrak{c}(\xi)-1\geqslant 2(n_{0}/2+1)-1=n_{0}+1>\mathfrak{c}(\beta). The last case in Lemma 3.6 yields (4.13). ∎

Lemma 4.6.

Write ξ=χ𝒪𝐅×\displaystyle\xi=\chi\mid_{\mathcal{O}_{\mathbf{F}}^{\times}}. For any ϵ>0\displaystyle\epsilon>0 sufficiently small we have the bound

|h~c+(χ)|2e(n02)+3n<ρ(Πξ)|h~n+(χ)|+nρ(Πχ)|h~n+(χ)|ϵ𝐂(Π)ϵqn0e+1e2𝟙max(n0e,2𝔠(Π))(𝔠(ξ)).\left\lvert\widetilde{h}_{c}^{+}(\chi)\right\rvert\leqslant\sideset{}{{}_{\frac{2}{e}(n_{0}-2)+3\leqslant n<\mathfrak{\rho}(\Pi\otimes\xi)}}{\sum}\left\lvert\widetilde{h}_{n}^{+}(\chi)\right\rvert+\sideset{}{{}_{n\geqslant\mathfrak{\rho}(\Pi\otimes\chi)}}{\sum}\left\lvert\widetilde{h}_{n}^{+}(\chi)\right\rvert\\ \ll_{\epsilon}\mathbf{C}(\Pi)^{\epsilon}q^{-\frac{n_{0}}{e}+\frac{1-e}{2}}\cdot\mathbbm{1}_{\leqslant\max\left(\frac{n_{0}}{e},2\mathfrak{c}(\Pi)\right)}(\mathfrak{c}(\xi)).
Proof.

Write the relevant L\displaystyle L-functions of Πχ\displaystyle\Pi\otimes\chi as

L(s,Πχ)=j=1d(1ajqs)1\displaystyle L(s,\Pi\otimes\chi)=\sideset{}{{}_{j=1}^{d}}{\prod}(1-a_{j}q^{-s})^{-1}

where d=d(Πχ){0,1,2,3}\displaystyle d=d(\Pi\otimes\chi)\in\{0,1,2,3\} is the degree, and aj=aj(Πχ)\displaystyle a_{j}=a_{j}(\Pi\otimes\chi) are the generalized Satake parameters of Πχ\displaystyle\Pi\otimes\chi satisfying |aj|{1,q12,q1}\displaystyle\lvert a_{j}\rvert\in\{1,q^{-\frac{1}{2}},q^{-1}\} by temperedness. We can rewrite (4.10) as

(4.14) fn(X;χ,H)=εn(χ,H)ε(12,Πχ,ψ)q𝔠(Πχ)2nXnρ(Πχ)j=1dXaj1aj¯q1X.f_{n}(X;\chi,H)=\varepsilon_{n}(\chi,H)\varepsilon\left(\tfrac{1}{2},\Pi\otimes\chi,\psi\right)q^{\frac{\mathfrak{c}(\Pi\otimes\chi)}{2}-n}X^{n-\mathfrak{\rho}(\Pi\otimes\chi)}\cdot\sideset{}{{}_{j=1}^{d}}{\prod}\tfrac{X-a_{j}}{1-\overline{a_{j}}q^{-1}X}.

If nρ:=ρ(Πχ)\displaystyle n\geqslant\mathfrak{\rho}:=\mathfrak{\rho}(\Pi\otimes\chi), then fn(X;χ,H)0\displaystyle f_{n}(X;\chi,H)\neq 0 implies εn(χ,H)0\displaystyle\varepsilon_{n}(\chi,H)\neq 0, hence χ𝒪𝐅×𝒜n\displaystyle\chi\mid_{\mathcal{O}_{\mathbf{F}}^{\times}}\in\mathcal{A}_{n} by definition. The Laurent expansion of fn(X;χ,H)\displaystyle f_{n}(X;\chi,H) contains only non-negative powers of X\displaystyle X. Therefore the right hand side is equal to h~n(s,χ)\displaystyle\widetilde{h}_{n}(s,\chi) by definition (4.8). Consequently we get the formula and deduce the bound

h~n(s,χ)=εn(χ,H)ε(12,Πχ,ψ)q𝔠(Πχ)2nqs(n𝔠(Πχ))j=1d1ajqs1aj¯qs1,\displaystyle\widetilde{h}_{n}(s,\chi)=\varepsilon_{n}(\chi,H)\varepsilon\left(\tfrac{1}{2},\Pi\otimes\chi,\psi\right)q^{\frac{\mathfrak{c}(\Pi\otimes\chi)}{2}-n}q^{s(n-\mathfrak{c}(\Pi\otimes\chi))}\cdot\sideset{}{{}_{j=1}^{d}}{\prod}\tfrac{1-a_{j}q^{-s}}{1-\overline{a_{j}}q^{s-1}},
(4.15) |h~n+(χ)|=|h~n(1/2,χ)|ϑ3𝟙𝒜n(χ𝒪𝐅×)qn2.\left\lvert\widetilde{h}_{n}^{+}(\chi)\right\rvert=\left\lvert\widetilde{h}_{n}(1/2,\chi)\right\rvert\ll_{\vartheta_{3}}\mathbbm{1}_{\mathcal{A}_{n}}(\chi\mid_{\mathcal{O}_{\mathbf{F}}^{\times}})\cdot q^{-\frac{n}{2}}.

If n<ρ\displaystyle n<\mathfrak{\rho}, then the Laurent expansion of fn(X;χ,H)\displaystyle f_{n}(X;\chi,H) contains non-negative powers of X\displaystyle X only if d=d(Πχ)>0\displaystyle d=d(\Pi\otimes\chi)>0, i.e. χ𝒪𝐅×E(Π)\displaystyle\chi\mid_{\mathcal{O}_{\mathbf{F}}^{\times}}\in\mathrm{E}(\Pi) where E(Π)\displaystyle\mathrm{E}(\Pi) is the set of exponents of Π\displaystyle\Pi introduced in Definition 4.3. By Lemma 4.4 and the summation range of n\displaystyle n we have

(4.16) 2e(n02)+3n2𝔠(Π)+2.\tfrac{2}{e}(n_{0}-2)+3\leqslant n\leqslant 2\mathfrak{c}(\Pi)+2.

Note that the Laurent expansion of fn(X;χ,H)\displaystyle f_{n}(X;\chi,H) is absolutely convergent for |X|<q\displaystyle\lvert X\rvert<q. Note also that Xfn(X;χ,H)\displaystyle Xf_{n}^{\prime}(X;\chi,H) and fn(X;χ,H)\displaystyle f_{n}(X;\chi,H) have the same type of bound for |X|=q12\displaystyle\lvert X\rvert=q^{\frac{1}{2}}. By Lemma 4.2 (1) and (4.8) we deduce

(4.17) |h~n+(χ)|=|h~n(1/2,χ)||fn(q12;χ,H)|+sup|X|=q12|Xfn(X;χ,H)|𝟙2e(n02)+3nρ(Πχ)1𝟙E(Π)(χ𝒪𝐅×)qn2(1+ρ(Πχ)n)𝟙2e(n02)+3n2𝔠(Π)+2𝟙E(Π)(χ𝒪𝐅×)qn2(2𝔠(Π)+3).\left\lvert\widetilde{h}_{n}^{+}(\chi)\right\rvert=\left\lvert\widetilde{h}_{n}(1/2,\chi)\right\rvert\leqslant\left\lvert f_{n}(q^{\frac{1}{2}};\chi,H)\right\rvert+\sideset{}{{}_{\lvert X\rvert=q^{\frac{1}{2}}}}{\sup}\left\lvert Xf_{n}^{\prime}(X;\chi,H)\right\rvert\\ \ll\mathbbm{1}_{\frac{2}{e}(n_{0}-2)+3\leqslant n\leqslant\mathfrak{\rho}(\Pi\otimes\chi)-1}\mathbbm{1}_{\mathrm{E}(\Pi)}(\chi\mid_{\mathcal{O}_{\mathbf{F}}^{\times}})\cdot q^{-\frac{n}{2}}\cdot\left(1+\mathfrak{\rho}(\Pi\otimes\chi)-n\right)\\ \ll\mathbbm{1}_{\frac{2}{e}(n_{0}-2)+3\leqslant n\leqslant 2\mathfrak{c}(\Pi)+2}\mathbbm{1}_{\mathrm{E}(\Pi)}(\chi\mid_{\mathcal{O}_{\mathbf{F}}^{\times}})\cdot q^{-\frac{n}{2}}\cdot(2\mathfrak{c}(\Pi)+3).

by Lemma 3.6 & 4.4. We conclude by summing over n\displaystyle n (for 2en\displaystyle 2\mid en) the bounds (4.15) and (4.17), taking into account the bound for 𝟙𝒜n\displaystyle\mathbbm{1}_{\mathcal{A}_{n}} given by Lemma 4.5. ∎

4.3. Negative Part

We turn to the study of a “trivial” bound of h~n(χ)\displaystyle\widetilde{h}_{n}^{-}(\chi). We introduce

(4.18) n=n(β,Π):={ξ𝒪𝐅×^|ρ(Πξ)>n,εn(ξ,H)0}.\mathcal{B}_{n}=\mathcal{B}_{n}(\beta,\Pi):=\left\{\xi\in\widehat{\mathcal{O}_{\mathbf{F}}^{\times}}\ \middle|\ \mathfrak{\rho}(\Pi\otimes\xi)>n,\ \varepsilon_{n}(\xi,H)\neq 0\right\}.
Lemma 4.7.

Recall e=e(𝐋/𝐅)\displaystyle e=e(\mathbf{L}/\mathbf{F}). We have n{ξ𝒪𝐅×^|𝔠(ξ)n/2}\displaystyle\mathcal{B}_{n}\subset\left\{\xi\in\widehat{\mathcal{O}_{\mathbf{F}}^{\times}}\ \middle|\ \mathfrak{c}(\xi)\leqslant n/2\right\} and the bound |n|qn2\displaystyle\left\lvert\mathcal{B}_{n}\right\rvert\ll q^{\frac{n}{2}}.

Proof.

For ξn\displaystyle\xi\in\mathcal{B}_{n}, the condition εn(ξ,H)0\displaystyle\varepsilon_{n}(\xi,H)\neq 0 implies 𝔠(ξ)n2\displaystyle\mathfrak{c}(\xi)\leqslant\tfrac{n}{2} by checking Lemma 3.6 case-by-case, taking into account (3.9). For example in the case 𝐋/𝐅\displaystyle\mathbf{L}/\mathbf{F} is ramified, we have nn0+1\displaystyle n\geqslant n_{0}+1 and

2𝔠(ξ)1𝔠(ξNr)max(𝔠(β),𝔠(β(ξNr)))=max(n0,n1)=n1.\displaystyle 2\mathfrak{c}(\xi)-1\leqslant\mathfrak{c}(\xi\circ{\rm Nr})\leqslant\max\left(\mathfrak{c}(\beta),\mathfrak{c}(\beta\cdot(\xi\circ{\rm Nr}))\right)=\max(n_{0},n-1)=n-1.

Therefore we get |n||𝒪𝐅×/(1+𝒫𝐅n/2)|qn2\displaystyle\left\lvert\mathcal{B}_{n}\right\rvert\leqslant\left\lvert\mathcal{O}_{\mathbf{F}}^{\times}/(1+\mathcal{P}_{\mathbf{F}}^{n/2})\right\rvert\ll q^{\frac{n}{2}}. ∎

Lemma 4.8.

We have the bound

h~n(χ)q12𝟙𝔠(χ)𝔠(Π)+n2+𝟙n2𝔠(Π)+2qn2m=12𝔠(Π)+3nqm2𝟙𝔠(χ)max(𝔠(Π),m).\displaystyle\widetilde{h}_{n}^{-}(\chi)\ll q^{-\frac{1}{2}}\cdot\mathbbm{1}_{\mathfrak{c}(\chi)\leqslant\mathfrak{c}(\Pi)+\frac{n}{2}}+\mathbbm{1}_{n\leqslant 2\mathfrak{c}(\Pi)+2}\cdot q^{-\frac{n}{2}}\sideset{}{{}_{m=1}^{2\mathfrak{c}(\Pi)+3-n}}{\sum}q^{-\frac{m}{2}}\mathbbm{1}_{\mathfrak{c}(\chi)\leqslant\max(\mathfrak{c}(\Pi),m)}.

Consequently if 𝔠(Π)>0\displaystyle\mathfrak{c}(\Pi)>0 and n0A𝔠(Π)\displaystyle n_{0}\leqslant A\mathfrak{c}(\Pi) for some constant A1\displaystyle A\geqslant 1 then we have

h~c(χ)n=2e(n02)+32en2n11|h~n(χ)|𝔠(Π)𝟙𝔠(χ)3A𝔠(Π).\displaystyle\widetilde{h}_{c}^{-}(\chi)\ll\sideset{}{{}_{\begin{subarray}{c}n=\frac{2}{e}(n_{0}-2)+3\\ 2\mid en\end{subarray}}^{2n_{1}-1}}{\sum}\left\lvert\widetilde{h}_{n}^{-}(\chi)\right\rvert\ll\mathfrak{c}(\Pi)\mathbbm{1}_{\mathfrak{c}(\chi)\leqslant 3A\mathfrak{c}(\Pi)}.
Proof.

Write χ0=χ𝒪𝐅×\displaystyle\chi_{0}=\chi\mid_{\mathcal{O}_{\mathbf{F}}^{\times}}. By definition and the Plancherel formula on 𝒪𝐅×\displaystyle\mathcal{O}_{\mathbf{F}}^{\times} we can write

h~n(χ)=Vol(𝒪𝐅×,d×t)1qnm=1χ(ϖ𝐅)mqm2Cm(n,χ01),\displaystyle\widetilde{h}_{n}^{-}(\chi)={\rm Vol}(\mathcal{O}_{\mathbf{F}}^{\times},\mathrm{d}^{\times}t)^{-1}q^{-n}\sideset{}{{}_{m=1}^{\infty}}{\sum}\chi(\varpi_{\mathbf{F}})^{m}q^{-\frac{m}{2}}\cdot C_{m}(n,\chi_{0}^{-1}),
Cm(n,χ01):=ξ𝒪𝐅×^am(ξ;n)bm(ξχ01),\displaystyle C_{m}(n,\chi_{0}^{-1}):=\sideset{}{{}_{\xi\in\widehat{\mathcal{O}_{\mathbf{F}}^{\times}}}}{\sum}a_{m}(\xi;n)\cdot b_{m}(\xi\chi_{0}^{-1}),

where we have put

am(ξ;n):=ϖ𝐅m𝒪𝐅×𝒱Π,ψ(Hn)(t)ξ1(t)d×t,bm(ξ):=𝒪𝐅×ψ(ϖ𝐅mt)ξ(t)d×t.\displaystyle a_{m}(\xi;n):=\int_{\varpi_{\mathbf{F}}^{-m}\mathcal{O}_{\mathbf{F}}^{\times}}\mathcal{VH}_{\Pi,\psi}(H_{n})(t)\xi^{-1}(t)\mathrm{d}^{\times}t,\quad b_{m}(\xi):=\int_{\mathcal{O}_{\mathbf{F}}^{\times}}\psi(-\varpi_{\mathbf{F}}^{-m}t)\xi(t)\mathrm{d}^{\times}t.

Since m1\displaystyle m\geqslant 1 the integral bm(ξ)\displaystyle b_{m}(\xi) is essentially a Gauss sum, which we can easily bound as

(4.19) bm(ξ)qm2𝟙𝔠(ξ)=m+𝟙m=1q1𝟙𝔠(ξ)=0qm2𝟙𝔠(ξ)m.b_{m}(\xi)\ll q^{-\frac{m}{2}}\cdot\mathbbm{1}_{\mathfrak{c}(\xi)=m}+\mathbbm{1}_{m=1}\cdot q^{-1}\cdot\mathbbm{1}_{\mathfrak{c}(\xi)=0}\ll q^{-\frac{m}{2}}\cdot\mathbbm{1}_{\mathfrak{c}(\xi)\leqslant m}.

The defining formula for am(ξ;n)\displaystyle a_{m}(\xi;n) makes sense for all m\displaystyle m\in\mathbb{Z}, and we have (writing X:=qs\displaystyle X:=q^{s})

mam(ξ;n)Xm=𝐅×𝒱Π,ψ(Hn)(t)ξ1(t)|t|sd×t=γ(s,Πξ,ψ)εn(ξ,H)Xn=εn(ξ,H)ε(12,Πξ,ψ)q𝔠(Πξ)2Xnρj=1dXaj1aj¯q1X,\sideset{}{{}_{m\in\mathbb{Z}}}{\sum}a_{m}(\xi;n)X^{-m}=\int_{\mathbf{F}^{\times}}\mathcal{VH}_{\Pi,\psi}(H_{n})(t)\xi^{-1}(t)\lvert t\rvert^{-s}\mathrm{d}^{\times}t=\gamma\left(s,\Pi\otimes\xi,\psi\right)\cdot\varepsilon_{n}(\xi,H)X^{n}\\ =\varepsilon_{n}(\xi,H)\varepsilon\left(\tfrac{1}{2},\Pi\otimes\xi,\psi\right)q^{\frac{\mathfrak{c}(\Pi\otimes\xi)}{2}}X^{n-\mathfrak{\rho}}\cdot\sideset{}{{}_{j=1}^{d}}{\prod}\tfrac{X-a_{j}}{1-\overline{a_{j}}q^{-1}X},

which is equivalent to (4.14). If ξE(Π)\displaystyle\xi\notin E(\Pi) then d:=d(Πξ)=0\displaystyle d:=d(\Pi\otimes\xi)=0, and we get for m1\displaystyle m\geqslant 1

(4.20) am(ξ;n)𝟙n(ξ)𝟙m+n(𝔠(Πξ))qm+n2.a_{m}(\xi;n)\ll\mathbbm{1}_{\mathcal{B}_{n}}(\xi)\cdot\mathbbm{1}_{m+n}(\mathfrak{c}(\Pi\otimes\xi))\cdot q^{\frac{m+n}{2}}.

If ξE(Π)\displaystyle\xi\in E(\Pi) then we apply the residue theorem (as |aj|{1,q12,q1}\displaystyle\lvert a_{j}\rvert\in\{1,q^{-\frac{1}{2}},q^{-1}\}) to get

am(ξ;n)\displaystyle\displaystyle a_{m}(\xi;n) =εn(ξ,H)ε(12,Πξ,ψ)q𝔠(Πξ)2|X|=q12Xn𝔠(Πξ)+m1j=1d1ajX11aj¯q1XdX2πi\displaystyle\displaystyle=\varepsilon_{n}(\xi,H)\varepsilon\left(\tfrac{1}{2},\Pi\otimes\xi,\psi\right)q^{\frac{\mathfrak{c}(\Pi\otimes\xi)}{2}}\int_{\lvert X\rvert=q^{\frac{1}{2}}}X^{n-\mathfrak{c}(\Pi\otimes\xi)+m-1}\sideset{}{{}_{j=1}^{d}}{\prod}\tfrac{1-a_{j}X^{-1}}{1-\overline{a_{j}}q^{-1}X}\frac{\mathrm{d}X}{2\pi i}
(4.21) 𝟙n(ξ)𝟙m+n(ρ(Πξ))qn+m2.\displaystyle\displaystyle\ll\mathbbm{1}_{\mathcal{B}_{n}}(\xi)\cdot\mathbbm{1}_{\geqslant m+n}(\mathfrak{\rho}(\Pi\otimes\xi))\cdot q^{\frac{n+m}{2}}.

Combining the bounds (4.19)-(4.21), Lemma 4.4 & 4.7 and 𝔠(Πξ)𝔠(Π)+3𝔠(ξ)\displaystyle\mathfrak{c}(\Pi\otimes\xi)\leqslant\mathfrak{c}(\Pi)+3\mathfrak{c}(\xi) (see [4]) we get

Cm(n,χ01)\displaystyle\displaystyle C_{m}(n,\chi_{0}^{-1}) qn2ξnE(Π)𝔠(ξχ01)m𝟙𝔠(Πξ)=m+n+qn2ξnE(Π)𝔠(ξχ01)m𝟙m+n(ρ(Πξ))\displaystyle\displaystyle\ll q^{\frac{n}{2}}\sideset{}{{}_{\begin{subarray}{c}\xi\in\mathcal{B}_{n}-E(\Pi)\\ \mathfrak{c}(\xi\chi_{0}^{-1})\leqslant m\end{subarray}}}{\sum}\mathbbm{1}_{\mathfrak{c}(\Pi\otimes\xi)=m+n}+q^{\frac{n}{2}}\sideset{}{{}_{\begin{subarray}{c}\xi\in\mathcal{B}_{n}\cap E(\Pi)\\ \mathfrak{c}(\xi\chi_{0}^{-1})\leqslant m\end{subarray}}}{\sum}\mathbbm{1}_{\geqslant m+n}(\mathfrak{\rho}(\Pi\otimes\xi))
qn2𝔠(ξ)n/2𝔠(ξχ01)m𝟙m𝔠(Π)+n2+qn2𝟙2𝔠(Π)+3(m+n)𝟙max(𝔠(Π),m)(𝔠(χ0))\displaystyle\displaystyle\ll q^{\frac{n}{2}}\sideset{}{{}_{\begin{subarray}{c}\mathfrak{c}(\xi)\leqslant n/2\\ \mathfrak{c}(\xi\chi_{0}^{-1})\leqslant m\end{subarray}}}{\sum}\mathbbm{1}_{m\leqslant\mathfrak{c}(\Pi)+\frac{n}{2}}+q^{\frac{n}{2}}\mathbbm{1}_{\leqslant 2\mathfrak{c}(\Pi)+3}(m+n)\mathbbm{1}_{\leqslant\max(\mathfrak{c}(\Pi),m)}(\mathfrak{c}(\chi_{0}))
qn2+min(n2,m)𝟙m𝔠(Π)+n2𝟙𝔠(Π)+n2(𝔠(χ0))+qn2𝟙2𝔠(Π)+3(m+n)𝟙max(𝔠(Π),m)(𝔠(χ0)),\displaystyle\displaystyle\leqslant q^{\frac{n}{2}+\min(\frac{n}{2},m)}\mathbbm{1}_{m\leqslant\mathfrak{c}(\Pi)+\frac{n}{2}}\mathbbm{1}_{\leqslant\mathfrak{c}(\Pi)+\frac{n}{2}}(\mathfrak{c}(\chi_{0}))+q^{\frac{n}{2}}\mathbbm{1}_{\leqslant 2\mathfrak{c}(\Pi)+3}(m+n)\mathbbm{1}_{\leqslant\max(\mathfrak{c}(\Pi),m)}(\mathfrak{c}(\chi_{0})),

and conclude the first bound. To derive the second bound, one simply notice n1\displaystyle n_{1} can be taken as 2A𝔠(Π)\displaystyle 2A\mathfrak{c}(\Pi), since under the condition 𝔠(Π)>0\displaystyle\mathfrak{c}(\Pi)>0 we have a(Π)𝔠(Π)\displaystyle a(\Pi)\leqslant\mathfrak{c}(\Pi) by the “moreover” part of Proposition 2.4. ∎

Remark 4.9.

The bound established in Lemma 4.8 is far from being optimal. For example in the case e=1\displaystyle e=1 for ξn\displaystyle\xi\in\mathcal{B}_{n} the typical size of 𝔠(Πξ)\displaystyle\mathfrak{c}(\Pi\otimes\xi) should be 3n/2\displaystyle 3n/2, hence the term q𝔠(Πξ)2\displaystyle q^{-\frac{\mathfrak{c}(\Pi\otimes\xi)}{2}} could be bounded as q3n4\displaystyle q^{-\frac{3n}{4}}. But even with this improvement the individual bound of h~n(χ)\displaystyle\widetilde{h}_{n}^{-}(\chi) is too weak to apply in the case n0𝔠(Π)\displaystyle n_{0}\gg\mathfrak{c}(\Pi). It would be interesting to recover the cancellation in the sum of am(ξ;n)bm(ξ)\displaystyle a_{m}(\xi;n)b_{m}(\xi) over ξ\displaystyle\xi by refining the above method.

4.4. For Unramified Characters

We turn to the unramified case χ=||𝐅s\displaystyle\chi=\lvert\cdot\rvert_{\mathbf{F}}^{s}. According to the decomposition (4.7) of H~\displaystyle\widetilde{H}, the notation H~(k;s0)\displaystyle\widetilde{H}_{\infty}(k;s_{0}), H~c(k;s0)\displaystyle\widetilde{H}_{c}(k;s_{0}) and H~n±(k;s0)\displaystyle\widetilde{H}_{n}^{\pm}(k;s_{0}) have obvious meaning and

H~(k;s0)=H~(k;s0)+±H~c±(k;s0),H~c±(k;s0)=n=2e(n02)+32en2n11H~n±(k;s0).\displaystyle\widetilde{H}(k;s_{0})=\widetilde{H}_{\infty}(k;s_{0})+\sideset{}{{}_{\pm}}{\sum}\widetilde{H}_{c}^{\pm}(k;s_{0}),\quad\widetilde{H}_{c}^{\pm}(k;s_{0})=\sideset{}{{}_{\begin{subarray}{c}n=\frac{2}{e}(n_{0}-2)+3\\ 2\mid en\end{subarray}}^{2n_{1}-1}}{\sum}\widetilde{H}_{n}^{\pm}(k;s_{0}).
Lemma 4.10.

For any k0\displaystyle k\in\mathbb{Z}_{\geqslant 0} we have the following bounds.

  • (1)

    H~(k;12)k,ϵqn1(1ϵ)\displaystyle\widetilde{H}_{\infty}(k;\tfrac{1}{2})\ll_{k,\epsilon}q^{-n_{1}(1-\epsilon)}, H~(k;12)k,ϵqn1ϵ\displaystyle\widetilde{H}_{\infty}(k;-\tfrac{1}{2})\ll_{k,\epsilon}q^{n_{1}\epsilon}.

  • (2)

    H~n+(k;±12)=0\displaystyle\widetilde{H}_{n}^{+}(k;\pm\tfrac{1}{2})=0 unless n=2n0e+e1\displaystyle n=\tfrac{2n_{0}}{e}+e-1. We have

    H~c+(k;12)k,ϵ𝐂(Π)12q(2n0e+e1)ϵ,H~c+(k;12)k,ϵ𝐂(Π)12q(2n0e+e1)(1ϵ).\displaystyle\widetilde{H}_{c}^{+}(k;\tfrac{1}{2})\ll_{k,\epsilon}\mathbf{C}(\Pi)^{\frac{1}{2}}\cdot q^{\left(\frac{2n_{0}}{e}+e-1\right)\epsilon},\quad\widetilde{H}_{c}^{+}(k;-\tfrac{1}{2})\ll_{k,\epsilon}\mathbf{C}(\Pi)^{\frac{1}{2}}\cdot q^{-\left(\frac{2n_{0}}{e}+e-1\right)(1-\epsilon)}.
Proof.

(1) By Lemma 4.1 we have H~(s)=q(12+s)n1L(12s,Π~)1\displaystyle\widetilde{H}_{\infty}(s)=q^{-\left(\frac{1}{2}+s\right)n_{1}}L\left(\tfrac{1}{2}-s,\widetilde{\Pi}\right)^{-1}. The stated bounds follow readily.

(2) By Lemma 3.6 and (4.10) we have

fn(q12;||𝐅s,H)=fn(qs+12;𝟙,H)=qn(s12)ε(s+12,Π,ψ)L(12s,Π~)L(12+s,Π){𝟙n=2n0ζ𝐅(1)χ0(1)if 𝐋/𝐅 split𝟙n=2n0e+e1ζ𝐅(1)ε(1/2,πβ,ψ)if 𝐋/𝐅 non-split.f_{n}(q^{\frac{1}{2}};\lvert\cdot\rvert_{\mathbf{F}}^{s},H)=f_{n}(q^{s+\frac{1}{2}};\mathbbm{1},H)=\\ q^{n\left(s-\frac{1}{2}\right)}\varepsilon\left(s+\tfrac{1}{2},\Pi,\psi\right)\frac{L(\tfrac{1}{2}-s,\widetilde{\Pi})}{L(\tfrac{1}{2}+s,\Pi)}\cdot\begin{cases}\mathbbm{1}_{n=2n_{0}}\cdot\zeta_{\mathbf{F}}(1)\chi_{0}(-1)&\text{if }\mathbf{L}/\mathbf{F}\text{ split}\\ \mathbbm{1}_{n=\frac{2n_{0}}{e}+e-1}\cdot\zeta_{\mathbf{F}}(1)\varepsilon(1/2,\pi_{\beta},\psi)&\text{if }\mathbf{L}/\mathbf{F}\text{ non-split}\end{cases}.

We introduce aj\displaystyle a_{j} (|aj|{1,q12,q1}\displaystyle\lvert a_{j}\rvert\in\{1,q^{-\frac{1}{2}},q^{-1}\}) as the parameters of L(s,Π)\displaystyle L(s,\Pi) so that L(12s,Π~)=P(X)1\displaystyle L(\tfrac{1}{2}-s,\widetilde{\Pi})=P(X)^{-1} with

P(X):=j=1d(Π)(1aj¯q12X),f(X):=fn(q12X;𝟙,H)=Xnρ(Π)j=1d(Π)Xajq121aj¯q12Xqn2δ\displaystyle P(X):=\sideset{}{{}_{j=1}^{d(\Pi)}}{\prod}\left(1-\overline{a_{j}}q^{-\frac{1}{2}}X\right),\quad f(X):=f_{n}(q^{\frac{1}{2}}X;\mathbbm{1},H)=X^{n-\mathfrak{\rho}(\Pi)}\sideset{}{{}_{j=1}^{d(\Pi)}}{\prod}\tfrac{X-a_{j}q^{-\frac{1}{2}}}{1-\overline{a_{j}}q^{-\frac{1}{2}}X}\cdot q^{-\frac{n}{2}}\delta

for some |δ|=1\displaystyle\lvert\delta\rvert=1. We then rewrite, with X=qs\displaystyle X=q^{s}

H~n+(||𝐅s)=f+(qs)L(12s,Π~)ζ𝐅(12+s)=(1q12X1)P(X)f+(X).\displaystyle\widetilde{H}_{n}^{+}(\lvert\cdot\rvert_{\mathbf{F}}^{s})=\tfrac{f_{+}(q^{s})}{L(\tfrac{1}{2}-s,\widetilde{\Pi})\zeta_{\mathbf{F}}(\tfrac{1}{2}+s)}=(1-q^{-\frac{1}{2}}X^{-1})P(X)f_{+}(X).

The bounds at s=1/2\displaystyle s=-1/2 then follows readily from Lemma 4.2 (1) via the bounds

sup|z|=1|Dkf(q12z)|k,ϵ𝐂(Π)12qn(1ϵ),Dk((1q12X1)P(X))X=q12k1.\displaystyle\sideset{}{{}_{\lvert z\rvert=1}}{\sup}\left\lvert D^{k}f(q^{-\frac{1}{2}}z)\right\rvert\ll_{k,\epsilon}\mathbf{C}(\Pi)^{\frac{1}{2}}q^{-n(1-\epsilon)},\quad D^{k}\left((1-q^{-\frac{1}{2}}X^{-1})P(X)\right)\mid_{X=q^{-\frac{1}{2}}}\ll_{k}1.

We now consider the bounds at s=1/2\displaystyle s=1/2. If n=2n0e+e1ρ(Π)\displaystyle n=\tfrac{2n_{0}}{e}+e-1\geqslant\mathfrak{\rho}(\Pi), then f+(X)=f(X)\displaystyle f_{+}(X)=f(X), implying

H~n+(||𝐅s)=qn2δXn𝔠(Π)(1q12X1)j=1d(Π)(1ajq12X1),\displaystyle\widetilde{H}_{n}^{+}(\lvert\cdot\rvert_{\mathbf{F}}^{s})=q^{-\frac{n}{2}}\delta\cdot X^{n-\mathfrak{c}(\Pi)}(1-q^{-\frac{1}{2}}X^{-1})\sideset{}{{}_{j=1}^{d(\Pi)}}{\prod}\left(1-a_{j}q^{-\frac{1}{2}}X^{-1}\right),
DkH~n+(||𝐅s)s=1/2k,ϵ𝐂(Π)12qnϵ.\displaystyle D^{k}\widetilde{H}_{n}^{+}(\lvert\cdot\rvert_{\mathbf{F}}^{s})\mid_{s=1/2}\ll_{k,\epsilon}\mathbf{C}(\Pi)^{-\frac{1}{2}}q^{n\epsilon}.

If n=2n0e+e1<ρ(Π)=:ρ\displaystyle n=\tfrac{2n_{0}}{e}+e-1<\mathfrak{\rho}(\Pi)=:\mathfrak{\rho}, then Q(X):=P(X)f(X)\displaystyle Q(X):=P(X)f(X) has highest term Xm\displaystyle X^{m} with m<d(Π)=degP\displaystyle m<d(\Pi)=\deg P. We also note that f+(X)=0\displaystyle f_{+}(X)=0 unless d:=d(Π)1\displaystyle d:=d(\Pi)\geqslant 1. We apply Lemma 4.2 (2) distinguishing cases:

(i) d=1\displaystyle d=1. Necessarily we have c:=c(Π)2\displaystyle c:=c(\Pi)\geqslant 2, nc0\displaystyle n-c\leqslant 0. We get and conclude the stated bounds by

P(X)f+(X)=a1¯cnδqc2(1|a1|2q1)𝐂(Π)12.\displaystyle P(X)f_{+}(X)=\overline{a_{1}}^{c-n}\delta\cdot q^{-\frac{c}{2}}(1-\lvert a_{1}\rvert^{2}q^{-1})\ll\mathbf{C}(\Pi)^{-\frac{1}{2}}.

(ii) d=2\displaystyle d=2. Necessarily Π=Stχχ2\displaystyle\Pi=\mathrm{St}_{\chi}\boxplus\chi^{-2} for some unramified unitary character χ\displaystyle\chi. We have c=1\displaystyle c=1 and nc1\displaystyle n-c\leqslant 1, and may assume |a1|=q12\displaystyle\lvert a_{1}\rvert=q^{-\frac{1}{2}} and |a2|=1\displaystyle\lvert a_{2}\rvert=1. We get the formula

P(X)f+(X)\displaystyle\displaystyle P(X)f_{+}(X) =a1¯cnδqc2(1|a1|2q1)(1a1¯a2q1)1a2¯a1¯1(1a2¯q12X)+\displaystyle\displaystyle=\overline{a_{1}}^{c-n}\delta\cdot q^{-\frac{c}{2}}\frac{(1-\lvert a_{1}\rvert^{2}q^{-1})(1-\overline{a_{1}}a_{2}q^{-1})}{1-\overline{a_{2}}\overline{a_{1}}^{-1}}(1-\overline{a_{2}}q^{-\frac{1}{2}}X)+
a2¯cnδqc2(1|a2|2q1)(1a2¯a1q1)1a1¯a2¯1(1a1¯q12X)\displaystyle\displaystyle\quad\overline{a_{2}}^{c-n}\delta\cdot q^{-\frac{c}{2}}\frac{(1-\lvert a_{2}\rvert^{2}q^{-1})(1-\overline{a_{2}}a_{1}q^{-1})}{1-\overline{a_{1}}\overline{a_{2}}^{-1}}(1-\overline{a_{1}}q^{-\frac{1}{2}}X)

and conclude the stated bounds by

(4.22) Dk(P(X)f+(X))X=q12k1.D^{k}\left(P(X)f_{+}(X)\right)\mid_{X=q^{\frac{1}{2}}}\ll_{k}1.

(iii) d=3\displaystyle d=3. Necessarily Π=μ1μ2μ3\displaystyle\Pi=\mu_{1}\boxplus\mu_{2}\boxplus\mu_{3} for some unramified unitary characters μi\displaystyle\mu_{i}. We have c=0\displaystyle c=0 and n2\displaystyle n\leqslant 2, and may assume |aj|=1\displaystyle\lvert a_{j}\rvert=1 with a1a2a3=1\displaystyle a_{1}a_{2}a_{3}=1. It is easy to see Cj,k1\displaystyle C_{j,k}\ll 1. Therefore (4.22) still holds and we conclude the stated bounds in the same way. ∎

Lemma 4.11.

For any k0\displaystyle k\in\mathbb{Z}_{\geqslant 0} we have the following bounds

H~n(k;12)k,ϵ𝐂(Π)ϵ,H~n(k;12)k,ϵ(𝐂(Π)2qn2)1+ϵ.\displaystyle\widetilde{H}_{n}^{-}(k;\tfrac{1}{2})\ll_{k,\epsilon}\mathbf{C}(\Pi)^{\epsilon},\quad\widetilde{H}_{n}^{-}(k;-\tfrac{1}{2})\ll_{k,\epsilon}\left(\mathbf{C}(\Pi)^{2}q^{\frac{n}{2}}\right)^{1+\epsilon}.

Consequently if 𝔠(Π)>0\displaystyle\mathfrak{c}(\Pi)>0 and n0A𝔠(Π)\displaystyle n_{0}\leqslant A\mathfrak{c}(\Pi) for some constant A1\displaystyle A\geqslant 1 then we have

H~c(k;δ12)n=2e(n02)+32en2n11|H~n(k;δ12)|ϵ{𝐂(Π)ϵif δ=1𝐂(Π)(2+2Ae1)(1+ϵ)if δ=1.\displaystyle\widetilde{H}_{c}^{-}(k;\delta\cdot\tfrac{1}{2})\leqslant\sideset{}{{}_{\begin{subarray}{c}n=\frac{2}{e}(n_{0}-2)+3\\ 2\mid en\end{subarray}}^{2n_{1}-1}}{\sum}\left\lvert\widetilde{H}_{n}^{-}(k;\delta\cdot\tfrac{1}{2})\right\rvert\ll_{\epsilon}\begin{cases}\mathbf{C}(\Pi)^{\epsilon}&\text{if }\delta=1\\ \mathbf{C}(\Pi)^{\left(2+2Ae^{-1}\right)(1+\epsilon)}&\text{if }\delta=-1\end{cases}.
Proof.

With similar argument as in the proof of Lemma 4.8 we get

h~n(||𝐅s)=qnm=1qm(12s)Cm(n),Cm(n)qn2+min(m,n2)𝟙m𝔠(Π)+n2+qn2𝟙m+n2𝔠(Π)+3.\displaystyle\widetilde{h}_{n}^{-}(\lvert\cdot\rvert_{\mathbf{F}}^{s})=q^{-n}\sideset{}{{}_{m=1}^{\infty}}{\sum}q^{m\left(\frac{1}{2}-s\right)}C_{m}(n),\quad C_{m}(n)\ll q^{\frac{n}{2}+\min(m,\frac{n}{2})}\mathbbm{1}_{m\leqslant\mathfrak{c}(\Pi)+\frac{n}{2}}+q^{\frac{n}{2}}\mathbbm{1}_{m+n\leqslant 2\mathfrak{c}(\Pi)+3}.

The stated bounds follow readily. ∎

5. Dual Weight Functions: Split and Unramified Cases

5.1. First Quadratic Elementary Functions

Definition 5.1.

For any n0\displaystyle n\in\mathbb{Z}_{\geqslant 0} we define the first quadratic elementary functions FnCc(𝐅×)\displaystyle F_{n}\in{\rm C}_{c}^{\infty}(\mathbf{F}^{\times}) by

Fn(y2):=𝟙v(y)=n±ψ(±y),\displaystyle F_{n}(y^{2}):=\mathbbm{1}_{v(y)=-n}\cdot\sideset{}{{}_{\pm}}{\sum}\psi(\pm y),

and are supported in the subset of square elements of 𝐅×\displaystyle\mathbf{F}^{\times}.

Definition 5.2.

Let η0\displaystyle\eta_{0} be the character of 𝐅×\displaystyle\mathbf{F}^{\times} associated with the (ramified) quadratic extension 𝐅[ϖ𝐅]/𝐅\displaystyle\mathbf{F}[\sqrt{-\varpi_{\mathbf{F}}}]/\mathbf{F}. Explicitly it is given by the following formula for all n\displaystyle n\in\mathbb{Z}

η0(ϖ𝐅nu)={1if u(𝒪𝐅×)21if u𝒪𝐅×(𝒪𝐅×)2.\displaystyle\eta_{0}(\varpi_{\mathbf{F}}^{n}u)=\begin{cases}1&\text{if }u\in(\mathcal{O}_{\mathbf{F}}^{\times})^{2}\\ -1&\text{if }u\in\mathcal{O}_{\mathbf{F}}^{\times}-(\mathcal{O}_{\mathbf{F}}^{\times})^{2}\end{cases}.

Denote by τ0=τ(η0,ψ;ϖ𝐅)\displaystyle\tau_{0}=\tau(\eta_{0},\psi;\varpi_{\mathbf{F}}) the quadratic Gauss sum given by

τ0:=𝒪𝐅×ψ(uϖ𝐅)η0(u)du=𝒪𝐅ψ(u2ϖ𝐅)du.\displaystyle\tau_{0}:=\int_{\mathcal{O}_{\mathbf{F}}^{\times}}\psi\left(\tfrac{u}{\varpi_{\mathbf{F}}}\right)\eta_{0}(u)\mathrm{d}u=\int_{\mathcal{O}_{\mathbf{F}}}\psi\left(\tfrac{u^{2}}{\varpi_{\mathbf{F}}}\right)\mathrm{d}u.
Lemma 5.3.

Let χ\displaystyle\chi be a quasi-character of 𝐅×\displaystyle\mathbf{F}^{\times}. We have

𝐅×Fn(y)χ(y)d×y={𝟙𝔠(χ2)=nζ𝐅(1)γ(1,χ2,ψ)if n2𝟙𝔠(χ2)=1ζ𝐅(1)γ(1,χ2,ψ)𝟙𝔠(χ2)=0ζ𝐅(1)q1χ(ϖ𝐅)2if n=1𝟙𝔠(χ2)=0if n=0.\displaystyle\int_{\mathbf{F}^{\times}}F_{n}(y)\chi(y)\mathrm{d}^{\times}y=\begin{cases}\mathbbm{1}_{\mathfrak{c}(\chi^{2})=n}\cdot\zeta_{\mathbf{F}}(1)\gamma(1,\chi^{-2},\psi)&\text{if }n\geqslant 2\\ \mathbbm{1}_{\mathfrak{c}(\chi^{2})=1}\cdot\zeta_{\mathbf{F}}(1)\gamma(1,\chi^{-2},\psi)-\mathbbm{1}_{\mathfrak{c}(\chi^{2})=0}\cdot\zeta_{\mathbf{F}}(1)q^{-1}\chi(\varpi_{\mathbf{F}})^{-2}&\text{if }n=1\\ \mathbbm{1}_{\mathfrak{c}(\chi^{2})=0}&\text{if }n=0\end{cases}.

Note that we can replace the condition 𝔠(χ2)=n\displaystyle\mathfrak{c}(\chi^{2})=n with 𝔠(χ)=n\displaystyle\mathfrak{c}(\chi)=n if n1\displaystyle n\geqslant 1.

Proof.

The case for n=0\displaystyle n=0 is simple and omitted. For n1\displaystyle n\geqslant 1 with the change of variables yy2\displaystyle y\to y^{2}

𝐅×Fn(y)χ(y)d×y=ϖ𝐅n𝒪𝐅×ψ(y)χ2(y)d×y,\displaystyle\int_{\mathbf{F}^{\times}}F_{n}(y)\chi(y)\mathrm{d}^{\times}y=\int_{\varpi_{\mathbf{F}}^{-n}\mathcal{O}_{\mathbf{F}}^{\times}}\psi(y)\chi^{2}(y)\mathrm{d}^{\times}y,

the desired formula follows readily from [5, Exercise 23.5]. ∎

Corollary 5.4.

Let na(Π)(2)\displaystyle n\geqslant a(\Pi)(\geqslant 2), the stability barrier of Π\displaystyle\Pi (see Proposition 2.4). We have

𝒱Π,ψ(Fn)(y)=𝟙ϖ𝐅n𝒪𝐅×(y){q3n2ψ(4y)if 2nτ0q3n2ψ(4y)η0(4y)if 2n.\displaystyle\mathcal{VH}_{\Pi,\psi}(F_{n})(y)=\mathbbm{1}_{\varpi_{\mathbf{F}}^{-n}\mathcal{O}_{\mathbf{F}}^{\times}}(y)\cdot\begin{cases}q^{\frac{3n}{2}}\psi(4y)&\text{if }2\mid n\\ \tau_{0}q^{\lceil\frac{3n}{2}\rceil}\psi(4y)\eta_{0}(4y)&\text{if }2\nmid n\end{cases}.
Proof.

By the local functional equation, Lemma 5.3 and Proposition 2.4 we have for any χ𝒪𝐅×^\displaystyle\chi\in\widehat{\mathcal{O}_{\mathbf{F}}^{\times}}

𝐅×𝒱Π,ψ(Fn)(y)χ(y)1|y|sd×y=ζ𝐅(1)𝟙𝔠(χ)=nγ(s,χ,ψ)3γ(12s,χ2,ψ)=ζ𝐅(1)𝟙𝔠(χ)=nγ(1,χ,ψ)3γ(1,χ2,ψ)q3nns.\int_{\mathbf{F}^{\times}}\mathcal{VH}_{\Pi,\psi}(F_{n})(y)\chi(y)^{-1}\lvert y\rvert^{-s}\mathrm{d}^{\times}y=\zeta_{\mathbf{F}}(1)\mathbbm{1}_{\mathfrak{c}(\chi)=n}\cdot\gamma(s,\chi,\psi)^{3}\gamma(1-2s,\chi^{-2},\psi)\\ =\zeta_{\mathbf{F}}(1)\mathbbm{1}_{\mathfrak{c}(\chi)=n}\gamma(1,\chi,\psi)^{3}\gamma(1,\chi^{-2},\psi)\cdot q^{3n-ns}.

Therefore the support of 𝒱Π,ψ(Fn)\displaystyle\mathcal{VH}_{\Pi,\psi}(F_{n}) is contained in ϖ𝐅n𝒪𝐅×\displaystyle\varpi_{\mathbf{F}}^{-n}\mathcal{O}_{\mathbf{F}}^{\times}. Applying [5, Exercise 23.5] we find

ζ𝐅(1)𝟙𝔠(χ)=nγ(1,χ,ψ)3γ(1,χ2,ψ)q3n=𝟙𝔠(χ)=nq3n(𝒪𝐅×)4ψ(t1+t2+t3+t4ϖ𝐅n)χ1(t1t2t3ϖ𝐅nt42)dt,\displaystyle\zeta_{\mathbf{F}}(1)\mathbbm{1}_{\mathfrak{c}(\chi)=n}\gamma(1,\chi,\psi)^{3}\gamma(1,\chi^{-2},\psi)\cdot q^{3n}=\mathbbm{1}_{\mathfrak{c}(\chi)=n}\cdot q^{3n}\int_{(\mathcal{O}_{\mathbf{F}}^{\times})^{4}}\psi\left(\tfrac{t_{1}+t_{2}+t_{3}+t_{4}}{\varpi_{\mathbf{F}}^{n}}\right)\chi^{-1}\left(\tfrac{t_{1}t_{2}t_{3}}{\varpi_{\mathbf{F}}^{n}t_{4}^{2}}\right)\mathrm{d}\vec{t},

where we have written dt:=dt1dt2dt3dt4\displaystyle\mathrm{d}\vec{t}:=\mathrm{d}t_{1}\mathrm{d}t_{2}\mathrm{d}t_{3}\mathrm{d}t_{4} for simplicity. The Fourier inversion on 𝒪𝐅×\displaystyle\mathcal{O}_{\mathbf{F}}^{\times} yields for y𝒪𝐅×\displaystyle y\in\mathcal{O}_{\mathbf{F}}^{\times}

𝒱Π,ψ(Fn)(yϖ𝐅n)=ζ𝐅(1)q3nχ𝒪𝐅×^𝔠(χ)=n(𝒪𝐅×)4ψ(t1+t2+t3+t4ϖ𝐅n)χ(t42yt1t2t3)dt=ζ𝐅(1)q3n(𝒪𝐅×)4ψ(t1+t2+t3+t4ϖ𝐅n)(|(𝒪𝐅/𝒫𝐅n)×|𝟙1+𝒫𝐅n|(𝒪𝐅/𝒫𝐅n1)×|𝟙1+𝒫𝐅n1)(t42yt1t2t3)dt=q3n(𝒪𝐅×)3ψ(t2+t3+t4+t21t31t42yϖ𝐅n)dt2dt3dt4q4n1(𝒪𝐅×)3×𝒫𝐅n1ψ(t2+t3+t4+t21t31t42yϖ𝐅n)ψ(t21t31t42yuϖ𝐅n)dt2dt3dt4du.\mathcal{VH}_{\Pi,\psi}(F_{n})\left(\tfrac{y}{\varpi_{\mathbf{F}}^{n}}\right)=\zeta_{\mathbf{F}}(1)q^{3n}\sideset{}{{}_{\begin{subarray}{c}\chi\in\widehat{\mathcal{O}_{\mathbf{F}}^{\times}}\\ \mathfrak{c}(\chi)=n\end{subarray}}}{\sum}\int_{(\mathcal{O}_{\mathbf{F}}^{\times})^{4}}\psi\left(\tfrac{t_{1}+t_{2}+t_{3}+t_{4}}{\varpi_{\mathbf{F}}^{n}}\right)\chi\left(\tfrac{t_{4}^{2}y}{t_{1}t_{2}t_{3}}\right)\mathrm{d}\vec{t}\\ =\zeta_{\mathbf{F}}(1)q^{3n}\int_{(\mathcal{O}_{\mathbf{F}}^{\times})^{4}}\psi\left(\tfrac{t_{1}+t_{2}+t_{3}+t_{4}}{\varpi_{\mathbf{F}}^{n}}\right)\left(\left\lvert\left(\mathcal{O}_{\mathbf{F}}/\mathcal{P}_{\mathbf{F}}^{n}\right)^{\times}\right\rvert\mathbbm{1}_{1+\mathcal{P}_{\mathbf{F}}^{n}}-\left\lvert\left(\mathcal{O}_{\mathbf{F}}/\mathcal{P}_{\mathbf{F}}^{n-1}\right)^{\times}\right\rvert\mathbbm{1}_{1+\mathcal{P}_{\mathbf{F}}^{n-1}}\right)\left(\tfrac{t_{4}^{2}y}{t_{1}t_{2}t_{3}}\right)\mathrm{d}\vec{t}\\ =q^{3n}\int_{(\mathcal{O}_{\mathbf{F}}^{\times})^{3}}\psi\left(\tfrac{t_{2}+t_{3}+t_{4}+t_{2}^{-1}t_{3}^{-1}t_{4}^{2}y}{\varpi_{\mathbf{F}}^{n}}\right)\mathrm{d}t_{2}\mathrm{d}t_{3}\mathrm{d}t_{4}-\\ q^{4n-1}\int_{(\mathcal{O}_{\mathbf{F}}^{\times})^{3}\times\mathcal{P}_{\mathbf{F}}^{n-1}}\psi\left(\tfrac{t_{2}+t_{3}+t_{4}+t_{2}^{-1}t_{3}^{-1}t_{4}^{2}y}{\varpi_{\mathbf{F}}^{n}}\right)\psi\left(\tfrac{t_{2}^{-1}t_{3}^{-1}t_{4}^{2}yu}{\varpi_{\mathbf{F}}^{n}}\right)\mathrm{d}t_{2}\mathrm{d}t_{3}\mathrm{d}t_{4}\mathrm{d}u.

The last integral is vanishing because 𝔠(ψ)=0\displaystyle\mathfrak{c}(\psi)=0. Re-numbering the variables we get

𝒱Π,ψ(Fn)(yϖ𝐅n)=q3n(𝒪𝐅×)3ψ(t1+t2+t3+t11t21t32yϖ𝐅n)dt1dt2dt3.\displaystyle\mathcal{VH}_{\Pi,\psi}(F_{n})\left(\tfrac{y}{\varpi_{\mathbf{F}}^{n}}\right)=q^{3n}\int_{(\mathcal{O}_{\mathbf{F}}^{\times})^{3}}\psi\left(\tfrac{t_{1}+t_{2}+t_{3}+t_{1}^{-1}t_{2}^{-1}t_{3}^{2}y}{\varpi_{\mathbf{F}}^{n}}\right)\mathrm{d}t_{1}\mathrm{d}t_{2}\mathrm{d}t_{3}.

Performing the level n2\displaystyle\lceil\tfrac{n}{2}\rceil regularization to dtj\displaystyle\mathrm{d}t_{j} we see that the non-vanishing of the above integral implies

t1t32yt1t2𝒫𝐅n2,t2t32yt1t2𝒫𝐅n2,t3+2t32yt1t2𝒫𝐅n2.\displaystyle t_{1}-\tfrac{t_{3}^{2}y}{t_{1}t_{2}}\in\mathcal{P}_{\mathbf{F}}^{\lfloor\frac{n}{2}\rfloor},\quad t_{2}-\tfrac{t_{3}^{2}y}{t_{1}t_{2}}\in\mathcal{P}_{\mathbf{F}}^{\lfloor\frac{n}{2}\rfloor},\quad t_{3}+\tfrac{2t_{3}^{2}y}{t_{1}t_{2}}\in\mathcal{P}_{\mathbf{F}}^{\lfloor\frac{n}{2}\rfloor}.

Writing a=t32yt1t2𝒪𝐅×\displaystyle a=\tfrac{t_{3}^{2}y}{t_{1}t_{2}}\in\mathcal{O}_{\mathbf{F}}^{\times}, we see that the above conditions imply a4y(mod𝒫𝐅n2)\displaystyle a\equiv 4y\pmod{\mathcal{P}_{\mathbf{F}}^{\lfloor\frac{n}{2}\rfloor}}, hence

t14y(1+𝒫𝐅n2),t24y(1+𝒫𝐅n2),t38y(1+𝒫𝐅n2).\displaystyle t_{1}\in 4y\left(1+\mathcal{P}_{\mathbf{F}}^{\lfloor\frac{n}{2}\rfloor}\right),\quad t_{2}\in 4y\left(1+\mathcal{P}_{\mathbf{F}}^{\lfloor\frac{n}{2}\rfloor}\right),\quad t_{3}\in-8y\left(1+\mathcal{P}_{\mathbf{F}}^{\lfloor\frac{n}{2}\rfloor}\right).

The change of variables t1=4y(1+u1)\displaystyle t_{1}=4y(1+u_{1}), t2=4y(1+u2)\displaystyle t_{2}=4y(1+u_{2}) and t3=8y(1+u3)\displaystyle t_{3}=-8y(1+u_{3}) with uj𝒫𝐅n2\displaystyle u_{j}\in\mathcal{P}_{\mathbf{F}}^{\lfloor\frac{n}{2}\rfloor} gives

𝒱Π,ψ(Fn)(yϖ𝐅n)=q3nuj𝒫𝐅n2ψ(4y(u1+u22u3)+4(1+u3)2y(1+u1)(1+u2)ϖ𝐅n)du1du2du3=q3nψ(4yϖ𝐅n)uj𝒫𝐅n2ψ(4yϖ𝐅n(u12+u22+u32+u1u22u1u32u2u3))du1du2du3=q3nψ(4yϖ𝐅n)uj𝒫𝐅n2ψ(4yϖ𝐅n(u12+u22+u1u2u1u3))du1du2du3=q3nn2ψ(4yϖ𝐅n)uj𝒫𝐅n2ψ(4yϖ𝐅n(u12+u22+u1u2))𝟙𝒫𝐅n2(u1)du1du2=q2nψ(4yϖ𝐅n)𝒫𝐅n2ψ(4yϖ𝐅nu22)du2,\mathcal{VH}_{\Pi,\psi}(F_{n})\left(\tfrac{y}{\varpi_{\mathbf{F}}^{n}}\right)=q^{3n}\int_{u_{j}\in\mathcal{P}_{\mathbf{F}}^{\lfloor\frac{n}{2}\rfloor}}\psi\left(\frac{4y(u_{1}+u_{2}-2u_{3})+\frac{4(1+u_{3})^{2}y}{(1+u_{1})(1+u_{2})}}{\varpi_{\mathbf{F}}^{n}}\right)\mathrm{d}u_{1}\mathrm{d}u_{2}\mathrm{d}u_{3}\\ =q^{3n}\psi\left(\tfrac{4y}{\varpi_{\mathbf{F}}^{n}}\right)\int_{u_{j}\in\mathcal{P}_{\mathbf{F}}^{\lfloor\frac{n}{2}\rfloor}}\psi\left(\tfrac{4y}{\varpi_{\mathbf{F}}^{n}}\left(u_{1}^{2}+u_{2}^{2}+u_{3}^{2}+u_{1}u_{2}-2u_{1}u_{3}-2u_{2}u_{3}\right)\right)\mathrm{d}u_{1}\mathrm{d}u_{2}\mathrm{d}u_{3}\\ =q^{3n}\psi\left(\tfrac{4y}{\varpi_{\mathbf{F}}^{n}}\right)\int_{u_{j}\in\mathcal{P}_{\mathbf{F}}^{\lfloor\frac{n}{2}\rfloor}}\psi\left(\tfrac{4y}{\varpi_{\mathbf{F}}^{n}}\left(u_{1}^{2}+u_{2}^{2}+u_{1}u_{2}-u_{1}u_{3}\right)\right)\mathrm{d}u_{1}\mathrm{d}u_{2}\mathrm{d}u_{3}\\ =q^{3n-\lfloor\frac{n}{2}\rfloor}\psi\left(\tfrac{4y}{\varpi_{\mathbf{F}}^{n}}\right)\int_{u_{j}\in\mathcal{P}_{\mathbf{F}}^{\lfloor\frac{n}{2}\rfloor}}\psi\left(\tfrac{4y}{\varpi_{\mathbf{F}}^{n}}\left(u_{1}^{2}+u_{2}^{2}+u_{1}u_{2}\right)\right)\mathbbm{1}_{\mathcal{P}_{\mathbf{F}}^{\lceil\frac{n}{2}\rceil}}(u_{1})\mathrm{d}u_{1}\mathrm{d}u_{2}\\ =q^{2n}\psi\left(\tfrac{4y}{\varpi_{\mathbf{F}}^{n}}\right)\int_{\mathcal{P}_{\mathbf{F}}^{\lfloor\frac{n}{2}\rfloor}}\psi\left(\tfrac{4y}{\varpi_{\mathbf{F}}^{n}}u_{2}^{2}\right)\mathrm{d}u_{2},

where we made the change of variables u2u2+u3\displaystyle u_{2}\mapsto u_{2}+u_{3} in the third line. The desired formula follows. ∎

Remark 5.5.

If we compute the Mellin transform of 𝒱Π,ψ(Fn)(y)\displaystyle\mathcal{VH}_{\Pi,\psi}(F_{n})(y) with the given formula in Corollary 5.4, we get in the case 2n2\displaystyle 2\mid n\geqslant 2 and for 𝔠(χ)=n\displaystyle\mathfrak{c}(\chi)=n an interesting equation

χ(4)γ(12,χ1,ψ)=γ(12,χ,ψ)3γ(12,χ2,ψ).\displaystyle\chi(4)\gamma\left(\tfrac{1}{2},\chi^{-1},\psi\right)=\gamma\left(\tfrac{1}{2},\chi,\psi\right)^{3}\gamma\left(\tfrac{1}{2},\chi^{-2},\psi\right).
Corollary 5.6.

Suppose Π=μ1μ2μ3\displaystyle\Pi=\mu_{1}\boxplus\mu_{2}\boxplus\mu_{3} with 𝔠(μj)=0\displaystyle\mathfrak{c}(\mu_{j})=0 and μ1μ2μ3=𝟙\displaystyle\mu_{1}\mu_{2}\mu_{3}=\mathbbm{1}. Let Ei(y):=μi1(y)|y|𝟙𝒪𝐅(y)\displaystyle E_{i}(y):=\mu_{i}^{-1}(y)\lvert y\rvert\mathbbm{1}_{\mathcal{O}_{\mathbf{F}}}(y) and fi:=(1μi(ϖ𝐅)𝔱(ϖ𝐅)).Ei\displaystyle f_{i}:=(1-\mu_{i}(\varpi_{\mathbf{F}})\mathfrak{t}(\varpi_{\mathbf{F}})).E_{i}. For f,gL1(𝐅×)\displaystyle f,g\in{\rm L}^{1}(\mathbf{F}^{\times}) define fg(y):=𝐅×f(yt1)g(t)d×t\displaystyle f*g(y):=\int_{\mathbf{F}^{\times}}f(yt^{-1})g(t)\mathrm{d}^{\times}t. We have

𝒱Π,ψ(F0)(y)\displaystyle\displaystyle\mathcal{VH}_{\Pi,\psi}(F_{0})(y) =(f1f2f3)(y)+τ0q2η0(y)𝟙ϖ𝐅3𝒪𝐅×(y),\displaystyle\displaystyle=(f_{1}*f_{2}*f_{3})(y)+\tau_{0}q^{2}\cdot\eta_{0}(-y)\mathbbm{1}_{\varpi_{\mathbf{F}}^{-3}\mathcal{O}_{\mathbf{F}}^{\times}}(y),
𝒱Π,ψ(F1)(y)\displaystyle\displaystyle\mathcal{VH}_{\Pi,\psi}(F_{1})(y) =q1(f1f2f3)(ϖ𝐅2y)ζ𝐅(1)𝟙ϖ𝐅1𝒪𝐅×(y)\displaystyle\displaystyle=-q^{-1}\cdot(f_{1}*f_{2}*f_{3})(\varpi_{\mathbf{F}}^{-2}y)-\zeta_{\mathbf{F}}(1)\cdot\mathbbm{1}_{\varpi_{\mathbf{F}}^{-1}\mathcal{O}_{\mathbf{F}}^{\times}}(y)
+τ0q𝟙ϖ𝐅1𝒪𝐅×(y)(ϖ𝐅1𝒪𝐅×)2ψ(t1+t2t1t24y)η0(4yt1t2)dt1dt2.\displaystyle\displaystyle\quad+\tau_{0}q\mathbbm{1}_{\varpi_{\mathbf{F}}^{-1}\mathcal{O}_{\mathbf{F}}^{\times}}(y)\int_{(\varpi_{\mathbf{F}}^{-1}\mathcal{O}_{\mathbf{F}}^{\times})^{2}}\psi\left(t_{1}+t_{2}-\tfrac{t_{1}t_{2}}{4y}\right)\eta_{0}\left(\tfrac{4y}{t_{1}t_{2}}\right)\mathrm{d}t_{1}\mathrm{d}t_{2}.
Proof.

By the local functional equation and Lemma 5.3 we have for any χ𝐅×^\displaystyle\chi\in\widehat{\mathbf{F}^{\times}}

𝐅×𝒱Π,ψ(F0)(y)χ(y)1|y|sd×y={i=13L(1s,μi1χ1)L(s,μiχ)if 𝔠(χ)=0q3(12s)γ(12,χ,ψ)3if 𝔠(χη0)=00otherwise.\displaystyle\int_{\mathbf{F}^{\times}}\mathcal{VH}_{\Pi,\psi}(F_{0})(y)\chi(y)^{-1}\lvert y\rvert^{-s}\mathrm{d}^{\times}y=\begin{cases}\sideset{}{{}_{i=1}^{3}}{\prod}\tfrac{L(1-s,\mu_{i}^{-1}\chi^{-1})}{L(s,\mu_{i}\chi)}&\text{if }\mathfrak{c}(\chi)=0\\ q^{3(\frac{1}{2}-s)}\gamma\left(\tfrac{1}{2},\chi,\psi\right)^{3}&\text{if }\mathfrak{c}(\chi\eta_{0})=0\\ 0&\text{otherwise}\end{cases}.

Therefore we can write 𝒱Π,ψ(F0)=𝒱Π,ψ(F0)0+𝒱Π,ψ(F0)1\displaystyle\mathcal{VH}_{\Pi,\psi}(F_{0})=\mathcal{VH}_{\Pi,\psi}(F_{0})_{0}+\mathcal{VH}_{\Pi,\psi}(F_{0})_{1} with the properties

𝒱Π,ψ(F0)0(yδ)=𝒱Π,ψ(F0)0(y),𝒱Π,ψ(F0)1(yδ)=𝒱Π,ψ(F0)1(y)η0(δ),y𝐅×,δ𝒪𝐅×;\displaystyle\mathcal{VH}_{\Pi,\psi}(F_{0})_{0}(y\delta)=\mathcal{VH}_{\Pi,\psi}(F_{0})_{0}(y),\quad\mathcal{VH}_{\Pi,\psi}(F_{0})_{1}(y\delta)=\mathcal{VH}_{\Pi,\psi}(F_{0})_{1}(y)\eta_{0}(\delta),\quad\forall y\in\mathbf{F}^{\times},\delta\in\mathcal{O}_{\mathbf{F}}^{\times};
𝐅×𝒱Π,ψ(F0)0(y)|y|sd×y=i=13L(1s,μi1)L(s,μi)=i=13𝐅×fi(y)|y|sd×y,\displaystyle\int_{\mathbf{F}^{\times}}\mathcal{VH}_{\Pi,\psi}(F_{0})_{0}(y)\lvert y\rvert^{-s}\mathrm{d}^{\times}y=\sideset{}{{}_{i=1}^{3}}{\prod}\tfrac{L(1-s,\mu_{i}^{-1})}{L(s,\mu_{i})}=\sideset{}{{}_{i=1}^{3}}{\prod}\int_{\mathbf{F}^{\times}}f_{i}(y)\lvert y\rvert^{-s}\mathrm{d}^{\times}y,
𝐅×𝒱Π,ψ(F0)0(y)η0(y)|y|sd×y=q3(12s)γ(12,η0,ψ)3=η0(1)τ0q2ϖ𝐅3𝒪𝐅×|y|sd×y.\displaystyle\int_{\mathbf{F}^{\times}}\mathcal{VH}_{\Pi,\psi}(F_{0})_{0}(y)\eta_{0}(y)\lvert y\rvert^{-s}\mathrm{d}^{\times}y=q^{3(\frac{1}{2}-s)}\gamma\left(\tfrac{1}{2},\eta_{0},\psi\right)^{3}=\eta_{0}(-1)\tau_{0}q^{2}\cdot\int_{\varpi_{\mathbf{F}}^{-3}\mathcal{O}_{\mathbf{F}}^{\times}}\lvert y\rvert^{-s}\mathrm{d}^{\times}y.

Hence we identify 𝒱Π,ψ(F0)0=f1f2f3\displaystyle\mathcal{VH}_{\Pi,\psi}(F_{0})_{0}=f_{1}*f_{2}*f_{3} and 𝒱Π,ψ(F0)1(y)=τ0q2η0(y)𝟙ϖ𝐅3𝒪𝐅×(y)\displaystyle\mathcal{VH}_{\Pi,\psi}(F_{0})_{1}(y)=\tau_{0}q^{2}\cdot\eta_{0}(-y)\mathbbm{1}_{\varpi_{\mathbf{F}}^{-3}\mathcal{O}_{\mathbf{F}}^{\times}}(y). Similarly we can write 𝒱Π,ψ(F1)=𝒱Π,ψ(F1)0+𝒱Π,ψ(F1)1\displaystyle\mathcal{VH}_{\Pi,\psi}(F_{1})=\mathcal{VH}_{\Pi,\psi}(F_{1})_{0}+\mathcal{VH}_{\Pi,\psi}(F_{1})_{1} with the properties

𝒱Π,ψ(F1)0(yδ)=𝒱Π,ψ(F1)0(y),δ𝒪𝐅×;\displaystyle\mathcal{VH}_{\Pi,\psi}(F_{1})_{0}(y\delta)=\mathcal{VH}_{\Pi,\psi}(F_{1})_{0}(y),\quad\forall\delta\in\mathcal{O}_{\mathbf{F}}^{\times};
𝐅×𝒱Π,ψ(F1)0(y)|y|sd×y=q2s1i=13𝐅×fi(y)|y|sd×y,\displaystyle\int_{\mathbf{F}^{\times}}\mathcal{VH}_{\Pi,\psi}(F_{1})_{0}(y)\lvert y\rvert^{-s}\mathrm{d}^{\times}y=-q^{2s-1}\sideset{}{{}_{i=1}^{3}}{\prod}\int_{\mathbf{F}^{\times}}f_{i}(y)\lvert y\rvert^{-s}\mathrm{d}^{\times}y,
𝐅×𝒱Π,ψ(F1)1(y)χ1(y)|y|sd×y=𝟙𝔠(χ)=1q3sϖ𝐅1𝒪𝐅×ψ(y)χ2(y)d×y(ϖ𝐅1𝒪𝐅×ψ(y)χ1(y)d×y)3.\displaystyle\int_{\mathbf{F}^{\times}}\mathcal{VH}_{\Pi,\psi}(F_{1})_{1}(y)\chi^{-1}(y)\lvert y\rvert^{-s}\mathrm{d}^{\times}y=\mathbbm{1}_{\mathfrak{c}(\chi)=1}\cdot q^{3-s}\int_{\varpi_{\mathbf{F}}^{-1}\mathcal{O}_{\mathbf{F}}^{\times}}\psi(y)\chi^{2}(y)\mathrm{d}^{\times}y\cdot\left(\int_{\varpi_{\mathbf{F}}^{-1}\mathcal{O}_{\mathbf{F}}^{\times}}\psi(y)\chi^{-1}(y)\mathrm{d}^{\times}y\right)^{3}.

We easily identify 𝒱Π,ψ(F1)0(y)=q1(f1f2f3)(ϖ𝐅2y)\displaystyle\mathcal{VH}_{\Pi,\psi}(F_{1})_{0}(y)=-q^{-1}\cdot(f_{1}*f_{2}*f_{3})(\varpi_{\mathbf{F}}^{-2}y). The argument in the proof of Corollary 5.4 shows that supp(𝒱Π,ψ(F1)1)ϖ𝐅1𝒪𝐅×\displaystyle\mathrm{supp}\left(\mathcal{VH}_{\Pi,\psi}(F_{1})_{1}\right)\subset\varpi_{\mathbf{F}}^{-1}\mathcal{O}_{\mathbf{F}}^{\times}, and for y𝒪𝐅×\displaystyle y\in\mathcal{O}_{\mathbf{F}}^{\times} that

𝒱Π,ψ(F1)1(yϖ𝐅)=q3(𝒪𝐅×)3ψ(t2+t3+t4+t21t31t42yϖ𝐅)dt2dt3dt4ζ𝐅(1)q3(𝒪𝐅×)4ψ(t1+t2+t3+t4ϖ𝐅)dt=τ0q3(𝒪𝐅×)2ψ(t2+t3t2t34yϖ𝐅)η0(yt2t3)dt2dt3q2(𝒪𝐅×)2ψ(t2+t3ϖ𝐅)dt2dt3ζ𝐅(1)q1=τ0q3(𝒪𝐅×)2ψ(t2+t3t2t34yϖ𝐅)η0(yt2t3)dt2dt3ζ𝐅(1).\mathcal{VH}_{\Pi,\psi}(F_{1})_{1}\left(\tfrac{y}{\varpi_{\mathbf{F}}}\right)=q^{3}\int_{(\mathcal{O}_{\mathbf{F}}^{\times})^{3}}\psi\left(\tfrac{t_{2}+t_{3}+t_{4}+t_{2}^{-1}t_{3}^{-1}t_{4}^{2}y}{\varpi_{\mathbf{F}}}\right)\mathrm{d}t_{2}\mathrm{d}t_{3}\mathrm{d}t_{4}-\zeta_{\mathbf{F}}(1)q^{3}\int_{(\mathcal{O}_{\mathbf{F}}^{\times})^{4}}\psi\left(\tfrac{t_{1}+t_{2}+t_{3}+t_{4}}{\varpi_{\mathbf{F}}}\right)\mathrm{d}\vec{t}\\ =\tau_{0}q^{3}\int_{(\mathcal{O}_{\mathbf{F}}^{\times})^{2}}\psi\left(\tfrac{t_{2}+t_{3}-\frac{t_{2}t_{3}}{4y}}{\varpi_{\mathbf{F}}}\right)\eta_{0}\left(\tfrac{y}{t_{2}t_{3}}\right)\mathrm{d}t_{2}\mathrm{d}t_{3}-q^{2}\int_{(\mathcal{O}_{\mathbf{F}}^{\times})^{2}}\psi\left(\tfrac{t_{2}+t_{3}}{\varpi_{\mathbf{F}}}\right)\mathrm{d}t_{2}\mathrm{d}t_{3}-\zeta_{\mathbf{F}}(1)q^{-1}\\ =\tau_{0}q^{3}\int_{(\mathcal{O}_{\mathbf{F}}^{\times})^{2}}\psi\left(\tfrac{t_{2}+t_{3}-\frac{t_{2}t_{3}}{4y}}{\varpi_{\mathbf{F}}}\right)\eta_{0}\left(\tfrac{y}{t_{2}t_{3}}\right)\mathrm{d}t_{2}\mathrm{d}t_{3}-\zeta_{\mathbf{F}}(1).

Re-numbering the variables and inserting the formula of 𝒱Π,ψ(F0)\displaystyle\mathcal{VH}_{\Pi,\psi}(F_{0}) we get the formula of 𝒱Π,ψ(F1)\displaystyle\mathcal{VH}_{\Pi,\psi}(F_{1}). ∎

5.2. Further Reductions

The functions Fn\displaystyle F_{n} are “building blocks” of our test functions Hc\displaystyle H_{c} in (4.3) when 𝐋/𝐅\displaystyle\mathbf{L}/\mathbf{F} is not ramified. In fact writing ε𝐋:=η𝐋/𝐅(ϖ𝐅){±1}\displaystyle\varepsilon_{\mathbf{L}}:=\eta_{\mathbf{L}/\mathbf{F}}(\varpi_{\mathbf{F}})\in\{\pm 1\} and applying Lemma 3.10 we can rewrite the summands of Hc\displaystyle H_{c} as

(5.1) H2n+1=0,\displaystyle\displaystyle H_{2n+1}=0,
(5.2) H2n={Enif 2n0nn11(ε𝐋q)n𝐋1𝒪𝐋Tr(α)2(1+𝒫𝐅2(n01))β(α)(𝔱(Tr(α)2).Fn)dαif n=2n01(ε𝐋q)n𝐋1𝒪𝐋Tr(α)2(1+ϖ𝐅2(nn0)𝒪𝐅×)β(α)(𝔱(Tr(α)2).Fn)dαif n0+1n<2n01.\displaystyle\displaystyle H_{2n}=\begin{cases}E_{n}&\text{if }2n_{0}\leqslant n\leqslant n_{1}-1\\ (\varepsilon_{\mathbf{L}}q)^{n}\int_{\begin{subarray}{c}\mathbf{L}^{1}\cap\mathcal{O}_{\mathbf{L}}\\ {\rm Tr}(\alpha)\in 2(1+\mathcal{P}_{\mathbf{F}}^{2(n_{0}-1)})\end{subarray}}\beta(\alpha)\cdot\left(\mathfrak{t}\left({\rm Tr}(\alpha)^{2}\right).F_{n}\right)\mathrm{d}\alpha&\text{if }n=2n_{0}-1\\ (\varepsilon_{\mathbf{L}}q)^{n}\int_{\begin{subarray}{c}\mathbf{L}^{1}\cap\mathcal{O}_{\mathbf{L}}\\ {\rm Tr}(\alpha)\in 2(1+\varpi_{\mathbf{F}}^{2(n-n_{0})}\mathcal{O}_{\mathbf{F}}^{\times})\end{subarray}}\beta(\alpha)\cdot\left(\mathfrak{t}\left({\rm Tr}(\alpha)^{2}\right).F_{n}\right)\mathrm{d}\alpha&\text{if }n_{0}+1\leqslant n<2n_{0}-1\end{cases}.

The decomposition of H2n0\displaystyle H_{2n_{0}} is subtler and really goes in the direction of expression in terms of the quadratic elementary functions. We shall write (the second numeric parameter m\displaystyle m in subscript always indicates the parameter of the relevant quadratic elementary function)

H2n0\displaystyle\displaystyle H_{2n_{0}} =H2n0a+H2n0b,\displaystyle\displaystyle=H_{2n_{0}}^{a}+H_{2n_{0}}^{b},
(5.3) H2n0a\displaystyle\displaystyle H_{2n_{0}}^{a} =(ε𝐋q)n02𝟙η0(1)=ε𝐋m=0n01H2n0,ma,1+(ε𝐋q)n02𝟙η0(1)=ε𝐋m=0n01H2n0,ma,ε,\displaystyle\displaystyle=\tfrac{(\varepsilon_{\mathbf{L}}q)^{n_{0}}}{2}\mathbbm{1}_{\eta_{0}(-1)=\varepsilon_{\mathbf{L}}}\cdot\sideset{}{{}_{m=0}^{n_{0}-1}}{\sum}H_{2n_{0},m}^{a,1}+\tfrac{(\varepsilon_{\mathbf{L}}q)^{n_{0}}}{2}\mathbbm{1}_{\eta_{0}(-1)=-\varepsilon_{\mathbf{L}}}\cdot\sideset{}{{}_{m=0}^{n_{0}-1}}{\sum}H_{2n_{0},m}^{a,\varepsilon},
(5.4) H2n0b\displaystyle\displaystyle H_{2n_{0}}^{b} =(ε𝐋q)n02{H2n0,n0b,ε+H2n0,n0b,1},\displaystyle\displaystyle=\tfrac{(\varepsilon_{\mathbf{L}}q)^{n_{0}}}{2}\cdot\left\{H_{2n_{0},n_{0}}^{b,\varepsilon}+H_{2n_{0},n_{0}}^{b,1}\right\},

where the summands are given by (below we assume m1\displaystyle m\geqslant 1 and τ{1,ε}\displaystyle\tau\in\{1,\varepsilon\})

H2n0,0a,τ=𝐋1𝒪𝐋Tr(xτα)𝒫𝐅n0β(xτα)dα(𝔱(τ1ϖ𝐅2n0).F0),\displaystyle H_{2n_{0},0}^{a,\tau}=\int_{\begin{subarray}{c}\mathbf{L}^{1}\cap\mathcal{O}_{\mathbf{L}}\\ {\rm Tr}(x_{\tau}\alpha)\in\mathcal{P}_{\mathbf{F}}^{n_{0}}\end{subarray}}\beta(x_{\tau}\alpha)\mathrm{d}\alpha\cdot\left(\mathfrak{t}\left(\tau^{-1}\varpi_{\mathbf{F}}^{2n_{0}}\right).F_{0}\right),
H2n0,ma,τ=𝐋1𝒪𝐋Tr(xτα)ϖ𝐅n0m𝒪𝐅×β(xτα)(𝔱(τ1Tr(xτα)2).Fm)dα;\displaystyle H_{2n_{0},m}^{a,\tau}=\int_{\begin{subarray}{c}\mathbf{L}^{1}\cap\mathcal{O}_{\mathbf{L}}\\ {\rm Tr}(x_{\tau}\alpha)\in\varpi_{\mathbf{F}}^{n_{0}-m}\mathcal{O}_{\mathbf{F}}^{\times}\end{subarray}}\beta(x_{\tau}\alpha)\cdot\left(\mathfrak{t}\left(\tau^{-1}{\rm Tr}(x_{\tau}\alpha)^{2}\right).F_{m}\right)\mathrm{d}\alpha;
H2n0,n0b,ε=𝐋1𝒪𝐋Tr(xεα)𝒪𝐅×β(xεα)(𝔱(ε1Tr(xεα)2).Fn0)dα,\displaystyle H_{2n_{0},n_{0}}^{b,\varepsilon}=\int_{\begin{subarray}{c}\mathbf{L}^{1}\cap\mathcal{O}_{\mathbf{L}}\\ {\rm Tr}(x_{\varepsilon}\alpha)\in\mathcal{O}_{\mathbf{F}}^{\times}\end{subarray}}\beta(x_{\varepsilon}\alpha)\cdot\left(\mathfrak{t}\left(\varepsilon^{-1}{\rm Tr}(x_{\varepsilon}\alpha)^{2}\right).F_{n_{0}}\right)\mathrm{d}\alpha,
H2,1b,1=𝐋1𝒪𝐋Tr(α)𝒪𝐅×β(α)(𝔱(Tr(α)2).F1)dα,\displaystyle H_{2,1}^{b,1}=\int_{\begin{subarray}{c}\mathbf{L}^{1}\cap\mathcal{O}_{\mathbf{L}}\\ {\rm Tr}(\alpha)\in\mathcal{O}_{\mathbf{F}}^{\times}\end{subarray}}\beta(\alpha)\cdot\left(\mathfrak{t}\left({\rm Tr}(\alpha)^{2}\right).F_{1}\right)\mathrm{d}\alpha,
H2n0,n0b,1=𝐋1𝒪𝐋Tr(α)2𝒪𝐅×4(1+𝒫𝐅)β(α)(𝔱(Tr(α)2).Fn0)dα,n02.\displaystyle H_{2n_{0},n_{0}}^{b,1}=\int_{\begin{subarray}{c}\mathbf{L}^{1}\cap\mathcal{O}_{\mathbf{L}}\\ {\rm Tr}(\alpha)^{2}\in\mathcal{O}_{\mathbf{F}}^{\times}-4(1+\mathcal{P}_{\mathbf{F}})\end{subarray}}\beta(\alpha)\cdot\left(\mathfrak{t}\left({\rm Tr}(\alpha)^{2}\right).F_{n_{0}}\right)\mathrm{d}\alpha,\ n_{0}\geqslant 2.
Lemma 5.7.

Let χ\displaystyle\chi be a (unitary) character of 𝐅×\displaystyle\mathbf{F}^{\times} with 𝔠(χ)=n\displaystyle\mathfrak{c}(\chi)=n. Recall the additive parameter cβ\displaystyle c_{\beta} (resp. cχ\displaystyle c_{\chi}) in Lemma 3.8 (resp. Remark 3.9).

(1) Assume 0n<n0\displaystyle 0\leqslant n<n_{0}. For 𝐋/𝐅\displaystyle\mathbf{L}/\mathbf{F} split resp. unramified we have

𝒪𝐅×χ0(1+ϖ𝐅n0nt1ϖ𝐅n0nt)χ(t)dtresp.𝒪𝐅×β(1+ϖ𝐅n0ntε)χ(t)dtqn2.\displaystyle\int_{\mathcal{O}_{\mathbf{F}}^{\times}}\chi_{0}\left(\tfrac{1+\varpi_{\mathbf{F}}^{n_{0}-n}t}{1-\varpi_{\mathbf{F}}^{n_{0}-n}t}\right)\chi(t)\mathrm{d}t\quad{\rm resp.}\quad\int_{\mathcal{O}_{\mathbf{F}}^{\times}}\beta(1+\varpi_{\mathbf{F}}^{n_{0}-n}t\sqrt{\varepsilon})\chi(t)\mathrm{d}t\ll q^{-\frac{n}{2}}.

(2) Assume n=n02\displaystyle n=n_{0}\geqslant 2. For 𝐋/𝐅\displaystyle\mathbf{L}/\mathbf{F} split resp. unramified we have for any k{0,1}\displaystyle k\in\{0,1\}

𝒪𝐅×±(±1+𝒫𝐅)χ0(1+t1t)χ(t)η0k(1t2)dtresp.𝒪𝐅×β(1+tε)χ(t)η0k(1t2ε)dt{qn02if χ𝓔(β)qn02+12if χ𝓔(β),\int_{\mathcal{O}_{\mathbf{F}}^{\times}-\sideset{}{{}_{\pm}}{\bigcup}(\pm 1+\mathcal{P}_{\mathbf{F}})}\chi_{0}\left(\tfrac{1+t}{1-t}\right)\chi(t)\eta_{0}^{k}(1-t^{2})\mathrm{d}t\quad{\rm resp.}\quad\int_{\mathcal{O}_{\mathbf{F}}^{\times}}\beta(1+t\sqrt{\varepsilon})\chi(t)\eta_{0}^{k}(1-t^{2}\varepsilon)\mathrm{d}t\\ \ll\begin{cases}q^{-\frac{n_{0}}{2}}&\text{if }\chi\notin\boldsymbol{\mathcal{E}}(\beta)\\ q^{-\frac{n_{0}}{2}+\frac{1}{2}}&\text{if }\chi\in\boldsymbol{\mathcal{E}}(\beta)\end{cases},

where 𝓔(β)\displaystyle\boldsymbol{\mathcal{E}}(\beta)\neq\emptyset only if 2n0\displaystyle 2\nmid n_{0} and η0(1)=ε𝐋\displaystyle\eta_{0}(-1)=\varepsilon_{\mathbf{L}}, under which condition it is given by

𝓔(β)={{χ|(cχcβ1)21(mod𝒫𝐅)}if 𝐋/𝐅 split{χ|(cχcβ1)2ε(mod𝒫𝐅)}if 𝐋/𝐅 unramified.\displaystyle\boldsymbol{\mathcal{E}}(\beta)=\begin{cases}\left\{\chi\ \middle|\ (c_{\chi}c_{\beta}^{-1})^{2}\equiv-1\pmod{\mathcal{P}_{\mathbf{F}}}\right\}&\text{if }\mathbf{L}/\mathbf{F}\text{ split}\\ \left\{\chi\ \middle|\ (c_{\chi}c_{\beta}^{-1})^{2}\equiv-\varepsilon\pmod{\mathcal{P}_{\mathbf{F}}}\right\}&\text{if }\mathbf{L}/\mathbf{F}\text{ unramified}\end{cases}.
Proof.

We omit the proof of the split case, which is similar and simpler. The case of n=0<n0\displaystyle n=0<n_{0} is easy and omitted, since the integrand is constant.

(1) If n=1\displaystyle n=1, then tβ(1+ϖ𝐅n0ntε)\displaystyle t\mapsto\beta\left(1+\varpi_{\mathbf{F}}^{n_{0}-n}t\sqrt{\varepsilon}\right) is a non-trivial additive character of 𝒪𝐅\displaystyle\mathcal{O}_{\mathbf{F}} and the bound follows from the one for Gauss sums. Assume n2\displaystyle n\geqslant 2. We perform a level n2\displaystyle\lceil\tfrac{n}{2}\rceil regularization to dt\displaystyle\mathrm{d}t and get

𝒪𝐅×β(1+ϖ𝐅n0ntε)χ(t)dt=𝒪𝐅×β(1+ϖ𝐅n0ntε)χ(t)(𝒪𝐅β(1+ϖ𝐅n0n21+ϖ𝐅n0ntεtuε)χ(1+ϖ𝐅n2u)du)𝑑t.\int_{\mathcal{O}_{\mathbf{F}}^{\times}}\beta\left(1+\varpi_{\mathbf{F}}^{n_{0}-n}t\sqrt{\varepsilon}\right)\chi(t)\mathrm{d}t\\ =\int_{\mathcal{O}_{\mathbf{F}}^{\times}}\beta\left(1+\varpi_{\mathbf{F}}^{n_{0}-n}t\sqrt{\varepsilon}\right)\chi(t)\cdot\left(\oint_{\mathcal{O}_{\mathbf{F}}}\beta\left(1+\tfrac{\varpi_{\mathbf{F}}^{n_{0}-\lfloor\frac{n}{2}\rfloor}}{1+\varpi_{\mathbf{F}}^{n_{0}-n}t\sqrt{\varepsilon}}tu\sqrt{\varepsilon}\right)\chi(1+\varpi_{\mathbf{F}}^{\lceil\frac{n}{2}\rceil}u)\mathrm{d}u\right)dt.

The integrands of the inner integral, denoted by I(t;n)\displaystyle I(t;n), are additive characters of 𝒪𝐅\displaystyle\mathcal{O}_{\mathbf{F}}. We have

I(t;n)=𝒪𝐅ψ(1ϖ𝐅n22cβtuε1ϖ𝐅2(n0n)t2ε)ψ(cχuϖ𝐅n2)du.\displaystyle I(t;n)=\oint_{\mathcal{O}_{\mathbf{F}}}\psi\left(\tfrac{1}{\varpi_{\mathbf{F}}^{\lfloor\frac{n}{2}\rfloor}}\cdot\tfrac{2c_{\beta}tu\varepsilon}{1-\varpi_{\mathbf{F}}^{2(n_{0}-n)}t^{2}\varepsilon}\right)\psi\left(\tfrac{c_{\chi}u}{\varpi_{\mathbf{F}}^{\lfloor\frac{n}{2}\rfloor}}\right)\mathrm{d}u.

The non-vanishing of I(t;n)\displaystyle I(t;n) implies the congruence condition

2cβtε1ϖ𝐅2(n0n)t2ε+cχ𝒫𝐅n22cβtε+cχ(1ϖ𝐅2(n0n)t2ε)𝒫𝐅n2,\displaystyle\tfrac{2c_{\beta}t\varepsilon}{1-\varpi_{\mathbf{F}}^{2(n_{0}-n)}t^{2}\varepsilon}+c_{\chi}\in\mathcal{P}_{\mathbf{F}}^{\lfloor\frac{n}{2}\rfloor}\quad\Leftrightarrow\quad 2c_{\beta}t\varepsilon+c_{\chi}(1-\varpi_{\mathbf{F}}^{2(n_{0}-n)}t^{2}\varepsilon)\in\mathcal{P}_{\mathbf{F}}^{\lfloor\frac{n}{2}\rfloor},

which has a unique solution tt0+𝒫𝐅n2\displaystyle t\in t_{0}+\mathcal{P}_{\mathbf{F}}^{\lfloor\frac{n}{2}\rfloor} with t0𝒪𝐅×\displaystyle t_{0}\in\mathcal{O}_{\mathbf{F}}^{\times} by Hensel’s lemma. Consequently we get

(5.5) 𝒪𝐅×β(1+ϖ𝐅n0ntε)χ(t)dt=t0+𝒫𝐅n2β(1+ϖ𝐅n0ntε)χ(t)dtqn2.\int_{\mathcal{O}_{\mathbf{F}}^{\times}}\beta\left(1+\varpi_{\mathbf{F}}^{n_{0}-n}t\sqrt{\varepsilon}\right)\chi(t)\mathrm{d}t=\int_{t_{0}+\mathcal{P}_{\mathbf{F}}^{\lfloor\frac{n}{2}\rfloor}}\beta\left(1+\varpi_{\mathbf{F}}^{n_{0}-n}t\sqrt{\varepsilon}\right)\chi(t)\mathrm{d}t\ll q^{-\lfloor\frac{n}{2}\rfloor}.

If 2n\displaystyle 2\mid n then we are done (for both (1) and (2)). Otherwise let n=2m+1\displaystyle n=2m+1. We may assume

2cβt0ε+cχ(1ϖ𝐅2(n0n)t02ε)=0,\displaystyle 2c_{\beta}t_{0}\varepsilon+c_{\chi}(1-\varpi_{\mathbf{F}}^{2(n_{0}-n)}t_{0}^{2}\varepsilon)=0,

make the change of variables t=t0(1+ϖ𝐅mu)\displaystyle t=t_{0}(1+\varpi_{\mathbf{F}}^{m}u), and continue (5.5) to conclude by

𝒪𝐅×β(1+ϖ𝐅n0ntε)χ(t)dt=qmβ(1+ϖ𝐅n0nt0ε)χ(t0)𝒪𝐅ψ(cχu22ϖ𝐅)𝑑uqn2.\displaystyle\int_{\mathcal{O}_{\mathbf{F}}^{\times}}\beta\left(1+\varpi_{\mathbf{F}}^{n_{0}-n}t\sqrt{\varepsilon}\right)\chi(t)\mathrm{d}t=q^{-m}\beta(1+\varpi_{\mathbf{F}}^{n_{0}-n}t_{0}\sqrt{\varepsilon})\chi(t_{0})\int_{\mathcal{O}_{\mathbf{F}}}\psi\left(-\tfrac{c_{\chi}u^{2}}{2\varpi_{\mathbf{F}}}\right)du\ll q^{-\frac{n}{2}}.

(2) Let n=2m+1\displaystyle n=2m+1 with m1\displaystyle m\geqslant 1. With the change of variables t=t0(1+ϖ𝐅mu)\displaystyle t=t_{0}(1+\varpi_{\mathbf{F}}^{m}u) for t0𝒪𝐅×\displaystyle t_{0}\in\mathcal{O}_{\mathbf{F}}^{\times} satisfying

(5.6) 2cβt0ε+cχ(1t02ε)=0,2c_{\beta}t_{0}\varepsilon+c_{\chi}(1-t_{0}^{2}\varepsilon)=0,

which is solvable only if η0(1)=ε𝐋\displaystyle\eta_{0}(-1)=\varepsilon_{\mathbf{L}}, we get the equation

𝒪𝐅×β(1+tε)χ(t)η0k(1t2ε)dt=t0β(1+t0ε)χ(t0)η0k(1t02ε)𝒪𝐅ψ(u2ϖ(2cβt03ε(1t02ε)2cχ2))du=t0β(1+t0ε)χ(t0)η0k(1t02ε)𝒪𝐅ψ(cχu22ϖ(cχcβt01))du.\int_{\mathcal{O}_{\mathbf{F}}^{\times}}\beta(1+t\sqrt{\varepsilon})\chi(t)\eta_{0}^{k}(1-t^{2}\varepsilon)\mathrm{d}t=\sideset{}{{}_{t_{0}}}{\sum}\beta(1+t_{0}\sqrt{\varepsilon})\chi(t_{0})\eta_{0}^{k}(1-t_{0}^{2}\varepsilon)\cdot\int_{\mathcal{O}_{\mathbf{F}}}\psi\left(\tfrac{u^{2}}{\varpi}\left(\tfrac{2c_{\beta}t_{0}^{3}\varepsilon}{(1-t_{0}^{2}\varepsilon)^{2}}-\tfrac{c_{\chi}}{2}\right)\right)\mathrm{d}u\\ =\sideset{}{{}_{t_{0}}}{\sum}\beta(1+t_{0}\sqrt{\varepsilon})\chi(t_{0})\eta_{0}^{k}(1-t_{0}^{2}\varepsilon)\cdot\int_{\mathcal{O}_{\mathbf{F}}}\psi\left(\tfrac{c_{\chi}u^{2}}{2\varpi}\left(\tfrac{c_{\chi}}{c_{\beta}}t_{0}-1\right)\right)\mathrm{d}u.

The above integral is q12\displaystyle\ll q^{-\frac{1}{2}} unless t0cχ1cβ(mod𝒫𝐅)\displaystyle t_{0}\equiv c_{\chi}^{-1}c_{\beta}\pmod{\mathcal{P}_{\mathbf{F}}}, in which case (5.6) implies χ𝓔(β)\displaystyle\chi\in\boldsymbol{\mathcal{E}}(\beta). ∎

Lemma 5.8.

Suppose n0+1n2n01\displaystyle n_{0}+1\leqslant n\leqslant 2n_{0}-1 (hence n02\displaystyle n_{0}\geqslant 2) and na(Π)\displaystyle n\geqslant a(\Pi). Then we have

h~2n(χ)qn0𝟙2n0n(𝔠(χη0n))+qn0𝟙n=2n01𝟙𝔠(χη0)=0.\displaystyle\widetilde{h}_{2n}^{-}(\chi)\ll q^{-n_{0}}\mathbbm{1}_{2n_{0}-n}(\mathfrak{c}(\chi\eta_{0}^{n}))+q^{-n_{0}}\mathbbm{1}_{n=2n_{0}-1}\mathbbm{1}_{\mathfrak{c}(\chi\eta_{0})=0}.
Proof.

(1) First consider n<2n01\displaystyle n<2n_{0}-1. By Corollary 5.4 we have for any δ1+ϖ𝐅2(nn0)𝒪𝐅×\displaystyle\delta\in 1+\varpi_{\mathbf{F}}^{2(n-n_{0})}\mathcal{O}_{\mathbf{F}}^{\times}

(5.7) 𝐅𝒪𝐅𝒱Π,ψ𝔪1(𝔱(4δ).Fn)(t)ψ(t)χ1(t)|t|12d×t=qnζ𝐅(1){χ1(δ1δ)γ(1,χ,ψ)𝟙2n0n(𝔠(χ))if 2nχ1η0(δ1δ)τ0q12γ(1,χη0,ψ)𝟙2n0n(𝔠(χη0))if 2n,\int_{\mathbf{F}-\mathcal{O}_{\mathbf{F}}}\mathcal{VH}_{\Pi,\psi}\circ\mathfrak{m}_{-1}\left(\mathfrak{t}\left(4\delta\right).F_{n}\right)(t)\cdot\psi(-t)\chi^{-1}(t)\lvert t\rvert^{-\frac{1}{2}}\mathrm{d}^{\times}t\\ =q^{-n}\zeta_{\mathbf{F}}(1)\cdot\begin{cases}\chi^{-1}\left(\tfrac{\delta}{1-\delta}\right)\cdot\gamma(1,\chi,\psi)\mathbbm{1}_{2n_{0}-n}(\mathfrak{c}(\chi))&\text{if }2\mid n\\ \chi^{-1}\eta_{0}\left(\tfrac{\delta}{1-\delta}\right)\cdot\tau_{0}q^{\frac{1}{2}}\gamma(1,\chi\eta_{0},\psi)\mathbbm{1}_{2n_{0}-n}(\mathfrak{c}(\chi\eta_{0}))&\text{if }2\nmid n\end{cases},

where we used η0(δ)=1\displaystyle\eta_{0}(\delta)=1. Inserting (5.7) with δ=41Tr(α)2\displaystyle\delta=4^{-1}{\rm Tr}(\alpha)^{2} into (5.2) we get

(5.8) h~2n(χ)=𝐅𝒪𝐅𝒱Π,ψ𝔪1(H2n)(t)ψ(t)χ1(t)|t|12d×t|𝐋1𝒪𝐋Tr(α)2(1+ϖ𝐅2(nn0)𝒪𝐅×)β(α)χ1η0n(Tr(α)24Tr(α)2)dα|q2n0n2𝟙2n0n(𝔠(χη0n)).\widetilde{h}_{2n}^{-}(\chi)=\int_{\mathbf{F}-\mathcal{O}_{\mathbf{F}}}\mathcal{VH}_{\Pi,\psi}\circ\mathfrak{m}_{-1}(H_{2n})(t)\cdot\psi(-t)\chi^{-1}(t)\lvert t\rvert^{-\frac{1}{2}}\mathrm{d}^{\times}t\ll\\ \left\lvert\int_{\begin{subarray}{c}\mathbf{L}^{1}\cap\mathcal{O}_{\mathbf{L}}\\ {\rm Tr}(\alpha)\in 2(1+\varpi_{\mathbf{F}}^{2(n-n_{0})}\mathcal{O}_{\mathbf{F}}^{\times})\end{subarray}}\beta(\alpha)\cdot\chi^{-1}\eta_{0}^{n}\left(\tfrac{{\rm Tr}(\alpha)^{2}}{4-{\rm Tr}(\alpha)^{2}}\right)\mathrm{d}\alpha\right\rvert\cdot q^{-\frac{2n_{0}-n}{2}}\mathbbm{1}_{2n_{0}-n}(\mathfrak{c}(\chi\eta_{0}^{n})).

In the case of split, resp. unramified 𝐋/𝐅\displaystyle\mathbf{L}/\mathbf{F} we apply the change of variables t=αα1α+α1\displaystyle t=\tfrac{\alpha-\alpha^{-1}}{\alpha+\alpha^{-1}}, resp. t=αα1α+α11ε\displaystyle t=\tfrac{\alpha-\alpha^{-1}}{\alpha+\alpha^{-1}}\tfrac{1}{\sqrt{\varepsilon}}. The inner integral in (5.8) becomes

2χη0n(1)ϖ𝐅nn0𝒪𝐅×χ0(1+t1t)χ2(t)dtresp.χη0n(ε)ϖ𝐅nn0𝒪𝐅×β(1+tε)χ2(t)dt.\displaystyle 2\chi\eta_{0}^{n}(-1)\int_{\varpi_{\mathbf{F}}^{n-n_{0}}\mathcal{O}_{\mathbf{F}}^{\times}}\chi_{0}\left(\tfrac{1+t}{1-t}\right)\chi^{2}(t)\mathrm{d}t\quad{\rm resp.}\quad\chi\eta_{0}^{n}(-\varepsilon)\int_{\varpi_{\mathbf{F}}^{n-n_{0}}\mathcal{O}_{\mathbf{F}}^{\times}}\beta(1+t\sqrt{\varepsilon})\chi^{2}(t)\mathrm{d}t.

We apply Lemma 5.7 (1) to bound the above integrals and conclude.

(2) Consider n=2n01\displaystyle n=2n_{0}-1. We have for any δ1+𝒫𝐅2(n01)\displaystyle\delta\in 1+\mathcal{P}_{\mathbf{F}}^{2(n_{0}-1)}

(5.9) 𝐅𝒪𝐅𝒱Π,ψ𝔪1(𝔱(4δ).Fn)(t)ψ(t)χ1(t)|t|12d×t=q(2n01)τ0q12{χη(ϖ𝐅)2n01𝟙0(𝔠(χη0))if δ1+𝒫𝐅2n01χ1η0(δ1δ)ζ𝐅(1){γ(1,χη0,ψ)𝟙1(𝔠(χη0))q1𝟙0(𝔠(χη0))}if δ1+ϖ𝐅2(n01)𝒪𝐅×.\int_{\mathbf{F}-\mathcal{O}_{\mathbf{F}}}\mathcal{VH}_{\Pi,\psi}\circ\mathfrak{m}_{-1}\left(\mathfrak{t}\left(4\delta\right).F_{n}\right)(t)\cdot\psi(-t)\chi^{-1}(t)\lvert t\rvert^{-\frac{1}{2}}\mathrm{d}^{\times}t=q^{-(2n_{0}-1)}\cdot\tau_{0}q^{\frac{1}{2}}\cdot\\ \begin{cases}\chi\eta(\varpi_{\mathbf{F}})^{2n_{0}-1}\cdot\mathbbm{1}_{0}(\mathfrak{c}(\chi\eta_{0}))&\text{if }\delta\in 1+\mathcal{P}_{\mathbf{F}}^{2n_{0}-1}\\ \chi^{-1}\eta_{0}\left(\tfrac{\delta}{1-\delta}\right)\cdot\zeta_{\mathbf{F}}(1)\left\{\gamma(1,\chi\eta_{0},\psi)\mathbbm{1}_{1}(\mathfrak{c}(\chi\eta_{0}))-q^{-1}\mathbbm{1}_{0}(\mathfrak{c}(\chi\eta_{0}))\right\}&\text{if }\delta\in 1+\varpi_{\mathbf{F}}^{2(n_{0}-1)}\mathcal{O}_{\mathbf{F}}^{\times}\end{cases}.

Inserting (5.9) with δ=41Tr(α)2\displaystyle\delta=4^{-1}{\rm Tr}(\alpha)^{2} into (5.2) we get

(5.10) h~2n(χ)=𝐅𝒪𝐅𝒱Π,ψ𝔪1(H2n)(t)ψ(t)χ1(t)|t|12d×t|𝐋1𝒪𝐋Tr(α)2(1+ϖ𝐅2(n01)𝒪𝐅×)β(α)χ1η0(Tr(α)24Tr(α)2)dα|q12𝟙1(𝔠(χη0))+|𝐋1𝒪𝐋Tr(α)2(1+𝒫𝐅2n01)β(α)dαχη0(ϖ𝐅)1q1𝐋1𝒪𝐋Tr(α)2(1+ϖ𝐅2(n01)𝒪𝐅×)β(α)dα|𝟙0(𝔠(χη0)).\widetilde{h}_{2n}^{-}(\chi)=\int_{\mathbf{F}-\mathcal{O}_{\mathbf{F}}}\mathcal{VH}_{\Pi,\psi}\circ\mathfrak{m}_{-1}(H_{2n})(t)\cdot\psi(-t)\chi^{-1}(t)\lvert t\rvert^{-\frac{1}{2}}\mathrm{d}^{\times}t\ll\\ \left\lvert\int_{\begin{subarray}{c}\mathbf{L}^{1}\cap\mathcal{O}_{\mathbf{L}}\\ {\rm Tr}(\alpha)\in 2(1+\varpi_{\mathbf{F}}^{2(n_{0}-1)}\mathcal{O}_{\mathbf{F}}^{\times})\end{subarray}}\beta(\alpha)\cdot\chi^{-1}\eta_{0}\left(\tfrac{{\rm Tr}(\alpha)^{2}}{4-{\rm Tr}(\alpha)^{2}}\right)\mathrm{d}\alpha\right\rvert\cdot q^{-\frac{1}{2}}\mathbbm{1}_{1}(\mathfrak{c}(\chi\eta_{0}))+\\ \left\lvert\int_{\begin{subarray}{c}\mathbf{L}^{1}\cap\mathcal{O}_{\mathbf{L}}\\ {\rm Tr}(\alpha)\in 2(1+\mathcal{P}_{\mathbf{F}}^{2n_{0}-1})\end{subarray}}\beta(\alpha)\mathrm{d}\alpha-\tfrac{\chi\eta_{0}(\varpi_{\mathbf{F}})^{-1}}{q-1}\int_{\begin{subarray}{c}\mathbf{L}^{1}\cap\mathcal{O}_{\mathbf{L}}\\ {\rm Tr}(\alpha)\in 2(1+\varpi_{\mathbf{F}}^{2(n_{0}-1)}\mathcal{O}_{\mathbf{F}}^{\times})\end{subarray}}\beta(\alpha)\mathrm{d}\alpha\right\rvert\cdot\mathbbm{1}_{0}(\mathfrak{c}(\chi\eta_{0})).

The first summand is bounded the same way as before. Note that Tr(α)2(1+𝒫𝐅2n01)\displaystyle{\rm Tr}(\alpha)\in 2(1+\mathcal{P}_{\mathbf{F}}^{2n_{0}-1}) is equivalent to α1+𝒫𝐋n0\displaystyle\alpha\in 1+\mathcal{P}_{\mathbf{L}}^{n_{0}}. With the same change of variables the inner integrals in the second summand become

2qn02χη0(ϖ𝐅)1q1ϖ𝐅n01𝒪𝐅×χ0(1+t1t)dtresp.qn0χη0(ϖ𝐅)1q1ϖ𝐅n01𝒪𝐅×β(1+tε)dt.\displaystyle 2q^{-n_{0}}-\tfrac{2\chi\eta_{0}(\varpi_{\mathbf{F}})^{-1}}{q-1}\int_{\varpi_{\mathbf{F}}^{n_{0}-1}\mathcal{O}_{\mathbf{F}}^{\times}}\chi_{0}\left(\tfrac{1+t}{1-t}\right)\mathrm{d}t\quad{\rm resp.}\quad q^{-n_{0}}-\tfrac{\chi\eta_{0}(\varpi_{\mathbf{F}})^{-1}}{q-1}\int_{\varpi_{\mathbf{F}}^{n_{0}-1}\mathcal{O}_{\mathbf{F}}^{\times}}\beta(1+t\sqrt{\varepsilon})\mathrm{d}t.

They are of size O(qn0)\displaystyle O(q^{-n_{0}}) since the integrands are additive characters of conductor exponent n0\displaystyle n_{0}. ∎

Lemma 5.9.

Suppose (2)a(Π)m<n0\displaystyle(2\leqslant)a(\Pi)\leqslant m<n_{0}. Then we have for unitary χ\displaystyle\chi and τ{1,ε}\displaystyle\tau\in\{1,\varepsilon\}

h~2n0,ma,τ(χ):=𝐅𝒪𝐅𝒱Π,ψ𝔪1(H2n0,ma,τ)(t)ψ(t)χ1(t)|t|12d×tq2n0𝟙m>2n03𝟙m(𝔠(χ)).\displaystyle\widetilde{h}_{2n_{0},m}^{a,\tau}(\chi):=\int_{\mathbf{F}-\mathcal{O}_{\mathbf{F}}}\mathcal{VH}_{\Pi,\psi}\circ\mathfrak{m}_{-1}\left(H_{2n_{0},m}^{a,\tau}\right)(t)\psi(-t)\chi^{-1}(t)\lvert t\rvert^{-\frac{1}{2}}\mathrm{d}^{\times}t\ll q^{-2n_{0}}\mathbbm{1}_{m>\frac{2n_{0}}{3}}\mathbbm{1}_{m}(\mathfrak{c}(\chi)).
Proof.

We first use Corollary 5.4 to obtain for any δϖ𝐅2(n0m)𝒪𝐅×\displaystyle\delta\in\varpi_{\mathbf{F}}^{2(n_{0}-m)}\mathcal{O}_{\mathbf{F}}^{\times} (note that η0(4δ)=1\displaystyle\eta_{0}(4-\delta)=1)

(5.11) 𝐅𝒪𝐅𝒱Π,ψ𝔪1(𝔱(δ).Fm)(t)ψ(t)χ1(t)|t|12d×t=qn0ζ𝐅(1)𝟙m>2n03{χ(4δδ)γ(1,χ,ψ)𝟙m(𝔠(χ))if 2mχ(4δδ)τ0q12γ(1,χη0,ψ)𝟙m(𝔠(χ))if 2m.\int_{\mathbf{F}-\mathcal{O}_{\mathbf{F}}}\mathcal{VH}_{\Pi,\psi}\circ\mathfrak{m}_{-1}\left(\mathfrak{t}\left(\delta\right).F_{m}\right)(t)\cdot\psi(-t)\chi^{-1}(t)\lvert t\rvert^{-\frac{1}{2}}\mathrm{d}^{\times}t\\ =q^{-n_{0}}\zeta_{\mathbf{F}}(1)\mathbbm{1}_{m>\frac{2n_{0}}{3}}\cdot\begin{cases}\chi\left(\tfrac{4-\delta}{\delta}\right)\cdot\gamma(1,\chi,\psi)\mathbbm{1}_{m}(\mathfrak{c}(\chi))&\text{if }2\mid m\\ \chi\left(\tfrac{4-\delta}{\delta}\right)\cdot\tau_{0}q^{\frac{1}{2}}\gamma(1,\chi\eta_{0},\psi)\mathbbm{1}_{m}(\mathfrak{c}(\chi))&\text{if }2\nmid m\end{cases}.

Inserting (5.11) with δ=τ1Tr(xτα)2\displaystyle\delta=\tau^{-1}{\rm Tr}(x_{\tau}\alpha)^{2} into the integral representation of H2n0,ma,τ\displaystyle H_{2n_{0},m}^{a,\tau} we get

(5.12) h~2n0,ma,τ(χ)𝟙m>2n03|𝐋1𝒪𝐋Tr(xτα)ϖ𝐅n0m𝒪𝐅×β(xτα)χ(4τTr(xτα)2Tr(xτα)2)dα|qn0m2𝟙m(𝔠(χ)).\widetilde{h}_{2n_{0},m}^{a,\tau}(\chi)\ll\mathbbm{1}_{m>\frac{2n_{0}}{3}}\cdot\left\lvert\int_{\begin{subarray}{c}\mathbf{L}^{1}\cap\mathcal{O}_{\mathbf{L}}\\ {\rm Tr}(x_{\tau}\alpha)\in\varpi_{\mathbf{F}}^{n_{0}-m}\mathcal{O}_{\mathbf{F}}^{\times}\end{subarray}}\beta(x_{\tau}\alpha)\cdot\chi\left(\tfrac{4\tau-{\rm Tr}(x_{\tau}\alpha)^{2}}{{\rm Tr}(x_{\tau}\alpha)^{2}}\right)\mathrm{d}\alpha\right\rvert\cdot q^{-n_{0}-\frac{m}{2}}\mathbbm{1}_{m}(\mathfrak{c}(\chi)).

In the case of split, resp. unramified 𝐋/𝐅\displaystyle\mathbf{L}/\mathbf{F} we apply the change of variables t=α+α1αα1\displaystyle t=\tfrac{\alpha+\alpha^{-1}}{\alpha-\alpha^{-1}}, resp. t=α+α1αα11ε\displaystyle t=\tfrac{\alpha+\alpha^{-1}}{\alpha-\alpha^{-1}}\tfrac{1}{\sqrt{\varepsilon}} for h~2n0,ma,1(χ)\displaystyle\widetilde{h}_{2n_{0},m}^{a,1}(\chi); t=α+εα1αεα1\displaystyle t=\tfrac{\alpha+\varepsilon\alpha^{-1}}{\alpha-\varepsilon\alpha^{-1}}, resp. t=xεα+xεα¯xεαxεα¯1ε\displaystyle t=\tfrac{x_{\varepsilon}\alpha+\overline{x_{\varepsilon}\alpha}}{x_{\varepsilon}\alpha-\overline{x_{\varepsilon}\alpha}}\tfrac{1}{\sqrt{\varepsilon}} for h~2n0,ma,ε(χ)\displaystyle\widetilde{h}_{2n_{0},m}^{a,\varepsilon}(\chi). The inner integrals in (5.12) become

2χ0χ(1)ϖ𝐅n0m𝒪𝐅×χ0(1+t1t)χ2(t)dtresp.2β(ε)χ1(ε)ϖ𝐅n0m𝒪𝐅×β(1+tε)χ2(t)dt.\displaystyle 2\chi_{0}\chi(-1)\int_{\varpi_{\mathbf{F}}^{n_{0}-m}\mathcal{O}_{\mathbf{F}}^{\times}}\chi_{0}\left(\tfrac{1+t}{1-t}\right)\chi^{-2}(t)\mathrm{d}t\quad{\rm resp.}\quad 2\beta(\sqrt{\varepsilon})\chi^{-1}(-\varepsilon)\int_{\varpi_{\mathbf{F}}^{n_{0}-m}\mathcal{O}_{\mathbf{F}}^{\times}}\beta(1+t\sqrt{\varepsilon})\chi^{-2}(t)\mathrm{d}t.

We apply Lemma 5.7 (1) to bound the above integrals and conclude the desired inequalities. ∎

Lemma 5.10.

Suppose n0a(Π)(2)\displaystyle n_{0}\geqslant a(\Pi)(\geqslant 2). Let τ{1,ε}\displaystyle\tau\in\{1,\varepsilon\}. With 𝓔(β)\displaystyle\boldsymbol{\mathcal{E}}(\beta) defined in Lemma 5.7 (2) we have

h~2n0,n0b,τ(χ):=𝐅𝒪𝐅𝒱Π,ψ𝔪1(H2n0,n0b,τ)(t)ψ(t)χ1(t)|t|12d×tq2n0𝟙n0(𝔠(χ))(𝟙𝓔(β)c(χ2)+q12𝟙𝓔(β)(χ2)).\widetilde{h}_{2n_{0},n_{0}}^{b,\tau}(\chi):=\int_{\mathbf{F}-\mathcal{O}_{\mathbf{F}}}\mathcal{VH}_{\Pi,\psi}\circ\mathfrak{m}_{-1}\left(H_{2n_{0},n_{0}}^{b,\tau}\right)(t)\psi(-t)\chi^{-1}(t)\lvert t\rvert^{-\frac{1}{2}}\mathrm{d}^{\times}t\\ \ll q^{-2n_{0}}\mathbbm{1}_{n_{0}}(\mathfrak{c}(\chi))\left(\mathbbm{1}_{\boldsymbol{\mathcal{E}}(\beta)^{c}}(\chi^{2})+q^{\frac{1}{2}}\mathbbm{1}_{\boldsymbol{\mathcal{E}}(\beta)}(\chi^{2})\right).
Proof.

We first use Corollary 5.4 to obtain for any δ𝒪𝐅×4(1+𝒫𝐅)\displaystyle\delta\in\mathcal{O}_{\mathbf{F}}^{\times}-4(1+\mathcal{P}_{\mathbf{F}})

(5.13) 𝐅𝒪𝐅𝒱Π,ψ𝔪1(𝔱(δ).Fm)(t)ψ(t)χ1(t)|t|12d×t=qn0ζ𝐅(1){χ(4δδ)γ(1,χ,ψ)𝟙n0(𝔠(χ))if 2n0η0(4δ)χ(4δδ)τ0q12γ(1,χη0,ψ)𝟙n0(𝔠(χ))if 2n0.\int_{\mathbf{F}-\mathcal{O}_{\mathbf{F}}}\mathcal{VH}_{\Pi,\psi}\circ\mathfrak{m}_{-1}\left(\mathfrak{t}\left(\delta\right).F_{m}\right)(t)\cdot\psi(-t)\chi^{-1}(t)\lvert t\rvert^{-\frac{1}{2}}\mathrm{d}^{\times}t\\ =q^{-n_{0}}\zeta_{\mathbf{F}}(1)\cdot\begin{cases}\chi\left(\tfrac{4-\delta}{\delta}\right)\cdot\gamma(1,\chi,\psi)\mathbbm{1}_{n_{0}}(\mathfrak{c}(\chi))&\text{if }2\mid n_{0}\\ \eta_{0}(4-\delta)\chi\left(\tfrac{4-\delta}{\delta}\right)\cdot\tau_{0}q^{\frac{1}{2}}\gamma(1,\chi\eta_{0},\psi)\mathbbm{1}_{n_{0}}(\mathfrak{c}(\chi))&\text{if }2\nmid n_{0}\end{cases}.

Inserting (5.13) with δ=τ1Tr(xτα)2\displaystyle\delta=\tau^{-1}{\rm Tr}(x_{\tau}\alpha)^{2} into the integral representation of H2n0,n0b,τ\displaystyle H_{2n_{0},n_{0}}^{b,\tau} we get

(5.14) h~2n0,n0b,1(χ)|𝐋1𝒪𝐋Tr(α)2𝒪𝐅×4(1+𝒫𝐅)β(α)χ(4Tr(α)2Tr(α)2)η0n0(4Tr(α)2)dα|q3n02𝟙n0(𝔠(χ)).\widetilde{h}_{2n_{0},n_{0}}^{b,1}(\chi)\ll\left\lvert\int_{\begin{subarray}{c}\mathbf{L}^{1}\cap\mathcal{O}_{\mathbf{L}}\\ {\rm Tr}(\alpha)^{2}\in\mathcal{O}_{\mathbf{F}}^{\times}-4(1+\mathcal{P}_{\mathbf{F}})\end{subarray}}\beta(\alpha)\cdot\chi\left(\tfrac{4-{\rm Tr}(\alpha)^{2}}{{\rm Tr}(\alpha)^{2}}\right)\eta_{0}^{n_{0}}\left(4-{\rm Tr}(\alpha)^{2}\right)\mathrm{d}\alpha\right\rvert\cdot q^{-\frac{3n_{0}}{2}}\mathbbm{1}_{n_{0}}(\mathfrak{c}(\chi)).
(5.15) h~2n0,n0b,ε(χ)|𝐋1𝒪𝐋Tr(xεα)𝒪𝐅×β(xεα)χ(4τTr(xεα)2Tr(xεα)2)η0n0(4τ1Tr(xεα)2)dα|q3n02𝟙n0(𝔠(χ)).\widetilde{h}_{2n_{0},n_{0}}^{b,\varepsilon}(\chi)\ll\left\lvert\int_{\begin{subarray}{c}\mathbf{L}^{1}\cap\mathcal{O}_{\mathbf{L}}\\ {\rm Tr}(x_{\varepsilon}\alpha)\in\mathcal{O}_{\mathbf{F}}^{\times}\end{subarray}}\beta(x_{\varepsilon}\alpha)\cdot\chi\left(\tfrac{4\tau-{\rm Tr}(x_{\varepsilon}\alpha)^{2}}{{\rm Tr}(x_{\varepsilon}\alpha)^{2}}\right)\eta_{0}^{n_{0}}\left(4-\tau^{-1}{\rm Tr}(x_{\varepsilon}\alpha)^{2}\right)\mathrm{d}\alpha\right\rvert\cdot q^{-\frac{3n_{0}}{2}}\mathbbm{1}_{n_{0}}(\mathfrak{c}(\chi)).

In the case of split, resp. unramified 𝐋/𝐅\displaystyle\mathbf{L}/\mathbf{F} we apply the change of variables t=α+α1αα1\displaystyle t=\tfrac{\alpha+\alpha^{-1}}{\alpha-\alpha^{-1}}, resp. t=α+α1αα11ε\displaystyle t=\tfrac{\alpha+\alpha^{-1}}{\alpha-\alpha^{-1}}\tfrac{1}{\sqrt{\varepsilon}} for h~2n0,n0b,1(χ)\displaystyle\widetilde{h}_{2n_{0},n_{0}}^{b,1}(\chi); t=α+εα1αεα1\displaystyle t=\tfrac{\alpha+\varepsilon\alpha^{-1}}{\alpha-\varepsilon\alpha^{-1}}, resp. t=xεα+xεα¯xεαxεα¯1ε\displaystyle t=\tfrac{x_{\varepsilon}\alpha+\overline{x_{\varepsilon}\alpha}}{x_{\varepsilon}\alpha-\overline{x_{\varepsilon}\alpha}}\tfrac{1}{\sqrt{\varepsilon}} for h~2n0,n0b,ε(χ)\displaystyle\widetilde{h}_{2n_{0},n_{0}}^{b,\varepsilon}(\chi). The inner integrals in (5.14) and (5.15), denoted by I1\displaystyle I_{1} and Iε\displaystyle I_{\varepsilon} respectively, become

I1=Iε=2χ0χ(1)𝒪𝐅×±(±1+𝒫𝐅)χ0(1+t1t)χ2(t)η0n0(1t2)dtresp.{I1=2β(ε)χ1(ε)η0n0(ε)𝒪𝐅×1εt2ε(𝒪𝐅×)2β(1+tε)χ2(t)dtIε=2β(ε)χ1(ε)η0n0(1)𝒪𝐅×1εt2(𝒪𝐅×)2β(1+tε)χ2(t)dt.I_{1}=I_{\varepsilon}=2\chi_{0}\chi(-1)\int_{\mathcal{O}_{\mathbf{F}}^{\times}-\sideset{}{{}_{\pm}}{\bigcup}(\pm 1+\mathcal{P}_{\mathbf{F}})}\chi_{0}\left(\tfrac{1+t}{1-t}\right)\chi^{-2}(t)\eta_{0}^{n_{0}}(1-t^{2})\mathrm{d}t\quad{\rm resp.}\\ \begin{cases}I_{1}=2\beta(\sqrt{\varepsilon})\chi^{-1}(-\varepsilon)\eta_{0}^{n_{0}}(-\varepsilon)\int_{\begin{subarray}{c}\mathcal{O}_{\mathbf{F}}^{\times}\\ 1-\varepsilon t^{2}\in-\varepsilon(\mathcal{O}_{\mathbf{F}}^{\times})^{2}\end{subarray}}\beta(1+t\sqrt{\varepsilon})\chi^{-2}(t)\mathrm{d}t\\ I_{\varepsilon}=2\beta(\sqrt{\varepsilon})\chi^{-1}(-\varepsilon)\eta_{0}^{n_{0}}(-1)\int_{\begin{subarray}{c}\mathcal{O}_{\mathbf{F}}^{\times}\\ 1-\varepsilon t^{2}\in-(\mathcal{O}_{\mathbf{F}}^{\times})^{2}\end{subarray}}\beta(1+t\sqrt{\varepsilon})\chi^{-2}(t)\mathrm{d}t\end{cases}.

We apply Lemma 5.7 (2) to bound the above integrals and conclude the desired inequalities. ∎

Lemma 5.11.

Let n0max(2𝔠(Π),𝔠(Π)+2)\displaystyle n_{0}\geqslant\max(2\mathfrak{c}(\Pi),\mathfrak{c}(\Pi)+2) and m𝔠(Π)\displaystyle m\leqslant\mathfrak{c}(\Pi). Then for any δϖ𝐅2(n0m)𝒪𝐅×\displaystyle\delta\in\varpi_{\mathbf{F}}^{2(n_{0}-m)}\mathcal{O}_{\mathbf{F}}^{\times} we have

𝐅𝒪𝐅𝒱Π,ψ𝔪1(𝔱(δ).Fm)(t)ψ(t)χ1(t)|t|12d×t=0.\displaystyle\int_{\mathbf{F}-\mathcal{O}_{\mathbf{F}}}\mathcal{VH}_{\Pi,\psi}\circ\mathfrak{m}_{-1}\left(\mathfrak{t}(\delta).F_{m}\right)(t)\cdot\psi(-t)\chi^{-1}(t)\lvert t\rvert^{-\frac{1}{2}}\mathrm{d}^{\times}t=0.

Consequently, we get for τ{1,ε}\displaystyle\tau\in\{1,\varepsilon\} and any unitary χ\displaystyle\chi the vanishing of h~2n0,ma,τ(χ)=0\displaystyle\widetilde{h}_{2n_{0},m}^{a,\tau}(\chi)=0.

Proof.

For simplicity we only consider the case m2\displaystyle m\geqslant 2. By the local functional equation and Lemma 5.3 we have (recall the parameters of L(s,Πξ)\displaystyle L(s,\Pi\otimes\xi) introduced in (4.14))

𝐅×𝒱Π,ψ(Fm)(t)ξ1(t)|t|sd×t=𝟙𝔠(ξ)=mζ𝐅(1)γ(1,ξ2,ψ)ε(12,Πξ,ψ)q𝔠(Πξ)2q(2mρ(Πξ))sj=1d(Πξ)qsaj1aj¯q1+s.\int_{\mathbf{F}^{\times}}\mathcal{VH}_{\Pi,\psi}(F_{m})(t)\xi^{-1}(t)\lvert t\rvert^{-s}\mathrm{d}^{\times}t=\mathbbm{1}_{\mathfrak{c}(\xi)=m}\cdot\zeta_{\mathbf{F}}(1)\gamma\left(1,\xi^{-2},\psi\right)\varepsilon\left(\tfrac{1}{2},\Pi\otimes\xi,\psi\right)q^{\frac{\mathfrak{c}(\Pi\otimes\xi)}{2}}\\ \cdot q^{(2m-\mathfrak{\rho}(\Pi\otimes\xi))s}\sideset{}{{}_{j=1}^{d(\Pi\otimes\xi)}}{\prod}\frac{q^{s}-a_{j}}{1-\overline{a_{j}}q^{-1+s}}.

As a Laurent series in X:=qs\displaystyle X:=q^{s}, the right hand side contains Xk\displaystyle X^{k} only for k2mρ(Πξ)\displaystyle k\geqslant 2m-\mathfrak{\rho}(\Pi\otimes\xi). We have

ρ(Πξ){=𝔠(Πξ)𝔠(Π)+3𝔠(ξ)4𝔠(Π)if ξE(Π)2𝔠(Π)+3if ξE(Π)\displaystyle\mathfrak{\rho}(\Pi\otimes\xi)\begin{cases}=\mathfrak{c}(\Pi\otimes\xi)\leqslant\mathfrak{c}(\Pi)+3\mathfrak{c}(\xi)\leqslant 4\mathfrak{c}(\Pi)&\text{if }\xi\notin E(\Pi)\\ \leqslant 2\mathfrak{c}(\Pi)+3&\text{if }\xi\in E(\Pi)\end{cases}

by [4] and Lemma 4.4 respectively. Therefore the support of 𝒱Π,ψ𝔪1(𝔱(δ).Fm)\displaystyle\mathcal{VH}_{\Pi,\psi}\circ\mathfrak{m}_{-1}\left(\mathfrak{t}(\delta).F_{m}\right) is contained in

δ𝒫𝐅2mmax(4𝔠(Π),2𝔠(Π)+3)=𝒫𝐅2n0max(4𝔠(Π),2𝔠(Π)+3)𝒪𝐅,\displaystyle\delta\mathcal{P}_{\mathbf{F}}^{2m-\max(4\mathfrak{c}(\Pi),2\mathfrak{c}(\Pi)+3)}=\mathcal{P}_{\mathbf{F}}^{2n_{0}-\max(4\mathfrak{c}(\Pi),2\mathfrak{c}(\Pi)+3)}\subset\mathcal{O}_{\mathbf{F}},

proving the desired vanishing. ∎

5.3. The Bounds of Dual Weight

Lemma 5.12.

Suppose 𝔠(Π)=0\displaystyle\mathfrak{c}(\Pi)=0 and n0=1\displaystyle n_{0}=1. Then we have h~2(χ)q32𝟙𝔠(χ)=0+q1𝟙𝔠(χ)=1\displaystyle\widetilde{h}_{2}^{-}(\chi)\leqslant q^{-\frac{3}{2}}\mathbbm{1}_{\mathfrak{c}(\chi)=0}+q^{-1}\mathbbm{1}_{\mathfrak{c}(\chi)=1}.

Proof.

Necessarily we have Π=μ1μ2μ3\displaystyle\Pi=\mu_{1}\boxplus\mu_{2}\boxplus\mu_{3} for unramified μj\displaystyle\mu_{j}. Note that H2a\displaystyle H_{2}^{a} (resp. H2b\displaystyle H_{2}^{b}) is related to F0\displaystyle F_{0} (resp. F1\displaystyle F_{1}) by the formulae:

H2a=ε𝐋𝟙η0(1)=ε𝐋β(1)(𝔱(ϖ𝐅2).F0)+ε𝐋𝟙η0(1)=ε𝐋β(ε)(𝔱(ε1ϖ𝐅2).F0),\displaystyle H_{2}^{a}=\varepsilon_{\mathbf{L}}\mathbbm{1}_{\eta_{0}(-1)=\varepsilon_{\mathbf{L}}}\beta(\sqrt{-1})\cdot\left(\mathfrak{t}(\varpi_{\mathbf{F}}^{2}).F_{0}\right)+\varepsilon_{\mathbf{L}}\mathbbm{1}_{\eta_{0}(-1)=-\varepsilon_{\mathbf{L}}}\beta(\sqrt{-\varepsilon})\cdot\left(\mathfrak{t}(\varepsilon^{-1}\varpi_{\mathbf{F}}^{2}).F_{0}\right),
H2b=ε𝐋q2𝐋1𝒪𝐋Tr(α)𝒪𝐅×β(α)(𝔱(Tr(α)2).F1)dα+ε𝐋q2𝐋1𝒪𝐋Tr(xεα)𝒪𝐅×β(xεα)(𝔱(ε1Tr(xεα)2).F1)dα.\displaystyle H_{2}^{b}=\tfrac{\varepsilon_{\mathbf{L}}q}{2}\int_{\begin{subarray}{c}\mathbf{L}^{1}\cap\mathcal{O}_{\mathbf{L}}\\ {\rm Tr}(\alpha)\in\mathcal{O}_{\mathbf{F}}^{\times}\end{subarray}}\beta(\alpha)\cdot\left(\mathfrak{t}({\rm Tr}(\alpha)^{2}).F_{1}\right)\mathrm{d}\alpha+\tfrac{\varepsilon_{\mathbf{L}}q}{2}\int_{\begin{subarray}{c}\mathbf{L}^{1}\cap\mathcal{O}_{\mathbf{L}}\\ {\rm Tr}(x_{\varepsilon}\alpha)\in\mathcal{O}_{\mathbf{F}}^{\times}\end{subarray}}\beta(x_{\varepsilon}\alpha)\cdot\left(\mathfrak{t}(\varepsilon^{-1}{\rm Tr}(x_{\varepsilon}\alpha)^{2}).F_{1}\right)\mathrm{d}\alpha.

For any δ0ϖ𝐅2𝒪𝐅×\displaystyle\delta_{0}\in\varpi_{\mathbf{F}}^{2}\mathcal{O}_{\mathbf{F}}^{\times} and δ1𝒪𝐅×\displaystyle\delta_{1}\in\mathcal{O}_{\mathbf{F}}^{\times} we have

𝐅𝒪𝐅𝒱Π,ψ𝔪1(𝔱(δ0).F0)(t)ψ(t)χ(t)|t|12d×t=q52ϖ𝐅1𝒪𝐅×𝒱Π,ψ(F0)(δ01t)ψ(t)χ(t)d×t,\displaystyle\int_{\mathbf{F}-\mathcal{O}_{\mathbf{F}}}\mathcal{VH}_{\Pi,\psi}\circ\mathfrak{m}_{-1}\left(\mathfrak{t}(\delta_{0}).F_{0}\right)(t)\psi(-t)\chi(t)\lvert t\rvert^{-\frac{1}{2}}\mathrm{d}^{\times}t=q^{-\frac{5}{2}}\int_{\varpi_{\mathbf{F}}^{-1}\mathcal{O}_{\mathbf{F}}^{\times}}\mathcal{VH}_{\Pi,\psi}(F_{0})(\delta_{0}^{-1}t)\psi(-t)\chi(t)\mathrm{d}^{\times}t,
𝐅𝒪𝐅𝒱Π,ψ𝔪1(𝔱(δ1).F1)(t)ψ(t)χ(t)|t|12d×t=q52ϖ𝐅1𝒪𝐅×𝒱Π,ψ(F1)(δ11t)ψ(t)χ(t)d×t,\displaystyle\int_{\mathbf{F}-\mathcal{O}_{\mathbf{F}}}\mathcal{VH}_{\Pi,\psi}\circ\mathfrak{m}_{-1}\left(\mathfrak{t}(\delta_{1}).F_{1}\right)(t)\psi(-t)\chi(t)\lvert t\rvert^{-\frac{1}{2}}\mathrm{d}^{\times}t=q^{-\frac{5}{2}}\int_{\varpi_{\mathbf{F}}^{-1}\mathcal{O}_{\mathbf{F}}^{\times}}\mathcal{VH}_{\Pi,\psi}(F_{1})(\delta_{1}^{-1}t)\psi(-t)\chi(t)\mathrm{d}^{\times}t,

by inspecting the supports of 𝒱Π,ψ(F0)\displaystyle\mathcal{VH}_{\Pi,\psi}(F_{0}) and 𝒱Π,ψ(F1)\displaystyle\mathcal{VH}_{\Pi,\psi}(F_{1}) given in Corollary 5.6. Define

K(y):=τ0q𝟙ϖ𝐅1𝒪𝐅×(y)(ϖ𝐅1𝒪𝐅×)2ψ(t1+t2t1t24y)η0(4yt1t2)dt1dt2,\displaystyle K(y):=\tau_{0}q\mathbbm{1}_{\varpi_{\mathbf{F}}^{-1}\mathcal{O}_{\mathbf{F}}^{\times}}(y)\int_{(\varpi_{\mathbf{F}}^{-1}\mathcal{O}_{\mathbf{F}}^{\times})^{2}}\psi\left(t_{1}+t_{2}-\tfrac{t_{1}t_{2}}{4y}\right)\eta_{0}\left(\tfrac{4y}{t_{1}t_{2}}\right)\mathrm{d}t_{1}\mathrm{d}t_{2},
K~(δ,χ):=q52ϖ𝐅1𝒪𝐅×K(δ1t)ψ(t)χ(t)d×t.\displaystyle\widetilde{K}(\delta,\chi):=q^{-\frac{5}{2}}\int_{\varpi_{\mathbf{F}}^{-1}\mathcal{O}_{\mathbf{F}}^{\times}}K(\delta^{-1}t)\psi(-t)\chi(t)\mathrm{d}^{\times}t.
I1(χ):=𝐋1𝒪𝐋Tr(α)𝒪𝐅×β(α)K~(Tr(α)2,χ)dα,I2(χ):=𝐋1𝒪𝐋Tr(xεα)𝒪𝐅×β(xεα)K~(ε1Tr(xεα)2,χ)dα.\displaystyle I_{1}(\chi):=\int_{\begin{subarray}{c}\mathbf{L}^{1}\cap\mathcal{O}_{\mathbf{L}}\\ {\rm Tr}(\alpha)\in\mathcal{O}_{\mathbf{F}}^{\times}\end{subarray}}\beta(\alpha)\widetilde{K}({\rm Tr}(\alpha)^{2},\chi)\mathrm{d}\alpha,\quad I_{2}(\chi):=\int_{\begin{subarray}{c}\mathbf{L}^{1}\cap\mathcal{O}_{\mathbf{L}}\\ {\rm Tr}(x_{\varepsilon}\alpha)\in\mathcal{O}_{\mathbf{F}}^{\times}\end{subarray}}\beta(x_{\varepsilon}\alpha)\widetilde{K}(\varepsilon^{-1}{\rm Tr}(x_{\varepsilon}\alpha)^{2},\chi)\mathrm{d}\alpha.

The formulae of 𝒱Π,ψ(F0)\displaystyle\mathcal{VH}_{\Pi,\psi}(F_{0}) and 𝒱Π,ψ(F1)\displaystyle\mathcal{VH}_{\Pi,\psi}(F_{1}) in Corollary 5.6 easily imply the bound

(5.16) h~2(χ)q52𝟙𝔠(χ)1+q|I1(χ)+I2(χ)|.\widetilde{h}_{2}^{-}(\chi)\ll q^{-\frac{5}{2}}\mathbbm{1}_{\mathfrak{c}(\chi)\leqslant 1}+q\left\lvert I_{1}(\chi)+I_{2}(\chi)\right\rvert.

An easy change of variables gives

(5.17) K~(δ,χ)=τ0q12ζ𝐅(1)(𝒪𝐅×)3ψ(t1+t2t1t2δ4ttϖ𝐅)η0(4tδt1t2)χ(tϖ𝐅)dt1dt2dt=τ0q32ζ𝐅(1)(𝒪𝐅×)4t1t24δ1t3t4(mod𝒫𝐅)ψ(t1+t2t3t4ϖ𝐅)χ(t3ϖ𝐅)η0(t4)dt1dt2dt3dt4.\widetilde{K}(\delta,\chi)=\tau_{0}q^{\frac{1}{2}}\zeta_{\mathbf{F}}(1)\int_{(\mathcal{O}_{\mathbf{F}}^{\times})^{3}}\psi\left(\tfrac{t_{1}+t_{2}-\frac{t_{1}t_{2}\delta}{4t}-t}{\varpi_{\mathbf{F}}}\right)\eta_{0}\left(\tfrac{4t}{\delta t_{1}t_{2}}\right)\chi\left(\tfrac{t}{\varpi_{\mathbf{F}}}\right)\mathrm{d}t_{1}\mathrm{d}t_{2}\mathrm{d}t\\ =\tau_{0}q^{\frac{3}{2}}\zeta_{\mathbf{F}}(1)\int_{\begin{subarray}{c}(\mathcal{O}_{\mathbf{F}}^{\times})^{4}\\ t_{1}t_{2}\equiv 4\delta^{-1}t_{3}t_{4}\pmod{\mathcal{P}_{\mathbf{F}}}\end{subarray}}\psi\left(\tfrac{t_{1}+t_{2}-t_{3}-t_{4}}{\varpi_{\mathbf{F}}}\right)\chi\left(\tfrac{t_{3}}{\varpi_{\mathbf{F}}}\right)\eta_{0}(t_{4})\mathrm{d}t_{1}\mathrm{d}t_{2}\mathrm{d}t_{3}\mathrm{d}t_{4}.

Note that the function tK(δ1t)ψ(t)\displaystyle t\mapsto K(\delta^{-1}t)\psi(-t) is invariant by 1+𝒫𝐅\displaystyle 1+\mathcal{P}_{\mathbf{F}}, hence K~(δ,χ)\displaystyle\widetilde{K}(\delta,\chi) is non-zero only if 𝔠(χ)1\displaystyle\mathfrak{c}(\chi)\leqslant 1. If 𝔠(χ)=0\displaystyle\mathfrak{c}(\chi)=0 then one simplifies (5.17) by performing the integrals over t1,t3,t4,t2\displaystyle t_{1},t_{3},t_{4},t_{2} in order

(5.18) K~(δ,χ)=χ(ϖ𝐅)1ζ𝐅(1)(q52+q32𝟙𝒪𝐅×(δ4)η0(δδ4)).\widetilde{K}(\delta,\chi)=\chi(\varpi_{\mathbf{F}})^{-1}\zeta_{\mathbf{F}}(1)\left(q^{-\frac{5}{2}}+q^{-\frac{3}{2}}\cdot\mathbbm{1}_{\mathcal{O}_{\mathbf{F}}^{\times}}(\delta-4)\eta_{0}\left(\tfrac{\delta}{\delta-4}\right)\right).

Inserting (5.18) into (5.16) we see that the integrands K~()\displaystyle\widetilde{K}(\cdot) are constant, equal to χ(ϖ𝐅)1ζ𝐅(1)q32(q1+ε𝐋)\displaystyle\chi(\varpi_{\mathbf{F}})^{-1}\zeta_{\mathbf{F}}(1)q^{-\frac{3}{2}}(q^{-1}+\varepsilon_{\mathbf{L}}), in both integrals. We obtain the bound h~2(χ)q32\displaystyle\widetilde{h}_{2}^{-}(\chi)\ll q^{-\frac{3}{2}} by

𝐋1𝒪𝐋Tr(α)𝒪𝐅×β(α)dα=q1𝟙η0(1)=ε𝐋β(1),𝐋1𝒪𝐋Tr(xεα)𝒪𝐅×β(xεα)dα=q1𝟙η0(1)=ε𝐋β(ε).\displaystyle\int_{\begin{subarray}{c}\mathbf{L}^{1}\cap\mathcal{O}_{\mathbf{L}}\\ {\rm Tr}(\alpha)\in\mathcal{O}_{\mathbf{F}}^{\times}\end{subarray}}\beta(\alpha)\mathrm{d}\alpha=-q^{-1}\mathbbm{1}_{\eta_{0}(-1)=\varepsilon_{\mathbf{L}}}\beta(\sqrt{-1}),\quad\int_{\begin{subarray}{c}\mathbf{L}^{1}\cap\mathcal{O}_{\mathbf{L}}\\ {\rm Tr}(x_{\varepsilon}\alpha)\in\mathcal{O}_{\mathbf{F}}^{\times}\end{subarray}}\beta(x_{\varepsilon}\alpha)\mathrm{d}\alpha=-q^{-1}\mathbbm{1}_{\eta_{0}(-1)=-\varepsilon_{\mathbf{L}}}\beta(\sqrt{-\varepsilon}).

If 𝔠(χ)=1\displaystyle\mathfrak{c}(\chi)=1, then regarding χ\displaystyle\chi and η0\displaystyle\eta_{0} as characters of 𝔽q×\displaystyle\mathbb{F}_{q}^{\times} we rewrite K~\displaystyle\widetilde{K} as

K~(δ,χ)=τ0q1ζ𝐅(1)H(4δ1,q;(𝟙,𝟙),(χ1,η0)),\displaystyle\widetilde{K}(\delta,\chi)=-\tau_{0}q^{-1}\zeta_{\mathbf{F}}(1)\cdot H(4\delta^{-1},q;(\mathbbm{1},\mathbbm{1}),(\chi^{-1},\eta_{0})),

where H()\displaystyle H(\cdot) is precisely the hypergeometric sum of Katz appeared in [21, §5.4]. In the case of split, resp. unramified 𝐋/𝐅\displaystyle\mathbf{L}/\mathbf{F} we apply the change of variables t=αα1α+α1\displaystyle t=\tfrac{\alpha-\alpha^{-1}}{\alpha+\alpha^{-1}}, resp. t=αα1α+α1ε\displaystyle t=\tfrac{\alpha-\alpha^{-1}}{\alpha+\alpha^{-1}}\sqrt{\varepsilon} for I1(χ)\displaystyle I_{1}(\chi); t=αεα1α+εα1\displaystyle t=\tfrac{\alpha-\varepsilon\alpha^{-1}}{\alpha+\varepsilon\alpha^{-1}}, resp. t=xεαxεα¯xεα+xεα¯ε\displaystyle t=\tfrac{x_{\varepsilon}\alpha-\overline{x_{\varepsilon}\alpha}}{x_{\varepsilon}\alpha+\overline{x_{\varepsilon}\alpha}}\sqrt{\varepsilon} for I2(χ)\displaystyle I_{2}(\chi). These integrals become

I1(χ)=I2(χ)=2𝒪𝐅±(±1+𝒫𝐅)χ0(1+t1t)K~(41t2,χ)dtresp.{I1(χ)=2β(ε)1𝒪𝐅β(t+ε)K~(41ε1t2,χ)𝟙(𝒪𝐅×)2(1ε1t2)dtI2(χ)=2β(ε)1𝒪𝐅β(t+ε)K~(41ε1t2,χ)𝟙ε(𝒪𝐅×)2(1ε1t2)dt.I_{1}(\chi)=I_{2}(\chi)=2\int_{\mathcal{O}_{\mathbf{F}}-\sideset{}{{}_{\pm}}{\bigcup}(\pm 1+\mathcal{P}_{\mathbf{F}})}\chi_{0}\left(\tfrac{1+t}{1-t}\right)\widetilde{K}\left(\tfrac{4}{1-t^{2}},\chi\right)\mathrm{d}t\quad{\rm resp.}\\ \begin{cases}I_{1}(\chi)=2\beta(\sqrt{\varepsilon})^{-1}\int_{\mathcal{O}_{\mathbf{F}}}\beta(t+\sqrt{\varepsilon})\widetilde{K}\left(\tfrac{4}{1-\varepsilon^{-1}t^{2}},\chi\right)\mathbbm{1}_{(\mathcal{O}_{\mathbf{F}}^{\times})^{2}}(1-\varepsilon^{-1}t^{2})\mathrm{d}t\\ I_{2}(\chi)=2\beta(\sqrt{\varepsilon})^{-1}\int_{\mathcal{O}_{\mathbf{F}}}\beta(t+\sqrt{\varepsilon})\widetilde{K}\left(\tfrac{4}{1-\varepsilon^{-1}t^{2}},\chi\right)\mathbbm{1}_{\varepsilon(\mathcal{O}_{\mathbf{F}}^{\times})^{2}}(1-\varepsilon^{-1}t^{2})\mathrm{d}t\end{cases}.

In the case of unramified 𝐋/𝐅\displaystyle\mathbf{L}/\mathbf{F} we recognize and get the bound (with H()\displaystyle H(\cdot) defined in [21, §5.3.2])

q|I1(χ)+I2(χ)|=2ζ𝐅(1)q32|α𝔽qβ(α+ε)H(1α2ε1,q;(𝟙,𝟙),(χ1,η0))|q1\displaystyle q\left\lvert I_{1}(\chi)+I_{2}(\chi)\right\rvert=2\zeta_{\mathbf{F}}(1)q^{-\frac{3}{2}}\left\lvert\sideset{}{{}_{\alpha\in\mathbb{F}_{q}}}{\sum}\beta(\alpha+\sqrt{\varepsilon})H(1-\alpha^{2}\varepsilon^{-1},q;(\mathbbm{1},\mathbbm{1}),(\chi^{-1},\eta_{0}))\right\rvert\ll q^{-1}

by [21, Lemma 5.10]. In the case of split 𝐋/𝐅\displaystyle\mathbf{L}/\mathbf{F} we recognize and claim a bound of

(5.19) q|I1(χ)+I2(χ)|=4ζ𝐅(1)q32|α𝔽q{±1}χ0(α+1α1)H(1α2,q;(𝟙,𝟙),(χ1,η0))|q1.q\left\lvert I_{1}(\chi)+I_{2}(\chi)\right\rvert=4\zeta_{\mathbf{F}}(1)q^{-\frac{3}{2}}\left\lvert\sideset{}{{}_{\alpha\in\mathbb{F}_{q}-\{\pm 1\}}}{\sum}\chi_{0}\left(\tfrac{\alpha+1}{\alpha-1}\right)H(1-\alpha^{2},q;(\mathbbm{1},\mathbbm{1}),(\chi^{-1},\eta_{0}))\right\rvert\ll q^{-1}.

The method of the proof of [21, Lemma 5.10] works through to bound the above sum. In fact the essence of that proof is to show:

  • (1)

    the \displaystyle\ell-adic sheaf associated with the function αH(1α2ε1,q;(𝟙,𝟙),(χ1,η0))\displaystyle\alpha\mapsto H(1-\alpha^{2}\varepsilon^{-1},q;(\mathbbm{1},\mathbbm{1}),(\chi^{-1},\eta_{0})) has rank 2\displaystyle 2,

  • (2)

    while the sheaf associated with the function αβ(α+ε)\displaystyle\alpha\mapsto\beta(\alpha+\sqrt{\varepsilon}) has rank 1\displaystyle 1.

The analogues for the above split case are

  • (1’)

    the \displaystyle\ell-adic sheaf associated with the function αH(1α2,q;(𝟙,𝟙),(χ1,η0))\displaystyle\alpha\mapsto H(1-\alpha^{2},q;(\mathbbm{1},\mathbbm{1}),(\chi^{-1},\eta_{0})) has rank 2\displaystyle 2,

  • (2’)

    while the sheaf associated with the function αχ0(α+1α1)\displaystyle\alpha\mapsto\chi_{0}\left(\tfrac{\alpha+1}{\alpha-1}\right) has rank 1\displaystyle 1.

Note that (1’) and (1) follow from the same argument that the sheaf associated with tH(t,q;)\displaystyle t\mapsto H(t,q;\cdots) has geometric monodromy group SL2\displaystyle{\rm SL}_{2}, which does not admit finite index subgroup; while (2’) is trivially true because the Kummer sheaf has rank 1\displaystyle 1. In fact, both (2) and (2’) are known to Weil [18, Appendix \@slowromancapv@]. We conclude the claimed bound (5.19). ∎

Proposition 5.13.

With the test function H\displaystyle H given by (3.5) the dual weight function is bounded as

h~(χ)ϵ𝐂(Π)2+ϵqn0+ϵ(𝟙max(n0,6𝔠(Π))(𝔠(χ))+q12𝟙η0(1)=ε𝐋𝟙2n03𝟙𝔠(χ)=n0𝟙𝓔(β)(χ2)),\displaystyle\widetilde{h}(\chi)\ll_{\epsilon}\mathbf{C}(\Pi)^{2+\epsilon}q^{-n_{0}+\epsilon}\left(\mathbbm{1}_{\leqslant\max(n_{0},6\mathfrak{c}(\Pi))}(\mathfrak{c}(\chi))+q^{\frac{1}{2}}\mathbbm{1}_{\eta_{0}(-1)=\varepsilon_{\mathbf{L}}}\mathbbm{1}_{2\nmid n_{0}\geqslant 3}\mathbbm{1}_{\mathfrak{c}(\chi)=n_{0}}\mathbbm{1}_{\boldsymbol{\mathcal{E}}(\beta)}(\chi^{2})\right),

where we recall ε𝐋=η𝐋/𝐅(ϖ𝐅)\displaystyle\varepsilon_{\mathbf{L}}=\eta_{\mathbf{L}/\mathbf{F}}(\varpi_{\mathbf{F}}), and 𝓔(β)\displaystyle\boldsymbol{\mathcal{E}}(\beta) is given in Lemma 5.7 (2).

Proof.

We distinguish two cases 𝔠(Π)>0\displaystyle\mathfrak{c}(\Pi)>0 and 𝔠(Π)=0\displaystyle\mathfrak{c}(\Pi)=0.

(1) Suppose 𝔠(Π)>0\displaystyle\mathfrak{c}(\Pi)>0. We have a(Π)𝔠(Π)\displaystyle a(\Pi)\leqslant\mathfrak{c}(\Pi) by Proposition 2.4. If n02𝔠(Π)\displaystyle n_{0}\leqslant 2\mathfrak{c}(\Pi), then we may take n1=4𝔠(Π)\displaystyle n_{1}=4\mathfrak{c}(\Pi). Applying Lemma 4.1, 4.6 and 4.8 we deduce

h~(χ)|h~(χ)|+|h~c+(χ)|+|h~c(χ)|𝔠(Π)𝟙6𝔠(Π)(𝔠(χ)),\displaystyle\widetilde{h}(\chi)\ll\left\lvert\widetilde{h}_{\infty}(\chi)\right\rvert+\left\lvert\widetilde{h}_{c}^{+}(\chi)\right\rvert+\left\lvert\widetilde{h}_{c}^{-}(\chi)\right\rvert\ll\mathfrak{c}(\Pi)\mathbbm{1}_{\leqslant 6\mathfrak{c}(\Pi)}(\mathfrak{c}(\chi)),

the “worst” bound being offered by Lemma 4.8. If n02𝔠(Π)\displaystyle n_{0}\geqslant 2\mathfrak{c}(\Pi), then we may take n1=2n0\displaystyle n_{1}=2n_{0}. Applying Lemma 4.1, 4.6 and Lemma 5.8-5.11 we deduce

h~(χ)qn0(𝟙n0(𝔠(χ))+q12𝟙η0(1)=ε𝐋𝟙2n03𝟙𝔠(χ)=n0𝟙𝓔(β)(χ)).\displaystyle\widetilde{h}(\chi)\ll q^{-n_{0}}\left(\mathbbm{1}_{\leqslant n_{0}}(\mathfrak{c}(\chi))+q^{\frac{1}{2}}\mathbbm{1}_{\eta_{0}(-1)=\varepsilon_{\mathbf{L}}}\mathbbm{1}_{2\nmid n_{0}\geqslant 3}\mathbbm{1}_{\mathfrak{c}(\chi)=n_{0}}\mathbbm{1}_{\boldsymbol{\mathcal{E}}(\beta)}(\chi)\right).

The stated bound is simply a common upper bound of the above two.

(2) Suppose 𝔠(Π)=0\displaystyle\mathfrak{c}(\Pi)=0. Necessarily we have Π=μ1μ2μ3\displaystyle\Pi=\mu_{1}\boxplus\mu_{2}\boxplus\mu_{3} for unramified μj\displaystyle\mu_{j} and a(Π)=2\displaystyle a(\Pi)=2. The argument of the second part in (1) works through also for n02\displaystyle n_{0}\geqslant 2. It remains to consider the case n0=1\displaystyle n_{0}=1. Note that n1=2\displaystyle n_{1}=2, hence the contribution of H\displaystyle H_{\infty} is h~(χ)q1𝟙𝔠(χ)=0\displaystyle\widetilde{h}_{\infty}(\chi)\ll q^{-1}\mathbbm{1}_{\mathfrak{c}(\chi)=0} by Lemma 4.1. We have

h~(χ)=h~(χ)+h~2+(χ)+h~2(χ)\displaystyle\widetilde{h}(\chi)=\widetilde{h}_{\infty}(\chi)+\widetilde{h}_{2}^{+}(\chi)+\widetilde{h}_{2}^{-}(\chi)

by (4.3) and the condition 2n\displaystyle 2\mid n following (5.1). By Lemma 4.6 we have h~2+(χ)ϵq1+ϵ𝟙𝔠(χ)=0\displaystyle\widetilde{h}_{2}^{+}(\chi)\ll_{\epsilon}q^{-1+\epsilon}\mathbbm{1}_{\mathfrak{c}(\chi)=0}. By Lemma 5.12 we have h~2(χ)q32𝟙𝔠(χ)=0+q1𝟙𝔠(χ)=1\displaystyle\widetilde{h}_{2}^{-}(\chi)\ll q^{-\frac{3}{2}}\mathbbm{1}_{\mathfrak{c}(\chi)=0}+q^{-1}\mathbbm{1}_{\mathfrak{c}(\chi)=1}. The stated bound follows. ∎

5.4. The Bounds of Unramified Dual Weight

Lemma 5.14.

Suppose n0+1n2n01\displaystyle n_{0}+1\leqslant n\leqslant 2n_{0}-1 and na(Π)\displaystyle n\geqslant a(\Pi). Then the function h~2n(||𝐅s)0\displaystyle\widetilde{h}_{2n}^{-}(\lvert\cdot\rvert_{\mathbf{F}}^{s})\neq 0 is non-vanishing only if n=2n01\displaystyle n=2n_{0}-1. Moreover for any k0\displaystyle k\in\mathbb{Z}_{\geqslant 0} we have

H~2(2n01)(k;12)k,ϵq12+(2n01)ϵ,H~2(2n01)(k;12)k,ϵq122n0+(2n01)ϵ.\displaystyle\widetilde{H}_{2(2n_{0}-1)}^{-}(k;\tfrac{1}{2})\ll_{k,\epsilon}q^{-\frac{1}{2}+\left(2n_{0}-1\right)\epsilon},\quad\widetilde{H}_{2(2n_{0}-1)}^{-}(k;-\tfrac{1}{2})\ll_{k,\epsilon}q^{\frac{1}{2}-2n_{0}+\left(2n_{0}-1\right)\epsilon}.
Proof.

By Corollary 5.4 and (5.2) the support of 𝒱Π,ψ(Fn)\displaystyle\mathcal{VH}_{\Pi,\psi}(F_{n}), hence also 𝒱Π,ψ𝔪1(H2n)\displaystyle\mathcal{VH}_{\Pi,\psi}\circ\mathfrak{m}_{-1}(H_{2n}) are contained in ϖ𝐅n𝒪𝐅×\displaystyle\varpi_{\mathbf{F}}^{-n}\mathcal{O}_{\mathbf{F}}^{\times}. Therefore we get by (5.8) the relation h~2n(||𝐅s)=qnsh~2n(𝟙)\displaystyle\widetilde{h}_{2n}^{-}(\lvert\cdot\rvert_{\mathbf{F}}^{s})=q^{ns}\widetilde{h}_{2n}^{-}(\mathbbm{1}). The stated results follow readily from Lemma 5.8. ∎

Lemma 5.15.

Suppose 𝔠(Π)=0\displaystyle\mathfrak{c}(\Pi)=0 and n0=1\displaystyle n_{0}=1. Then we have for any k0\displaystyle k\in\mathbb{Z}_{\geqslant 0}

H~2(k;12)k,ϵq(1ϵ),H~2(k;12)k,ϵq2(1ϵ).\displaystyle\widetilde{H}_{2}^{-}(k;\tfrac{1}{2})\ll_{k,\epsilon}q^{-(1-\epsilon)},\quad\widetilde{H}_{2}^{-}(k;-\tfrac{1}{2})\ll_{k,\epsilon}q^{-2(1-\epsilon)}.
Proof.

Similarly as the previous lemma, we have h~2(||𝐅s)=qsh~2(𝟙)\displaystyle\widetilde{h}_{2}^{-}(\lvert\cdot\rvert_{\mathbf{F}}^{s})=q^{s}\cdot\widetilde{h}_{2}^{-}(\mathbbm{1}) by the inspection of the supports of 𝒱Π,ψ𝔪1(H2)\displaystyle\mathcal{VH}_{\Pi,\psi}\circ\mathfrak{m}_{-1}(H_{2}) via Corollary 5.6 in the proof of Lemma 5.12, from which the stated bounds follow. ∎

Proposition 5.16.

With the test function H\displaystyle H given by (3.5) we have for any k0\displaystyle k\in\mathbb{Z}_{\geqslant 0}

H~(k;12)k,ϵ𝐂(Π)12q2n0ϵ,H~(k;12)k,ϵ𝐂(Π)4+ϵq2n0ϵ.\displaystyle\widetilde{H}(k;\tfrac{1}{2})\ll_{k,\epsilon}\mathbf{C}(\Pi)^{\frac{1}{2}}\cdot q^{2n_{0}\epsilon},\quad\widetilde{H}(k;-\tfrac{1}{2})\ll_{k,\epsilon}\mathbf{C}(\Pi)^{4+\epsilon}\cdot q^{2n_{0}\epsilon}.
Proof.

The argument is quite similar to Proposition 5.13. We distinguish two cases 𝔠(Π)>0\displaystyle\mathfrak{c}(\Pi)>0 and 𝔠(Π)=0\displaystyle\mathfrak{c}(\Pi)=0.

(1) Suppose 𝔠(Π)>0\displaystyle\mathfrak{c}(\Pi)>0. We take n1=4𝔠(Π)\displaystyle n_{1}=4\mathfrak{c}(\Pi) if n02𝔠(Π)\displaystyle n_{0}\leqslant 2\mathfrak{c}(\Pi) and apply Lemma 4.10 & 4.11. If n02𝔠(Π)\displaystyle n_{0}\geqslant 2\mathfrak{c}(\Pi) we take n1=2n0\displaystyle n_{1}=2n_{0}. The proofs of Lemma 5.8-5.11 imply h~2n0(||𝐅s)=0\displaystyle\widetilde{h}_{2n_{0}}^{-}(\lvert\cdot\rvert_{\mathbf{F}}^{s})=0. Together with Lemma 4.10 & 5.14 we conclude the stated bounds in this case.

(2) Suppose 𝔠(Π)=0\displaystyle\mathfrak{c}(\Pi)=0. Necessarily we have Π=μ1μ2μ3\displaystyle\Pi=\mu_{1}\boxplus\mu_{2}\boxplus\mu_{3} for unramified μj\displaystyle\mu_{j} and a(Π)=2\displaystyle a(\Pi)=2. The argument of the second part in (1) works through also for n02\displaystyle n_{0}\geqslant 2. It remains to consider the case n0=1\displaystyle n_{0}=1. Note that n1=2\displaystyle n_{1}=2. We conclude the stated bounds by Lemma 4.10 & 5.15. ∎

6. Dual Weight Functions: Ramified Cases

6.1. Second Quadratic Elementary Functions

Definition 6.1.

Let 𝐋=𝐅[ϖ𝐅]\displaystyle\mathbf{L}=\mathbf{F}[\sqrt{\varpi_{\mathbf{F}}}] and recall the associated quadratic character η𝐋/𝐅\displaystyle\eta_{\mathbf{L}/\mathbf{F}}. For any n0\displaystyle n\in\mathbb{Z}_{\geqslant 0} we define the second quadratic elementary functions GnCc(𝐅×)\displaystyle G_{n}\in{\rm C}_{c}^{\infty}(\mathbf{F}^{\times}) by

Gn(y2):=𝟙v(y)=n±η𝐋/𝐅(±y)ψ(±y),\displaystyle G_{n}(y^{2}):=\mathbbm{1}_{v(y)=-n}\cdot\sideset{}{{}_{\pm}}{\sum}\eta_{\mathbf{L}/\mathbf{F}}(\pm y)\psi(\pm y),

and are supported in the subset of square elements of 𝐅×\displaystyle\mathbf{F}^{\times}.

Lemma 6.2.

Let χ\displaystyle\chi be a quasi-character of 𝐅×\displaystyle\mathbf{F}^{\times}. We have

𝐅×Gn(y)χ(y)d×y={𝟙𝔠(χ2)=nζ𝐅(1)γ(1,η𝐋/𝐅χ2,ψ)if n2𝟙𝔠(η𝐋/𝐅χ2)=1ζ𝐅(1)γ(1,η𝐋/𝐅χ2,ψ)𝟙𝔠(η𝐋/𝐅χ2)=0ζ𝐅(1)q1(η𝐋/𝐅χ2)(ϖ𝐅)if n=1𝟙𝔠(η𝐋/𝐅χ2)=0if n=0.\int_{\mathbf{F}^{\times}}G_{n}(y)\chi(y)\mathrm{d}^{\times}y=\\ \begin{cases}\mathbbm{1}_{\mathfrak{c}(\chi^{2})=n}\cdot\zeta_{\mathbf{F}}(1)\gamma(1,\eta_{\mathbf{L}/\mathbf{F}}\chi^{-2},\psi)&\text{if }n\geqslant 2\\ \mathbbm{1}_{\mathfrak{c}(\eta_{\mathbf{L}/\mathbf{F}}\chi^{2})=1}\cdot\zeta_{\mathbf{F}}(1)\gamma(1,\eta_{\mathbf{L}/\mathbf{F}}\chi^{-2},\psi)-\mathbbm{1}_{\mathfrak{c}(\eta_{\mathbf{L}/\mathbf{F}}\chi^{2})=0}\cdot\zeta_{\mathbf{F}}(1)q^{-1}(\eta_{\mathbf{L}/\mathbf{F}}\chi^{-2})(\varpi_{\mathbf{F}})&\text{if }n=1\\ \mathbbm{1}_{\mathfrak{c}(\eta_{\mathbf{L}/\mathbf{F}}\chi^{2})=0}&\text{if }n=0\end{cases}.
Proof.

The proof is quite similar to Lemma 5.3 and is omitted. ∎

Corollary 6.3.

Let na(Π)(2)\displaystyle n\geqslant a(\Pi)(\geqslant 2), the stability barrier of Π\displaystyle\Pi (see Proposition 2.4). We have

𝒱Π,ψ(Gn)(y)=η0(1)nη0(2)𝟙ϖ𝐅n𝒪𝐅×(y){q3n2ψ(4y)η0(4y)if 2nτ0q3n2ψ(4y)if 2n.\displaystyle\mathcal{VH}_{\Pi,\psi}(G_{n})(y)=\eta_{0}(-1)^{n}\eta_{0}(-2)\mathbbm{1}_{\varpi_{\mathbf{F}}^{-n}\mathcal{O}_{\mathbf{F}}^{\times}}(y)\cdot\begin{cases}q^{\frac{3n}{2}}\psi(4y)\eta_{0}(4y)&\text{if }2\mid n\\ \tau_{0}q^{\lceil\frac{3n}{2}\rceil}\psi(4y)&\text{if }2\nmid n\end{cases}.
Proof.

The proof is quite similar to Corollary 5.4. We only emphasize the difference. We get

𝒱Π,ψ(Gn)(yϖ𝐅n)=q3n(𝒪𝐅×)3η𝐋/𝐅(t3ϖ𝐅n)ψ(t1+t2+t3+t11t21t32yϖ𝐅n)dt1dt2dt3.\displaystyle\mathcal{VH}_{\Pi,\psi}(G_{n})\left(\tfrac{y}{\varpi_{\mathbf{F}}^{n}}\right)=q^{3n}\int_{(\mathcal{O}_{\mathbf{F}}^{\times})^{3}}\eta_{\mathbf{L}/\mathbf{F}}\left(\tfrac{t_{3}}{\varpi_{\mathbf{F}}^{n}}\right)\psi\left(\tfrac{t_{1}+t_{2}+t_{3}+t_{1}^{-1}t_{2}^{-1}t_{3}^{2}y}{\varpi_{\mathbf{F}}^{n}}\right)\mathrm{d}t_{1}\mathrm{d}t_{2}\mathrm{d}t_{3}.

Performing the level n2\displaystyle\lceil\tfrac{n}{2}\rceil regularization to dtj\displaystyle\mathrm{d}t_{j} we see that the non-vanishing of the above integral implies t1=4y(1+u1)\displaystyle t_{1}=4y(1+u_{1}), t2=4y(1+u2)\displaystyle t_{2}=4y(1+u_{2}) and t3=8y(1+u3)\displaystyle t_{3}=-8y(1+u_{3}) with uj𝒫𝐅n2\displaystyle u_{j}\in\mathcal{P}_{\mathbf{F}}^{\lfloor\frac{n}{2}\rfloor} and obtain

𝒱Π,ψ(Gn)(yϖ𝐅n)=q2nη𝐋/𝐅(8yϖ𝐅n)ψ(4yϖ𝐅n)𝒫𝐅n2ψ(4yϖ𝐅nu22)du2.\displaystyle\mathcal{VH}_{\Pi,\psi}(G_{n})\left(\tfrac{y}{\varpi_{\mathbf{F}}^{n}}\right)=q^{2n}\eta_{\mathbf{L}/\mathbf{F}}\left(\tfrac{-8y}{\varpi_{\mathbf{F}}^{n}}\right)\psi\left(\tfrac{4y}{\varpi_{\mathbf{F}}^{n}}\right)\int_{\mathcal{P}_{\mathbf{F}}^{\lfloor\frac{n}{2}\rfloor}}\psi\left(\tfrac{4y}{\varpi_{\mathbf{F}}^{n}}u_{2}^{2}\right)\mathrm{d}u_{2}.

We conclude by noting η𝐋/𝐅(ϖ𝐅ny)=η0(1)nη0(y)\displaystyle\eta_{\mathbf{L}/\mathbf{F}}\left(\varpi_{\mathbf{F}}^{-n}y\right)=\eta_{0}(-1)^{n}\eta_{0}(y) for y𝒪𝐅×\displaystyle y\in\mathcal{O}_{\mathbf{F}}^{\times}. ∎

Corollary 6.4.

Suppose Π=μ1μ2μ3\displaystyle\Pi=\mu_{1}\boxplus\mu_{2}\boxplus\mu_{3} with 𝔠(μj)=0\displaystyle\mathfrak{c}(\mu_{j})=0 and μ1μ2μ3=𝟙\displaystyle\mu_{1}\mu_{2}\mu_{3}=\mathbbm{1}. Let Ei(y):=μi1(y)|y|𝟙𝒪𝐅(y)\displaystyle E_{i}(y):=\mu_{i}^{-1}(y)\lvert y\rvert\mathbbm{1}_{\mathcal{O}_{\mathbf{F}}}(y) and fi:=(1μi(ϖ𝐅)𝔱(ϖ𝐅)).Ei\displaystyle f_{i}:=(1-\mu_{i}(\varpi_{\mathbf{F}})\mathfrak{t}(\varpi_{\mathbf{F}})).E_{i}. For f,gL1(𝐅×)\displaystyle f,g\in{\rm L}^{1}(\mathbf{F}^{\times}) define fg(y):=𝐅×f(yt1)g(t)d×t\displaystyle f*g(y):=\int_{\mathbf{F}^{\times}}f(yt^{-1})g(t)\mathrm{d}^{\times}t.

  • (1)

    We have G0=0\displaystyle G_{0}=0 unless 4q1\displaystyle 4\mid q-1, in which case {χ𝒪𝐅×^|η0=χ2}={η1,η11}\displaystyle\left\{\chi\in\widehat{\mathcal{O}_{\mathbf{F}}^{\times}}\ \middle|\ \eta_{0}=\chi^{2}\right\}=\{\eta_{1},\eta_{1}^{-1}\}. Extending η1\displaystyle\eta_{1} to 𝐅×\displaystyle\mathbf{F}^{\times} by η1(ϖ𝐅)=1\displaystyle\eta_{1}(\varpi_{\mathbf{F}})=1 we get

    𝒱Π,ψ(G0)(y)=𝟙ϖ𝐅3𝒪𝐅×(y)q32{γ(12,η1,ψ)3η1(y)+γ(12,η11,ψ)3η11(y)}.\displaystyle\mathcal{VH}_{\Pi,\psi}(G_{0})(y)=\mathbbm{1}_{\varpi_{\mathbf{F}}^{-3}\mathcal{O}_{\mathbf{F}}^{\times}}(y)\cdot q^{\frac{3}{2}}\cdot\left\{\gamma(\tfrac{1}{2},\eta_{1},\psi)^{3}\eta_{1}(y)+\gamma(\tfrac{1}{2},\eta_{1}^{-1},\psi)^{3}\eta_{1}^{-1}(y)\right\}.
  • (2)

    We have the formula

    𝒱Π,ψ(G1)(y)=ζ𝐅(1)γ(1,η𝐋/𝐅,ψ)(f1f2f3)(ϖ𝐅2y)+η𝐋/𝐅(1)𝟙ϖ𝐅1𝒪𝐅×(y){(ϖ𝐅1𝒪𝐅×)3η0(t3)ψ(t1+t2+t3+t32yt1t2)dt+ζ𝐅(1)τ0}.\mathcal{VH}_{\Pi,\psi}(G_{1})(y)=\zeta_{\mathbf{F}}(1)\gamma(1,\eta_{\mathbf{L}/\mathbf{F}},\psi)\cdot\left(f_{1}*f_{2}*f_{3}\right)(\varpi_{\mathbf{F}}^{-2}y)+\\ \eta_{\mathbf{L}/\mathbf{F}}(-1)\mathbbm{1}_{\varpi_{\mathbf{F}}^{-1}\mathcal{O}_{\mathbf{F}}^{\times}}(y)\cdot\left\{\int_{(\varpi_{\mathbf{F}}^{-1}\mathcal{O}_{\mathbf{F}}^{\times})^{3}}\eta_{0}(t_{3})\psi\left(t_{1}+t_{2}+t_{3}+\tfrac{t_{3}^{2}y}{t_{1}t_{2}}\right)\mathrm{d}\vec{t}+\zeta_{\mathbf{F}}(1)\tau_{0}\right\}.
Proof.

(1) From Lemma 6.2 and the local functional equation we deduce for any χ𝒪𝐅×^\displaystyle\chi\in\widehat{\mathcal{O}_{\mathbf{F}}^{\times}}

𝐅×𝒱Π,ψ(G0)(y)χ1(y)|y|sd×y=𝟙χ=η1±1q3(12s)γ(12,χ,ψ)3.\displaystyle\int_{\mathbf{F}^{\times}}\mathcal{VH}_{\Pi,\psi}(G_{0})(y)\chi^{-1}(y)\lvert y\rvert^{-s}\mathrm{d}^{\times}y=\mathbbm{1}_{\chi=\eta_{1}^{\pm 1}}\cdot q^{3\left(\frac{1}{2}-s\right)}\gamma\left(\tfrac{1}{2},\chi,\psi\right)^{3}.

We readily deduce the desired formula for 𝒱Π,ψ(G0)\displaystyle\mathcal{VH}_{\Pi,\psi}(G_{0}).

(2) We can write 𝒱Π,ψ(G1)=𝒱Π,ψ(G1)0+𝒱Π,ψ(G1)1\displaystyle\mathcal{VH}_{\Pi,\psi}(G_{1})=\mathcal{VH}_{\Pi,\psi}(G_{1})_{0}+\mathcal{VH}_{\Pi,\psi}(G_{1})_{1} with the properties

𝒱Π,ψ(G1)0(yδ)=𝒱Π,ψ(G1)0(y),y𝐅×,δ𝒪𝐅×;\displaystyle\mathcal{VH}_{\Pi,\psi}(G_{1})_{0}(y\delta)=\mathcal{VH}_{\Pi,\psi}(G_{1})_{0}(y),\quad\forall y\in\mathbf{F}^{\times},\delta\in\mathcal{O}_{\mathbf{F}}^{\times};
𝐅×𝒱Π,ψ(G1)0(y)|y|sd×y=ζ𝐅(1)γ(1,η𝐋/𝐅,ψ)q2si=13𝐅×fi(y)|y|sd×y,\displaystyle\int_{\mathbf{F}^{\times}}\mathcal{VH}_{\Pi,\psi}(G_{1})_{0}(y)\lvert y\rvert^{-s}\mathrm{d}^{\times}y=\zeta_{\mathbf{F}}(1)\gamma(1,\eta_{\mathbf{L}/\mathbf{F}},\psi)\cdot q^{2s}\cdot\sideset{}{{}_{i=1}^{3}}{\prod}\int_{\mathbf{F}^{\times}}f_{i}(y)\lvert y\rvert^{-s}\mathrm{d}^{\times}y,
𝐅×𝒱Π,ψ(G1)1(y)χ1(y)|y|sd×y=𝟙𝔠(χ)=1q3sϖ𝐅1𝒪𝐅×ψ(y)η𝐋/𝐅χ2(y)d×y(ϖ𝐅1𝒪𝐅×ψ(y)χ1(y)d×y)3.\int_{\mathbf{F}^{\times}}\mathcal{VH}_{\Pi,\psi}(G_{1})_{1}(y)\chi^{-1}(y)\lvert y\rvert^{-s}\mathrm{d}^{\times}y=\\ \mathbbm{1}_{\mathfrak{c}(\chi)=1}\cdot q^{3-s}\int_{\varpi_{\mathbf{F}}^{-1}\mathcal{O}_{\mathbf{F}}^{\times}}\psi(y)\eta_{\mathbf{L}/\mathbf{F}}\chi^{2}(y)\mathrm{d}^{\times}y\cdot\left(\int_{\varpi_{\mathbf{F}}^{-1}\mathcal{O}_{\mathbf{F}}^{\times}}\psi(y)\chi^{-1}(y)\mathrm{d}^{\times}y\right)^{3}.

We easily identify 𝒱Π,ψ(G1)0(y)=ζ𝐅(1)γ(1,η𝐋/𝐅,ψ)(f1f2f3)(ϖ𝐅2y)\displaystyle\mathcal{VH}_{\Pi,\psi}(G_{1})_{0}(y)=\zeta_{\mathbf{F}}(1)\gamma(1,\eta_{\mathbf{L}/\mathbf{F}},\psi)\cdot\left(f_{1}*f_{2}*f_{3}\right)(\varpi_{\mathbf{F}}^{-2}y). We also deduce that supp(𝒱Π,ψ(G1)1)ϖ𝐅1𝒪𝐅×\displaystyle\mathrm{supp}(\mathcal{VH}_{\Pi,\psi}(G_{1})_{1})\subset\varpi_{\mathbf{F}}^{-1}\mathcal{O}_{\mathbf{F}}^{\times}, and for y𝒪𝐅×\displaystyle y\in\mathcal{O}_{\mathbf{F}}^{\times}

𝒱Π,ψ(G1)1(yϖ𝐅)=q3(𝒪𝐅×)3η𝐋/𝐅(t4ϖ𝐅)ψ(t2+t3+t4+t21t31t42yϖ𝐅)dt2dt3dt4ζ𝐅(1)q3(𝒪𝐅×)4η𝐋/𝐅(t4ϖ𝐅)ψ(t1+t2+t3+t4ϖ𝐅)dt.\mathcal{VH}_{\Pi,\psi}(G_{1})_{1}\left(\tfrac{y}{\varpi_{\mathbf{F}}}\right)=q^{3}\int_{(\mathcal{O}_{\mathbf{F}}^{\times})^{3}}\eta_{\mathbf{L}/\mathbf{F}}\left(\tfrac{t_{4}}{\varpi_{\mathbf{F}}}\right)\psi\left(\tfrac{t_{2}+t_{3}+t_{4}+t_{2}^{-1}t_{3}^{-1}t_{4}^{2}y}{\varpi_{\mathbf{F}}}\right)\mathrm{d}t_{2}\mathrm{d}t_{3}\mathrm{d}t_{4}\\ -\zeta_{\mathbf{F}}(1)q^{3}\int_{(\mathcal{O}_{\mathbf{F}}^{\times})^{4}}\eta_{\mathbf{L}/\mathbf{F}}\left(\tfrac{t_{4}}{\varpi_{\mathbf{F}}}\right)\psi\left(\tfrac{t_{1}+t_{2}+t_{3}+t_{4}}{\varpi_{\mathbf{F}}}\right)\mathrm{d}\vec{t}.

The second summand is equal to ζ𝐅(1)η𝐋/𝐅(1)τ0\displaystyle\zeta_{\mathbf{F}}(1)\eta_{\mathbf{L}/\mathbf{F}}(-1)\tau_{0}. We conclude by re-numbering the variables. ∎

6.2. Further Reductions

The functions Gn\displaystyle G_{n} are “building blocks” of our test functions Hc\displaystyle H_{c} in (4.3) when 𝐋/𝐅\displaystyle\mathbf{L}/\mathbf{F} is ramified. In fact writing λ𝐋:=λ(𝐋/𝐅,ψ)η𝐋/𝐅(2)\displaystyle\lambda_{\mathbf{L}}:=\lambda(\mathbf{L}/\mathbf{F},\psi)\eta_{\mathbf{L}/\mathbf{F}}(2) and applying Lemma 3.10 we can rewrite the summands of Hc\displaystyle H_{c} as

(6.1) H2n+1=0if nn0/2,\displaystyle\displaystyle H_{2n+1}=0\quad\text{if }n\neq n_{0}/2,
(6.2) H2n={Enif n0+1nn11λ𝐋qn𝐋1Tr(α)2(1+𝒫𝐅n01)β(α)(𝔱(Tr(α)2).Gn)dαif n=n0λ𝐋qn𝐋1Tr(α)2(1+ϖ𝐅2nn01𝒪𝐅×)β(α)(𝔱(Tr(α)2).Gn)dαif n02+1nn01.\displaystyle\displaystyle H_{2n}=\begin{cases}E_{n}&\text{if }n_{0}+1\leqslant n\leqslant n_{1}-1\\ \lambda_{\mathbf{L}}q^{n}\int_{\begin{subarray}{c}\mathbf{L}^{1}\\ {\rm Tr}(\alpha)\in 2(1+\mathcal{P}_{\mathbf{F}}^{n_{0}-1})\end{subarray}}\beta(\alpha)\cdot\left(\mathfrak{t}\left({\rm Tr}(\alpha)^{2}\right).G_{n}\right)\mathrm{d}\alpha&\text{if }n=n_{0}\\ \lambda_{\mathbf{L}}q^{n}\int_{\begin{subarray}{c}\mathbf{L}^{1}\\ {\rm Tr}(\alpha)\in 2(1+\varpi_{\mathbf{F}}^{2n-n_{0}-1}\mathcal{O}_{\mathbf{F}}^{\times})\end{subarray}}\beta(\alpha)\cdot\left(\mathfrak{t}\left({\rm Tr}(\alpha)^{2}\right).G_{n}\right)\mathrm{d}\alpha&\text{if }\tfrac{n_{0}}{2}+1\leqslant n\leqslant n_{0}-1\end{cases}.

The decomposition of Hn0+1\displaystyle H_{n_{0}+1} is subtler and really goes in the direction of expression in terms of the quadratic elementary functions. We shall write (the second numeric parameter m\displaystyle m in subscript always indicates the parameter of the relevant quadratic elementary function)

(6.3) Hn0+1=λ(𝐋/𝐅,ψ)2qn0+12m=0n02Hn0+1,m,H_{n_{0}+1}=\tfrac{\lambda(\mathbf{L}/\mathbf{F},\psi)}{2}q^{\frac{n_{0}+1}{2}}\cdot\sideset{}{{}_{m=0}^{\frac{n_{0}}{2}}}{\sum}H_{n_{0}+1,m},

where the summands are given by (below m1\displaystyle m\geqslant 1)

Hn0+1,m=𝐋1Tr(ϖ𝐋α)ϖ𝐅n0/2+1m𝒪𝐅×β(ϖ𝐋α)η𝐋/𝐅(Tr(ϖ𝐋α))𝔱(ϖ𝐅1Tr(ϖ𝐋α)2).Gmdα,\displaystyle H_{n_{0}+1,m}=\int_{\begin{subarray}{c}\mathbf{L}^{1}\\ {\rm Tr}(\varpi_{\mathbf{L}}\alpha)\in\varpi_{\mathbf{F}}^{n_{0}/2+1-m}\mathcal{O}_{\mathbf{F}}^{\times}\end{subarray}}\beta(\varpi_{\mathbf{L}}\alpha)\cdot\eta_{\mathbf{L}/\mathbf{F}}\left({\rm Tr}(\varpi_{\mathbf{L}}\alpha)\right)\cdot\mathfrak{t}\left(-\varpi_{\mathbf{F}}^{-1}{\rm Tr}(\varpi_{\mathbf{L}}\alpha)^{2}\right).G_{m}\mathrm{d}\alpha,
Hn0+1,0=𝐋1Tr(ϖ𝐋α)𝒫𝐅n0/2+1β(ϖ𝐋α)dα𝔱((ϖ𝐅)n0+1).G0.\displaystyle H_{n_{0}+1,0}=\int_{\begin{subarray}{c}\mathbf{L}^{1}\\ {\rm Tr}(\varpi_{\mathbf{L}}\alpha)\in\mathcal{P}_{\mathbf{F}}^{n_{0}/2+1}\end{subarray}}\beta(\varpi_{\mathbf{L}}\alpha)\mathrm{d}\alpha\cdot\mathfrak{t}\left((-\varpi_{\mathbf{F}})^{n_{0}+1}\right).G_{0}.
Lemma 6.5.

Let χ\displaystyle\chi be a (unitary) character of 𝐅×\displaystyle\mathbf{F}^{\times} with 𝔠(χ)=nn02\displaystyle\mathfrak{c}(\chi)=n\leqslant\tfrac{n_{0}}{2}. Recall the additive parameter cβ\displaystyle c_{\beta} (resp. cχ\displaystyle c_{\chi}) in Lemma 3.8 (resp. Remark 3.9). We have

𝒪𝐅×β(1+ϖ𝐅n02ntϖ𝐋)χ(t)dtqn2.\displaystyle\int_{\mathcal{O}_{\mathbf{F}}^{\times}}\beta\left(1+\varpi_{\mathbf{F}}^{\frac{n_{0}}{2}-n}t\varpi_{\mathbf{L}}\right)\chi(t)\mathrm{d}t\ll q^{-\frac{n}{2}}.
Proof.

Since the proof is quite similar to the one of Lemma 5.7 (1), we skip some details. The case for n=0\displaystyle n=0 is easy. If n=1\displaystyle n=1, then tβ(1+ϖ𝐅n0ntε)\displaystyle t\mapsto\beta\left(1+\varpi_{\mathbf{F}}^{n_{0}-n}t\sqrt{\varepsilon}\right) is a non-trivial additive character of 𝒪𝐅\displaystyle\mathcal{O}_{\mathbf{F}} and the bound follows from the one for Gauss sums. Assume n2\displaystyle n\geqslant 2. We perform a level n2\displaystyle\lceil\tfrac{n}{2}\rceil regularization to dt\displaystyle\mathrm{d}t and get the non-vanishing condition

2cβt1ϖ𝐅n0+12nt2+cχ𝒫𝐅n22cβt+cχ(1ϖ𝐅n0+12nt2)𝒫𝐅n2,\displaystyle\tfrac{2c_{\beta}t}{1-\varpi_{\mathbf{F}}^{n_{0}+1-2-n}t^{2}}+c_{\chi}\in\mathcal{P}_{\mathbf{F}}^{\lfloor\frac{n}{2}\rfloor}\quad\Leftrightarrow\quad 2c_{\beta}t+c_{\chi}(1-\varpi_{\mathbf{F}}^{n_{0}+1-2n}t^{2})\in\mathcal{P}_{\mathbf{F}}^{\lfloor\frac{n}{2}\rfloor},

which has a unique solution tt0+𝒫𝐅n2\displaystyle t\in t_{0}+\mathcal{P}_{\mathbf{F}}^{\lfloor\frac{n}{2}\rfloor} with t0𝒪𝐅×\displaystyle t_{0}\in\mathcal{O}_{\mathbf{F}}^{\times} by Hensel’s lemma. Consequently we get

(6.4) 𝒪𝐅×β(1+ϖ𝐅n02ntϖ𝐋)χ(t)dt=t0+𝒫𝐅n2β(1+ϖ𝐅n02ntϖ𝐋)χ(t)dtqn2.\int_{\mathcal{O}_{\mathbf{F}}^{\times}}\beta\left(1+\varpi_{\mathbf{F}}^{\frac{n_{0}}{2}-n}t\varpi_{\mathbf{L}}\right)\chi(t)\mathrm{d}t=\int_{t_{0}+\mathcal{P}_{\mathbf{F}}^{\lfloor\frac{n}{2}\rfloor}}\beta\left(1+\varpi_{\mathbf{F}}^{\frac{n_{0}}{2}-n}t\varpi_{\mathbf{L}}\right)\chi(t)\mathrm{d}t\ll q^{-\lfloor\frac{n}{2}\rfloor}.

If 2n\displaystyle 2\mid n then we are done (for both (1) and (2)). Otherwise let n=2m+1\displaystyle n=2m+1. We may assume

2cβt0+cχ(1ϖ𝐅n0+12nt02ε)=0,\displaystyle 2c_{\beta}t_{0}+c_{\chi}(1-\varpi_{\mathbf{F}}^{n_{0}+1-2n}t_{0}^{2}\varepsilon)=0,

make the change of variables t=t0(1+ϖ𝐅mu)\displaystyle t=t_{0}(1+\varpi_{\mathbf{F}}^{m}u), and continue (6.4) as

𝒪𝐅×β(1+ϖ𝐅n02ntϖ𝐋)χ(t)dt=qmβ(1+ϖ𝐅n02nt0ϖ𝐋)χ(t0)𝒪𝐅ψ(cχu22ϖ𝐅)𝑑uqn2\displaystyle\int_{\mathcal{O}_{\mathbf{F}}^{\times}}\beta\left(1+\varpi_{\mathbf{F}}^{\frac{n_{0}}{2}-n}t\varpi_{\mathbf{L}}\right)\chi(t)\mathrm{d}t=q^{-m}\beta\left(1+\varpi_{\mathbf{F}}^{\frac{n_{0}}{2}-n}t_{0}\varpi_{\mathbf{L}}\right)\chi(t_{0})\int_{\mathcal{O}_{\mathbf{F}}}\psi\left(-\tfrac{c_{\chi}u^{2}}{2\varpi_{\mathbf{F}}}\right)du\ll q^{-\frac{n}{2}}

and conclude the proof. ∎

Lemma 6.6.

Suppose n02+1nn0\displaystyle\tfrac{n_{0}}{2}+1\leqslant n\leqslant n_{0} and na(Π)\displaystyle n\geqslant a(\Pi). Then we have

h~2n(χ)qn0+12𝟙n0+1n(𝔠(χη0n+1))+qn0+12𝟙n=n0𝟙𝔠(χη0)=0.\displaystyle\widetilde{h}_{2n}^{-}(\chi)\ll q^{-\frac{n_{0}+1}{2}}\mathbbm{1}_{n_{0}+1-n}(\mathfrak{c}(\chi\eta_{0}^{n+1}))+q^{-\frac{n_{0}+1}{2}}\mathbbm{1}_{n=n_{0}}\mathbbm{1}_{\mathfrak{c}(\chi\eta_{0})=0}.
Proof.

(1) First consider n<n0\displaystyle n<n_{0}. By Corollary 6.3 we have for any δ1+ϖ𝐅2nn01𝒪𝐅×\displaystyle\delta\in 1+\varpi_{\mathbf{F}}^{2n-n_{0}-1}\mathcal{O}_{\mathbf{F}}^{\times}

(6.5) 𝐅𝒪𝐅𝒱Π,ψ𝔪1(𝔱(4δ).Gn)(t)ψ(t)χ1(t)|t|12d×t=qnη0(2(1)n+1)ζ𝐅(1){χ1η0(δ1δ)γ(1,χη0,ψ)𝟙n0+1n(𝔠(χη0))if 2nχ1(δ1δ)τ0q12γ(1,χ,ψ)𝟙n0+1n(𝔠(χ))if 2n,\int_{\mathbf{F}-\mathcal{O}_{\mathbf{F}}}\mathcal{VH}_{\Pi,\psi}\circ\mathfrak{m}_{-1}\left(\mathfrak{t}\left(4\delta\right).G_{n}\right)(t)\cdot\psi(-t)\chi^{-1}(t)\lvert t\rvert^{-\frac{1}{2}}\mathrm{d}^{\times}t=\\ q^{-n}\eta_{0}(2(-1)^{n+1})\zeta_{\mathbf{F}}(1)\cdot\begin{cases}\chi^{-1}\eta_{0}\left(\tfrac{\delta}{1-\delta}\right)\cdot\gamma(1,\chi\eta_{0},\psi)\mathbbm{1}_{n_{0}+1-n}(\mathfrak{c}(\chi\eta_{0}))&\text{if }2\mid n\\ \chi^{-1}\left(\tfrac{\delta}{1-\delta}\right)\cdot\tau_{0}q^{\frac{1}{2}}\cdot\gamma(1,\chi,\psi)\mathbbm{1}_{n_{0}+1-n}(\mathfrak{c}(\chi))&\text{if }2\nmid n\end{cases},

where we used η0(δ)=1\displaystyle\eta_{0}(\delta)=1. Inserting (6.5) with δ=41Tr(α)2\displaystyle\delta=4^{-1}{\rm Tr}(\alpha)^{2} into (6.2) we get

(6.6) h~2n(χ)=𝐅𝒪𝐅𝒱Π,ψ𝔪1(H2n)(t)ψ(t)χ1(t)|t|12d×t|𝐋1Tr(α)2(1+ϖ𝐅2nn01𝒪𝐅×)β(α)χ1η0n+1(Tr(α)24Tr(α)2)dα|qn0+1n2𝟙n0+1n(𝔠(χη0n+1)).\widetilde{h}_{2n}^{-}(\chi)=\int_{\mathbf{F}-\mathcal{O}_{\mathbf{F}}}\mathcal{VH}_{\Pi,\psi}\circ\mathfrak{m}_{-1}(H_{2n})(t)\cdot\psi(-t)\chi^{-1}(t)\lvert t\rvert^{-\frac{1}{2}}\mathrm{d}^{\times}t\ll\\ \left\lvert\int_{\begin{subarray}{c}\mathbf{L}^{1}\\ {\rm Tr}(\alpha)\in 2(1+\varpi_{\mathbf{F}}^{2n-n_{0}-1}\mathcal{O}_{\mathbf{F}}^{\times})\end{subarray}}\beta(\alpha)\cdot\chi^{-1}\eta_{0}^{n+1}\left(\tfrac{{\rm Tr}(\alpha)^{2}}{4-{\rm Tr}(\alpha)^{2}}\right)\mathrm{d}\alpha\right\rvert\cdot q^{-\frac{n_{0}+1-n}{2}}\mathbbm{1}_{n_{0}+1-n}(\mathfrak{c}(\chi\eta_{0}^{n+1})).

Applying the change of variables t=αα1α+α11ϖ𝐋\displaystyle t=\tfrac{\alpha-\alpha^{-1}}{\alpha+\alpha^{-1}}\tfrac{1}{\varpi_{\mathbf{L}}} the inner integral in (6.6) becomes, taking into account the measure normalization in Proposition 3.5 (3)

q12χη0n+1(1)ϖ𝐅nn021𝒪𝐅×β(1+tϖ𝐋)χ2(t)dt.\displaystyle q^{-\frac{1}{2}}\cdot\chi\eta_{0}^{n+1}(-1)\int_{\varpi_{\mathbf{F}}^{n-\frac{n_{0}}{2}-1}\mathcal{O}_{\mathbf{F}}^{\times}}\beta(1+t\varpi_{\mathbf{L}})\chi^{2}(t)\mathrm{d}t.

We apply Lemma 6.5 (1) to bound the above integrals and conclude.

(2) Consider n=n0\displaystyle n=n_{0}. We have for any δ1+𝒫𝐅n01\displaystyle\delta\in 1+\mathcal{P}_{\mathbf{F}}^{n_{0}-1}

(6.7) 𝐅𝒪𝐅𝒱Π,ψ𝔪1(𝔱(4δ).Gn)(t)ψ(t)χ1(t)|t|12d×t=qn0η0(2)τ0q12{χ1η0(δ1δ)ζ𝐅(1){γ(1,χη0,ψ)𝟙1(𝔠(χη0))q1𝟙0(𝔠(χη0))}if δ1+ϖ𝐅n01𝒪𝐅×χη0(ϖ𝐅)n0𝟙0(𝔠(χη0))if δ1+𝒫𝐅n0.\int_{\mathbf{F}-\mathcal{O}_{\mathbf{F}}}\mathcal{VH}_{\Pi,\psi}\circ\mathfrak{m}_{-1}\left(\mathfrak{t}\left(4\delta\right).G_{n}\right)(t)\cdot\psi(-t)\chi^{-1}(t)\lvert t\rvert^{-\frac{1}{2}}\mathrm{d}^{\times}t=q^{-n_{0}}\eta_{0}(-2)\cdot\tau_{0}q^{\frac{1}{2}}\cdot\\ \begin{cases}\chi^{-1}\eta_{0}\left(\tfrac{\delta}{1-\delta}\right)\cdot\zeta_{\mathbf{F}}(1)\left\{\gamma(1,\chi\eta_{0},\psi)\mathbbm{1}_{1}(\mathfrak{c}(\chi\eta_{0}))-q^{-1}\mathbbm{1}_{0}(\mathfrak{c}(\chi\eta_{0}))\right\}&\text{if }\delta\in 1+\varpi_{\mathbf{F}}^{n_{0}-1}\mathcal{O}_{\mathbf{F}}^{\times}\\ \chi\eta_{0}(\varpi_{\mathbf{F}})^{n_{0}}\cdot\mathbbm{1}_{0}(\mathfrak{c}(\chi\eta_{0}))&\text{if }\delta\in 1+\mathcal{P}_{\mathbf{F}}^{n_{0}}\end{cases}.

Inserting (6.7) with δ=41Tr(α)2\displaystyle\delta=4^{-1}{\rm Tr}(\alpha)^{2} into (6.2) we get

(6.8) h~2n(χ)=𝐅𝒪𝐅𝒱Π,ψ𝔪1(H2n)(t)ψ(t)χ1(t)|t|12d×t|𝐋1Tr(α)2(1+ϖ𝐅n01𝒪𝐅×)β(α)χ1η0(Tr(α)24Tr(α)2)dα|q12𝟙1(𝔠(χη0))+|𝐋1Tr(α)2(1+𝒫𝐅n0)β(α)dαχη0(ϖ𝐅)1q1𝐋1Tr(α)2(1+ϖ𝐅n01𝒪𝐅×)β(α)dα|𝟙0(𝔠(χη0)).\widetilde{h}_{2n}^{-}(\chi)=\int_{\mathbf{F}-\mathcal{O}_{\mathbf{F}}}\mathcal{VH}_{\Pi,\psi}\circ\mathfrak{m}_{-1}(H_{2n})(t)\cdot\psi(-t)\chi^{-1}(t)\lvert t\rvert^{-\frac{1}{2}}\mathrm{d}^{\times}t\ll\\ \left\lvert\int_{\begin{subarray}{c}\mathbf{L}^{1}\\ {\rm Tr}(\alpha)\in 2(1+\varpi_{\mathbf{F}}^{n_{0}-1}\mathcal{O}_{\mathbf{F}}^{\times})\end{subarray}}\beta(\alpha)\cdot\chi^{-1}\eta_{0}\left(\tfrac{{\rm Tr}(\alpha)^{2}}{4-{\rm Tr}(\alpha)^{2}}\right)\mathrm{d}\alpha\right\rvert\cdot q^{-\frac{1}{2}}\mathbbm{1}_{1}(\mathfrak{c}(\chi\eta_{0}))+\\ \left\lvert\int_{\begin{subarray}{c}\mathbf{L}^{1}\\ {\rm Tr}(\alpha)\in 2(1+\mathcal{P}_{\mathbf{F}}^{n_{0}})\end{subarray}}\beta(\alpha)\mathrm{d}\alpha-\tfrac{\chi\eta_{0}(\varpi_{\mathbf{F}})^{-1}}{q-1}\int_{\begin{subarray}{c}\mathbf{L}^{1}\\ {\rm Tr}(\alpha)\in 2(1+\varpi_{\mathbf{F}}^{n_{0}-1}\mathcal{O}_{\mathbf{F}}^{\times})\end{subarray}}\beta(\alpha)\mathrm{d}\alpha\right\rvert\cdot\mathbbm{1}_{0}(\mathfrak{c}(\chi\eta_{0})).

The first summand is bounded the same way as before. With the same change of variables the inner integrals of the second summand become

q12(qn02χη0(ϖ𝐅)1q1ϖ𝐅n021𝒪𝐅×β(1+tϖ𝐋)dt).\displaystyle q^{-\frac{1}{2}}\cdot\left(q^{-\frac{n_{0}}{2}}-\tfrac{\chi\eta_{0}(\varpi_{\mathbf{F}})^{-1}}{q-1}\int_{\varpi_{\mathbf{F}}^{\frac{n_{0}}{2}-1}\mathcal{O}_{\mathbf{F}}^{\times}}\beta(1+t\varpi_{\mathbf{L}})\mathrm{d}t\right).

It is of size O(qn0+12)\displaystyle O(q^{-\frac{n_{0}+1}{2}}) since the integrand is an additive character of conductor exponent n0/2\displaystyle n_{0}/2. ∎

Lemma 6.7.

Suppose (2)a(Π)mn02\displaystyle(2\leqslant)a(\Pi)\leqslant m\leqslant\tfrac{n_{0}}{2}. Then we have for unitary χ\displaystyle\chi

h~n0+1,m(χ):=𝐅𝒪𝐅𝒱Π,ψ𝔪1(Hn0+1,m)(t)ψ(t)χ1(t)|t|12d×tqn01𝟙m>n0+13𝟙m(𝔠(χ)).\displaystyle\widetilde{h}_{n_{0}+1,m}(\chi):=\int_{\mathbf{F}-\mathcal{O}_{\mathbf{F}}}\mathcal{VH}_{\Pi,\psi}\circ\mathfrak{m}_{-1}\left(H_{n_{0}+1,m}\right)(t)\psi(-t)\chi^{-1}(t)\lvert t\rvert^{-\frac{1}{2}}\mathrm{d}^{\times}t\ll q^{-n_{0}-1}\mathbbm{1}_{m>\frac{n_{0}+1}{3}}\mathbbm{1}_{m}(\mathfrak{c}(\chi)).
Proof.

We first use Corollary 6.3 to obtain for any δϖ𝐅n0+12m𝒪𝐅×\displaystyle\delta\in\varpi_{\mathbf{F}}^{n_{0}+1-2m}\mathcal{O}_{\mathbf{F}}^{\times} (note that η0(4δ)=1\displaystyle\eta_{0}(4-\delta)=1)

(6.9) 𝐅𝒪𝐅𝒱Π,ψ𝔪1(𝔱(δ).Gm)(t)ψ(t)χ1(t)|t|12d×t=qn0+12η0(2(1)m+1)ζ𝐅(1)𝟙m>n0+13{χ(4δδ)γ(1,χη0,ψ)𝟙m(𝔠(χ))if 2mχ(4δδ)τ0q12γ(1,χ,ψ)𝟙m(𝔠(χ))if 2m.\int_{\mathbf{F}-\mathcal{O}_{\mathbf{F}}}\mathcal{VH}_{\Pi,\psi}\circ\mathfrak{m}_{-1}\left(\mathfrak{t}\left(\delta\right).G_{m}\right)(t)\cdot\psi(-t)\chi^{-1}(t)\lvert t\rvert^{-\frac{1}{2}}\mathrm{d}^{\times}t\\ =q^{-\frac{n_{0}+1}{2}}\eta_{0}(2(-1)^{m+1})\zeta_{\mathbf{F}}(1)\mathbbm{1}_{m>\frac{n_{0}+1}{3}}\cdot\begin{cases}\chi\left(\tfrac{4-\delta}{\delta}\right)\cdot\gamma(1,\chi\eta_{0},\psi)\mathbbm{1}_{m}(\mathfrak{c}(\chi))&\text{if }2\mid m\\ \chi\left(\tfrac{4-\delta}{\delta}\right)\cdot\tau_{0}q^{\frac{1}{2}}\gamma(1,\chi,\psi)\mathbbm{1}_{m}(\mathfrak{c}(\chi))&\text{if }2\nmid m\end{cases}.

Inserting (6.9) with δ=ϖ𝐅1Tr(ϖ𝐋α)2\displaystyle\delta=-\varpi_{\mathbf{F}}^{-1}{\rm Tr}(\varpi_{\mathbf{L}}\alpha)^{2} into the integral representation of Hn0+1,m\displaystyle H_{n_{0}+1,m} we get

(6.10) h~n0+1,m(χ)𝟙m>n0+13qn0+1+m2𝟙m(𝔠(χ))|𝐋1Tr(ϖ𝐋α)ϖ𝐅n02+1m𝒪𝐅×β(ϖ𝐋α)η𝐋/𝐅(Tr(ϖ𝐋α))χ(4ϖ𝐅+Tr(ϖ𝐋α)2Tr(ϖ𝐋α)2)dα|.\widetilde{h}_{n_{0}+1,m}(\chi)\ll\mathbbm{1}_{m>\frac{n_{0}+1}{3}}\cdot q^{-\frac{n_{0}+1+m}{2}}\mathbbm{1}_{m}(\mathfrak{c}(\chi))\cdot\\ \left\lvert\int_{\begin{subarray}{c}\mathbf{L}^{1}\\ {\rm Tr}(\varpi_{\mathbf{L}}\alpha)\in\varpi_{\mathbf{F}}^{\frac{n_{0}}{2}+1-m}\mathcal{O}_{\mathbf{F}}^{\times}\end{subarray}}\beta(\varpi_{\mathbf{L}}\alpha)\eta_{\mathbf{L}/\mathbf{F}}\left({\rm Tr}(\varpi_{\mathbf{L}}\alpha)\right)\cdot\chi\left(\tfrac{4\varpi_{\mathbf{F}}+{\rm Tr}(\varpi_{\mathbf{L}}\alpha)^{2}}{-{\rm Tr}(\varpi_{\mathbf{L}}\alpha)^{2}}\right)\mathrm{d}\alpha\right\rvert.

Applying the change of variables t=ϖ𝐋α+ϖ𝐋α¯ϖ𝐋αϖ𝐋α¯1ϖ𝐋\displaystyle t=\tfrac{\varpi_{\mathbf{L}}\alpha+\overline{\varpi_{\mathbf{L}}\alpha}}{\varpi_{\mathbf{L}}\alpha-\overline{\varpi_{\mathbf{L}}\alpha}}\tfrac{1}{\varpi_{\mathbf{L}}} the inner integral in (6.10) becomes, taking into account the measure normalization in Proposition 3.5 (3)

q122β(ϖ𝐋)η𝐋/𝐅(2)χ1(ϖ𝐅)ϖ𝐅n02m𝒪𝐅×β(1+tϖ𝐋)η𝐋/𝐅χ2(t)dt.\displaystyle q^{-\frac{1}{2}}\cdot 2\beta(\varpi_{\mathbf{L}})\eta_{\mathbf{L}/\mathbf{F}}(-2)\chi^{-1}(-\varpi_{\mathbf{F}})\int_{\varpi_{\mathbf{F}}^{\frac{n_{0}}{2}-m}\mathcal{O}_{\mathbf{F}}^{\times}}\beta(1+t\varpi_{\mathbf{L}})\eta_{\mathbf{L}/\mathbf{F}}\chi^{-2}(t)\mathrm{d}t.

We apply Lemma 6.5 (1) to bound the above integrals and conclude the desired inequalities. ∎

Lemma 6.8.

Let n0max(4𝔠(Π),2𝔠(Π)+2)\displaystyle n_{0}\geqslant\max(4\mathfrak{c}(\Pi),2\mathfrak{c}(\Pi)+2) and m𝔠(Π)\displaystyle m\leqslant\mathfrak{c}(\Pi). Then for any δϖ𝐅n0+12m𝒪𝐅×\displaystyle\delta\in\varpi_{\mathbf{F}}^{n_{0}+1-2m}\mathcal{O}_{\mathbf{F}}^{\times} we have

𝐅𝒪𝐅𝒱Π,ψ𝔪1(𝔱(δ).Gm)(t)ψ(t)χ1(t)|t|12d×t=0.\displaystyle\int_{\mathbf{F}-\mathcal{O}_{\mathbf{F}}}\mathcal{VH}_{\Pi,\psi}\circ\mathfrak{m}_{-1}\left(\mathfrak{t}(\delta).G_{m}\right)(t)\cdot\psi(-t)\chi^{-1}(t)\lvert t\rvert^{-\frac{1}{2}}\mathrm{d}^{\times}t=0.

Consequently, we get for any unitary χ\displaystyle\chi the vanishing of h~n0+1,m(χ)=0\displaystyle\widetilde{h}_{n_{0}+1,m}(\chi)=0.

Proof.

The proof is quite similar to Lemma 5.11. One shows that the support of 𝒱Π,ψ𝔪1(𝔱(δ).Gm)\displaystyle\mathcal{VH}_{\Pi,\psi}\circ\mathfrak{m}_{-1}\left(\mathfrak{t}(\delta).G_{m}\right) is contained in δ𝒫𝐅2mmax(4𝔠(Π),2𝔠(Π)+3)𝒪𝐅\displaystyle\delta\mathcal{P}_{\mathbf{F}}^{2m-\max(4\mathfrak{c}(\Pi),2\mathfrak{c}(\Pi)+3)}\subset\mathcal{O}_{\mathbf{F}} under the assumption. ∎

6.3. The Bounds of Dual Weight

Lemma 6.9.

Suppose 𝔠(Π)=0\displaystyle\mathfrak{c}(\Pi)=0 and n0=2\displaystyle n_{0}=2. We have h~3(χ)=0\displaystyle\widetilde{h}_{3}^{-}(\chi)=0 for any unitary character χ\displaystyle\chi.

Proof.

Necessarily we have Π=μ1μ2μ3\displaystyle\Pi=\mu_{1}\boxplus\mu_{2}\boxplus\mu_{3} for unramified μj\displaystyle\mu_{j}. Note that H3,0\displaystyle H_{3,0} (resp. H3,1\displaystyle H_{3,1}) is related to G0\displaystyle G_{0} (resp. G1\displaystyle G_{1}) by the formulae

H3,0=𝐋1Tr(ϖ𝐋α)𝒫𝐅2β(ϖ𝐋α)dα𝔱((ϖ𝐅)3).G0,\displaystyle H_{3,0}=\int_{\begin{subarray}{c}\mathbf{L}^{1}\\ {\rm Tr}(\varpi_{\mathbf{L}}\alpha)\in\mathcal{P}_{\mathbf{F}}^{2}\end{subarray}}\beta(\varpi_{\mathbf{L}}\alpha)\mathrm{d}\alpha\cdot\mathfrak{t}\left((-\varpi_{\mathbf{F}})^{3}\right).G_{0},
H3,1=𝐋1Tr(ϖ𝐋α)ϖ𝐅𝒪𝐅×β(ϖ𝐋α)η𝐋/𝐅(Tr(ϖ𝐋α))𝔱(ϖ𝐅1Tr(ϖ𝐋α)2).G1dα.\displaystyle H_{3,1}=\int_{\begin{subarray}{c}\mathbf{L}^{1}\\ {\rm Tr}(\varpi_{\mathbf{L}}\alpha)\in\varpi_{\mathbf{F}}\mathcal{O}_{\mathbf{F}}^{\times}\end{subarray}}\beta(\varpi_{\mathbf{L}}\alpha)\cdot\eta_{\mathbf{L}/\mathbf{F}}\left({\rm Tr}(\varpi_{\mathbf{L}}\alpha)\right)\cdot\mathfrak{t}\left(-\varpi_{\mathbf{F}}^{-1}{\rm Tr}(\varpi_{\mathbf{L}}\alpha)^{2}\right).G_{1}\mathrm{d}\alpha.

Inspecting the supports of G0\displaystyle G_{0} and G1\displaystyle G_{1} given in Corollary 6.4 we see supp(H3,m)𝒪𝐅\displaystyle\mathrm{supp}(H_{3,m})\subset\mathcal{O}_{\mathbf{F}} for m{0,1}\displaystyle m\in\{0,1\}. ∎

Proposition 6.10.

With the test function H\displaystyle H given by (3.5) the dual weight function is bounded as

h~(χ)ϵ𝐂(Π)2+ϵqn0+12+ϵ𝟙max(n02,6𝔠(Π))(𝔠(χ)).\displaystyle\widetilde{h}(\chi)\ll_{\epsilon}\mathbf{C}(\Pi)^{2+\epsilon}q^{-\frac{n_{0}+1}{2}+\epsilon}\mathbbm{1}_{\leqslant\max(\frac{n_{0}}{2},6\mathfrak{c}(\Pi))}(\mathfrak{c}(\chi)).
Proof.

The argument is quite similar to and simpler than Proposition 5.13. In the case 𝔠(Π)>0\displaystyle\mathfrak{c}(\Pi)>0 we distinguish n04𝔠(Π)\displaystyle n_{0}\leqslant 4\mathfrak{c}(\Pi), resp. n04𝔠(Π)\displaystyle n_{0}\geqslant 4\mathfrak{c}(\Pi). We apply Lemma 4.1, 4.6 and 4.8 (with A=4\displaystyle A=4), resp. Lemma 4.1, 4.6 and Lemma 6.6-6.8. In the case 𝔠(Π)=0\displaystyle\mathfrak{c}(\Pi)=0, we apply Lemma 4.1, 4.6 and 6.9. We leave the details to the reader. ∎

6.4. The Bounds of Unramified Dual Weight

Lemma 6.11.

Suppose n02+1nn0\displaystyle\tfrac{n_{0}}{2}+1\leqslant n\leqslant n_{0} and na(Π)\displaystyle n\geqslant a(\Pi). Then the function h~2n(||𝐅s)0\displaystyle\widetilde{h}_{2n}^{-}(\lvert\cdot\rvert_{\mathbf{F}}^{s})\neq 0 is non-vanishing only if n=n0\displaystyle n=n_{0}. Moreover for any k0\displaystyle k\in\mathbb{Z}_{\geqslant 0} we have

H~2n0(k;12)k,ϵq12+n0ϵ,H~2n0(k;12)k,ϵqn012+n0ϵ.\displaystyle\widetilde{H}_{2n_{0}}^{-}(k;\tfrac{1}{2})\ll_{k,\epsilon}q^{-\frac{1}{2}+n_{0}\epsilon},\quad\widetilde{H}_{2n_{0}}^{-}(k;-\tfrac{1}{2})\ll_{k,\epsilon}q^{-n_{0}-\frac{1}{2}+n_{0}\epsilon}.
Proof.

By Corollary 6.3 and (6.2) the support of 𝒱Π,ψ(Gn)\displaystyle\mathcal{VH}_{\Pi,\psi}(G_{n}), hence also 𝒱Π,ψ𝔪1(H2n)\displaystyle\mathcal{VH}_{\Pi,\psi}\circ\mathfrak{m}_{-1}(H_{2n}) are contained in ϖ𝐅n𝒪𝐅×\displaystyle\varpi_{\mathbf{F}}^{-n}\mathcal{O}_{\mathbf{F}}^{\times}. Therefore we get by (6.6) the relation h~2n(||𝐅s)=qnsh~2n(𝟙)\displaystyle\widetilde{h}_{2n}^{-}(\lvert\cdot\rvert_{\mathbf{F}}^{s})=q^{ns}\widetilde{h}_{2n}^{-}(\mathbbm{1}). The stated results follow readily from Lemma 6.6. ∎

Proposition 6.12.

With the test function H\displaystyle H given by (3.5) we have for any k0\displaystyle k\in\mathbb{Z}_{\geqslant 0}

H~(k;12)k,ϵ𝐂(Π)12q(n0+1)ϵ,H~(k;12)k,ϵ𝐂(Π)4+ϵq(n0+1)ϵ.\displaystyle\widetilde{H}(k;\tfrac{1}{2})\ll_{k,\epsilon}\mathbf{C}(\Pi)^{\frac{1}{2}}\cdot q^{(n_{0}+1)\epsilon},\quad\widetilde{H}(k;-\tfrac{1}{2})\ll_{k,\epsilon}\mathbf{C}(\Pi)^{4+\epsilon}\cdot q^{(n_{0}+1)\epsilon}.
Proof.

The proof is quite similar to and simpler than the one of Proposition 5.16. We simply note that we can take n1=n0+1\displaystyle n_{1}=n_{0}+1, and leave the details to the reader. ∎

Appendix A Relation with Petrow–Young’s Exponential Sums

For simplicity of notation we shall write t\displaystyle\sideset{}{{}_{t}}{\sum} for t𝔽q\displaystyle\sideset{}{{}_{t\in\mathbb{F}_{q}}}{\sum}, and use the convention ρ(0)=0\displaystyle\rho(0)=0 for any character ρ\displaystyle\rho of 𝔽q×\displaystyle\mathbb{F}_{q}^{\times} (even if ρ=𝟙\displaystyle\rho=\mathbbm{1} is the trivial one). In (5.19) we have encountered the following algebraic exponential sum

(A.1) S=S(χ0,χ):=α𝔽q{±1}χ0(α+1α1)H(1α2,q;(𝟙,𝟙),(χ1,η)),S=S(\chi_{0},\chi):=\sideset{}{{}_{\alpha\in\mathbb{F}_{q}-\{\pm 1\}}}{\sum}\chi_{0}\left(\tfrac{\alpha+1}{\alpha-1}\right)H(1-\alpha^{2},q;(\mathbbm{1},\mathbbm{1}),(\chi^{-1},\eta)),

where χ0,χ\displaystyle\chi_{0},\chi are non-trivial characters of 𝔽q×\displaystyle\mathbb{F}_{q}^{\times} and η\displaystyle\eta is the unique non-trivial quadratic character of 𝔽q×\displaystyle\mathbb{F}_{q}^{\times}. Note that the setting of the relevant dual weight specializes to the Petrow–Young’s [14] upon taking Π=𝟙𝟙𝟙\displaystyle\Pi=\mathbbm{1}\boxplus\mathbbm{1}\boxplus\mathbbm{1}. It is a natural question to relate S\displaystyle S with their algebraic exponential sum

(A.2) T=T(χ0,χ):=u,vχ(u(u+1)v(v+1))χ0(uv1).T=T(\chi_{0},\chi):=\sideset{}{{}_{u,v}}{\sum}\chi\left(\tfrac{u(u+1)}{v(v+1)}\right)\chi_{0}(uv-1).

The purpose of this appendix is to give such an explicit relation. Our approach will take into account the recent discovery of Xi [22], which relates T\displaystyle T to a special value of a hypergeometric sum of Katz.

We need to rewrite a hyper-Kloosterman sum via the duplication formula of Gauss sums.

Definition A.1.

For a character ρ\displaystyle\rho of 𝔽q×\displaystyle\mathbb{F}_{q}^{\times} and a non-trivial character ψ\displaystyle\psi of 𝔽q\displaystyle\mathbb{F}_{q} we have the Gauss sum

τ(ρ)=τ(ρ,ψ):=tρ(t)ψ(t).\displaystyle\tau(\rho)=\tau(\rho,\psi):=\sideset{}{{}_{t}}{\sum}\rho(t)\psi(t).
Lemma A.2.

We have the relation τ(ρ2)τ(η)=ρ(4)τ(ρ)τ(ρη)\displaystyle\tau(\rho^{2})\tau(\eta)=\rho(4)\tau(\rho)\tau(\rho\eta) for any character ρ\displaystyle\rho of 𝔽q×\displaystyle\mathbb{F}_{q}^{\times}.

Proof.

If ρ=𝟙\displaystyle\rho=\mathbbm{1} or η\displaystyle\eta, the stated relation trivially holds true. Assume ρ𝟙,η\displaystyle\rho\neq\mathbbm{1},\eta. We have the classical relation of Gauss and Jacobi sums [9, §8.3 Theorem 1]

τ(ρ)2τ(ρ2)=J(ρ,ρ)=tρ(t(1t)),τ(ρ)τ(η)τ(ρη)=J(ρ,η)=tη(t)ρ(1t).\displaystyle\tfrac{\tau(\rho)^{2}}{\tau(\rho^{2})}=J(\rho,\rho)=\sideset{}{{}_{t}}{\sum}\rho(t(1-t)),\quad\tfrac{\tau(\rho)\tau(\eta)}{\tau(\rho\eta)}=J(\rho,\eta)=\sideset{}{{}_{t}}{\sum}\eta(t)\rho(1-t).

Since q\displaystyle q is odd, we can rewrite by completing the square

J(ρ,ρ)=tρ(t(1t))=ρ(4)1tρ(1t2)=ρ(4)1t(1+η(t))ρ(1t)=ρ(4)1tη(t)ρ(1t)=ρ(4)1J(ρ,η).J(\rho,\rho)=\sideset{}{{}_{t}}{\sum}\rho(t(1-t))=\rho(4)^{-1}\sideset{}{{}_{t}}{\sum}\rho(1-t^{2})\\ =\rho(4)^{-1}\sideset{}{{}_{t}}{\sum}(1+\eta(t))\rho(1-t)=\rho(4)^{-1}\sideset{}{{}_{t}}{\sum}\eta(t)\rho(1-t)=\rho(4)^{-1}J(\rho,\eta).

The stated relation follows readily since τ(ρ)0\displaystyle\tau(\rho)\neq 0. ∎

Corollary A.3.

For any δ𝔽q×\displaystyle\delta\in\mathbb{F}_{q}^{\times} we have the equality

u,t0η(1u)ψ(δt2u+2t)=x1x2x3=δψ(x1+x2+x3).\displaystyle\sideset{}{{}_{u,t\neq 0}}{\sum}\eta\left(1-u\right)\psi\left(\tfrac{\delta}{t^{2}u}+2t\right)=\sideset{}{{}_{x_{1}x_{2}x_{3}=\delta}}{\sum}\psi(x_{1}+x_{2}+x_{3}).
Proof.

The right hand side is Kl3(δ)\displaystyle\mathrm{Kl}_{3}(\delta), a hyper-Kloosterman sum, whose Mellin transform satisfies

δ𝔽q×Kl3(δ)ρ(δ)=τ(ρ)3,ρ𝔽q×^.\displaystyle\sideset{}{{}_{\delta\in\mathbb{F}_{q}^{\times}}}{\sum}\mathrm{Kl}_{3}(\delta)\rho(\delta)=\tau(\rho)^{3},\quad\forall\rho\in\widehat{\mathbb{F}_{q}^{\times}}.

It suffices to identify the Mellin transform of the left hand side as τ(ρ)3\displaystyle\tau(\rho)^{3}. We have

δu,t0η(1u)ψ(δt2u+2t)ρ(δ)=τ(ρ)u,tη(1u)ψ(2t)ρ(t2u)=ρ(4)1τ(ρ)τ(ρ2)uη(1u)ρ(u)=τ(ρ)2τ(ρ2)τ(η)ρ(4)τ(ρη).\sideset{}{{}_{\delta}}{\sum}\sideset{}{{}_{u,t\neq 0}}{\sum}\eta\left(1-u\right)\psi\left(\tfrac{\delta}{t^{2}u}+2t\right)\rho(\delta)=\tau(\rho)\sideset{}{{}_{u,t}}{\sum}\eta\left(1-u\right)\psi\left(2t\right)\rho(t^{2}u)\\ =\rho(4)^{-1}\tau(\rho)\tau(\rho^{2})\sideset{}{{}_{u}}{\sum}\eta\left(1-u\right)\rho(u)=\tfrac{\tau(\rho)^{2}\tau(\rho^{2})\tau(\eta)}{\rho(4)\tau(\rho\eta)}.

Lemma A.2 identifies the above right hand side precisely as τ(ρ)3\displaystyle\tau(\rho)^{3}. ∎

For S=S(χ0,χ)\displaystyle S=S(\chi_{0},\chi), we make the change of variables u=1+α\displaystyle u=1+\alpha and v=1α\displaystyle v=1-\alpha, and detect the condition u+v=2\displaystyle u+v=2 with the additive character ψ\displaystyle\psi to get

S\displaystyle\displaystyle S =q32χ0(1)u+v=2χ0(u)χ0¯(v)xi,yix1x2=uvy1y2χ(y1)η(y2)ψ(x1+x2y1y2)\displaystyle\displaystyle=-q^{-\frac{3}{2}}\chi_{0}(-1)\sideset{}{{}_{u+v=2}}{\sum}\chi_{0}(u)\overline{\chi_{0}}(v)\sideset{}{{}_{\begin{subarray}{c}x_{i},y_{i}\\ x_{1}x_{2}=uvy_{1}y_{2}\end{subarray}}}{\sum}\chi(y_{1})\eta(y_{2})\psi(x_{1}+x_{2}-y_{1}-y_{2})
=q52χ0(1)txi,yi,u,vx1x2=uvy1y2χ0(u)χ0¯(v)χ(y1)η(y2)ψ(x1+x2y1y2+tu+tv2t)\displaystyle\displaystyle=-q^{-\frac{5}{2}}\chi_{0}(-1)\sideset{}{{}_{t}}{\sum}\sideset{}{{}_{\begin{subarray}{c}x_{i},y_{i},u,v\\ x_{1}x_{2}=uvy_{1}y_{2}\end{subarray}}}{\sum}\chi_{0}(u)\overline{\chi_{0}}(v)\chi(y_{1})\eta(y_{2})\psi(x_{1}+x_{2}-y_{1}-y_{2}+tu+tv-2t)
=q52(S1+S2).\displaystyle\displaystyle=-q^{-\frac{5}{2}}(S_{1}+S_{2}).

In the above the sum S1\displaystyle S_{1} is defined and computed as

S1\displaystyle\displaystyle S_{1} :=χ0(1)t𝔽q×xi,yi,u,vx1x2=uvy1y2χ0(u)χ0¯(v)χ(y1)η(y2)ψ(x1+x2y1y2tutv+2t)\displaystyle\displaystyle:=\chi_{0}(-1)\sideset{}{{}_{t\in\mathbb{F}_{q}^{\times}}}{\sum}\sideset{}{{}_{\begin{subarray}{c}x_{i},y_{i},u,v\\ x_{1}x_{2}=uvy_{1}y_{2}\end{subarray}}}{\sum}\chi_{0}(u)\overline{\chi_{0}}(v)\chi(y_{1})\eta(y_{2})\psi(x_{1}+x_{2}-y_{1}-y_{2}-tu-tv+2t)
=χ0(1)t𝔽q×xi,yi,u,vt2x1x2=uvy1y2χ0(u)χ0¯(v)χ(y1)η(y2)ψ(x1+x2y1y2uv+2t)\displaystyle\displaystyle=\chi_{0}(-1)\sideset{}{{}_{t\in\mathbb{F}_{q}^{\times}}}{\sum}\sideset{}{{}_{\begin{subarray}{c}x_{i},y_{i},u,v\\ t^{2}x_{1}x_{2}=uvy_{1}y_{2}\end{subarray}}}{\sum}\chi_{0}(u)\overline{\chi_{0}}(v)\chi(y_{1})\eta(y_{2})\psi(x_{1}+x_{2}-y_{1}-y_{2}-u-v+2t)
=χ0(1)t𝔽q×xi,yi,u,vt2x1x2=uvy1η(y2)ψ(y2(x21))χ0(u)χ0¯(v)χ(y1)ψ(x1+2tuvy1)\displaystyle\displaystyle=\chi_{0}(-1)\sideset{}{{}_{t\in\mathbb{F}_{q}^{\times}}}{\sum}\sideset{}{{}_{\begin{subarray}{c}x_{i},y_{i},u,v\\ t^{2}x_{1}x_{2}=uvy_{1}\end{subarray}}}{\sum}\eta(y_{2})\psi(y_{2}(x_{2}-1))\cdot\chi_{0}(u)\overline{\chi_{0}}(v)\chi(y_{1})\psi(x_{1}+2t-u-v-y_{1})
=χ0(1)τ(η)t𝔽q×xi,yit2x1x2=y1y2y3η(x21)χ(y1)χ0(y2)χ0¯(y3)ψ(x1+2ty1y2y3)\displaystyle\displaystyle=\chi_{0}(-1)\tau(\eta)\sideset{}{{}_{t\in\mathbb{F}_{q}^{\times}}}{\sum}\sideset{}{{}_{\begin{subarray}{c}x_{i},y_{i}\\ t^{2}x_{1}x_{2}=y_{1}y_{2}y_{3}\end{subarray}}}{\sum}\eta(x_{2}-1)\chi(y_{1})\chi_{0}(y_{2})\overline{\chi_{0}}(y_{3})\psi(x_{1}+2t-y_{1}-y_{2}-y_{3})
=χ0η(1)τ(η)yiχ(y1)χ0(y2)χ0¯(y3)ψ(y1y2y3)x2,t0η(1x2)ψ(y1y2y3t2x2+2t).\displaystyle\displaystyle=\chi_{0}\eta(-1)\tau(\eta)\sideset{}{{}_{y_{i}}}{\sum}\chi(y_{1})\chi_{0}(y_{2})\overline{\chi_{0}}(y_{3})\psi(-y_{1}-y_{2}-y_{3})\sideset{}{{}_{x_{2},t\neq 0}}{\sum}\eta(1-x_{2})\psi\left(\tfrac{y_{1}y_{2}y_{3}}{t^{2}x_{2}}+2t\right).

Re-naming the variable u=x2\displaystyle u=x_{2} we identify the inner sum as Kl3(y1y2y3)\displaystyle\mathrm{Kl}_{3}(y_{1}y_{2}y_{3}) by Corollary A.3 so that

S1\displaystyle\displaystyle S_{1} =χ0η(1)τ(η)xi,yix1x2x3=y1y2y3χ(y1)χ0(y2)χ0¯(y3)ψ(x1+x2+x3y1y2y3)\displaystyle\displaystyle=\chi_{0}\eta(-1)\tau(\eta)\sideset{}{{}_{\begin{subarray}{c}x_{i},y_{i}\\ x_{1}x_{2}x_{3}=y_{1}y_{2}y_{3}\end{subarray}}}{\sum}\chi(y_{1})\chi_{0}(y_{2})\overline{\chi_{0}}(y_{3})\psi(x_{1}+x_{2}+x_{3}-y_{1}-y_{2}-y_{3})
=q52χ0η(1)τ(η)H(1,q;(𝟙,𝟙,𝟙),(χ¯,χ0,χ0¯))=q2T(χ0,χ),\displaystyle\displaystyle=-q^{\frac{5}{2}}\chi_{0}\eta(-1)\tau(\eta)H(1,q;(\mathbbm{1},\mathbbm{1},\mathbbm{1}),(\overline{\chi},\chi_{0},\overline{\chi_{0}}))=-q^{2}\cdot T(\chi_{0},\chi),

where we have applied [22, Theorem 1.1] to get the last equality. The sum S2\displaystyle S_{2} is defined as

S2:=χ0(1)xi,yi,u,vx1x2=uvy1y2χ0(u)χ0¯(v)χ(y1)η(y2)ψ(x1+x2y1y2).\displaystyle S_{2}:=\chi_{0}(-1)\sideset{}{{}_{\begin{subarray}{c}x_{i},y_{i},u,v\\ x_{1}x_{2}=uvy_{1}y_{2}\end{subarray}}}{\sum}\chi_{0}(u)\overline{\chi_{0}}(v)\chi(y_{1})\eta(y_{2})\psi(x_{1}+x_{2}-y_{1}-y_{2}).

Let t=uv\displaystyle t=uv, so u=tv1\displaystyle u=tv^{-1}, and sum first over v𝔽q×\displaystyle v\in\mathbb{F}_{q}^{\times}. We see that S20\displaystyle S_{2}\neq 0 only if χ0=η\displaystyle\chi_{0}=\eta, and

S2\displaystyle\displaystyle S_{2} =δχ0=ηη(1)(q1)t,xi,yix1x2=ty1y2χ(y1)η(ty2)ψ(x1+x2y1y2)\displaystyle\displaystyle=\delta_{\chi_{0}=\eta}\cdot\eta(-1)(q-1)\sideset{}{{}_{\begin{subarray}{c}t,x_{i},y_{i}\\ x_{1}x_{2}=ty_{1}y_{2}\end{subarray}}}{\sum}\chi(y_{1})\eta(ty_{2})\psi(x_{1}+x_{2}-y_{1}-y_{2})
=δχ0=ηη(1)(q1)xi,yix1x2=y1y2χ(y1)η(y2)ψ(x1+x2y1)t0ψ(t1y2)\displaystyle\displaystyle=\delta_{\chi_{0}=\eta}\cdot\eta(-1)(q-1)\sideset{}{{}_{\begin{subarray}{c}x_{i},y_{i}\\ x_{1}x_{2}=y_{1}y_{2}\end{subarray}}}{\sum}\chi(y_{1})\eta(y_{2})\psi(x_{1}+x_{2}-y_{1})\sideset{}{{}_{t\neq 0}}{\sum}\psi(-t^{-1}y_{2})
=δχ0=ηη(1)(q1)xi,yix1x2=y1y2χ(y1)η(y2)ψ(x1+x2y1)\displaystyle\displaystyle=-\delta_{\chi_{0}=\eta}\cdot\eta(-1)(q-1)\sideset{}{{}_{\begin{subarray}{c}x_{i},y_{i}\\ x_{1}x_{2}=y_{1}y_{2}\end{subarray}}}{\sum}\chi(y_{1})\eta(y_{2})\psi(x_{1}+x_{2}-y_{1})
=δχ0=ηη(1)(q1)x1,x2,y1χ(y1)η(x1x2y1)ψ(x1+x2y1)\displaystyle\displaystyle=-\delta_{\chi_{0}=\eta}\cdot\eta(-1)(q-1)\sideset{}{{}_{x_{1},x_{2},y_{1}}}{\sum}\chi(y_{1})\eta\left(\tfrac{x_{1}x_{2}}{y_{1}}\right)\psi(x_{1}+x_{2}-y_{1})
=δχ0=ηη(1)(q1)τ(η)2τ(ηχ1)¯=δχ0=ηq(q1)τ(ηχ1)¯.\displaystyle\displaystyle=-\delta_{\chi_{0}=\eta}\cdot\eta(-1)(q-1)\tau(\eta)^{2}\overline{\tau(\eta\chi^{-1})}=-\delta_{\chi_{0}=\eta}\cdot q(q-1)\overline{\tau(\eta\chi^{-1})}.

We summarize the above computation as the following result.

Proposition A.4.

The two algebraic exponential sums S(χ0,χ)\displaystyle S(\chi_{0},\chi) in (A.1) and T(χ0,χ)\displaystyle T(\chi_{0},\chi) in (A.2) satisfy

S(χ0,χ)=q12T(χ0,χ)+δχ0=ηq12(1q1)τ(ηχ1)¯.\displaystyle S(\chi_{0},\chi)=q^{-\frac{1}{2}}\cdot T(\chi_{0},\chi)+\delta_{\chi_{0}=\eta}\cdot q^{-\frac{1}{2}}(1-q^{-1})\overline{\tau(\eta\chi^{-1})}.

References

  • [1] Balkanova, O., Frolenkov, D., and Wu, H. On Weyl’s subconvex bound for cube-free Hecke characters: totally real case. arXiv: 2108.12283, 2021.
  • [2] Baruch, E. M., and Mao, Z. Bessel identities in the Waldspurger correspondence over the real numbers. Israel Journal of Mathematics 145 (2005), 1–81.
  • [3] Baruch, E. M., and Snitz, K. A note on Bessel functions for supercuspidal representations of GL(2)\displaystyle\mathrm{GL}(2) over a p\displaystyle p-adic field. Algebra Colloquium 18, Spec 1 (2011), 733–738.
  • [4] Bushnell, C. J., and Henniart, G. An upper bound on conductors for pairs. Journal of Number Theory 65 (1997), 183–196.
  • [5] Bushnell, C. J., and Henniart, G. The Local Langlands Conjecture for GL(2). No. 335 in Grundlehren der mathematischen Wissenschaften. Springer-Verlag, 2006.
  • [6] Bushnell, C. J., Henniart, G., and Kutzko, P. Local Rankin-Selberg convolutions for GLn\displaystyle\mathrm{GL}_{n}: explicit conductor formula. Journal of the American Mathematical Society 11, 3 (1998), 703–730.
  • [7] Chai, J., and Cong, X. A note on Weil index. Science in China Series A: Mathematics 50, 7 (July 2007), 951–956.
  • [8] Hu, Y., Petrow, I., and Young, M. P. The cubic moment of L\displaystyle L-functions for specified local component families. arXiv: 2506.14741, 2025.
  • [9] Ireland, K., and Rosen, M. A Classical Introduction to Modern Number Theory, 2nd ed., vol. 84 of Graduate Texts in Mathematics. Springer-Verlag, 1990.
  • [10] Jacquet, H., and Langlands, R. P. Automorphic Forms on GL(2). No. 114 in Lecture Notes in Mathematics. Springer-Verlag, 1970.
  • [11] Jacquet, H., and Shalika, J. A lemma on highly ramified ε\displaystyle\varepsilon-factors. Mathematische Annalen 271 (1985), 319–332.
  • [12] Moy, A., and Prasad, G. Unrefined minimal 𝕂\displaystyle\mathbb{K}-types for p\displaystyle p-adic groups. Inventiones mathematicae 116 (1994), 393–408.
  • [13] Nelson, P. Eisenstein series and the cubic moment for PGL2\displaystyle\mathrm{PGL}_{2}. arXiv: 1911.06310, 2020.
  • [14] Petrow, I., and Young, M. P. The Weyl bound for Dirichlet L\displaystyle L-functions of cube-free conductor. Annals of Mathematics 192, 2 (2020), 437–486.
  • [15] Rohrlich, D. Elliptic curves and the Weil-Deligne group. In Elliptic curves and related topics (1994), vol. 4 of CRM Proc. Lect. Notes, pp. 125–157.
  • [16] Serre, J.-P. Local Fields (translated from the French by Marvin Jay Greenberg). No. 67 in Graduate Texts in Mathematics. Springer-Verlag, 1979.
  • [17] Soudry, D. The L\displaystyle L and γ\displaystyle\gamma factors for generic representations of GSp(4,k)×GL(2,k)\displaystyle GSp(4,k)\times GL(2,k) over a local nonarchimedean field k\displaystyle k. Duke Mathematical Journal 51, 2 (June 1984), 355–394.
  • [18] Weil, A. Basic Number Theory, 3rd ed. Springer-Verlag, 1974.
  • [19] Wu, H. On a generalization of Motohashi’s formula. arXiv:2310.08236.
  • [20] Wu, H. Burgess-like subconvex bounds for GL2×GL1\displaystyle GL_{2}\times GL_{1}. Geometric and Functional Analysis 24, 3 (2014), 968–1036.
  • [21] Wu, H., and Xi, P. A uniform Weyl bound for L\displaystyle L-functions of Hilbert modular forms. arXiv:2302.14652, 2023.
  • [22] Xi, P. A double character sum of Conrey–Iwaniec and Petrow–Young. arXiv:2302.14681, 2023.