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

The subconvexity problem for Rankin–Selberg and triple product L-functions

Yueke Hu YMSC, Tsinghua University [email protected] Philippe Michel EPFL [email protected]  and  Paul Nelson AArhus University [email protected]
Abstract.

In this paper we study the subconvexity problem for the Rankin-Selberg L-function and triple product L-function, allowing joint ramifications and conductor dropping range. We first extend the method of Michel-Venkatesh to reduce the bounds for L-functions to local conjectures on test vectors, then verify these local conjectures under certain conditions, giving new subconvexity bounds as long as the representations are not completely related.

1. Introduction

Let FF be a number field with ring of adeles 𝔸{\mathbb{A}}, and let Π\Pi be an automorphic representation of a reductive group GG. Let L(Π,s)L(\Pi,s) be the L-function associated to Π\Pi. The subconvexity problem for the central value L(Π,1/2)L(\Pi,1/2) is to obtain a (non-trivial) bound of the shape

L(Π,1/2)FC(Π)1/4δL(\Pi,1/2)\ll_{F}C(\Pi)^{1/4-\delta}

where δ>0\delta>0 is some positive absolute constant (independent of Π\Pi) and C(Π)C(\Pi) is the analytic conductor of Π\Pi which is a product over all places of the local analytic conductors Cv(Π)C_{v}(\Pi).

In the case G=GL1G={\text{GL}}_{1}, GL2{\text{GL}}_{2}, the subconvexity problem has now been solved completely through the efforts of many people over the years, starting with the work of Weyl [Weyl] and ending with [MVIHES] with important intermediary results along the way such as [Bu, HB, Good, DFI1, DFI2, DFI3, BHM, Ven].

Beyond GL2{\text{GL}}_{2} this problem is far from well-understood, even for the Rankin–Selberg case GL2×GL2{\text{GL}}_{2}\times{\text{GL}}_{2}. Here we briefly summarize the progress for Rankin–Selberg L-functions.

The problem is solved if one assumes that the conductor of one of the representations, C(π1)C(\pi_{1}) say, is fixed, or more generally, is bounded by a small enough but fixed positive power of the conductor of the other (see [KMV, Sarnak, HaM, Ven, MVIHES, hu_triple_2017]. Some sporadic cases are known when both conductors vary simultaneously and have comparable sizes, as in [HoM, HolTem, ZYe, LMY]. These works usually assume disjoint ramifications.

In contrast the problem remains open when π2,vπ~1,v\pi_{2,v}\simeq\tilde{\pi}_{1,v} at places with ramifications, as in the QUE case. We shall refer to this case as the QUE-like case later on.

It is natural to ask if one can bridge the gap between the known cases and the QUE-like cases, allowing joint ramifications and some conductor dropping away from QUE-like cases for πi\pi_{i}.

Alternatively one can understand this problem using the Langlands functoriality and view, for example, the Rankin–Selberg L-function as a special case of standard L-functions on GL4{\text{GL}}_{4}. There has been some important progresses in the subconvexity problem for L-functions on higher rank groups as in [Li11, Mu15], and most recently [Ne21]. But the available results are specific to the “uniform growth” cases where the conductor does not drop. In practice, the techniques employed in these papers depend polynomially upon the extent to which the conductor drops, so they can be extended to the range where the conductor drops a little. However, it is generally much more difficult to let the conductor drop substantially.

In this paper we obtain new cases of subconvexity bounds for the Rankin–Selberg and triple product L-functions of GL2{\text{GL}}_{2}, allowing general number fields, joint ramifications and conductor dropping range away from the QUE-like cases. In the non-archimedean aspect a special case of our main results states the following.

Theorem 1.1.

Let π1\pi_{1} and π2\pi_{2} be cuspidal automorphic representations of GL2,F{\text{GL}}_{2,F} both with trivial central character whose archimedean components are bounded. Then

(1.1) L(π1×π2,1/2)C(π1×π2)1/4+ε(C(π1×π2)C(π2×π~2))δ.L(\pi_{1}\times\pi_{2},1/2)\ll C(\pi_{1}\times\pi_{2})^{1/4+\varepsilon}\left(\frac{C(\pi_{1}\times\pi_{2})}{C(\pi_{2}\times\tilde{\pi}_{2})}\right)^{-\delta}.

By re-indexing if necessary, the above upper bound provides a non-trivial power saving whenever there exists fixed γ>0\gamma>0 such that

(1.2) mini=1,2{C(πi×π~i)}C(π1×π2)1γ.\min_{i=1,2}\{C(\pi_{i}\times\tilde{\pi}_{i})\}\leq C(\pi_{1}\times\pi_{2})^{1-\gamma}.

Thus we obtain a subconvexity bound for the Rankin–Selberg L-function on PGL2(F){\mathrm{PGL}}_{2}(F) even when the conductor drops all the way down except the QUE-like cases.

Note that when Cv(π1×π2)C_{v}(\pi_{1}\times\pi_{2}) is large enough, we have by [BH17, Corollary 3.1]

Cv(π1×π2)maxi=1,2{Cv(πi×π~i)},C_{v}(\pi_{1}\times\pi_{2})\geq\max_{i=1,2}\{C_{v}(\pi_{i}\times\tilde{\pi}_{i})\},

with equality in the QUE-like case.

1.1. Conductor dropping range

Here we specify the meaning of conductor dropping range locally in terms of the Langlands correspondence. We fix a local field for the discussion.

In general if σ\sigma is an n-dimensional local Weil–Deligne representation which decomposes into irreducible subrepresentations as σi\oplus\sigma_{i}, then we say σ\sigma is in the conductor dropping range if any one of the normalized conductors C(σi)1dimσiC(\sigma_{i})^{\frac{1}{\dim\sigma_{i}}} has significantly different size compared with others. In the Rankin–Selberg/triple product case, let σi,v\sigma_{i,v} be associated to πi,v\pi_{i,v} via the local Langlands correspondence, and we say {πi,v}\{\pi_{i,v}\} are in the conductor dropping range if σ=iσi,v\sigma=\bigotimes\limits_{i}\sigma_{i,v} is so. As we shall also consider horizontal aspect uniformly, any fixed power of pp will also be considered as a significant difference.

Example 1.2.

Consider the case where ν=p\nu=p is a finite prime, and for i=1,2i=1,2,

πi,pπ(χi,χi1)\pi_{i,p}\simeq\pi(\chi_{i},\chi_{i}^{-1})

are principal series representations. Suppose for simplicity that p>2p>2 and χi\chi_{i} are not quadratic. The corresponding Weil–Deligne representation i=1,2σi,p\otimes_{i=1,2}\sigma_{i,p} decomposes into a sum of four characters (of the Galois group of p{\mathbb{Q}}_{p}), corresponding to the characters χ1±1χ2±1\chi_{1}^{\pm 1}\chi_{2}^{\pm 1} via Class Field Theory. Given a character χ\chi of p×{\mathbb{Q}}_{p}^{\times}, let c(χ)c(\chi) be its conductor exponent (that is the smallest integer such that χ\chi is trivial on 1+pc(χ)p1+p^{c(\chi)}{\mathbb{Z}}_{p}, and c(χ)=0c(\chi)=0 when χ\chi is unramified).

The conductor dropping range occurs when

c(χ1)=c(χ2)>c(χ1±1χ2±1)c(\chi_{1})=c(\chi_{2})>c(\chi^{\pm 1}_{1}\chi^{\pm 1}_{2})

for some (±1,±1)(\pm 1,\pm 1); for example it occurs when χ2=χ1±1\chi_{2}=\chi_{1}^{\pm 1}. This is also the range where

(1.3) C(π1×π2)<maxi{C(πi)2}.C(\pi_{1}\times\pi_{2})<\max\limits_{i}\{C(\pi_{i})^{2}\}.

In comparison, when c(χ1)c(χ2)c(\chi_{1})\neq c(\chi_{2}) or c(χ1)=c(χ2)=c(χ1±1χ2±1)c(\chi_{1})=c(\chi_{2})=c(\chi^{\pm 1}_{1}\chi^{\pm 1}_{2}), we have

C(π1×π2)=maxi{C(πi)2}.C(\pi_{1}\times\pi_{2})=\max\limits_{i}\{C(\pi_{i})^{2}\}.

There are similar examples when both π1,p,π2,p\pi_{1,p},\ \pi_{2,p} are supercuspidal representations constructed from related datum.

Consider now the case where the ramification of π\pi comes from essentially a single place, that is,

C(πi)C(πi,p)C(\pi_{i})\asymp C(\pi_{i,p})

for a large enough prime pp, and that π1,p,π2,p\pi_{1,p},\ \pi_{2,p} are principal series as above, then condition (1.2) is automatic if C(π1,p)C(π2,p)C(\pi_{1,p})\neq C(\pi_{2,p}). On the other hand if c(χ1)=c(χ2)2c(\chi_{1})=c(\chi_{2})\geq 2, (1.2) is satisfied if and only if

(1.4) min{c(χ1χ2),c(χ1χ21)}γc(χi2)\min\{c(\chi_{1}\chi_{2}),\ c(\chi_{1}\chi_{2}^{-1})\}\geq\gamma^{\prime}c(\chi_{i}^{2})

for some constant γ>0\gamma^{\prime}>0. (Actually γ=γ1γ\gamma^{\prime}=\frac{\gamma}{1-\gamma} with γ<1\gamma<1 in the current case.) This is because we have max{c(χ1χ2),c(χ1χ21)}=c(χi2)=12c(πi,p×πi,p)\max\{c(\chi_{1}\chi_{2}),\ c(\chi_{1}\chi_{2}^{-1})\}=c(\chi_{i}^{2})=\frac{1}{2}c(\pi_{i,p}\times\pi_{i,p}), and c(π1,p×π2,p)=2(c(χ1χ2)+c(χ1χ21))c(\pi_{1,p}\times\pi_{2,p})=2(c(\chi_{1}\chi_{2})+c(\chi_{1}\chi_{2}^{-1})). Here c(Πp)c(\Pi_{p}) is the exponent of C(Πp)C(\Pi_{p}).

On the other hand min{c(χ1χ2),c(χ1χ21)}=0\min\{c(\chi_{1}\chi_{2}),\ c(\chi_{1}\chi_{2}^{-1})\}=0 corresponds to the QUE-like case.

We leave it to readers to check that (1.2) and (1.4) remain equivalent when c(χ1)=c(χ2)=1c(\chi_{1})=c(\chi_{2})=1, in which case it could happen that c(χi2)=0c(\chi_{i}^{2})=0 for quadratic characters.

In practice, one can also use (1.3) as a working definition for the conductor dropping range when we work with representations of trivial central characters in this paper.

1.2. Main result and strategy of proof

To state our results for both the Rankin–Selberg and the triple product L-functions, we denote

Q=QQf=νQν=νCν(π1×π2×π3),Q=Q_{\infty}Q_{f}=\prod_{\nu}Q_{\nu}=\prod_{\nu}C_{\nu}(\pi_{1}\times\pi_{2}\times\pi_{3}),

and

P=PPf=vPv=vQv1/2maxi=2,3{Cv(πi×π~i)}.P=P_{\infty}P_{f}=\prod_{v}P_{v}=\prod_{v}\frac{Q_{v}^{1/2}}{\max\limits_{i=2,3}\{C_{v}(\pi_{i}\times\tilde{\pi}_{i})\}}.

Here in the Rankin–Selberg case, we take π3=π(1,1)\pi_{3}=\pi(1,1). Note that Qv=Pv=1Q_{v}=P_{v}=1 for almost every ν\nu, and Pv1P_{v}\geq 1 at least when QvQ_{v} is large enough. The main result we shall prove in this paper is the following theorem:

Theorem 1.3.

Suppose that π1,π2\pi_{1},\ \pi_{2} are cuspidal automorphic representations with trivial central characters. Suppose further that (Cf(π2),Cf(π3))=1(C_{f}(\pi_{2}),C_{f}(\pi_{3}))=1 (this is automatic in the Rankin–Selberg case since Cf(π3)=1C_{f}(\pi_{3})=1). Then there exists an absolute constant δ>0\delta>0 such that for any ε>0\varepsilon>0, one has

L(π1×π2×π3,1/2)\displaystyle L(\pi_{1}\times\pi_{2}\times\pi_{3},1/2) C(π1×π2×π3)1/4+ε1Pδ,\displaystyle\ll C(\pi_{1}\times\pi_{2}\times\pi_{3})^{1/4+\varepsilon}\frac{1}{P^{\delta}},
L(π1×π2,1/2)\displaystyle L(\pi_{1}\times\pi_{2},1/2) C(π1×π2)1/4+ε1Pδ.\displaystyle\ll C(\pi_{1}\times\pi_{2})^{1/4+\varepsilon}\frac{1}{P^{\delta}}.

Here the implicit constants depend on ε,F\varepsilon,F and (continuously) on the archimedean conductors C(πi)C_{\infty}(\pi_{i}) for i=1,2,3i=1,2,3.

In the proof below we actually start with more general situations and formulate a conjecture which potentially allows one to remove/relax the technical assumptions in this result. Currently we stick to the setting of Theorem 1.3.

The main global strategy follows and extends that of [MVIHES]. Note first that the cases of Rankin–Selberg L-function and triple product L-function can be put into a uniform framework in the sense that they have analogous integral representations: for G=GL2G={\text{GL}}_{2} or B×B^{\times} for BB some suitable quaternion algebra, depending on the (global and local) root numbers ϵ(π1×π2×π3,1/2)\epsilon(\pi_{1}\times\pi_{2}\times\pi_{3},1/2), the Rankin–Selberg theory/triple product formula provides an identity of the shape

(1.5) |[G]i=13φi(g)dg|2=|φ1,φ2φ3¯|2L(π1×π2×π3,1/2)vIv(φ1,φ2,φ3).\left|\int\limits_{[G]}\prod\limits_{i=1}^{3}\varphi_{i}(g)dg\right|^{2}=\left|\langle\varphi_{1},\overline{\varphi_{2}\varphi_{3}}\rangle\right|^{2}\sim L(\pi_{1}\times\pi_{2}\times\pi_{3},1/2)\cdot\prod_{v}I_{v}(\varphi_{1},\varphi_{2},\varphi_{3}).

Here [G]=ZG(𝔸)G(F)\G(𝔸)[G]=Z_{G}({\mathbb{A}})G(F)\backslash G({\mathbb{A}}); \sim means equality up to some unimportant factors; φiπiG\varphi_{i}\in\pi_{i}^{G} are L2L^{2}-normalized automorphic forms in the image of πi\pi_{i} under the Jacquet-Langlands correspondence; ,\langle\cdot,\cdot\rangle is the unitary integral pairing for automorphic forms on GG with the same central characters; the local integral Iv(φ1,φ2,φ3)I_{v}(\varphi_{1},\varphi_{2},\varphi_{3}) is an integral of products of local matrix coefficients in case of triple product L-function, or the local Rankin–Selberg integrals in case of Rankin–Selberg L-function. Furthermore in the Rankin–Selberg case, G=GL2G={\text{GL}}_{2}, the Jacquet-Langlands correspondence is the identity and φ3\varphi_{3} is an Eisenstein series constructed out of π(1,1)\pi(1,1).

On the one hand we need to control Iv(φ1,φ2,φ3)I_{v}(\varphi_{1},\varphi_{2},\varphi_{3}) from below; this is done by choosing appropriate factorable vectors φi\varphi_{i} which also captures the information of Q=C(π1×π2×π3)Q=C(\pi_{1}\times\pi_{2}\times\pi_{3}) and detects the conductor dropping range.

On the other hand, we also have to control

|φ1,φ2φ3¯|2|\langle\varphi_{1},\overline{\varphi_{2}\varphi_{3}}\rangle|^{2}

from above; we do this by applying the Cauchy–Schwarz inequality, and then the Plancherel formula (the regularized version in the Rankin–Selberg case) to get

(1.6) |φ1,φ2φ3¯|2φ2φ2¯,φ3φ3¯=πψB(π)φ2φ2¯,ψφ3φ3¯,ψ¯.\displaystyle|\langle\varphi_{1},\overline{\varphi_{2}\varphi_{3}}\rangle|^{2}\leq\langle\varphi_{2}\overline{\varphi_{2}},\varphi_{3}\overline{\varphi_{3}}\rangle=\int\limits_{\pi^{\prime}}\sum\limits_{\psi\in B(\pi^{\prime})}\langle\varphi_{2}\overline{\varphi_{2}},\psi\rangle\overline{\langle\varphi_{3}\overline{\varphi_{3}},\psi\rangle}.

Note that the first inequality is the source of asymmetry in the definition of PP.

When all πi\pi_{i} are varying, the length of spectral sum/integral could however become very long. Denote

Mv:=min{Cv(π2),Cv(π3)}M:=vMv.M_{v}:=\min\{C_{v}(\pi_{2}),C_{v}(\pi_{3})\}\hbox{,\ \ }M:=\prod_{v}M_{v}.

We further assume that there exist test vectors φi\varphi_{i} such that for every place vv, if jv=2j_{v}=2 or 33 (which may depend on vv) is the index ii of πi,v\pi_{i,v} which obtains the minimum MvM_{v}, that is,

Cv(πjv)=Mv,C_{v}(\pi_{j_{v}})=M_{v},

then the complexity of φjv,v\varphi_{j_{v},v} is controlled by MvM_{v}. See Assumption 3.10 for the precise meaning. Then one can restrict the spectral sum/integral in (1.6) to those π\pi^{\prime} and ψ\psi that are also controlled in terms of MvM_{v}, in which range we control the contributions term-wisely.

To study the individual terms φ2φ2¯,ψφ3φ3¯,ψ¯\langle\varphi_{2}\overline{\varphi_{2}},\psi\rangle\overline{\langle\varphi_{3}\overline{\varphi_{3}},\psi\rangle} for non-residual spectrum, we apply the period relation (1.5) again and the known convexity bound for the resulting LL-functions L(πi×π~i×π,1/2)L(\pi_{i}\times\tilde{\pi}_{i}\times\pi^{\prime},1/2). One also needs a reasonable upper bound for Iv(φi,φi¯,ψ)I_{v}(\varphi_{i},\overline{\varphi_{i}},\psi), by which we expect a power saving in terms of PP for non-residual spectrum which, combined with the amplification method, should give a power saving for the initial global period φ1,φ2φ3¯\langle\varphi_{1},\overline{\varphi_{2}\varphi_{3}}\rangle.

To summarize, we expect the global consideration above to reduce the subconvexity problem to the following test vector problem in our setting:

Conjecture 1.4.

There exist factorable (normalized) test vectors φiπi\varphi_{i}\in\pi_{i} satisfying the following three local requirements.

  1. (0)

    The local component φjv,v\varphi_{j_{v},v} should be controlled by MvM_{v}, to control the length of the spectral sum/integral;

  2. (1)

    Iv(φ1,φ2,φ3)I_{v}(\varphi_{1},\varphi_{2},\varphi_{3}) is bounded from below by 1Qv1/4;\frac{1}{Q_{v}^{1/4}};

  3. (2)

    Iv(φi,φi¯,ψ)I_{v}(\varphi_{i},\overline{\varphi_{i}},\psi) is controlled from above to provide power saving in terms of PvP_{v}, while ψ\psi is controlled in terms of MM.

See Conjecture 3.17 below for more precise formulation. Note that scaling the test vectors would change the requirements for (1) and (2) simultaneously. So one can assume without loss of generality that φi\varphi_{i} are L2L^{2}-normalized.

The main challenge of the test vector problem is the lower bound in item (1). The existence of such test vectors is partially supported by previous know cases away from the conductor dropping range. Within the conductor dropping range however, the required lower bound for Iv(φ1,φ2,φ3)I_{v}(\varphi_{1},\varphi_{2},\varphi_{3}) becomes substantially larger compared with the non-conductor dropping range. Whether this is possible is not clear at all from, for example, [MVIHES] or other literature.

There are two main sources of test vectors, coming from translates of classical newforms, or localized vectors (or minimal vectors) as used in [Ne18, HNS, HN18, NV, Ne21]. The newforms are commonly used in the cases with disjoint ramifications, but they have more complicated matrix coefficients and Whittaker functions when involved in the conductor dropping range. On the other hand the localized vectors have simpler description of matrix coefficients/Whittaker functions, but they do not work well for representations with small levels, requiring more case by case discussions.

To our surprise, in the setting of Theorem 1.3 (in level aspect, with trivial central character, and M=1M=1), we are able to find test vectors uniformly using only diagonal translates of newforms. The complicated nature of matrix coefficients/Whittaker functions of newforms from different πi\pi_{i} turns out to be reflecting the conductor dropping range in a relatively simple fashion. The extent of diagonal translates would then have to match the extent of conductor dropping to achieve item (1) of Conjecture 1.4.

We also have some sporadic evidences for the conjecture when MM is square-free, or in certain archimedean aspects, but we shall skip these cases for the sake of conciseness. In more general situations, the localized vectors may also turn out to be useful.

1.3. Main local ingredient

Here we briefly discuss the main local ingredient to verify item (1) of Conjecture 1.4, in the conductor dropping range where c=c(π1,v)=c(π2,v)c=c(\pi_{1,v})=c(\pi_{2,v}) is the exponent of the local conductors at vv.

Consider for simplicity the Rankin–Selberg case. The local representations πi,v,i=1,2\pi_{i,v},i=1,2 of PGL2(Fv){\mathrm{PGL}}_{2}(F_{v}), in most cases, can be associated to some characters θi\theta_{i} over some étale quadratic algebra Ei/FvE_{i}/F_{v} by the local Langlands correspondence. Then the evaluation of the local period integral IvI_{v} can be closely related to the following correlation between twisted Gauss sum/integrals:

Pk(θ1,θ2)=c(χ)=ckθ11(u)χE1(u)ψE1(u)d×uθ21(w)χE2(w)ψE2(w)d×w¯.P_{k}(\theta_{1},\theta_{2})=\sum\limits_{c(\chi)=c-k}\int\ \theta_{1}^{-1}(u)\chi_{E_{1}}(u)\psi_{E_{1}}(u)d^{\times}u\overline{\int\theta_{2}^{-1}(w)\chi_{E_{2}}(w)\psi_{E_{2}}(w)d^{\times}w}.

Here ψEi=ψTrEi/Fv\psi_{E_{i}}=\psi\circ\text{Tr}_{E_{i}/F_{v}} is an additive character on EiE_{i}, χ\chi is a multiplicative character of Fv×F_{v}^{\times} with specified level, and χEi=χNmEi/Fv\chi_{E_{i}}=\chi\circ\text{Nm}_{E_{i}/F_{v}}.

An interesting phenomenon we find is that this quantity PkP_{k} is exactly capturing the conductor dropping range: when kc(π1,v×π2,v)/2k\geq c(\pi_{1,v}\times\pi_{2,v})/2, there is no cancellation in the sum over χ\chi. On the other hand when c/2k<c(π1,v×π2,v)/2c/2\leq k<c(\pi_{1,v}\times\pi_{2,v})/2, there are heavy cancellations, and PkP_{k} becomes much smaller and equals zero in many cases. The choice of test vector is then reduced to maximizing the contributions from those PkP_{k} with kc(π1,v×π2,v)/2k\geq c(\pi_{1,v}\times\pi_{2,v})/2. This phenomenon may also have independent interest.

1.4. The structure of the paper

  1. -

    After setting up notations in Section 2, we recall in Section 3 the triple product formula and the amplification method, and prove the main result Theorem 3.14 that reduces the subconvexity problem for L(π1×π2×π3,1/2)L(\pi_{1}\times\pi_{2}\times\pi_{3},1/2) to Conjecture 3.17, which becomes a local problem.

  2. -

    We recall in Section 4 some more local preparations including the local Whittaker functions. In Section 5 we prove Proposition 5.4 for the partial pairing PkP_{k} which detects the conductor dropping range. Finally in Section 6 we verify Conjecture 3.17 for the setting in Theorem 1.3.

1.5. Acknowledgment

The first author is supported by the National Key Research and Development Program of China (No. 2021YFA1000700).

2. Notations

Let FF be a number field. We denote by 𝒱\mathcal{V} the set of places of FF, by 𝒱f\mathcal{V}_{f} and 𝒱\mathcal{V}_{\infty} the finite and archimedean places; we denote by FvF_{v} the local field at vv. We denote by 𝔸=vFv{\mathbb{A}}=\mathop{{\prod}^{\mathbf{{}^{\prime}}}}\limits_{v}F_{v} the ring of adèles of FF and for any S𝒱S\subset\mathcal{V},

𝔸S=vSFv,𝔸(S)=vSFv.{\mathbb{A}}_{S}=\mathop{{\prod}^{\mathbf{{}^{\prime}}}}\limits_{v\in S}F_{v},\ {\mathbb{A}}^{(S)}=\mathop{{\prod}^{\mathbf{{}^{\prime}}}}\limits_{v\not\in S}F_{v}.

We denote by ψ:F\𝔸×\psi:F\backslash{\mathbb{A}}\rightarrow{\mathbb{C}}^{\times} the additive character ψ(x)=ψ(trF/(x))\psi(x)=\psi_{\mathbb{Q}}({\rm tr}_{F/{\mathbb{Q}}}(x)) where ψ\psi_{\mathbb{Q}} is the additive character on \𝔸{\mathbb{Q}}\backslash{\mathbb{A}}_{\mathbb{Q}} whose restriction to {\mathbb{R}} is exp(2πix)\exp(2\pi ix_{\mathbb{R}}); the character ψ\psi decomposes into a product of characters of FvF_{v}: ψ=vψv.\psi=\prod_{v}\psi_{v}.

Let G{\mathrm{G}} be a reductive group over FF. For any place v𝒱v\in\mathcal{V} we denote by Gv=G(Fv){{\mathrm{G}}_{v}}={\mathrm{G}}(F_{v}) the groups of FvF_{v} points, by G(𝔸)=v𝒱Gv{{\mathrm{G}}({\mathbb{A}})}=\mathop{{\prod}^{\mathbf{{}^{\prime}}}}\limits_{v\in\mathcal{V}}{{\mathrm{G}}_{v}} the group of adelic points, and more generally for any S𝒱S\subset\mathcal{V}, we denote

G(𝔸S)=vSGv,G(𝔸(S))=vSGv.{\mathrm{G}}({\mathbb{A}}_{S})=\mathop{{\prod}^{\mathbf{{}^{\prime}}}}\limits_{v\in S}G_{v},\ {\mathrm{G}}({\mathbb{A}}^{(S)})=\mathop{{\prod}^{\mathbf{{}^{\prime}}}}\limits_{v\not\in S}G_{v}.

We set

[Gv]:=Fv×\Gv,[G]:=G(F)𝔸×\G(𝔸).[{{\mathrm{G}}_{v}}]:=F_{v}^{\times}\backslash{{\mathrm{G}}_{v}},\ [{\mathrm{G}}]:={{\mathrm{G}}(F)}{\mathbb{A}}^{\times}\backslash{{\mathrm{G}}({\mathbb{A}})}.

We choose for each place vv a maximal compact subgroup KvGvK_{v}\subset{{\mathrm{G}}_{v}} and for vv finite a decreasing family of principal congruence subgroups Kv[m]Kv,m𝐍K_{v}[m]\subset K_{v},\ m\in{\bf{N}} as well as Haar measures on Gv{{\mathrm{G}}_{v}}, G(𝔸){{\mathrm{G}}({\mathbb{A}})} and [G][{\mathrm{G}}] as in [MVIHES, §2.1]. For vv non-archimedean, KvK_{v} has measure 11.

In this paper G{\mathrm{G}} will most of the time be equal to B×{{\mathrm{B}}^{\times}} for B{\mathrm{B}} some quaternion algebra over FF (or product of such groups) and for any vv such that Gv=GL2,v{{\mathrm{G}}_{v}}={\text{GL}}_{2,v} we make the same choice of Haar measures as in [MVIHES, §3.1.5]. The choice of a maximal order 𝒪BB\mathscr{O}_{\mathrm{B}}\subset{\mathrm{B}} and of a {\mathbb{Z}}-basis define local norms on Bv{\mathrm{B}}_{v} for every vv. For u=(uv)B×(𝔸)u=(u_{v})\in{{\mathrm{B}}^{\times}({\mathbb{A}})} we denote by

(2.1) u=vuvv\|u\|=\prod_{v}\|u_{v}\|_{v}

the operator norm of Adu\mathrm{Ad}u, the action of uu on B(𝔸){\mathrm{B}}({\mathbb{A}}) by conjugation. The infinite product in the definition of u\|u\| is justified by the following straightforward result:

Lemma 2.1.

Suppose Gv=GL2,vG_{v}={\text{GL}}_{2,v} and uv=zk1a(x)k2u_{v}=zk_{1}a(x)k_{2} in the local Cartan decomposition. Then

uvvmax{|x|v,|x|v1}\|u_{v}\|_{v}\asymp\max\{|x|_{v},|x|_{v}^{-1}\}

with absolute implied constant. Furthermore if vv is a p-adic place and xOv×x\in O_{v}^{\times},

uvv=1.\|u_{v}\|_{v}=1.

On the other hand if GvG_{v} comes from a local division algebra, we have

uvv1\|u_{v}\|_{v}\asymp 1

by compactness.

2.1. Automorphic representations

Let G=B×{\mathrm{G}}={{\mathrm{B}}^{\times}}; given χ\chi a unitary character of F×\𝔸×F^{\times}\backslash{\mathbb{A}}^{\times} we denote by 𝒜(G,χ)\mathcal{A}({\mathrm{G}},\chi) the set of (unitary) automorphic representations of G(𝔸){{\mathrm{G}}({\mathbb{A}})} with central character χ\chi. Given such a representation we denote by (πv)v𝒱(\pi_{v})_{v\in\mathcal{V}} the sequence of its local constituents: these are unitary Gv{{\mathrm{G}}_{v}}-representation all but finitely many of which are unramified principal series representations (of Gv=GL2(Fv){{\mathrm{G}}_{v}}={\text{GL}}_{2}(F_{v}) for vv finite), and one has πvπv\pi\simeq\bigotimes_{v}\pi_{v} the restricted tensor product being taken with respect to a sequence of spherical vectors (φv)v(\varphi^{\circ}_{v})_{v} indexed by the finite places where π\pi is unramified (admits a KvK_{v}-invariant vector). We denote by unram(π)𝒱f\mathrm{unram}(\pi)\subset\mathcal{V}_{f} the set of (finite) unramified places of π\pi and ram(π)𝒱\mathrm{ram}(\pi)\subset\mathcal{V} its complement111 by convention infinite places are contained in ram(π)\mathrm{ram}(\pi) even the ones for which πv\pi_{v} is spherical. At any place where Gv=GL2,v{{\mathrm{G}}_{v}}={\text{GL}}_{2,v}, we denote by 𝒲(πv){\mathcal{W}}(\pi_{v}) the Whittaker model of πv\pi_{v} relative to the additive character ψv\psi_{v} and for vunram(π)v\in\mathrm{unram}(\pi), we take the spherical vector W𝒲(πv)W^{\circ}\in{\mathcal{W}}(\pi_{v}) such that W(1)=1W^{\circ}(1)=1.

We define inner products on π\pi and πv\pi_{v} as follows: For φiπ\varphi_{i}\in\pi, i=1,2i=1,2,

φ1,φ2=[G]φ1(g)φ2(g)¯𝑑g;\langle\varphi_{1},\varphi_{2}\rangle=\int\limits_{[G]}\varphi_{1}(g)\overline{\varphi_{2}(g)}dg;

For φi,vπv\varphi_{i,v}\in\pi_{v} and Wi,v=Wφi,vW_{i,v}=W_{\varphi_{i},v} the associated Whittaker functions in the case Bv×=GL2(Fv)B_{v}^{\times}={\text{GL}}_{2}({F}_{v}),

φ1,v,φ2,v=Fv×W1,v(a(x))W2,v(a(x))¯d×x.\langle\varphi_{1,v},\varphi_{2,v}\rangle=\int\limits_{F_{v}^{\times}}W_{1,v}(a(x))\overline{W_{2,v}(a(x))}d^{\times}x.

The L2L^{2}-norms of φ\varphi and φv\varphi_{v} are defined with respect to these pairings. Inner products can also be defined for compactly induced representations when BvB_{v} is a division algebra, though we do not need the details in this paper.

We denote by π~𝒜(G,χ1)\tilde{\pi}\in\mathcal{A}({\mathrm{G}},\chi^{-1}) the contragredient representation of π\pi, and by πJL𝒜(GL2,χ)\pi^{JL}\in\mathcal{A}({\mathrm{GL}_{2}},\chi) the automorphic representation corresponding to π\pi under the Jacquet-Langlands correspondence (one has πvπvJL\pi_{v}\simeq\pi^{JL}_{v} for vram(π)v\not\in\mathrm{ram}(\pi)).

Convention 1.

Unless stated otherwise a local unitary representation of Gv{{\mathrm{G}}_{v}} for some vv will always be understood as a local constituent of some global automorphic representation. In particular, the Langlands parameters (when the representation is unramified) satisfy the best known approximation towards the Ramanujan-Petersson conjecture.

Convention 2.

To ease notations (and as long as the context is clear) we will not distinguish between π\pi and the Jacquet-Langlands constituent πJL\pi^{JL} at least notationally: suppose we have a quantity, say L(πJL)L(\pi^{JL}), which a priori is constructed out of πJL\pi^{JL}: we will denote it indifferently L(πJL)L(\pi^{JL}) or L(π)L(\pi). This convention will be in use especially to denote L-functions and related quantities like conductors.

2.2. L-functions

Let ΠvΠv\Pi\simeq\bigotimes_{v}\Pi_{v} be an automorphic representation of some reductive group G{\mathrm{G}} and ρ\rho a representation of its dual group GL{\mathrm{G}}^{L} of dimension dd; to these data, one associates an L-function given by an Euler product of local LL-factors of degree dd converging in some halfspace {s,Res1}\{s,\ \operatorname{Re}s\gg 1\}

L(Π,ρ,s)=v𝒱fLv(Π,ρ,s),L(\Pi,\rho,s)=\prod_{v\in\mathcal{V}_{f}}L_{v}(\Pi,\rho,s),

where Lv(Π,ρ,s)=i=1d(1αv,i(Π,ρ)qvs)1L_{v}(\Pi,\rho,s)=\prod_{i=1}^{d}(1-\frac{\alpha_{v,i}(\Pi,\rho)}{q_{v}^{s}})^{-1} at unramified places for αv,i(Π,ρ)\alpha_{v,i}(\Pi,\rho) the eigenvalues of ρ(c(Πv))\rho(c(\Pi_{v})) where c(Πv)c(\Pi_{v}) is the Satake matrix associated to Πv\Pi_{v}. At ramified places, LvL_{v} is often defined as the common denominator of the relevant local period integrals, or using the local Langlands correspondence.

The archimedean factors are defined similarly as

L(Π,ρ,s)=v𝒱Lv(Π,ρ,s),Lv(Π,ρ,s)=i=1dΓFv(s+αv,i(Π,ρ)).L_{\infty}(\Pi,\rho,s)=\prod_{v\in\mathcal{V}_{\infty}}L_{v}(\Pi,\rho,s),\ L_{v}(\Pi,\rho,s)=\prod_{i=1}^{d}\Gamma_{F_{v}}(s+\alpha_{v,i}(\Pi,\rho)).

In several cases one can prove that L(Π,ρ,s)L(\Pi,\rho,s) admits analytic continuation to {\mathbb{C}} with at most a finite number of poles located on the line Res=1\operatorname{Re}s=1 and that it admits a functional equation of the shape

Λ(Π,ρ,s)=ε(Π,ρ)Cf(Π,ρ)12sΛ(Π,ρ,1s¯)¯\Lambda(\Pi,\rho,s)=\varepsilon(\Pi,\rho)C_{f}(\Pi,\rho)^{\frac{1}{2}-s}\overline{\Lambda(\Pi,\rho,1-\overline{s})}

where

Λ(Π,ρ,s)=L(Π,ρ,s)L(Π,ρ,s),|ε(Π,ρ)|=1,Cf(Π,ρ)𝐍1.\Lambda(\Pi,\rho,s)=L_{\infty}(\Pi,\rho,s)L(\Pi,\rho,s),\ |\varepsilon(\Pi,\rho)|=1,\ C_{f}(\Pi,\rho)\in{\bf{N}}_{\geq 1}.

The factor ε(Π,ρ)\varepsilon(\Pi,\rho) is the root number and Cf(Π,ρ)C_{f}(\Pi,\rho) is the arithmetic conductor. Both quantities can also be factored into a product of local root numbers (upon choosing a non-trivial additive character of F\𝔸F\backslash{\mathbb{A}}) and local conductors which are almost everywhere equal to 11 and arise from local functional equations:

ε(Π,ρ)=vεv(Π,ρ),Cf(Π,ρ)=v𝒱FCv(Π,ρ).\varepsilon(\Pi,\rho)=\prod_{v}\varepsilon_{v}(\Pi,\rho),\ C_{f}(\Pi,\rho)=\prod_{v\in\mathcal{V}_{F}}C_{v}(\Pi,\rho).

One define the analytic conductor of (Π,ρ)(\Pi,\rho) by completing the arithmetic conductor with archimedean local conductors:

C(Π,ρ)=Cf(Π,ρ)C(Π,ρ)=Cf(Π,ρ)v𝒱Cv(Π,ρ)C(\Pi,\rho)=C_{f}(\Pi,\rho)C_{\infty}(\Pi,\rho)=C_{f}(\Pi,\rho)\prod_{v\in\mathcal{V}_{\infty}}C_{v}(\Pi,\rho)
Cv(Π,ρ)=i=1d(1+|αv,i(Π,ρ)|)[Fv:],v𝒱.C_{v}(\Pi,\rho)=\prod_{i=1}^{d}(1+|\alpha_{v,i}(\Pi,\rho)|)^{[F_{v}:{\mathbb{R}}]},\ v\in\mathcal{V}_{\infty}.

2.2.1. Convexity bound

The following bound is also sometimes known: for any ε>0\varepsilon>0, one has

L(Π,ρ,1/2)F,d,εC(Π,ρ)1/4+ε.L(\Pi,\rho,1/2)\ll_{F,d,\varepsilon}C(\Pi,\rho)^{1/4+\varepsilon}.

The subconvexity problem consist in improving the exponent 1/41/4 to one strictly smaller.

All the properties mentionned above are known for the following L-functions

  • -

    Hecke-Godement-Jacquet L-functions: n1n\geq 1, G=GLn{\mathrm{G}}={\text{GL}}_{n}, ρ=Std\rho=\mathrm{Std}, d=nd=n. The L-function will be noted L(Π,s)L(\Pi,s).

  • -

    Rankin–Selberg L-functions: G=GL2×GL2{\mathrm{G}}={\mathrm{GL}_{2}}\times{\mathrm{GL}_{2}}, ρ=StdStd\rho=\mathrm{Std}\otimes\mathrm{Std}, d=4d=4. If Π=π1×π2\Pi=\pi_{1}\times\pi_{2}, L-function will be noted L(π1×π2,s)L(\pi_{1}\times\pi_{2},s).

  • -

    Adjoint L-functions: G=GL2{\mathrm{G}}={\mathrm{GL}_{2}}, ρ=Ad\rho=\mathrm{Ad}, d=3d=3. The L-function will be noted L(Π,Ad,s)L(\Pi,\mathrm{Ad},s).

  • -

    Triple product L-function: G=GL2×GL2×GL2{\mathrm{G}}={\mathrm{GL}_{2}}\times{\mathrm{GL}_{2}}\times{\mathrm{GL}_{2}}, ρ=StdStdStd\rho=\mathrm{Std}\otimes\mathrm{Std}\otimes\mathrm{Std}, d=8d=8. If Π=π1×π2×π3\Pi=\pi_{1}\times\pi_{2}\times\pi_{3}, the L-function will be noted L(π1×π2×π3,s)L(\pi_{1}\times\pi_{2}\times\pi_{3},s).

In the following we will write

(2.2) C(π1×π2×π3)=vCv(π1×π2×π3)=Q=vQv.C(\pi_{1}\times\pi_{2}\times\pi_{3})=\prod_{v}C_{v}(\pi_{1}\times\pi_{2}\times\pi_{3})=Q=\prod_{v}Q_{v}.

2.2.2. Bound for adjoint L-functions

Given π𝒜(GL2,χ)\pi\in\mathcal{A}({\mathrm{GL}_{2}},\chi) it follows from [GHL] that

L(π,Ad,1)=C(π)o(1) as C(π).L(\pi,\mathrm{Ad},1)=C(\pi)^{o(1)}\hbox{ as }C(\pi)\rightarrow\infty.

2.3. Convention for inequalities

For two functions f1f_{1} f2f_{2} of variables a,b,ca,b,c,etc., by

f1(a,b,)a,bf2(a,b,),f_{1}(a,b,\cdots)\ll_{a,b}f_{2}(a,b,\cdots),

we mean f1(a,b,)Ca,bf2(a,b,)f_{1}(a,b,\cdots)\leq C_{a,b}f_{2}(a,b,\cdots) for some implied constant Ca,bC_{a,b} depending only on aa and bb, independent of other variables. This convention is commonly used among analytic number theorists.

In the following we will omit additional variables from notations without confusion. To deal with infinite products over all places, we also introduce the following notation:

Definition 2.2.

Let f1,f2f_{1},f_{2} be two functions of v𝒱v\in\mathcal{V} and other variables a,b,ca,b,c, etc. By

f1v,af2,f_{1}\leq_{v,a}f_{2},

we mean f1v,af2f_{1}\ll_{v,a}f_{2} with implied constant Cv,aC_{v,a}, and Cv,a=1C_{v,a}=1 if the cardinality of residue field qvNaq_{v}\geq N_{a} where NaN_{a} depends only on aa.

We define f1v,af2f_{1}\geq_{v,a}f_{2} similarly. By f1v,af2f_{1}\asymp_{v,a}f_{2}, we mean f1v,af2f_{1}\leq_{v,a}f_{2} and f1v,af2f_{1}\geq_{v,a}f_{2}.

The following lemma is straightforward:

Lemma 2.3.

If f1(v)=f2(v)=1f_{1}(v)=f_{2}(v)=1 for almost all vv, and f1v,af2f_{1}\leq_{v,a}f_{2}, then

vf1avf2.\prod\limits_{v}f_{1}\ll_{a}\prod\limits_{v}f_{2}.

The nuance here is that the number of local factors fi(v)f_{i}(v) which are not 11 may depend on other variables.

Example 2.4.

Let f(v)=ζv(1)f(v)=\zeta_{v}(1) for v|Qfv|Q_{f} and f(v)=1f(v)=1 otherwise. Then

Qvϵv,ϵf(v)v,ϵQvϵ,Q_{v}^{-\epsilon}\leq_{v,\epsilon}f(v)\leq_{v,\epsilon}Q_{v}^{\epsilon},
Qϵϵvf(v)ϵQϵ.Q^{-\epsilon}\ll_{\epsilon}\prod\limits_{v}f(v)\ll_{\epsilon}Q^{\epsilon}.

Same conclusions hold if f(v)=Cf(v)=C some fixed constant for v|Qfv|Q_{f} and f(v)=1f(v)=1 otherwise.

3. bounds for global periods

Let πi,i=1,2,3\pi_{i},\ i=1,2,3 be three generic automorphic representations of GL2(𝔸){\mathrm{GL}_{2}}({\mathbb{A}}) whose product of central characters is trivial; we will moreover assume that π1\pi_{1} and π2\pi_{2} are cuspidal.

The basic analytic properties stated in §2.2 (including the convexity bound) are known for the associated triple product L-function L(π1×π2×π3,s)L(\pi_{1}\times\pi_{2}\times\pi_{3},s). If π3\pi_{3} is a principal series representation this is the Rankin–Selberg theory; when π3\pi_{3} is cuspidal these are non-trivial facts which are consequences of the work of several people including Garrett, Piatestsky-Schapiro–Rallis and Ramakrishnan [Garrett, PSR, Ram]. L(π1×π2×π3,s)L(\pi_{1}\times\pi_{2}\times\pi_{3},s) is holomorphic unless π3\pi_{3} is a principal series and π1,π2\pi_{1},\pi_{2} are contragredient to one another up to a twist. Note that the convexity bound follows from Ramakrishnan’s proof that L(π1×π2,s)L(\pi_{1}\times\pi_{2},s) is GL4{\text{GL}}_{4}-automorphic and from an argument of Molteni [Mol].

The condition on the product of central characters being trivial implies that L(π1×π2×π3,s)L(\pi_{1}\times\pi_{2}\times\pi_{3},s) has real coefficients and that the global and local root numbers

ε(π1×π2×π3)=v𝒱εv(π1×π2×π3)\varepsilon(\pi_{1}\times\pi_{2}\times\pi_{3})=\prod_{v\in\mathcal{V}}\varepsilon_{v}(\pi_{1}\times\pi_{2}\times\pi_{3})

are all ±1\pm 1.

The subconvexity problem we are studying in this paper amounts to finding an absolute δ>0\delta>0 such that

L(π1×π2×π3,1/2)FC(π1×π2×π3)1/4δ.L(\pi_{1}\times\pi_{2}\times\pi_{3},1/2)\ll_{F}C(\pi_{1}\times\pi_{2}\times\pi_{3})^{1/4-\delta}.

3.1. The triple product formula

We may assume that

(3.1) ε(π1×π2×π3)=+1,\varepsilon(\pi_{1}\times\pi_{2}\times\pi_{3})=+1,

as otherwise L(π1×π2×π3,1/2)=0L(\pi_{1}\times\pi_{2}\times\pi_{3},1/2)=0 and we are done. This implies that the set of places of FF at which the local root number equals 1-1 has even cardinality; this set determines a unique quaternion algebra B{\mathrm{B}} (the one ramified precisely at this even set of places) and three B×{{\mathrm{B}}^{\times}}-automorphic representation (π1B,π2B,π3B)(\pi^{\mathrm{B}}_{1},\pi^{\mathrm{B}}_{2},\pi^{\mathrm{B}}_{3}) whose images under the Jacquet-Langlands correspondence are (π1,π2,π3)(\pi_{1},\pi_{2},\pi_{3}). More precisely, at any place for which B{\mathrm{B}} is unramified, Bv×=GL2,v{\mathrm{B}}^{\times}_{v}={\text{GL}}_{2,v} and πi,vBπi,v\pi_{i,v}^{\mathrm{B}}\simeq\pi_{i,v} while for any place where B{\mathrm{B}} is ramified, πi,vB\pi_{i,v}^{\mathrm{B}} is finite dimensional and πi,v\pi_{i,v} is the image of πi,vB\pi_{i,v}^{\mathrm{B}} under the local Jacquet-Langlands correspondence; note that in the Rankin–selberg case, when π3\pi_{3} is a principal series representation, B×=GL2{{\mathrm{B}}^{\times}}={\mathrm{GL}_{2}} and the correspondence is the identity.

From now on, the quaternion algebra B{\mathrm{B}} will be considered fixed; we write

π=π1B𝒜(G,χ1),Π=π2Bπ3B𝒜(G×G,χ11).\pi=\pi^{{\mathrm{B}}}_{1}\in\mathcal{A}({\mathrm{G}},\chi_{1}),\ \Pi=\pi^{{\mathrm{B}}}_{2}\otimes\pi^{{\mathrm{B}}}_{3}\in\mathcal{A}({\mathrm{G}}\times{\mathrm{G}},\chi^{-1}_{1}).

The later is an automorphic representation of G(𝔸)×G(𝔸){{\mathrm{G}}({\mathbb{A}})}\times{{\mathrm{G}}({\mathbb{A}})} on which G(𝔸){{\mathrm{G}}({\mathbb{A}})} acts via the diagonal embedding GG×G{\mathrm{G}}\hookrightarrow{\mathrm{G}}\times{\mathrm{G}}.

It follows from the work of Prasad [Pra] that for any place vv the space

HomGv(Πv,π~v)\operatorname{Hom}_{{\mathrm{G}}_{v}}(\Pi_{v},\tilde{\pi}_{v}) of Gv{{\mathrm{G}}_{v}}-invariant intertwiners is one dimensional and a generator is given by the functional, PrΠvπv\mathrm{Pr}^{\pi_{v}}_{\Pi_{v}} say, well defined up to a scalar of modulus 11 satisfying for any φvΦvπvΠv\varphi_{v}\otimes\varPhi_{v}\in\pi_{v}\otimes\Pi_{v}

|PrΠvπv(Φv)(φv)|2=Lv(π,Ad,1)Lv(Π,Ad,1)ζFv(2)2Lv(π×Π,1/2)[Gv]φv,g.φvΦv,g.Φvdg.\left|\mathrm{Pr}^{\pi_{v}}_{\Pi_{v}}(\varPhi_{v})(\varphi_{v})\right|^{2}=\frac{L_{v}(\pi,\mathrm{Ad},1)L_{v}(\Pi,\mathrm{Ad},1)}{\zeta_{F_{v}}(2)^{2}L_{v}(\pi\times\Pi,1/2)}\int_{[{{\mathrm{G}}_{v}}]}\langle\varphi_{v},g.\varphi_{v}\rangle\langle\varPhi_{v},g.\varPhi_{v}\rangle dg.
Remark 3.1.

Observe that the Gv{{\mathrm{G}}_{v}}-integral is absolutely converging due to the known bounds for matrix coefficient [KiSa, BlBr] and the assumption that πv\pi_{v} and Πv\Pi_{v} are local components of global automorphic representations.

Furthermore by work of Garrett, Piatetsky-Shapiro–Rallis, Harris-Kudla, Watson and Ichino [Garrett, PSR, HK, Watson, Ichino], the space of global G(𝔸){{\mathrm{G}}({\mathbb{A}})}-equivariant intertwiners HomG(𝔸)(Π,π~)\operatorname{Hom}_{{\mathrm{G}}({\mathbb{A}})}(\Pi,\tilde{\pi}) is non-zero (and therefore one dimensional) if an only if

L(π×Π,1/2)0L(\pi\times\Pi,1/2)\not=0

and then is generated by PrΠπ\mathrm{Pr}^{\pi}_{\Pi} given by

(3.2) PrΠπ(Φ)(φ)\displaystyle\mathrm{Pr}^{\pi}_{\Pi}(\varPhi)(\varphi) =[G]φ(g)Φ(g)𝑑g\displaystyle=\int_{[{\mathrm{G}}]}\varphi(g)\varPhi(g)dg
PrΠπ(Φ)\displaystyle\mathrm{Pr}^{\pi}_{\Pi}(\varPhi) =φ(π)PrΠπ(Φ)(φ)\displaystyle=\sum\limits_{\varphi\in\mathcal{B}(\pi)}\mathrm{Pr}^{\pi}_{\Pi}(\varPhi)(\varphi)

where Φ\varPhi is viewed as a function on G(𝔸)×G(𝔸){{\mathrm{G}}({\mathbb{A}})}\times{{\mathrm{G}}({\mathbb{A}})}, gΦ(g)g\mapsto\varPhi(g) denotes the restriction of Φ\varPhi to diagonally-embedded G(𝔸){{\mathrm{G}}({\mathbb{A}})} in G(𝔸)×G(𝔸){{\mathrm{G}}({\mathbb{A}})}\times{{\mathrm{G}}({\mathbb{A}})} and (π)\mathcal{B}(\pi) denotes an orthonormal basis for π\pi; note that the integral is converging since π\pi is cuspidal. More precisely the following period formula relating global and local intertwiners holds ([Ichino]):

Theorem 3.2.

For φΦvφvΦv\varphi\otimes\varPhi\simeq\bigotimes_{v}\varphi_{v}\otimes\varPhi_{v} any non-zero factorable global vector in πΠ\pi\otimes\Pi, one has

(3.3) |[G]φ(g)Φ(g)𝑑g|2φ,φΦ,Φ=CFL(π×Π,1/2)L(π,Ad,1)L(Π,Ad,1)v|PrΠvπv(Φv)(φv)|2φv,φvΦv,Φv,\frac{\left|\int_{[{\mathrm{G}}]}\varphi(g)\varPhi(g)dg\right|^{2}}{\langle\varphi,\varphi\rangle\langle\varPhi,\varPhi\rangle}=C_{F}\frac{L(\pi\times\Pi,1/2)}{L(\pi,\mathrm{Ad},1)L(\Pi,\mathrm{Ad},1)}\prod_{v}\frac{\left|\mathrm{Pr}^{\pi_{v}}_{\Pi_{v}}(\varPhi_{v})(\varphi_{v})\right|^{2}}{\langle\varphi_{v},\varphi_{v}\rangle\langle\varPhi_{v},\varPhi_{v}\rangle},

where CF>0C_{F}>0 is an absolute constant.

Remark 3.3.

If G=GL2{\mathrm{G}}={\mathrm{GL}_{2}} and π3\pi_{3} is a principal series representation the above formula amounts to the Rankin–Selberg theory, and PrΠvπv(Φv)(φv)\mathrm{Pr}^{\pi_{v}}_{\Pi_{v}}(\varPhi_{v})(\varphi_{v}) becomes the standard local Rankin–Selberg integral.

Assuming that φ,Φ\varphi,\varPhi are both globally and locally L2L^{2}-normalized and using §2.2.2 we rewrite this formula in the following less precise form

(3.4) |[G]φ(g)Φ(g)𝑑g|2=C(π×Π)o(1)L(π×Π,1/2)v|PrΠvπv(Φv)(φv)|2.{\left|\int_{[{\mathrm{G}}]}\varphi(g)\varPhi(g)dg\right|^{2}}=C(\pi\times\Pi)^{o(1)}{L(\pi\times\Pi,1/2)}\prod_{v}{\left|\mathrm{Pr}^{\pi_{v}}_{\Pi_{v}}(\varPhi_{v})(\varphi_{v})\right|^{2}}.

3.2. Test vectors

Following a method initiated in [Ven], we use this last formula to get our hands on the size of the central value L(π×Π,1/2)L(\pi\times\Pi,1/2).

By (2.2) we rewrite (3.4) in the following form

(3.5) L(π×Π,1/2)C(π×Π)1/4+ϵ=|[G]φ(g)Φ(g)𝑑g|2v|PrΠvπv(Φv)(φv)|2Qv1/4.\frac{L(\pi\times\Pi,1/2)}{C(\pi\times\Pi)^{1/4+\epsilon}}=\frac{\left|\int_{[{\mathrm{G}}]}\varphi(g)\varPhi(g)dg\right|^{2}}{\prod_{v}\left|\mathrm{Pr}^{\pi_{v}}_{\Pi_{v}}(\varPhi_{v})(\varphi_{v})\right|^{2}Q_{v}^{1/4}}.

From this we see that in order to solve the subconvexity problem it will be sufficient to show the existence of factorable vectors φ\varphi, Φ\varPhi for which the global period is small, that is,

|[G]φ(g)Φ(g)𝑑g|2Qδ\left|\int_{[{\mathrm{G}}]}\varphi(g)\varPhi(g)dg\right|^{2}\ll Q^{-\delta}

for some δ>0\delta>0 and for which the product of local periods

(3.6) v|PrΠvπv(Φv)(φv)|2Qv1/4Qϵ\prod_{v}\left|\mathrm{Pr}^{\pi_{v}}_{\Pi_{v}}(\varPhi_{v})(\varphi_{v})\right|^{2}Q_{v}^{1/4}\gg Q^{-\epsilon}

is not too small.

Remark 3.4.

Observe that this product is converging: when vv is finite, πv\pi_{v} and Πv\Pi_{v} are unramified principal series and φv,Φv\varphi^{\circ}_{v},\ \varPhi^{\circ}_{v} are the spherical vectors, we have

|PrΠvπv(Φv)(φv)|2=Qv1/4=1.\left|\mathrm{Pr}^{\pi_{v}}_{\Pi_{v}}(\varPhi^{\circ}_{v})(\varphi^{\circ}_{v})\right|^{2}=Q_{v}^{1/4}=1.

We will search for vectors of the following form:

Definition 3.5.

For i=1,2,3i=1,2,3 a test vector for πiBvπi,vB\pi^{\mathrm{B}}_{i}\simeq\bigotimes_{v}\pi^{\mathrm{B}}_{i,v} is a factorable vector φi=vφi,v\varphi_{i}=\otimes_{v}\varphi_{i,v} such that

  1. (1)

    For all vv, φi,v=1\|\varphi_{i,v}\|=1.

  2. (2)

    For all d𝐍d\in{\bf{N}}, 𝒮dπv(φi,v)dQvA(d){\mathcal{S}}_{d}^{\pi_{v}}(\varphi_{i,v})\ll_{d}Q_{v}^{A(d)} for some AA depending on dd only.

  3. (3)

    For vi=1,2,3ram(πiB)v\not\in\bigcup_{i=1,2,3}\mathrm{ram}(\pi^{\mathrm{B}}_{i}), φi,v=φi,v\varphi_{i,v}=\varphi^{\circ}_{i,v} is a spherical element.

Here 𝒮dπv{\mathcal{S}}_{d}^{\pi_{v}} is the Sobolev norm defined as in, for example, [MVIHES, Section 2.3.5]. A test vector of Π=π2Bπ3B\Pi=\pi_{2}^{\mathrm{B}}\otimes\pi_{3}^{\mathrm{B}} is a tensor product of test vectors Φ=φ2φ3\varPhi=\varphi_{2}\otimes\varphi_{3}.

The control of local periods from below is essentially a local problem (finding for each ramified place vv a quantitatively "good" vector for the intertwiner PrΠvπv\mathrm{Pr}^{\pi_{v}}_{\Pi_{v}}), but the explicit choice of test vectors is also related to the global upper bound. We make the following definition:

Definition 3.6.

Given φi,i=1,2,3\varphi_{i},\ i=1,2,3 test vectors of πiB\pi_{i}^{\mathrm{B}}, the feasibility index of Φ=φ2φ3\varPhi=\varphi_{2}\otimes\varphi_{3} with respect to φ=φ1\varphi=\varphi_{1} at the place vv is defined as

Feasv(Φ|φ)=|PrΠvπv(Φv)(φv)|2Qv1/4,\mathrm{Feas}_{v}(\varPhi|\varphi)=\left|\mathrm{Pr}_{\Pi_{v}}^{\pi_{v}}(\varPhi_{v})(\varphi_{v})\right|^{2}Q_{v}^{1/4},

measuring if (3.6) can be achieved for proper L2L^{2}-normalized local test vectors; Define the total feasibility index to be the product

Feas(Φ|φ)=vFeasv(Φ|φ).\mathrm{Feas}(\varPhi|\varphi)=\prod_{v}\mathrm{Feas}_{v}(\varPhi|\varphi).

Finally the feasibility index of Φ\varPhi with respect to the representation π1B\pi_{1}^{\mathrm{B}} is

Feas(Φ|π):=supφFeas(Φ|φ)\mathrm{Feas}(\varPhi|\pi):=\sup_{\varphi}\mathrm{Feas}(\varPhi|\varphi)

where the supremum is taken over test vectors in π1B\pi_{1}^{\mathrm{B}}.

3.3. The amplification method

To give global upper bound we proceed as in [MVIHES] by first convolving φ\varphi by some amplifier σ\sigma with the following properties:

Lemma 3.7.

Let L1L\geq 1 be some parameter (to be chosen for optimization) , there exist a complex-valued measure σ\sigma on G(𝔸){{\mathrm{G}}({\mathbb{A}})} satisfying

  1. (1)

    σ\sigma is compactly supported on G(𝔸SL){\mathrm{G}}({\mathbb{A}}_{S_{L}}) where SLS_{L} is a set of finite places of norm L\leq L at which φ\varphi, Φ\varPhi and ψ\psi are all unramified;

  2. (2)

    For any usupp(σ)u\in\mathrm{supp}(\sigma) we have uL\|u\|\leq L where u\|u\| is the operator norm as in (2.1).

  3. (3)

    Let |σ||\sigma| denote the total variation measure. Then the total mass of |σ||\sigma| is bounded above by LBL^{B}, for some absolute constant BB. Moreover, with |σ|(2)=|σ||σˇ||\sigma|^{(2)}=|\sigma|\star|\check{\sigma}|, one has for any γ>1/2\gamma>1/2

    uγd|σ|(2)(u)Lη,for some η=η(γ)>0.\int\|u\|^{-\gamma}d|\sigma|^{(2)}(u)\leq L^{-\eta},\ \hbox{for some }\eta=\eta(\gamma)>0.
  4. (4)

    φσ=λφ\varphi\star\sigma=\lambda\varphi with λF,εQε\lambda\gg_{F,\varepsilon}Q^{-\varepsilon} for any ε>0\varepsilon>0.

Since at least one of π2\pi_{2} and π3\pi_{3} is cuspidal, Φ\varPhi is rapidly decreasing. As φ\varphi is L2L^{2}-normalized, the Cauchy–Schwarz inequality implies that:

(3.7) |λ|2|[G]φ(h)Φ(h)𝑑h|2=\displaystyle|\lambda|^{2}\left|\int_{[{\mathrm{G}}]}\varphi(h)\varPhi(h)dh\right|^{2}= |[G](φσ)(g)Φ(g)𝑑g|2\displaystyle\left|\int_{[{\mathrm{G}}]}(\varphi\star\sigma)(g)\varPhi(g)dg\right|^{2}
\displaystyle\leq [G]|Φσˇ(g)|2𝑑g=(φ2φ3)(σ¯σˇ),φ2φ3.\displaystyle\int_{[{\mathrm{G}}]}\left|\varPhi\star\check{\sigma}(g)\right|^{2}dg=\langle(\varphi_{2}\varphi_{3})\star(\overline{\sigma}\star\check{\sigma}),\varphi_{2}\varphi_{3}\rangle.

The later inner product decomposes into a sum of terms of the shape

u.φ2φ2¯,φ3u.φ3¯=Φ2,u|G,Φ3,u|G\langle u.\varphi_{2}\ \overline{\varphi_{2}},\varphi_{3}\ u.\overline{\varphi_{3}}\rangle=\langle\varPhi_{2,u}|_{\mathrm{G}},\varPhi_{3,u}|_{\mathrm{G}}\rangle

where usupp(σ¯σˇ)u\in\mathrm{supp}(\overline{\sigma}\star\check{\sigma}), Φ2,u,Φ3,u\varPhi_{2,u},\varPhi_{3,u} are the functions on G(𝔸)×G(𝔸){{\mathrm{G}}({\mathbb{A}})}\times{{\mathrm{G}}({\mathbb{A}})} given by

(3.8) Φ2,u(g1,g2)=u.φ2(g1)×φ¯2(g2),Φ3,u(g1,g2)=φ3(g1)×u.φ¯3(g3),\varPhi_{2,u}(g_{1},g_{2})=u.\varphi_{2}(g_{1})\times\overline{\varphi}_{2}(g_{2}),\varPhi_{3,u}(g_{1},g_{2})=\varphi_{3}(g_{1})\times u.\overline{\varphi}_{3}(g_{3}),

and |G|_{\mathrm{G}} denote the restriction to G(𝔸)G(𝔸)×G(𝔸){{\mathrm{G}}({\mathbb{A}})}\hookrightarrow{{\mathrm{G}}({\mathbb{A}})}\times{{\mathrm{G}}({\mathbb{A}})} (diagonally embedded). The function Φ2,u\varPhi_{2,u} and Φ3,u\varPhi_{3,u} are factorable vectors in the G(𝔸)2{{\mathrm{G}}({\mathbb{A}})}^{2}-automorphic representations

Π2=π2π~2,Π3=π3π~3.\Pi_{2}=\pi_{2}\otimes\tilde{\pi}_{2},\ \Pi_{3}=\pi_{3}\otimes\tilde{\pi}_{3}.

To simplify notations we will omit the restriction |G|_{\mathrm{G}} of Φ2\varPhi_{2} and Φ3\varPhi_{3} and write Φ2,u,Φ3,u\langle\varPhi_{2,u},\varPhi_{3,u}\rangle in place of Φ2,u|G,Φ3,u|G\langle\varPhi_{2,u}|_{\mathrm{G}},\varPhi_{3,u}|_{\mathrm{G}}\rangle. Thus our discussion so far gives

(3.9) |λ|2|[G]φ(h)Φ(h)𝑑h|2uΦ2,u,Φ3,ud(σ¯σˇ)(u)\displaystyle|\lambda|^{2}\left|\int_{[{\mathrm{G}}]}\varphi(h)\varPhi(h)dh\right|^{2}\leq\int\limits_{u}\langle\varPhi_{2,u},\varPhi_{3,u}\rangle d(\overline{\sigma}\star\check{\sigma})(u)

which is essentially a finite sum in uu.

3.4. Applying the Plancherel formula

Suppose first that π3\pi_{3} is a cuspidal automorphic representation, then the restrictions to G(𝔸){{\mathrm{G}}({\mathbb{A}})} of Φ2\varPhi_{2} and Φ3\varPhi_{3} are rapidly decreasing modulo 𝔸×G(F){\mathbb{A}}^{\times}{{\mathrm{G}}(F)} and by the Plancherel formula the above inner product decomposes as

Φ2,u,Φ3,u=\displaystyle\langle\varPhi_{2,u},\varPhi_{3,u}\rangle= π𝒜(G,1)ψ(π)Φ2,u,ψψ,Φ3,udμPl(π)\displaystyle\int_{\pi^{\prime}\in\mathcal{A}({\mathrm{G}},1)}\sum_{\psi\in\mathcal{B}(\pi^{\prime})}\langle\varPhi_{2,u},\psi\rangle\langle\psi,\varPhi_{3,u}\rangle\ d\mu_{\mathrm{Pl}}(\pi^{\prime})
=\displaystyle= π𝒜(G,1)ψ(π)PrΠ2π(Φ2,u)(ψ)PrΠ3π(Φ3,u)(ψ)¯dμPl(π)\displaystyle\int_{\pi^{\prime}\in\mathcal{A}({\mathrm{G}},1)}\sum_{\psi\in\mathcal{B}(\pi^{\prime})}\mathrm{Pr}^{\pi^{\prime}}_{\Pi_{2}}(\varPhi_{2,u})(\psi)\overline{\mathrm{Pr}_{\Pi_{3}}^{\pi^{\prime}}(\varPhi_{3,u})(\psi)}\ d\mu_{\mathrm{Pl}}(\pi^{\prime})
=\displaystyle= π𝒜(G,1)PrΠ2π(Φ2,u),PrΠ3π(Φ3,u)𝑑μPl(π)\displaystyle\int_{\pi^{\prime}\in\mathcal{A}({\mathrm{G}},1)}\langle\mathrm{Pr}^{\pi^{\prime}}_{\Pi_{2}}(\varPhi_{2,u}),{\mathrm{Pr}_{\Pi_{3}}^{\pi^{\prime}}(\varPhi_{3,u})}\rangle d\mu_{\mathrm{Pl}}(\pi^{\prime})

where π\pi^{\prime} runs over the automorphic representations of B×{{\mathrm{B}}^{\times}} with trivial central character, (π)\mathcal{B}(\pi^{\prime}) is an orthonormal basis of the space of π\pi^{\prime}, and PrΠiπ,i=1,2\mathrm{Pr}^{\pi^{\prime}}_{\Pi_{i}},\ i=1,2 is defined as in (3.2).

Restricting the above integral to the finite and the generic spectrum we write

Φ2,u,Φ3,u=Φ2,u,Φ3,ufinite+Φ2,u,Φ3,ugeneric.\langle\varPhi_{2,u},\varPhi_{3,u}\rangle=\langle\varPhi_{2,u},\varPhi_{3,u}\rangle_{\mathrm{finite}}+\langle\varPhi_{2,u},\varPhi_{3,u}\rangle_{\mathrm{generic}}.

The contribution of the finite spectrum (characters of order at most 22) equals a finite sum of products of matrix coefficients:

Φ2,u,Φ3,ufinite=χ2=1u.φ2,χφ2u.φ3,χφ3Qϵuγ,\langle\varPhi_{2,u},\varPhi_{3,u}\rangle_{\mathrm{finite}}=\sum_{\chi^{2}=1}\langle u.\varphi_{2},\chi\varphi_{2}\rangle\langle u.\varphi_{3},\chi\varphi_{3}\rangle\ll Q^{\epsilon}\|u\|^{-\gamma},

where γ=12θ\gamma=1-2\theta with θ<7/64\theta<7/64 ([KiSa, BlBr]). Indeed the above sum comprises at most two non-zero terms (the trivial character and the unique quadratic character (if any) such that π2χπ2\pi_{2}\simeq\chi\pi_{2} and π3χπ3\pi_{3}\simeq\chi\pi_{3}); The bound then follows from bounds for spherical matrix coefficients, the support for uu, and the bound towards Ramanujan-Petersson conjecture [KiSa, BlBr].

Using this bound together with item (3) of Lemma 3.7 we obtain that

(3.10) |σ¯||σˇ|(u|Φ2,u,Φ3,ufinite|)QϵLδ|\overline{\sigma}|\star|\check{\sigma}|(u\mapsto|\langle\varPhi_{2,u},\varPhi_{3,u}\rangle_{\mathrm{finite}}|)\ll Q^{\epsilon}L^{-\delta}

for some absolute δ>0\delta>0.

3.5. The generic contribution

It remains to bound the generic spectrum contribution

(3.11) Φ2,u,Φ3,ugeneric=πgenericPrΠ2π(Φ2,u),PrΠ3π(Φ3,u)𝑑μPl(π).\langle\varPhi_{2,u},\varPhi_{3,u}\rangle_{\mathrm{generic}}=\int_{\pi^{\prime}\ \mathrm{generic}}\langle\mathrm{Pr}^{\pi^{\prime}}_{\Pi_{2}}(\varPhi_{2,u}),{\mathrm{Pr}_{\Pi_{3}}^{\pi^{\prime}}(\varPhi_{3,u})}\rangle d\mu_{\mathrm{Pl}}(\pi^{\prime}).

In this paper we will not try to exploit cancellations between the terms of the π\pi^{\prime}-integral and will rather focus on controlling the length of spectral sum and individual terms.

3.5.1. Triple product L-functions and Quantum unique ergodicity

We first look at the decay properties of the inner product

PrΠ2π(Φ2,u),PrΠ3π(Φ3,u)=ψ(π)PrΠ2π(Φ2,u)(ψ)PrΠ3π(Φ3,u)(ψ)¯.\langle\mathrm{Pr}^{\pi^{\prime}}_{\Pi_{2}}(\varPhi_{2,u}),{\mathrm{Pr}_{\Pi_{3}}^{\pi^{\prime}}(\varPhi_{3,u})}\rangle=\sum_{\psi\in\mathcal{B}(\pi^{\prime})}\mathrm{Pr}^{\pi^{\prime}}_{\Pi_{2}}(\varPhi_{2,u})(\psi)\overline{\mathrm{Pr}_{\Pi_{3}}^{\pi^{\prime}}(\varPhi_{3,u})(\psi)}.

Consider the case u=1u=1, then

PrΠiπ(Φi,u)(ψ)=[G]ψ(g)|φi(g)|2𝑑g=μ|φi|2(ψ)\mathrm{Pr}^{\pi^{\prime}}_{\Pi_{i}}(\varPhi_{i,u})(\psi)=\int_{[{\mathrm{G}}]}\psi(g)|\varphi_{i}(g)|^{2}dg=\mu_{|\varphi_{i}|^{2}}(\psi)

where μ|φi|2\mu_{|\varphi_{i}|^{2}} is the probability measure with density |φi(g)|2dg|\varphi_{i}(g)|^{2}dg. The convergence of this measure to the standard measure for varying φi\varphi_{i} is interpreted as an equidistribution property of the "mass" of φi\varphi_{i} on (quotients) [G][{\mathrm{G}}]: in other terms a form of Quantum Unique Ergodicity (QUE) in the sense of Rudnick-Sarnak [RS]. QUE has now been established in a number of contexts by different methods [LuoSar, Sarnak, Lin, HolLin, ANP, YH18]. Our present problem is a slight variant of the above. The comment justifies the choice of notations in Definition 3.8 below.

By (3.5) and bounding terms individually, we obtain that

(3.12) |PrΠ2π(Φ2,u),PrΠ3π(Φ3,u)|ψ(π)|PrΠ2π(Φ2,u)(ψ)||PrΠ3π(Φ3,u)(ψ)|\displaystyle\left|\langle\mathrm{Pr}^{\pi^{\prime}}_{\Pi_{2}}(\varPhi_{2,u}),{\mathrm{Pr}_{\Pi_{3}}^{\pi^{\prime}}(\varPhi_{3,u})}\rangle\right|\leq\sum\limits_{\psi\in\mathcal{B}(\pi^{\prime})}\left|\mathrm{Pr}^{\pi^{\prime}}_{\Pi_{2}}(\varPhi_{2,u})(\psi)\right|\left|\mathrm{Pr}^{\pi^{\prime}}_{\Pi_{3}}(\varPhi_{3,u})(\psi)\right|
=\displaystyle= (C(π)Q)o(1)ψ(π)L(π×Π2,1/2)1/2C(π×Π2)1/8v|PrΠ2π(Φ2,u,v)(ψv)|Cv(π×Π2)1/8\displaystyle(C(\pi^{\prime})Q)^{o(1)}\sum\limits_{\psi\in\mathcal{B}(\pi^{\prime})}\frac{L(\pi^{\prime}\times\Pi_{2},1/2)^{1/2}}{C(\pi^{\prime}\times\Pi_{2})^{1/8}}\prod_{v}{\left|\mathrm{Pr}^{\pi^{\prime}}_{\Pi_{2}}(\varPhi_{2,u,v})(\psi_{v})\right|}{C_{v}(\pi^{\prime}\times\Pi_{2})^{1/8}}
×L(π×Π3,1/2)1/2C(π×Π3)1/8v|PrΠ3π(Φ3,u,v)(ψv)|Cv(π×Π3)1/8\displaystyle\times\frac{L(\pi^{\prime}\times\Pi_{3},1/2)^{1/2}}{C(\pi^{\prime}\times\Pi_{3})^{1/8}}\prod_{v}{\left|\mathrm{Pr}^{\pi^{\prime}}_{\Pi_{3}}(\varPhi_{3,u,v})(\psi_{v})\right|}{C_{v}(\pi^{\prime}\times\Pi_{3})^{1/8}}
=\displaystyle= (C(π)Q)o(1)(L(π×Π2,1/2)L(π×Π3,1/2)C(π×Π2)1/4C(π×Π3)1/4)1/2\displaystyle(C(\pi^{\prime})Q)^{o(1)}\left(\frac{L(\pi^{\prime}\times\Pi_{2},1/2)L(\pi^{\prime}\times\Pi_{3},1/2)}{C(\pi^{\prime}\times\Pi_{2})^{1/4}C(\pi^{\prime}\times\Pi_{3})^{1/4}}\right)^{1/2}
×ψ(π)v|PrΠ2π(Φ2,u,v)(ψv)||PrΠ3π(Φ3,u,v)(ψv)|(Cv(π×Π2)Cv(π×Π3))1/8.\displaystyle\times\sum\limits_{\psi\in\mathcal{B}(\pi^{\prime})}\prod_{v}\left|\mathrm{Pr}^{\pi^{\prime}}_{\Pi_{2}}(\varPhi_{2,u,v})(\psi_{v})\right|\left|\mathrm{Pr}^{\pi^{\prime}}_{\Pi_{3}}(\varPhi_{3,u,v})(\psi_{v})\right|\left(C_{v}(\pi^{\prime}\times\Pi_{2})C_{v}(\pi^{\prime}\times\Pi_{3})\right)^{1/8}.
Definition 3.8.
  1. -

    We call the quantity

    E(Π|π)=L(π×Π,s)C(π×Π)1/4\mathrm{E}(\Pi|\pi^{\prime})=\frac{L(\pi^{\prime}\times\Pi,s)}{C(\pi^{\prime}\times\Pi)^{1/4}}

    the global equidistribution index of Π\Pi with respect to π\pi^{\prime}.

  2. -

    Given some test vector ΦΠ\varPhi\in\Pi we define the local equidistribution index

    ev(Φ|ψ)=|PrΠvπv(Φv)(ψv)|2Cv(π×Π)1/4.\mathrm{e}_{v}(\varPhi|\psi)=\left|\mathrm{Pr}^{\pi^{\prime}_{v}}_{\Pi_{v}}(\varPhi_{v})(\psi_{v})\right|^{2}C_{v}(\pi^{\prime}\times\Pi)^{1/4}.
  3. -

    For any S𝒱S\subset\mathcal{V} we define

    eS(Φ|ψ):=vSev(Φ|ψ).\mathrm{e}_{S}(\varPhi|\psi):=\prod_{v\in S}\mathrm{e}_{v}(\varPhi|\psi).
  4. -

    For Π23=Π2Π3\Pi_{23}=\Pi_{2}\boxplus\Pi_{3}, define L(π×Π23,s)=L(π×Π2,s)L(π×Π3,s)L(\pi^{\prime}\times\Pi_{23},s)=L(\pi^{\prime}\times\Pi_{2},s)L(\pi^{\prime}\times\Pi_{3},s) and similarly C(π×Π23)\ C(\pi^{\prime}\times\Pi_{23}), E(Π23|π)\mathrm{E}(\Pi_{23}|\pi^{\prime}). For test vector Φ23=Φ2Φ3Π23,\varPhi_{23}=\varPhi_{2}\boxplus\varPhi_{3}\in\Pi_{23}, denote eS(Φ23|ψ)=eS(Φ2|ψ)eS(Φ3|ψ)e_{S}(\varPhi_{23}|\psi)=e_{S}(\varPhi_{2}|\psi)e_{S}(\varPhi_{3}|\psi).

Remark 3.9.

At places vv where Φ\varPhi and ψ\psi are spherical, one has

ev(Φ|ψ)=1.\mathrm{e}_{v}(\varPhi|\psi)=1.

Thus the product eS(Φ|ψ)\mathrm{e}_{S}(\varPhi|\psi) is always finite.

Now take the vector

Φ23,u:=Φ2,uΦ3,uΠ23\varPhi_{23,u}:=\varPhi_{2,u}\boxplus\varPhi_{3,u}\in\Pi_{23}

(cf. (3.8)) for usupp(σ¯σˇ)u\in\mathrm{supp}(\overline{\sigma}\star\check{\sigma}). Using the above notations, we can rewrite (3.12) as

(3.13) |PrΠ2π(Φ2,u),PrΠ3π(Φ3,u)|(C(π)Q)o(1)E(Π23|π)1/2ψ(π)e𝒱(Φ23,u|ψ)1/2,\left|\langle\mathrm{Pr}^{\pi^{\prime}}_{\Pi_{2}}(\varPhi_{2,u}),{\mathrm{Pr}_{\Pi_{3}}^{\pi^{\prime}}(\varPhi_{3,u})}\rangle\right|\leq(C(\pi^{\prime})Q)^{o(1)}\mathrm{E}(\Pi_{23}|\pi^{\prime})^{1/2}\sum\limits_{\psi\in\mathcal{B}(\pi^{\prime})}\mathrm{e}_{\mathcal{V}}(\varPhi_{23,u}|\psi)^{1/2},

and thus

(3.14) |Φ2,u,Φ3,ugeneric|\displaystyle\left|\langle\varPhi_{2,u},\varPhi_{3,u}\rangle_{\mathrm{generic}}\right|
\displaystyle\leq (C(π)Q)o(1)πgenericE(Π23|π)1/2ψ(π)e𝒱(Φ23,u|ψ)1/2dμPl(π).\displaystyle(C(\pi^{\prime})Q)^{o(1)}\int_{\pi^{\prime}\ \mathrm{generic}}\mathrm{E}(\Pi_{23}|\pi^{\prime})^{1/2}\sum\limits_{\psi\in\mathcal{B}(\pi^{\prime})}\mathrm{e}_{\mathcal{V}}(\varPhi_{23,u}|\psi)^{1/2}d\mu_{\mathrm{Pl}}(\pi^{\prime}).

3.5.2. The global equidistribution index

By the convexity bound for L-functions, we have

(3.15) E(Π23|π)ϵC(π×Π23)ϵϵQϵ.\mathrm{E}(\Pi_{23}|\pi^{\prime})\ll_{\epsilon}C(\pi^{\prime}\times\Pi_{23})^{\epsilon}\ll_{\epsilon}Q^{\epsilon}.

The Generalized Lindelöf hypothesis predicts that it is much smaller, controlled by C(π×Π23)1/4+o(1)C(\pi^{\prime}\times\Pi_{23})^{-1/4+o(1)}. The resolution of the subconvexity problem for L(π×Π2,1/2)L(π×Π3,1/2)L(\pi^{\prime}\times\Pi_{2},1/2)L(\pi^{\prime}\times\Pi_{3},1/2) is equivalent to

E(Π23|π)C(π×Π23)δ\mathrm{E}(\Pi_{23}|\pi^{\prime})\ll C(\pi^{\prime}\times\Pi_{23})^{-\delta}

for some absolute constant δ>0\delta>0. In some special cases this bound is known (for example when π2\pi_{2} or π3\pi_{3} is either an Eisenstein series representation or a dihedral representation.) Our result uses only (3.15) and does not rely on this saving, but can be greatly improved if it is known.

3.5.3. The generic spectral length

For a factorable automorphic form φ\varphi, let ram(φ)\mathrm{ram}(\varphi) denote the set of finite places vv where φv\varphi_{v} is not spherical.

For each place vv, let

Mv=min{C(π2,v),C(π3,v)},M_{v}=\min\{C(\pi_{2,v}),C(\pi_{3,v})\},
M=vMv,Mf=vMv.M=\prod\limits_{v}M_{v},\ M_{f}=\prod\limits_{v\nmid\infty}M_{v}.
Assumption 3.10.

For each v|Mfv|M_{f}\infty, let jv=2j_{v}=2 or 33 so that C(πjv,v)=MvC(\pi_{j_{v},v})=M_{v}. We assume the following

  1. (1)

    For v|Mfv|M_{f}, φjv,v\varphi_{j_{v},v} is K(Mv)K(M^{\prime}_{v})-invariant where

    Mv=MvAM^{\prime}_{v}=M_{v}^{A}

    for some absolute constant AA.

  2. (2)

    For v|v|\infty, we assume that 𝒮dπjv,v(φjv,v)MvA(d){\mathcal{S}}^{\pi_{j_{v},v}}_{d}\left(\varphi_{j_{v},v}\right)\ll M_{v}^{A(d)} for some constants A(d)A(d) depending only on dd.

Note that this assumption will be incorporated into Conjecture 3.17 later on.

To simplify the notations, we shall omit subscript vv from jj from now on, though it can change for different vv.

Lemma 3.11.

Let M=v finite(uvMv)AM^{\prime}=\prod\limits_{v\text{\ finite}}(\|u\|_{v}M_{v})^{A} for some absolute constant AA. With Assumption 3.10 above, we have

(3.16) |Φ2,u,Φ3,ugeneric|\displaystyle\left|\langle\varPhi_{2,u},\varPhi_{3,u}\rangle_{\mathrm{generic}}\right|
ϵ\displaystyle\ll_{\epsilon} (LQ)ϵπgenericE(Π23|π)1/2ψ((π)K(M))e𝒱(Φ23,u|ψ)1/2dμPl(π)\displaystyle(LQ)^{\epsilon}\int_{\pi^{\prime}\ \mathrm{generic}}\mathrm{E}(\Pi_{23}|\pi^{\prime})^{1/2}\sum\limits_{\psi\in\mathcal{B}((\pi^{\prime})^{K(M^{\prime})})}\mathrm{e}_{\mathcal{V}}(\varPhi_{23,u}|\psi)^{1/2}d\mu_{\mathrm{Pl}}(\pi^{\prime})
ϵ\displaystyle\ll_{\epsilon} Qϵ(LM)Asupπ,ψE(Π23|π)1/2e𝒱(Φ23,u|ψ)1/2,\displaystyle Q^{\epsilon}(LM)^{A}\sup\limits_{\pi^{\prime},\psi}\mathrm{E}(\Pi_{23}|\pi^{\prime})^{1/2}\mathrm{e}_{\mathcal{V}}(\varPhi_{23,u}|\psi)^{1/2},

where the integral/supremum in π\pi^{\prime} is for those π\pi^{\prime} with C(πv)C(\pi^{\prime}_{v}) bounded by (uvMv)A(\|u\|_{v}M_{v})^{A} for all vv; The sum/supremum in ψ\psi is over orthonormal elements in π\pi^{\prime} such that ψv\psi_{v} at finite places is K(Mv)K(M_{v}^{\prime})-invariant, and ψv\psi_{v} at archimedean places satisfies 𝒮dπv(ψv)MvA(d){\mathcal{S}}^{\pi_{v}^{\prime}}_{d}\left(\psi_{v}\right)\ll M_{v}^{A(d)}.

Proof.

At archimedean places, the local period integral PrΠjπ(Φj,v)(ψv)\mathrm{Pr}^{\pi^{\prime}}_{\Pi_{j}}(\varPhi_{j,v})(\psi_{v}) is rapidly decaying by [Ne19, Lemma 6.9, Section 6.11] with respect to the Sobolev norm of ψv\psi_{v}. Thus up to a factor of size QϵMAQ^{\epsilon}M^{A}, one can restrict the summation in ψ\psi to those with 𝒮dπv(ψv)MvA(d){\mathcal{S}}^{\pi^{\prime}_{v}}_{d}(\psi_{v})\ll M_{v}^{A(d)}, and the integral in π\pi^{\prime} to those with analytic conductor at vv bounded by MvAM_{v}^{A} according to [Ne19, Lemma 6.6]. The dimension of such ψπ\psi\in\pi^{\prime} are controlled by v|MvA\prod\limits_{v|\infty}M_{v}^{A} according to [MVIHES, Lemma 2.6.3].

At non-archimedean places, the restriction on ramification comes from that PrΠjπ(Φj,u,v)(ψv)\mathrm{Pr}^{\pi^{\prime}}_{\Pi_{j}}(\varPhi_{j,u,v})(\psi_{v}) would be vanishing when Φj,u,v\varPhi_{j,u,v} is K(Mv)K(M^{\prime}_{v})-invariant while ψv\psi_{v} is in the orthogonal complement of (πv)K(Mv)(\pi_{v}^{\prime})^{K(M^{\prime}_{v})}, the subspace of K(Mv)K(M^{\prime}_{v})-invariant elements.

The second line of (3.16) follows from the Weyl law with the bounded local conductors for π\pi^{\prime} and ψ\psi. ∎

Example 3.12.

Consider the simplest case where φ3\varphi_{3} is spherical. Then one only has to consider those π\pi^{\prime}/ψ\psi in Lemma 3.11 whose ramifications at finite places happen only at vv where uv1u_{v}\neq 1, and are controlled there by uvA\|u_{v}\|^{A}.

3.5.4. The local equidistribution index

Take S=ram(Φ23,u)S=\mathrm{ram}(\varPhi_{23,u}). For relevant π\pi^{\prime} and ψ\psi as in Lemma 3.11, we have

e𝒱(Φ23,u|ψ)=eS(Φ23,u|ψ).\mathrm{e}_{\mathcal{V}}(\varPhi_{23,u}|\psi)=\mathrm{e}_{S}(\varPhi_{23,u}|\psi).

One has the relation ram(Φ23,u)=ram(Φ23){v,uv1}\mathrm{ram}(\varPhi_{23,u})=\mathrm{ram}(\varPhi_{23})\sqcup\{v,\ u_{v}\not=1\}. The study of ev(Φ23,u|ψ)e_{v}(\varPhi_{23,u}|\psi) at vram(Φ23)v\in\mathrm{ram}(\varPhi_{23}) will be left to later sections. Here we control the local equidistribution index at vv where uv1u_{v}\neq 1.

Lemma 3.13.

Let vv be such that uv1u_{v}\not=1, then for ψ\psi as in Lemma 3.11,

ev(Φ23,u|ψ)Cv(π)2uvA\mathrm{e}_{v}(\varPhi_{23,u}|\psi)\ll C_{v}(\pi^{\prime})^{2}\|u_{v}\|^{A^{\prime}}

where the implied constant and AA^{\prime} are absolute.

Proof.

Since uv1u_{v}\not=1, Π2\Pi_{2} and Π3\Pi_{3} are unramified and

Cv(π×Π2)Cv(π×Π3)=Cv(π)8.C_{v}(\pi^{\prime}\times\Pi_{2})C_{v}(\pi^{\prime}\times\Pi_{3})=C_{v}(\pi^{\prime})^{8}.

Recall that

Φ2,u(g1,g2)=u.φ2(g1)×φ¯2(g2),Φ3,u(g1,g2)=φ3(g1)×u.φ¯3(g2).\varPhi_{2,u}(g_{1},g_{2})=u.\varphi_{2}(g_{1})\times\overline{\varphi}_{2}(g_{2}),\varPhi_{3,u}(g_{1},g_{2})=\varphi_{3}(g_{1})\times u.\overline{\varphi}_{3}(g_{2}).

It then follows from [MVIHES, Lemma 3.5.2] and the distortion property for Sobolev norms [MVIHES, S1b] that for some absolute constants AA and AA^{\prime}

PrΠ2π(Φ2,u,v)×PrΠ3π(Φ3,u,v)𝒮Aπ2,v(φ2)𝒮Aπ2,v(uv.φ2)𝒮Aπ3,v(φ3)𝒮Aπ3,v(uv.φ3)uvA.\|{\mathrm{Pr}^{\pi^{\prime}}_{\Pi_{2}}(\varPhi_{2,u,v})\|\times\|\mathrm{Pr}^{\pi^{\prime}}_{\Pi_{3}}(\varPhi_{3,u,v})\|}\ll{\mathcal{S}}^{\pi_{2,v}}_{A}(\varphi_{2}^{\circ}){\mathcal{S}}^{\pi_{2,v}}_{A}(u_{v}.\varphi_{2}^{\circ}){\mathcal{S}}^{\pi_{3,v}}_{A}(\varphi_{3}^{\circ}){\mathcal{S}}^{\pi_{3,v}}_{A}(u_{v}.\varphi_{3}^{\circ})\ll\|u_{v}\|^{A^{\prime}}.

The lemma follows by dropping all but one term. ∎

From this bound and Lemma 3.11 on the restriction for π\pi^{\prime}, we deduce that

(3.17) e𝒱(Φ23,u|ψ)uAeram(Φ23)(Φ23|ψ)\mathrm{e}_{\mathcal{V}}(\varPhi_{23,u}|\psi)\ll\|u\|^{A}\mathrm{e}_{\mathrm{ram}(\varPhi_{23})}(\varPhi_{23}|\psi)

for some absolute constant AA.

3.6. The main global bound

By combining (3.5), (3.7), Lemma 3.7(4), (3.10), Lemma 3.11 and (3.17) we obtain after integration over uu that there exist absolute constant δ,A>0\delta,A>0 such that

L(π×π2×π3,1/2)\displaystyle L(\pi\times\pi_{2}\times\pi_{3},1/2)
ϵ\displaystyle\ll_{\epsilon} C(π1×π2×π2)1/4+ϵFeas(φ2φ3|π)(Lδ+(LM)Asupπ,ψE(Π23|π)1/2eram(Φ23)(Φ23|ψ)1/2).\displaystyle\frac{C(\pi_{1}\times\pi_{2}\times\pi_{2})^{1/4+\epsilon}}{\mathrm{Feas}(\varphi_{2}\otimes\varphi_{3}|\pi)}\left(L^{-\delta}+(LM)^{A}\sup\limits_{\pi^{\prime},\psi}\mathrm{E}(\Pi_{23}|\pi^{\prime})^{1/2}\mathrm{e}_{\mathrm{ram}(\varPhi_{23})}(\varPhi_{23}|\psi)^{1/2}\right).

Here the supremum is taken over π\pi^{\prime} and ψ\psi as in Lemma 3.11.

Choosing L>1L>1 optimally if

(3.18) 𝔼(Φ23):=supπ,ψE(Π23|π)1/2eram(Φ23)(Φ23|ψ)1/21,{\mathbb{E}}(\varPhi_{23}):=\sup\limits_{\pi^{\prime},\psi}\mathrm{E}(\Pi_{23}|\pi^{\prime})^{1/2}\mathrm{e}_{\mathrm{ram}(\varPhi_{23})}(\varPhi_{23}|\psi)^{1/2}\leq 1,

and L=1L=1 otherwise, we obtain (under the assumption that π1,π2,π3\pi_{1},\pi_{2},\pi_{3} are all cuspidal) the following result:

Theorem 3.14.

Under Assumption 3.10, there exist absolute constants δ,A>0\delta,A>0 such that

L(π×π2×π3,1/2)ϵMAC(π1×π2×π2)1/4+ϵ1Feas(φ2φ3|π)min(1,𝔼(Φ23)δ).L(\pi\times\pi_{2}\times\pi_{3},1/2)\ll_{\epsilon}M^{A}C(\pi_{1}\times\pi_{2}\times\pi_{2})^{1/4+\epsilon}\frac{1}{\mathrm{Feas}(\varphi_{2}\otimes\varphi_{3}|\pi)}\min\left(1,{\mathbb{E}}(\varPhi_{23})^{\delta}\right).

3.7. The Rankin–Selberg case

In this paper we are also interested in the critical value of the Rankin–Selberg L-function L(π1×π2,1/2)L(\pi_{1}\times\pi_{2},1/2), in which case φ3\varphi_{3} is an Eisenstein series in the above formulation. In this section we show that Theorem 3.14 continue to hold in this case. We use the regularization process as developed in [MVIHES], and refer to §4.3 of that paper for the detailed definitions and notations.

We assume for simplicity that π3=1χ3\pi_{3}=1\boxplus\chi_{3} (with χ3=(χ1χ2)1\chi_{3}=(\chi_{1}\chi_{2})^{-1}, χi\chi_{i} being central character of πi\pi_{i}). We have

L(π1×π2×π3,1/2)=|L(π1×π2,1/2)|2L(\pi_{1}\times\pi_{2}\times\pi_{3},1/2)=\left|L(\pi_{1}\times\pi_{2},1/2)\right|^{2}
ε(π1×π2×π3)=|ε(π1×π2)|2=1\varepsilon(\pi_{1}\times\pi_{2}\times\pi_{3})=\left|\varepsilon(\pi_{1}\times\pi_{2})\right|^{2}=1

so that G=GL2{\mathrm{G}}={\mathrm{GL}_{2}} and the Jacquet-Langlands correspondence is just the identity.

We take φ3\varphi_{3} of the shape

φ3=Eis(f3)\varphi_{3}=\mathrm{Eis}(f_{3})

for f3:N(𝔸)B(F)\G(𝔸)f_{3}:{\mathrm{N}}({\mathbb{A}})B(F)\backslash{{\mathrm{G}}({\mathbb{A}})}\rightarrow{\mathbb{C}} an element in π(1,χ3)\pi(1,\chi_{3}). One can try to proceed exactly as in the beginning of this section up to §3.4 and the evaluation of linear combination of inner products of the shape Φ2,u,Φ3,u\langle\varPhi_{2,u},\varPhi_{3,u}\rangle. However since φ3\varphi_{3} is not rapidly decreasing we have to use the regularized version of Plancherel’s formula proven in [MVIHES, §4.3]. In order to apply this formula, we have to deform slightly the vector Φ3,u\varPhi_{3,u}: given tt\in{\mathbb{C}} let

π3(t)=||it||itχ3,φ3(t)=Eis(f3(t)),\pi_{3}(t)=|\cdot|^{it}\boxplus|\cdot|^{-it}\chi_{3},\ \varphi_{3}(t)=\mathrm{Eis}(f_{3}(t)),

where f3(t)f_{3}(t) is the section whose restriction to the maximal compact subgroup vKv\prod_{v}K_{v} coincide with f3f_{3}. Set Π3(t)=π3(t)π~3\Pi_{3}(t)=\pi_{3}(t)\otimes\tilde{\pi}_{3} and define the vector

Φ3,u(t):(g1,g2)φ3(t)(g1)×u.φ¯3(g2).\displaystyle\varPhi_{3,u}(t):(g_{1},g_{2})\mapsto\varphi_{3}(t)(g_{1})\times u.\overline{\varphi}_{3}(g_{2}).

As φ2\varphi_{2} is cuspidal, we observe that the function

tΦ2,u,Φ3,u(t)t\mapsto\langle\varPhi_{2,u},\varPhi_{3,u}(t)\rangle

is anti-holomorphic in the tt-variable near t=0t=0, and to bound the value at t=0t=0 it is sufficient to bound it uniformly on some circle 𝒞{\mathcal{C}} of fixed radius around t=0t=0.

The set of exponents of Φ3,u(t)\varPhi_{3,u}(t) is

T={||𝔸1+it,||𝔸1itχ3,||𝔸1+itχ¯3,||𝔸1it}T=\{|\cdot|^{1+it}_{\mathbb{A}},|\cdot|^{1-it}_{\mathbb{A}}\chi_{3},|\cdot|^{1+it}_{\mathbb{A}}\overline{\chi}_{3},|\cdot|^{1-it}_{\mathbb{A}}\}

and the squares of the above exponents are not ||𝔸2|\cdot|^{2}_{\mathbb{A}} for t𝒞t\in{\mathcal{C}} for almost all radius of 𝒞{\mathcal{C}}. We fix such 𝒞{\mathcal{C}} with small enough radius in the following.

Then as in the regularized version of Plancherel’s formula [MVIHES, Prop. 4.3.8] we have the following:

(3.19) Φ2,u,Φ3,u(t)\displaystyle\langle\varPhi_{2,u},\varPhi_{3,u}(t)\rangle =Φ2,u,3(t)+πgenericψ(π)Φ2,u,ψψ,Φ3,u(t)regdμPl(π),\displaystyle=\langle\varPhi_{2,u},{\mathcal{E}}_{3}(t)\rangle+\int\limits_{\pi^{\prime}\ \mathrm{generic}}\sum_{\psi\in\mathcal{B}(\pi^{\prime})}\langle\varPhi_{2,u},\psi\rangle\langle\psi,\varPhi_{3,u}(t)\rangle_{reg}\ d\mu_{\mathrm{Pl}}(\pi^{\prime}),

which we rewrite as

Φ2,u,Φ3,u(t):=Φ2,u,Φ3,u(t)deg+Φ2,u,Φ3,u(t)generic.\langle\varPhi_{2,u},\varPhi_{3,u}(t)\rangle:=\langle\varPhi_{2,u},\varPhi_{3,u}(t)\rangle_{\mathrm{deg}}+\langle\varPhi_{2,u},\varPhi_{3,u}(t)\rangle_{\mathrm{generic}}.

In these expressions,

3=Eis(Φ3(t)N){\mathcal{E}}_{3}=\mathrm{Eis}(\varPhi_{3}(t)_{N}^{*})

is the Eisenstein series formed out of the exponents of Φ3,u(t)\varPhi_{3,u}(t), ψ,Φ3,u(t)reg\langle\psi,\varPhi_{3,u}(t)\rangle_{reg} denote the regularized inner product (which is the regular inner product if π\pi^{\prime} is cuspidal).

The term

Φ2,u,Φ3,u(t)deg=Φ2,u,3(t)\langle\varPhi_{2,u},\varPhi_{3,u}(t)\rangle_{\mathrm{deg}}=\langle\varPhi_{2,u},{\mathcal{E}}_{3}(t)\rangle

is called the "degenerate term"; in a sense it replaces the finite spectrum contribution which vanishes for t0t\neq 0 in this formula [MVIHES, §5.2.7]. Indeed the degenerate term can be bounded by the method of [MVIHES, Lemma 5.2.9], satisfying

(3.20) Φ2,u,Φ3,u(t)deg(C(π2)C(π3))o(1)uγ\langle\varPhi_{2,u},\varPhi_{3,u}(t)\rangle_{\mathrm{deg}}\ll(C(\pi_{2})C(\pi_{3}))^{o(1)}\|u\|^{-\gamma}

for some absolute constant γ>1/2\gamma>1/2, similar to the finite spectrum. One can then control its contribution similarly as in (3.10).

To be self-contained we briefly describe the argument for (3.20) here. The degenerate term can be unfolded and related to the Rankin–Selberg L-function for each Eisenstein series formed out of the exponents. For example the exponent ||𝔸1+itT|\cdot|_{{\mathbb{A}}}^{1+it}\in T gives rise to a product of L(π2×π~2,1+t)L(\pi_{2}\times\tilde{\pi}_{2},1+t) with local period integrals at finite number of places, and some other unimportant factors. It suffices to control them separately. The LL-function factorizes as L(π2,Ad,1+t)ζ(1+t)L(\pi_{2},\mathrm{Ad},1+t)\zeta(1+t), and L(π2,Ad,1+t)C(π2)o(1)L(\pi_{2},\mathrm{Ad},1+t)\ll C(\pi_{2})^{o(1)} near t=0t=0, while ζ(1+t)\zeta(1+t) (despite having a pole at t=0t=0) can be uniformly bounded on fixed circle 𝒞{\mathcal{C}} round t=0t=0, independent of π2\pi_{2}.

On the other hand the local period integrals can be bounded by uγ\|u\|^{-\gamma} as in [MVIHES, Section 5.2.10]. Indeed for vram(Φ23)v\in\mathrm{ram}(\varPhi_{23}), the bound in, for example, [MVIHES, (5.22)] holds uniformly regardless of the ramifications.

Using [MVIHES, Lemma 4.4.3], the generic contribution can be treated exactly as in §3.5, and the analogues of Lemma 3.11 and (3.17) hold. Hence Theorem 3.14 also hold in the Rankin–Selberg case.

3.8. Conjecture on test vectors and subconvexity bound

To further obtain subconvexity bound from Theorem 3.14, we first formulate a local conjecture on the existence of proper test vectors.

Let vv be a place of FF. Recall that |PrΠvπv|2\left|\mathrm{Pr}_{\Pi_{v}}^{\pi_{v}}\right|^{2} is either a local integral of product of matrix coefficients if all the global representations are cuspidal, or the absolute value squared of local Rankin–Selberg integral if one of the representations is an Eisenstein serie. Recall that

Qv=Cv(π1×π2×π3),Mv=min{C(π2,v),C(π3,v)}.Q_{v}=C_{v}(\pi_{1}\times\pi_{2}\times\pi_{3}),\ M_{v}=\min\{C(\pi_{2,v}),C(\pi_{3,v})\}.

Recall the notation v,a\leq_{v,a} and relevant results from Section 2.3.

Definition 3.15.

Denote

P=vPv=vCv(π1×π2×π3)1/2maxi=2,3{Cv(πi×πi)}.P=\prod_{v}P_{v}=\prod_{v}\frac{C_{v}(\pi_{1}\times\pi_{2}\times\pi_{3})^{1/2}}{\max\limits_{i=2,3}\{C_{v}(\pi_{i}\times\pi_{i})\}}.
Remark 3.16.

One can check case by case that

maxi=2,3{Cv(πi×πi)}Cv(π1×π2×π3)1/2,\max\limits_{i=2,3}\{C_{v}(\pi_{i}\times\pi_{i})\}\leq C_{v}(\pi_{1}\times\pi_{2}\times\pi_{3})^{1/2},

so Pv1P_{v}\geq 1 is always true.

For φi,vπi,v\varphi_{i,v}\in\pi_{i,v}, where π3,v\pi_{3,v} is a parabolically induced representation, denote by

(3.21) IvRS(φ1,v,φ2,v,φ3,v)=Z(Fv)N\GL2(Fv)Wφ1,v(g)Wφ2,v(g)¯φ3,v(g)𝑑gI_{v}^{\text{RS}}\left(\varphi_{1,v},\varphi_{2,v},\varphi_{3,v}\right)=\int\limits_{{Z({F}_{v})N}\backslash{\text{GL}}_{2}{({F}_{v})}}W_{\varphi_{1,v}}\left(g\right)\overline{W_{\varphi_{2,v}}\left(g\right)}\varphi_{3,v}\left(g\right)dg

the local Rankin–Selberg integral. Here WφvW_{\varphi_{v}} is the Whittaker function associated to φv\varphi_{v} with respect to a fixed additive character ψv\psi_{v}. We also note that Wφv¯=Wφv\overline{W_{\varphi_{v}}}=W^{-}_{\varphi_{v}} where WφvW^{-}_{\varphi_{v}} is the Whittaker function for the additive character ψv(x)=ψv(x)\psi_{v}^{-}(x)=\psi_{v}(-x).

In general for φi,vπi,vB\varphi_{i,v}\in\pi_{i,v}^{B}, denote by

(3.22) IvT(φ1,v,φ2,v,φ3,v)=Fv×\GL2(Fv)i=13Φφi,v(g)dgI_{v}^{\text{T}}\left(\varphi_{1,v},\varphi_{2,v},\varphi_{3,v}\right)=\int\limits_{{F}_{v}^{\times}\backslash{\text{GL}}_{2}{({F}_{v})}}\prod\limits_{i=1}^{3}\Phi_{\varphi_{i,v}}\left(g\right)dg

the local integral for the triple product formula, where Φφv\Phi_{\varphi_{v}} is the matrix coefficient associated to φv\varphi_{v}.

Let

Iv=IvT or |IvRS|2I_{v}=I_{v}^{T}\hbox{ or }\left|I_{v}^{\text{RS}}\right|^{2}

depending on whether φ3\varphi_{3} or ψ\psi is an Eisenstein series. We formulate the following local conjecture, for which we skip subscript vv’s.

Conjecture 3.17.

Let ν\nu be a place of FF. We assume that if π3\pi_{3} or π\pi^{\prime} is an Eisenstein series (so that Iv=|IvRS|2I_{v}=\left|I_{v}^{\text{RS}}\right|^{2}) the corresponding local representation π3,v\pi_{3,v} or πv\pi_{v}^{\prime} is tempered. Then there exists normalized test vectors φi,vπi,vB\varphi_{i,v}\in\pi_{i,v}^{B} satisfying the following properties:

  1. (0)

    If Cv(πj)=MvC_{v}(\pi_{j})=M_{v} for j=2j=2 or 33, then φj,v\varphi_{j,v} is K(Mv)K(M_{v}^{\prime})-invariant if vv is non-archimedean, and 𝒮dπj,v(φj,v)MvA(d){\mathcal{S}}^{\pi_{j,v}}_{d}\left(\varphi_{j,v}\right)\ll M_{v}^{A(d)} if vv is archimedean as in Assumption 3.10.

  2. (1)

    There exists some constant AA such that

    I(φ1,v,φ2,v,φ3,v)ϵQvϵMvA1Qv1/4,I(\varphi_{1,v},\varphi_{2,v},\varphi_{3,v})\geq_{\epsilon}Q_{v}^{\epsilon}M_{v}^{A}\frac{1}{Q_{v}^{1/4}},
  3. (2)

    There exists some absolute constant AA (independent of vv) such that for ψvπv\psi_{v}\in\pi_{v}^{\prime} controlled by MvM_{v} as in Lemma 3.11, and jj^{\prime} the other index in {2,3}\{2,3\} different from jj,

    Iv(φj,v,φj,v,ψv)ϵQvϵMvA1maxi=2,3{Cv(πi×πi)}1/21Pv1/2θ.I_{v}(\varphi_{j^{\prime},v},\varphi_{j^{\prime},v},\psi_{v})\leq_{\epsilon}Q_{v}^{\epsilon}M_{v}^{A}\frac{1}{\max\limits_{i=2,3}\{C_{v}(\pi_{i}\times\pi_{i})\}^{1/2}}\frac{1}{P_{v}^{1/2-\theta}}.

Here θ<1/2\theta<1/2 is any bound towards the global Ramanujan conjecture.

Remark 3.18.

Note that for jj as above, we have

|Iv(φj,v,φj,v,ψv)|v,ϵMvAQvϵ,\left|I_{v}(\varphi_{j,v},\varphi_{j,v},\psi_{v})\right|\leq_{v,\epsilon}M_{v}^{A}Q_{v}^{\epsilon},

as πj,v\pi_{j,v}, φj,v\varphi_{j,v}, πv\pi_{v}^{\prime}, ψv\psi_{v} are all controlled in terms of MvM_{v}.

From the definitions of feasibility index and equidistribution indices in Definition 3.6, 3.8, we immediately obtain the following:

Theorem 3.19.

Suppose that Conjecture 3.17 is true. Then there exists constants δ,A>0\delta,A>0 such that

(3.23) L(π1×π2×π3,1/2)ϵMAC(π1×π2×π3)1/4+ϵ1Pδ.L(\pi_{1}\times\pi_{2}\times\pi_{3},1/2)\ll_{\epsilon}M^{A}C(\pi_{1}\times\pi_{2}\times\pi_{3})^{1/4+\epsilon}\frac{1}{P^{\delta}}.
Remark 3.20.

In particular this is a subconvexity bound if MϵQϵM\ll_{\epsilon}Q^{\epsilon} and there exists a constant γ>0\gamma>0 such that

vmaxi=2,3{Cv(πi×πi)}MAC(π1×π2×π3)1/2γ.\prod_{v}\max\limits_{i=2,3}\{C_{v}(\pi_{i}\times\pi_{i})\}\leq M^{A}C(\pi_{1}\times\pi_{2}\times\pi_{3})^{1/2-\gamma}.
Proof of Theorem 3.19.

We first remark that |PrΠvπv|2\left|\Pr_{\Pi_{v}}^{\pi_{v}}\right|^{2} differs from IvI_{v} only by some absolute constant and normalizing L-factors, which can be controlled by QϵQ^{\epsilon} in the sense of Example 2.4. So we shall not distinguish them below.

The temperedness condition in Conjecture 3.17 is satisfied as when φ3\varphi_{3}/ψ\psi is a unitary Eisenstein series in the Rankin–Selberg case, the associated local component π3,v\pi_{3,v}/ πv\pi^{\prime}_{v} is indeed tempered.

Let j,jj,j^{\prime} be as in Conjecture 3.17. For the test vectors φi\varphi_{i} which are specified in Definition 3.5 and satisfy Conjecture 3.17, it follows from Theorem 3.14 (which requires item (0) of Conjecture 3.17), Definition 3.6 and item (1) of Conjecture 3.17 that

L(π×π2×π3,1/2)ϵMAC(π1×π2×π2)1/4+ϵmin(1,𝔼(Φ23)δ).L(\pi\times\pi_{2}\times\pi_{3},1/2)\ll_{\epsilon}M^{A}C(\pi_{1}\times\pi_{2}\times\pi_{2})^{1/4+\epsilon}\min\left(1,{\mathbb{E}}(\varPhi_{23})^{\delta}\right).

For 𝔼(Φ23){\mathbb{E}}(\varPhi_{23}) as in (3.18), using Definition 3.8, bound (3.15), item (2) of Conjecture 3.17 and Remark 3.18, we get

𝔼(Φ23)ϵMAQϵ1P(1/2θ)δ.{\mathbb{E}}(\varPhi_{23})\ll_{\epsilon}M^{A}Q^{\epsilon}\frac{1}{P^{(1/2-\theta)\delta}}.

Then the theorem follows for a different δ>0\delta>0 as P1P\geq 1 by Remark 3.16. ∎

Remark 3.21.

Our formulation also implies that if Conjecture 3.17 is true and the subconvexity bound for L(πj×πj×π,1/2)L(\pi_{j^{\prime}}\times\pi_{j^{\prime}}\times\pi^{\prime},1/2) holds for controlled π\pi^{\prime}, then the subconvexity bound for L(π1×π2×π3,1/2)L(\pi_{1}\times\pi_{2}\times\pi_{3},1/2) always holds up to a factor MAM^{A}.

The remaining of this paper is devoted to partially verifying Conjecture 3.17. In particular we prove the following result.

Theorem 3.22.

Suppose that πi\pi_{i} have trivial central characters with bounded archimedean components, and Mf=1M_{f}=1. Then Conjecture 3.17 is true.

Remark 3.23.

One can immediately obtain the same result if each πi\pi_{i} is further twisted by some character ηi\eta_{i} as long as ηi=1\prod\eta_{i}=1, allowing more general central characters.

In general there are some sporadic cases (for example, in specific archimedean aspect, or for square-free MfM_{f}) where one can also verify the conjecture. For the conciseness of this paper we limit ourselves to the case in Theorem 3.22.

4. Local preparations

The remaining sections are purely local. For the sake of conciseness we shall omit subscript vv from the notations.

4.1. Basics

We collect some basic definitions and results here.

Let FF be a p-adic local field, with ring of integers OFO_{F}, uniformizer ϖ\varpi, valuation vv, residue field kFk_{F}, q=|kF|q=|k_{F}| and pp be the characteristic of kFk_{F}. Let EE be an étale quadratic algebra over FF. When EE is a field, we denote OEO_{E} to be the ring of integers, ϖE\varpi_{E} to be a uniformizer, and eEe_{E} to be the ramification index of EE. We normalize the valuation on EE such that vE(ϖE)=1v_{E}(\varpi_{E})=1. Define UE(i)=1+ϖiOEU_{E}(i)=1+\varpi^{i}O_{E}. To be uniform, we also take the following conventions when working with the case where EE splits.

Definition 4.1.

Let EE be a split quadratic extension over FF identified with F×FF\times F. Take eE=1e_{E}=1 in this case. Define OE=OF×OFO_{E}=O_{F}\times O_{F}, OE×=OF××OF×O_{E}^{\times}=O_{F}^{\times}\times O_{F}^{\times}, UE(i)=1+ϖiOEU_{E}(i)=1+\varpi^{i}O_{E} for i>0i>0. Also we write vE(x)=iv_{E}(x)=i when xϖiOE×x\in\varpi^{i}O_{E}^{\times}. (In particular vEv_{E} is not defined for all EE in this case.)

For a non-trivial additive character ψ\psi over FF, denote

c(ψ)=i,c(\psi)=i,

if i0i\geq 0 is the smallest integer such that ψ|ϖFiOF=1.\psi|_{\varpi_{F}^{i}O_{F}}=1. This definition also works for additive characters over EE. For a multiplicative character χ\chi of F×F^{\times}, define

c(χ)={0, if ψ(OF×)=1;i, if i>0 is smallest integer such that χ|UF(i)=1.c(\chi)=\begin{cases}0,&\text{\ if }\psi(O_{F}^{\times})=1;\\ i,&\text{\ if $i>0$ is smallest integer such that }\chi|_{U_{F}(i)}=1.\end{cases}

Let ψ\psi now be an additive character over FF with c(ψ)=0c(\psi)=0. Denote ψE=ψTrE/F\psi_{E}=\psi\circ\text{Tr}_{E/F}. Then c(ψE)=eE+1c(\psi_{E})=-e_{E}+1. For a multiplicative character χ\chi on F×F^{\times}, we can associate a character χE=χNE/F\chi_{E}=\chi\circ N_{E/F} on E×E^{\times}.

For multiplicative characters over F×F^{\times}, we introduce an equivalence relation here.

Definition 4.2.

For any two characters χi\chi_{i} of F×F^{\times}, i=1,2i=1,2, χ1jχ2\chi_{1}\sim_{j}\chi_{2} if and only if c(χ1χ21)jc(\chi_{1}\chi_{2}^{-1})\leq j.

We collect the following straightforward lemmas.

Lemma 4.3.

Suppose that 2q2\nmid q. Let E,EE,E^{\prime} be two non-isomorphic étale quadratic algebra over FF, α1E\alpha_{1}\in E, α2E\alpha_{2}\in E^{\prime} be trace 0 elements with v(NE/F(α1))=v(NE/F(α2))v(N_{E/F}(\alpha_{1}))=v(N_{E/F}(\alpha_{2})). Let α0F\alpha_{0}\in F with v(NE/F(α1))v(NE/F(α0))v(N_{E/F}(\alpha_{1}))\leq v(N_{E/F}(\alpha_{0})). Then we have

NE/F(α1+α0)NE/F(α2+α0)1modϖ.\frac{N_{E/F}(\alpha_{1}+\alpha_{0})}{N_{E^{\prime}/F}(\alpha_{2}+\alpha_{0})}\not\equiv 1\mod\varpi.
Proof.

Since αi\alpha_{i} are trace 0 for i=1,2i=1,2 and α0F\alpha_{0}\in F, we have

N(αi+α0)=N(αi)+N(α0).N(\alpha_{i}+\alpha_{0})=N(\alpha_{i})+N(\alpha_{0}).

By the condition v(NE/F(α1))v(NE/F(α0))v(N_{E/F}(\alpha_{1}))\leq v(N_{E/F}(\alpha_{0})), we are reduced to prove

NE/F(α1)NE/F(α2)1modϖ.\frac{N_{E/F}(\alpha_{1})}{N_{E^{\prime}/F}(\alpha_{2})}\not\equiv 1\mod\varpi.

This follows directly from that E,EE,E^{\prime} are not isomorphic and 2q2\nmid q. ∎

Lemma 4.4.

Let mFm\in F such that v(m)=j<0v(m)=-j<0, and μ\mu be a character of OF×O_{F}^{\times} of level k>0k>0. Then

(4.1) |v(x)=0ψ(mx)μ1(x)d×x|={q(q1)2qk1=ζF(1)qk/2, if j=k;0, otherwise.\left|\int\limits_{v(x)=0}\psi(mx)\mu^{-1}(x)d^{\times}x\right|=\begin{cases}\sqrt{\frac{q}{(q-1)^{2}q^{k-1}}}=\zeta_{F}(1)q^{-k/2},&\text{\ if\ }j=k;\\ 0,&\text{\ otherwise.}\end{cases}
Lemma 4.5.

Let θ\theta be a character over an étale quadratic algebra E/FE/F, such that eE=1e_{E}=1, c(θ)=1c(\theta)=1 θ|F×=1\theta|_{F^{\times}}=1. Then for any trace 0 element α\alpha with vE(α)=0v_{E}(\alpha)=0,

|xkF×,x+αOE×θ(x+α)|2.\left|\sum\limits_{x\in k_{F}^{\times},x+\alpha\in O_{E}^{\times}}\theta(x+\alpha)\right|\leq 2.

Here we have identified OF×/UF(1)O_{F}^{\times}/U_{F}(1) with kF×k_{F}^{\times} without confusion.

Proof.

Consider the case EE is a field first. As c(θ)=1c(\theta)=1, we have

x,ykF,(x,y)(0,0)θ(x+yα)=0.\sum\limits_{x,y\in k_{F},(x,y)\neq(0,0)}\theta(x+y\alpha)=0.

Using that θ|F×=1\theta|_{F^{\times}}=1, we get that

xkF×θ(x+α)=1q1x,ykF×θ(x+yα)=1θ(α).\sum\limits_{x\in k_{F}^{\times}}\theta(x+\alpha)=\frac{1}{q-1}\sum\limits_{x,y\in k_{F}^{\times}}\theta(x+y\alpha)=-1-\theta(\alpha).

The claim follows. On the other hand if EF×FE\simeq F\times F, θ=(μ,μ1)\theta=(\mu,\mu^{-1}), α=(y,y)\alpha=(y,-y), then

xkF×,x+αOE×θ(x+α)=xkF×,xy,yμ(x+yxy)=zkF×,z±1μ(z)=μ(1)μ(1).\sum\limits_{x\in k_{F}^{\times},x+\alpha\in O_{E}^{\times}}\theta(x+\alpha)=\sum\limits_{x\in k_{F}^{\times},x\not\equiv y,-y}\mu\left(\frac{x+y}{x-y}\right)=\sum\limits_{z\in k_{F}^{\times},z\not\equiv\pm 1}\mu(z)=-\mu(1)-\mu(-1).

Again the claim follows. ∎

Lemma 4.6.

For a character μ\mu over F×{F}^{\times} with c(μ)2c(\mu)\geq 2, there exists αμF×\alpha_{\mu}\in{F}^{\times} with vF(αμ)=c(μ)+c(ψF)v_{F}(\alpha_{\mu})=-c(\mu)+c(\psi_{F}) such that:

(4.2) μ(1+u)=ψF(αμu) for any uϖFc(μ)/2OF.\mu(1+u)=\psi_{F}(\alpha_{\mu}u)\text{\ \ for any $u\in\varpi_{F}^{\lceil c(\mu)/2\rceil}O_{F}$}.

Here we keep track of c(ψF)c(\psi_{F}) so we can also apply the lemma directly to a character θ\theta defined over an étale quadratic algebra EE. Also note that if θ\theta is a character over an étale quadratic algebra E/FE/F such that θ|F×=1\theta|_{F^{\times}}=1, then the associated element αθ\alpha_{\theta} can be chosen as a trace 0 element in EE.

Corollary 4.7.

Let E/FE/F be an étale quadratic algebra. χ\chi is defined over F×F^{\times} with c(χ)2c(\chi)\geq 2, and is associated to α\alpha by χ(1+x)=ψ(αx)\chi(1+x)=\psi(\alpha x) for v(x)c(χ)/2v(x)\geq c(\chi)/2 and v(α)=c(χ)v(\alpha)=-c(\chi).

Then α\alpha is also associated to χE\chi_{E}, in the sense that

χNE/F(1+y)=ψE(αy),yϖEeEc(χ)eE+12OE.\chi\circ N_{E/F}(1+y)=\psi_{E}(\alpha y),\forall y\in\varpi_{E}^{\lceil\frac{e_{E}c(\chi)-e_{E}+1}{2}\rceil}O_{E}.

Note that when p2p\neq 2,

c(χE)=eEc(χ)eE+1.c(\chi_{E})=e_{E}c(\chi)-e_{E}+1.
Lemma 4.8.

Let μ\mu, χ\chi be multiplicative characters on F×F^{\times} and ψ\psi be an additive character on FF. Suppose that c(μ)2c(χ)c(\mu)\geq 2c(\chi), and αμ\alpha_{\mu} is associated to μ\mu by Lemma 4.6. Then

(4.3) v(x)=c(μ)+c(ψ)μ(x)χ(x)ψ(x)d×x=χ(αμ)v(x)=c(μ)+c(ψ)μ(x)ψ(x)d×x.\int\limits_{v(x)=-c(\mu)+c(\psi)}\mu(x)\chi(x)\psi(x)d^{\times}x=\chi(-\alpha_{\mu})\int\limits_{v(x)=-c(\mu)+c(\psi)}\mu(x)\psi(x)d^{\times}x.

This result follows immediately from the p-adic stationary phase analysis.

Let K1(ϖc)K_{1}(\varpi^{c}) denote the compact subgroup of GL2{\text{GL}}_{2} whose elements are congruent to (01)mod(ϖc)\begin{pmatrix}*&*\\ 0&1\end{pmatrix}\text{mod}{(\varpi^{c})}. Now we record some basic facts about integrals on GL2(F){\text{GL}}_{2}(F).

Lemma 4.9.

For every positive integer cc,

GL2(F)=0icB(10ϖi1)K1(ϖc).{\text{GL}}_{2}(F)=\coprod\limits_{0\leq i\leq c}B\begin{pmatrix}1&0\\ \varpi^{i}&1\end{pmatrix}K_{1}(\varpi^{c}).

Here BB is the Borel subgroup of GL2{\text{GL}}_{2}.

We normalize the Haar measure on GL2(F){\text{GL}}_{2}({F}) such that its maximal compact open subgroup KK has volume 1. Then we have the following easy result (see, for example, [YH13, Appendix A]).

Lemma 4.10.

Locally let ff be a K1(ϖc)K_{1}(\varpi^{c})-invariant function, on which the center acts trivially. Then

(4.4) F×\GL2(F)f(g)𝑑g=0icAiF×\B(F)f(b(10ϖi1))𝑑b.\int\limits_{F^{\times}\backslash{\text{GL}}_{2}({F})}f(g)dg=\sum\limits_{0\leq i\leq c}A_{i}\int\limits_{{F}^{\times}\backslash B({F})}f\left(b\begin{pmatrix}1&0\\ \varpi^{i}&1\end{pmatrix}\right)db.

Here dbdb is the left Haar measure on F×\B(F){F}^{\times}\backslash B({F}), and

A0=pp+1Ac=1(p+1)pc1Ai=p1(p+1)pi for 0<i<c.A_{0}=\frac{p}{p+1}\text{,\ \ \ }A_{c}=\frac{1}{(p+1)p^{c-1}}\text{,\ \ \ }A_{i}=\frac{p-1}{(p+1)p^{i}}\text{\ \ for\ }0<i<c.

4.2. More preparations for Whittaker model and matrix coefficient

Here we briefly review known results on the Whittaker function and matrix coefficient for newforms. Let π\pi be an irreducible smooth representation of GL2(F){\text{GL}}_{2}(F) with trivial central character. In particular π\pi is unitary. Any element φπ\varphi\in\pi can be associated to a Whittaker function WφW_{\varphi} in the Whittaker model of π\pi. For asymptotic purposes, we assume either 2q2\nmid q, or c(π)c(\pi) is large enough when 2|q2|q.

Lemma 4.11.

The unitary pairings on π\pi can be given in the Whittaker model as follows:

W1,W2=xF×W1(a(x))W2(a(x))¯d×x.\langle W_{1},W_{2}\rangle=\int\limits_{x\in{F}^{\times}}W_{1}(a(x))\overline{W_{2}(a(x))}d^{\times}x.

In the following we give explicit formulae for Whittaker functions.

Definition 4.12.

For WπW\in\pi a Whittaker function, denote

W(j)(α)=W((α1)(1ϖj1)).W^{(j)}(\alpha)=W\left(\begin{pmatrix}\alpha&\\ &1\end{pmatrix}\begin{pmatrix}1&\\ \varpi^{j}&1\end{pmatrix}\right).
Remark 4.13.

If φ0\varphi_{0} is the newform, and Wφ0W_{\varphi_{0}} is the associated Whittaker function, then by Lemma 4.9 and the basic properties of Whittaker newforms, Wφ0W_{\varphi_{0}} is determined by Wφ0(j)W_{\varphi_{0}}^{(j)} for 0jc(π)0\leq j\leq c(\pi).

4.2.1. Unramified representations and Special unramified representations

Lemma 4.14.

Suppose μi\mu_{i} are unramified (that is, c(μi)=0c(\mu_{i})=0) and π=π(μ1,μ2)\pi=\pi(\mu_{1},\mu_{2}). Let φ0π\varphi_{0}\in\pi be a newform and Wφ0W_{\varphi_{0}} be its associated Whittaker function normalized so that Wφ0(1)=1W_{\varphi_{0}}(1)=1. Then Wφ0W_{\varphi_{0}} is invariant under the maximal compact open subgroup and

(4.5) Wφ0(0)(α)={|α|1/2μ1(ϖα)μ2(ϖα)μ1(ϖ)μ2(ϖ),if v(α)0;0,otherwise.W_{\varphi_{0}}^{(0)}(\alpha)=\begin{cases}|\alpha|^{1/2}\frac{\mu_{1}(\varpi\alpha)-\mu_{2}(\varpi\alpha)}{\mu_{1}(\varpi)-\mu_{2}(\varpi)},&\text{if }v(\alpha)\geq 0;\\ 0,&\text{otherwise}.\end{cases}
Remark 4.15.

Note that when v(α)0v(\alpha)\geq 0, the numerator contains the denominator as a factor and can be canceled. In this sense the formula still holds when μ1(ϖ)=μ2(ϖ)\mu_{1}(\varpi)=\mu_{2}(\varpi). Also note that the above expression for Wφ0W_{\varphi_{0}} is not L2L^{2}-normalized, but differ only by a factor which can be controlled globally by QϵQ^{\epsilon}.

Lemma 4.16.

Let π=σ(μ||1/2,μ||1/2)\pi=\sigma(\mu|\cdot|^{1/2},\mu|\cdot|^{-1/2}) be a special unramified representation, where μ\mu is unramified with μ2=1\mu^{2}=1.

The Whittaker function associated to the newform φ0π\varphi_{0}\in\pi is given by

(4.6) Wφ0(1)(α)={μ(α)|α|, if v(α)0;0, otherwise.W_{\varphi_{0}}^{(1)}\left(\alpha\right)=\begin{cases}\mu(\alpha)|\alpha|,&\text{\ if $v(\alpha)\geq 0$;}\\ 0,&\text{\ otherwise}.\end{cases}
(4.7) Wφ0(0)(α)={q1μ(α)|α|ψ(α), if v(α)1;0, otherwise.W_{\varphi_{0}}^{(0)}\left(\alpha\right)=\begin{cases}-q^{-1}\mu(\alpha)|\alpha|\psi(\alpha),&\text{\ if $v(\alpha)\geq-1$;}\\ 0,&\text{\ otherwise}.\end{cases}

Again the expressions are not L2L^{2}-normalized but differ only by a constant controlled by QϵQ^{\epsilon}. Lemma 4.14 and 4.16 are well-known except (4.7), which can be obtained by reformulating [YH13, Corollary 5.7].

4.2.2. Supercuspidal representation case

Suppose that p2p\neq 2, or c(π)c(\pi) large enough when p=2p=2. Then a supercuspidal representation π\pi can be associated with a character θ\theta over a quadratic field extension EE by Local Langlands Correspondence, and are called dihedral.

The levels of dihedral supercuspidal representations can be associated to the levels of θ\theta by the following relations.

  1. Case 1.

    c(π)=2n+1c\left(\pi\right)=2n+1 corresponds to eE=2e_{E}=2 and c(θ)=2nc\left(\theta\right)=2n .

  2. Case 2.

    c(π)=4nc\left(\pi\right)=4n corresponds to eE=1e_{E}=1 and c(θ)=2nc\left(\theta\right)=2n.

  3. Case 3.

    c(π)=4n+2c\left(\pi\right)=4n+2 corresponds to eE=1e_{E}=1 and c(θ)=2n+1c\left(\theta\right)=2n+1 .

Recall from [HuSa:19, Lemma 5.7] the following result, which is a reformulation of [Assing, Lemma 3.1] and holds actually for all dihedral supercuspidal representations.

Lemma 4.17.

Let π\pi be a dihedral supercuspidal representation with trivial central character. Let c=c(π)c=c(\pi). Denote

C0=v𝔼(u)=c(θ)e𝔼+1θ1(u)ψ𝔼(u)d×u.C_{0}=\int\limits_{v_{\mathbb{E}}(u)=-c(\theta)-e_{\mathbb{E}}+1}\theta^{-1}(u)\psi_{\mathbb{E}}(u)d^{\times}u.

As a function in xx, W(i)(x)W^{(i)}(x) is supported on v(x)=min{0,2ic}v(x)=\min\{0,2i-c\}, consisting only of level cic-i components (in the sense of Mellin transform), except when i=c1i=c-1 where it consists of level 1\leq 1 components. In particular when ic/2i\geq c/2, W(i)(x)W^{(i)}(x) is supported on v(x)=0v(x)=0, and on the support,

W(i)(x)=C01v𝔼(u)=c(θ)e𝔼+1θ1(u)ψ(1xϖiN𝔼/F(u))ψ𝔼(u)d×u.W^{(i)}(x)=C_{0}^{-1}\int\limits_{v_{\mathbb{E}}(u)=-c(\theta)-e_{\mathbb{E}}+1}\theta^{-1}(u)\psi\left(-\frac{1}{x}\varpi^{i}N_{{\mathbb{E}}/{F}}(u)\right)\psi_{\mathbb{E}}(u)d^{\times}u.

The normalization by C0C_{0} guarantees that W(c)(1)=1W^{(c)}(1)=1. Note that it is also possible to give an expression in the case i<c/2i<c/2 by using Atkin-Lehner symmetry. Though we do not need such explicit formula.

Corollary 4.18.

Suppose ic/2i\geq c/2. Let χ\chi be a character over F×F^{\times} such that χ(ϖ)=1\chi(\varpi)=1, c(χ)=cic(\chi)=c-i when ic1i\neq c-1, or c(χ)1c(\chi)\leq 1 when i=c1i=c-1. Then

xOF×W(i)(x)χ(x)d×x\displaystyle\int\limits_{x\in O_{F}^{\times}}W^{(i)}(x)\chi(x)d^{\times}x
=\displaystyle= C01xOF×ψ(ϖicx)χ1(x)d×xvE(u)=c(θ)e𝔼+1θ1(u)χE(u)ψ𝔼(u)d×u.\displaystyle C_{0}^{-1}\int\limits_{x\in O_{F}^{\times}}\psi(-\varpi^{i-c}x)\chi^{-1}(x)d^{\times}x\int\limits_{v_{E}(u)=-c(\theta)-e_{\mathbb{E}}+1}\theta^{-1}(u)\chi_{E}(u)\psi_{\mathbb{E}}(u)d^{\times}u.

Here we note that v(NE/F(u))=cv(N_{E/{F}}(u))=-c when vE(u)=c(θ)e𝔼+1v_{E}(u)=-c(\theta)-e_{\mathbb{E}}+1.

Note that by [BH06, Chapter 11] there are (exceptional) supercuspidal representations over p=2p=2 which are not directly related to some θ\theta over a quadratic extension EE, but they have bounded conductors by [BH06, Corollary 45.6].

For non-dihedral supercuspidal representations, we only need the following less precise result coming from [hu_triple_2017, Corollary 2.18]:

Lemma 4.19.

Let π\pi be a supercuspidal representation with trivial central character. Then

W(c)=char(OF×),W^{(c)}={\text{char}}({O_{F}^{\times}}),

and W(i)(x)W^{(i)}(x) consists of level cic-i components (and level 0 component when i=c1i=c-1).

4.2.3. Principal series representation case

Let π=π(μ1,μ)\pi=\pi(\mu^{-1},\mu) be a principal series representation with c(μ)=c0=c/2c(\mu)=c_{0}=c/2. In this case denote

C=uϖc0OF×μ(u)ψ(u)𝑑u.C=\int\limits_{u\in\varpi^{-c_{0}}O_{F}^{\times}}\mu(u)\psi(-u)du.

Using Lemma 4.4 we have that |C|qc0/2.|C|\asymp q^{c_{0}/2}.

By [hu_triple_2017, Lemma 2.12], we have

Lemma 4.20.

As a function in xx, W(i)(x)W^{(i)}(x) is supported on v(x)=min{0,2ic}v(x)=\min\{0,2i-c\}, except when i=c/2i=c/2, where W(i)(x)W^{(i)}(x) can be supported on v(x)0v(x)\geq 0. W(i)(x)W^{(i)}(x) consists only of level cic-i components, except when i=c1i=c-1 where it consists of level 1\leq 1 components. More explicitly,

  1. (1)

    When c0<ic(π)=2c0c_{0}<i\leq c(\pi)=2c_{0}, W(i)(x)W^{(i)}(x) is supported on xOF×x\in O_{F}^{\times}, where

    W(i)(x)=C1uϖc0OF×μ(1+uϖi)μ(xu)ψ(xu)𝑑u.W^{(i)}(x)=C^{-1}\int\limits_{u\in\varpi^{-c_{0}}O_{F}^{\times}}\mu(1+u\varpi^{i})\mu(xu)\psi(-xu)du.
  2. (2)

    If i=c0i=c_{0}, W(c0)(x)W^{(c_{0})}(x) is supported on xOFx\in O_{F}, where

    W(c0)(x)\displaystyle W^{(c_{0})}(x)
    =\displaystyle= C1v(u)c0,uϖc0(1+ϖOF)μ(1+uϖc0)μ(xu)|1uϖc0(1+uϖc0)|1/2ψ(xu)qv(x)/2𝑑u.\displaystyle C^{-1}\int\limits_{v(u)\leq-c_{0},u\notin\varpi^{-c_{0}}(-1+\varpi O_{F})}\mu(1+u\varpi^{c_{0}})\mu(xu)\left|\frac{1}{u\varpi^{c_{0}}(1+u\varpi^{c_{0}})}\right|^{1/2}\psi(-xu)q^{-v(x)/2}du.

As a function in xx, W(i)W^{(i)} consists only of level cic-i components, except when i=c1i=c-1 when it consists of level 1\leq 1 components.

Recall the convention for split quadratic algebra in Definition 4.1. It allows us to reformulate the above result similarly as in the field extension case. Denote θ=μ1μ\theta=\mu^{-1}\otimes\mu in this case. Then principal series representations can be parameterized by θ\theta over split quadratic algebra.

Corollary 4.21.

Denote C0=uϖc0OE×θ1(u)ψE(u)d×uC_{0}=\int\limits_{u\in\varpi^{-c_{0}}O_{E}^{\times}}\theta^{-1}(u)\psi_{E}(u)d^{\times}u. Then we have

  1. (1)

    When c0<ic(π)=2c0c_{0}<i\leq c(\pi)=2c_{0}, W(i)(x)W^{(i)}(x) is supported on xOF×x\in O_{F}^{\times}, where

    W(i)(x)=C01uϖc0OE×θ1(u)ψ(ϖixNE/F(u))ψE(u)d×u.W^{(i)}(x)=C_{0}^{-1}\int\limits_{u\in\varpi^{-c_{0}}O_{E}^{\times}}\theta^{-1}(u)\psi\left(-\frac{\varpi^{i}}{x}N_{E/F}(u)\right)\psi_{E}(u)d^{\times}u.
  2. (2)

    Denote ϖi,j=(ϖi,ϖj)\varpi^{i,j}=(\varpi^{i},\varpi^{j}) as an element in EE. For fixed k=v(x)k=v(x), we have

    W(c0)(x)\displaystyle W^{(c_{0})}(x)
    =\displaystyle= 0jkqk/2μ(ϖk2j)C0(u,v)ϖc0OE×θ1((u,v))ψ(ϖc0+kuvx)ψE((ϖkju,ϖjv))d×ud×v\displaystyle\sum\limits_{0\leq j\leq k}\frac{q^{-k/2}\mu(\varpi^{k-2j})}{C_{0}}\iint\limits_{(u,v)\in\varpi^{-c_{0}}O_{E}^{\times}}\theta^{-1}((u,v))\psi\left(-\frac{\varpi^{c_{0}+k}uv}{x}\right)\psi_{E}((\varpi^{k-j}u,\varpi^{j}v))d^{\times}ud^{\times}v
    =\displaystyle= 0jkqk/2μ(ϖk2j)C0uϖc0OE×θ1(u)ψ(ϖc0+kNE/F(u)x)ψE(ϖkj,ju)d×u.\displaystyle\sum\limits_{0\leq j\leq k}\frac{q^{-k/2}\mu(\varpi^{k-2j})}{C_{0}}\iint\limits_{u\in\varpi^{-c_{0}}O_{E}^{\times}}\theta^{-1}(u)\psi\left(-\frac{\varpi^{c_{0}+k}N_{E/F}(u)}{x}\right)\psi_{E}(\varpi^{k-j,j}u)d^{\times}u.
Proof.

Note that by Lemma 4.20, we have for c0<i2c0c_{0}<i\leq 2c_{0},

W(i)(x)\displaystyle W^{(i)}(x) =C1uϖc0OF×μ(1uϖi)μ(xu)ψ(xu)𝑑u\displaystyle=C^{-1}\int\limits_{u\in\varpi^{-c_{0}}O_{F}^{\times}}\mu(1-u\varpi^{i})\mu(-xu)\psi(xu)du
=1Cvϖc0OF×μ1(v)ψ(v)𝑑vu,vϖc0OF×μ1(v)ψ((1uϖi)v)μ(xu)ψ(xu)𝑑u𝑑v\displaystyle=\frac{1}{C\int\limits_{v\in\varpi^{-c_{0}}O_{F}^{\times}}\mu^{-1}(v)\psi(v)dv}\iint\limits_{u,v\in\varpi^{-c_{0}}O_{F}^{\times}}\mu^{-1}(v)\psi((1-u\varpi^{i})v)\mu(-xu)\psi(xu)dudv
=1μψ𝑑uμ1ψ𝑑vu,vϖc0OF×μ1(v)μ(u)ψ(ϖiuvx)ψ(u+v)𝑑u𝑑v.\displaystyle=\frac{1}{\int\mu\psi du\int\mu^{-1}\psi dv}\iint\limits_{u,v\in\varpi^{-c_{0}}O_{F}^{\times}}\mu^{-1}(v)\mu(u)\psi\left(-\frac{\varpi^{i}uv}{x}\right)\psi(u+v)dudv.

We can then change the Haar measures in the numerator and the denominator simultaneously. Part (1) is indeed as claimed. Part (2) can be proven similarly, with extra care about valuations. Note that for fixed k=v(x)k=v(x), the domain for uu need to satisfy v(u)c0kv(u)\geq-c_{0}-k to have nonzero integrals. Denote v(u)=c0jv(u)=-c_{0}-j for 0jk0\leq j\leq k. Then we have

W(c0)(x)\displaystyle W^{(c_{0})}(x)
=\displaystyle= 1μψ𝑑uv(u)c0,uϖc0(1+ϖOF)μ(1uϖc0)μ(xu)ψ(xu)qv(u)+c0k/2𝑑u\displaystyle\frac{1}{\int\mu\psi du}\int\limits_{v(u)\leq-c_{0},u\notin\varpi^{-c_{0}}(1+\varpi O_{F})}\mu(1-u\varpi^{c_{0}})\mu(xu)\psi(xu)q^{v(u)+c_{0}-k/2}du
=\displaystyle= 0jkqjk/2μ(ϖj)μψ𝑑uμ1ψ𝑑v\displaystyle\sum\limits_{0\leq j\leq k}\frac{q^{-j-k/2}\mu(\varpi^{-j})}{\int\mu\psi du\int\mu^{-1}\psi dv}
vϖc0OF×v(u)=c0j,uϖc0(1+ϖOF)μ1(v)ψ((1uϖc0)ϖjv)μ(xu)ψ(xu)𝑑u𝑑v\displaystyle\ \ \int\limits_{v\in\varpi^{-c_{0}}O_{F}^{\times}}\int\limits_{v(u)=-c_{0}-j,u\notin\varpi^{-c_{0}}(1+\varpi O_{F})}\mu^{-1}(v)\psi((1-u\varpi^{c_{0}})\varpi^{j}v)\mu(xu)\psi(xu)dudv
=\displaystyle= 0jkqjk/2μ(ϖkj)μψ𝑑uμ1ψ𝑑v\displaystyle\sum\limits_{0\leq j\leq k}\frac{q^{-j-k/2}\mu(\varpi^{k-j})}{\int\mu\psi du\int\mu^{-1}\psi dv}
vϖc0OF×v(u)=c0j,uxϖc0k(1+ϖOF)μ1(v)ψ((1uϖc0+kx)ϖjv)μ(u)ψ(ϖku)𝑑u𝑑v\displaystyle\ \ \int\limits_{v\in\varpi^{-c_{0}}O_{F}^{\times}}\int\limits_{v(u)=-c_{0}-j,u\notin x\varpi^{-c_{0}-k}(1+\varpi O_{F})}\mu^{-1}(v)\psi\left(\left(1-\frac{u\varpi^{c_{0}+k}}{x}\right)\varpi^{j}v\right)\mu(u)\psi(\varpi^{k}u)dudv
=\displaystyle= 0jkqk/2μ(ϖk2j)μψ𝑑uμ1ψ𝑑v(u,v)ϖc0OE×θ1((u,v))ψ(ϖc0+kuvx)ψE((ϖkju,ϖjv))𝑑u𝑑v.\displaystyle\sum\limits_{0\leq j\leq k}\frac{q^{-k/2}\mu(\varpi^{k-2j})}{\int\mu\psi du\int\mu^{-1}\psi dv}\iint\limits_{(u,v)\in\varpi^{-c_{0}}O_{E}^{\times}}\theta^{-1}((u,v))\psi\left(-\frac{\varpi^{c_{0}+k}uv}{x}\right)\psi_{E}((\varpi^{k-j}u,\varpi^{j}v))dudv.

In the fourth equality we removed the restriction on uu, as it only affects the case j=0j=0, where if uxϖc0(1+ϖOF)u\in x\varpi^{-c_{0}}(1+\varpi O_{F}), the integral in vv will be zero. ∎

Parallel to Corollary 4.18, we have

Corollary 4.22.

Suppose ic/2i\geq c/2. Let χ\chi be a character over F×F^{\times} such that χ(ϖ)=1\chi(\varpi)=1, c(χ)=cic(\chi)=c-i when ic1i\neq c-1 or c(χ)1c(\chi)\leq 1 when i=c1i=c-1. Then if i>c/2i>c/2,

xOF×W(i)(x)χ(x)d×x\displaystyle\int\limits_{x\in O_{F}^{\times}}W^{(i)}(x)\chi(x)d^{\times}x
=\displaystyle= C01xOF×ψ(ϖicx)χ1(x)d×xv(v)=c(θ)μ1(v)χ(v)ψ(v)d×vv(u)=c(θ)μ(u)χ(u)ψ(u)d×u\displaystyle C_{0}^{-1}\int\limits_{x\in O_{F}^{\times}}\psi(-\varpi^{i-c}x)\chi^{-1}(x)d^{\times}x\int\limits_{v(v)=-c(\theta)}\mu^{-1}(v)\chi(v)\psi(v)d^{\times}v\int\limits_{v(u)=-c(\theta)}\mu(u)\chi(u)\psi(u)d^{\times}u
=\displaystyle= C01xOF×ψ(ϖicx)χ1(x)d×xuϖc0OE×θ1(u)χE(u)ψE(u)d×u.\displaystyle C_{0}^{-1}\int\limits_{x\in O_{F}^{\times}}\psi(-\varpi^{i-c}x)\chi^{-1}(x)d^{\times}x\int\limits_{u\in\varpi^{-c_{0}}O_{E}^{\times}}\theta^{-1}(u)\chi_{E}(u)\psi_{E}(u)d^{\times}u.

If i=c/2i=c/2, k0k\geq 0 then

xOF×W(c/2)(ϖkx)χ(x)d×x\displaystyle\int\limits_{x\in O_{F}^{\times}}W^{(c/2)}(\varpi^{k}x)\chi(x)d^{\times}x
=\displaystyle= 0jkqk/2μ(ϖk2j)C0xOF×ψ(ϖc0x)χ1(x)d×xv(v)=c0μ1(v)χ(v)ψ(ϖjv)d×v\displaystyle\sum\limits_{0\leq j\leq k}\frac{q^{-k/2}\mu(\varpi^{k-2j})}{C_{0}}\int\limits_{x\in O_{F}^{\times}}\psi(-\varpi^{-c_{0}}x)\chi^{-1}(x)d^{\times}x\int\limits_{v(v)=-c_{0}}\mu^{-1}(v)\chi(v)\psi(\varpi^{j}v)d^{\times}v
v(u)=c0μ(u)χ(u)ψ(ϖkju)d×u.\displaystyle\int\limits_{v(u)=-c_{0}}\mu(u)\chi(u)\psi(\varpi^{k-j}u)d^{\times}u.

Furthermore if μ\mu is not an unramified twist of a quadratic character (or equivalently, c(μ2)0c(\mu^{2})\neq 0), we have

(4.8) xOF×W(c/2)(ϖkx)χ(x)d×x\displaystyle\int\limits_{x\in O_{F}^{\times}}W^{(c/2)}(\varpi^{k}x)\chi(x)d^{\times}x
=\displaystyle= j=0,kqk/2μ(ϖk2j)C0xOF×ψ(ϖc0x)χ1(x)d×xuϖEc0OE×θ1(u)χE(u)ψE(ϖkj,ju)d×u.\displaystyle\sum\limits_{j=0,k}\frac{q^{-k/2}\mu(\varpi^{k-2j})}{C_{0}}\int\limits_{x\in O_{F}^{\times}}\psi(-\varpi^{-c_{0}}x)\chi^{-1}(x)d^{\times}x\int\limits_{u\in\varpi_{E}^{-c_{0}}O_{E}^{\times}}\theta^{-1}(u)\chi_{E}(u)\psi_{E}(\varpi^{k-j,j}u)d^{\times}u.
Proof.

The only tricky part is the last statement. When the condition for μ\mu is satisfied and 2q2\nmid q, we have used that for any χ\chi with c(χ)=c(μ)c(\chi)=c(\mu), at least one of c(μ1χ)c(\mu^{-1}\chi), c(μχ)c(\mu\chi) is c(μ)c(\mu). So j=0j=0 or kk for the Gauss integrals to be nonzero. ∎

Corollary 4.23.

Suppose p2p\neq 2, θ=μ1μ\theta=\mu^{-1}\otimes\mu with c(μ2)1c(\mu^{2})\geq 1. Then

L2(W,k):=xOF×|W(c/2)(ϖkx)|d×x={q3q1, if k=0;2qk, if k1.\displaystyle L^{2}(W,k):=\int\limits_{x\in O_{F}^{\times}}\left|W^{(c/2)}(\varpi^{k}x)\right|d^{\times}x=\begin{cases}\frac{q-3}{q-1},\text{\ if $k=0$};\\ \frac{2}{q^{k}},\text{\ if $k\geq 1$}.\end{cases}
Proof.

Using the spectral decomposition on OF×O_{F}^{\times}, we have for c4c\geq 4,

L2(W,k)=\displaystyle L^{2}(W,k)= c(χ)=c/2|xOF×W(i)(ϖkx)χ(x)d×x|2\displaystyle\sum\limits_{c(\chi)=c/2}\left|\int\limits_{x\in O_{F}^{\times}}W^{(i)}(\varpi^{k}x)\chi(x)d^{\times}x\right|^{2}
=\displaystyle= c(χ)=c/2i,j=0,kμ(ϖk2i)μ(ϖk2j)¯q2k|C0|2(q1)2qc/2uϖEc0OE×θ1(u)χE(u)ψE(ϖki,iu)d×u\displaystyle\sum\limits_{c(\chi)=c/2}\sum\limits_{i,j=0,k}\frac{\mu(\varpi^{k-2i})\overline{\mu(\varpi^{k-2j})}q^{2-k}}{|C_{0}|^{2}(q-1)^{2}q^{c/2}}\int\limits_{u\in\varpi_{E}^{-c_{0}}O_{E}^{\times}}\theta^{-1}(u)\chi_{E}(u)\psi_{E}(\varpi^{k-i,i}u)d^{\times}u
wϖEc0OE×θ1(w)χE(w)ψE(ϖkj,jw)d×w¯.\displaystyle\overline{\int\limits_{w\in\varpi_{E}^{-c_{0}}O_{E}^{\times}}\theta^{-1}(w)\chi_{E}(w)\psi_{E}(\varpi^{k-j,j}w)d^{\times}w}.

When k=0k=0 and c4c\geq 4, the Gauss integral uϖEc0OE×θ1(u)χE(u)ψE(u)d×u\int\limits_{u\in\varpi_{E}^{-c_{0}}O_{E}^{\times}}\theta^{-1}(u)\chi_{E}(u)\psi_{E}(u)d^{\times}u is non-vanishing if and only if c(χμ1)=c(χμ11)=c/2c(\chi\mu_{1})=c(\chi\mu_{1}^{-1})=c/2. The number of such characters is (q3)(q1)qc/22(q-3)(q-1)q^{c/2-2}. Then we have by Lemma 4.4

L2(W,0)=q3q1.L^{2}(W,0)=\frac{q-3}{q-1}.

When 0<k<c/210<k<c/2-1, uϖEc0OE×θ1(u)χE(u)ψE(ϖki,iu)d×u\int\limits_{u\in\varpi_{E}^{-c_{0}}O_{E}^{\times}}\theta^{-1}(u)\chi_{E}(u)\psi_{E}(\varpi^{k-i,i}u)d^{\times}u is non-vanishing if and only if c(μ1χ)=c/2kc(\mu^{-1}\chi)=c/2-k and i=ki=k, or c(μχ)=c/2kc(\mu\chi)=c/2-k and i=0i=0. Thus for fixed χ\chi, only i=j=0i=j=0 or i=j=ki=j=k contributes. The number of possible χ\chi is 2(q1)2qc/2k22(q-1)^{2}q^{c/2-k-2}, and

L2(W,k)=2qk.L^{2}(W,k)=\frac{2}{q^{k}}.

When k=c/21k=c/2-1, the discussion is similar, except that unramified χ\chi also contributes. The number of level 11 characters is q2q-2. Thus

L2(W,c1)=2qc1[(q2)q(q1)2+(1q1)2]=2qc1.L^{2}(W,c-1)=\frac{2}{q^{c-1}}\left[(q-2)\frac{q}{(q-1)^{2}}+\left(-\frac{1}{q-1}\right)^{2}\right]=\frac{2}{q^{c-1}}.

When kc/2k\geq c/2, uϖEc0OE×θ1(u)χE(u)ψE(u)d×u\int\limits_{u\in\varpi_{E}^{-c_{0}}O_{E}^{\times}}\theta^{-1}(u)\chi_{E}(u)\psi_{E}(u)d^{\times}u is non-vanishing if and only if c(μ1χ)=0c(\mu^{-1}\chi)=0 and i=ki=k, or c(μχ)=0c(\mu\chi)=0 and i=0i=0. Then we have

L2(W,k)=2qk.L^{2}(W,k)=\frac{2}{q^{k}}.

The total L2L^{2} mass k0L2(W,k)=1\sum\limits_{k\geq 0}L^{2}(W,k)=1, which is expected.

Consider now the case c=2c=2, and c(μ2)=1c(\mu^{2})=1. When k1k\geq 1, the contribution comes from c(μ1χ)=0c(\mu^{-1}\chi)=0 or c(μχ)=0c(\mu\chi)=0, and

L2(W,k)=2qk.L^{2}(W,k)=\frac{2}{q^{k}}.

When k=0k=0, the contribution comes from all χ\chi with c(χ)1c(\chi)\leq 1, and

L2(W,0)=[(q4)q(q1)2+1(q1)2+21(q1)2]=q3q1.L^{2}(W,0)=\left[(q-4)\frac{q}{(q-1)^{2}}+\frac{1}{(q-1)^{2}}+2\frac{1}{(q-1)^{2}}\right]=\frac{q-3}{q-1}.

Here the first term comes from χμ,μ1,1\chi\neq\mu,\mu^{-1},1, the middle term comes from χ=1\chi=1, and the last term comes from χ=μ,μ1\chi=\mu,\mu^{-1}.

4.3. List of levels

Recall that when π\pi is a special unramified representation, c(π)=1c(\pi)=1. For the remaining cases, π\pi corresponds to a character θ\theta over étale quadratic algebra EE, and we write π=πθ\pi=\pi_{\theta} in that case. One can uniformly describe its level by the formula

c(πθ)=2c(θ)eE+eE1.c(\pi_{\theta})=\frac{2c(\theta)}{e_{E}}+e_{E}-1.

In the Rankin–Selberg/triple product case, one can use the local Langlands correspondence to get the following:

Lemma 4.24.

Suppose that the central characters of πi\pi_{i} are trivial.

  1. (1)

    If c(πi)=1c(\pi_{i})=1, then c(π1×π2)=2c(\pi_{1}\times\pi_{2})=2;

  2. (2)

    If c(π1)c(π2)c(\pi_{1})\neq c(\pi_{2}), or πi\pi_{i} are associated to different étale quadratic algebras, then

    c(π1×π2)=max{c(π1)2,c(π2)2};c(\pi_{1}\times\pi_{2})=\max\{c(\pi_{1})^{2},c(\pi_{2})^{2}\};
  3. (3)

    If πi\pi_{i} are associated to θi\theta_{i} over the same étale quadratic algebra EE, then

    c(πθ1×πθ2)=c(πθ1θ2)+c(πθ1θ2¯);c(\pi_{\theta_{1}}\times\pi_{\theta_{2}})=c(\pi_{\theta_{1}\theta_{2}})+c(\pi_{\theta_{1}\overline{\theta_{2}}});
  4. (4)

    If c(π3)=0c(\pi_{3})=0, then

    c(π1×π2×π3)=c(π1×π2)2.c(\pi_{1}\times\pi_{2}\times\pi_{3})=c(\pi_{1}\times\pi_{2})^{2}.

5. Partial orthogonality for Whittaker functions

Let c(Π)c(\Pi) be the exponent of finite conductor C(Π)C(\Pi). In this section we shall focus on the case c(π1)=c(π2)c(\pi_{1})=c(\pi_{2}), c(π3)=0c(\pi_{3})=0.

Definition 5.1.

For two Whittaker functions WiπiW_{i}\in\pi_{i}, normalized so that Wi(1)=1W_{i}(1)=1, we denote

Pi(W1,W2,γ)=xF×W1(i)(x)W2(i)(x)¯|x|1/2+γd×xP_{i}(W_{1},W_{2},\gamma)=\int\limits_{x\in{F}^{\times}}W_{1}^{(i)}(x)\overline{W_{2}^{(i)}(x)}|x|^{-1/2+\gamma}d^{\times}x

where γ\gamma is a parameter such that |γ|<7/64|\gamma|<7/64. In the following we skip the parameter γ\gamma from PiP_{i} without confusion.

As we shall see in the next section, this quantity is vital in evaluating the local period integrals. So our goal in this section is to evaluate or give proper bounds for it. The simplest case is where c(πi)=1c(\pi_{i})=1.

Lemma 5.2.

Suppose that πi=σ(μi||1/2,μi||1/2)\pi_{i}=\sigma(\mu_{i}|\cdot|^{1/2},\mu_{i}|\cdot|^{-1/2}) are special unramified representations for c(μi)=0c(\mu_{i})=0 and μi2=1\mu_{i}^{2}=1. Then for WiπiW_{i}\in\pi_{i} the Whittaker functions associated to newforms,

P1(W1,W2)=11μ1μ2(ϖ)q3/2γ.P_{1}(W_{1},W_{2})=\frac{1}{1-\mu_{1}\mu_{2}(\varpi)q^{-3/2-\gamma}}.

This follows directly from the definition of PiP_{i} and Lemma 4.16.

We introduce the following technical assumption which we shall circumvent later on.

Assumption 5.3.

If c(π1)=c(π2)>0c(\pi_{1})=c(\pi_{2})>0, then πi,i=1,2\pi_{i},i=1,2 are not twists of unramified representations by quadratic characters.

The main result of this section is the following.

Proposition 5.4.

Suppose that p2p\neq 2, c(π1)=c(π2)=c2c(\pi_{1})=c(\pi_{2})=c\geq 2 satisfying Assumption 5.3, so that cc(π1×π2)2cc\leq c(\pi_{1}\times\pi_{2})\leq 2c. Let WiW_{i} be the Whittaker functions associated to newforms φiπi\varphi_{i}^{\circ}\in\pi_{i}. Then for p2p\neq 2, ic/2>1i\geq c/2>1, we have

(5.1) {Pi(W1,W2)=1, if ic(π1×π2)/2, except when i=c/2 and  πi are principal series;|Pi(W1,W2)q3q1|2q1/2+Re(γ)1, if i=c(π1×π2)/2=c/2πi are principal series;Pi(W1,W2)=1q1, if i=c(π1×π2)/21>c/2;|Pi(W1,W2)|2q1, if i=c(π1×π2)/21=c/2;Pi(W1,W2)=0, otherwise.\begin{cases}P_{i}(W_{1},W_{2})=1,&\text{\ if $i\geq c(\pi_{1}\times\pi_{2})/2$, except when $i=c/2$ and }\\ &\text{\ $\pi_{i}$ are principal series;}\\ \left|P_{i}(W_{1},W_{2})-\frac{q-3}{q-1}\right|\leq\frac{2}{q^{1/2+\operatorname{Re}(\gamma)}-1},&\text{\ if $i=c(\pi_{1}\times\pi_{2})/2=c/2$, $\pi_{i}$ are principal series;}\\ P_{i}(W_{1},W_{2})=-\frac{1}{q-1},&\text{\ if }i=c(\pi_{1}\times\pi_{2})/2-1>c/2;\\ |P_{i}(W_{1},W_{2})|\leq\frac{2}{q-1},&\text{\ if }i=c(\pi_{1}\times\pi_{2})/2-1=c/2;\\ P_{i}(W_{1},W_{2})=0,&\text{\ otherwise.}\par\end{cases}

When pp is large enough and c=2c=2, the first two lines of (5.1) still holds, while in the case c(π1×π2)=4c(\pi_{1}\times\pi_{2})=4 and i=1i=1, we have

P1(W1,W2)1q.P_{1}(W_{1},W_{2})\ll\frac{1}{q}.
Remark 5.5.

When i=c(π1×π2)/2=c/2i=c(\pi_{1}\times\pi_{2})/2=c/2 and πi\pi_{i} are principal series, the reason for difference is that some of the L2L^{2}-mass of WiW_{i} goes to v(x)>0v(x)>0.

Note that Assumption 5.3 also allows us to apply (4.8) and Corollary 4.23.

The main ingredient for proving the above proposition is the following result on the relation between Gauss sums/integrals, which may have independent interest.

Proposition 5.6.

Suppose that p2p\neq 2, c(πθ1)=c(πθ2)=c2c(\pi_{\theta_{1}})=c(\pi_{\theta_{2}})=c\geq 2, where πi=πθi\pi_{i}=\pi_{\theta_{i}} satisfy Assumption 5.3. Denote

(5.2) Pi(θ1,θ2)\displaystyle P_{i}(\theta_{1},\theta_{2})
=\displaystyle= c(χ)=civE(u)=c(θ)e𝔼+1θ11(u)χE(u)ψE(u)d×uvE(w)=c(θ)eE+1θ21(w)χE(w)ψE(w)d×w¯,\displaystyle\sum\limits_{c(\chi)=c-i}\int\limits_{v_{E}(u)=-c(\theta)-e_{\mathbb{E}}+1}\theta_{1}^{-1}(u)\chi_{E}(u)\psi_{E}(u)d^{\times}u\overline{\int\limits_{v_{E^{\prime}}(w)=-c(\theta)-e_{E^{\prime}}+1}\theta_{2}^{-1}(w)\chi_{E^{\prime}}(w)\psi_{E^{\prime}}(w)d^{\times}w},

and denote

Pi(θ1,θ2)\displaystyle P_{i}^{\circ}(\theta_{1},\theta_{2})
=\displaystyle= Pi(θ1,θ2){c(χ)=ci}vE(u)=c(θ)e𝔼+1θ11(u)ψE(u)d×uvE(w)=c(θ)eE+1θ21(w)ψE(w)d×w¯.\displaystyle\frac{P_{i}(\theta_{1},\theta_{2})}{\sharp\{c(\chi)=c-i\}\int\limits_{v_{E}(u)=-c(\theta)-e_{\mathbb{E}}+1}\theta_{1}^{-1}(u)\psi_{E}(u)d^{\times}u\overline{\int\limits_{v_{E^{\prime}}(w)=-c(\theta)-e_{E^{\prime}}+1}\theta_{2}^{-1}(w)\psi_{E^{\prime}}(w)d^{\times}w}}.

Then for p2p\neq 2 and ic/2>1i\geq c/2>1,

(5.3) {Pi(θ1,θ2)=1, if ic(π1×π2)/2, except when i=c/2 and  πi are principal series;Pi(θ1,θ2)=q3q1, if i=c(π1×π2)/2=c/2 and πi are principal series;Pi(θ1,θ2)=1q1, if i=c(π1×π2)/21>c/2;|Pi(θ1,θ2)|2q1, if i=c(π1×π2)/21=c/2;Pi(θ1,θ2)=0, otherwise.\begin{cases}P_{i}^{\circ}(\theta_{1},\theta_{2})=1,&\text{\ if $i\geq c(\pi_{1}\times\pi_{2})/2$, except when $i=c/2$ and }\\ &\text{\ \ $\pi_{i}$ are principal series;}\\ P_{i}^{\circ}(\theta_{1},\theta_{2})=\frac{q-3}{q-1},&\text{\ if $i=c(\pi_{1}\times\pi_{2})/2=c/2$ and $\pi_{i}$ are principal series;}\\ P_{i}^{\circ}(\theta_{1},\theta_{2})=-\frac{1}{q-1},&\text{\ if $i=c(\pi_{1}\times\pi_{2})/2-1>c/2$};\\ |P_{i}^{\circ}(\theta_{1},\theta_{2})|\leq\frac{2}{q-1},&\text{\ if }i=c(\pi_{1}\times\pi_{2})/2-1=c/2;\\ P_{i}^{\circ}(\theta_{1},\theta_{2})=0,&\text{\ otherwise.}\end{cases}

When pp is large enough and c=2c=2, the first two lines of (5.3) still holds, while in the case c(π1×π2)=4c(\pi_{1}\times\pi_{2})=4 and i=1i=1, we have

(5.4) Pi(θ1,θ2)1q.P_{i}^{\circ}(\theta_{1},\theta_{2})\ll\frac{1}{q}.

In the following, suppose that π1\pi_{1} is associated to a character θ1\theta_{1} over quadratic algebra EE, and π2\pi_{2} is associated to θ2\theta_{2} over EE^{\prime}. We break the proof into two main parts. The first part, consisting of Section 5.1, 5.2, 5.3, assume that c3c\geq 3 and discuss according to whether EEE\neq E^{\prime}, E=EE=E^{\prime} are fields, or E=EE=E^{\prime} split. One of the basic tools for these cases is the p-adic stationary phase analysis relying on Lemma 4.6.

The second part in Section 5.4 covers the case c=2c=2, where we can not apply Lemma 4.6 and need different but elementary approaches. It may also be possible to approach this case using methods from algebraic geometry.

In Section 5.5 we give partial results in the p=2p=2 case when c(πi)c(\pi_{i}) are large enough, which is however sufficient for later purposes.

5.1. Case EEE\neq E^{\prime}, c3c\geq 3.

Let C0C_{0} C0C_{0}^{\prime} be the corresponding constant for π1\pi_{1}, π2\pi_{2} as in Lemma 4.17 or Corollary 4.21.

In this case, according to the beginning of Section 4.2.2, either c(πi)c(\pi_{i}) is even, where one of EE, EE^{\prime} is inert field extension while the other one splits; or c(πi)c(\pi_{i}) is odd and EE EE^{\prime} are different ramified field extensions. Also by the condition c3c\geq 3, we have c(θi)2c(\theta_{i})\geq 2.

In either cases, by Lemma 4.17, the support of the integral for Pi(W1,W2)P_{i}(W_{1},W_{2}) is v(x)=0v(x)=0. Using the spectral decomposition on OF×O_{F}^{\times}, we get that

(5.5) Pi(W1,W2)=χxOF×W1(i)(x)χ(x)d×xyOF×W2(i)(y)χ(y)d×y¯.P_{i}(W_{1},W_{2})=\sum\limits_{\chi}\int\limits_{x\in O_{F}^{\times}}W_{1}^{(i)}(x)\chi(x)d^{\times}x\overline{\int\limits_{y\in O_{F}^{\times}}W_{2}^{(i)}(y)\chi(y)d^{\times}y}.

By Lemma 4.17, the sum in χ\chi is over those with c(χ)=cic(\chi)=c-i when ic1i\neq c-1, and c(χ)1c(\chi)\leq 1 when i=c1i=c-1.

5.1.1. Case i=ci=c

In this case, it is well known that Wj(c)W_{j}^{(c)} are both characteristic functions on OF×O_{F}^{\times}, thus Pc(W1,W2)=1P_{c}(W_{1},W_{2})=1.

5.1.2. Case i=c1i=c-1

Note that the contributions come from c(χ)1c(\chi)\leq 1. Using Lemma 4.4, Corollary 4.18, 4.22, we have

Pc1(W1,W2)=\displaystyle P_{c-1}(W_{1},W_{2})= 1C0C0¯[1(q1)2Pc(θ1,θ2)+q(q1)2Pc1(θ1,θ2)].\displaystyle\frac{1}{C_{0}\overline{C_{0}^{\prime}}}\left[\frac{1}{(q-1)^{2}}P_{c}(\theta_{1},\theta_{2})+\frac{q}{(q-1)^{2}}P_{c-1}(\theta_{1},\theta_{2})\right].

Let αi\alpha_{i} be associated to θi\theta_{i} using Lemma 4.6. Using the definition for C0,C0C_{0},C_{0}^{\prime}, formula (5.2) and that c(θi)2c(χE)c(\theta_{i})\geq 2c(\chi_{E}), we have by Lemma 4.8,

Pc1(W1,W2)=1(q1)2+q(q1)2c(χ)=1χ(NE/F(α1)NE/F(α2))=1q1.\displaystyle P_{c-1}(W_{1},W_{2})=\frac{1}{(q-1)^{2}}+\frac{q}{(q-1)^{2}}\sum\limits_{c(\chi)=1}\chi\left(\frac{N_{E/F}(\alpha_{1})}{N_{E^{\prime}/F}(\alpha_{2})}\right)=-\frac{1}{q-1}.

Here we have used that NE/F(α1)NE/F(α2)1\frac{N_{E/F}(\alpha_{1})}{N_{E^{\prime}/F}(\alpha_{2})}\not\equiv 1 by Lemma 4.3.

5.1.3. Case c/2i<c1c/2\leq i<c-1, c3c\geq 3.

Recall the notation that χ1jχ2\chi_{1}\sim_{j}\chi_{2} if and only if c(χ1χ21)jc(\chi_{1}\chi_{2}^{-1})\leq j.

Consider first the case c/2<i<c1c/2<i<c-1. Using Lemma 4.4, Corollary 4.18, 4.22, we have

Pi(W1,W2)=1C0C0¯|ψχ|2Pi(θ1,θ2)=1C0C0¯(q1)2qci2Pi(θ1,θ2).\displaystyle P_{i}(W_{1},W_{2})=\frac{1}{C_{0}\overline{C_{0}^{\prime}}}\left|\int\psi\chi\right|^{2}P_{i}(\theta_{1},\theta_{2})=\frac{1}{C_{0}\overline{C_{0}^{\prime}}(q-1)^{2}q^{c-i-2}}P_{i}(\theta_{1},\theta_{2}).

We break the sum over χ\chi in Pi(θ1,θ2)P_{i}(\theta_{1},\theta_{2}) into a double sum of χ0\chi_{0} over the set of characters of level cic-i modulo 1\sim_{1}, and the sum of characters η\eta of level 1\leq 1. To each χ0\chi_{0}, we associate α0\alpha_{0} as in Lemma 4.6. Then using Lemma 4.8, Pi(θ1,θ2)P_{i}(\theta_{1},\theta_{2}) can be rewritten as

χ0/1c(η)1η(NE/F(α1+α0)NE/F(α2+α0))\displaystyle\sum\limits_{\chi_{0}/\sim_{1}}\sum\limits_{c(\eta)\leq 1}\eta\left(\frac{N_{E/F}(\alpha_{1}+\alpha_{0})}{N_{E^{\prime}/F}(\alpha_{2}+\alpha_{0})}\right) vE(u)=c(θ)e𝔼+1θ11(u)χ0,E(u)ψ𝔼(u)d×u\displaystyle\int\limits_{v_{E}(u)=-c(\theta)-e_{\mathbb{E}}+1}\theta_{1}^{-1}(u)\chi_{0,E}(u)\psi_{{\mathbb{E}}}(u)d^{\times}u
×\displaystyle\times v𝔼(u)=c(θ)e𝔼+1θ21(u)χ0,𝔼(u)ψ𝔼(u)d×u¯.\displaystyle\overline{\int\limits_{v_{{\mathbb{E}}^{\prime}}(u)=-c(\theta)-e_{{\mathbb{E}}^{\prime}}+1}\theta_{2}^{-1}(u)\chi_{0,{\mathbb{E}}^{\prime}}(u)\psi_{{\mathbb{E}}^{\prime}}(u)d^{\times}u}.

Using Lemma 4.3, the inner sum in η\eta is vanishing, thus Pi(W1,W2)=0P_{i}(W_{1},W_{2})=0 in this case.

Now consider the case i=c/2i=c/2. In this case, we suppose without loss of generality that EE is split and EE^{\prime} is inert. The expression for xOF×W1(i)(x)χ(x)d×x\int\limits_{x\in O_{F}^{\times}}W_{1}^{(i)}(x)\chi(x)d^{\times}x is a little more complicated according to Corollary 4.22, but since W2W_{2} is supported only at v(x)=0v(x)=0, we only need to consider the case k=j=0k=j=0 in Corollary 4.22. The computation is then similar to the c/2<i<c1c/2<i<c-1 case, and we have again Pi(W1,W2)=0P_{i}(W_{1},W_{2})=0.

5.2. Case E=EE=E^{\prime} is a field extension, c3c\geq 3.

Let πj\pi_{j} be associated to θj\theta_{j} over the same quadratic field extension EE, with c(θ1)=c(θ2)c(\theta_{1})=c(\theta_{2}). Note that θ(x)¯=θ1(x)\overline{\theta(x)}=\theta^{-1}(x) by the condition that θ|F×=1\theta|_{F^{\times}}=1. Without loss of generality, we assume that

l:=c(θ2θ11)c(θ1θ2).l:=c(\theta_{2}\theta_{1}^{-1})\leq c(\theta_{1}\theta_{2}).

5.2.1. Case ic(π1×π2)/2i\geq c(\pi_{1}\times\pi_{2})/2

If l=0l=0, we have

Pi(θ1,θ2)\displaystyle P_{i}(\theta_{1},\theta_{2})
=\displaystyle= c(χ)=civE(u)=c(θ)e𝔼+1θ11(u)χE(u)ψE(u)d×uvE(w)=c(θ)eE+1θ21(w)χE(w)ψE(w)d×w¯\displaystyle\sum\limits_{c(\chi)=c-i}\int\limits_{v_{E}(u)=-c(\theta)-e_{\mathbb{E}}+1}\theta_{1}^{-1}(u)\chi_{E}(u)\psi_{E}(u)d^{\times}u\overline{\int\limits_{v_{E^{\prime}}(w)=-c(\theta)-e_{E^{\prime}}+1}\theta_{2}^{-1}(w)\chi_{E^{\prime}}(w)\psi_{E^{\prime}}(w)d^{\times}w}
=\displaystyle= c(χ)=ciθ2θ11(ϖE)c(θ)eE+1|vE(u)=c(θ)e𝔼+1θ11(u)χE(u)ψE(u)d×u|2.\displaystyle\sum\limits_{c(\chi)=c-i}\theta_{2}\theta_{1}^{-1}(\varpi_{E})^{-c(\theta)-e_{E}+1}\left|\ \int\limits_{v_{E}(u)=-c(\theta)-e_{\mathbb{E}}+1}\theta_{1}^{-1}(u)\chi_{E}(u)\psi_{E}(u)d^{\times}u\right|^{2}.

There is no cancellation and one can easily get that

Pi(W1,W2)=1.P_{i}(W_{1},W_{2})=1.

Consider the case l>0l>0 now. We can rewrite Pi(θ1,θ2)P_{i}(\theta_{1},\theta_{2}) as

(5.6) Pi(θ1,θ2)\displaystyle P_{i}(\theta_{1},\theta_{2})
=\displaystyle= c(χ)=civE(u),vE(w)=c(θ1)e𝔼+1θ11(u)θ2(w)χE(uw)ψE(uw)d×ud×w\displaystyle\sum\limits_{c(\chi)=c-i}\iint\limits_{v_{E}(u),v_{E}(w)=-c(\theta_{1})-e_{\mathbb{E}}+1}\theta_{1}^{-1}(u)\theta_{2}(w)\chi_{E}\left(\frac{u}{w}\right)\psi_{E}(u-w)d^{\times}ud^{\times}w
=\displaystyle= c(χ)=cixOE×,vE(w)=c(θ1)e𝔼+1θ11(x)θ2θ11(w)χE(x)ψE((x1)w)d×wd×x.\displaystyle\sum\limits_{c(\chi)=c-i}\iint\limits_{x\in O_{E}^{\times},v_{E}(w)=-c(\theta_{1})-e_{\mathbb{E}}+1}\theta_{1}^{-1}(x)\theta_{2}\theta_{1}^{-1}(w)\chi_{E}\left(x\right)\psi_{E}((x-1)w)d^{\times}wd^{\times}x.

As c(θ2θ11)=lc(\theta_{2}\theta_{1}^{-1})=l, the inner integral in ww is non-vanishing if and only if x1+ϖEc(θ1)lOE×x\in 1+\varpi_{E}^{c(\theta_{1})-l}O_{E}^{\times}. Writing x=1+ϖEc(θ1)lyx=1+\varpi_{E}^{c(\theta_{1})-l}y for yOE×y\in O_{E}^{\times}, qE=|ϖE|E1q_{E}=|\varpi_{E}|_{E}^{-1},

(5.7) Pi(θ1,θ2)\displaystyle P_{i}(\theta_{1},\theta_{2})
=\displaystyle= θ1θ21(ϖEc(θ1)l)qEc(θ1)lvE(w)=leE+1θ2θ11(w)ψE(w)d×w\displaystyle\frac{\theta_{1}\theta_{2}^{-1}(\varpi_{E}^{c(\theta_{1})-l})}{q_{E}^{c(\theta_{1})-l}}\int\limits_{v_{E}(w)=-l-e_{E}+1}\theta_{2}\theta_{1}^{-1}(w)\psi_{E}(w)d^{\times}w
c(χ)=ciyOE×θ11χE(1+ϖEc(θ1)ly)θ1θ21(y)d×y.\displaystyle\sum\limits_{c(\chi)=c-i}\int\limits_{y\in O_{E}^{\times}}\theta_{1}^{-1}\chi_{E}(1+\varpi_{E}^{c(\theta_{1})-l}y)\theta_{1}\theta_{2}^{-1}(y)d^{\times}y.

Note that c(χE)=eEc(χ)eE+1c(\chi_{E})=e_{E}c(\chi)-e_{E}+1, so when

(5.8) eE(ci)eE+1c(θ1)l,e_{E}(c-i)-e_{E}+1\leq c(\theta_{1})-l,

we have

Pi(θ1,θ2)=\displaystyle P_{i}(\theta_{1},\theta_{2})= θ1θ21(ϖEc(θ1)l)qEc(θ1)lvE(w)=leE+1θ2θ11(w)ψE(w)d×w\displaystyle\frac{\theta_{1}\theta_{2}^{-1}(\varpi_{E}^{c(\theta_{1})-l})}{q_{E}^{c(\theta_{1})-l}}\int\limits_{v_{E}(w)=-l-e_{E}+1}\theta_{2}\theta_{1}^{-1}(w)\psi_{E}(w)d^{\times}w
c(χ)=ciyOE×θ11(1+ϖEc(θ1)ly)θ1θ21(y)d×y\displaystyle\sum\limits_{c(\chi)=c-i}\int\limits_{y\in O_{E}^{\times}}\theta_{1}^{-1}(1+\varpi_{E}^{c(\theta_{1})-l}y)\theta_{1}\theta_{2}^{-1}(y)d^{\times}y
=\displaystyle= C0C0¯(q1)2qci2.\displaystyle C_{0}\overline{C_{0}^{\prime}}(q-1)^{2}q^{c-i-2}.

Here we arrived at the second line by working backward to (5.6) without the twists by χ\chi. The condition (5.8) translates into the following: when eE=1e_{E}=1, we have ic(θ1)+li\geq c(\theta_{1})+l; when eE=2e_{E}=2, we have ic(θ1)+l+22i\geq\frac{c(\theta_{1})+l+2}{2} as both c(θ1)c(\theta_{1}) and ll must be even. In both cases, we can rewrite the condition as

ic(π1×π2)2.i\geq\frac{c(\pi_{1}\times\pi_{2})}{2}.

We can then proceed as in EEE\neq E^{\prime} case to get

Pi(W1,W2)=1P_{i}(W_{1},W_{2})=1

when ic(πθ1×πθ2)2i\geq\frac{c(\pi_{\theta_{1}}\times\pi_{\theta_{2}})}{2}.

5.2.2. Case c/2i<c(π1×π2)/2c/2\leq i<c(\pi_{1}\times\pi_{2})/2

There are many possible situations in this case. We shall discuss case by case according to l=c(θ2θ11)c(θ1θ2)l=c(\theta_{2}\theta_{1}^{-1})\leq c(\theta_{1}\theta_{2}).

Note that l=0l=0 does not happen as there will not be an integer ii with c/2i<c(π1×π2)/2c/2\leq i<c(\pi_{1}\times\pi_{2})/2 in this case.

  1. (1)

    Consider first the case where l=c(θ1)l=c(\theta_{1}). By the assumption c(θ2θ11)c(θ1θ2)c(\theta_{2}\theta_{1}^{-1})\leq c(\theta_{1}\theta_{2}), this implies that α1±α2\alpha_{1}\not\equiv\pm\alpha_{2}. As a result, we have NE/F(α1)NE/F(α2)1\frac{N_{E/F}(\alpha_{1})}{N_{E^{\prime}/F}(\alpha_{2})}\not\equiv 1 and NE/F(α1+α0)NE/F(α2+α0)1\frac{N_{E/F}(\alpha_{1}+\alpha_{0})}{N_{E^{\prime}/F}(\alpha_{2}+\alpha_{0})}\not\equiv 1, and the result follows similarly as in Section 5.1.2, 5.1.3.

  2. (2)

    Suppose now l=1l=1, then automatically eE=1e_{E}=1, and i=c/2i=c/2 in this case. From (5.7), we break the sum over χ\chi into a double sum of χ0\chi_{0} modulo c/21\sim_{c/2-1}, and the sum of characters η\eta of level c/21\leq c/2-1, with χ=χ0η\chi=\chi_{0}\eta. For each such χ0\chi_{0}, we can associated α0\alpha_{0} by Lemma 4.6, then the sum in χ0\chi_{0} corresponds to a sum in α0ϖc/2OF×modϖc/2+1OF\alpha_{0}\in\varpi^{-c/2}O_{F}^{\times}\mod\varpi^{-c/2+1}O_{F}. The last line of (5.7) becomes

    c(χ)=ciyOE×θ11χE(1+ϖEc(θ1)ly)θ1θ21(y)d×y\displaystyle\sum\limits_{c(\chi)=c-i}\int\limits_{y\in O_{E}^{\times}}\theta_{1}^{-1}\chi_{E}(1+\varpi_{E}^{c(\theta_{1})-l}y)\theta_{1}\theta_{2}^{-1}(y)d^{\times}y
    =(q1)qc/22χ0yOE×ψE((α1+α0)ϖEc(θ1)ly)θ1θ21(y)d×y\displaystyle=(q-1)q^{c/2-2}\sum\limits_{\chi_{0}}\int\limits_{y\in O_{E}^{\times}}\psi_{E}((-\alpha_{1}+\alpha_{0})\varpi_{E}^{c(\theta_{1})-l}y)\theta_{1}\theta_{2}^{-1}(y)d^{\times}y
    =(q1)qc/22yϖE1OE×ψE(y)θ1θ21(y)d×ya0θ2θ11((α1+α0)ϖEc(θ1)1).\displaystyle=(q-1)q^{c/2-2}\int\limits_{y\in\varpi_{E}^{-1}O_{E}^{\times}}\psi_{E}(y)\theta_{1}\theta_{2}^{-1}(y)d^{\times}y\sum\limits_{a_{0}}\theta_{2}\theta_{1}^{-1}\left((-\alpha_{1}+\alpha_{0})\varpi_{E}^{c(\theta_{1})-1}\right).

    Here in the second equality we have used Corollary 4.7, and that the integral is independent of η\eta. Using Lemma 4.4, 4.5, we have

    |c(χ)=ciyOE×θ11χE(1+ϖEc(θ1)ly)θ1θ21(y)d×y|2qc/21q+1,\left|\sum\limits_{c(\chi)=c-i}\int\limits_{y\in O_{E}^{\times}}\theta_{1}^{-1}\chi_{E}(1+\varpi_{E}^{c(\theta_{1})-l}y)\theta_{1}\theta_{2}^{-1}(y)d^{\times}y\right|\leq\frac{2q^{c/2-1}}{q+1},
    |Pc/2(θ1,θ2)|2q2(q+1)(q21)qc/2,\left|P_{c/2}(\theta_{1},\theta_{2})\right|\leq\frac{2q^{2}}{(q+1)(q^{2}-1)q^{c/2}},
    |Pc/2(W1,W2)|\displaystyle\left|P_{c/2}(W_{1},W_{2})\right| =|1C0C0¯(q1)2qci2Pc/2(θ1,θ2)|\displaystyle=\left|\frac{1}{C_{0}\overline{C_{0}^{\prime}}(q-1)^{2}q^{c-i-2}}P_{c/2}(\theta_{1},\theta_{2})\right|
    (1q2)2qc(q1)2qc/222q2(q+1)(q21)qc/2=2q1.\displaystyle\leq\frac{(1-q^{-2})^{2}q^{c}}{(q-1)^{2}q^{c/2-2}}\frac{2q^{2}}{(q+1)(q^{2}-1)q^{c/2}}=\frac{2}{q-1}.
  3. (3)

    If 1<l<c(θ1)1<l<c(\theta_{1}), we write y=y0(1+y1)y=y_{0}(1+y_{1}) for y0OE×/1+ϖEl1OEy_{0}\in O_{E}^{\times}/1+\varpi_{E}^{l-1}O_{E}, y1ϖEl1OEy_{1}\in\varpi_{E}^{l-1}O_{E}. Let α3\alpha_{3} be the constant associated to θ1θ21\theta_{1}\theta_{2}^{-1}. Then

    yOE×θ11χE(1+ϖEc(θ1)ly)θ1θ21(y)d×y\displaystyle\int\limits_{y\in O_{E}^{\times}}\theta_{1}^{-1}\chi_{E}(1+\varpi_{E}^{c(\theta_{1})-l}y)\theta_{1}\theta_{2}^{-1}(y)d^{\times}y
    =\displaystyle= 11qE1y0OE×/1+ϖEl1OEθ11χE(1+ϖEc(θ1)ly0)θ1θ21(y0)\displaystyle\frac{1}{1-q_{E}^{-1}}\sum\limits_{y_{0}\in O_{E}^{\times}/1+\varpi_{E}^{l-1}O_{E}}\theta_{1}^{-1}\chi_{E}(1+\varpi_{E}^{c(\theta_{1})-l}y_{0})\theta_{1}\theta_{2}^{-1}(y_{0})
    y1ϖEl1OEψE((α1+α0)ϖEc(θ1)ly0y11+ϖEc(θ1)ly0+α3y1)𝑑y1.\displaystyle\int\limits_{y_{1}\in\varpi_{E}^{l-1}O_{E}}\psi_{E}\left((-\alpha_{1}+\alpha_{0})\frac{\varpi_{E}^{c(\theta_{1})-l}y_{0}y_{1}}{1+\varpi_{E}^{c(\theta_{1})-l}y_{0}}+\alpha_{3}y_{1}\right)dy_{1}.

    The integral in y1y_{1} is nonvanishing only if y0y_{0} satisfies

    (α1+α0)ϖEc(θ1)ly01+ϖEc(θ1)ly0+α30modϖEleE+2.(-\alpha_{1}+\alpha_{0})\frac{\varpi_{E}^{c(\theta_{1})-l}y_{0}}{1+\varpi_{E}^{c(\theta_{1})-l}y_{0}}+\alpha_{3}\equiv 0\mod\varpi_{E}^{-l-e_{E}+2}.

    Note that this is only possible if c(χE)c(θ1)c(\chi_{E})\leq c(\theta_{1}). Then vE(α1+α0)=c(θ1)eE+1,v_{E}(-\alpha_{1}+\alpha_{0})=-c(\theta_{1})-e_{E}+1, vE(α3)=leE+1>c(θ1)eE+1v_{E}(\alpha_{3})=-l-e_{E}+1>-c(\theta_{1})-e_{E}+1. As c(θ1)>lc(\theta_{1})>l we get

    y0α3(α1α0)ϖEc(θ1)lmodϖE.y_{0}\equiv\frac{\alpha_{3}}{(\alpha_{1}-\alpha_{0})\varpi_{E}^{c(\theta_{1})-l}}\mod\varpi_{E}.

    As α1,α3E\alpha_{1},\alpha_{3}\in E are trace 0 elements and α0F\alpha_{0}\in F, we have

    (5.9) NE/F(1+ϖEc(θ1)ly0)1+2α3α11modϖ(c(θ1)l)/eE+1.N_{E/F}(1+\varpi_{E}^{c(\theta_{1})-l}y_{0})\equiv 1+\frac{2\alpha_{3}}{\alpha_{1}}\not\equiv 1\mod\varpi^{(c(\theta_{1})-l)/e_{E}+1}.

    We can then rewrite (5.7) as follows by imposing the condition on y0y_{0}

    Pi(θ1,θ2)=\displaystyle P_{i}(\theta_{1},\theta_{2})= θ1θ21(ϖEc(θ1)l)qEc(θ1)lwθ2θ11ψEc(χ)=ci\displaystyle\frac{\theta_{1}\theta_{2}^{-1}(\varpi_{E}^{c(\theta_{1})-l})}{q_{E}^{c(\theta_{1})-l}}\int\limits_{w}\theta_{2}\theta_{1}^{-1}\psi_{E}\sum\limits_{c(\chi)=c-i}
    yα3(α1α0)ϖEc(θ1)lmodϖEθ11χE(1+ϖEc(θ1)ly)θ1θ21(y)d×y.\displaystyle\int\limits_{y\equiv\frac{\alpha_{3}}{(\alpha_{1}-\alpha_{0})\varpi_{E}^{c(\theta_{1})-l}}\mod\varpi_{E}}\theta_{1}^{-1}\chi_{E}(1+\varpi_{E}^{c(\theta_{1})-l}y)\theta_{1}\theta_{2}^{-1}(y)d^{\times}y.
    1. (3i)

      If ci>(c(θ1)l)/eE+1c-i>(c(\theta_{1})-l)/e_{E}+1, which is equivalent to i<c(π1×π2)/21i<c(\pi_{1}\times\pi_{2})/2-1, we can break the sum of χ=χ0η\chi=\chi_{0}\eta into the sum of χ0\chi_{0} modulo (c(θ1)l)/eE+1\sim_{(c(\theta_{1})-l)/e_{E}+1}, and the sum of η\eta with c(η)(c(θ1)l)/eE+1c(\eta)\leq(c(\theta_{1})-l)/e_{E}+1. Note that the domain yα3(α1α0)ϖEc(θ1)lmodϖEy\equiv\frac{\alpha_{3}}{(\alpha_{1}-\alpha_{0})\varpi_{E}^{c(\theta_{1})-l}}\mod\varpi_{E} is independent of η\eta. Then the sum in η\eta first is vanishing for any yα3(α1α0)ϖEc(θ1)lmodϖEy\equiv\frac{\alpha_{3}}{(\alpha_{1}-\alpha_{0})\varpi_{E}^{c(\theta_{1})-l}}\mod\varpi_{E} using (5.9), and

      Pi(W1,W2)=Pi(θ1,θ2)=0.P_{i}(W_{1},W_{2})=P_{i}(\theta_{1},\theta_{2})=0.
    2. (3ii)

      If ci=(c(θ1)l)/eE+1c-i=(c(\theta_{1})-l)/e_{E}+1, which is equivalent to that i=c(π1×π2)/21i=c(\pi_{1}\times\pi_{2})/2-1, then the domain of integral is actually yα3α1ϖEc(θ1)lmodϖEy\equiv\frac{\alpha_{3}}{\alpha_{1}\varpi_{E}^{c(\theta_{1})-l}}\mod\varpi_{E}. We take the sum in χ\chi first. Using that

      c(χ)ciχE(1+α3α1)=0\sum\limits_{c(\chi)\leq c-i}\chi_{E}\left(1+\frac{\alpha_{3}}{\alpha_{1}}\right)=0

      in this case, and χE(1+ϖEc(θ1)ly)=1\chi_{E}(1+\varpi_{E}^{c(\theta_{1})-l}y)=1 when c(χ)<cic(\chi)<c-i, we obtain that

      Pi(θ1,θ2)\displaystyle P_{i}(\theta_{1},\theta_{2})
      =\displaystyle= c(χ)<ciθ1θ21(ϖEc(θ1)l)qc(θ1)lwθ2θ11ψEyα3α1ϖEc(θ1)lmodϖEθ11(1+ϖEc(θ1)ly)θ1θ21(y)d×y.\displaystyle-\sum\limits_{c(\chi)<c-i}\frac{\theta_{1}\theta_{2}^{-1}(\varpi_{E}^{c(\theta_{1})-l})}{q^{c(\theta_{1})-l}}\int\limits_{w}\theta_{2}\theta_{1}^{-1}\psi_{E}\int\limits_{y\equiv\frac{\alpha_{3}}{\alpha_{1}\varpi_{E}^{c(\theta_{1})-l}}\mod\varpi_{E}}\theta_{1}^{-1}(1+\varpi_{E}^{c(\theta_{1})-l}y)\theta_{1}\theta_{2}^{-1}(y)d^{\times}y.

      By working backwards without χ\chi twists, we get

      Pi(W1,W2)=1q1.P_{i}(W_{1},W_{2})=-\frac{1}{q-1}.

5.3. Case E=EE=E^{\prime} splits

Most of the discussions in Section 5.2 hold for the split case. The only difference is the case i=c/2i=c/2 due to the slightly more complicated expression in Corollary 4.22. In that case, we have

Pc/2(W1,W2)\displaystyle P_{c/2}(W_{1},W_{2})
=\displaystyle= k0q(1/2γ)kc(χ)=c/2xOF×W1(c/2)(ϖkx)χ(x)d×xyOF×W2(c/2)(ϖky)χ(y)d×y¯\displaystyle\sum\limits_{k\geq 0}q^{(1/2-\gamma)k}\sum\limits_{c(\chi)=c/2}\int\limits_{x\in O_{F}^{\times}}W_{1}^{(c/2)}(\varpi^{k}x)\chi(x)d^{\times}x\overline{\int\limits_{y\in O_{F}^{\times}}W_{2}^{(c/2)}(\varpi^{k}y)\chi(y)d^{\times}y}
=\displaystyle= k0q(1/2γ)kqkC0C0¯(q1)2qc/22c(χ)=c/2i,j=0,kμ1(ϖk2i)μ2(ϖk2j)¯\displaystyle\sum\limits_{k\geq 0}q^{(1/2-\gamma)k}\frac{q^{-k}}{C_{0}\overline{C_{0}^{\prime}}(q-1)^{2}q^{c/2-2}}\sum\limits_{c(\chi)=c/2}\sum\limits_{i,j=0,k}\mu_{1}(\varpi^{k-2i})\overline{\mu_{2}(\varpi^{k-2j})}
uϖEc0OE×θ11(u)χE(u)ψE(ϖki,iu)d×uwϖEc0OE×θ21(w)χE(w)ψE(ϖkj,jw)d×w¯.\displaystyle\int\limits_{u\in\varpi_{E}^{-c_{0}}O_{E}^{\times}}\theta_{1}^{-1}(u)\chi_{E}(u)\psi_{E}(\varpi^{k-i,i}u)d^{\times}u\overline{\int\limits_{w\in\varpi_{E}^{-c_{0}}O_{E}^{\times}}\theta_{2}^{-1}(w)\chi_{E}(w)\psi_{E}(\varpi^{k-j,j}w)d^{\times}w}.

Here θ1=(μ11,μ1)\theta_{1}=(\mu_{1}^{-1},\mu_{1}), θ2=(μ21,μ2)\theta_{2}=(\mu_{2}^{-1},\mu_{2}). Denote ϖE=(ϖ,ϖ)\varpi_{E}=(\varpi,\varpi). Denote by Pc/2,k(W1,W2)P_{c/2,k}(W_{1},W_{2}) the corresponding summand in Pc/2(W1,W2)P_{c/2}(W_{1},W_{2}) for any fixed k0k\geq 0.

5.3.1. Case l=0l=0

If l=0l=0, then C0=C0C_{0}=C_{0}^{\prime},

uϖEc0OE×θ21(u)χE(u)ψE(ϖki,iu)d×u\displaystyle\int\limits_{u\in\varpi_{E}^{-c_{0}}O_{E}^{\times}}\theta_{2}^{-1}(u)\chi_{E}(u)\psi_{E}(\varpi^{k-i,i}u)d^{\times}u
=\displaystyle= θ2θ11(ϖE)c0wϖEc0OE×θ11(w)χE(w)ψE(ϖki,iw)d×w.\displaystyle\theta_{2}\theta_{1}^{-1}(\varpi_{E})^{c_{0}}\int\limits_{w\in\varpi_{E}^{-c_{0}}O_{E}^{\times}}\theta_{1}^{-1}(w)\chi_{E}(w)\psi_{E}(\varpi^{k-i,i}w)d^{\times}w.

There is no cancellation in the summation over χ\chi for Pc/2,k(θ1,θ2)P_{c/2,k}(\theta_{1},\theta_{2}). Thus when k=0k=0, we have by Corollary 4.23,

Pc/2,0(W1,W2)=L2(W1,0)=q3q1,P_{c/2,0}(W_{1},W_{2})=L^{2}(W_{1},0)=\frac{q-3}{q-1},

When k>0k>0, we use Corollary 4.23 and Cauchy–Schwarz inequality to get

|Pc/2,k(W1,W2)|2q1/2kqRe(γ)k.\left|P_{c/2,k}(W_{1},W_{2})\right|\leq\frac{2}{q^{1/2k}}q^{-\operatorname{Re}(\gamma)k}.

Adding up all k0k\geq 0, we get

|Pc/2(W1,W2)q3q1|2q1/2+Re(γ)1.\left|P_{c/2}(W_{1},W_{2})-\frac{q-3}{q-1}\right|\leq\frac{2}{q^{1/2+\operatorname{Re}(\gamma)}-1}.

In particular when qq is large enough, we have

Pc/2(W1,W2)1.P_{c/2}(W_{1},W_{2})\gg 1.

5.3.2. Case l>1l>1.

As π(μ,μ1)π(μ1,μ)\pi(\mu,\mu^{-1})\simeq\pi(\mu^{-1},\mu), we may assume without loss of generality that c(μ1μ21)=lc(\mu_{1}\mu_{2}^{-1})=l.

  1. (1)

    One can similarly relate Pc/2,0(W1,W2)P_{c/2,0}(W_{1},W_{2}) with Pc/2(θ1,θ2)P_{c/2}(\theta_{1},\theta_{2}) as in Section 5.2.2, with the difference that the contribution to Pc/2,0(W1,W2)P_{c/2,0}(W_{1},W_{2}) comes from χ\chi with c(χμ1)=c(χμ11)=c(χμ2)=c(χμ21)=c/2c(\chi\mu_{1})=c(\chi\mu_{1}^{-1})=c(\chi\mu_{2})=c(\chi\mu_{2}^{-1})=c/2. Since c/2>1c/2>1 by our condition, the method for the case (1) and (3i) in Section 5.2.2 applies here and

    Pc/2,0(W1,W2)=0.P_{c/2,0}(W_{1},W_{2})=0.
  2. (2)

    When 0<kc/2l0<k\leq c/2-l, we use similar strategy as in Section 5.2.2 with slight modifications. Note that the case l=c/2l=c/2 is excluded here. By the proof of Corollary 4.23, we can separate the contribution of χ\chi with c(μi1χ)=c/2k<c/2c(\mu_{i}^{-1}\chi)=c/2-k<c/2, from that of χ\chi with c(μiχ)<c/2c(\mu_{i}\chi)<c/2. Since these two cases are parallel, we focus on the case c(μi1χ)<c/2c(\mu_{i}^{-1}\chi)<c/2, and consider

    (5.10) Pc/2,k(θ1,θ2)\displaystyle P_{c/2,k}(\theta_{1},\theta_{2})
    :=\displaystyle:= χuϖEc0OE×θ11(u)χE(u)ψE(ϖ0,ku)d×uwϖEc0OE×θ21(w)χE(w)ψE(ϖ0,kw)d×w¯.\displaystyle\sum\limits_{\chi}\int\limits_{u\in\varpi_{E}^{-c_{0}}O_{E}^{\times}}\theta_{1}^{-1}(u)\chi_{E}(u)\psi_{E}(\varpi^{0,k}u)d^{\times}u\overline{\int\limits_{w\in\varpi_{E}^{-c_{0}}O_{E}^{\times}}\theta_{2}^{-1}(w)\chi_{E}(w)\psi_{E}(\varpi^{0,k}w)d^{\times}w}.
    =\displaystyle= χxOE×,wϖEc0OE×θ11(x)θ2θ11(w)χE(x)ψE(ϖ0,k(x1)w)d×wd×x.\displaystyle\sum\limits_{\chi}\iint\limits_{x\in O_{E}^{\times},w\in\varpi_{E}^{-c_{0}}O_{E}^{\times}}\theta_{1}^{-1}(x)\theta_{2}\theta_{1}^{-1}(w)\chi_{E}\left(x\right)\psi_{E}(\varpi^{0,k}(x-1)w)d^{\times}wd^{\times}x.

    Here the sum in χ\chi is for those with χ(ϖ)=1\chi(\varpi)=1, c(χ)=c/2c(\chi)=c/2 and c(μi1χ)=c/2kc(\mu_{i}^{-1}\chi)=c/2-k.

    For the integral in ww to be nonvanishing, we get that x1+ϖc/2l,c/2klOE×x\in 1+\varpi^{c/2-l,c/2-k-l}O_{E}^{\times}. Then as in Section 5.2.2, we write x=1+ϖc/2l,c/2klyx=1+\varpi^{c/2-l,c/2-k-l}y with yOE×y\in O_{E}^{\times} when k<c/2lk<c/2-l, or y=(y1,y2)y=(y_{1},y_{2}) for yiOF×y_{i}\in O_{F}^{\times}, y21modϖy_{2}\not\equiv-1\mod\varpi when k=c/2lk=c/2-l. We can uniformly describe the domain for yy as y=(y1,y2)OE×,y2ϖc/2kly=(y_{1},y_{2})\in O_{E}^{\times},y_{2}\not\equiv-\varpi^{c/2-k-l}. Then

    (5.11) Pc/2,k(θ1,θ2)=\displaystyle P_{c/2,k}(\theta_{1},\theta_{2})= θ1θ21(ϖEc/2l)qc2lkwϖElOE×θ2θ11(w)ψE(w)d×wc(μi1χ)=c/2k\displaystyle\frac{\theta_{1}\theta_{2}^{-1}(\varpi_{E}^{c/2-l})}{q^{c-2l-k}}\int\limits_{w\in\varpi_{E}^{-l}O_{E}^{\times}}\theta_{2}\theta_{1}^{-1}(w)\psi_{E}(w)d^{\times}w\sum\limits_{c(\mu_{i}^{-1}\chi)=c/2-k}
    y=(y1,y2)OE×,y2ϖc/2klθ11χE(1+ϖc/2l,c/2kly)θ1θ21(y)d×y.\displaystyle\int\limits_{y=(y_{1},y_{2})\in O_{E}^{\times},y_{2}\not\equiv-\varpi^{c/2-k-l}}\theta_{1}^{-1}\chi_{E}(1+\varpi^{c/2-l,c/2-k-l}y)\theta_{1}\theta_{2}^{-1}(y)d^{\times}y.

    Unlike case (3) in Section 5.2.2, we do not need stationary phase analysis in this case to see cancellations. Indeed one can take the sum in χ\chi first and break the sum of χ=χ0χ1\chi=\chi_{0}\chi_{1} into a sum over χ0\chi_{0} up to c/2kl+1\sim_{c/2-k-l+1}, and a sum over χ1\chi_{1} with c(χ1)c/2kl+1c(\chi_{1})\leq c/2-k-l+1. The sum in χ1\chi_{1} first gives

    χ1χ1,E(1+ϖc/2l,c/2kly)=0,yOE×.\sum\limits_{\chi_{1}}\chi_{1,E}(1+\varpi^{c/2-l,c/2-k-l}y)=0,\forall y\in O_{E}^{\times}.

    Note that the first component is always trivial as c/2lc/2kl+1c(χ1)c/2-l\geq c/2-k-l+1\geq c(\chi_{1}). The cancellation comes from the second component. Thus when 0<kc/2l0<k\leq c/2-l we get

    Pc/2,k(W1,W2)=0.P_{c/2,k}(W_{1},W_{2})=0.
  3. (3)

    If k>c/2lk>c/2-l and c(μ1μ21)=lc(\mu_{1}\mu_{2}^{-1})=l, we can not simultaneously have c(μ1χ)c(\mu_{1}\chi) and c(μ2χ)c/2kc(\mu_{2}\chi)\leq c/2-k, or c(μ11χ)c(\mu_{1}^{-1}\chi) and c(μ21χ)c/2kc(\mu_{2}^{-1}\chi)\leq c/2-k. Thus

    Pc/2,k(W1,W2)=0.P_{c/2,k}(W_{1},W_{2})=0.

Adding up all pieces, we get that when l>1l>1,

Pc/2(W1,W2)=Pc/2,0(W1,W2)+2k0Pc/2,k(W1,W2)=0.P_{c/2}(W_{1},W_{2})=P_{c/2,0}(W_{1},W_{2})+2\sum\limits_{k\geq 0}P_{c/2,k}(W_{1},W_{2})=0.

Here the multiple 22 comes from the dichotomy that either c(μi1χ)=c/2kc(\mu_{i}^{-1}\chi)=c/2-k or c(μiχ)=c/2kc(\mu_{i}\chi)=c/2-k.

5.3.3. The case l=1l=1.

Here we can apply the method in Section 5.2.2 case (2).

  1. (1)

    If k<c/2lk<c/2-l, we keep working with the case c(μi1χ)=c/2kc(\mu_{i}^{-1}\chi)=c/2-k and can still start with the last integral in (5.11). Using Lemma 4.6, we have

    (5.12) yOE×θ11χE(1+ϖc/2l,c/2kly)θ1θ21(y)d×y\displaystyle\int\limits_{y\in O_{E}^{\times}}\theta_{1}^{-1}\chi_{E}(1+\varpi^{c/2-l,c/2-k-l}y)\theta_{1}\theta_{2}^{-1}(y)d^{\times}y
    =\displaystyle= y1OF×μ1χ(1+ϖc/21y1)μ11μ2(y1)d×y1y2OF×μ11χ(1+ϖc/2k1y2)μ1μ21(y2)d×y2\displaystyle\int\limits_{y_{1}\in O_{F}^{\times}}\mu_{1}\chi(1+\varpi^{c/2-1}y_{1})\mu_{1}^{-1}\mu_{2}(y_{1})d^{\times}y_{1}\int\limits_{y_{2}\in O_{F}^{\times}}\mu_{1}^{-1}\chi(1+\varpi^{c/2-k-1}y_{2})\mu_{1}\mu_{2}^{-1}(y_{2})d^{\times}y_{2}
    =\displaystyle= y1OF×ψ(bϖc/21y1)μ11μ2(y1)d×y1y2OF×ψ(aϖc/2k1y2)μ1μ21(y2)d×y2\displaystyle\int\limits_{y_{1}\in O_{F}^{\times}}\psi(b\varpi^{c/2-1}y_{1})\mu_{1}^{-1}\mu_{2}(y_{1})d^{\times}y_{1}\int\limits_{y_{2}\in O_{F}^{\times}}\psi(a\varpi^{c/2-k-1}y_{2})\mu_{1}\mu_{2}^{-1}(y_{2})d^{\times}y_{2}
    =\displaystyle= μ11μ2(abϖk)y1OF×ψ(ϖ1y1)μ11μ2(y1)d×y1y1OF×ψ(ϖ1y2)μ1μ21(y2)d×y2.\displaystyle\mu_{1}^{-1}\mu_{2}\left(\frac{a}{b}\varpi^{-k}\right)\int\limits_{y_{1}\in O_{F}^{\times}}\psi(\varpi^{-1}y_{1})\mu_{1}^{-1}\mu_{2}(y_{1})d^{\times}y_{1}\int\limits_{y_{1}\in O_{F}^{\times}}\psi(\varpi^{-1}y_{2})\mu_{1}\mu_{2}^{-1}(y_{2})d^{\times}y_{2}.

    Here aa is associated to μ11χ\mu_{1}^{-1}\chi with v(a)=c/2+kv(a)=-c/2+k, and bb is associated to μ1χ\mu_{1}\chi with v(b)=c/2v(b)=-c/2. The expression depends only on amodϖc/2+k+1a\mod\varpi^{-c/2+k+1} and bmodϖc/2+1b\mod\varpi^{-c/2+1}.

    If k>0k>0, when we take the sum over χ\chi with c(μ11χ)=c/2kc(\mu_{1}^{-1}\chi)=c/2-k, the congruence class for bb is not affected, while the congruence class for aa runs over ϖc/2+kOF×modϖc/2+k+1\varpi^{-c/2+k}O_{F}^{\times}\mod\varpi^{-c/2+k+1}. As c(μ11μ2)=l=1c(\mu_{1}^{-1}\mu_{2})=l=1, the sum in χ\chi with c(μ11χ)=c/2kc(\mu_{1}^{-1}\chi)=c/2-k for μ11μ2(abϖk)\mu_{1}^{-1}\mu_{2}\left(\frac{a}{b}\varpi^{-k}\right) is thus vanishing, and

    Pc/2,k(W1,W2)=0P_{c/2,k}(W_{1},W_{2})=0

    in this case.

    If k=0k=0, we take the sum in χ\chi with c(χ)=c(μiχ)=c(μi1χ)=c/2c(\chi)=c(\mu_{i}\chi)=c(\mu_{i}^{-1}\chi)=c/2. We break the sum in χ=χ0η\chi=\chi_{0}\eta into a sum of χ0c/21\chi_{0}\sim_{c/2-1} and a sum of η\eta with c(η)c/21c(\eta)\leq c/2-1. The sum in η\eta gives a counting constant as ηE(1+ϖc/21,c/21y)=1\eta_{E}(1+\varpi^{c/2-1,c/2-1}y)=1 in (5.12). Let α0\alpha_{0} be associated to χ0\chi_{0} while αi\alpha_{i} be associated to μi\mu_{i} for i=1,2i=1,2, then we have aα0α1a\equiv\alpha_{0}-\alpha_{1}, bα0+α1b\equiv\alpha_{0}+\alpha_{1}. The sum of χ0\chi_{0} is then equivalent to a sum of

    t=abα0α1α0+α1t=\frac{a}{b}\equiv\frac{\alpha_{0}-\alpha_{1}}{\alpha_{0}+\alpha_{1}}

    running through all OF×modϖOFO_{F}^{\times}\mod\varpi O_{F} except ±1\pm 1. As

    tOF×μ11μ2(t)=0,\sum\limits_{t\in O_{F}^{\times}}\mu_{1}^{-1}\mu_{2}(t)=0,

    we have

    (5.13) χyOE×θ11χE(1+ϖEc/2ly)θ1θ21(y)d×y=(q1)qc/22(1μ11μ2(1))q(q1)2.\sum\limits_{\chi}\int\limits_{y\in O_{E}^{\times}}\theta_{1}^{-1}\chi_{E}(1+\varpi_{E}^{c/2-l}y)\theta_{1}\theta_{2}^{-1}(y)d^{\times}y=(q-1)q^{c/2-2}(-1-\mu_{1}^{-1}\mu_{2}(-1))\frac{q}{(q-1)^{2}}.

    Recall that

    Pc/2,0(θ1,θ2)\displaystyle P_{c/2,0}(\theta_{1},\theta_{2})
    =\displaystyle= θ1θ21(ϖEc/2l)qc2lvE(w)=lθ2θ11(w)ψE(w)d×wχyOE×θ11χE(1+ϖEc/2ly)θ1θ21(y)d×y.\displaystyle\frac{\theta_{1}\theta_{2}^{-1}(\varpi_{E}^{c/2-l})}{q^{c-2l}}\int\limits_{v_{E}(w)=-l}\theta_{2}\theta_{1}^{-1}(w)\psi_{E}(w)d^{\times}w\sum\limits_{\chi}\int\limits_{y\in O_{E}^{\times}}\theta_{1}^{-1}\chi_{E}(1+\varpi_{E}^{c/2-l}y)\theta_{1}\theta_{2}^{-1}(y)d^{\times}y.

    Thus by (5.13) and Lemma 4.4,

    |Pc/2,0(θ1,θ2)|2(q1)3qc/22,\left|P_{c/2,0}(\theta_{1},\theta_{2})\right|\leq\frac{2}{(q-1)^{3}q^{c/2-2}},
    |Pc/2,0(W1,W2)|=||ψχ|2C0C0¯Pc/2,0(θ1,θ2)|2q1.\left|P_{c/2,0}(W_{1},W_{2})\right|=\left|\frac{\left|\int\psi\chi\right|^{2}}{C_{0}\overline{C_{0}^{\prime}}}P_{c/2,0}(\theta_{1},\theta_{2})\right|\leq\frac{2}{q-1}.
  2. (2)

    If k=c/21k=c/2-1, we again focus on the sum in χ\chi with c(μi1χ)1c(\mu_{i}^{-1}\chi)\leq 1. Recall that θi=(μi1,μi)\theta_{i}=(\mu_{i}^{-1},\mu_{i}). Then the integral

    c(μi1χ)1y=(y1,y2)OE×,y2ϖc/2klθ11χE(1+ϖc/2l,c/2kly)θ1θ21(y)d×y\sum\limits_{c(\mu_{i}^{-1}\chi)\leq 1}\int\limits_{y=(y_{1},y_{2})\in O_{E}^{\times},y_{2}\not\equiv-\varpi^{c/2-k-l}}\theta_{1}^{-1}\chi_{E}(1+\varpi^{c/2-l,c/2-k-l}y)\theta_{1}\theta_{2}^{-1}(y)d^{\times}y

    as in (5.11) can be explicitly written as

    c(μi1χ)1y1OF×μ1χ(1+ϖc/21y1)μ11μ2(y1)d×y1y2OF×,y21μ11χ(1+y2)μ1μ21(y2)d×y2\displaystyle\sum\limits_{c(\mu_{i}^{-1}\chi)\leq 1}\int\limits_{y_{1}\in O_{F}^{\times}}\mu_{1}\chi(1+\varpi^{c/2-1}y_{1})\mu_{1}^{-1}\mu_{2}(y_{1})d^{\times}y_{1}\int\limits_{y_{2}\in O_{F}^{\times},y_{2}\not\equiv-1}\mu_{1}^{-1}\chi(1+y_{2})\mu_{1}\mu_{2}^{-1}(y_{2})d^{\times}y_{2}
    =\displaystyle= μ1μ21(bϖc/2)y1OF×ψ(ϖ1y1)μ11μ2(y1)d×y1y2OF×,y21c(μi1χ)1μ11χ(1+y2)μ1μ21(y2)d×y2\displaystyle\mu_{1}\mu_{2}^{-1}(b\varpi^{c/2})\int\limits_{y_{1}\in O_{F}^{\times}}\psi(\varpi^{-1}y_{1})\mu_{1}^{-1}\mu_{2}(y_{1})d^{\times}y_{1}\int\limits_{y_{2}\in O_{F}^{\times},y_{2}\not\equiv-1}\sum\limits_{c(\mu_{i}^{-1}\chi)\leq 1}\mu_{1}^{-1}\chi(1+y_{2})\mu_{1}\mu_{2}^{-1}(y_{2})d^{\times}y_{2}
    =\displaystyle= 0.\displaystyle 0.

    Here bϖc/2OF×b\in\varpi^{-c/2}O_{F}^{\times} is the constant associated to μ1χ\mu_{1}\chi and is independent from χ\chi with c(μi1χ)1c(\mu_{i}^{-1}\chi)\leq 1. Thus again

    Pc/2,k(W1,W2)=0.P_{c/2,k}(W_{1},W_{2})=0.
  3. (3)

    If k>c/2lk>c/2-l, similar to l>1l>1 case we also have

    Pc/2,k(W1,W2)=0.P_{c/2,k}(W_{1},W_{2})=0.

    Adding up all pieces, we get that when l=1l=1,

    |Pc/2(W1,W2)|2q1.\left|P_{c/2}(W_{1},W_{2})\right|\leq\frac{2}{q-1}.

5.4. c(πi)=2c(\pi_{i})=2 case

Note that in this case πi\pi_{i} is either a supercuspidal representation constructed from an inert field extension, or a principal series representation from a split extension. In this case the associated characters satisfy c(θi)=1c(\theta_{i})=1, so we can no longer apply Lemma 4.6 and p-adic stationary phase analysis for the local integrals.

Consider first the computation of P2(W1,W2)P_{2}(W_{1},W_{2}), or that of P1(W1,W2)P_{1}(W_{1},W_{2}) when E=EE=E^{\prime} and c(θ1θ21)=0c(\theta_{1}\theta_{2}^{-1})=0. The discussions in Section 5.1.1, 5.2.1 and 5.3.1 still hold, as we didn’t use Lemma 4.6 there. Thus we always have

P2(W1,W2)=1.P_{2}(W_{1},W_{2})=1.

When E=EE=E^{\prime} are fields, we have

P1(W1,W2)=1.P_{1}(W_{1},W_{2})=1.

When E=EE=E^{\prime} split, we have

|P1(W1,W2)q3q1|2q1/2+Re(γ)1.\left|P_{1}(W_{1},W_{2})-\frac{q-3}{q-1}\right|\leq\frac{2}{q^{1/2+\operatorname{Re}(\gamma)}-1}.

For the remaining computations, we first prove the following lemma:

Lemma 5.7.

Let θ\theta be a nontrivial charcter on kE×k_{E}^{\times} with θ|kF×=1\theta|_{k_{F}^{\times}}=1, where kEk_{E} is a quadratic field extension over kFk_{F}. Without loss of generality we assume that kE=kF[D]k_{E}=k_{F}[\sqrt{D}]. Then

|akFθ(1+aD)|=1.\left|\sum\limits_{a\in k_{F}}\theta(1+a\sqrt{D})\right|=1.
Proof.

Indeed as θ\theta is nontrivial, we have

(a,b)kF2,(a,b)0θ(a+bD)=0.\sum\limits_{(a,b)\in k_{F}^{2},(a,b)\neq 0}\theta(a+b\sqrt{D})=0.

Using that θ|kF×=1\theta|_{k_{F}^{\times}}=1, we get that

(a,b)kE2,(a,b)0θ(a+bD)=(q1)akFθ(1+aD)+(q1)θ(D).\sum\limits_{(a,b)\in k_{E}^{2},(a,b)\neq 0}\theta(a+b\sqrt{D})=(q-1)\sum\limits_{a\in k_{F}}\theta(1+a\sqrt{D})+(q-1)\theta(\sqrt{D}).

The lemma follows. ∎

We summarize the remaining computations in the following.

Lemma 5.8.

Suppose that pp is large enough. Suppose that θi\theta_{i} are defined over étale quadratic algebras EE and EE^{\prime} with c(θi)=1c(\theta_{i})=1, such that either EEE\neq E^{\prime}, or E=EE=E^{\prime} and c(θ1θ2)=c(θ1θ21)=1c(\theta_{1}\theta_{2})=c(\theta_{1}\theta_{2}^{-1})=1. Then we have

(5.14) P1(θ1,θ2)\displaystyle P_{1}(\theta_{1},\theta_{2})
=\displaystyle= c(χ)1vE(u)=1θ11(u)χE(u)ψE(u)d×uvE(w)=1θ21(w)χE(w)ψE(w)d×w¯1q2,\displaystyle\sum\limits_{c(\chi)\leq 1}\int\limits_{v_{E}(u)=-1}\theta_{1}^{-1}(u)\chi_{E}(u)\psi_{E}(u)d^{\times}u\overline{\int\limits_{v_{E^{\prime}}(w)=-1}\theta_{2}^{-1}(w)\chi_{E^{\prime}}(w)\psi_{E^{\prime}}(w)d^{\times}w}\ll\frac{1}{q^{2}},
P1(W1,W2)1q.\displaystyle P_{1}(W_{1},W_{2})\ll\frac{1}{q}.

Note that if we directly apply the bound for each Gauss integrals, we will get

P1(θ1,θ2)1q,P_{1}(\theta_{1},\theta_{2})\ll\frac{1}{q},

which is also sharp if E=EE=E^{\prime} and θ1=θ2±1\theta_{1}=\theta_{2}^{\pm 1}. The lemma claims additional power saving when θi\theta_{i} are not related.

Note that as pp is large enough, we can freely add or subtract the term χ=1\chi=1 in the definition of P1(θ1,θ2)P_{1}(\theta_{1},\theta_{2}) without affecting the final asymptotic estimate, as

|vE(u)=1θ11(u)ψE(u)d×u|1q.\left|\int\limits_{v_{E}(u)=-1}\theta_{1}^{-1}(u)\psi_{E}(u)d^{\times}u\right|\ll\frac{1}{q}.
Proof.

Here we take a case by case approach. There might be more uniform way to obtain the power saving.

Suppose first that E=EE=E^{\prime}. By a change of variable u=wxu=wx, we have

P1(θ1,θ2)\displaystyle P_{1}(\theta_{1},\theta_{2}) =c(χ)1vE(w)=1vE(x)=0θ11θ2(w)θ11(x)χE(x)ψE((x1)w)d×xd×w\displaystyle=\sum\limits_{c(\chi)\leq 1}\int\limits_{v_{E}(w)=-1}\int\limits_{v_{E}(x)=0}\theta_{1}^{-1}\theta_{2}(w)\theta_{1}^{-1}(x)\chi_{E}(x)\psi_{E}((x-1)w)d^{\times}xd^{\times}w
=vE(w)=1vE(x)=0θ11θ2(w)θ11(x)c(χ)1χE(x)ψE((x1)w)d×xd×w\displaystyle=\int\limits_{v_{E}(w)=-1}\int\limits_{v_{E}(x)=0}\theta_{1}^{-1}\theta_{2}(w)\theta_{1}^{-1}(x)\sum\limits_{c(\chi)\leq 1}\chi_{E}(x)\psi_{E}((x-1)w)d^{\times}xd^{\times}w
=(q1)vE(w)=1Nm(x)1modpθ11θ2(w)θ11(x)ψE((x1)w)d×xd×w.\displaystyle=(q-1)\int\limits_{v_{E}(w)=-1}\int\limits_{Nm(x)\equiv 1\mod p}\theta_{1}^{-1}\theta_{2}(w)\theta_{1}^{-1}(x)\psi_{E}((x-1)w)d^{\times}xd^{\times}w.

Here q1q-1 comes from the number of χ\chi. The integral in ww is nonvanishing if and only if x1OE×x-1\in O_{E}^{\times}, in which case we can apply a change of variable v=(x1)wv=(x-1)w and get

P1(θ1,θ2)\displaystyle P_{1}(\theta_{1},\theta_{2})
=\displaystyle= (q1)vE(v)=1θ11θ2(v)ψE(v)d×vNm(x)1modp,x1OE×θ1θ21(x1)θ11(x)d×x.\displaystyle(q-1)\int\limits_{v_{E}(v)=-1}\theta_{1}^{-1}\theta_{2}(v)\psi_{E}(v)d^{\times}v\int\limits_{Nm(x)\equiv 1\mod p,x-1\in O_{E}^{\times}}\theta_{1}\theta_{2}^{-1}(x-1)\theta_{1}^{-1}(x)d^{\times}x.

Note that

vE(v)=1θ11θ2(v)ψE(v)d×v1q.\int\limits_{v_{E}(v)=-1}\theta_{1}^{-1}\theta_{2}(v)\psi_{E}(v)d^{\times}v\ll\frac{1}{q}.

Thus to verify the lemma in this case, it suffices to prove that

Nm(x)1modp,x1OE×θ1θ21(x1)θ11(x)d×x1q2.\int\limits_{Nm(x)\equiv 1\mod p,x-1\in O_{E}^{\times}}\theta_{1}\theta_{2}^{-1}(x-1)\theta_{1}^{-1}(x)d^{\times}x\ll\frac{1}{q^{2}}.

Using Hilbert 90, we write x=y/y¯x=y/\overline{y} for ykF×\kE×y\in k_{F}^{\times}\backslash k_{E}^{\times}. Using that c(θi)=1c(\theta_{i})=1, we have

Nm(x)1modp,x1OE×θ1θ21(x1)θ11(x)d×x\displaystyle\int\limits_{Nm(x)\equiv 1\mod p,x-1\in O_{E}^{\times}}\theta_{1}\theta_{2}^{-1}(x-1)\theta_{1}^{-1}(x)d^{\times}x
=\displaystyle= Vol(1+ϖEOE,d×x)ykF×\kE×,yy¯kE×θ1θ21(yy¯)θ11(y)θ2(y¯),\displaystyle\text{Vol}(1+\varpi_{E}O_{E},d^{\times}x)\sum\limits_{y\in k_{F}^{\times}\backslash k_{E}^{\times},y-\overline{y}\in k_{E}^{\times}}\theta_{1}\theta_{2}^{-1}(y-\overline{y})\theta_{1}^{-1}(y)\theta_{2}(\overline{y}),

where

Vol(1+ϖEOE,d×x)1q2.\text{Vol}(1+\varpi_{E}O_{E},d^{\times}x)\asymp\frac{1}{q^{2}}.

Now if kE=kF[D]k_{E}=k_{F}[\sqrt{D}] is a field, for the sum in yy with yy¯y\not\equiv\overline{y}, we can choose a special set of representatives y=1+aDy=1+a\sqrt{D} or D\sqrt{D}. y=Dy=\sqrt{D} contributes a term of absolute value 11, while the sum

akFθ1θ21(yy¯)θ11(y)θ2(y¯)\displaystyle\sum\limits_{a\in k_{F}}\theta_{1}\theta_{2}^{-1}(y-\overline{y})\theta_{1}^{-1}(y)\theta_{2}(\overline{y}) =akFθ1θ21(2aD)θ11θ21(1+aD)\displaystyle=\sum\limits_{a\in k_{F}}\theta_{1}\theta_{2}^{-1}(2a\sqrt{D})\theta_{1}^{-1}\theta_{2}^{-1}(1+a\sqrt{D})
=θ1θ21(D)akFθ11θ21(1+aD)\displaystyle=\theta_{1}\theta_{2}^{-1}(\sqrt{D})\sum\limits_{a\in k_{F}}\theta_{1}^{-1}\theta_{2}^{-1}(1+a\sqrt{D})
1.\displaystyle\ll 1.

Here in the first two equalities we have used that θi|F×=1\theta_{i}|_{F^{\times}}=1. In the last line we applied Lemma 5.7. The result follows in this case.

On the other hand if kE=kF×kFk_{E}=k_{F}\times k_{F}, for the sum in yy with yy¯kE×y-\overline{y}\in k_{E}^{\times}, we can choose a set of representatives y=(a,1)y=(a,1) with akF×a\in k_{F}^{\times}, a1a\neq 1. Then

ykF×\kE×,yy¯kE×θ1θ21(yy¯)θ11(y)θ2(y¯)\displaystyle\sum\limits_{y\in k_{F}^{\times}\backslash k_{E}^{\times},y-\overline{y}\in k_{E}^{\times}}\theta_{1}\theta_{2}^{-1}(y-\overline{y})\theta_{1}^{-1}(y)\theta_{2}(\overline{y}) =θ1θ21(1,1)akF×,a1μ1μ2(a)\displaystyle=\theta_{1}\theta_{2}^{-1}(1,-1)\sum\limits_{a\in k_{F}^{\times},a\neq 1}\mu_{1}\mu_{2}(a)
=θ1θ21(1,1).\displaystyle=-\theta_{1}\theta_{2}^{-1}(1,-1).

Here in the second equality we have used that c(μ1μ2)=1c(\mu_{1}\mu_{2})=1 and thus

akF×μ1μ2(a)=0.\sum\limits_{a\in k_{F}^{\times}}\mu_{1}\mu_{2}(a)=0.

Again the result follows.

Suppose without loss of generality now that EE is a field extension while E=F×FE^{\prime}=F\times F. By writing w=(y1,y2)ϖ1OE×w=(y_{1},y_{2})\in\varpi^{-1}O_{E}^{\times} in (5.14) and sum in χ\chi first, we get that the sum in χ\chi is vanishing unless y2NE/F(u)y1modϖOFy_{2}\equiv\frac{N_{E/F}(u)}{y_{1}}\mod\varpi O_{F}. Thus

P1(θ1,θ2)\displaystyle P_{1}(\theta_{1},\theta_{2})
=\displaystyle= vE(u)=1θ11(u)ψE(u)vF(y1)=1μ2(y1)μ21(NE/F(u)y1)ψ(y1+NE/F(u)y1)d×y1¯d×u.\displaystyle\int\limits_{v_{E}(u)=-1}\theta_{1}^{-1}(u)\psi_{E}(u)\overline{\int\limits_{v_{F}(y_{1})=-1}\mu_{2}(y_{1})\mu_{2}^{-1}\left(\frac{N_{E/F}(u)}{y_{1}}\right)\psi\left(y_{1}+\frac{N_{E/F}(u)}{y_{1}}\right)d^{\times}y_{1}}d^{\times}u.

Here the number of χ\chi cancels with the volume of y2y_{2} with the above mentioned congruence requirement. Make a change of variable u=y1vu=y_{1}v, we get

P1(θ1,θ2)\displaystyle P_{1}(\theta_{1},\theta_{2}) =vE(v)=0vF(y1)=1θ11(v)μ21(NE/F(v))¯ψ((TrE/F(v)1NE/F(v))y1)d×y1d×v.\displaystyle=\int\limits_{v_{E}(v)=0}\int\limits_{v_{F}(y_{1})=-1}\theta_{1}^{-1}(v)\overline{\mu_{2}^{-1}\left(N_{E/F}(v)\right)}\psi\left((Tr_{E/F}(v)-1-N_{E/F}(v))y_{1}\right)d^{\times}y_{1}d^{\times}v.
=vE(v)=0vF(y1)=1θ11(v)μ21(NE/F(v))¯ψ(NE/F(v1)y1)d×y1d×v.\displaystyle=\int\limits_{v_{E}(v)=0}\int\limits_{v_{F}(y_{1})=-1}\theta_{1}^{-1}(v)\overline{\mu_{2}^{-1}\left(N_{E/F}(v)\right)}\psi\left(-N_{E/F}(v-1)y_{1}\right)d^{\times}y_{1}d^{\times}v.

Here in the first equality we have used that θ1|F×=1\theta_{1}|_{{F}^{\times}}=1. Note that as vOE×v\in O_{E}^{\times}, v(NE/F(v1))=2vE(v1)v(N_{E/F}(v-1))=2v_{E}(v-1). We can then integrate in y1y_{1} first and break the integral in vv according to whether v1modϖOEv\equiv 1\mod\varpi O_{E},

P1(θ1,θ2)\displaystyle P_{1}(\theta_{1},\theta_{2}) =vE(v)=0,v1θ11(v)μ2,E1(v)¯d×v1q1vE(v)=0,v1θ11(v)μ2,E1(v)¯d×v\displaystyle=\int\limits_{v_{E}(v)=0,v\equiv 1}\theta_{1}^{-1}(v)\overline{\mu_{2,E}^{-1}(v)}d^{\times}v-\frac{1}{q-1}\int\limits_{v_{E}(v)=0,v\not\equiv 1}\theta_{1}^{-1}(v)\overline{\mu_{2,E}^{-1}(v)}d^{\times}v
=qq1vE(v)=0,v1θ11(v)μ2,E1(v)¯d×v1q1vE(v)=0θ11(v)μ2,E1(v)¯d×v.\displaystyle=\frac{q}{q-1}\int\limits_{v_{E}(v)=0,v\equiv 1}\theta_{1}^{-1}(v)\overline{\mu_{2,E}^{-1}(v)}d^{\times}v-\frac{1}{q-1}\int\limits_{v_{E}(v)=0}\theta_{1}^{-1}(v)\overline{\mu_{2,E}^{-1}(v)}d^{\times}v.

Note that Vol(1+ϖEOE,d×v)=1q21\text{Vol}(1+\varpi_{E}O_{E},d^{\times}v)=\frac{1}{q^{2}-1}, so the first term can be directly controlled by 1q2\frac{1}{q^{2}}. On the other hand θ11μ2,E1¯\theta_{1}^{-1}\overline{\mu_{2,E}^{-1}} is a nontrivial character on OE×O_{E}^{\times}, so the second term is vanishing. Thus we get in the last case

P1(θ1,θ2)1q2.P_{1}(\theta_{1},\theta_{2})\ll\frac{1}{q^{2}}.

5.5. p=2p=2 case

In addition to possible non-dihedral supercuspidal representations, there are several other potential issues in this case. For example, (5.9) may not be true. c(θ11θ2)c(\theta_{1}^{-1}\theta_{2}) and c(θ1θ2)c(\theta_{1}\theta_{2}) may both be smaller than c(θi)c(\theta_{i}) when E=EE=E^{\prime}, though we have following slightly weaker result.

Lemma 5.9.

Suppose that c(θi)c(\theta_{i}) are large enough, and l=c(θ11θ2)c(θ1θ2)l=c(\theta_{1}^{-1}\theta_{2})\leq c(\theta_{1}\theta_{2}). Then

c(θ1θ2)=c(θi)+O(1).c(\theta_{1}\theta_{2})=c(\theta_{i})+O(1).
Proof.

By the assumption lc(θ1θ2)l\leq c(\theta_{1}\theta_{2}), we only have to consider the case where ll is sufficiently smaller than c(θi)c(\theta_{i}). Let αi\alpha_{i} be associated to θi\theta_{i} by Corollary 4.7. Then by c(θ11θ2)=lc(\theta_{1}^{-1}\theta_{2})=l,

vE(α2α1)l+c(ψE).v_{E}(\alpha_{2}-\alpha_{1})\geq-l+c(\psi_{E}).

θ1θ2\theta_{1}\theta_{2} is then associated to α1+α2\alpha_{1}+\alpha_{2} with

vE(α1+α2)=vE(α1α2+2α2)=vE(α2)+vE(2),v_{E}(\alpha_{1}+\alpha_{2})=v_{E}(\alpha_{1}-\alpha_{2}+2\alpha_{2})=v_{E}(\alpha_{2})+v_{E}(2),

from which the lemma follows. ∎

For the partial pairing Pi(W1,W2)P_{i}(W_{1},W_{2}), the following weaker result suffices for our purposes while avoiding many technique issues.

Lemma 5.10.

Suppose that p=2p=2 and c(π1)=c(π2)c(\pi_{1})=c(\pi_{2}) is large enough. Let WiW_{i} be the Whittaker functions associated to newforms in πi\pi_{i}. Then there exists an absolutely bounded positive integer aa such that

Pi(W1,W2)=1P_{i}(W_{1},W_{2})=1

for any ic(π1×π2)/2+ai\geq c(\pi_{1}\times\pi_{2})/2+a.

Here we allow ic(π1)i\geq c(\pi_{1}), in which case one directly have Pi(W1,W2)=1P_{i}(W_{1},W_{2})=1 from the newform theory.

The assumption that c(π1)=c(π2)c(\pi_{1})=c(\pi_{2}) large enough excludes the case where πi\pi_{i} could be non-dihedral supercuspidal representations. This result can be then proven similarly as in Section 5.1.1, 5.2.1. Picking sufficiently large but bounded aa also avoids the case i=c/2i=c/2 for principal series representations according to Lemma 5.9.

6. Bounds for local period integrals

For η\eta a character of F×{F}^{\times}, Recall that an element φ3π3=π(η,η1)\varphi_{3}\in\pi_{3}=\pi(\eta,\eta^{-1}) satisfies

(6.1) φ3((a1n0a2)g)=η(a1)η1(a2)|a1a2|1/2φ(g).\varphi_{3}\left(\begin{pmatrix}a_{1}&n\\ 0&a_{2}\end{pmatrix}g\right)=\eta(a_{1})\eta^{-1}(a_{2})\left|\frac{a_{1}}{a_{2}}\right|^{1/2}\varphi(g).

Denote by

(6.2) IRS(φ1,φ2,φ3)=Z(F)N\GL2(F)Wφ1(g)Wφ2(g)¯φ3(g)𝑑gI^{\text{RS}}\left(\varphi_{1},\varphi_{2},\varphi_{3}\right)=\int\limits_{{Z({F})N}\backslash{\text{GL}}_{2}{({F})}}W_{\varphi_{1}}\left(g\right)\overline{W_{\varphi_{2}}\left(g\right)}\varphi_{3}\left(g\right)dg

the local integral for the Rankin–Selberg integral. Here WφW_{\varphi} is the Whittaker function associated to φ\varphi with respect to the fixed additive character ψ\psi. We also note that Wφ¯=Wφ\overline{W_{\varphi}}=W^{-}_{\varphi} where WφW^{-}_{\varphi} is for ψ(x)=ψ(x)\psi^{-}(x)=\psi(-x).

In general for φiπi\varphi_{i}\in\pi_{i}, denote by

(6.3) IT(φ1,φ2,φ3)=F×\GL2(F)i=13Φφi(g)dgI^{\text{T}}\left(\varphi_{1},\varphi_{2},\varphi_{3}\right)=\int\limits_{{F}^{\times}\backslash{\text{GL}}_{2}{({F})}}\prod\limits_{i=1}^{3}\Phi_{\varphi_{i}}\left(g\right)dg

the local integral for the triple product formula, where Φφi\Phi_{\varphi_{i}} is the matrix coefficient associated to φi\varphi_{i}.

This section focuses on giving lower and upper bound for IRSI^{\text{RS}} and ITI^{T} with proper choice of L2L^{2}-normalized test vectors. For conciseness we omit the terms QϵQ^{\epsilon} from the computations.

6.1. Set up and preparations

By the setting in Theorem 3.22, we assume without loss of generality that η\eta and π3\pi_{3} are unramified. We shall search for test vectors of the following shape: φi=φi\varphi_{i}=\varphi_{i}^{\circ} are normalized newforms for i=1,2i=1,2, and φ3=π3((ϖn1))φ3\varphi_{3}=\pi_{3}\left(\begin{pmatrix}\varpi^{-n}&\\ &1\end{pmatrix}\right)\varphi_{3}^{\circ} for c/2ncc/2\leq n\leq c where c=max{c(π1),c(π2)}c=\max\{c(\pi_{1}),c(\pi_{2})\}.

Suppose that η=||γ\eta=|\cdot|^{\gamma} with |Re(γ)|<7/64|\operatorname{Re}(\gamma)|<7/64.

The main strategy for the computations follows that of [YH20], which we briefly recall here. As all integrand are invariant by K0(ϖc)K_{0}(\varpi^{c}) by our setting, we have

IRS(φ1,φ2,φ3)=i=0cAiaF×Wφ1(i)Wφ2(i)¯(a)φ3((aϖn001)(10ϖin1))|a|1d×aI^{\text{RS}}\left(\varphi_{1},\varphi_{2},\varphi_{3}\right)=\sum\limits_{i=0}^{c}A_{i}\int\limits_{a\in{F}^{\times}}W_{\varphi_{1}}^{(i)}\overline{W_{\varphi_{2}}^{(i)}}(a)\varphi_{3}^{\circ}\left(\begin{pmatrix}a\varpi^{-n}&0\\ 0&1\end{pmatrix}\begin{pmatrix}1&0\\ \varpi^{i-n}&1\end{pmatrix}\right)|a|^{-1}d^{\times}a

for the constants AiA_{i} as in Lemma 4.10.

The Iwasawa decomposition (in the sense of Lemma 4.9) for (aϖn001)(10ϖin1)\begin{pmatrix}a\varpi^{-n}&0\\ 0&1\end{pmatrix}\begin{pmatrix}1&0\\ \varpi^{i-n}&1\end{pmatrix} can be done as follows: When ini\geq n, it is already in the standard form. When i<ni<n, we have

(6.4) (aϖn001)(10ϖin1)=(aϖiaϖn0ϖin)(011ϖni).\begin{pmatrix}a\varpi^{-n}&0\\ 0&1\end{pmatrix}\begin{pmatrix}1&0\\ \varpi^{i-n}&1\end{pmatrix}=\begin{pmatrix}a\varpi^{-i}&a\varpi^{-n}\\ 0&\varpi^{i-n}\end{pmatrix}\begin{pmatrix}0&-1\\ 1&\varpi^{n-i}\end{pmatrix}.

Thus by the property of φ3\varphi_{3}^{\circ} with normalization φ3(1)=1\varphi_{3}^{\circ}(1)=1, we get

(6.5) IRS(φ1,φ2,φ3)\displaystyle I^{\text{RS}}\left(\varphi_{1},\varphi_{2},\varphi_{3}\right)
=\displaystyle= i=ncAiqn(1/2+γ)aF×Wφ1(i)Wφ2(i)¯(a)|a|γ1/2d×a+i=0n1Aiq(2in)(1/2+γ)aF×Wφ1(i)Wφ2(i)¯(a)|a|γ1/2d×a\displaystyle\sum\limits_{i=n}^{c}A_{i}q^{n(1/2+\gamma)}\int\limits_{a\in{F}^{\times}}W_{\varphi_{1}}^{(i)}\overline{W_{\varphi_{2}}^{(i)}}(a)|a|^{\gamma-1/2}d^{\times}a+\sum\limits_{i=0}^{n-1}A_{i}q^{(2i-n)(1/2+\gamma)}\int\limits_{a\in{F}^{\times}}W_{\varphi_{1}}^{(i)}\overline{W_{\varphi_{2}}^{(i)}}(a)|a|^{\gamma-1/2}d^{\times}a
=\displaystyle= i=ncAiqn(1/2+γ)Pi(Wφ1,Wφ2)+i=0n1Aiq(2in)(1/2+γ)Pi(Wφ1,Wφ2).\displaystyle\sum\limits_{i=n}^{c}A_{i}q^{n(1/2+\gamma)}P_{i}(W_{\varphi_{1}},W_{\varphi_{2}})+\sum\limits_{i=0}^{n-1}A_{i}q^{(2i-n)(1/2+\gamma)}P_{i}(W_{\varphi_{1}},W_{\varphi_{2}}).

The following lemma relates Pi(Wφ1,Wφ2)P_{i}(W_{\varphi_{1}},W_{\varphi_{2}}) with i<c/2i<c/2 to those with i>c/2i>c/2 using the Atkin-Lehner operator when c(π1)=c(π2)c(\pi_{1})=c(\pi_{2}).

Lemma 6.1.

Suppose that c=c(π1)=c(π2)c=c(\pi_{1})=c(\pi_{2}), and i<c/2i<c/2. Then

|Pi(Wφ1,Wφ2)|=q(c2i)(Re(γ)1/2)|Pci(Wφ1,Wφ2)|.\left|P_{i}(W_{\varphi_{1}},W_{\varphi_{2}})\right|=q^{(c-2i)(\operatorname{Re}(\gamma)-1/2)}\left|P_{c-i}(W_{\varphi_{1}},W_{\varphi_{2}})\right|.
Proof.

Recall that for ωc=(01ϖc0)\omega_{c}=\begin{pmatrix}0&1\\ -\varpi^{c}&0\end{pmatrix} which stabilises the congruence subgroup K0(ϖc)K_{0}\left(\varpi^{c}\right), we have by the uniqueness of the newform,

πi(ωc)φi=aiφi\pi_{i}\left(\omega_{c}\right)\varphi_{i}^{\circ}=a_{i}\varphi_{i}^{\circ}

for i=1,2i=1,2 and some constants ai=±1a_{i}=\pm 1. It is also straightforward to verify that

(1ϖi1)(01ϖc0)=(ϖci1ϖi)(1ϖci1)(11)\begin{pmatrix}1&\\ \varpi^{i}&1\end{pmatrix}\begin{pmatrix}0&1\\ -\varpi^{c}&0\end{pmatrix}=\begin{pmatrix}-\varpi^{c-i}&1\\ &\varpi^{i}\end{pmatrix}\begin{pmatrix}1&\\ \varpi^{c-i}&1\end{pmatrix}\begin{pmatrix}-1&\\ &1\end{pmatrix}

As a result for i<c/2i<c/2 and newforms φπ1\varphi\in\pi_{1} or π2\pi_{2},

Wφ(i)(a)=aiWφ(ci)(aϖc2i)ψ(ϖia).W_{\varphi}^{(i)}(a)=a_{i}W_{\varphi}^{(c-i)}(-a\varpi^{c-2i})\psi(\varpi^{-i}a).

By Lemma 4.17, 4.20, Wφ(ci)W_{\varphi}^{(c-i)} is supported only at v(a)=0v(a)=0, thus Wφ(i)(a)W_{\varphi}^{(i)}(a) is supported only at v(a)=2icv(a)=2i-c. Thus

Pi(Wφ1,Wφ2)\displaystyle P_{i}(W_{\varphi_{1}},W_{\varphi_{2}}) =aF×Wφ1(i)Wφ2(i)¯(a)|a|γ1/2d×a\displaystyle=\int\limits_{a\in{F}^{\times}}W_{\varphi_{1}}^{(i)}\overline{W_{\varphi_{2}}^{(i)}}(a)|a|^{\gamma-1/2}d^{\times}a
=a1a2¯q(c2i)(γ1/2)aF×Wφ1(ci)Wφ2(ci)¯(a)|a|γ1/2d×a\displaystyle=a_{1}\overline{a_{2}}q^{(c-2i)(\gamma-1/2)}\int\limits_{a\in{F}^{\times}}W_{\varphi_{1}}^{(c-i)}\overline{W_{\varphi_{2}}^{(c-i)}}(a)|a|^{\gamma-1/2}d^{\times}a

The lemma now follows by taking absolute values on both sides. ∎

The following result relates the local triple product integral with the local Rankin–Selberg integral, which is originally due to [MVIHES] and is later extended in other works in for example [HS19, Proposition 5.1].

Lemma 6.2.

Suppose that π3\pi_{3} is a parabolically-induced representation, and πi\pi_{i} satisfies the bound θ<1/6\theta{<1/6} towards the Ramanujan conjecture. Let πi~\tilde{\pi_{i}} be the contragredient representation of πi\pi_{i}, φiπi\varphi_{i}\in\pi_{i} and φi~πi~\tilde{\varphi_{i}}\in\tilde{\pi_{i}}. Let (,)(\cdot,\cdot) be the natural GL2{\text{GL}}_{2}-invariant pairing between πi\pi_{i} and πi~\tilde{\pi_{i}}. Suppose furthermore that φ3\varphi_{3} belongs to the model π(η,η1)\pi(\eta,\eta^{-1}) and φ3~\tilde{\varphi_{3}} belongs to the model π(η1,η)\pi(\eta^{-1},\eta). Then

(6.6) Z\GL2(F)i(πi(g)φi,φi~)dg\displaystyle\int\limits_{Z\backslash{\text{GL}}_{2}({F})}\prod_{i}(\pi_{i}(g)\varphi_{i},\tilde{\varphi_{i}})dg
=\displaystyle= ζF(1)Z(F)N\GL2(F)Wφ1(g)Wφ2(Jg)φ3(g)𝑑gZ(F)N\GL2(F)Wφ1~(Jg)Wφ2~(g)φ3~(g)𝑑g.\displaystyle\zeta_{F}(1)\int\limits_{{Z({F})N}\backslash{\text{GL}}_{2}{({F})}}W_{\varphi_{1}}\left(g\right)W_{\varphi_{2}}\left(Jg\right)\varphi_{3}\left(g\right)dg\int\limits_{{Z({F})N}\backslash{\text{GL}}_{2}{({F})}}W_{\tilde{\varphi_{1}}}\left(Jg\right)W_{\tilde{\varphi_{2}}}\left(g\right)\tilde{\varphi_{3}}\left(g\right)dg.

Here J=(1001)J=\begin{pmatrix}-1&0\\ 0&1\end{pmatrix} and Wφ(Jg)W_{\varphi}(Jg) is the Whittaker function associated to φ\varphi with respect to ψ\psi^{-}.

In our case, we have πi~πi\tilde{\pi_{i}}\simeq\pi_{i} as the central characters are trivial, and (,)(\cdot,\cdot) is the usual unitary pairing. We can take φi~=φi\tilde{\varphi_{i}}=\varphi_{i}, but note that φ3\varphi_{3} and φ3~\tilde{\varphi_{3}} are in different models.

6.2. Upper bounds

Recall the notation v,ϵ\leq_{v,\epsilon} from Section 2.3. We need a special case of the upper bounds obtained in [YH20] here. Note however that [YH20, Corollary 3.17] contains an error in the non-tempered case, leading to a weaker upper bound (while the main conclusions there should still hold). This is corrected in the following result:

Proposition 6.3.

Suppose that π3\pi_{3} is unramified, and c=c(π1)=c(π2)c=c(\pi_{1})=c(\pi_{2}). Denote c=c(π2×π2)c^{\prime}=c(\pi_{2}\times\pi_{2}). Suppose that πi\pi_{i} satisfies the bound θ<7/64\theta<7/64 towards the Ramanujan conjecture, for i=1,2,3i=1,2,3. Let φiπi\varphi_{i}^{\circ}\in\pi_{i} be normalized newforms. Suppose that nc/2n\geq c/2, and π3=π(η,η1)\pi_{3}=\pi(\eta,\eta^{-1}) with η=||γ\eta=|\cdot|^{\gamma} in the case of local Rankin–Selberg integral.

When Re(γ)0\operatorname{Re}(\gamma)\geq 0,

(6.7) |IRS(φ1,φ2,π3((ϖn001))φ3)|v,ϵ1q(1/2|Re(γ)|ϵ)n.\left|I^{\text{RS}}\left(\varphi_{1}^{\circ},\varphi_{2}^{\circ},\pi_{3}\left(\begin{pmatrix}\varpi^{-n}&0\\ 0&1\end{pmatrix}\right)\varphi_{3}^{\circ}\right)\right|\leq_{v,\epsilon}\frac{1}{q^{\left(1/2-|\operatorname{Re}(\gamma)|-\epsilon\right)n}}.

When Re(γ)0\operatorname{Re}(\gamma)\leq 0,

(6.8) |IRS(φ1,φ2,π3((ϖn001))φ3)|v,ϵ1qc/21q(1/2|Re(γ)|ϵ)(nc).\left|I^{\text{RS}}\left(\varphi_{1}^{\circ},\varphi_{2}^{\circ},\pi_{3}\left(\begin{pmatrix}\varpi^{-n}&0\\ 0&1\end{pmatrix}\right)\varphi_{3}^{\circ}\right)\right|\leq_{v,\epsilon}\frac{1}{q^{c^{\prime}/2}}\frac{1}{q^{\left(1/2-|\operatorname{Re}(\gamma)|-\epsilon\right)(n-c^{\prime})}}.

For general γ\gamma,

(6.9) |IT(φ1,φ2,π3((ϖn001))φ3)|v,ϵ1qc/21q(1/2|Re(γ)|ϵ)(2nc).\left|I^{T}\left(\varphi_{1}^{\circ},\varphi_{2}^{\circ},\pi_{3}\left(\begin{pmatrix}\varpi^{-n}&0\\ 0&1\end{pmatrix}\right)\varphi_{3}^{\circ}\right)\right|\leq_{v,\epsilon}\frac{1}{q^{c^{\prime}/2}}\frac{1}{q^{\left(1/2-|\operatorname{Re}(\gamma)|-\epsilon\right)(2n-c^{\prime})}}.
Proof.

Note that (6.9) follows from (6.7), (6.8) and Lemma 6.2, as η1=||γ\eta^{-1}=|\cdot|^{-\gamma}.

On the other hand, (6.7) and (6.8) follows from (6.5), and that Pi(W1,W2)1P_{i}(W_{1},W_{2})\ll 1 for ic/2i\geq c/2 according to Lemma 5.2 Proposition 5.4, and Lemma 6.1 for i<c/2i<c/2 with a direct computation. When |Re(γ)|<7/64|\operatorname{Re}(\gamma)|<7/64, the main contributions come from i=ni=n when Re(γ)>0\operatorname{Re}(\gamma)>0, c/2inc/2\leq i\leq n when Re(γ)=0\operatorname{Re}(\gamma)=0, and i=c/2=c/2i=\lceil c/2\rceil=c^{\prime}/2 when Re(γ)<0\operatorname{Re}(\gamma)<0. Note that the conclusions hold even when n>cn>c. ∎

6.3. Lower bounds

Our main ingredient for the lower bound of period integrals is Proposition 5.4, thus we assume for now Assumption 5.3. In that case we choose the following test vectors:

Choice of test vectors.

In all cases, we take φi=φi\varphi_{i}=\varphi_{i}^{\circ} to be newforms for i=1,2i=1,2. Take φ3=π3((ϖn1))φ3\varphi_{3}=\pi_{3}\left(\begin{pmatrix}\varpi^{-n}&\\ &1\end{pmatrix}\right)\varphi_{3}^{\circ} with n=c(π1×π2)/2n=c(\pi_{1}\times\pi_{2})/2 when pp is large enough, and n=c(π1×π2)/2+an=c(\pi_{1}\times\pi_{2})/2+a for some non-negative absolutely bounded integer aa when pp is bounded.

Lemma 6.4.

Suppose that c(π1)<c(π2)=cc(\pi_{1})<c(\pi_{2})=c and φiπi\varphi_{i}\in\pi_{i} are L2L^{2}-normalized newforms for i=1,2i=1,2. Suppose that either pp or cc is large enough. Then

Pc(Wφ1,Wφ2)1,|Pc1(Wφ1,Wφ2)|1q1|Pc(Wφ1,Wφ2)|P_{c}(W_{\varphi_{1}},W_{\varphi_{2}})\asymp 1,\left|P_{c-1}(W_{\varphi_{1}},W_{\varphi_{2}})\right|\leq\frac{1}{q-1}\left|P_{c}(W_{\varphi_{1}},W_{\varphi_{2}})\right|

with implied constants controlled by QϵQ^{\epsilon}.

Proof.

It suffices to do case-by-case computations. For φ1\varphi_{1}, we only need its information Wφ1(c(π1))W^{(c(\pi_{1}))}_{\varphi_{1}} to compute PcP_{c} and Pc1P_{c-1} as c(π1)<cc(\pi_{1})<c. This is provided by Lemma 4.14, 4.16, 4.19, and Wφ1(c(π1))=char(OF×)W^{(c(\pi_{1}))}_{\varphi_{1}}={\text{char}}(O_{F}^{\times}) whenever c(π1)2c(\pi_{1})\geq 2 (this is true even when π1\pi_{1} is a non-dihedral supercuspidal representation). For φ2\varphi_{2}, we apply Lemma 4.16 when c=1c=1, and Lemma 4.17, Corollary 4.18/Corollary 4.21, 4.22 when c2c\geq 2, as we avoid the non-dihedral supercuspidal representations by the assumption pp or cc large enough.

Here we only verify the case where c(π1)=1c(\pi_{1})=1, c2c\geq 2, and left the remaining cases to readers. In this case, as Wφ2(c)W^{(c)}_{\varphi_{2}} is again char(OF×){\text{char}}(O_{F}^{\times}),

P2(Wφ1,Wφ2)\displaystyle P_{2}(W_{\varphi_{1}},W_{\varphi_{2}}) =xF×Wφ1(2)(x)Wφ2(2)(x)¯|x|1/2+γd×x=xOF×Wφ1(c(π1))(x)d×x=1\displaystyle=\int\limits_{x\in{F}^{\times}}W_{\varphi_{1}}^{(2)}(x)\overline{W_{\varphi_{2}}^{(2)}(x)}|x|^{-1/2+\gamma}d^{\times}x=\int\limits_{x\in O_{F}^{\times}}W_{\varphi_{1}}^{(c(\pi_{1}))}(x)d^{\times}x=1

by Lemma 4.16. As remarked before the formula in Lemma 4.16 is not L2L^{2}-normalized but the resulting difference is controlled by QϵQ^{\epsilon}.

On the other hand by Lemma 4.16, we have

P1(Wφ1,Wφ2)\displaystyle P_{1}(W_{\varphi_{1}},W_{\varphi_{2}}) =j0q(1/2γ)jv(x)=jWφ1(1)(x)Wφ2(2)(x)¯d×x\displaystyle=\sum\limits_{j\geq 0}q^{(1/2-\gamma)j}\int\limits_{v(x)=j}W_{\varphi_{1}}^{(1)}(x)\overline{W_{\varphi_{2}}^{(2)}(x)}d^{\times}x
=j0μ(ϖ)jq(1/2γ)jv(x)=jWφ2(2)(x)d×x¯.\displaystyle=\sum\limits_{j\geq 0}\mu(\varpi)^{j}q^{(-1/2-\gamma)j}\overline{\int\limits_{v(x)=j}W_{\varphi_{2}}^{(2)}(x)d^{\times}x}.

By taking χ=1\chi=1 in Corollary 4.18/ 4.22, we have

v(x)=jWφ2(2)(x)d×x={1q1, if j=0,0, otherwise.\int\limits_{v(x)=j}W_{\varphi_{2}}^{(2)}(x)d^{\times}x=\begin{cases}-\frac{1}{q-1},\text{\ if $j=0$},\\ 0,\text{\ otherwise}.\end{cases}

Thus P1(Wφ1,Wφ2)=1q1P_{1}(W_{\varphi_{1}},W_{\varphi_{2}})=-\frac{1}{q-1} in this case. ∎

Proposition 6.5.

Suppose that π1\pi_{1} π2\pi_{2} satisfy Assumption 5.3. Suppose that either pp is large enough or c(π1×π2)c(\pi_{1}\times\pi_{2}) is large enough. Then there exists a choice of test vectors as specified above, such that IRS(φ1,φ2,φ3)I^{\text{RS}}\left(\varphi_{1},\varphi_{2},\varphi_{3}\right) and IT(φ1,φ2,φ3)I^{T}\left(\varphi_{1},\varphi_{2},\varphi_{3}\right) are non-zero and satisfy the lower bounds

(6.10) |IRS(φ1,φ2,φ3)|v,ϵ1q(1/2Re(γ))c(π1×π2)/2,\left|I^{\text{RS}}\left(\varphi_{1},\varphi_{2},\varphi_{3}\right)\right|\geq_{v,\epsilon}\frac{1}{q^{(1/2-\operatorname{Re}(\gamma))c(\pi_{1}\times\pi_{2})/2}},
(6.11) |IT(φ1,φ2,φ3)|v,ϵ1qc(π1×π2)/2.\left|I^{T}\left(\varphi_{1},\varphi_{2},\varphi_{3}\right)\right|\geq_{v,\epsilon}\frac{1}{q^{c(\pi_{1}\times\pi_{2})/2}}.

Note that when pp and c(π1×π2)c(\pi_{1}\times\pi_{2}) are both bounded, the size of local integral does not affect the asymptotic analysis.

We shall see below that the ambiguity for the choice of nn when pp is small is due to the fact that the supposed main term in the computation could be smaller than the supposed error terms. But at least one of the choices for nn will give the right size of the local integrals.

Proof.

(6.11) follows from Lemma 6.2 and applying (6.10) for γ\gamma and γ-\gamma. In the following we focus on proving (6.10).

If c(π1)c(π2)c(\pi_{1})\neq c(\pi_{2}), we can assume without loss of generality that c(π1)<c(π2)=cc(\pi_{1})<c(\pi_{2})=c. In this case c(π1×π2)=2cc(\pi_{1}\times\pi_{2})=2c, and we take

φ3=π3((ϖc1))φ3.\varphi_{3}=\pi_{3}\left(\begin{pmatrix}\varpi^{-c}&\\ &1\end{pmatrix}\right)\varphi_{3}^{\circ}.

We start with (6.5). By Lemma 4.14, 4.16, 4.17, 4.19 and 4.20, Pi(Wφ1,Wφ2)=0P_{i}(W_{\varphi_{1}},W_{\varphi_{2}})=0 for any i<c1i<c-1 due to the difference of levels when c2c\geq 2, or empty set when c=1c=1. Thus when pp is large enough or cc is large enough, we have by Lemma 6.4

IRS(φ1,φ2,φ3)\displaystyle I^{\text{RS}}(\varphi_{1},\varphi_{2},\varphi_{3}) v,ϵAcqc(1/2+γ)Ac1q(c2)(1/2+γ)1q1\displaystyle\gg_{v,\epsilon}A_{c}q^{c(1/2+\gamma)}-A_{c-1}q^{(c-2)(1/2+\gamma)}\frac{1}{q-1}
=\displaystyle= 1(q+1)qc1qc(1/2+γ)(11q1+2γ)1q(1/2Re(γ))c.\displaystyle\frac{1}{(q+1)q^{c-1}}q^{c(1/2+\gamma)}\left(1-\frac{1}{q^{1+2\gamma}}\right)\gg\frac{1}{q^{(1/2-\operatorname{Re}(\gamma))c}}.

If c(π1)=c(π2)c(\pi_{1})=c(\pi_{2}), we again start with (6.5). There are more cases to consider here.

  1. (1)

    Case c(πi)=1c(\pi_{i})=1 with pp large enough. Suppose that πi=σ(μi||1/2,μi||1/2)\pi_{i}=\sigma(\mu_{i}|\cdot|^{1/2},\mu_{i}|\cdot|^{-1/2}) for c(μi)=0c(\mu_{i})=0 and μi2=1\mu_{i}^{2}=1. Using Lemma 4.16, 5.2 and 6.1 we have for n=1n=1,

    IRS=A1q1/2+γP1(W1,W2)+A0q1/2γP0(W1,W2)q1/2+Re(γ).I^{\text{RS}}=A_{1}q^{1/2+\gamma}P_{1}(W_{1},W_{2})+A_{0}q^{-1/2-\gamma}P_{0}(W_{1},W_{2})\gg q^{-1/2+\operatorname{Re}(\gamma)}.

    Note that c(π1×π2)=2c(\pi_{1}\times\pi_{2})=2 in this case. Thus (6.10) is true.

  2. (2)

    Case c(π1×π2)/2c/2+1c(\pi_{1}\times\pi_{2})/2\geq c/2+1, pp large enough. Using Proposition 5.4 and Lemma 4.10, we have for n=c(π1×π2)/2n=c(\pi_{1}\times\pi_{2})/2,

    I1:=|i=ncAiqn(1/2+γ)Pi(Wφ1,Wφ2)|=qn/2+1qnRe(γ)q+1,I_{1}:=\left|\sum\limits_{i=n}^{c}A_{i}q^{n(1/2+\gamma)}P_{i}(W_{\varphi_{1}},W_{\varphi_{2}})\right|=\frac{q^{-n/2+1}q^{n\operatorname{Re}(\gamma)}}{q+1},

    as Pi(Wφ1,Wφ2)=1P_{i}(W_{\varphi_{1}},W_{\varphi_{2}})=1 for each ii in the range. This part is expected to be the main term. On the other hand using Lemma 6.1,

    I2:=|i=0cnAiq(2in)(1/2+γ)Pi(Wφ1,Wφ2)|\displaystyle I_{2}:=\left|\sum\limits_{i=0}^{c-n}A_{i}q^{(2i-n)(1/2+\gamma)}P_{i}(W_{\varphi_{1}},W_{\varphi_{2}})\right|
    \displaystyle\leq qcn+1q+1qn(1/2+Re(γ))qc(1/2Re(γ))=qc/23n/2+1q(cn)Re(γ)q+1.\displaystyle\frac{q^{c-n+1}}{q+1}q^{-n(1/2+\operatorname{Re}(\gamma))}q^{-c(1/2-\operatorname{Re}(\gamma))}=\frac{q^{c/2-3n/2+1}q^{(c-n)\operatorname{Re}(\gamma)}}{q+1}.

    There will also be contributions coming from i=n1i=n-1 and i=cn+1i=c-n+1 in this case, which are

    I3:\displaystyle I_{3}: =|An1q(n2)(1/2+γ)Pn1(Wφ1,Wφ2)+Acn+1q(2c3n+2)(1/2+γ)Pcn+1(Wφ1,Wφ2)|\displaystyle=\left|A_{n-1}q^{(n-2)(1/2+\gamma)}P_{n-1}(W_{\varphi_{1}},W_{\varphi_{2}})+A_{c-n+1}q^{(2c-3n+2)(1/2+\gamma)}P_{c-n+1}(W_{\varphi_{1}},W_{\varphi_{2}})\right|
    qn/2q(n2)Re(γ)q+1+qc/23n/2+1q(cn)Re(γ)q+1\displaystyle\leq\frac{q^{-n/2}q^{(n-2)\operatorname{Re}(\gamma)}}{q+1}+\frac{q^{c/2-3n/2+1}q^{(c-n)\operatorname{Re}(\gamma)}}{q+1}

    Note that this upper bound is still correct when c(π1×π2)/2=c/2+1c(\pi_{1}\times\pi_{2})/2=c/2+1. The remaining ii’s (if exist) have zero contribution by Proposition 5.4. Then we have

    |IRS|I1I2I3=qn/2qnRe(γ)q+1(qq2Re(γ)2qc/2n+1q(c2n)Re(γ)).\displaystyle\left|I^{\text{RS}}\right|\geq I_{1}-I_{2}-I_{3}=\frac{q^{-n/2}q^{n\operatorname{Re}(\gamma)}}{q+1}\left(q-q^{-2\operatorname{Re}(\gamma)}-2q^{c/2-n+1}q^{(c-2n)\operatorname{Re}(\gamma)}\right).

    When nc/2+1n\geq c/2+1, and p5p\geq 5, one can easily verify that

    |IRS|qn/2qnRe(γ).\left|I^{\text{RS}}\right|\gg q^{-n/2}q^{n\operatorname{Re}(\gamma)}.

    On the other hand, if p=3p=3 and nc+32n\geq\frac{c+3}{2}, then one can also directly verify that the same bound holds.

  3. (3)

    Case cc is odd, c(π1×π2)/2=c+12>1c(\pi_{1}\times\pi_{2})/2=\frac{c+1}{2}>1 and pp large enough. This case happens when both πi\pi_{i} are supercuspidal representations associated to characters over ramified field extensions. We directly have for n=c(π1×π2)/2n=c(\pi_{1}\times\pi_{2})/2 that

    |IRS|I1I2qn/2qnRe(γ)q+1(qq1/2Re(γ))qn/2+nRe(γ)\left|I^{\text{RS}}\right|\geq I_{1}-I_{2}\geq\frac{q^{-n/2}q^{n\operatorname{Re}(\gamma)}}{q+1}\left(q-q^{1/2-\operatorname{Re}(\gamma)}\right)\gg q^{-n/2+n\operatorname{Re}(\gamma)}

    when 2p2\nmid p. Here I1I_{1}, I2I_{2} are defined similarly as in the previous case.

  4. (4)

    Case cc is even, c(π1×π2)/2=c/2c(\pi_{1}\times\pi_{2})/2=c/2 and pp large enough. If πi\pi_{i} are supercuspidal representations, we have by Proposition 5.4 that Pi(Wφ1,Wφ2)=1P_{i}(W_{\varphi_{1}},W_{\varphi_{2}})=1 for ic/2i\geq c/2. So for n=c/2n=c/2, we have

    I1:=i=c/2+1cAiqn(1/2+Re(γ))Pi(Wφ1,Wφ2)=qc/4qcRe(γ)/2q+1,I_{1}:=\sum\limits_{i=c/2+1}^{c}A_{i}q^{n(1/2+\operatorname{Re}(\gamma))}P_{i}(W_{\varphi_{1}},W_{\varphi_{2}})=\frac{q^{-c/4}q^{c\operatorname{Re}(\gamma)/2}}{q+1},
    I2:=|i=0c/21Aiq(2in)(1/2+γ)Pi(Wφ1,Wφ2)|qc/4qcRe(γ)/2q+1,I_{2}:=\left|\sum\limits_{i=0}^{c/2-1}A_{i}q^{(2i-n)(1/2+\gamma)}P_{i}(W_{\varphi_{1}},W_{\varphi_{2}})\right|\leq\frac{q^{-c/4}q^{c\operatorname{Re}(\gamma)/2}}{q+1},
    I3:=Ac/2qc/2(1/2+Re(γ))Pc/2(Wφ1,Wφ2)=(q1)qc/4qcRe(γ)/2q+1.I_{3}:=A_{c/2}q^{c/2(1/2+\operatorname{Re}(\gamma))}P_{c/2}(W_{\varphi_{1}},W_{\varphi_{2}})=\frac{(q-1)q^{-c/4}q^{c\operatorname{Re}(\gamma)/2}}{q+1}.

    Thus for any 2p2\nmid p, I3I_{3} is the main term and we have

    |IRS|I1+I3I2qc/4qcRe(γ)/2.\left|I^{\text{RS}}\right|\geq I_{1}+I_{3}-I_{2}\gg q^{-c/4}q^{c\operatorname{Re}(\gamma)/2}.

    When πi\pi_{i} are principal series, I1,I_{1}, I2I_{2} can be controlled similarly, while I3I_{3} is slightly more complicated. By Proposition 5.4, we have

    |I3|(q1)qc/4qcRe(γ)/2q+1(q3q12q1/2+Re(γ)1).\left|I_{3}\right|\geq\frac{(q-1)q^{-c/4}q^{c\operatorname{Re}(\gamma)/2}}{q+1}\left(\frac{q-3}{q-1}-\frac{2}{q^{1/2+\operatorname{Re}(\gamma)}-1}\right).

    One can check that when, for example, p23p\geq 23, we have

    |I3|qc/4qcRe(γ)/2,|IRS|I1+I3I2qc/4qcRe(γ)/2.|I_{3}|\gg q^{-c/4}q^{c\operatorname{Re}(\gamma)/2},\ \left|I^{\text{RS}}\right|\gg I_{1}+I_{3}-I_{2}\gg q^{-c/4}q^{c\operatorname{Re}(\gamma)/2}.
  5. (5)

    Consider now the case pp is bounded (p23p\leq 23 for example, including the case p=2p=2), and cc is large enough. Denote n0=c(π1×π2)/2+an_{0}=c(\pi_{1}\times\pi_{2})/2+a, where aa is as in Lemma 5.10 when p=2p=2, or a=1a=1 when p>2p>2. Thus Pi(W1,W2)=1P_{i}(W_{1},W_{2})=1 whenever in0i\geq n_{0} by Lemma 5.10 or Proposition 5.4.

    Let c=max{c,n0+1}c^{\prime}=\max\{c,n_{0}+1\}. Using Lemma 4.10 for cc^{\prime}, we compare the calculation of (6.5) for n=n0n=n_{0} and n=n0+1n=n_{0}+1.

    In0:\displaystyle I_{n_{0}}: =IRS(φ1,φ2,π3((ϖn01))φ3)\displaystyle=I^{\text{RS}}\left(\varphi_{1},\varphi_{2},\pi_{3}\left(\begin{pmatrix}\varpi^{-n_{0}}&\\ &1\end{pmatrix}\right)\varphi_{3}^{\circ}\right)
    =i=n0cAiqn0(1/2+γ)Pi(Wφ1,Wφ2)+i=0n01Aiq(2in0)(1/2+γ)Pi(Wφ1,Wφ2)\displaystyle=\sum\limits_{i=n_{0}}^{c^{\prime}}A_{i}q^{n_{0}(1/2+\gamma)}P_{i}(W_{\varphi_{1}},W_{\varphi_{2}})+\sum\limits_{i=0}^{n_{0}-1}A_{i}q^{(2i-n_{0})(1/2+\gamma)}P_{i}(W_{\varphi_{1}},W_{\varphi_{2}})
    =i=n0+1cAiqn0(1/2+γ)Pi(Wφ1,Wφ2)+i=0n0Aiq(2in0)(1/2+γ)Pi(Wφ1,Wφ2),\displaystyle=\sum\limits_{i=n_{0}+1}^{c^{\prime}}A_{i}q^{n_{0}(1/2+\gamma)}P_{i}(W_{\varphi_{1}},W_{\varphi_{2}})+\sum\limits_{i=0}^{n_{0}}A_{i}q^{(2i-n_{0})(1/2+\gamma)}P_{i}(W_{\varphi_{1}},W_{\varphi_{2}}),
    In0+1:\displaystyle I_{n_{0}+1}: =IRS(φ1,φ2,π3((ϖn011))φ3)\displaystyle=I^{\text{RS}}\left(\varphi_{1},\varphi_{2},\pi_{3}\left(\begin{pmatrix}\varpi^{-n_{0}-1}&\\ &1\end{pmatrix}\right)\varphi_{3}^{\circ}\right)
    =i=n0+1cAiq(n0+1)(1/2+γ)Pi(Wφ1,Wφ2)+i=0n0Aiq(2in01)(1/2+γ)Pi(Wφ1,Wφ2).\displaystyle=\sum\limits_{i=n_{0}+1}^{c^{\prime}}A_{i}q^{(n_{0}+1)(1/2+\gamma)}P_{i}(W_{\varphi_{1}},W_{\varphi_{2}})+\sum\limits_{i=0}^{n_{0}}A_{i}q^{(2i-n_{0}-1)(1/2+\gamma)}P_{i}(W_{\varphi_{1}},W_{\varphi_{2}}).

    We can do a linear combination of two equations to cancel the terms with in0i\leq n_{0}. In particular we have

    In0q1/2+γIn0+1=i=n0+1cAiqn0(1/2+γ)(1q1+γ)Pi(Wφ1,Wφ2).I_{n_{0}}-q^{1/2+\gamma}I_{n_{0}+1}=\sum\limits_{i=n_{0}+1}^{c^{\prime}}A_{i}q^{n_{0}(1/2+\gamma)}(1-q^{1+\gamma})P_{i}(W_{\varphi_{1}},W_{\varphi_{2}}).

    Using that Pi(Wφ1,Wφ2)=1P_{i}(W_{\varphi_{1}},W_{\varphi_{2}})=1 for in0i\geq n_{0}, and that qq is bounded in this case, we get that

    |In0q1/2+γIn0+1|qn0(1/2+Re(γ)).\left|I_{n_{0}}-q^{1/2+\gamma}I_{n_{0}+1}\right|\gg q^{n_{0}(-1/2+\operatorname{Re}(\gamma))}.

    Then we either have

    |In0|qn0(1/2+Re(γ)),\left|I_{n_{0}}\right|\gg q^{n_{0}(-1/2+\operatorname{Re}(\gamma))},

    or

    |In0+1|qn0(1/2+Re(γ)).\left|I_{n_{0}+1}\right|\gg q^{n_{0}(-1/2+\operatorname{Re}(\gamma))}.

6.4. Proof of Theorem 3.22

We can now prove Theorem 3.22, that is, verify Conjecture 3.17 under the conditions that πi\pi_{i} have trivial central characters and bounded Archimedean components, and Mf=1M_{f}=1 at finite places. Recall that I=ITI=I^{T} or |IRS|2\left|I^{\text{RS}}\right|^{2} depending on whether φ3\varphi_{3} or ψ\psi comes from an Eisenstein series.

Note that the formulations in Conjecture 3.17 are symmetric in index i=2,3i=2,3 and are local, we may assume without loss of generality that C(π3)=1C(\pi_{3})=1.

We first address the issue about potential conflicts with Assumption 5.3. Suppose now without loss of generality that c(π1)=c(π2)=2c(\pi_{1})=c(\pi_{2})=2, π1=π1χ\pi_{1}=\pi_{1}^{\prime}\otimes\chi for a quadratic character χ\chi and an unramified representation π1\pi_{1}^{\prime}, not satisfying Assumption 5.3. Let π2=π2χ\pi_{2}^{\prime}=\pi_{2}\otimes\chi. Then πi\pi_{i}^{\prime} still have trivial central characters, and

C(πi×πj)=C(πi×πj)C(\pi_{i}\times\pi_{j})=C(\pi_{i}^{\prime}\times\pi_{j}^{\prime})

for i,j=1,2i,j=1,2. One can now reduce the test vector problem to that of the triple π1,π2,π3\pi_{1}^{\prime},\pi_{2}^{\prime},\pi_{3}, which satisfies Assumption 5.3 now.

With Assumption 5.3 we take now φi=πi((ϖn1))φi\varphi_{i}=\pi_{i}\left(\begin{pmatrix}\varpi^{n}&\\ &1\end{pmatrix}\right)\varphi_{i}^{\circ} for i=1,2i=1,2, and φ3=φ3\varphi_{3}=\varphi_{3}^{\circ}, where nn is as in Proposition 6.5. Thus φ3\varphi_{3} satisfies item (0) of Conjecture 3.17.

Up to a uniform translates of test vectors which does not change the period integrals, the required lower bound in item (1) of Conjecture 3.17 is provided by Proposition 6.5. Note that in the case of Rankin–Selberg L-function, we can take Re(γ)=0\operatorname{Re}(\gamma)=0 for (6.10) when we assume π3\pi_{3} to be tempered.

Now in the spectral expansion in Lemma 3.11, π\pi^{\prime} must be unramified and ψ\psi must be spherical for vram(Φ23)v\in\mathrm{ram}(\varPhi_{23}) as Mf=1M_{f}=1. The required upper bound in item (2) of Conjecture 3.17 follows from Proposition 6.3. Indeed, in the triple product case and ψ=ψ\psi=\psi^{\circ}, we have

I(φ2,φ2,ψ)\displaystyle I(\varphi_{2},\varphi_{2},\psi) =IT(π2((ϖn1))φ2,π2((ϖn1))φ2,ψ)\displaystyle=I^{T}\left(\pi_{2}\left(\begin{pmatrix}\varpi^{n}&\\ &1\end{pmatrix}\right)\varphi_{2}^{\circ},\pi_{2}\left(\begin{pmatrix}\varpi^{n}&\\ &1\end{pmatrix}\right)\varphi_{2}^{\circ},\psi^{\circ}\right)
=IT(φ2,φ2,π((ϖn1))ψ)\displaystyle=I^{T}\left(\varphi_{2}^{\circ},\varphi_{2}^{\circ},\pi^{\prime}\left(\begin{pmatrix}\varpi^{-n}&\\ &1\end{pmatrix}\right)\psi^{\circ}\right)
ϵQϵqθc(π2×π2)1Q1/2(1/2θ)Qϵmaxi=2,3{C(πi×πi)}1/21P1/2θ.\displaystyle\leq_{\epsilon}\frac{Q^{\epsilon}}{q^{\theta c(\pi_{2}\times\pi_{2})}}\frac{1}{Q^{1/2(1/2-\theta)}}\leq\frac{Q^{\epsilon}}{\max\limits_{i=2,3}\{C(\pi_{i}\times\pi_{i})\}^{1/2}}\frac{1}{P^{1/2-\theta}}.

Here in the third line we have applied (6.9) and that q2nϵQ1/2+ϵq^{2n}\asymp_{\epsilon}Q^{1/2+\epsilon} for the choice of nn above.

The Rankin–Selberg case can be verified similarly using (6.7) from Proposition 6.3 and taking Re(γ)=0\operatorname{Re}(\gamma)=0.

References