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

Short Second Moment Bound for GL(2) LL-functions in qq-Aspect

Agniva Dasgupta
Abstract.

We prove a Lindelöf-on-average upper bound for the second moment of the LL-functions associated to a level 1 holomorphic cusp form, twisted along a coset of subgroup of the characters modulo q2/3q^{2/3} (where q=p3q=p^{3} for some odd prime pp). This result should be seen as a qq-aspect analogue of Anton Good’s (1982) result on upper bounds of the second moment of cusp forms in short intervals.

1. Introduction

1.1. Statement of Results

The study of moments of LL-functions at the central point, s=12s=\frac{1}{2}, has been an important area of research in analytic number theory. While originally this interest grew due to connections with the Lindelöf Hypothesis, estimation of similar moments for a family of LL-functions has now become an interesting point of study on its own.

One key result in this area is the following proposition, due to Iwaniec in 1978 (Theorem 33 in [Iwa80]).

Proposition 1.1.

For T2T\geq 2 and ε>0\varepsilon>0,

(1.1) TT+T23|ζ(12+it)|4𝑑tεT23+ε.\int_{T}^{T+T^{\frac{2}{3}}}\lvert\zeta\left(\tfrac{1}{2}+it\right)\rvert^{4}dt\ll_{\varepsilon}T^{\frac{2}{3}+\varepsilon}.

Proposition 1.1 is one of the first examples of an upper bound on a ‘short’ moment of LL-functions. Most of the earlier results were concerned with estimating integrals of the type 0T|ζ(12+it)|k𝑑t\int_{0}^{T}\lvert\zeta\left(\frac{1}{2}+it\right)\rvert^{k}dt, for some positive even integer kk.

Note that (1.1) is consistent with the Lindelöf Hypothesis, and is an example of a Lindelöf-on-average upper bound. For level 11 holomorphic cusp forms, an analogous result follows from Good [Goo82],

Proposition 1.2.

Let ff be a fixed level 11 holomorphic cusp form. For T2T\geq 2 and ε>0\varepsilon>0,

(1.2) TT+T23|L(f,12+it)|2𝑑tf,εT23+ε.\int_{T}^{T+T^{\frac{2}{3}}}\lvert L\left(f,\tfrac{1}{2}+it\right)\rvert^{2}dt\ll_{f,\varepsilon}T^{\frac{2}{3}+\varepsilon}.

Equations (1.1) and (1.2) imply a Weyl-type subconvexity bound for the respective LL functions, ζ()\zeta(\cdot), and L(f,)L(f,\cdot). In fact, for holomorphic cusp forms, [Goo82] was the first instance where such a result was obtained.

In their paper on the Weyl bound for Dirichlet LL-functions [PY23], Petrow and Young, proved a qq-aspect analogue of Proposition 1.1. A special case of Theorem 1.41.4 (with q=p3,d=p2q=p^{3},d=p^{2}) in [PY23] can be stated as follows.

Proposition 1.3.

Let q=p3q=p^{3}, for an odd prime pp. Let α\alpha be a fixed primitive character mod qq. For any ε>0\varepsilon>0, we have

(1.3) ψ(modq23)|L(αψ,12)|4εq23+ε.\sum_{\psi\negthickspace\negthickspace\negthickspace\pmod{q^{\frac{2}{3}}}}\lvert L\left(\alpha\cdot\psi,\tfrac{1}{2}\right)\rvert^{4}\ll_{\varepsilon}q^{\frac{2}{3}+\varepsilon}.

We discuss this analogy in more detail in Section 1.4.

In this paper, following up on the ideas in [PY23], we derive a qq-aspect analogue of Good’s result in Proposition 1.2. We prove the following theorem.

Theorem 1.4.

Let ff be a level 11 cusp form. Let q=p3q=p^{3}, for an odd prime pp, and let α\alpha be a primitive character modulo qq. For any ε>0\varepsilon>0,

(1.4) ψ(modq23)|L(f(αψ),12)|2f,εq23+ε.\sum_{\psi(\mathrm{mod}\ q^{\frac{2}{3}})}\lvert L\left(f\otimes\left(\alpha\cdot\psi\right),\tfrac{1}{2}\right)\rvert^{2}\ll_{f,\varepsilon}q^{\frac{2}{3}+\varepsilon}.

We note that, similar to Propositions 1.1, 1.2, 1.3, this is also a Lindelöf-on-average bound.

Also, even though Theorem 1.4 is stated for q=p3q=p^{3}, we expect this to generalise to a short moment exactly analogous to the Theorem 1.41.4 [PY23]. While the authors in [PY23] needed the fully general result to prove the Weyl bound for all Dirichlet LL-functions (see also [PY20]), we do not have any such demand. So, for the sake of simplicity, we choose to work with q=p3q=p^{3}.

Using a lower bound on the associated first moment, we can deduce from Theorem 1.4 the following result about non-vanishing of LL-functions within this family.

Theorem 1.5.

Let f,p,q,αf,p,q,\alpha be as in Theorem 1.4. Then for ε>0\varepsilon>0,

(1.5) #{ψ(modq23);L(f(αψ),12)0}εp2ε.\#\{\psi\negthickspace\negthickspace\negthickspace\pmod{q^{\frac{2}{3}}};\ L\left(f\otimes\left(\alpha\cdot\psi\right),\tfrac{1}{2}\right)\neq 0\}\gg_{\varepsilon}p^{2-\varepsilon}.

Moments of a family of LL-functions encode a lot of information about the individual members of the family. For one such example, the strength of the second moment bound in (1.4)\eqref{eq:MainThm} is enough to immediately deduce a Weyl-type subconvexity bound for individual LL-functions in this family. This result was originally proven by Munshi and Singh in 2019 (Theorem 1.11.1 in [MS19] with r=1,t=0r=1,t=0). The authors use a completely different approach. They do not compute moments for an associated family; relying instead on a novel variant of the circle method, first introduced in [Mun14].

We state their result as a corollary.

Corollary 1.6.

Let f,p,q,αf,p,q,\alpha be as in Theorem 1.4. Let ψ\psi be a character modulo q23q^{\frac{2}{3}}. Then for any ε>0\varepsilon>0,

(1.6) L(f(αψ),12)f,εq13+ε.L\left(f\otimes\left(\alpha\cdot\psi\right),\tfrac{1}{2}\right)\ll_{f,\varepsilon}q^{\frac{1}{3}+\varepsilon}.

1.2. Notations

We use standard conventions present in analytic number theory. We use ε\varepsilon to denote an arbitrarily small positive constant. For brevity of notation, we allow ε\varepsilon to change depending on the context.
The expression FGF\ll G implies there exists some constant kk for which |F|kG\lvert F\rvert\leq k\cdot G for all the relevant FF and GG. We use FεGF\ll_{\varepsilon}G to emphasize that the implied constant kk depends on ε\varepsilon (it may also depend on other parameters). For error terms, we often use the big OO notation, so f(x)=O(g(x))f(x)=O(g(x)) implies that f(x)g(x)f(x)\ll g(x) for sufficiently large xx. We use the term ‘small’ or ‘very small’ to refer to error terms which are of the size OA(pA)O_{A}(p^{-A}), for any arbitrarily large AA.
By a dyadic interval, we mean an interval of the type [2k2M,2k+12M][2^{\frac{k}{2}}M,2^{\frac{k+1}{2}}M] for some kk\in\mathbb{Z}, MM\in\mathbb{R}. We also use mM0m\asymp M_{0} to mean mm ranges over the dyadic interval [M0,2M0][M_{0},2M_{0}].
We also use n\sideset{}{{}^{*}}{\sum}_{n} to denote ngcd(n,p)=1\smashoperator[]{\sum_{\begin{subarray}{c}n\\ \text{gcd}(n,p)=1\end{subarray}}^{}} . Similarly, a(modc)\sideset{}{{}^{*}}{\sum}_{a(\mathrm{mod}\ c)} is used for a(modc)gcd(a,c)=1\smashoperator[]{\sum_{\begin{subarray}{c}a(\mathrm{mod}\ c)\\ \text{gcd}(a,c)=1\end{subarray}}^{}} , and χ(p)\sideset{}{{}^{*}}{\sum}_{\chi(p)} for χ(modp)χ primitive\smashoperator[]{\sum_{\begin{subarray}{c}\chi(\mathrm{mod}\ p)\\ \chi\text{ primitive}\end{subarray}}^{}}.
As usual, e(x)=e2πixe(x)=e^{2\pi ix}. Also, ep(x)=e(xp)=e2πixpe_{p}(x)=e(\frac{x}{p})=e^{2\pi i\frac{x}{p}}.

1.3. Sketch of Proof

We give a brief sketch of Theorem 1.4 in this section.
In this sketch, we use the symbol ‘\approx’ between two expressions to mean the expressions on the left side can be written as an expression similar to the right side, along with an acceptable error term.

Using an approximate functional equation and orthogonality of characters, (see Section 3), it suffices to prove the following result on an associated shifted convolution problem.

Theorem 1.7.

Let f,p,q,αf,p,q,\alpha be as in Theorem 1.4. Let λf()\lambda_{f}(\cdot) be the coefficients of the associated Dirichlet series. Let

(1.7) S(N,α)l,nλf(n+p2l)λf¯(n)α(n+p2l)α(n)¯wN(n+p2l)wN(n),S(N,\alpha)\coloneqq\sideset{}{{}^{*}}{\sum}_{l,n}\lambda_{f}(n+p^{2}l)\overline{\lambda_{f}}(n)\alpha(n+p^{2}l)\overline{\alpha(n)}w_{N}(n+p^{2}l)w_{N}(n),

where Nεp3+εN\ll_{\varepsilon}p^{3+\varepsilon}, and wN()w_{N}(\cdot) is some smooth function supported on [N,2N][N,2N] satisfying wN(j)(x)Njw_{N}^{(j)}(x)\ll N^{-j}. We then have

(1.8) S(N,α)f,εNpε.S(N,\alpha)\ll_{f,\varepsilon}Np^{\varepsilon}.
Remark 1.8.

As wN()w_{N}(\cdot) is supported in [N,2N][N,2N], we have that, in (1.7), 0<lNp20<l\leq\frac{N}{p^{2}}, and nNn\asymp N.

The trivial bound on S(N,α)S(N,\alpha) is N2p2pεNp1+ε\frac{N^{2}}{p^{2}}p^{\varepsilon}\ll Np^{1+\varepsilon}. To improve on this, we first note that S(N,α)S(N,\alpha) displays a conductor dropping phenomenon, notably

(1.9) α(n+p2l)α(n)¯=ep(aαln¯).\alpha(n+p^{2}l)\overline{\alpha(n)}=e_{p}(a_{\alpha}l\overline{n}).

for some non-zero aα(modp)a_{\alpha}(\mathrm{mod}\ p).

Notice that, when plp\mid l in (1.7), S(N,α)S(N,\alpha) does not have any cancellations. However, we are saved by the fact that the number of such terms is O(Npε)O(Np^{\varepsilon}). So for the rest of the sketch, we assume (l,p)=1(l,p)=1.

In order to separate the variables nn and (n+p2l)(n+p^{2}l) we introduce a delta symbol (see Section 2.3). This introduces a new averaging variable cc, and once again we split the resulting expression into whether gcd(c,p)=1(c,p)=1, or not.
The former requires more work, and we focus on this term for the sketch. We have

(1.10) S(N,α)lNp2cNa(modc)gc(v)e(p2lv)mNλf(m)e(amc)wN(m)e(mv)nNλ¯f(n)ep(aαln¯)e(anc)wN(n)e(nv)dv.S(N,\alpha)\approx\sideset{}{{}^{*}}{\sum}_{l\leq\frac{N}{p^{2}}}\sideset{}{{}^{*}}{\sum}_{c\leq\sqrt{N}}\ \sideset{}{{}^{*}}{\sum}_{a\negthickspace\negthickspace\negthickspace\pmod{c}}\int_{-\infty}^{\infty}g_{c}(v)e(-p^{2}lv)\\ \cdot\sum_{m\asymp N}\lambda_{f}(m)e\left(\frac{am}{c}\right)w_{N}(m)e(mv)\sideset{}{{}^{*}}{\sum}_{n\asymp N}\overline{\lambda}_{f}(n)e_{p}(a_{\alpha}l\overline{n})e\left(\frac{-an}{c}\right)w_{N}(n)e(-nv)\ dv.

Here, gc(v)g_{c}(v) is a smooth function which is small when v1cNv\gg\frac{1}{c\sqrt{N}}.

We can now use the Voronoi summation formula (see Sec. 2.5), for the mm and nn sums in (1.10). Note that, as there is an extra additive character ep(aαln¯)e_{p}(a_{\alpha}l\overline{n}) in the nn sum, we need to use a modified Voronoi summation formula (see Prop 2.17) here. While this modification does introduce some extra terms to the final expression, these terms are pretty easily dealt with (see Section 6.1). We get that

(1.11) S(N,α)lNp2mpεnp2+ελf(m)λf¯(n)p2cN1c2S(p¯2nm,p2l;c)Kl3(nc¯2aαl,1,1;p)IN(c,l,m,n),S(N,\alpha)\approx\sideset{}{{}^{*}}{\sum}_{\begin{subarray}{c}l\leq\frac{N}{p^{2}}\\ m\ll p^{\varepsilon}\\ n\ll p^{2+\varepsilon}\end{subarray}}\frac{\lambda_{f}(m)\overline{\lambda_{f}}(n)}{p^{2}}\sideset{}{{}^{*}}{\sum}_{c\leq\sqrt{N}}\frac{1}{c^{2}}S(\overline{p}^{2}n-m,-p^{2}l;c)\text{Kl}_{3}(-n\overline{c}^{2}a_{\alpha}l,1,1;p)I_{N}(c,l,m,n),

with S(a,b;c)=mn1(modc)e(am+bnc)S(a,b;c)=\sum_{mn\equiv 1(\mathrm{mod}\ c)}e\left(\frac{am+bn}{c}\right), and Kl3(x,y,z;c)=pqr1(modc)e(px+qy+rzc)\text{Kl}_{3}(x,y,z;c)=\sum_{pqr\equiv 1(\mathrm{mod}\ c)}e\left(\frac{px+qy+rz}{c}\right) denoting the Kloosterman and hyper-Kloosterman sums respectively. Also, IN()I_{N}(\cdot) is an integral of special functions (see (4.24)) which trivially satisfies IN(c,l,m,n)f,ε1cN32+εI_{N}(c,l,m,n)\ll_{f,\varepsilon}\frac{1}{c}N^{\frac{3}{2}+\varepsilon}.

Here, we should indicate that while the original sums were of length NN each (and Np3+εN\ll p^{3+\varepsilon}), the dual sums are significantly shorter - one is mpεm\ll p^{\varepsilon}, and the other np2+εn\ll p^{2+\varepsilon}. Even though this represents significant savings (by a factor of N2p2\frac{N^{2}}{p^{2}}), it is still not sufficient. If we use Weil’s bound for the Kloosterman sum, and Deligne’s bound for the hyper-Kloosterman sum, along with the trivial bound on IN()I_{N}(\cdot), we get (1.11) is bounded by Np34+εNp^{\frac{3}{4}+\varepsilon}. While this beats the trivial bound for (1.7), it still falls short of the desired bound by a factor of p34p^{-\frac{3}{4}}.

In order to get more cancellations, we want to use a spectral decomposition for the cc-sum. We do this by using the Bruggeman Kuznetsov formula (Prop. 2.18). Note that if gcd(r,p)=1(r,p)=1, the hyper-Kloosterman sum can be decomposed as

Kl3(r,1,1;p)=1ϕ(p)χ(p)τ(χ)3χ(r).\text{Kl}_{3}(r,1,1;p)=\frac{1}{\phi(p)}\sum_{\chi(p)}\tau(\chi)^{3}\chi(r).

Using this we rewrite (1.11) as

lNp2mpεnp2+ελf(m)λf¯(n)ϕ(p)p2χ(p)τ(χ)3χ(naαl)c1c2S(p¯(np2m),pl;c)χ¯2(c)IN(c,l,m,n).\sideset{}{{}^{*}}{\sum}_{\begin{subarray}{c}l\leq\frac{N}{p^{2}}\\ m\ll p^{\varepsilon}\\ n\ll p^{2+\varepsilon}\end{subarray}}\frac{\lambda_{f}(m)\overline{\lambda_{f}}(n)}{\phi(p)p^{2}}\sum_{\chi(p)}\tau(\chi)^{3}\chi(-na_{\alpha}l)\sideset{}{{}^{*}}{\sum}_{c}\frac{1}{c^{2}}S(\overline{p}(n-p^{2}m),-pl;c)\overline{\chi}^{2}(c)I_{N}(c,l,m,n).

We now use the Bruggeman-Kuznetsov formula for Γ0(p)\Gamma_{0}(p) with central character χ¯2\overline{\chi}^{2} at the cusps \infty and 0. We use two different versions of the integral transforms, depending on whether the integral IN(c,l,m,n)I_{N}(c,l,m,n) has oscillatory behavior, or not (see Section 2.6). We focus on the non-oscillatory case here, the other one is similar. We have

(1.12) S(N,α)Np2χ(p)χ(aα)τ(χ)3p3tjπitj(p,χ2)L(12,π¯χ)mpεnp2+ελ¯f(n)λ¯π(|p2mn|)χ(n)|p2mn|+(Holomorphic terms)+(Eisenstein terms).S(N,\alpha)\approx\frac{N}{p^{2}}\sideset{}{{}^{*}}{\sum}_{\chi(p)}\chi(a_{\alpha})\frac{\tau(\chi)^{3}}{p^{3}}\sum_{t_{j}}\sum_{\pi\in\mathcal{H}_{it_{j}}(p,\chi^{2})}L(\tfrac{1}{2},\overline{\pi}\otimes\chi)\sum_{\begin{subarray}{c}m\ll p^{\varepsilon}\\ n\ll p^{2+\varepsilon}\end{subarray}}\frac{\overline{\lambda}_{f}(n)\overline{\lambda}_{\pi}(\lvert p^{2}m-n\rvert)\chi(-n)}{\sqrt{\lvert p^{2}m-n\rvert}}\\ +\text{(Holomorphic terms)}+\text{(Eisenstein terms)}.

Here it(n,ψ)\mathcal{H}_{it}(n,\psi) denotes the set of Hecke-Maass newforms of conductor nn, central character ψ\psi, and spectral parameter itit. Analysing the associated integral transforms, it suffices to restrict (1.12) to when tjpεt_{j}\ll p^{\varepsilon}.

The contribution of the holomorphic and Eisenstein terms are similar to the Maass form term written down here.
Using the Cauchy-Schwarz inequality, it then suffices to bound the sums

tjpεχ(p)πitj(p,χ2)|L(12,π¯χ)|2,tjpεχ(p)πitj(p,χ2)np2+ε|λ¯f(n)λ¯π(|p2mn|)χ(n)|p2mn||2.\sum_{t_{j}\ll p^{\varepsilon}}\sideset{}{{}^{*}}{\sum}_{\chi(p)}\sum_{\pi\in\mathcal{H}_{it_{j}}(p,\chi^{2})}\lvert L(\tfrac{1}{2},\overline{\pi}\otimes\chi)\rvert^{2}\ ,\ \ \sum_{t_{j}\ll p^{\varepsilon}}\sideset{}{{}^{*}}{\sum}_{\chi(p)}\sum_{\pi\in\mathcal{H}_{it_{j}}(p,\chi^{2})}\sum_{n\ll p^{2+\varepsilon}}\left|\frac{\overline{\lambda}_{f}(n)\overline{\lambda}_{\pi}(\lvert p^{2}m-n\rvert)\chi(-n)}{\sqrt{\lvert p^{2}m-n\rvert}}\right|^{2}.

Using the fact that π¯χitj(p2,1)\overline{\pi}\otimes\chi\in\mathcal{H}_{it_{j}}(p^{2},1), we can get the required bounds by applying the spectral large sieve inequality on each of the two sums. The details can be seen in Section 5.

If we compare our proof and the proof of Prop 1.3, we can see that the authors in [PY23] do not use a delta symbol for the shifted convolution sum at the beginning, instead relying on an approximate functional equation-type formula for the divisor function. A more significant difference can be observed later, by considering the shape of the terms in (1.12) and equation (1.20) in [PY23]. Unlike our case, the authors get a product of three LL-functions that they then bound using Holder’s inequality. This reduces their problem to bounding the fourth moment of L -functions, which is accomplished via the use of the spectral large sieve inequality.

Remark 1.9.

We expect the proof steps outlined here to easily generalise to the case when ff is a Hecke-Maass cusp form. We do not use any result which is known for holomorphic cusp forms, but not for Maass forms. In particular, our proof does not need Deligne’s bound on the Fourier coefficients of holomorphic cusp forms.

1.4. Short Moments

Theorem 1.4 joins a growing list of results on short moments of families of LL-functions. The general idea here is to work with a subfamily of LL-functions that exhibit a conductor lowering phenomenon. Say, we start with a family \mathcal{F} of analytic LL-functions; we choose a subfamily 0\mathcal{F}_{0}, such that the quantity C(π1π2¯)C(\pi_{1}\otimes\overline{\pi_{2}}), where C(π)C(\pi) denotes the analytic conductor of π\pi, is of a comparatively smaller size, when π1,π2\pi_{1},\pi_{2} are in 0\mathcal{F}_{0}. (See Section 1.5 in [PY23] for a more detailed discussion on this).

We see an archimedean version of this in Propositions 1.1 and 1.2. The full family of LL-functions considered in Proposition 1.2, for example, is ={L(f||it,12),Tt 2T}.\mathcal{F}=\{L(f\otimes\lvert\cdot\rvert^{it},\tfrac{1}{2}),T\leq t\leq\ 2T\}. Here ff is a level 11 cusp form. The analytic conductor of L(f||it1f||it2¯)L(f\otimes\lvert\cdot\rvert^{it_{1}}\otimes\overline{f\otimes\lvert\cdot\rvert^{it_{2}}}) is proportional to |t1t2|4\lvert t_{1}-t_{2}\rvert^{4}. Now, for arbitrary t1,t2t_{1},t_{2} in [T,2T][T,2T], this can be as high as T4T^{4}. However, if we restrict to the subfamily, 0={L(f||it,12),TtT+T23}\mathcal{F}_{0}=\{L(f\otimes\lvert\cdot\rvert^{it},\tfrac{1}{2}),T\leq t\leq\ T+T^{\tfrac{2}{3}}\} - as considered in (1.2), we have that |t1t2|4T83\lvert t_{1}-t_{2}\rvert^{4}\leq T^{\tfrac{8}{3}}, which is significantly smaller than T4T^{4}. This leads to a conductor lowering analogous to (1.9).

For our work, we start with the family, ={L(fχ,12),χ(modq)}\mathcal{F}=\{L(f\otimes\chi,\tfrac{1}{2}),\chi(\mathrm{mod}\ q)\}, where q=p3q=p^{3}. Using some of the results in [Li75], we know that the analytic conductor of L((fχ1(fχ2)¯,12)L\left((f\otimes\chi_{1}\otimes\overline{(f\otimes\chi_{2})},\tfrac{1}{2}\right) depends on the fourth power of the conductor of (χ1χ¯2)(\chi_{1}\overline{\chi}_{2}). Now, for arbitrary χ1\chi_{1}, and χ2\chi_{2}, this can be as high as q4q^{4}. To get a short moment in the level aspect, we choose a subfamily of \mathcal{F} that lowers this significantly. We consider the subfamily 0={L(f(αψ),12),ψ(modq23)},\mathcal{F}_{0}=\{L(f\otimes(\alpha\cdot\psi),\tfrac{1}{2}),\psi(\mathrm{mod}\ q^{\tfrac{2}{3}})\}, for a fixed primitive character α(modq)\alpha(\mathrm{mod}\ q).Now, the analytic conductor of L((f(αψ1))(f(αψ2))¯,12)L\left((f\otimes(\alpha\cdot\psi_{1}))\otimes\overline{(f\otimes(\alpha\cdot\psi_{2}))},\tfrac{1}{2}\right) (more accurately, of L((ff¯(ψ1ψ¯2),12)L\left((f\otimes\overline{f}\otimes(\psi_{1}\overline{\psi}_{2}),\tfrac{1}{2}\right)) depends on (conductor(ψ1ψ2¯))4q83\left(\text{conductor}(\psi_{1}\overline{\psi_{2}})\right)^{4}\leq q^{\frac{8}{3}}, which is significantly lower than q4.q^{4}.

Short moments in the level aspect, in a GL(1)GL(1) setting, have also been considered in [Nun21], and in [MW21]. Another similar application was considered along Galois orbits in [KMN16].

1.5. Acknowledgements

I am deeply grateful to my doctoral advisor, Matthew P. Young, for suggesting this problem, and for engaging in extensive discussions on this subject throughout the process of working on this paper. I would also like to thank Peter Humphries and Chung-Hang Kwan for some insightful discussions regarding this work.

2. Preliminary Results from Analytic Number Theory

In this section, we note down some of the analytic number theory preliminaries we will use later. We omit proofs in most cases. Interested readers can check the references listed next to the results.

2.1. Approximate Functional Equation

We state a version of the approximate functional equation for analytic LL-functions. For proofs, see Theorem 5.35.3 and Proposition 5.45.4 in [IK04].

Proposition 2.1.

Let X>0X>0, and L(f,s)L(f,s) be an LL-function given by L(f,s)=n1λf(n)nsL(f,s)=\sum_{n\geq 1}\frac{\lambda_{f}(n)}{n^{s}} when (s)>1\Re{(s)}>1. If the completed LL-function Λ(f,s)\Lambda(f,s) is entire, then we have the following approximation for L(f,s)L\left(f,s\right), when 0(s)10\leq\Re{(s)}\leq 1 -

(2.1) L(f,s)=nλf(n)nsV(nXq)+ε(f,s)nλf¯(n)n1sV(nXq).L\left(f,s\right)=\sum_{n}\frac{\lambda_{f}(n)}{n^{s}}V\left(\frac{n}{X\sqrt{q}}\right)+\varepsilon\left(f,s\right)\sum_{n}\frac{\overline{\lambda_{f}}(n)}{n^{1-s}}V\left(\frac{nX}{\sqrt{q}}\right).

Here, qq is the conductor of L(f,s)L(f,s), and V(y)V(y) is a smooth function that satisfies the following bounds -

(2.2) V(y)(1+y𝔮)A,V(y)\ll\left(1+\frac{y}{\sqrt{\mathfrak{q}_{\infty}}}\right)^{-A},

where q𝔮q\cdot\mathfrak{q}_{\infty} is the analytic conductor of L(f,s)L(f,s) and A>0A>0.

As |ε(f,12)|=1\lvert\varepsilon\left(f,\tfrac{1}{2}\right)\rvert=1, we have the following immediate corollary (using s=12,X=1s=\frac{1}{2},X=1),

Corollary 2.2.

Let L(f,s),q,λL(f,s),q,\lambda be as above. For any fixed ε>0,A>0\varepsilon>0,A>0,

(2.3) |L(f,12)|24|nq1+ελf(n)nV(nq)|2+O(qA).\left\lvert L\left(f,\tfrac{1}{2}\right)\right\rvert^{2}\leq 4\left\lvert\sum_{n\leq q^{1+\varepsilon}}\frac{\lambda_{f}(n)}{\sqrt{n}}V\left(\frac{n}{\sqrt{q}}\right)\right\rvert^{2}+O(q^{-A}).

2.2. Postnikov Formula

We will need to use a particular case of the Postnikov formula stated below.

Proposition 2.3.

Let pp be an odd prime, α\alpha be a primitive character modulo p3p^{3} and l>0l\in\mathbb{Z}_{>0}. Then for all nn with gcd(n,p)1,α(n+p2l)α(n)¯=ep(aαln¯),(n,p)\neq 1,\ \alpha(n+p^{2}l)\overline{\alpha(n)}=e_{p}(a_{\alpha}l\overline{n}), for some aα(/p)×,a_{\alpha}\in\left(\mathbb{Z}/{p\mathbb{Z}}\right)^{\times}, independent of nn.

Proof.

Define a function ψ\psi on \mathbb{Z} as ψ(n)=α(1+p2n)\psi(n)=\alpha(1+p^{2}n). We have

ψ(m+n)=α(1+p2(m+n))=α(1+p2(m+n)+p4mn)=α(1+p2n)α(1+p2m)=ψ(m)ψ(n).\psi(m+n)=\alpha(1+p^{2}(m+n))=\alpha(1+p^{2}(m+n)+p^{4}mn)=\alpha(1+p^{2}n)\cdot\alpha(1+p^{2}m)=\psi(m)\cdot\psi(n).

Hence, ψ\psi is an additive character modulo pp, and ψ(n)=ep(aαn)\psi(n)=e_{p}(a_{\alpha}n), for some aα(/p)×a_{\alpha}\in\left(\mathbb{Z}/{p\mathbb{Z}}\right)^{\times} (as α\alpha is primitive, aαa_{\alpha} cannot be 0).
Now, when gcd(n,p)=1(n,p)=1, we have α(n+p2l)α(n)¯=α(1+p2ln¯)=ψ(ln¯)=ep(aαln¯)\alpha(n+p^{2}l)\overline{\alpha(n)}=\alpha(1+p^{2}l\overline{n})=\psi(l\overline{n})=e_{p}(a_{\alpha}l\overline{n}). ∎

2.3. Delta Symbol

This section is a brief review of the delta-symbol method from Section 20.520.5 in [IK04].

Let w(u)w(u) be a smooth, compactly supported function on [C,2C][C,2C] for some C>0C>0. Additionally suppose that q=1w(q)=1\sum_{q=1}^{\infty}w(q)=1. Then we can rewrite the Kronecker delta function at zero as

(2.4) δ(n)=qn(w(q)w(|n|q)).\delta(n)=\sum_{q\mid n}\left(w(q)-w\left(\frac{\lvert n\rvert}{q}\right)\right).
Proposition 2.4.

Let δ\delta be as in (2.4). Using orthogonality of the additive characters modulo qq, we can rewrite (2.4) as

(2.5) δ(n)=c=1S(0,n;c)Δc(n).\delta(n)=\sum_{c=1}^{\infty}S(0,n;c)\Delta_{c}(n).

Here,

S(0,n;c)=d(modc)e(dnc),and Δc(u)=r=11cr(w(cr)w(|u|cr)).S(0,n;c)=\sideset{}{{}^{*}}{\sum}_{d\negthickspace\negthickspace\negthickspace\pmod{c}}e\left(\frac{dn}{c}\right),\ \text{and }\Delta_{c}(u)=\sum_{r=1}^{\infty}\frac{1}{cr}\left(w(cr)-w\left(\frac{\lvert u\rvert}{cr}\right)\right).

Since Δc(u)\Delta_{c}(u) is not compactly supported, we multiply it by a smooth function ff supported on [2N,2N][-2N,2N] (N>0N>0) such that f(0)f(0) = 1. We also assume, for any aa\in\mathbb{N}, f(a)(u)Naf^{(a)}(u)\ll N^{-a}.

Thus, we have

(2.6) δ(n)=c=1S(0,n;c)Δc(n)f(n).\delta(n)=\sum_{c=1}^{\infty}S(0,n;c)\Delta_{c}(n)f(n).

As ww is supported on [C,2C][C,2C] and ff is supported on [N,2N][N,2N], the sum on the right hand side is only upto c2 max (C,NC)=Xc\leq 2\text{ max }\left(C,\frac{N}{C}\right)=X. The optimal choice would thus be to take C=NC=\ \sqrt{N}, giving X=2C=2NX=2C=2\sqrt{N}.

Thus, (2.6) becomes

(2.7) δ(n)=c2CS(0,n;c)Δc(n)f(n).\delta(n)=\sum_{c\leq 2C}S(0,n;c)\Delta_{c}(n)f(n).

We can get additional bounds on the derivatives of Δc(u)\Delta_{c}(u) if we assume more conditions on ww.

Proposition 2.5.

Suppose w(u)w(u) is smooth, compactly supported in the segment Cu2CC\leq u\leq 2C with C1C\geq 1. Additionally, suppose that q=1w(q)=1\sum_{q=1}^{\infty}w(q)=1, and w(u)w(u) has derivatives satisfying

(2.8) w(a)(u)Ca1,w^{(a)}(u)\ll C^{-a-1},

for arbitrarily large aa\in\mathbb{N}. Then for any c1c\geq 1 and uu\in\mathbb{R} we have

(2.9) Δc(u)1(c+C)C+1(|u|+cC),\Delta_{c}(u)\ll\frac{1}{(c+C)C}+\frac{1}{\left(\lvert u\rvert+cC\right)},
(2.10) Δc(a)(n)(cC)1(|u|+cC)a.\Delta_{c}^{(a)}(n)\ll(cC)^{-1}(\lvert u\rvert+cC)^{-a}.

Finally, it is often useful to express Δc(n)f(n)\Delta_{c}(n)f(n) in terms of additive characters, by considering its Fourier transform.
Let

(2.11) gc(v)=Δc(u)f(u)e(uv)𝑑u,g_{c}(v)=\int_{-\infty}^{\infty}\Delta_{c}(u)f(u)e(-uv)du,

be the Fourier transform of Δc(u)f(u)\Delta_{c}(u)f(u). By the Fourier inversion formula we have

(2.12) Δc(u)f(u)=gc(v)e(uv)𝑑v.\Delta_{c}(u)f(u)=\int_{-\infty}^{\infty}g_{c}(v)e(uv)dv.

Using this in (2.7) , we have

(2.13) δ(n)=c2CS(0,n;c)gc(v)e(nv)𝑑v.\delta(n)=\sum_{c\leq 2C}S(0,n;c)\int_{-\infty}^{\infty}g_{c}(v)e(nv)dv.

we have the following bounds on gc(v)g_{c}(v):

Proposition 2.6.

Suppose gc(v)g_{c}(v) is defined as in (2.11), and C=NC=\sqrt{N}. Then

(2.14) gc(v)=1+O(1cC(cC+|v|cC)),g_{c}(v)=1+O\left(\frac{1}{cC}\left(\frac{c}{C}+\lvert v\rvert cC\right)\right),

and

(2.15) gc(v)(|v|cC)a.g_{c}(v)\ll(\lvert v\rvert cC)^{-a}.

2.4. Inert Functions and Stationary Phase

We mention some properties of inert functions in this section. Inert functions are special families of smooth functions characterised by certain derivative bounds. See Sections 22 and 33 in [KPY19], and Section 44 in [KrY21], for proofs.

Definition 2.7.

Let \mathcal{F} be an index set. A family {wT}T\{w_{T}\}_{T\in\mathcal{F}} of smooth function supported on a product of dyadic intervals in >0d\mathbb{R}_{>0}^{d} is called XX-inert if for each j=(j1,j2,,jd)>0dj=(j_{1},j_{2},\cdots,j_{d})\in\mathbb{Z}_{>0}^{d}, we have

C(j1,j2,,jd)supTsup(x1,x2,,xd)>0dXj1jd|x1j1xdjdwT(j1,,jd)(x1,,xd)|<.C(j_{1},j_{2},\cdots,j_{d})\coloneqq\sup_{T\in\mathcal{F}}\ \sup_{(x_{1},x_{2},\cdots,x_{d})\in\mathbb{R}_{>0}^{d}}X^{-j_{1}-\cdots-j_{d}}\left|x_{1}^{j_{1}}\cdots x_{d}^{j_{d}}{w_{T}}^{(j_{1},\cdots,j_{d})}(x_{1},\cdots,x_{d})\right|<\infty.

We will often denote the sequence of constants C(j1,j2,,jd)C(j_{1},j_{2},\dots,j_{d}) associated with this inert function as CC_{\mathcal{F}}.

We note that the requirements for the functions to be supported on dyadic intervals can be easily achieved by applying a dyadic partition of unity.

We give one simple example to highlight how such families can be constructed.

Example 2.8.

Let w(x1,,xd)w(x_{1},\cdots,x_{d}) be a fixed smooth function that is supported on [1,2]d[1,2]^{d}, and define,

(2.16) wX1,,Xd(x1,,xd)=w(x1X1,,xdXd).w_{X_{1},\cdots,X_{d}}(x_{1},\cdots,x_{d})=w\left(\frac{x_{1}}{X_{1}},\cdots,\frac{x_{d}}{X_{d}}\right).

Then with ={T=(X1,,Xd)>0d}\mathcal{F}=\{T=(X_{1},\cdots,X_{d})\in\mathbb{R}_{>0}^{d}\}, the family {wT}T\{w_{T}\}_{T\in\mathcal{F}} is 11-inert.

The following propositions can be checked immediately:

Proposition 2.9.

Let a,ba,b\in\mathbb{R} with b>0b>0. If wT(x)w_{T}(x) is a family of XX-inert functions supported on x[N,2n]x\in[N,2n], then the family WT(x)W_{T}(x) given by, WT(x)=wT(bxa)W_{T}(x)=w_{T}(bx^{a}) is also XX-inert, with support [(Nb)1a,2(Nb)1a]\left[\left(\frac{N}{b}\right)^{\frac{1}{a}},2\left(\frac{N}{b}\right)^{\frac{1}{a}}\right].

Proposition 2.10.

Let aa\in\mathbb{R}. If wT(x)w_{T}(x) is a family of XX-inert functions supported on xNx\asymp N, then so is the family WT(x)W_{T}(x) given by, WT(x)=(xN)awT(x)W_{T}(x)=(\frac{x}{N})^{-a}w_{T}(x).

Proposition 2.11.

If ww is an XX-inert function and vv is a YY-inert function, then their product wvw\cdot v is a max(X,Y)\max{(X,Y)}-inert function.

Suppose that wT(x1,,xd)w_{T}(x_{1},\cdots,x_{d}) is XX-inert and is supported on xiXix_{i}\asymp X_{i}. Let

(2.17) w^T(t1,x2,,xd)=wT(x1,,xd)e(x1t1)𝑑x1,\widehat{w}_{T}(t_{1},x_{2},\cdots,x_{d})=\int_{-\infty}^{\infty}w_{T}(x_{1},\cdots,x_{d})e(-x_{1}t_{1})dx_{1},

denote its Fourier transform in the x1x_{1}-variable.

We state the following result (Prop 2.6 in [KPY19]) regarding the Fourier transform of inert functions.

Proposition 2.12.

Suppose that {wT:T}\{w_{T}:T\in\mathcal{F}\} is a family of XX-inert functions such that x1x_{1} is supported on x1X1x_{1}\asymp X_{1}, and {w±Y1:Y1(0,)}\{w_{\pm Y_{1}}:Y_{1}\in(0,\infty)\} is a 11-inert family of functions with support on ±t1Y1\pm t_{1}\asymp Y_{1}. Then the family {X11w±Y1(t1)w^T(t1,x2,,xd):(T,±Y1)×±(0,)}\{X_{1}^{-1}w_{\pm Y_{1}}(t_{1})\widehat{w}_{T}(t_{1},x_{2},\cdots,x_{d})\ :\ (T,\pm Y_{1})\in\mathcal{F}\times\pm(0,\infty)\} is XX-inert. Furthermore if Y1pεXX1Y_{1}\gg p^{\varepsilon}\frac{X}{X_{1}}, then for any A>0A>0, we have

(2.18) X11w±Y1(t1)w^T(t1,x2,,xd)ε,ApA.X_{1}^{-1}w_{\pm Y_{1}}(t_{1})\widehat{w}_{T}(t_{1},x_{2},\cdots,x_{d})\ll_{\varepsilon,A}p^{-A}.

We can similarly describe the Mellin transform of such functions as well. We state the following result (Lemma 4.2 in [KrY21]).

Proposition 2.13.

Suppose that {wT(x1,x2,,xd):T}\{w_{T}(x_{1},x_{2},\cdots,x_{d}):T\in\mathcal{F}\} is a family of XX-inert functions such that x1x_{1} is supported on x1X1x_{1}\asymp X_{1}. Let

(2.19) w~T(s,x2,,xd)=0wT(x,x2,,xd)xsdxx.\widetilde{w}_{T}(s,x_{2},\cdots,x_{d})=\int_{0}^{\infty}w_{T}(x,x_{2},\cdots,x_{d})x^{s}\frac{dx}{x}.

Then we have w~T(s,x2,,xd)=X1sWT(s,x2,,xd)\widetilde{w}_{T}(s,x_{2},\cdots,x_{d})={X_{1}}^{s}W_{T}(s,x_{2},\cdots,x_{d}) where WT(s,)W_{T}(s,\cdots) is a family of XX-inert functions in x2,,xdx_{2},\cdots,x_{d}, which is entire in ss, and has rapid decay for |(s)|X11+ε\lvert\Im(s)\rvert\gg{X_{1}}^{1+\varepsilon}.

The following is a restatement of Lemma 3.1 in [KPY19].

Proposition 2.14.

Suppose that ww is an XX-inert function, with compact support on [Z,2Z][Z,2Z], so that w(j)(t)(Z/X)jw^{(j)}(t)\ll(Z/X)^{-j}. And suppose that ϕ\phi is smooth and satisfies ϕ(j)(t)YZj\phi^{(j)}(t)\ll\frac{Y}{Z^{j}} for some Y/X2R1Y/X^{2}\geq R\geq 1, and for all t[Z,2Z]t\in[Z,2Z]. Let

(2.20) I=w(t)eiϕ(t)𝑑t.I=\int_{-\infty}^{\infty}w(t)e^{i\phi(t)}dt.

We then have

  1. (a)

    If |ϕ(t)|YZ\lvert\phi^{\prime}(t)\rvert\geq\frac{Y}{Z} for all t[Z,2Z]t\in[Z,2Z], then IZRAI\ll ZR^{-A} for AA arbitrary large.

  2. (b)

    If ϕ′′(t)YZ2\phi^{\prime\prime}(t)\gg\frac{Y}{Z^{2}} for all t[Z,2Z]t\in[Z,2Z], and there exists a (necessarily unique) t0t_{0}\in\mathbb{R} such that ϕ(t0)=0\phi^{\prime}(t_{0})=0, then

    (2.21) I=eiϕ(t0)ϕ′′(t0)F(t0)+OA(ZRA),I=\frac{e^{i\phi(t_{0})}}{\sqrt{\phi^{\prime\prime}(t_{0})}}F(t_{0})+O_{A}(ZR^{-A}),

    where FF is an XX-inert function supported on t0Zt_{0}\asymp Z.

The previous result has a natural generalisation (Main Theorem in [KPY19]).

Proposition 2.15.

Suppose that ww is an XX-inert function in t1,,tdt_{1},\cdots,t_{d}, supported on t1Zt_{1}\asymp Z and tiXit_{i}\asymp X_{i} for i=2,3,,di=2,3,\dots,d. Suppose that the smooth function ϕ\phi satisfies

(2.22) a1+a2++adt1a1t2a2tdadϕ(t1,t2,,td)CYZa11X2a2Xdad,\frac{\partial^{a_{1}+a_{2}+\cdots+a_{d}}}{\partial t_{1}^{a_{1}}\partial t_{2}^{a_{2}}\dots\partial t_{d}^{a_{d}}}\phi(t_{1},t_{2},\dots,t_{d})\ll_{C}\frac{Y}{Z^{a_{1}}}\frac{1}{X_{2}^{a_{2}}\dots X_{d}^{a_{d}}},

for all a1,a2,,ada_{1},a_{2},\dots,a_{d}\in\mathbb{N}.
Now, suppose that ϕ(t1,t2,,td)YZ2\phi^{\prime}(t_{1},t_{2},\dots,t_{d})\gg\frac{Y}{Z^{2}} (here ϕ′′\phi^{\prime\prime} and ϕ\phi^{\prime} denote the derivatives of ϕ\phi with respect to t1t_{1}), for all t1,t2,,tdt_{1},t_{2},\dots,t_{d} in the support of ww, and there exists a (necessarily unique) t0t_{0}\in\mathbb{R} such that ϕ(t0)=0\phi^{\prime}(t_{0})=0. Suppose also that YX2R1\frac{Y}{X^{2}}\geq R\geq 1, then

(2.23) I=eiϕ(t1,,td)w(t1,,td)dt1=ZYeiϕ(t0,t2,,td)W(t2,,td)+OA(ZRA),I=\int_{\mathbb{R}}e^{i\phi(t_{1},\dots,t_{d})w(t_{1},\dots,t_{d})dt_{1}}=\frac{Z}{\sqrt{Y}}e^{i\phi(t_{0},t_{2},\dots,t_{d})}W(t_{2},\dots,t_{d})+O_{A}(ZR^{-A}),

for some XXinert function WW, and AA can be taken to be arbitrarily large. The implied constant depends on AA and CC.

2.5. Voronoi Summation Formula

We will need the following two Voronoi-type summation formulas.

Proposition 2.16.

Let ff be a cusp form of weight k1k\geq 1, level 11 and trivial central character. Suppose c1c\geq 1, and ad1(modc)ad\equiv 1(\mathrm{mod}\ c). Then for any smooth, compactly supported function gg on +\mathbb{R}^{+}, we have

(2.24) m=1λf(m)e(amc)g(m)=1cn=1λf(n)e(dnc)H(n),\sum_{m=1}^{\infty}\lambda_{f}(m)e\left(\frac{am}{c}\right)g(m)=\frac{1}{c}\sum_{n=1}^{\infty}\lambda_{f}(n)e\left(\frac{-dn}{c}\right)H(n),

where

(2.25) H(n)=2πik0g(x)Jk1(4πcxn)𝑑x.H(n)=2\pi i^{k}\int_{0}^{\infty}g(x)J_{k-1}\left(\frac{4\pi}{c}\sqrt{xn}\right)dx.

We will also use a modified Voronoi summation formula, which we state and prove below.

Proposition 2.17.

Let f,g,a,c,f,g,a,c, and dd be as in Proposition 2.16. Additionally, assume that ff is a Hecke eigenform. Let pp be a prime with gcd(c,p)=1(c,p)=1 and fix a constant r0(modp)r\not\equiv 0(\mathrm{mod}\ p). We then have

(2.26) gcd(m,p)=1λf¯(m)e(rm¯p)e(amc)g(m)=gcd(n,p)=1λf¯(n)H1(n)p2ce(ap2¯nc)Kl3(nc¯2r,1,1;p)+gcd(n,p2)=pλf¯(n)H1(n)p2ce(ap2¯nc)+n=1λf¯(np)λf¯(p)H1(p2n)p2ce(a¯nc)(1+1p)n=1λf¯(n)H1(p2n)pce(a¯nc).\sum_{\text{gcd}(m,p)=1}\overline{\lambda_{f}}(m)e\left(\frac{r\overline{m}}{p}\right)e\left(-\frac{am}{c}\right)g(m)=\\ \sum_{\text{gcd}(n,p)=1}\frac{\overline{\lambda_{f}}(n)H_{1}(n)}{p^{2}c}e\left(\frac{\overline{ap^{2}}n}{c}\right)\text{Kl}_{3}(-n\overline{c}^{2}r,1,1;p)+\sum_{\text{gcd}(n,p^{2})=p}\frac{\overline{\lambda_{f}}(n)H_{1}(n)}{p^{2}c}e\left(\frac{\overline{ap^{2}}n}{c}\right)\\ +\sum_{n=1}^{\infty}\frac{\overline{\lambda_{f}}(np)\overline{\lambda_{f}}(p)H_{1}(p^{2}n)}{p^{2}c}e\left(\frac{\overline{a}n}{c}\right)-\left(1+\frac{1}{p}\right)\sum_{n=1}^{\infty}\frac{\overline{\lambda_{f}}(n)H_{1}(p^{2}n)}{pc}e\left(\frac{\overline{a}n}{c}\right).

Here,

(2.27) Kl3(x,y,x;p)=abc1(modp)e(ax+by+czp),\text{Kl}_{3}(x,y,x;p)=\sum_{abc\equiv 1(\mathrm{mod}\ p)}e\left(\frac{ax+by+cz}{p}\right),

and

(2.28) H1(y)=2πikoJk1(4πxypc)g(x)𝑑x.H_{1}(y)=2\pi i^{k}\int_{o}^{\infty}J_{k-1}\left(\frac{4\pi\sqrt{xy}}{pc}\right)g(x)dx.
Proof.

Let

(2.29) S=gcd(m,p)=1λf¯(m)e(rm¯p)e(amc)g(m).S=\sum_{\text{gcd}(m,p)=1}\overline{\lambda_{f}}(m)e\left(\frac{r\overline{m}}{p}\right)e\left(-\frac{am}{c}\right)g(m).

From the orthogonality of characters modulo pp, we know that when gcd(m,p)=1(m,p)=1,

(2.30) e(rm¯p)=1pt(modp)h(modp)e(rh¯ht+mtp).e\left(\frac{r\overline{m}}{p}\right)=\frac{1}{p}\sum_{t\negthickspace\negthickspace\negthickspace\pmod{p}}\ \sideset{}{{}^{*}}{\sum}_{h\negthickspace\negthickspace\negthickspace\pmod{p}}e\left(\frac{r\overline{h}-ht+mt}{p}\right).

We note that if gcd(m,p)1(m,p)\neq 1, then the right hand side of (2.30) is zero. We thus have

(2.31) S=1pt(modp)h(modp)e(rh¯htp)m=1λf¯(m)g(m)e(amc+tmp).S=\frac{1}{p}\sum_{t\negthickspace\negthickspace\negthickspace\pmod{p}}\ \sideset{}{{}^{*}}{\sum}_{h\negthickspace\negthickspace\negthickspace\pmod{p}}e\left(\frac{r\overline{h}-ht}{p}\right)\sum_{m=1}^{\infty}\overline{\lambda_{f}}(m)g(m)e\left(\frac{-am}{c}+\frac{tm}{p}\right).

Separating the t0(modp)t\equiv 0(\mathrm{mod}\ p) term, (2.31) can be rewritten as

(2.32) S=S1+S2,S=S_{1}+S_{2},

where

(2.33) S1=1ph,t(modp)e(rh¯htp)m=1g(m)λf¯(m)e(m(ap+tc)pc),S_{1}=\frac{1}{p}\ \sideset{}{{}^{*}}{\sum}_{h,t(\mathrm{mod}\ p)}e\left(\frac{r\overline{h}-ht}{p}\right)\sum_{m=1}^{\infty}g(m)\overline{\lambda_{f}}(m)e\left(\frac{m(-ap+tc)}{pc}\right),

and

(2.34) S2=1ph(modp)e(rh¯p)m=1g(m)λf¯(m)e(amc).S_{2}=\frac{1}{p}\ \sideset{}{{}^{*}}{\sum}_{h(\mathrm{mod}\ p)}e\left(\frac{r\overline{h}}{p}\right)\sum_{m=1}^{\infty}g(m)\overline{\lambda_{f}}(m)e\left(\frac{-am}{c}\right).

We can now use Proposition 2.16 to simplify S1S_{1} and S2S_{2}. Note that gcd(a,c)=1(a,c)=1 and gcd(ap+tc,pc)=1(-ap+tc,pc)=1 if gcd(t,p)=1(t,p)=1. Thus (2.33) becomes

(2.35) S1=1ph,t(modp)e(rh¯htp)(1pcn=1λf¯(n)e(tcap¯pcn)H1(n)),S_{1}=\frac{1}{p}\ \sideset{}{{}^{*}}{\sum}_{h,t(\mathrm{mod}\ p)}e\left(\frac{r\overline{h}-ht}{p}\right)\left(\frac{1}{pc}\sum_{n=1}^{\infty}\overline{\lambda_{f}}(n)e\left(-\frac{\overline{tc-ap}}{pc}\cdot n\right)H_{1}(n)\right),

where H1(y)H_{1}(y) is defined in (2.28).

We now use the Chinese Remainder Theorem to simplify (2.35) further.

(2.36) S1=1p2cn=1λf¯(n)H1(n)(h,t(modp)e(rh¯htp)e(p¯(tcap¯)nc)e(c¯(tcap¯)np))=1p2cn=1λf¯(n)H1(n)e(ap2¯nc)(h,t(modp)e(rh¯httc2¯np))=n=1λf¯(n)H1(n)p2ce(ap2¯nc)Kl3(nc¯2r,1,1;p),S_{1}=\frac{1}{p^{2}c}\sum_{n=1}^{\infty}\overline{\lambda_{f}}(n)H_{1}(n)\left(\ \sideset{}{{}^{*}}{\sum}_{h,t(\mathrm{mod}\ p)}e\left(\frac{r\overline{h}-ht}{p}\right)e\left(\frac{-\overline{p}(\overline{tc-ap})n}{c}\right)e\left(\frac{-\overline{c}(\overline{tc-ap})n}{p}\right)\right)\\ =\frac{1}{p^{2}c}\sum_{n=1}^{\infty}\overline{\lambda_{f}}(n)H_{1}(n)e\left(\frac{\overline{ap^{2}}n}{c}\right)\left(\ \sideset{}{{}^{*}}{\sum}_{h,t(\mathrm{mod}\ p)}e\left(\frac{r\overline{h}-ht-\overline{tc^{2}}n}{p}\right)\right)\\ =\sum_{n=1}^{\infty}\frac{\overline{\lambda_{f}}(n)H_{1}(n)}{p^{2}c}e\left(\frac{\overline{ap^{2}}n}{c}\right)\text{Kl}_{3}(-n\overline{c}^{2}r,1,1;p),

where Kl3()\text{Kl}_{3}(\cdot) is defined in (2.27).
Note that if pnp\mid n, then this hyper-Kloosterman sum can be simplified as

(2.37) Kl3(nc¯2r,1,1;p)=xyz1(modp)e(x(nc¯2r)+y+zp)=y,z(modp)e(y+zp)=1.\text{Kl}_{3}(-n\overline{c}^{2}r,1,1;p)=\sum_{xyz\equiv 1(\mathrm{mod}\ p)}e\left(\frac{x(-n\overline{c}^{2}r)+y+z}{p}\right)=\ \sideset{}{{}^{*}}{\sum}_{y,z(\mathrm{mod}\ p)}e\left(\frac{y+z}{p}\right)=1.

This allows us to rewrite (2.36) as,

(2.38) S1=S1+S1p+S1p2.S_{1}=S_{1}^{*}+S_{1}^{p}+S_{1}^{p^{2}}.

Here,

(2.39) S1=gcd(n,p)=1λf¯(n)H1(n)p2ce(ap2¯nc)Kl3(nc¯2r,1,1;p),S_{1}^{*}=\sum_{\text{gcd}(n,p)=1}\frac{\overline{\lambda_{f}}(n)H_{1}(n)}{p^{2}c}e\left(\frac{\overline{ap^{2}}n}{c}\right)\text{Kl}_{3}(-n\overline{c}^{2}r,1,1;p),
(2.40) S1p=gcd(n,p2)=pλf¯(n)H1(n)p2ce(ap2¯nc)Kl3(nc¯2r,1,1;p)=gcd(n,p2)=pλf¯(n)H1(n)p2ce(ap2¯nc),S_{1}^{p}=\sum_{\text{gcd}(n,p^{2})=p}\frac{\overline{\lambda_{f}}(n)H_{1}(n)}{p^{2}c}e\left(\frac{\overline{ap^{2}}n}{c}\right)\text{Kl}_{3}(-n\overline{c}^{2}r,1,1;p)=\sum_{\text{gcd}(n,p^{2})=p}\frac{\overline{\lambda_{f}}(n)H_{1}(n)}{p^{2}c}e\left(\frac{\overline{ap^{2}}n}{c}\right),

and

(2.41) S1p2=gcd(n,p2)=p2λf¯(n)H1(n)p2ce(ap2¯nc)Kl3(nc¯2r,1,1;p)=gcd(n,p2)=p2λf¯(n)H1(n)p2ce(ap2¯nc).S_{1}^{p^{2}}=\sum_{\text{gcd}(n,p^{2})=p^{2}}\frac{\overline{\lambda_{f}}(n)H_{1}(n)}{p^{2}c}e\left(\frac{\overline{ap^{2}}n}{c}\right)\text{Kl}_{3}(-n\overline{c}^{2}r,1,1;p)=\sum_{\text{gcd}(n,p^{2})=p^{2}}\frac{\overline{\lambda_{f}}(n)H_{1}(n)}{p^{2}c}e\left(\frac{\overline{ap^{2}}n}{c}\right).

Using Proposition 2.16 again on (2.34) we get

(2.42) S2=1ph(modp)e(rh¯p)1cn=1λf¯(n)e(a¯nc)H(n)=n=1λf¯(n)H(n)pce(a¯nc)S(r,0;p),S_{2}=\frac{1}{p}\ \sideset{}{{}^{*}}{\sum}_{h(\mathrm{mod}\ p)}e\left(\frac{r\overline{h}}{p}\right)\frac{1}{c}\sum_{n=1}^{\infty}\overline{\lambda_{f}}(n)e\left(\frac{\overline{a}n}{c}\right)H(n)=\sum_{n=1}^{\infty}\frac{\overline{\lambda_{f}}(n)H(n)}{pc}e\left(\frac{\overline{a}n}{c}\right)S(r,0;p),

where S(r,0;p)S(r,0;p) is the Ramanujan sum modulo p.H(y)p.\ H(y) is the same as in (2.25).
Note that S(r,0;p)=1S(r,0;p)=-1 as r0(modp)r\not\equiv 0(\mathrm{mod}\ p). Also, H(y)=H1(p2y)H(y)=H_{1}(p^{2}y).
Finally, as ff is a Hecke eigenform, λf(p2n)=λf(p)λf(np)λf(n)\lambda_{f}(p^{2}n)=\lambda_{f}(p)\lambda_{f}(np)-\lambda_{f}(n). Now, (2.32), (2.38) and (2.42), together imply (2.26). ∎

2.6. Bruggeman-Kuznetsov Formula

We state the version of Bruggeman-Kuznetsov formula that we will use. This is a restatement of Theorem 6.106.10 in [PY23].

Let χ\chi be an even Dirichlet character modulo qq and ΦCC(>0)\Phi\in C_{C}^{\infty}(\mathbb{R}_{>0}),

(2.43) V(q)=Vol (Γ0(q)\)=π3qpq(1+1p),V(q)=\text{Vol }(\Gamma_{0}(q)\backslash\mathcal{H})=\frac{\pi}{3}q\prod_{p\mid q}\left(1+\frac{1}{p}\right),

and

(2.44) 𝒦=(c,q)=1χ¯(c)S(q¯m,n,c)Φ(q12c),\mathcal{K}=\sum_{(c,q)=1}\overline{\chi}(c)S(\overline{q}m,n,c)\Phi(q^{\frac{1}{2}}c),

where S(a,b;c)S(a,b;c) denotes the usual Kloosterman sum.

Additionally, for fixed integers mm and nn, we define the following integral transforms -

(2.45) holΦ(k)=0Jk1(4π|mn|x)Φ(x)𝑑x.\mathcal{L}^{\text{hol}}\Phi(k)=\int_{0}^{\infty}J_{k-1}\left(\frac{4\pi\sqrt{\lvert mn\rvert}}{x}\right)\Phi(x)dx.

Also,

(2.46) ±Φ(t)=0B2it±(4π|mn|x)Φ(x)𝑑x,\mathcal{L}^{\pm}\Phi(t)=\int_{0}^{\infty}B_{2it}^{\pm}\left(\frac{4\pi\sqrt{\lvert mn\rvert}}{x}\right)\Phi(x)dx,

where

(2.47) B2it+(x)=i2sin(πt)(J2it(x)J2it(x)),B_{2it}^{+}(x)=\frac{i}{2\sin{(\pi t)}}\left(J_{2it}(x)-J_{-2it}(x)\right),
(2.48) B2it(x)=2πcosh(πt)K2it(x).B_{2it}^{-}(x)=\frac{2}{\pi}\cosh{(\pi t)}K_{2it}(x).

Here, JJ and KK denote the usual JJ-Bessel and KK-Bessel functions.
We could also use Mellin inversion formula to rewrite these integral transforms as-

(2.49) holΦ(k)=12πi(1)2s1Γ(s+k12)Γ(k+1s2)Φ~(s+1)(4π|mn|)s𝑑s,\mathcal{L}^{\text{hol}}\Phi(k)=\frac{1}{2\pi i}\int_{(1)}\frac{2^{s-1}\Gamma\left(\frac{s+k-1}{2}\right)}{\Gamma\left(\frac{k+1-s}{2}\right)}\tilde{\Phi}(s+1)(4\pi\sqrt{\lvert mn\rvert})^{-s}ds,
(2.50) ±Φ(t)=12πi(2)h±(s,t)Φ~(s+1)(4π|mn|)s𝑑s,\mathcal{L}^{\pm}\Phi(t)=\frac{1}{2\pi i}\int_{(2)}h_{\pm}(s,t)\tilde{\Phi}(s+1)(4\pi\sqrt{\lvert mn\rvert})^{-s}ds,

where

(2.51) h±(s,t)=2s1πΓ(s2+it)Γ(s2it){cos(πs/2),±=+cosh(πt),±=h_{\pm}(s,t)=\frac{2^{s-1}}{\pi}\Gamma(\frac{s}{2}+it)\Gamma(\frac{s}{2}-it)\begin{cases}\cos(\pi s/2),&\pm=+\\ \cosh(\pi t),&\pm=-\end{cases}

and Φ~()\tilde{\Phi}(\cdot) denotes the Mellin transform of Φ()\Phi(\cdot).

For any Hecke eigenform π\pi define,

(2.52) λπ(δ)(n)=dδd12xδ(d)λπ(n/d),\lambda_{\pi}^{(\delta)}(n)=\sum_{d\mid\delta}d^{\frac{1}{2}}x_{\delta}(d)\lambda_{\pi}(n/d),

where λπ(n/d)=0\lambda_{\pi}(n/d)=0 if nd\frac{n}{d} is not an integer. xδsx_{\delta}^{\prime}s are defined in the same way as equation 6.216.21 in [PY23].

Also, let it(m,ψ)\mathcal{H}_{it}(m,\psi) denote the set of Hecke-Maass newforms of conductor mm, central character ψ\psi, and spectral parameter itit. Similarly, we can define k(m,ψ)\mathcal{H}_{k}(m,\psi) as the set of Hecke newforms of conductor mm, central character ψ\psi and weight kk. We can also define it,Eis(m,ψ)\mathcal{H}_{it,\text{Eis}}(m,\psi) as the set of newform Eisenstein series of level mm and character χ\chi.

We can now state a version of the Bruggeman-Kuznetsov formula -

Proposition 2.18.

Let ΦCC(>0)\Phi\in C_{C}^{\infty}(\mathbb{R}_{>0}). We have

(2.53) gcd(c,q)=1χ¯(c)S(q¯m,n,c)Φ(q12c)=𝒦Maass+𝒦Eis+𝒦hol.\sum_{\text{gcd}(c,q)=1}\overline{\chi}(c)S(\overline{q}m,n,c)\Phi(q^{\frac{1}{2}}c)=\mathcal{K}_{\text{Maass}}+\mathcal{K}_{\text{Eis}}+\mathcal{K}_{\text{hol}}\ .

Here,

(2.54) 𝒦Maass=tj±Φ(tj)lr=qπitj(r,χ)4πϵπV(q)π(1)δlλ¯π(δ)(|m|)λ¯π(δ)(|n|),\mathcal{K}_{\text{Maass}}=\sum_{t_{j}}\mathcal{L}^{\pm}\Phi(t_{j})\sum_{lr=q}\sum_{\pi\in\mathcal{H}_{it_{j}}(r,\chi)}\frac{4\pi\epsilon_{\pi}}{V(q)\mathscr{L}_{\pi}^{*}(1)}\sum_{\delta\mid l}\overline{\lambda}_{\pi}^{(\delta)}(\lvert m\rvert)\overline{\lambda}_{\pi}^{(\delta)}(\lvert n\rvert),

and

(2.55) 𝒦Eis=14π±Φ(tj)lr=qπit,Eis(r,χ)4πϵπV(q)π(1)δlλ¯π(δ)(|m|)λ¯π(δ)(|n|)dt,\mathcal{K}_{\text{Eis}}=\frac{1}{4\pi}\int_{-\infty}^{\infty}\mathcal{L}^{\pm}\Phi(t_{j})\sum_{lr=q}\sum_{\pi\in\mathcal{H}_{it,\textrm{Eis}}(r,\chi)}\frac{4\pi\epsilon_{\pi}}{V(q)\mathscr{L}_{\pi}^{*}(1)}\sum_{\delta\mid l}\overline{\lambda}_{\pi}^{(\delta)}(\lvert m\rvert)\overline{\lambda}_{\pi}^{(\delta)}(\lvert n\rvert)dt,

where one takes Φ+\Phi+ (resp. Φ\Phi^{-}) if mn>0mn>0 (resp. mn<0mn<0), and ϵπ\epsilon_{\pi} is the finite root number of π\pi. Also,

(2.56) 𝒦hol=k>0, evenholΦ(k)lr=qπk(r,χ)4πϵπV(q)π(1)δlλ¯π(δ)(|m|)λ¯π(δ)(|n|),\mathcal{K}_{\text{hol}}=\sum_{k>0,\text{ even}}\mathcal{L}^{\text{hol}}\Phi(k)\sum_{lr=q}\sum_{\pi\in\mathcal{H}_{k}(r,\chi)}\frac{4\pi\epsilon_{\pi}}{V(q)\mathscr{L}_{\pi}^{*}(1)}\sum_{\delta\mid l}\overline{\lambda}_{\pi}^{(\delta)}(\lvert m\rvert)\overline{\lambda}_{\pi}^{(\delta)}(\lvert n\rvert),

if mn>0mn>0, and 𝒦hol=0\mathcal{K}_{\text{hol}}=0 if mn<0mn<0.

2.7. Spectral Large Sieve Inequality

We state the version of spectral large sieve inequality that we will use. This is a restatement of Lemma 7.47.4 in [PY23].

Let us denote by

T any of |tj|T,kT, or |t|T𝑑t,\int_{*\leq T}\ \text{ any of }\sum_{\lvert t_{j}\rvert\leq T},\sum_{k\leq T},\text{ or }\int_{\lvert t\rvert\leq T}dt,

according to whether =itj,k,*=it_{j},k, or it,Eisit,\text{Eis}.

Proposition 2.19.

For any sequence of complex numbers ana_{n}, we have

(2.57) Tπ(q)|nNanλπ(n)|2ε(T2q+N)(qTN)εnN|an|2.\int_{*\leq T}\sum_{\pi\in\mathcal{H}_{*}(q)}\left|\sum_{n\leq N}a_{n}\lambda_{\pi}(n)\right|^{2}\ll_{\varepsilon}(T^{2}q+N)(qTN)^{\varepsilon}\sum_{n\leq N}\lvert a_{n}\rvert^{2}.

2.8. Fourth Moment of Fourier Coefficients of Cusp Forms

We will use the following result on bounding the fourth moment of the Fourier coefficients of cusp forms.

Proposition 2.20.

Let λf(n)\lambda_{f}(n) denote the nthn^{\text{th}} normalised Fourier coefficient of a primitive holomorphic or Maass cusp form ff for SL2()SL_{2}(\mathbb{Z}). Let xx\in\mathbb{R} be positive. Then, for any ε>0\varepsilon>0,

(2.58) nx|λf(n)|4x1+ε.\sum_{n\leq x}\lvert\lambda_{f}(n)\rvert^{4}\ll x^{1+\varepsilon}.

Proposition 2.20 follows from Theorem 1.5 and Remark 1.7 in [LL11]. In fact, Moreno and Shahidi first obtained a similar bound for the fourth moment of the Ramanujan τ\tau-function in [MS83]. They were able to prove nxτ(n)4cxlogx\sum_{n\leq x}\tau(n)^{4}\sim cx\log{x}, for some positive constant cc.

This result was then extended to primitive holomorphic cusp forms by Lü (see [L0̈9]) and to primitive Maass cusp forms by Lau and Lü (see [LL11]).

3. Reduction of Theorem 1.4 to Theorem 1.7

In this section, we will prove Theorem 1.4, assuming Theorem 1.7 is true.

Using Corollary 2.2 and a dyadic partition of unity, it suffices to show (via an application of Cauchy-Schwarz inequality) that for any ε>0\varepsilon>0,

(3.1) S(N,α)ψ(modp2)|nNλf(n)ψ(n)α(n)wN(n)|2Np2+ε,S(N,\alpha)\coloneqq\sum_{\psi(\mathrm{mod}\ p^{2})}\left\lvert\sideset{}{{}^{*}}{\sum}_{n\asymp N}\lambda_{f}(n)\psi(n)\alpha(n)w_{N}(n)\right\rvert^{2}\ll Np^{2+\varepsilon},

where wNw_{N} is a smooth function supported on [N,2N][N,2N] satisfying wN(j)(x)Njw_{N}^{(j)}(x)\ll N^{-j} and Np3+εN\ll p^{3+\varepsilon}. Recall that n\sideset{}{{}^{*}}{\sum}_{n} denotes the sum is over values of nn relatively prime to pp.

Expanding the square in S(N,α)S(N,\alpha) and rearranging terms, we get

(3.2) S(N,α)=m,nλf(m)λf¯(n)α(m)α(n)¯wN(m)wN(n)ψ(modp2)ψ(mn¯).S(N,\alpha)=\sideset{}{{}^{*}}{\sum}_{m,n}\lambda_{f}(m)\overline{\lambda_{f}}(n)\alpha(m)\overline{\alpha(n)}w_{N}(m)w_{N}(n)\sum_{\psi(\mathrm{mod}\ p^{2})}\psi(m\overline{n}).

The inner sum in the right hand side of (3.2) vanishes unless mn(modp2)m\equiv n(\mathrm{mod}\ p^{2}), in which case it is equal to ϕ(p2)\phi(p^{2}). We can then rewrite S(N,α)S(N,\alpha) as a sum of diagonal and off-diagonal terms. We get

(3.3) S(N,α)=nϕ(p2)|λf(n)|2wn(N)2+S0(N,α),S(N,\alpha)=\sideset{}{{}^{*}}{\sum}_{n}\phi(p^{2})\left\lvert\lambda_{f}(n)\right\rvert^{2}w_{n}(N)^{2}+S_{0}(N,\alpha),

where

(3.4) S0(N,α)=mnmn(modp2)ϕ(p2)λf(m)λf¯(n)α(m)α(n)¯wN(m)wN(n),S_{0}(N,\alpha)=\sideset{}{{}^{*}}{\sum}_{\begin{subarray}{c}m\neq n\\ m\equiv n(\mathrm{mod}\ p^{2})\end{subarray}}\phi(p^{2})\lambda_{f}(m)\overline{\lambda_{f}}(n)\alpha(m)\overline{\alpha(n)}w_{N}(m)w_{N}(n),

is the off-diagonal term.

Using known bounds on wNw_{N} and λf(n)\lambda_{f}(n) the diagonal term is of the order O(Np2+ε)O(Np^{2+\varepsilon}). Thus (3.1) follows if we can show that off-diagonal terms (S0(N,α)S_{0}(N,\alpha)) satisfy the same bound.

Now, for S0(N,α)S_{0}(N,\alpha), we note that it suffices to consider the sum for the terms with m>nm>n, as the sum for the terms with m<nm<n is just the complex conjugate of the m>nm>n sum.

Letting l=mnp2l=\frac{m-n}{p^{2}}, and considering only positive values of ll, we get

(3.5) S0(N,α)ϕ(p2)lnλf(n+p2l)λf¯(n)α(n+p2l)α(n)¯wN(n+p2l)wN(n).S_{0}(N,\alpha)\ll\phi(p^{2})\sum_{l}\ \sideset{}{{}^{*}}{\sum}_{n}\lambda_{f}(n+p^{2}l)\overline{\lambda_{f}}(n)\alpha(n+p^{2}l)\overline{\alpha(n)}w_{N}(n+p^{2}l)w_{N}(n).

We define

(3.6) S1(N,α)n,lλf(n+p2l)λf¯(n)α(n+p2l)α(n)¯wN(n+p2l)wN(n),S_{1}(N,\alpha)\coloneqq\sideset{}{{}^{*}}{\sum}_{n,l}\lambda_{f}(n+p^{2}l)\overline{\lambda_{f}}(n)\alpha(n+p^{2}l)\overline{\alpha(n)}w_{N}(n+p^{2}l)w_{N}(n),

and

(3.7) S2(N,α)l0(modp)nλf(n+p2l)λf¯(n)α(n+p2l)α(n)¯wN(n+p2l)wN(n).S_{2}(N,\alpha)\coloneqq\sum_{l\equiv 0\negthickspace\negthickspace\negthickspace\pmod{p}}\ \sideset{}{{}^{*}}{\sum}_{n}\lambda_{f}(n+p^{2}l)\overline{\lambda_{f}}(n)\alpha(n+p^{2}l)\overline{\alpha(n)}w_{N}(n+p^{2}l)w_{N}(n).

Note that S1(N,α)S_{1}(N,\alpha) is precisely the left hand side in (1.7) in Theorem 1.7.

Then (3.5) can be written as

(3.8) S0(N,α)ϕ(p2)(S1(N,α)+S2(N,α)).S_{0}(N,\alpha)\ll\phi(p^{2})(S_{1}(N,\alpha)+S_{2}(N,\alpha)).

So, in order to prove Theorem 1.4, it suffices to show

(3.9) S1(N,α)Np2+εϕ(p2)Npε,S_{1}(N,\alpha)\ll\frac{Np^{2+\varepsilon}}{\phi(p^{2})}\ll Np^{\varepsilon},

and

(3.10) S2(N,α)Npε.S_{2}(N,\alpha)\ll Np^{\varepsilon}.

The bound in (3.9) is just a restatement of Theorem 1.7. As Np3+εN\ll p^{3+\varepsilon}, (3.10) can be obtained by just taking trivial bounds for all the terms. Thus, Theorem 1.4 follows.

4. Harmonic Analysis

We note that, bounding all the terms trivially, we get that S1(N,α)N2p2pεNp1+εS_{1}(N,\alpha)\ll\frac{N^{2}}{p^{2}}p^{\varepsilon}\ll Np^{1+\varepsilon}.
The proof for Theorem 1.7 starts with introducing the delta symbol, followed by using Voronoi summation to get additional cancellations on the terms that spring up. Barring a ‘main term’, this approach proves very fruitful, and we end up getting the desired upper bounds.

4.1. Application of the Delta Symbol

As gcd(n,p)=1(n,p)=1 in the definition of S1(N,α)S_{1}(N,\alpha), Proposition 2.3 guarantees the existence of a non-zero aα(modp)a_{\alpha}(\mathrm{mod}\ p) such that α(n+p2l)α(n)¯=ep(aαln¯)\alpha(n+p^{2}l)\overline{\alpha(n)}=e_{p}(a_{\alpha}l\overline{n}).

We recall that from Remark 1.8, we know that 0<lNp20<l\leq\frac{N}{p^{2}}, and Nn2NN\leq n\leq 2N in (3.6). Using m=n+p2lm=n+p^{2}l, we can rewrite (3.6) as

(4.1) S1(N,α)=ml,nλf(n+p2l)λf¯(n)ep(aαln¯)wN(m)wN(n)δ(mnp2l).S_{1}(N,\alpha)=\sum_{m}\sideset{}{{}^{*}}{\sum}_{l,n}\lambda_{f}(n+p^{2}l)\overline{\lambda_{f}}(n)e_{p}(a_{\alpha}l\overline{n})w_{N}(m)w_{N}(n)\delta(m-n-p^{2}l).

We can now introduce the δ\delta-symbol in the nn-sum and get via (2.13) (using C=NC=\sqrt{N}),

(4.2) S1(N,α)=ml,nλf(n+p2l)λf¯(n)ep(aαln¯)wN(m)wN(n)c2CS(0,mnp2l;c)gc(v)e((mnp2l)v)dv.S_{1}(N,\alpha)=\sum_{m}\sideset{}{{}^{*}}{\sum}_{l,n}\lambda_{f}(n+p^{2}l)\overline{\lambda_{f}}(n)e_{p}(a_{\alpha}l\overline{n})w_{N}(m)w_{N}(n)\\ \cdot\sum_{c\leq 2C}S(0,m-n-p^{2}l;c)\int_{-\infty}^{\infty}g_{c}(v)e\left((m-n-p^{2}l)v\right)dv.

Expanding out the Ramanujan sum and combining the mm and nn terms, we get

(4.3) S1(N,α)=lc2CVN,l(c).S_{1}(N,\alpha)=\sideset{}{{}^{*}}{\sum}_{l}\sum_{c\leq 2C}V_{N,l}(c).

Here,

(4.4) VN,l(c)=a(modc)e(ap2lc)gc(v)e(p2lv)T1(a,c,v)T2(a,c,v)𝑑v,V_{N,l}(c)=\sideset{}{{}^{*}}{\sum}_{a\negthickspace\negthickspace\negthickspace\pmod{c}}e\left(\frac{-ap^{2}l}{c}\right)\int_{-\infty}^{\infty}g_{c}(v)e(-p^{2}lv)\cdot T_{1}(a,c,v)\cdot T_{2}(a,c,v)dv,
(4.5) T1(a,c,v)=m=1λf(m)e(amc)wN(m)e(mv),T_{1}(a,c,v)=\sum_{m=1}^{\infty}\lambda_{f}(m)e\left(\frac{am}{c}\right)w_{N}(m)e(mv),
(4.6) T2(a,c,v)=gcd(n,p)=1λf¯(n)ep(aαln¯)e(anc)wN(n)e(nv).T_{2}(a,c,v)=\sum_{\text{gcd}(n,p)=1}\overline{\lambda_{f}}(n)e_{p}(a_{\alpha}l\overline{n})e\left(\frac{-an}{c}\right)w_{N}(n)e(-nv).

We further split (4.3) based on whether cc is relatively prime to pp, or not. We get

(4.7) S1(N,α)=S3(N,α)+S4(N,α),S_{1}(N,\alpha)=S_{3}(N,\alpha)+S_{4}(N,\alpha),

where

(4.8) S3(N,α)=lc2CVN,l(c),S_{3}(N,\alpha)=\sideset{}{{}^{*}}{\sum}_{l}\sideset{}{{}^{*}}{\sum}_{c\leq 2C}V_{N,l}(c),

and

(4.9) S4(N,α)=lc2CpcVN,l(c).S_{4}(N,\alpha)=\sideset{}{{}^{*}}{\sum}_{l}\sum_{\begin{subarray}{c}c\leq 2C\\ p\mid c\end{subarray}}V_{N,l}(c).

We have the following two lemmas.

Lemma 4.1.

Let ε>0\varepsilon>0 and Np3+εN\ll p^{3+\varepsilon}. Let S3(N,α)S_{3}(N,\alpha) be defined as in (4.8). Then

(4.10) S3(N,α)εNpε.S_{3}(N,\alpha)\ll_{\varepsilon}Np^{\varepsilon}.
Lemma 4.2.

Let ε>0\varepsilon>0 and Np3+εN\ll p^{3+\varepsilon}. Let S4(N,α)S_{4}(N,\alpha) be defined as in (4.9). Then

(4.11) S4(N,α)εNpε.S_{4}(N,\alpha)\ll_{\varepsilon}Np^{\varepsilon}.

It is clear that Theorem 1.7 follows from these two lemmas.

We proceed with the proof of Lemma 4.1 now, and delay the proof of Lemma 4.2 to Section 6.

4.2. Voronoi Summation

We recall that T1(a,c,v)T_{1}(a,c,v) and T2(a,c,v)T_{2}(a,c,v) are defined in (4.5) and (4.6).

We use Proposition 2.16 on T1(a,c,v)T_{1}(a,c,v) to get

(4.12) T1(a,c,v)=1cm=1λf(m)e(a¯mc)w~c,v,N(m).T_{1}(a,c,v)=\frac{1}{c}\sum_{m=1}^{\infty}\lambda_{f}(m)e\left(\frac{-\overline{a}m}{c}\right)\widetilde{w}_{c,v,N}(m).

Here,

(4.13) w~c,v,N(m)=2πikoJk1(4πmxc)e(xv)wN(x)𝑑x.\widetilde{w}_{c,v,N}(m)=2\pi i^{k}\int_{o}^{\infty}J_{k-1}\left(\frac{4\pi\sqrt{mx}}{c}\right)e(xv)w_{N}(x)dx.

For T2(a,c,v)T_{2}(a,c,v), (note that gcd(aαl,p)=1(a_{\alpha}l,p)=1) we can use Proposition 2.17, to get

(4.14) T2(a,c,v)=D0+D1+D2+D3T_{2}(a,c,v)=D_{0}+D_{1}+D_{2}+D_{3}

where

(4.15) D0=gcd(n,p)=1λf¯(n)w~pc,v,N(n)p2ce(ap2¯nc)Kl3(nc¯2aαl,1,1;p),D_{0}=\sum_{\text{gcd}(n,p)=1}\frac{\overline{\lambda_{f}}(n)\widetilde{w}_{pc,-v,N}(n)}{p^{2}c}e\left(\frac{\overline{ap^{2}}n}{c}\right)\text{Kl}_{3}(-n\overline{c}^{2}a_{\alpha}l,1,1;p),
(4.16) D1=gcd(n,p2)=pλf¯(n)w~pc,v,N(n)p2ce(ap2¯nc),D_{1}=\sum_{\text{gcd}(n,p^{2})=p}\frac{\overline{\lambda_{f}}(n)\widetilde{w}_{pc,-v,N}(n)}{p^{2}c}e\left(\frac{\overline{ap^{2}}n}{c}\right),
(4.17) D2=n=1λf¯(np)λf¯(p)w~pc,v,N(p2n)p2ce(a¯nc),D_{2}=\sum_{n=1}^{\infty}\frac{\overline{\lambda_{f}}(np)\overline{\lambda_{f}}(p)\widetilde{w}_{pc,-v,N}(p^{2}n)}{p^{2}c}e\left(\frac{\overline{a}n}{c}\right),

and

(4.18) D3=(1+1p)n=1λf¯(n)w~pc,v,N(p2n)pce(a¯nc).D_{3}=-\left(1+\frac{1}{p}\right)\sum_{n=1}^{\infty}\frac{\overline{\lambda_{f}}(n)\widetilde{w}_{pc,-v,N}(p^{2}n)}{pc}e\left(\frac{\overline{a}n}{c}\right).

Here, w~()\widetilde{w}_{-}(\cdot) is the same as in (4.13).

Thus, using (4.12) and (4.14), we have

(4.19) S3(N,α)=E0+E1+E2+E3+small errorS_{3}(N,\alpha)=E_{0}+E_{1}+E_{2}+E_{3}+\ \text{small error}

where

(4.20) E0=lc2Ca(modc)e(ap2lc)gc(v)e(p2lv)(1cmλf(m)e(a¯mc)w~c,v,N(m))(nλf¯(n)w~pc,v,N(n)p2ce(ap2¯nc)Kl3(nc¯2aαl,1,1;p))dv,E_{0}=\sideset{}{{}^{*}}{\sum}_{l}\ \sideset{}{{}^{*}}{\sum}_{c\leq 2C}\ \sideset{}{{}^{*}}{\sum}_{a\negthickspace\negthickspace\negthickspace\pmod{c}}e\left(\frac{-ap^{2}l}{c}\right)\int_{-\infty}^{\infty}g_{c}(v)e(-p^{2}lv)\left(\frac{1}{c}\sum_{m}\lambda_{f}(m)e\left(\frac{-\overline{a}m}{c}\right)\widetilde{w}_{c,v,N}(m)\right)\\ \left(\sideset{}{{}^{*}}{\sum}_{n}\frac{\overline{\lambda_{f}}(n)\widetilde{w}_{pc,-v,N}(n)}{p^{2}c}e\left(\frac{\overline{ap^{2}}n}{c}\right)\text{Kl}_{3}(-n\overline{c}^{2}a_{\alpha}l,1,1;p)\right)dv,

is the contribution to S3(N,α)S_{3}(N,\alpha) from D0D_{0}.
Similarly, E1,E2,E3E_{1},E_{2},E_{3} represent the contributions to EE from the remaining three terms of (4.14).

We have the two following lemmas.

Lemma 4.3.

Let ε>0\varepsilon>0 and let Np3+εN\ll p^{3+\varepsilon}. Let E0E_{0} be defined as in (4.19). We have

(4.21) E0εNpε.E_{0}\ll_{\varepsilon}Np^{\varepsilon}.
Lemma 4.4.

Let ε>0\varepsilon>0 and let Np3+εN\ll p^{3+\varepsilon}. Let E1,E2E_{1},E_{2} and E3E_{3} be defined as in (4.19). We have

(4.22) EjεN34pε.E_{j}\ll_{\varepsilon}N^{\frac{3}{4}}p^{\varepsilon}.

It is clear that together, these two lemmas imply Lemma 4.1. We prove Lemma 4.3 first, and postpone the proof of Lemma 4.4 to Section 6.

The main idea for proving Lemma 4.3 is via spectral analysis, using the Bruggeman-Kuznetsov formula, followed by applying spectral large sieve inequality, along with other results, to bound each of the resultant terms.

To set this up, note that the aa-sum in (4.20) forms a Kloosterman sum. Using this, we have

(4.23) E0=l,nmλf(m)λf¯(n)p2c2C1c2S(p¯2nm,p2l;c)Kl3(nc¯2aαl,1,1;p)IN(c,l,m,n),E_{0}=\sideset{}{{}^{*}}{\sum}_{l,n}\sum_{m}\frac{\lambda_{f}(m)\overline{\lambda_{f}}(n)}{p^{2}}\sideset{}{{}^{*}}{\sum}_{c\leq 2C}\frac{1}{c^{2}}S(\overline{p}^{2}n-m,-p^{2}l;c)\text{Kl}_{3}(-n\overline{c}^{2}a_{\alpha}l,1,1;p)I_{N}(c,l,m,n),

where

(4.24) IN(c,l,m,n)=gc(v)e(p2lv)w~c,v,N(m)w~pc,v,N(n)𝑑v.I_{N}(c,l,m,n)=\int_{-\infty}^{\infty}g_{c}(v)e(-p^{2}lv)\widetilde{w}_{c,v,N}(m)\widetilde{w}_{pc,-v,N}(n)dv.

Using the fact that,when gcd(r,p)=1(r,p)=1

(4.25) Kl3(r,1,1;p)=1ϕ(p)χ(p)τ(χ)3χ(r).\text{Kl}_{3}(r,1,1;p)=\frac{1}{\phi(p)}\sum_{\chi(p)}\tau(\chi)^{3}\chi(r).

we get that,

(4.26) E0=l,nmλf(m)λf¯(n)ϕ(p)p2χ(p)τ(χ)3χ(naαl)𝒦,E_{0}=\sideset{}{{}^{*}}{\sum}_{l,n}\sum_{m}\frac{\lambda_{f}(m)\overline{\lambda_{f}}(n)}{\phi(p)p^{2}}\sum_{\chi(p)}\tau(\chi)^{3}\chi(-na_{\alpha}l)\cdot\mathcal{K},

where

(4.27) 𝒦=c2C1c2S(p¯(np2m),pl;c)χ(c¯2)IN(c,l,m,n),\mathcal{K}=\sideset{}{{}^{*}}{\sum}_{c\leq 2C}\frac{1}{c^{2}}S(\overline{p}(n-p^{2}m),-pl;c)\chi(\overline{c}^{2})I_{N}(c,l,m,n),

We want to use the Bruggeman-Kuznetsov formula here. However, before that, we need to dyadically decompose (4.26).

4.3. Dyadic Partition of Unity

We want to modify (4.20) by applying a dyadic decomposition. We recall some definitions from Section 6.16.1 in [KrY21].
We define a number NN to be dyadic if N=2k2N=2^{\frac{k}{2}} for some kk\in\mathbb{Z}.
Let gg be a fixed smooth function supported on the interval [1,2][1,2]. A dyadic partition of unity is a partition of unity of the form kg(2k2x)1,\sum_{k\in\mathbb{Z}}g(2^{-\frac{k}{2}}x)\equiv 1, for x>0x>0.
The family gN(x)=g(xN)g_{N}(x)=g(\frac{x}{N}) forms a 11-inert family of functions.

We want to dyadically decomposition E0E_{0} (as defined in (4.26)) in the l,m,n,c,l,m,n,c, and vv variable. As l,m,n,cl,m,n,c are all positive integers, this decomposition is relatively simple, but a dyadic decomposition in the vv-variable is a bit more involved. We show that first.

4.3.1. Dyadic decomposition of IN(c,l,m,n)I_{N}(c,l,m,n)

Recall that gc(v)g_{c}(v) is defined in (2.11), and satisfies bounds given by (2.14) and (2.15). Also, w~c,v,N(m)\widetilde{w}_{c,v,N}(m) is defined in (4.13), and IN(c,l,m,n)I_{N}(c,l,m,n) is defined in (4.24).

We define,

(4.28) H(v)gc(v)e(p2lv)w~c,v,N(m)w~pc,v,N(n).H(v)\coloneqq g_{c}(v)e(-p^{2}lv)\widetilde{w}_{c,v,N}(m)\widetilde{w}_{pc,-v,N}(n).

Note that, we trivially have H(v)N2+εH(v)\ll N^{2+\varepsilon}.

(4.29) IN(c,l,m,n)=H(v)𝑑v.I_{N}(c,l,m,n)=\int_{-\infty}^{\infty}H(v)dv.

The bounds in (2.14) and (2.15) immediately imply that for any A>0A>0 arbitrarily large,

(4.30) IN(c,l,m,n)=|v|1cCH(v)𝑑v+O(pA).I_{N}(c,l,m,n)=\int_{\lvert v\rvert\ll\frac{1}{cC}}H(v)dv+O(p^{-A}).

We now state and prove the following lemma, that allows us to apply a dyadic decomposition.

Lemma 4.5.

Let A>0A>0 be arbitrarily large. Then there exists V0>0V_{0}>0, and a smooth function F(x)F(x) with |F(x)|1\lvert F(x)\rvert\leq 1 (both depending on AA), such that

(4.31) IN(c,l,m,n)=V0|v|1cCH(v)F(v)+O(pA).I_{N}(c,l,m,n)=\int_{V_{0}\leq\lvert v\rvert\ll\frac{1}{cC}}H(v)F(v)+O(p^{-A}).

Note that as the integral in (4.31) is defined away from zero, we can easily apply a dyadic decomposition to it.

Proof.

Fix A>0A>0. For any V0>0V_{0}>0, we can choose a smooth even function G(x)G(x) compactly supported on [2V0,2V0][-2V_{0},2V_{0}] such that G(x)1G(x)\leq 1 and G(x)=1G(x)=1 on [V0,V0][-V_{0},V_{0}]. Notice that 1G()1-G(\cdot) is supported on |x|>V0\lvert x\rvert>V_{0}.

Thus, for any V0<1cCV_{0}<\frac{1}{cC}, we can rewrite (4.29) as

(4.32) IN(c,l,m,n)=|v|1cCH(v)G(v)𝑑v+|v|1cCH(v)(1G(v))𝑑v+O(pA).I_{N}(c,l,m,n)=\int_{\lvert v\rvert\ll\frac{1}{cC}}H(v)G(v)dv+\int_{\lvert v\rvert\ll\frac{1}{cC}}H(v)(1-G(v))dv+O(p^{-A}).

Now, as GG is supported on [2V0,2V0][-2V_{0},2V_{0}] and equals 11 on [V0,V0][-V_{0},V_{0}], we can modify this further as,

(4.33) IN(c,l,m,n)=|v|2V0H(v)G(v)𝑑v+V0|v|1cCH(v)(1G)(v)𝑑v+O(pA).I_{N}(c,l,m,n)=\int_{\lvert v\rvert\leq 2V_{0}}H(v)G(v)dv+\int_{V_{0}\leq\lvert v\rvert\ll\frac{1}{cC}}H(v)(1-G)(v)dv+O(p^{-A}).

As Np3+εN\ll p^{3+\varepsilon}, and |v|2V0H(v)G(v)𝑑vV0N2+ε\int_{\lvert v\rvert\leq 2V_{0}}H(v)G(v)dv\ll V_{0}N^{2+\varepsilon}, if we choose V0=pA6V_{0}=p^{-A-6}, we will get that this integral is also O(pA)O(p^{-A}). Choose F(x)=1G(x)F(x)=1-G(x), completes the proof.

We can now apply a dyadic decomposition to (4.31), to get

(4.34) IN(c,l,m,n)=VdyadicV0|V|1cCIN,V(c,l,m,n)+O(pA),I_{N}(c,l,m,n)=\sum_{\begin{subarray}{c}V\text{dyadic}\\ V_{0}\leq\lvert V\rvert\ll\frac{1}{cC}\end{subarray}}I_{N,V}(c,l,m,n)+O(p^{-A}),

where

(4.35) IN,V(c,l,m,n)=VVgc(v)e(p2lv)w~c,v,N(m)w~pc,v,N(n)F(v)g(vV)𝑑v.I_{N,V}(c,l,m,n)=\int_{-V}^{V}g_{c}(v)e(-p^{2}lv)\widetilde{w}_{c,v,N}(m)\widetilde{w}_{pc,-v,N}(n)F(v)g\left(\frac{v}{V}\right)dv.

As F(v)F(v) and g(vV)g\left(\frac{v}{V}\right) are smooth, we can absorb them in w~c,v,N(m)\widetilde{w}_{c,v,N}(m), and by a slight abuse of notation, we denote the new function as w~c,v,N(m)\widetilde{w}_{c,v,N}(m) also.

4.3.2. Dyadic Decomposition of E0E_{0}

We can now use apply a dyadic decomposition to (4.26), to get

(4.36) E0=C0,L0,M0,N0,VdyadicEC0,L0,M0,N0,V+O(pA),E_{0}=\sum_{\begin{subarray}{c}C_{0},L_{0},M_{0},N_{0},V\\ \text{dyadic}\end{subarray}}E_{C_{0},L_{0},M_{0},N_{0},V}+O(p^{-A}),

where

(4.37) EC0,L0,M0,N0,V=ml,n,cλf(m)λf¯(n)ϕ(p)p2χ(p)τ(χ)3χ(naαl)𝒦C0,L0,M0,N0,V,E_{C_{0},L_{0},M_{0},N_{0},V}=\sum_{m}\sideset{}{{}^{*}}{\sum}_{l,n,c}\frac{\lambda_{f}(m)\overline{\lambda_{f}}(n)}{\phi(p)p^{2}}\sum_{\chi(p)}\tau(\chi)^{3}\chi(-na_{\alpha}l)\cdot\mathcal{K}_{C_{0},L_{0},M_{0},N_{0},V},

with

(4.38) 𝒦C0,L0,M0,N0,V=c1c2S(p¯(np2m),pl;c)χ2¯(c)IN,V(c,l,m,n)gT(l,m,n,c),\mathcal{K}_{C_{0},L_{0},M_{0},N_{0},V}=\sideset{}{{}^{*}}{\sum}_{c}\frac{1}{c^{2}}S(\overline{p}(n-p^{2}m),-pl;c)\overline{\chi^{2}}({c})I_{N,V}(c,l,m,n)g_{T}\left(l,m,n,c\right),

and

(4.39) gT(l,m,n,c)=g(lL0)g(mM0)g(nN0)g(cC0).g_{T}\left(l,m,n,c\right)=g\left(\frac{l}{L_{0}}\right)g\left(\frac{m}{M_{0}}\right)g\left(\frac{n}{N_{0}}\right)g\left(\frac{c}{C_{0}}\right).

Also, we claim the dyadic numbers satisfy

(4.40) 212M0pε, 212N0p2+ε, 212C02N, 212L0Np2,V0|V|1cC.2^{-\frac{1}{2}}\leq M_{0}\ll p^{\varepsilon},\ 2^{-\frac{1}{2}}\leq N_{0}\ll p^{2+\varepsilon},\ 2^{-\frac{1}{2}}\leq\ C_{0}\leq 2\sqrt{N},\ 2^{-\frac{1}{2}}\leq L_{0}\leq\frac{N}{p^{2}},\ V_{0}\leq\lvert V\rvert\ll\frac{1}{cC}.

The bounds for C0,L0C_{0},L_{0} and V0V_{0} are clear. The bounds for M0M_{0} and N0N_{0} follow as a consequence of Lemmas 4.6 and 4.8.

We will use the Bruggeman-Kuznetsov formula on (4.38) in Section 5. We finish this section by stating some results regarding the behaviour of IN,V(c,l,m,n)I_{N,V}(c,l,m,n).

4.4. Analysis of IN,V(c,l,m,n)I_{N,V}(c,l,m,n)

We use stationary phase methods to analyse the behaviour of some of the oscillatory integrals we have obtained so far. The general scheme is to identify regions when the integral is highly oscillatory, and when it is not.

4.4.1. Bounds for w~c,v,N(m)\widetilde{w}_{c,v,N}(m)

We want to use the properties of inert functions to analyse the behaviour of w~c,v,N(m)\widetilde{w}_{c,v,N}(m).

Lemma 4.6.

Let w~c,v,N(m)\widetilde{w}_{c,v,N}(m) be as in (4.13). Let l,m,n,c,vl,m,n,c,v be in dyadic intervals as before. We have

  • (a)

    (Non-Oscillatory) If M0NC0pε\frac{\sqrt{M_{0}N}}{C_{0}}\ll p^{\varepsilon}, then

    (4.41) w~c,v,N(m)=(M0NC0)k1NWT(c,m,v).\widetilde{w}_{c,v,N}(m)=\left(\frac{\sqrt{M_{0}N}}{C_{0}}\right)^{k-1}\cdot N\cdot W_{T}(c,m,v).

    Here, T=(V,M0,C0)T=(V,M_{0},C_{0}) and WT(c,m,v)W_{T}(c,m,v) is a pεp^{\varepsilon}-inert family in c,m,vc,m,v.
    Also, w~c,v,N(m)\widetilde{w}_{c,v,N}(m) is small unless N|V|pεN\lvert V\rvert\ll p^{\varepsilon}.

  • (b)

    (Oscillatory) If M0NC0pε\frac{\sqrt{M_{0}N}}{C_{0}}\gg p^{\varepsilon}, then

    (4.42) w~c,v,N(m)=NC0M0Ne(mc2v)WT(c,m,v)+O(pA).\widetilde{w}_{c,v,N}(m)=\frac{NC_{0}}{\sqrt{M_{0}N}}\cdot e\left(-\frac{m}{c^{2}v}\right)\cdot W_{T}(c,m,v)+O(p^{-A}).

    Here, T=(V,M0,C0)T=(V,M_{0},C_{0}) and WT(c,m,v)W_{T}(c,m,v) is pεp^{\varepsilon}-inert in c,m,vc,m,v. AA can be chosen to be arbitrarily large.
    Also, w~c,v,N(m)\widetilde{w}_{c,v,N}(m) is small unless N|V|M0NC0N\lvert V\rvert\asymp\frac{\sqrt{M_{0}N}}{C_{0}}.

Proof.

Note that, given the conditions of the lemma, and the fact that the integral in (4.13) can be truncated up to NN, we have 4πmxcM0NC0\frac{4\pi\sqrt{mx}}{c}\asymp\frac{\sqrt{M_{0}N}}{C_{0}}.

  • (a)

    If M0NC0pε\frac{\sqrt{M_{0}N}}{C_{0}}\ll p^{\varepsilon}, the Bessel function is not oscillatory, and we can write, Jk1(t)=tk1W(t)J_{k-1}(t)=t^{k-1}W(t), where tjW(j)(t)T0t^{j}W^{(j)}(t)\ll T_{0} with T0pεT_{0}\ll p^{\varepsilon}. This is the same derivative bound satisfied by a T0T_{0}-inert function (again, when YpεY\ll p^{\varepsilon}). Thus, we have

    (4.43) w~c,v,N(m)=2πik0(4πmxc)k1W(4πmxc)wN(x)e(xv)𝑑x.\widetilde{w}_{c,v,N}(m)=2\pi i^{k}\int_{0}^{\infty}\left(\frac{4\pi\sqrt{mx}}{c}\right)^{k-1}W\left(\frac{4\pi\sqrt{mx}}{c}\right)w_{N}(x)e(xv)dx.

    We can now use Proposition 2.10 and 2.12 to get (4.41). Proposition 2.12 also implies that w~c,v,N(m)\widetilde{w}_{c,v,N}(m) is small unless |V|1Npε\lvert V\rvert\ll\frac{1}{N}p^{\varepsilon}, or, N|V|pεN\lvert V\rvert\ll p^{\varepsilon}.

  • (b)

    We use the fact that when t1t\gg 1, the J-Bessel Function has an oscillatory behaviour and satisfies,

    (4.44) Jk1(t)=1t(eitW+(t)+eitWit),J_{k-1}(t)=\frac{1}{\sqrt{t}}(e^{it}W_{+}(t)+e^{-it}W_{it}),

    where W+W_{+} and WW_{-} satisfy the same derivative bounds as a 11-inert family of functions. Thus, we have (using t=4πmxc,T0=M0NC0t=\frac{4\pi\sqrt{mx}}{c},\ T_{0}=\frac{\sqrt{M_{0}N}}{C_{0}}),

    (4.45) w~c,v,N(m)=±0e(xv)wN(x)e±ittW±(t)𝑑x=±1T00eiϕ±(x)WN±(x)𝑑x,\widetilde{w}_{c,v,N}(m)=\sum_{\pm}\int_{0}^{\infty}e(xv)w_{N}(x)\frac{e^{\pm it}}{\sqrt{t}}W_{\pm}(t)dx=\sum_{\pm}\frac{1}{\sqrt{T_{0}}}\int_{0}^{\infty}e^{i\phi_{\pm}(x)}W_{N_{\pm}}(x)dx,

    where

    (4.46) ϕ±(x)=2πxv±4πmxc,\phi_{\pm}(x)=2\pi xv\pm 4\pi\frac{\sqrt{mx}}{c},

    and

    (4.47) WN±(x)=wN(x)W±(t)T0t.W_{N_{\pm}}(x)=\frac{w_{N}(x)W_{\pm}(t)\sqrt{T_{0}}}{\sqrt{t}}.

    We want to use Proposition 2.14 to analyse (4.45). Note that WN±(x)W_{N_{\pm}}(x) forms a pεp^{\varepsilon}-inert family.

    Notice that, as cNc\leq\sqrt{N}, when V>0V>0, |ϕ+(x)|=|2π(v+mcx)|1N\lvert\phi_{+}^{\prime}(x)\rvert=\lvert 2\pi\left(v+\frac{\sqrt{m}}{c\sqrt{x}}\right)\rvert\geq\frac{1}{N}. So we can use Proposition 2.14 (a) to conclude that the ‘+’ integral is small. Similarly, when V<0V<0, we can conclude that the ‘-’ integral is small.
    So, it suffices to consider the ‘-’ (resp. ‘+’) integral only when V>0V>0 (resp. V<0V<0). As the cases are similar, we only consider the first. So, assume V>0V>0.

    We can check that ϕ′′(x)=πmcxx1N2\phi_{-}^{\prime\prime}(x)=\frac{\pi\sqrt{m}}{cx\sqrt{x}}\gg\frac{1}{N^{2}}. Also, ϕ(x0)=0\phi_{-}^{\prime}(x_{0})=0, when x0=mc2v2x_{0}=\frac{m}{c^{2}v^{2}}; and ϕ′′(x0)=πc2v3m\phi^{\prime\prime}(x_{0})=\frac{\pi c^{2}v^{3}}{m}.
    Using Proposition 2.14 (b), we can rewrite (4.45) (when V>0V>0)as

    (4.48) w~c,v,N(m)=1T01ϕ′′(x0)eiϕ(x0)WT(x0)+O(pA).\widetilde{w}_{c,v,N}(m)=\frac{1}{\sqrt{T_{0}}}\frac{1}{\sqrt{\phi^{\prime\prime}(x_{0})}}e^{i\phi_{-}(x_{0})}\cdot W_{T}(x_{0})+O(p^{-A}).

    where T=(V,M0,C0)T=(V,M_{0},C_{0}) and WT(x0)W_{T}(x_{0}) is pεp^{\varepsilon}-inert in c,m,vc,m,v, and is supported on x0Nx_{0}\asymp N.

    The condition x0Nx_{0}\asymp N implies M0NC02V2\frac{M_{0}}{N{C_{0}}^{2}}\asymp V^{2}. This, along with the fact that T0=M0NC0T_{0}=\frac{\sqrt{M_{0}N}}{C_{0}} gives (4.42). This also implies the integral is small unless N|V|M0NC0N\lvert V\rvert\asymp\frac{\sqrt{M_{0}N}}{C_{0}}.

Remark 4.7.

We note that as |V|1C0C=1C0N\lvert V\rvert\leq\frac{1}{C_{0}C}=\frac{1}{C_{0}\sqrt{N}}, and C0NC_{0}\leq\sqrt{N}, N|V|M0NC0N\lvert V\rvert\asymp\frac{\sqrt{M_{0}N}}{C_{0}} is only possible when M0pεM_{0}\ll p^{\varepsilon}. As the condition in (a)(a) automatically implies M0pεM_{0}\ll p^{\varepsilon}, it suffices to only consider M0pεM_{0}\ll p^{\varepsilon} for (4.36).

We can also get a similar result for wpc,v,N(n)w_{pc,-v,N}(n) by following the same arguments. We state the result here. This also allows us to only consider N0p2+εN_{0}\ll p^{2+\varepsilon} for (4.36).

Lemma 4.8.

Let l,m,n,c,vl,m,n,c,v be in dyadic intervals at L0,M0,N0,C0,VL_{0},M_{0},N_{0},C_{0},V. We have

  • (a)

    (Non-Oscillatory) If N0NpC0pε\frac{\sqrt{N_{0}N}}{pC_{0}}\ll p^{\varepsilon}, then

    (4.49) w~pc,v,N(n)=(N0NpC0)k1NWT(c,n,v).\widetilde{w}_{pc,-v,N}(n)=\left(\frac{\sqrt{N_{0}N}}{pC_{0}}\right)^{k-1}\cdot N\cdot W_{T}(c,n,v).

    Here, T=(V,N0,C0)T=(V,N_{0},C_{0}) and WT(c,n,v)W_{T}(c,n,v) is a pεp^{\varepsilon}-inert family in c,n,vc,n,v.
    Also, w~pc,v,N(n)\widetilde{w}_{pc,-v,N}(n) is small unless N|V|pεN\lvert V\rvert\ll p^{\varepsilon}

  • (b)

    (Oscillatory) If N0NpC0pε\frac{\sqrt{N_{0}N}}{pC_{0}}\gg p^{\varepsilon}, then

    (4.50) w~pc,v,N(n)=NpC0N0Ne(np2c2v)WT(c,n,v)+O(pA).\widetilde{w}_{pc,-v,N}(n)=\frac{NpC_{0}}{\sqrt{N_{0}N}}\cdot e\left(\frac{n}{p^{2}c^{2}v}\right)\cdot W_{T}(c,n,v)+O(p^{-A}).

    Here, T=(V,N0,C0)T=(V,N_{0},C_{0}) and WT(c,n,v)W_{T}(c,n,v) is pεp^{\varepsilon}-inert in c,n,vc,n,v. AA can be chosen to be arbitrarily large.
    Also, w~pc,v,N(n)\widetilde{w}_{pc,-v,N}(n) is small unless N|V|N0NpC0N\lvert V\rvert\asymp\frac{\sqrt{N_{0}N}}{pC_{0}}.

4.4.2. Bounds on IN,V(c,l,m,n)I_{N,V}(c,l,m,n)

We can use Lemma 4.6 and Lemma 4.8, along with Proposition 2.14 to analyse the behaviour of

(4.51) IN,V(c,l,m,n)=V2Vgc(v)e(p2lv)w~c,v,N(m)w~pc,v,N(n)𝑑v.I_{N,V}(c,l,m,n)=\int_{V}^{2V}g_{c}(v)e(-p^{2}lv)\widetilde{w}_{c,v,N}(m)\widetilde{w}_{pc,-v,N}(n)dv.

We prove the following proposition.

Proposition 4.9.

Let IN,V(c,l,m,n)I_{N,V}(c,l,m,n) be as in (4.51). Let Let l,m,n,c,vl,m,n,c,v be in dyadic intervals at L0L_{0}, M0M_{0}, N0N_{0}, C0C_{0}, VV satisfying (4.40). Then

  • (a)

    (Non-oscillatory) If N|V|pεN\lvert V\rvert\ll p^{\varepsilon}, then

    (4.52) IN,V(c,l,m,n)=N2VWT(c,l,m,n).I_{N,V}(c,l,m,n)=N^{2}V\cdot W_{T}(c,l,m,n).

    Here, T=(C0,L0,M0,N0)T=(C_{0},L_{0},M_{0},N_{0}) and WT(c,l,m,n)W_{T}(c,l,m,n) is a pεp^{\varepsilon}-inert family in c,l,m,nc,l,m,n.
    Also, IN,V(c,l,m,n)I_{N,V}(c,l,m,n) is small unless C0N12εC_{0}\gg N^{\frac{1}{2}-\varepsilon}.

  • (b)

    (Oscillatory) If N|V|pεN\lvert V\rvert\gg p^{\varepsilon}, then IN,V(c,l,m,n)I_{N,V}(c,l,m,n) is small unless p2mn>0p^{2}m-n>0, and N|V|M0NC0N0NpC0N\lvert V\rvert\asymp\frac{\sqrt{M_{0}N}}{C_{0}}\asymp\frac{\sqrt{N_{0}N}}{pC_{0}}.
    If p2mn>0p^{2}m-n>0, and N|V|M0NC0N0NpC0N\lvert V\rvert\asymp\frac{\sqrt{M_{0}N}}{C_{0}}\asymp\frac{\sqrt{N_{0}N}}{pC_{0}},

    (4.53) IN,V(c,l,m,n)=N(N|V|)32e(2l(p2mn)c)WT(c,l,m,n)+O(pA).I_{N,V}(c,l,m,n)=\frac{N}{(N\lvert V\rvert)^{\frac{3}{2}}}\cdot e\left(\frac{2\sqrt{l(p^{2}m-n)}}{c}\right)\cdot W_{T}(c,l,m,n)+O(p^{-A}).

    Here, T=(C0,L0,M0,N0)T=(C_{0},L_{0},M_{0},N_{0}) and WT(c,l,m,n)W_{T}(c,l,m,n) is a pεp^{\varepsilon}-inert family in c,l,m,nc,l,m,n. AA can be chosen to be arbitrarily large.

Note that, in both cases we have the trivial bound, IN,V(c,l,m,n)NpεI_{N,V}(c,l,m,n)\ll Np^{\varepsilon}.

Proof.
  • (a)

    We have N|V|pεN\lvert V\rvert\ll p^{\varepsilon}. Now, using Lemma 4.6(b) (resp. Lemma 4.8(b)), if M0NC0pε\frac{\sqrt{M_{0}N}}{C_{0}}\gg p^{\varepsilon} (resp., N0NpC0pε\frac{\sqrt{N_{0}N}}{pC_{0}}\gg p^{\varepsilon}), then IN,V(c,l,m,n)I_{N,V}(c,l,m,n) is small, as N|V|N0NpC0N\lvert V\rvert\ll\frac{\sqrt{N_{0}N}}{pC_{0}} (resp. N|V|N0NpC0N\lvert V\rvert\ll\frac{\sqrt{N_{0}N}}{pC_{0}}).
    So, we must have M0NC0pε\frac{\sqrt{M_{0}N}}{C_{0}}\ll p^{\varepsilon} and N0NpC0pε\frac{\sqrt{N_{0}N}}{pC_{0}}\ll p^{\varepsilon}. Using Lemma 4.6(a) and Lemma 4.8(a), we get

    (4.54) IN,V(c,l,m,n)=N2(M0NC0)k1(N0NpC0)k1V2Vgc(v)e(p2lv)WT(c,m,n,v)𝑑v.I_{N,V}(c,l,m,n)=N^{2}\left(\frac{\sqrt{M_{0}N}}{C_{0}}\right)^{k-1}\left(\frac{\sqrt{N_{0}N}}{pC_{0}}\right)^{k-1}\int_{V}^{2V}g_{c}(v)e(-p^{2}lv)\cdot W_{T}(c,m,n,v)dv.

    Using the fact that gc(v)g_{c}(v) satisfies the same bounds as a 11-inert function, the integral is the Fourier transform of a pεp^{\varepsilon}-inert family of functions at p2lp^{2}l. As p2lN1|V|pεp^{2}l\leq N\ll\frac{1}{\lvert V\rvert}p^{\varepsilon}, we can use Proposition 2.12 to complete the proof. Also, as M0pεM_{0}\ll p^{\varepsilon}, M0NC0pε\frac{\sqrt{M_{0}N}}{C_{0}}\ll p^{\varepsilon} implies C0N12εC_{0}\gg N^{\frac{1}{2}-\varepsilon}.

  • (b)

    We have N|V|pεN\lvert V\rvert\gg p^{\varepsilon}. Using Lemma 4.6(a) and. Lemma 4.8(a), if M0NC0pε\frac{\sqrt{M_{0}N}}{C_{0}}\ll p^{\varepsilon} or if N0NpC0pε\frac{\sqrt{N_{0}N}}{pC_{0}}\ll p^{\varepsilon}, then IN,V(c,l,m,n)I_{N,V}(c,l,m,n) is small, as N|V|pεN\lvert V\rvert\gg p^{\varepsilon}. So, we must have M0NC0pε\frac{\sqrt{M_{0}N}}{C_{0}}\gg p^{\varepsilon} and N0NpC0pε\frac{\sqrt{N_{0}N}}{pC_{0}}\gg p^{\varepsilon}.
    Now, using Lemma 4.6(b) (resp. Lemma 4.8(b)) the integral then is small, unless N|V|M0NC0N0NpC0N\lvert V\rvert\asymp\frac{\sqrt{M_{0}N}}{C_{0}}\asymp\frac{\sqrt{N_{0}N}}{pC_{0}}. If N|V|M0NC0N0NpC0N\lvert V\rvert\asymp\frac{\sqrt{M_{0}N}}{C_{0}}\asymp\frac{\sqrt{N_{0}N}}{pC_{0}}, we have (for A>0A>0 arbitrarily large)

    (4.55) IN,V(c,l,m,n)=NpC02M0N0V2Vgc(v)e(np2c2vmc2vp2lv)WT(c,m,n,v)𝑑v+O(pA).I_{N,V}(c,l,m,n)=\frac{Np{C_{0}}^{2}}{\sqrt{M_{0}N_{0}}}\int_{V}^{2V}g_{c}(v)e\left(\frac{n}{p^{2}c^{2}v}-\frac{m}{c^{2}v}-p^{2}lv\right)W_{T}(c,m,n,v)dv+O(p^{-A}).

    We can then use Proposition 2.14(a) to conclude that the integral is small if p2mn0p^{2}m-n\leq 0. If p2mn>0p^{2}m-n>0, we can locate the stationary point and complete the proof using Proposition 2.14(b).

Remark 4.10.

In (4.38), we actually need to work with the product IN,V(c,l,m,n)gT(l,m,n,c)I_{N,V}(c,l,m,n)\cdot g_{T}(l,m,n,c). However, as gT(l,m,n,c)g_{T}(l,m,n,c) satisfies the same derivative bounds in l,m,n,cl,m,n,c as a 11-inert family of functions, we can use the results proven in Proposition 4.9 for the product IN,V(c,l,m,n)gT(l,m,n,c)I_{N,V}(c,l,m,n)\cdot g_{T}(l,m,n,c) also.

5. Spectral Analysis

We continue with our proof of Lemma 4.3. One simplification is immediate.

If χ\chi is trivial, then τ(χ)=1\tau(\chi)=1 and we can trivially bound all the terms in (4.37) (we can use IN,V(c,l,m,n)NpεI_{N,V}(c,l,m,n)\ll Np^{\varepsilon}) to get EC0,L0,M0,N0,V=O(Npε)E_{C_{0},L_{0},M_{0},N_{0},V}=O(Np^{\varepsilon}). As the number of dyadic terms is also O(pε)O(p^{\varepsilon}), we get the required bound for E0E_{0}. So, it suffices to work with χ\chi non-trivial in (4.37).

Recall 𝒦C0,L0,M0,N0,V\mathcal{K}_{C_{0},L_{0},M_{0},N_{0},V}, as defined in (4.38). Comparing with Proposition 2.18, we see that the parameters (q,m,n,χ,ϕ())(q,m,n,\chi,\phi(\cdot)) for the Bruggeman Kuznetsov formula in (2.53) takes the form (p,np2m,pl,χ2,IN,V())(p,n-p^{2}m,-pl,\chi^{2},I_{N,V}(\cdot)) in our case. Thus, from (4.38) and (2.53), we have

(5.1) 𝒦C0,L0,M0,N0,V=𝒦Maass+𝒦hol+𝒦Eis,\mathcal{K}_{C_{0},L_{0},M_{0},N_{0},V}=\mathcal{K}_{\text{Maass}}+\mathcal{K}_{\text{hol}}+\mathcal{K}_{\text{Eis}},

where

(5.2) 𝒦Maass=tj±Φ(tj)r1r2=pπit(r2,χ2)4πϵπV(p)π(1)δr1λ¯π(δ)(|np2m|)λ¯π(δ)(|pl|).\mathcal{K}_{\text{Maass}}=\sum_{t_{j}}\mathcal{L}^{\pm}\Phi(t_{j})\sum_{r_{1}r_{2}=p}\sum_{\pi\in\mathcal{H}_{it}(r_{2},\chi^{2})}\frac{4\pi\epsilon_{\pi}}{V(p)\mathscr{L}_{\pi}^{*}(1)}\sum_{\delta\mid r_{1}}\overline{\lambda}^{(\delta)}_{\pi}(\lvert n-p^{2}m\rvert)\overline{\lambda}^{(\delta)}_{\pi}(\lvert-pl\rvert).

𝒦hol\mathcal{K}_{\text{hol}} and 𝒦Eis\mathcal{K}_{\text{Eis}} are defined similarly, as in Proposition 2.18.

We can now rewrite (4.36) as

(5.3) EC0,L0,M0,N0,V=Maass+hol+Eis,E_{C_{0},L_{0},M_{0},N_{0},V}=\mathcal{M}_{\text{Maass}}+\mathcal{M}_{\text{hol}}+\mathcal{M}_{\text{Eis}},

where

(5.4) Maass=lL0,mM0,gcd(cln,p)=1λf(m)λf¯(n)ϕ(p)p2χ(p)τ(χ)3χ(naαl)𝒦Maass,\mathcal{M}_{\text{Maass}}=\sum_{\begin{subarray}{c}l\asymp L_{0},m\asymp M_{0},\cdots\\ \text{gcd}(cln,p)=1\end{subarray}}\frac{\lambda_{f}(m)\overline{\lambda_{f}}(n)}{\phi(p)p^{2}}\sideset{}{{}^{*}}{\sum}_{\chi(p)}\tau(\chi)^{3}\chi(-na_{\alpha}l)\cdot\mathcal{K}_{\text{Maass}},

is the contribution from the 𝒦Maass\mathcal{K}_{\text{Maass}} term. Similarly, hol\mathcal{M}_{\text{hol}} and Eis\mathcal{M}_{\text{Eis}} are the contributions from the 𝒦Eis\mathcal{K}_{\text{Eis}} and 𝒦hol\mathcal{K}_{\text{hol}} terms respectively.

We note that as the number of dyadic components in (4.36) is O(pε)O(p^{\varepsilon}), Lemma 4.3 follows if we can show each of these terms, Maass,Eis\mathcal{M}_{\text{Maass}},\ \mathcal{M}_{\text{Eis}} and hol\mathcal{M}_{\text{hol}} are O(Npε)O(Np^{\varepsilon}). This is what we will show.

We use two different expressions for the integral transforms holΦ()\mathcal{L}^{\text{hol}}\Phi(\cdot), and ±Φ()\mathcal{L}^{\pm}\Phi(\cdot). If N|V|pεN\lvert V\rvert\ll p^{\varepsilon} (non-oscillatory range), we use the ones given in (2.49) and (2.50). If N|V|pεN\lvert V\rvert\gg p^{\varepsilon} (oscillatory range), we use the ones given in (2.45) and (2.46).

We first work with the Maass form term.

5.1. Maass Form Term Analysis

As r1r2=pr_{1}r_{2}=p and δr1\delta\mid r_{1} in (5.2), the right side of (5.2) can be rewritten as a sum of three terms corresponding to (r1,r2,δ)=(1,p,1)(r_{1},r_{2},\delta)=(1,p,1), or (p,1,1)(p,1,1), or (p,1,p)(p,1,p).

Also, recall that λπ(δ)()\lambda_{\pi}^{(\delta)}(\cdot) is defined in (2.52). In particular, λπ(1)(m)=λπ(m)\lambda_{\pi}^{(1)}(m)=\lambda_{\pi}(m).

We can use this to split (5.4) as -

(5.5) Maass=Maass0+Maass1+Maass2,\mathcal{M}_{\text{Maass}}=\mathcal{M}_{\text{Maass}_{0}}+\mathcal{M}_{\text{Maass}_{1}}+\mathcal{M}_{\text{Maass}_{2}},

where

(5.6) Maass0=lL0,mM0,gcd(ln,p)=1λf(m)λf¯(n)ϕ(p)p2χ(p)τ(χ)3χ(naαl)tj±Φ(tj)πitj(p,χ2)4πϵπV(p)π(1)λ¯π(1)(|np2m|)λ¯π(1)(|pl|),\mathcal{M}_{\text{Maass}_{0}}=\sum_{\begin{subarray}{c}l\asymp L_{0},m\asymp M_{0},\cdots\\ \text{gcd}(ln,p)=1\end{subarray}}\frac{\lambda_{f}(m)\overline{\lambda_{f}}(n)}{\phi(p)p^{2}}\sideset{}{{}^{*}}{\sum}_{\chi(p)}\tau(\chi)^{3}\chi(-na_{\alpha}l)\\ \cdot\sum_{t_{j}}\mathcal{L}^{\pm}\Phi(t_{j})\sum_{\pi\in\mathcal{H}_{it_{j}}(p,\chi^{2})}\frac{4\pi\epsilon_{\pi}}{V(p)\mathscr{L}_{\pi}^{*}(1)}\overline{\lambda}^{(1)}_{\pi}(\lvert n-p^{2}m\rvert)\overline{\lambda}^{(1)}_{\pi}(\lvert-pl\rvert),

is the contribution from KMaassK_{\text{Maass}} when (r1,r2,δ)=(1,p,1)(r_{1},r_{2},\delta)=(1,p,1). The other two are defined similarly - Maass1\mathcal{M}_{\text{Maass}_{1}} corresponds to (r1,r2,δ)=(p,1,1)(r_{1},r_{2},\delta)=(p,1,1), and Maass2\mathcal{M}_{\text{Maass}_{2}} corresponds to (r1,r2,δ)=(p,1,p)(r_{1},r_{2},\delta)=(p,1,p).

We note that, for Maass1\mathcal{M}_{\text{Maass}_{1}} and Maass2\mathcal{M}_{\text{Maass}_{2}}, as r2=1r_{2}=1, we must have that χ2\chi^{2} is trivial. As χ\chi itself is non-trivial, χ\chi must be the unique quadratic character modulo pp (Legendre symbol).

We show that each of the three terms in (5.5) is O(Npε)O(Np^{\varepsilon}). We start with Maass0\mathcal{M}_{\text{Maass}_{0}}.

We split the analysis into two cases - non-oscillatory range (when N|V|pεN\lvert V\rvert\ll p^{\varepsilon}) and oscillatory range (when N|V|pεN\lvert V\rvert\gg p^{\varepsilon}).

5.1.1. Non-Oscillatory Case

We begin first by noting some properties of the functions Φ()\Phi(\cdot) and ±Φ()\mathcal{L}^{\pm}\Phi(\cdot).

Analysis of Φ()\Phi(\cdot):

We have

(5.7) Φ(y,)=py2IN,V(yp,l,m,n).\Phi(y,\cdot)=\frac{p}{y^{2}}I_{N,V}\left(\frac{y}{\sqrt{p}},l,m,n\right).

The Mellin transform of Φ\Phi is given by,

(5.8) Φ~(s,)=0Φ(y,)ys1𝑑y,\widetilde{\Phi}(s,\cdot)=\int_{0}^{\infty}\Phi(y,\cdot)y^{s-1}dy,

As N|V|pεN\lvert V\rvert\ll p^{\varepsilon}, Proposition 4.9(a) implies that IN,V(c,l,m,n)=N2VWT(c,l,m,n)I_{N,V}(c,l,m,n)=N^{2}\cdot V\cdot W_{T}(c,l,m,n), where WTW_{T} is pεp^{\varepsilon}-inert in the variables c,l,m,nc,l,m,n. We also recall that Proposition 4.9(a) allows us to restrict to C0N12εC_{0}\geq N^{\frac{1}{2}-\varepsilon}, as IN,V()I_{N,V}(\cdot) is small otherwise.

Thus, Φ~(s+1,)\widetilde{\Phi}(s+1,\cdot) is the Mellin transform of a pεp^{\varepsilon}-inert family of function at s1s-1. Proposition 2.13 then implies,

(5.9) Φ~(s+1,)=(Np)(N|V|)(pC0)s1WT(s1,l,m,n).\widetilde{\Phi}(s+1,\cdot)=(Np)(N\lvert V\rvert)(\sqrt{p}{C_{0}})^{s-1}W_{T}(s-1,l,m,n).

Here, WT()W_{T}(\cdot) is pεp^{\varepsilon}-inert in l,m,nl,m,n and is small when |Im(s1)|pε\lvert\text{Im}(s-1)\rvert\gg p^{\varepsilon}.

Analysis of ±Φ()\mathcal{L}^{\pm}\Phi(\cdot):

Recall h±(s,t)h_{\pm}(s,t) was defined in (2.51). We note that, for any σ0\sigma_{0} fixed and with d(σ02,0)1100d(\frac{\sigma_{0}}{2},\mathbb{Z}_{\leq 0})\geq\frac{1}{100}, we have the bound,

(5.10) h±(σ0+iv,t)(1+|t+v2|)σ012(1+|tv2|)σ012.h_{\pm}(\sigma_{0}+iv,t)\ll\left(1+\lvert t+\frac{v}{2}\rvert\right)^{\frac{\sigma_{0}-1}{2}}\left(1+\lvert t-\frac{v}{2}\rvert\right)^{\frac{\sigma_{0}-1}{2}}.

Also recall that ±Φ(t)\mathcal{L}^{\pm}\Phi(t) was defined in (2.50). We define,

(5.11) H±(s,t,l,m,n)=12πih±(s,t)Φ~(s+1,)(4π)s.H_{\pm}(s,t,l,m,n)=\frac{1}{2\pi i}h_{\pm}(s,t)\widetilde{\Phi}(s+1,\cdot)(4\pi)^{-s}.

So, we can rewrite (2.50) as

(5.12) ±Φ(t)=(s)=σH±(s,t,l,m,n)|(np2m)(pl)|s2𝑑s.\mathcal{L}^{\pm}\Phi(t)=\int_{\Re(s)=\sigma}H_{\pm}(s,t,l,m,n)\lvert(n-p^{2}m)(-pl)\rvert^{-\frac{s}{2}}ds.

We state the following lemma regarding the behaviour of ±Φ(t)\mathcal{L}^{\pm}\Phi(t).

Lemma 5.1.

Let H±(s,t,l,m,n)H_{\pm}(s,t,l,m,n) and ±Φ(t)\mathcal{L}^{\pm}\Phi(t) be as in (5.11) and (5.12). Let A>0A>0 be arbitrarily large.

  • (a)

    If |t|pε,\lvert t\rvert\gg p^{\varepsilon}, then

    (5.13) ±Φ(t)A,ε(1+|t|)A(Np)100.\mathcal{L}^{\pm}\Phi(t)\ll_{A,\varepsilon}(1+\lvert t\rvert)^{-A}(Np)^{-100}.
  • (b)

    If |t|pε,\lvert t\rvert\ll p^{\varepsilon}, Then for σ=(s)>14\sigma=\Re(s)>\frac{1}{4},

    (5.14) |H±(s,t)|(Np)1+ε(pC0)σ1(1+|s|)A.\lvert H_{\pm}(s,t)\rvert\ll(Np)^{1+\varepsilon}(\sqrt{p}{C_{0}})^{\sigma-1}(1+\lvert s\rvert)^{-A}.
Proof.
  • (a)

    If we take the contour of integration in (5.12) far to the left, we encounter poles at s2±it=0,1,2,\frac{s}{2}\pm it=0,-1,-2,\cdots. This means that |Im(s)||t|(Np)ε\lvert\text{Im}(s)\rvert\asymp\lvert t\rvert\gg(Np)^{\varepsilon}. Now, by (5.9), Φ~()\widetilde{\Phi}(\cdot) is very small at this height.

  • (b)

    This follows from repeated integration by parts on (5.8), and using Proposition 4.9(a) and (5.10).

Lemma 5.1 allows us to essentially restrict to |tj|pε\lvert t_{j}\rvert\ll p^{\varepsilon}.

Mellin Transform of H±(s,t,)H_{\pm}(s,t,\cdot):

Consider the functions H+()H_{+}(\cdot) as defined in (5.11). By the Mellin inversion theorem, we have for (uj)=σj>0\Re(u_{j})=\sigma_{j}>0, j=1,2,3j=1,2,3

(5.15) H+(s,t,l,m,n)=(uj)=σjlu1mu2nu3H~(s,t,u1,u2,u3)𝑑u1𝑑u2𝑑u3,H_{+}(s,t,l,m,n)=\int_{\Re(u_{j})=\sigma_{j}}l^{-u_{1}}m^{-u_{2}}n^{-u_{3}}\widetilde{H}(s,t,u_{1},u_{2},u_{3})\,du_{1}\,du_{2}\,du_{3},

where H~(s,t,u1,u2,u3)\widetilde{H}(s,t,u_{1},u_{2},u_{3}) is the (partial) Mellin transform of H+()H_{+}(\cdot) with respect to the variables l,m,nl,m,n. More specifically,

(5.16) H~(s,t,u1,u2,u3)=000H+(s,t,l,m,n)lu11mu21nu31𝑑l𝑑m𝑑n.\widetilde{H}(s,t,u_{1},u_{2},u_{3})=\int_{0}^{\infty}\int_{0}^{\infty}\int_{0}^{\infty}H_{+}(s,t,l,m,n)l^{u_{1}-1}m^{u_{2}-1}n^{u_{3}-1}\,dl\,dm\,dn.

Using the fact that Φ~()\widetilde{\Phi}(\cdot) is pεp^{\varepsilon}-inert in l,m,nl,m,n, when σ=(s)>1\sigma=\Re(s)>1 and σj=(uj)>0\sigma_{j}=\Re(u_{j})>0 for j=1,2,3j=1,2,3, we have

(5.17) H~(s,t,u1,u2,u3)(pC0)σ1(Np)1+εL0σ1M0σ2N0σ3(1+|s|)Aj=13(1+|uj|)A.\widetilde{H}(s,t,u_{1},u_{2},u_{3})\ll(\sqrt{p}C_{0})^{\sigma-1}(Np)^{1+\varepsilon}L_{0}^{\sigma_{1}}M_{0}^{\sigma_{2}}N_{0}^{\sigma_{3}}(1+\lvert s\rvert)^{-A}\prod_{j=1}^{3}(1+\lvert u_{j}\rvert)^{-A}.

Using (5.12) and (5.15), we get that,

(5.18) +Φ(t)=(s)=σ(uj)=σjlu1mu2nu3H~(s,t,u1,u2,u3)((p2mn)pl)s2𝑑s𝑑u1𝑑u2𝑑u3.\mathcal{L}^{+}\Phi(t)=\int_{\Re(s)=\sigma}\int_{\Re(u_{j})=\sigma_{j}}l^{-u_{1}}m^{-u_{2}}n^{-u_{3}}\widetilde{H}(s,t,u_{1},u_{2},u_{3})\left((p^{2}m-n)pl\right)^{-\frac{s}{2}}\,ds\,du_{1}\,du_{2}\,du_{3}.
Remark 5.2.

We can repeat these steps for H()H_{-}(\cdot), and get an analogous expression to (5.18) for Φ(t)\mathcal{L}^{-}\Phi(t).

Final Steps:

We show the required bounds for Maass0\mathcal{M}_{\text{Maass}_{0}} in the non-oscillatory region, when the sign is ++. The proof when the sign is - is very similar.
We recall that the sign in (5.6) depends on

sgn(pl(np2m))=sgn(p2mn)), as l>0.\text{sgn}(-pl(n-p^{2}m))=\text{sgn}(p^{2}m-n)),\text{ as }l>0.

So, we only consider terms in (5.6) with p2mn>0p^{2}m-n>0, and NVpεNV\ll p^{\varepsilon}.

Using (5.18), we can rewrite (5.6) as,

(5.19) Maass0=l,m,ngcd(ln,p)=1λf(m)λf¯(n)ϕ(p)p2χ(p)τ(χ)3χ(naαl)tjπitj(p,χ2)4πϵπV(p)π(1)λ¯π(p2mn)λ¯π(pl)(s)=σ(uj)=σjH~(s,t,u1,u2,u3)((p2mn)pl)s2dsdu1du2du3lu1mu2nu3.\mathcal{M}_{\text{Maass}_{0}}=\sum_{\begin{subarray}{c}l,m,n\\ \text{gcd}(ln,p)=1\end{subarray}}\frac{\lambda_{f}(m)\overline{\lambda_{f}}(n)}{\phi(p)p^{2}}\sideset{}{{}^{*}}{\sum}_{\chi(p)}\tau(\chi)^{3}\chi(-na_{\alpha}l)\\ \cdot\sum_{t_{j}}\sum_{\pi\in\mathcal{H}_{it_{j}}(p,\chi^{2})}\frac{4\pi\epsilon_{\pi}}{V(p)\mathscr{L}_{\pi}^{*}(1)}\overline{\lambda}_{\pi}(p^{2}m-n)\overline{\lambda}_{\pi}(pl)\int_{\begin{subarray}{c}\Re(s)=\sigma\\ \Re(u_{j})=\sigma_{j}\end{subarray}}\frac{\widetilde{H}(s,t,u_{1},u_{2},u_{3})}{\left((p^{2}m-n)pl\right)^{\frac{s}{2}}}\frac{ds\ du_{1}\ du_{2}\ du_{3}}{l^{u_{1}}m^{u_{2}}n^{u_{3}}}.

Rearranging the expression, we have

(5.20) Maass0=1ϕ(p)p2χ(p)tjπitj(p,χ2)(s)=σ(uj)=σjτ(χ)3χ(aα)4πϵπV(p)π(1)ps2H~(s,t,u1,u2,u3)(lgcd(l,p)=1λ¯π(pl)χ(l)ls2+u1)(m,ngcd(n,p)=1χ(n)λf(m)λf¯(n)λ¯π(p2mn)(p2mn)s2mu2nu3)dsdu1du2du3.\mathcal{M}_{\text{Maass}_{0}}=\frac{1}{\phi(p)p^{2}}\sideset{}{{}^{*}}{\sum}_{\chi(p)}\cdot\sum_{t_{j}}\sum_{\pi\in\mathcal{H}_{it_{j}}(p,\chi^{2})}\int_{\begin{subarray}{c}\Re(s)=\sigma\\ \Re(u_{j})=\sigma_{j}\end{subarray}}\frac{\tau(\chi)^{3}\chi(a_{\alpha})4\pi\epsilon_{\pi}}{V(p)\mathscr{L}_{\pi}^{*}(1)p^{\frac{s}{2}}}\widetilde{H}(s,t,u_{1},u_{2},u_{3})\\ \left(\sum_{\begin{subarray}{c}l\\ \text{gcd}(l,p)=1\end{subarray}}\frac{\overline{\lambda}_{\pi}(pl)\chi(l)}{l^{\frac{s}{2}+u_{1}}}\right)\left(\sum_{\begin{subarray}{c}m,n\\ \text{gcd}(n,p)=1\end{subarray}}\frac{\chi(-n)\lambda_{f}(m)\overline{\lambda_{f}}(n)\overline{\lambda}_{\pi}(p^{2}m-n)}{\left(p^{2}m-n\right)^{\frac{s}{2}}m^{u_{2}}n^{u_{3}}}\right)\ ds\ du_{1}\ du_{2}\ du_{3}.

We can combine the ll terms in (5.20) to form an L-function. In particular, we have

(5.21) lχ(l)λπ¯(pl)ls2+u1=λ¯π(p)lχ(l)λ¯π(l)ls2+u1=λ¯π(p)L(π¯χ,s2+u1).\sum_{l}\frac{\chi(l)\overline{\lambda_{\pi}}(pl)}{l^{\frac{s}{2}+u_{1}}}=\overline{\lambda}_{\pi}(p)\sum_{l}\frac{\chi(l)\overline{\lambda}_{\pi}(l)}{l^{\frac{s}{2}+u_{1}}}=\overline{\lambda}_{\pi}(p)L(\overline{\pi}\otimes\chi,\tfrac{s}{2}+u_{1}).

We bound (5.20) by taking the absolute value of the integrand. Note that as H~()\widetilde{H}(\cdot) decays rapidly along vertical lines (using (5.17)), we can restrict the integral up to |(s)|pε\lvert\Im(s)\rvert\ll p^{\varepsilon} and |(uj)|pε\lvert\Im(u_{j})\rvert\ll p^{\varepsilon}. We thus get, for arbitrarily large A>0A>0,

(5.22) Maass0(pC0)σ1(Np)1+εL0σ1M0σ2N0σ3ϕ(p)p2χ(p)tjπitj(p,χ2)|τ(χ)3χ(aα)4πϵπλ¯π(p)V(p)π(1)pσ2|(s)=σ(uj)=σj|m,ngcd(n,p)=1χ(n)λf(m)λf¯(n)λ¯π(p2mn)((p2mn))s2mu2nu3||L(π¯χ,s2+u1)|(1+|s|)Aj=13(1+|uj|)A𝑑s𝑑u1𝑑u2𝑑u3.\mathcal{M}_{\text{Maass}_{0}}\ll\frac{(\sqrt{p}C_{0})^{\sigma-1}(Np)^{1+\varepsilon}L_{0}^{\sigma_{1}}M_{0}^{\sigma_{2}}N_{0}^{\sigma_{3}}}{\phi(p)p^{2}}\sideset{}{{}^{*}}{\sum}_{\chi(p)}\cdot\sum_{t_{j}}\sum_{\pi\in\mathcal{H}_{it_{j}}(p,\chi^{2})}\left|\frac{\tau(\chi)^{3}\chi(a_{\alpha})4\pi\epsilon_{\pi}\overline{\lambda}_{\pi}(p)}{V(p)\mathscr{L}_{\pi}^{*}(1)p^{\frac{\sigma}{2}}}\right|\\ \int_{\begin{subarray}{c}\Re(s)=\sigma\\ \Re(u_{j})=\sigma_{j}\end{subarray}}\left|\sum_{\begin{subarray}{c}m,n\\ \text{gcd}(n,p)=1\end{subarray}}\frac{\chi(-n)\lambda_{f}(m)\overline{\lambda_{f}}(n)\overline{\lambda}_{\pi}(p^{2}m-n)}{\left((p^{2}m-n)\right)^{\frac{s}{2}}m^{u_{2}}n^{u_{3}}}\right|\frac{\lvert L(\overline{\pi}\otimes\chi,\tfrac{s}{2}+u_{1})\rvert}{(1+\lvert s\rvert)^{A}\prod_{j=1}^{3}(1+\lvert u_{j}\rvert)^{A}}\ ds\ du_{1}\ du_{2}\ du_{3}.

Lemma 5.1 and Proposition 4.9(a) allow us to restrict (up to a small error) to when |tj|pε\lvert t_{j}\rvert\ll p^{\varepsilon}, and N12εC0NN^{\frac{1}{2}-\varepsilon}\leq C_{0}\leq\sqrt{N}.

Additionally we have the bounds |τ(χ)|=p,|V(p)|p\lvert\tau(\chi)\rvert=\sqrt{p},\ \lvert V(p)\rvert\asymp p. Also, as πitj(p,χ2)\pi\in\mathcal{H}_{it_{j}}(p,\chi^{2}), |λπ(p)|1\lvert\lambda_{\pi}(p)\rvert\leq 1 (pp divides the level). These, along with (5.21), and the fact that χ(n)=χ(p2mn)\chi(-n)=\chi(p^{2}m-n) give us,

(5.23) Maass0p32L0σ1M0σ2N0σ3(Np)1+σ2pp3pσ2𝒮,\mathcal{M}_{\text{Maass}_{0}}\ll\ \frac{p^{\frac{3}{2}}L_{0}^{\sigma_{1}}M_{0}^{\sigma_{2}}N_{0}^{\sigma_{3}}(Np)^{\frac{1+\sigma}{2}}}{p\cdot p^{3}\cdot p^{\frac{\sigma}{2}}}\cdot\mathcal{S},

where

𝒮=pεtjpεχ(p)πitj(p,χ2)(s)=σ,(s)pε(uj)=σj(uj)pε|L(π¯χ,s2+u1)||mM0λf(m)mu2nN0λ¯f(n)λ¯π(p2mn)χ(p2mn)(p2mn)s2nu3|dsdu1du2du3.\mathcal{S}=p^{\varepsilon}\cdot\sum_{t_{j}\ll p^{\varepsilon}}\sideset{}{{}^{*}}{\sum}_{\chi(p)}\sum_{\pi\in\mathcal{H}_{it_{j}}(p,\chi^{2})}\int_{\begin{subarray}{c}\Re(s)=\sigma,\ \Im(s)\ll p^{\varepsilon}\\ \Re(u_{j})=\sigma_{j}\ \Im(u_{j})\ll p^{\varepsilon}\end{subarray}}\lvert L(\overline{\pi}\otimes\chi,\tfrac{s}{2}+u_{1})\rvert\\ \cdot\left|\sum_{m\asymp M_{0}}\frac{\lambda_{f}(m)}{m^{u_{2}}}\sum_{n\asymp N_{0}}\frac{\overline{\lambda}_{f}(n)\overline{\lambda}_{\pi}(p^{2}m-n)\chi(p^{2}m-n)}{(p^{2}m-n)^{\frac{s}{2}}n^{u_{3}}}\right|\ ds\ du_{1}\ du_{2}\ du_{3}.

Using the Cauchy-Schwarz inequality on 𝒮\mathcal{S} we get

(5.24) |𝒮|(𝒮1)12(𝒮2)12,\lvert\mathcal{S}\rvert\leq(\mathcal{S}_{1})^{\frac{1}{2}}\cdot(\mathcal{S}_{2})^{\frac{1}{2}},

where

(5.25) 𝒮1=pεtjpεχ(p)πitj(p,χ2)(s)=σ,(s)pε(uj)=σj(uj)pε|L(π¯χ,s2+u1)|2𝑑s𝑑u1𝑑u2𝑑u3,\mathcal{S}_{1}=p^{\varepsilon}\cdot\sum_{t_{j}\ll p^{\varepsilon}}\sideset{}{{}^{*}}{\sum}_{\chi(p)}\sum_{\pi\in\mathcal{H}_{it_{j}}(p,\chi^{2})}\int_{\begin{subarray}{c}\Re(s)=\sigma,\ \Im(s)\ll p^{\varepsilon}\\ \Re(u_{j})=\sigma_{j}\ \Im(u_{j})\ll p^{\varepsilon}\end{subarray}}\left|L(\overline{\pi}\otimes\chi,\tfrac{s}{2}+u_{1})\right|^{2}\ ds\ du_{1}\ du_{2}\ du_{3},

and

(5.26) 𝒮2=pεtjpεχ(p)πitj(p,χ2)(s)=σ,(s)pε(uj)=σj(uj)pε|mM0λf(m)mu2nN0λ¯f(n)λ¯π(p2mn)χ(p2mn)(p2mn)s2nu3|2dsdu1du2du3.\mathcal{S}_{2}=p^{\varepsilon}\cdot\sum_{t_{j}\ll p^{\varepsilon}}\sideset{}{{}^{*}}{\sum}_{\chi(p)}\sum_{\pi\in\mathcal{H}_{it_{j}}(p,\chi^{2})}\\ \cdot\int_{\begin{subarray}{c}\Re(s)=\sigma,\ \Im(s)\ll p^{\varepsilon}\\ \Re(u_{j})=\sigma_{j}\ \Im(u_{j})\ll p^{\varepsilon}\end{subarray}}\left|\sum_{m\asymp M_{0}}\frac{\lambda_{f}(m)}{m^{u_{2}}}\sum_{n\asymp N_{0}}\frac{\overline{\lambda}_{f}(n)\overline{\lambda}_{\pi}(p^{2}m-n)\chi(p^{2}m-n)}{(p^{2}m-n)^{\frac{s}{2}}n^{u_{3}}}\right|^{2}ds\ du_{1}\ du_{2}\ du_{3}.

Let (s)=v\Im(s)=v, and (u1)=v1\Im(u_{1})=v_{1}, and v0=v+2v12v_{0}=\tfrac{v+2v_{1}}{2}. We now choose σ=12,σ1=σ3=14+ε2,σ2=ε2\sigma=\frac{1}{2},\ \sigma_{1}=\sigma_{3}=\frac{1}{4}+\frac{\varepsilon}{2},\ \sigma_{2}=\frac{\varepsilon}{2}, and use the spectral large sieve inequality to get bounds on (5.25) and (5.26).

We note that the map (π,χ)π¯χ(\pi,\chi)\rightarrow\overline{\pi}\otimes\chi is an at most two-to-one map, and π¯χitj(p2,1)\overline{\pi}\otimes\chi\in\mathcal{H}_{it_{j}}(p^{2},1).

Using Proposition 2.1 (with X=1X=1) in (5.25), we get

(5.27) 𝒮1tjpεϕitj(p2,1)vpεv1pε|L(ϕ,1+ε2+iv0|2dvdv1vpεv1pεtjpεϕitj(p2,1)(|np1+ελϕ(n)n1+ε2+iv0V(np)|2+|np1+ελϕ¯(n)n1ε2iv0V(np)|2)dvdv1.\mathcal{S}_{1}\ll\sum_{t_{j}\ll p^{\varepsilon}}\sum_{\phi\in\mathcal{H}_{itj}(p^{2},1)}\int_{\begin{subarray}{c}v\ll p^{\varepsilon}\\ v_{1}\ll p^{\varepsilon}\end{subarray}}\lvert L(\phi,\tfrac{1+\varepsilon}{2}+iv_{0}\rvert^{2}\ dv\ dv_{1}\\ \ll\int_{\begin{subarray}{c}v\ll p^{\varepsilon}\\ v_{1}\ll p^{\varepsilon}\end{subarray}}\sum_{t_{j}\ll p^{\varepsilon}}\sum_{\phi\in\mathcal{H}_{it_{j}}(p^{2},1)}\left(\left|\sum_{n\ll p^{1+\varepsilon}}\frac{\lambda_{\phi}(n)}{n^{\tfrac{1+\varepsilon}{2}+iv_{0}}}V\left(\frac{n}{p}\right)\right|^{2}+\left|\sum_{n\ll p^{1+\varepsilon}}\frac{\overline{\lambda_{\phi}}(n)}{n^{\tfrac{1-\varepsilon}{2}-iv_{0}}}V\left(\frac{n}{p}\right)\right|^{2}\right)\ dv\ dv_{1}.

Using Proposition 2.19 twice in (5.27), we get that

(5.28) 𝒮1(p2+ε+p1+ε)1+ε(np1+ε1n1+ε(vpεv1pε|niv0V(np)|2dvdv1)+np1+ε1n1ε(vpεv1pε|niv0V(np)|2dvdv1))p2+ε.\mathcal{S}_{1}\ll\left(p^{2+\varepsilon}+p^{1+\varepsilon}\right)^{1+\varepsilon}\cdot\left(\sum_{n\leq p^{1+\varepsilon}}\frac{1}{n^{1+\varepsilon}}\left(\int_{\begin{subarray}{c}v\ll p^{\varepsilon}\\ v_{1}\ll p^{\varepsilon}\end{subarray}}\left\lvert n^{-iv_{0}}V\left(\frac{n}{p}\right)\right\rvert^{2}\ dv\ dv_{1}\right)\right.\\ +\left.\sum_{n\leq p^{1+\varepsilon}}\frac{1}{n^{1-\varepsilon}}\left(\int_{\begin{subarray}{c}v\ll p^{\varepsilon}\\ v_{1}\ll p^{\varepsilon}\end{subarray}}\left\lvert n^{-iv_{0}}V\left(\frac{n}{p}\right)\right\rvert^{2}\ dv\ dv_{1}\right)\right)\ll p^{2+\varepsilon}.

For the bound on 𝒮2\mathcal{S}_{2} we state and prove the following lemma.

Lemma 5.3.

For any ε>0\varepsilon>0,

(5.29) 𝒮2p2+ε.\mathcal{S}_{2}\ll p^{2+\varepsilon}.

Notice that assuming Lemma 5.3, we can complete the proof pretty easily. Using (5.24),(5.28) and (5.29) in (5.23), we have (with σ=12,σ1=σ3=14+ε2,σ2=ε2\sigma=\frac{1}{2},\ \sigma_{1}=\sigma_{3}=\frac{1}{4}+\frac{\varepsilon}{2},\sigma_{2}=\frac{\varepsilon}{2}),

(5.30) Maass0p32(L0M0N0)ε2(Np)1+ε2pp3p12p2+εNpε.\mathcal{M}_{\text{Maass}_{0}}\ll\frac{p^{\frac{3}{2}}(L_{0}M_{0}N_{0})^{\frac{\varepsilon}{2}}(Np)^{1+\frac{\varepsilon}{2}}}{p\cdot p^{3}\cdot p^{\frac{1}{2}}}\cdot p^{2+\varepsilon}\ll Np^{\varepsilon}.

Proof of Lemma 5.3

We first define,

(5.31) S2,N0(m)=nN0λ¯f(n)λ¯π(p2mn)χ(p2mn)(p2mn)s2nu3.S_{2,N_{0}}(m)=\sum_{n\asymp N_{0}}\frac{\overline{\lambda}_{f}(n)\overline{\lambda}_{\pi}(p^{2}m-n)\chi(p^{2}m-n)}{(p^{2}m-n)^{\frac{s}{2}}n^{u_{3}}}.

Now, consider the mm-sum in (5.26) (with σ2=ε2\sigma_{2}=\frac{\varepsilon}{2}). We recall that M0pεM_{0}\ll p^{\varepsilon}, and N0p2+εN_{0}\ll p^{2+\varepsilon}. Using the Cauchy-Schwarz inequality here, we see that

(5.32) |mM0λf(m)mu2S2,N0(m)|2mM0|λf(m)mε2|2mM0|S2,N0(m)|2pεmM0|S2,N0(m)|2.\left|\sum_{m\asymp M_{0}}\frac{\lambda_{f}(m)}{m^{u_{2}}}S_{2,N_{0}}(m)\right|^{2}\leq\sum_{m\asymp M_{0}}\left|\frac{\lambda_{f}(m)}{m^{\frac{\varepsilon}{2}}}\right|^{2}\cdot\sum_{m\asymp M_{0}}\lvert S_{2,N_{0}}(m)\rvert^{2}\ll p^{\varepsilon}\sum_{m\asymp M_{0}}\lvert S_{2,N_{0}}(m)\rvert^{2}.

Now,changing variables np2mnn\rightarrow p^{2}m-n, we can rewrite (5.31) as,

(5.33) S2,N0(m)=np2+ελ¯f(p2mn)(p2mn)u3λ¯πχ(n)ns2.S_{2,N_{0}}(m)=\sum_{n\ll p^{2+\varepsilon}}\frac{\overline{\lambda}_{f}(p^{2}m-n)}{(p^{2}m-n)^{u_{3}}}\cdot\frac{\overline{\lambda}_{\pi\otimes\chi}(n)}{n^{\frac{s}{2}}}.

Now, we can use (5.32), and (5.33) in (5.26), to get

(5.34) 𝒮2pε(s)=σ(uj)=σjmM0tjpεϕitj(p2,1)|np2+ελ¯f(p2mn)(p2mn)u3λϕ(n)ns2|2dsdu1du2du3.\mathcal{S}_{2}\ll p^{\varepsilon}\int_{\begin{subarray}{c}\Re(s)=\sigma\\ \Re(u_{j})=\sigma_{j}\end{subarray}}\sum_{m\asymp M_{0}}\sum_{t_{j}\ll p^{\varepsilon}}\sum_{\phi\in\mathcal{H}_{itj}(p^{2},1)}\left|\sum_{n\ll p^{2+\varepsilon}}\frac{\overline{\lambda}_{f}(p^{2}m-n)}{(p^{2}m-n)^{u_{3}}}\cdot\frac{\lambda_{\phi}(n)}{n^{\frac{s}{2}}}\right|^{2}\ ds\ du_{1}\ du_{2}\ du_{3}.

Using Proposition 2.19 in (5.34), we get

(5.35) 𝒮2(s)=σ(uj)=σj(p2+ε+p2+ε)1+εmM0np2+ε|λ¯f(p2mn)(p2mn)u31ns2|2dsdu1du2du3.\mathcal{S}_{2}\ll\int_{\begin{subarray}{c}\Re(s)=\sigma\\ \Re(u_{j})=\sigma_{j}\end{subarray}}\left(p^{2+\varepsilon}+p^{2+\varepsilon}\right)^{1+\varepsilon}\sum_{m\asymp M_{0}}\sum_{n\ll p^{2+\varepsilon}}\left|\frac{\overline{\lambda}_{f}(p^{2}m-n)}{(p^{2}m-n)^{u_{3}}}\cdot\frac{1}{n^{\frac{s}{2}}}\right|^{2}\ ds\ du_{1}\ du_{2}\ du_{3}.

Using the fact that the integrals can be restricted up to |(s)|pε\lvert\Im(s)\rvert\ll p^{\varepsilon} and |(uj)|pε\lvert\Im(u_{j})\rvert\ll p^{\varepsilon}, and that σ=12\sigma=\tfrac{1}{2}, and σ3=14+ε2\sigma_{3}=\tfrac{1}{4}+\tfrac{\varepsilon}{2}, we can simplify (5.35) as,

(5.36) 𝒮2(p2+ε+p2+ε)1+εmM0np2+ε|λ¯f(p2mn)(p2mn)14+ε21n14|2.\mathcal{S}_{2}\ll\left(p^{2+\varepsilon}+p^{2+\varepsilon}\right)^{1+\varepsilon}\sum_{m\asymp M_{0}}\sum_{n\ll p^{2+\varepsilon}}\left|\frac{\overline{\lambda}_{f}(p^{2}m-n)}{(p^{2}m-n)^{\frac{1}{4}+\frac{\varepsilon}{2}}}\cdot\frac{1}{n^{\frac{1}{4}}}\right|^{2}.

Let cf(m,n)|λf(p2mn)|2(p2mn)12+εn12c_{f}(m,n)\coloneqq\frac{\lvert\lambda_{f}(p^{2}m-n)\rvert^{2}}{(p^{2}m-n)^{\frac{1}{2}+\varepsilon}n^{\frac{1}{2}}}. It suffices to show that mM0npεcf(m,n)pε.\sum_{m\asymp M_{0}}\sum_{n\ll p^{\varepsilon}}c_{f}(m,n)\ll p^{\varepsilon}.

We separate the nn-sum into two parts depending on if np2m2n\geq\frac{p^{2}m}{2} or not.

In the former case,

(5.37) mM0p2m2npεcf(m,n)mM0p2m2np2+ε|λf(p2mn)|2(p2mn)12+ε(p2m)12mM0np2+ε|λf(n)|2n1+εM0pεpε.\sum_{m\asymp M_{0}}\sum_{\frac{p^{2}m}{2}\leq n\ll p^{\varepsilon}}c_{f}(m,n)\ll\sum_{m\asymp M_{0}}\sum_{\frac{p^{2}m}{2}\leq n\ll p^{2+\varepsilon}}\frac{\lvert\lambda_{f}(p^{2}m-n)\rvert^{2}}{(p^{2}m-n)^{\frac{1}{2}+\varepsilon}(p^{2}-m)^{\frac{1}{2}}}\\ \ll\sum_{m\asymp M_{0}}\sum_{n\ll p^{2+\varepsilon}}\frac{\lvert\lambda_{f}(n)\rvert^{2}}{n^{1+\varepsilon}}\ll M_{0}\cdot p^{\varepsilon}\ll p^{\varepsilon}.

The latter case (when n<p2m2n<\frac{p^{2}m}{2}), requires more work. We split the nn-sum into dyadic segments of size, nN1n\asymp N_{1}, N1<p2mN_{1}<p^{2}m.

(5.38) mM0n<p2m2cf(m,n)mM0N1 dyadicnN1|λf(p2mn)|2(p2mn)12+εn12mM0N1 dyadic1p1+εN112nN1|λf(p2mn)|2mM0N1 dyadic1pN112(nN1|λf(p2mn)|4)12N112.\sum_{m\asymp M_{0}}\sum_{n<\frac{p^{2}m}{2}}c_{f}(m,n)\ll\sum_{m\asymp M_{0}}\sum_{N_{1}\text{ dyadic}}\sum_{n\asymp N_{1}}\frac{\lvert\lambda_{f}(p^{2}m-n)\rvert^{2}}{(p^{2}m-n)^{\frac{1}{2}+\varepsilon}n^{\frac{1}{2}}}\\ \ll\sum_{\begin{subarray}{c}m\asymp M_{0}\\ N_{1}\text{ dyadic}\end{subarray}}\frac{1}{p^{1+\varepsilon}N_{1}^{\frac{1}{2}}}\sum_{n\asymp N_{1}}\lvert\lambda_{f}(p^{2}m-n)\rvert^{2}\ll\sum_{\begin{subarray}{c}m\asymp M_{0}\\ N_{1}\text{ dyadic}\end{subarray}}\frac{1}{pN_{1}^{\frac{1}{2}}}\left(\sum_{n\asymp N_{1}}\lvert\lambda_{f}(p^{2}m-n)\rvert^{4}\right)^{\frac{1}{2}}\cdot N_{1}^{\frac{1}{2}}.

Here the last inequality follows from an application of Cauchy’s inequality. Using (2.58), we can now get

(5.39) nN1|λf(p2mn)|4np2m|λf(n)|4(p2m)1+ε.\sum_{n\asymp N_{1}}\lvert\lambda_{f}(p^{2}m-n)\rvert^{4}\ll\sum_{n\ll p^{2}m}\lvert\lambda_{f}(n)\rvert^{4}\ll(p^{2}m)^{1+\varepsilon}.

Using (5.39) in (5.38), we get that,

(5.40) mM0n<p2m2cf(m,n)mM0N1 dyadicnN1|λf(p2mn)|2(p2mn)12n12+εpε.\sum_{m\asymp M_{0}}\sum_{n<\frac{p^{2}m}{2}}c_{f}(m,n)\ll\sum_{m\asymp M_{0}}\sum_{N_{1}\text{ dyadic}}\sum_{n\asymp N_{1}}\frac{\lvert\lambda_{f}(p^{2}m-n)\rvert^{2}}{(p^{2}m-n)^{\frac{1}{2}}n^{\frac{1}{2}+\varepsilon}}\ll p^{\varepsilon}.

Now, (5.37) and (5.40) jointly imply that mM0npεcf(m,n)pε.\sum_{m\asymp M_{0}}\sum_{n\ll p^{\varepsilon}}c_{f}(m,n)\ll p^{\varepsilon}. This completes the proof of the lemma.

5.1.2. Oscillatory Case

In this range, NVpεNV\gg p^{\varepsilon}. The analysis is similar to the non-oscillatory case, although we use the Bessel integral form for the Bruggeman-Kuznetsov formula. Once again, we begin by noting some properties of the functions Φ()\Phi(\cdot) and ±Φ()\mathcal{L}^{\pm}\Phi(\cdot).

Analysis of Φ()\Phi(\cdot):

Recall Φ(y,)\Phi(y,\cdot), as defined in (5.7). As N|V|pεN\lvert V\rvert\gg p^{\varepsilon}, Proposition 4.9(b) implies that IN,V(c,l,m,n)I_{N,V}(c,l,m,n) is small unless p2mn>0p^{2}m-n>0 and NVM0NC0N0NpC0NV\asymp\frac{\sqrt{M_{0}N}}{C_{0}}\asymp\frac{\sqrt{N_{0}N}}{pC_{0}}, in which case we have (restating (4.53))

(5.41) IN,V(c,l,m,n)=N(NV)32e(2l(p2mn)c)WT(l,m,n,c)+O(pA).I_{N,V}(c,l,m,n)=\frac{N}{(NV)^{\frac{3}{2}}}e\left(\frac{2\sqrt{l(p^{2}m-n)}}{c}\right)\cdot W_{T}(l,m,n,c)+O(p^{-A}).

We recall that the sign in (5.6) is sgn((np2m)(pl))=sgn(p2mn)\text{sgn}((n-p^{2}m)(-pl))=\text{sgn}(p^{2}m-n). Now, as IN,V(c,l,m,n)I_{N,V}(c,l,m,n) is small if p2mn<0p^{2}m-n<0, we only need to consider the case when the sign is ++.

Analysis of +Φ()\mathcal{L}^{+}\Phi(\cdot):

Recall that +Φ()\mathcal{L}^{+}\Phi(\cdot) is defined as in (2.46).
Let z=2pl(p2mn)z=2\sqrt{pl(p^{2}m-n)}. As p2mn>0p^{2}m-n>0, this is well defined. Note that z(NN0p)12z\asymp\left(\frac{NN_{0}}{p}\right)^{\frac{1}{2}}.
Now,

(5.42) J2ir(x)J2ir(x)sinh(πr)=2πicos(xcoshy)e(ryπ)𝑑y.\frac{J_{2ir(x)}-J_{-2ir}(x)}{\sinh\left(\pi r\right)}=\frac{2}{\pi i}\int_{-\infty}^{\infty}\cos(x\cosh y)e\left(\frac{ry}{\pi}\right)dy.

We can use (4.53) and (5.42) to rewrite +Φ()\mathcal{L}^{+}\Phi(\cdot) (up to a small error term) as,

(5.43) +Φ(tj)=Npπ(NV)320cos(2πzxcoshy)e(tjyπ)e(zx)WT(xp,)dydxx2.\mathcal{L}^{+}\Phi(t_{j})=\frac{Np}{\pi(NV)^{\frac{3}{2}}}\int_{0}^{\infty}\int_{-\infty}^{\infty}\cos\left(\frac{2\pi z}{x}\cosh y\right)e\left(\frac{t_{j}y}{\pi}\right)e\left(\frac{z}{x}\right)W_{T}(\frac{x}{\sqrt{p}},\cdot)\frac{dy\ dx}{x^{2}}.

Changing variables, xC0pxx\rightarrow\frac{C_{0}\sqrt{p}}{x}, this becomes,

(5.44) +Φ(tj)=NpπC0(NV)320cos(2πzxC0pcoshy)e(tjyπ+zxC0p)WT(x,)𝑑y𝑑x.\mathcal{L}^{+}\Phi(t_{j})=\frac{N\sqrt{p}}{\pi C_{0}(NV)^{\frac{3}{2}}}\int_{0}^{\infty}\int_{-\infty}^{\infty}\cos\left(\frac{2\pi zx}{C_{0}\sqrt{p}}\cosh y\right)e\left(\frac{t_{j}y}{\pi}+\frac{zx}{C_{0}\sqrt{p}}\right)W_{T^{\prime}}(x,\cdot)dy\ dx.

Here, WTW_{T^{\prime}} is another pεp^{\varepsilon}-inert family, supported on x1x\asymp 1.

We can extend the xx-integral to all of \mathbb{R} as WT()W_{T^{\prime}}(\cdot) vanishes in the negative reals. Also, as cosu=12(eiu+eiu)\cos{u}=\frac{1}{2}\left(e^{iu}+e^{-iu}\right), we can use Fubini’s theorem once to get

(5.45) +Φ(tj)=NpπC0(NV)32e(tjyπ)WT^(zC0p(1±coshy))𝑑y.\mathcal{L}^{+}\Phi(t_{j})=\frac{N\sqrt{p}}{\pi C_{0}(NV)^{\frac{3}{2}}}\int_{-\infty}^{\infty}e\left(\frac{t_{j}y}{\pi}\right)\widehat{W_{T^{\prime}}}\left(-\frac{z}{C_{0}\sqrt{p}}(1\pm\cosh y)\right)dy.

Here,

(5.46) WT^(t)=e(xt)WT(x,)𝑑x.\widehat{W_{T^{\prime}}}(t)=\int_{-\infty}^{\infty}e\left(-xt\right)W_{T^{\prime}}(x,\cdot)dx.

Using Proposition 2.12, we can show that the the right hand side in (5.46) is small unless zC0p(1±coshy)pε\frac{z}{C_{0}\sqrt{p}}(1\pm\cosh{y})\ll p^{\varepsilon}.

So, for the integral to not be small, the sign should be -.
As, 1coshy=(y22!+y44!+)1-\cosh{y}=-(\frac{y^{2}}{2!}+\frac{y^{4}}{4!}+\dots), we should have

(5.47) y(C0pz)12pε(C0pNN0)12pε(1NV)12pε.y\ll\left(\frac{C_{0}\sqrt{p}}{z}\right)^{\frac{1}{2}}p^{\varepsilon}\asymp\left(\frac{C_{0}p}{\sqrt{NN_{0}}}\right)^{\frac{1}{2}}p^{\varepsilon}\asymp\left(\frac{1}{NV}\right)^{\frac{1}{2}}p^{\varepsilon}.

We can use a Taylor series expansion of WT^()\widehat{W_{T^{\prime}}}(\cdot) , with the leading term being WT^(zC0py22)\widehat{W_{T^{\prime}}}\left(-\frac{z}{C_{0}\sqrt{p}}\frac{y^{2}}{2}\right), to rewrite (5.45) as,

(5.48) +Φ(tj)=NpπC0(NV)32e(tjyπ)WT^(zC0py22,)𝑑y+O(pA).\mathcal{L}^{+}\Phi(t_{j})=\frac{N\sqrt{p}}{\pi C_{0}(NV)^{\frac{3}{2}}}\int_{-\infty}^{\infty}e\left(\frac{t_{j}y}{\pi}\right)\widehat{W_{T^{\prime}}}\left(\frac{z}{C_{0}\sqrt{p}}\frac{y^{2}}{2},\cdot\right)dy\ +O(p^{-A}).

We define

(5.49) Qz2C0pNN0C0pNV.Q\coloneqq\frac{z}{2C_{0}\sqrt{p}}\asymp\frac{\sqrt{NN_{0}}}{C_{0}p}\asymp NV.

Using a change of variables, yQyy\rightarrow\sqrt{Q}y in (5.48), we get

(5.50) +Φ(tj)=NpπC0(NV)32Qe(tjyπQ)WT^(y2,)𝑑y.\mathcal{L}^{+}\Phi(t_{j})=\frac{N\sqrt{p}}{\pi C_{0}(NV)^{\frac{3}{2}}\sqrt{Q}}\int_{-\infty}^{\infty}e\left(\frac{t_{j}y}{\pi\sqrt{Q}}\right)\widehat{W_{T^{\prime}}}\left(y^{2},\cdot\right)dy.

Now, we can use Proposition 2.12 once again on (5.50), to get

(5.51) +Φ(tj)=NpπC0(NV)32QG(tjQ,),\mathcal{L}^{+}\Phi(t_{j})=\frac{N\sqrt{p}}{\pi C_{0}(NV)^{\frac{3}{2}}\sqrt{Q}}G\left(\frac{t_{j}}{\sqrt{Q}},\cdot\right),

where G(t,l,m,n)G(t,l,m,n) is a satisfies the same derivative bounds as a pεp^{\varepsilon}-inert family in the first variable. It is pεp^{\varepsilon}-inert in the other variables (l,m,nl,m,n). Also,

(5.52) G(tjQ,) is small unless |tj|Qpε(NV)12pε.\text{$G\left(\frac{t_{j}}{\sqrt{Q}},\cdot\right)$ is small unless }\lvert t_{j}\rvert\ll\sqrt{Q}p^{\varepsilon}\asymp(NV)^{\frac{1}{2}}p^{\varepsilon}.

Mellin Transform of G(t,)G(t,\cdot):

Similar to (5.18), we take a Mellin transform of G()G(\cdot) with respect to the variables, l,m,nl,m,n, and use Mellin inversion to rewrite (5.51) as

(5.53) +Φ(tj)=NpπC0(NV)2(u1)=σ1(u2)=σ2(u3)=σ3lu1mu2nu3G~(tj,u1,u2,u3)𝑑u1𝑑u2𝑑u3,\mathcal{L}^{+}\Phi(t_{j})=\frac{N\sqrt{p}}{\pi C_{0}(NV)^{2}}\int_{\begin{subarray}{c}\Re(u_{1})=\sigma_{1}\\ \Re(u_{2})=\sigma_{2}\\ \Re(u_{3})=\sigma_{3}\end{subarray}}l^{-u_{1}}m^{-u_{2}}n^{-u_{3}}\widetilde{G}(t_{j},u_{1},u_{2},u_{3})\ du_{1}\ du_{2}\ du_{3},

where

(5.54) G~(t,u1,u2,u3)=000G(t,l,m,n)lu11mu21nu31𝑑l𝑑m𝑑n,\widetilde{G}(t,u_{1},u_{2},u_{3})=\int_{0}^{\infty}\int_{0}^{\infty}\int_{0}^{\infty}G(t,l,m,n)l^{u_{1}-1}m^{u_{2}-1}n^{u_{3}-1}\ dl\ dm\ dn,

and σj>0\sigma_{j}>0, for j=1,2,3.j=1,2,3.

We note here that, G~(t,)\widetilde{G}(t,\cdot) is small unless |tj|Qpε(NV)12pε\lvert t_{j}\rvert\ll\sqrt{Q}p^{\varepsilon}\asymp(NV)^{\frac{1}{2}}p^{\varepsilon}, because of a similar bound on G(t,)G(t,\cdot). Also, similar to (5.17), we have that, for arbitrarily large A>0A>0,

(5.55) G~(t,u1,u2,u3)pεL0σ1M0σ2N0σ3j=13(1+|uj|)A.\widetilde{G}(t,u_{1},u_{2},u_{3})\ll p^{\varepsilon}L_{0}^{\sigma_{1}}M_{0}^{\sigma_{2}}N_{0}^{\sigma_{3}}\prod_{j=1}^{3}(1+\lvert u_{j}\rvert)^{-A}.

Final Steps:

Using (5.53), we can rewrite (5.6) (up to a small error term) as,

(5.56) Maass0=1ϕ(p)p2NpπC0(NV)2χ(p)tjπitj(p,χ2)(uj)=σjτ(χ)3χ(aα)4πϵπV(p)π(1)G~(tj,u1,u2,u3)(lgcd(l,p)=1λ¯(pl)χ(l)lu1)(m,ngcd(n,p)=1χ(n)λf(m)λf¯(n)λ¯π(p2mn)mu2nu3)du1du2du3.\mathcal{M}_{\text{Maass}_{0}}=\frac{1}{\phi(p)p^{2}}\frac{N\sqrt{p}}{\pi C_{0}(NV)^{2}}\sideset{}{{}^{*}}{\sum}_{\chi(p)}\sum_{t_{j}}\sum_{\pi\in\mathcal{H}_{it_{j}}(p,\chi^{2})}\int_{\Re(u_{j})=\sigma_{j}}\frac{\tau(\chi)^{3}\chi(a_{\alpha})4\pi\epsilon_{\pi}}{V(p)\mathscr{L}_{\pi}^{*}(1)}\widetilde{G}(t_{j},u_{1},u_{2},u_{3})\\ \cdot\left(\sum_{\begin{subarray}{c}l\\ \text{gcd}(l,p)=1\end{subarray}}\frac{\overline{\lambda}(pl)\chi(l)}{l^{u_{1}}}\right)\left(\sum_{\begin{subarray}{c}m,n\\ \text{gcd}(n,p)=1\end{subarray}}\frac{\chi(-n)\lambda_{f}(m)\overline{\lambda_{f}}(n)\overline{\lambda}_{\pi}(p^{2}m-n)}{m^{u_{2}}n^{u_{3}}}\right)\ du_{1}\ du_{2}\ du_{3}.

We can again combine the ll terms to get (similar to (5.21)),

(5.57) lχ(l)λ¯π(pl)lu1=λ¯π(p)lχ(l)λ¯π(l)lu1=λ¯π(p)L(π¯χ,u1).\sum_{l}\frac{\chi(l)\overline{\lambda}_{\pi}(pl)}{l^{u_{1}}}=\overline{\lambda}_{\pi}(p)\sum_{l}\frac{\chi(l)\overline{\lambda}_{\pi}(l)}{l^{u_{1}}}=\overline{\lambda}_{\pi}(p){L}(\overline{\pi}\otimes\chi,u_{1}).

We bound (5.56) by taking the absolute value of the integrand. Note that as G~()\widetilde{G}(\cdot) decays rapidly along vertical lines (using (5.55)), we can restrict the integral up to |(uj)|pε\lvert\Im(u_{j})\rvert\ll p^{\varepsilon}. Using (5.55) gives us (for arbitrarily large A>0A>0),

(5.58) Maass0L0σ1M0σ2N0σ3ϕ(p)p2NpπC0(NV)2χ(p)tjπitj(p,χ2)|τ(χ)3χ(aα)4πϵπλ¯π(p)V(p)π(1)|(uj)=σj|m,ngcd(n,p)=1χ(n)λf(m)λf¯(n)λ¯π(p2mn)mu2nu3||L(π¯χ,u1)|j=13(1+|uj|)A𝑑u1𝑑u2𝑑u3.\mathcal{M}_{\text{Maass}_{0}}\ll\frac{L_{0}^{\sigma_{1}}M_{0}^{\sigma_{2}}N_{0}^{\sigma_{3}}}{\phi(p)p^{2}}\frac{N\sqrt{p}}{\pi C_{0}(NV)^{2}}\sideset{}{{}^{*}}{\sum}_{\chi(p)}\sum_{t_{j}}\sum_{\pi\in\mathcal{H}_{it_{j}}(p,\chi^{2})}\left|\frac{\tau(\chi)^{3}\chi(a_{\alpha})4\pi\epsilon_{\pi}\overline{\lambda}_{\pi}(p)}{V(p)\mathscr{L}_{\pi}^{*}(1)}\right|\\ \int_{\Re(u_{j})=\sigma_{j}}\left|\sum_{\begin{subarray}{c}m,n\\ \text{gcd}(n,p)=1\end{subarray}}\frac{\chi(-n)\lambda_{f}(m)\overline{\lambda_{f}}(n)\overline{\lambda}_{\pi}(p^{2}m-n)}{m^{u_{2}}n^{u_{3}}}\right|\frac{\lvert{L}(\overline{\pi}\otimes\chi,u_{1})\rvert}{\prod_{j=1}^{3}(1+\lvert u_{j}\rvert)^{A}}\ du_{1}\ du_{2}\ du_{3}.

We can use (5.52) to restrict (up to a small error) to when |tj|Qpε\lvert t_{j}\rvert\ll\sqrt{Q}p^{\varepsilon}. Also, as πitj(p,χ2)\pi\in\mathcal{H}_{it_{j}}(p,\chi^{2}), |λπ(p)|1\lvert\lambda_{\pi}(p)\rvert\leq 1 (pp divides the level). Additionally we have the bounds |τ(χ)|=p,|V(p)|p\lvert\tau(\chi)\rvert=\sqrt{p},\ \lvert V(p)\rvert\asymp p. These, along with (5.57), and the fact that χ(n)=χ(p2mn)\chi(-n)=\chi(p^{2}m-n) give us,

(5.59) Maass0L0σ1M0σ2N0σ3Npp32πC0(NV)2p3p𝒮.\mathcal{M}_{\text{Maass}_{0}}\ll\frac{L_{0}^{\sigma_{1}}M_{0}^{\sigma_{2}}N_{0}^{\sigma_{3}}\cdot N\sqrt{p}\cdot p^{\frac{3}{2}}}{\pi C_{0}(NV)^{2}\cdot p^{3}\cdot p}\cdot\mathcal{S^{\prime}}.

with

𝒮=pεtjQpεχ(p)πitj(p,χ2)(uj)=σj(uj)pε|L(π¯χ,u1)||mM0λf(m)mu2nN0λ¯f(n)λ¯π(p2mn)χ(p2mn)nu3|du1du2du3.\mathcal{S^{\prime}}=p^{\varepsilon}\cdot\sum_{t_{j}\ll\sqrt{Q}p^{\varepsilon}}\sideset{}{{}^{*}}{\sum}_{\chi(p)}\sum_{\pi\in\mathcal{H}_{it_{j}}(p,\chi^{2})}\int_{\begin{subarray}{c}\Re(u_{j})=\sigma_{j}\\ \Im(u_{j})\ll p^{\varepsilon}\end{subarray}}\left|L(\overline{\pi}\otimes\chi,u_{1})\right|\\ \cdot\left|\sum_{m\asymp M_{0}}\frac{\lambda_{f}(m)}{m^{u_{2}}}\sum_{n\asymp N_{0}}\frac{\overline{\lambda}_{f}(n)\overline{\lambda}_{\pi}(p^{2}m-n)\chi(p^{2}m-n)}{n^{u_{3}}}\right|\ du_{1}\ du_{2}\ du_{3}.

Once again, we can use Cauchy-Schwarz inequality on 𝒮\mathcal{S^{\prime}} to get

(5.60) 𝒮(𝒮1)12(𝒮2)12,\mathcal{S^{\prime}}\leq(\mathcal{S^{\prime}}_{1})^{\frac{1}{2}}\cdot(\mathcal{S^{\prime}}_{2})^{\frac{1}{2}},

where

(5.61) 𝒮1=pεtjQpεχ(p)πitj(p,χ2)(uj)=σj(uj)pε|L(π¯χ,u1)|2𝑑u1𝑑u2𝑑u3,\mathcal{S^{\prime}}_{1}=p^{\varepsilon}\cdot\sum_{t_{j}\ll\sqrt{Q}p^{\varepsilon}}\sideset{}{{}^{*}}{\sum}_{\chi(p)}\sum_{\pi\in\mathcal{H}_{it_{j}}(p,\chi^{2})}\int_{\begin{subarray}{c}\Re(u_{j})=\sigma_{j}\\ \Im(u_{j})\ll p^{\varepsilon}\end{subarray}}\left|L(\overline{\pi}\otimes\chi,u_{1})\right|^{2}\ du_{1}\ du_{2}\ du_{3},

and

(5.62) 𝒮2=pεtjQpεχ(p)πitj(p,χ2)(uj)=σj(uj)pε|mM0λf(m)mu2nN0λ¯f(n)λ¯π(p2mn)χ(p2mn)nu3|2du1du2du3.\mathcal{S^{\prime}}_{2}=p^{\varepsilon}\cdot\sum_{t_{j}\ll\sqrt{Q}p^{\varepsilon}}\sideset{}{{}^{*}}{\sum}_{\chi(p)}\sum_{\pi\in\mathcal{H}_{it_{j}}(p,\chi^{2})}\\ \cdot\int_{\begin{subarray}{c}\Re(u_{j})=\sigma_{j}\\ \Im(u_{j})\ll p^{\varepsilon}\end{subarray}}\left|\sum_{m\asymp M_{0}}\frac{\lambda_{f}(m)}{m^{u_{2}}}\sum_{n\asymp N_{0}}\frac{\overline{\lambda}_{f}(n)\overline{\lambda}_{\pi}(p^{2}m-n)\chi(p^{2}m-n)}{n^{u_{3}}}\right|^{2}\ du_{1}\ du_{2}\ du_{3}.

We can now take σ1=σ3=12\sigma_{1}=\sigma_{3}=\frac{1}{2}, and σ2=ε\sigma_{2}=\varepsilon, and similar to (5.28) and (5.36), use the spectral large sieve inequality to bound 𝒮1\mathcal{S^{\prime}}_{1} and 𝒮2\mathcal{S^{\prime}}_{2}. The analysis is actually simpler, especially for 𝒮2\mathcal{S^{\prime}}_{2}, with the only difference being tjQpεt_{j}\ll\sqrt{Q}p^{\varepsilon} here. The analogous bounds then become,

(5.63) 𝒮1(Qp2+ε+p1+ε)1+εnp1+ε1npεQp2+ε,\mathcal{S^{\prime}}_{1}\ll\left(Qp^{2+\varepsilon}+p^{1+\varepsilon}\right)^{1+\varepsilon}\cdot\sum_{n\leq p^{1+\varepsilon}}\frac{1}{n}\cdot p^{\varepsilon}\ll Qp^{2+\varepsilon},

and

(5.64) 𝒮2(Qp2+ε+p2+ε)1+εmM0np2+ε|λf(n)|2nQp2+ε.\mathcal{S^{\prime}}_{2}\ll\left(Qp^{2+\varepsilon}+p^{2+\varepsilon}\right)^{1+\varepsilon}\sum_{m\asymp M_{0}}\sum_{n\ll p^{2+\varepsilon}}\frac{\lvert\lambda_{f}(n)\rvert^{2}}{n}\ll Qp^{2+\varepsilon}.

Using (5.60), (5.63), and (5.64) in (5.59) we get (with σ1=σ3=12,σ2=ε\sigma_{1}=\sigma_{3}=\frac{1}{2},\sigma_{2}=\varepsilon,)

(5.65) Maass0L012M0εN012Npp32πC0(NV)2p3pQp2+ε.\mathcal{M}_{\text{Maass}_{0}}\ll\frac{L_{0}^{\tfrac{1}{2}}M_{0}^{\varepsilon}N_{0}^{\tfrac{1}{2}}\cdot N\sqrt{p}\cdot p^{\frac{3}{2}}}{\pi C_{0}(NV)^{2}\cdot p^{3}\cdot p}\cdot Qp^{2+\varepsilon}.

Using the fact that QNN0pC0NVQ\asymp\frac{\sqrt{NN_{0}}}{pC_{0}}\asymp NV, and L0Np2,L_{0}\leq\frac{N}{p^{2}}, and N0p2+εN_{0}\ll p^{2+\varepsilon}, taking absolute values in (5.65)\eqref{eq:M0Maass3} we get that,

(5.66) Maass0Npε.\mathcal{M}_{\text{Maass}_{0}}\ll Np^{\varepsilon}.

5.1.3. Maass1\mathcal{M}_{\text{Maass}_{1}} and Maass2\mathcal{M}_{\text{Maass}_{2}}

We briefly go over the process for bounding Maass1\mathcal{M}_{\text{Maass}_{1}} and Maass2\mathcal{M}_{\text{Maass}_{2}}. Notice that, in this case χ=χ0\chi=\chi_{0} is fixed, it is the unique quadratic character modulo pp.

So compared to (5.6), the corresponding expressions for Maass1\mathcal{M}_{\text{Maass}_{1}} and Maass2\mathcal{M}_{\text{Maass}_{2}} do not have a sum over the characters χ\chi modulo pp; and the Maass forms are level 11, instead of pp. Both of these facts lead to additional cancellations.

However, we can no longer use the bound |λπ(p)|1\lvert\lambda_{\pi}(p)\rvert\leq 1. Nevertheless the trivial bound |λπ(p)|p\lvert\lambda_{\pi}(p)\rvert\leq\sqrt{p} is sufficient for our result.

For Maass1\mathcal{M}_{\text{Maass}_{1}}, similar to (5.6), we have the expression,

(5.67) Maass1=l,m,ngcd(ln,p)=1λf(m)λf¯(n)ϕ(p)p2τ(χ0)3χ0(naαl)tj±Φ(tj)πitj(1,1)4πϵπV(p)π(1)λ¯π(p2mn)λ¯π(pl).\mathcal{M}_{\text{Maass}_{1}}=\sum_{\begin{subarray}{c}l,m,n\\ \text{gcd}(ln,p)=1\end{subarray}}\frac{\lambda_{f}(m)\overline{\lambda_{f}}(n)}{\phi(p)p^{2}}\tau(\chi_{0})^{3}\chi_{0}(-na_{\alpha}l)\\ \cdot\sum_{t_{j}}\mathcal{L}^{\pm}\Phi(t_{j})\sum_{\pi\in\mathcal{H}_{it_{j}}(1,1)}\frac{4\pi\epsilon_{\pi}}{V(p)\mathscr{L}_{\pi}^{*}(1)}\overline{\lambda}_{\pi}(p^{2}m-n)\overline{\lambda}_{\pi}(pl).

We can now repeat the same steps as in Maass0\mathcal{M}_{\text{Maass}_{0}} case and in fact get an even better bound of O(Np12+ε)O(Np^{-\frac{1}{2}+\varepsilon}). (We save a factor of p2p^{2} in the corresponding expressions for 𝒮1\mathcal{S}_{1} and 𝒮1\mathcal{S}_{1}^{\prime})

For Maass2\mathcal{M}_{\text{Maass}_{2}}, we have the expression,

(5.68) Maass2=lL0,mM0,gcd(cln,p)=1λf(m)λf¯(n)ϕ(p)p2τ(χ)3χ(naαl)tj±Φ(tj)πitj(1,1)4πϵπV(p)π(1)λ¯π(p)(p2mn)λ¯π(p)(pl),\mathcal{M}_{\text{Maass}_{2}}=\sum_{\begin{subarray}{c}l\asymp L_{0},m\asymp M_{0},\cdots\\ \text{gcd}(cln,p)=1\end{subarray}}\frac{\lambda_{f}(m)\overline{\lambda_{f}}(n)}{\phi(p)p^{2}}\tau(\chi)^{3}\chi(-na_{\alpha}l)\\ \cdot\sum_{t_{j}}\mathcal{L}^{\pm}\Phi(t_{j})\sum_{\pi\in\mathcal{H}_{it_{j}}(1,1)}\frac{4\pi\epsilon_{\pi}}{V(p)\mathscr{L}_{\pi}^{*}(1)}\overline{\lambda}^{(p)}_{\pi}(p^{2}m-n)\overline{\lambda}^{(p)}_{\pi}(pl),

where (recalling from (2.52))

(5.69) λπ(p)(m)=xp(1)λπ(m)+pxp(p)λπ(mp).\lambda_{\pi}^{(p)}(m)=x_{p}(1)\lambda_{\pi}(m)+\sqrt{p}\ x_{p}(p)\lambda_{\pi}\left(\frac{m}{p}\right).

Here, λπ(mp)=0\lambda_{\pi}(\frac{m}{p})=0 if pmp\nmid m.

Notice that, as gcd(n,p)=1(n,p)=1, p(np2m)p\nmid(n-p^{2}m). Also, xp()pεx_{p}(\cdot)\ll p^{\varepsilon}.

Comparing with (5.67), we see that there’s an additional p\sqrt{p} factor. However, the cancellation we get from not having the χ\chi-sum and the level of the Maass forms dropping to 11 is more than enough to take care of this. We can once again proceed as before and show that Maass2Npε\mathcal{M}_{\text{Maass}_{2}}\ll Np^{\varepsilon}.

Remark 5.4.

We could use non trivial bounds for |λπ(p)|\lvert\lambda_{\pi}(p)\rvert while analysing Maass1\mathcal{M}_{\text{Maass}_{1}} and Maass2\mathcal{M}_{\text{Maass}_{2}}. In that case we would get Maass1=O(Np1+θ+ε)\mathcal{M}_{\text{Maass}_{1}}=O\left(Np^{-1+\theta+\varepsilon}\right), and Maass2=O(Np12+θ+ε)\mathcal{M}_{\text{Maass}_{2}}=O\left(Np^{-\frac{1}{2}+\theta+\varepsilon}\right). Here, θ\theta is current progress towards the Ramanujan Peterson conjecture, so for instance, θ764\theta\leq\frac{7}{64}.

5.2. Holomorphic Term Analysis

For the non-oscillatory case in hol\mathcal{M}_{\text{hol}}, we can very easily prove a variant of Lemma 5.1 (with tt replaced by kk), by choosing

(5.70) hhol(s,k)=2s1πΓ(s+k12)Γ(ks+12),h_{\text{hol}}(s,k)=\frac{2^{s-1}}{\pi}\frac{\Gamma\left(\frac{s+k-1}{2}\right)}{\Gamma\left(\frac{k-s+1}{2}\right)},

and

(5.71) Hhol(s,k,l,m,n)=12πihhol(s,t)Φ~(s+1,)(4π)s.H_{\text{hol}}(s,k,l,m,n)=\frac{1}{2\pi i}h_{\text{hol}}(s,t)\widetilde{\Phi}(s+1,\cdot)(4\pi)^{-s}.

The rest of the steps are identical to the corresponding Maass form case.

For the oscillatory case, instead of (5.42), we use

(5.72) Jl1(x)=±ei(l1)π2π0π2cos((l1)θ)e±ixcosθ𝑑θ.J_{l-1}(x)=\sum_{\pm}\frac{e^{\mp i(l-1)\frac{\pi}{2}}}{\pi}\int_{0}^{\frac{\pi}{2}}\cos\left((l-1)\theta\right)e^{\pm ix\cos\theta}\,d\theta.

We then proceed similarly, the only difference being, we use the power series expansion of cos(y)\cos(y) instead of coshy\cosh{y} in (5.48)\eqref{eq:L0+ver4}.

5.3. Eisenstein Term Analysis

The kernel function for Eis\mathcal{M}_{\text{Eis}} is the same as Maass\mathcal{M}_{\text{Maass}}, and hence, the analysis is pretty similar, except for replacing tj()\sum_{t_{j}}(\cdot) with ()𝑑t\int_{-\infty}^{\infty}(\cdot)dt. Similar to (5.2), the corresponding Eisenstein expression is,

(5.73) 𝒦Eis=14π±Φ(t)r1r2=pπit,Eis(r2,χ2)4πϵπV(p)π(1)δr1λ¯π(δ)(|np2m|)λ¯π(δ)(|pl|)dt\mathcal{K}_{\text{Eis}}=\frac{1}{4\pi}\int_{-\infty}^{\infty}\mathcal{L}^{\pm}\Phi(t)\sum_{r_{1}r_{2}=p}\sum_{\pi\in\mathcal{H}_{it,Eis}(r_{2},\chi^{2})}\frac{4\pi\epsilon_{\pi}}{V(p)\mathscr{L}_{\pi}^{*}(1)}\sum_{\delta\mid r_{1}}\overline{\lambda}^{(\delta)}_{\pi}(\lvert n-p^{2}m\rvert)\overline{\lambda}^{(\delta)}_{\pi}(\lvert-pl\rvert)dt

Here,

(5.74) it,Eis(r2,χ2)={Eχ1,χ2(z,12+it),χi(modqi), for i=1,2;q1q2=r2,χ1χ2¯χ2},\mathcal{H}_{it,Eis}(r_{2},\chi^{2})=\{E_{\chi_{1},\chi_{2}}\left(z,\frac{1}{2}+it\right),\chi_{i}\negthickspace\negthickspace\negthickspace\pmod{q_{i}},\text{ for }i=1,2;\ q_{1}q_{2}=r_{2},\ \chi_{1}\overline{\chi_{2}}\simeq\chi^{2}\},

where we use the notation ψχ\psi\simeq\chi to mean that ψ\psi and χ\chi share the same underlying primitive character.

Also, for π=Eχ1,χ2(z,12+it)\pi=E_{\chi_{1},\chi_{2}}\left(z,\frac{1}{2}+it\right), we have

(5.75) λπ(n)=λχ1,χ2,t(n)=χ2(sgn(n))ab=|n|χ1(a)χ¯2(b)aitbit.\lambda_{\pi}(n)=\lambda_{\chi_{1},\chi_{2},t}(n)=\chi_{2}(\text{sgn}(n))\sum_{ab=\lvert n\rvert}\frac{\chi_{1}(a)\overline{\chi}_{2}(b)}{a^{it}b^{-it}}.

As r1r2=pr_{1}r_{2}=p in (5.73), the only possibilities are r2=1r_{2}=1, or r2=pr_{2}=p. Again, r2=1r_{2}=1 only if χ\chi is quadratic (χ\chi cannot be trivial). Hence, the only possibilities for (χ1,χ2)(\chi_{1},\chi_{2}) are (1,1),(1,1), or (1,χ¯2)(1,\overline{\chi}^{2}), or (χ2,1)(\chi^{2},1).
Thus, it,Eis(1,1)={E1,1(z,12+it)}\mathcal{H}_{it,Eis}(1,1)=\{E_{1,1}\left(z,\frac{1}{2}+it\right)\}, and it,Eis(p,χ2)={E1,χ¯2(z,12+it),Eχ2,1(z,12+it)}\mathcal{H}_{it,Eis}(p,\chi^{2})=\{E_{1,\overline{\chi}^{2}}\left(z,\frac{1}{2}+it\right),E_{\chi^{2},1}\left(z,\frac{1}{2}+it\right)\}.

We can now proceed exactly similar to the Maass form case. The only difference we encounter is for the corresponding expressions for (5.21) (and (5.57) for the oscillatory case). Here, we get (for the non-oscillatory case),

(5.76) lχ(l)λπ¯(pl)lσ2+σ1=λπ¯(p)lχ(l)λπ¯(l)lσ2+σ1=λπ¯(p)abχ1¯(a)χ2(b)χ(ab)(ab)σ2+σ1aitbit=λ¯π(p)L(σ2+σ1it,χχ1¯)L(σ2+σ1+it,χχ2).\sum_{l}\frac{\chi(l)\overline{\lambda_{\pi}}(pl)}{l^{\frac{\sigma}{2}+\sigma_{1}}}=\overline{\lambda_{\pi}}(p)\sum_{l}\frac{\chi(l)\overline{\lambda_{\pi}}(l)}{l^{\frac{\sigma}{2}+\sigma_{1}}}=\overline{\lambda_{\pi}}(p)\sum_{a}\sum_{b}\frac{\overline{\chi_{1}}(a)\chi_{2}(b)\chi(ab)}{(ab)^{\frac{\sigma}{2}+\sigma_{1}}a^{-it}b^{it}}\\ =\overline{\lambda}_{\pi}(p){L}(\frac{\sigma}{2}+\sigma_{1}-it,\chi\cdot\overline{\chi_{1}}){L}(\frac{\sigma}{2}+\sigma_{1}+it,\chi\cdot\chi_{2}).

The last expression can be simplified further depending on if (χ1,χ2)=(1,1),(\chi_{1},\chi_{2})=(1,1), or (1,χ¯2)(1,\overline{\chi}^{2}), or (χ2,1)(\chi^{2},1). In each case, we get a product of Dirichlet LL-functions for the characters χ\chi and χ¯\overline{\chi}.

The rest of the analysis is the same, we just swap a degree 22 LL-function with a product of two degree 11 LL-functions.

6. Remaining Terms

We complete the proof of Theorem 1.7 by proving Lemma 4.2 (bounding the terms introduced by the delta symbol with cc divisible by pp), and Lemma 4.4 (bounding the smaller terms introduced after Voronoi summation). We start with the latter.

6.1. Proof of Lemma 4.4

Recall the term E1E_{1} was defined in (4.19). This can be rewritten as,

(6.1) E1=lNp2gcd(l,p)=1mpεnp2+εgcd(n,p2)=pcgcd(c,p)=1λf(m)λf¯(n)p2c2IN(c,l,m,n)a<cgcd(a,c)=1e(ap2lc)e(am¯c)e(ap2¯nc),E_{1}=\sum_{\begin{subarray}{c}l\leq\frac{N}{p^{2}}\\ \text{gcd}(l,p)=1\end{subarray}}\sum_{m\ll p^{\varepsilon}}\sum_{\begin{subarray}{c}n\ll p^{2+\varepsilon}\\ \text{gcd}(n,p^{2})=p\end{subarray}}\sum_{\begin{subarray}{c}c\\ \text{gcd}(c,p)=1\end{subarray}}\frac{\lambda_{f}(m)\overline{\lambda_{f}}(n)}{p^{2}c^{2}}I_{N}(c,l,m,n)\\ \cdot\ \sum_{\begin{subarray}{c}a<c\\ \text{gcd}(a,c)=1\end{subarray}}e\left(\frac{-ap^{2}l}{c}\right)e\left(\frac{-a\overline{m}}{c}\right)e\left(\frac{\overline{ap^{2}}n}{c}\right),

where

(6.2) IN(c,l,m,n)=gc(v)e(p2lv)w~c,v,N(m)w~pc,v,N(n)𝑑v.I_{N}(c,l,m,n)=\int_{-\infty}^{\infty}g_{c}(v)e(-p^{2}lv)\widetilde{w}_{c,v,N}(m)\widetilde{w}_{pc,-v,N}(n)dv.

Here, w~c,v,N()\widetilde{w}_{c,v,N}(\cdot) is given by (4.13). Similar to Lemma 6.1, we claim that, IN(c,l,m,n)NpεI_{N}(c,l,m,n)\ll Np^{\varepsilon}.

The aa-sum at the end is actually a Kloosterman sum,

(6.3) a<cgcd(a,c)=1e(ap2lc)e(am¯c)e(ap2¯nc)=S(np¯2m,p2l;c)=S(p¯(np2m),pl;c).\sum_{\begin{subarray}{c}a<c\\ \text{gcd}(a,c)=1\end{subarray}}e\left(\frac{-ap^{2}l}{c}\right)e\left(\frac{-a\overline{m}}{c}\right)e\left(\frac{\overline{ap^{2}}n}{c}\right)=S(n\overline{p}^{2}-m,-p^{2}l;c)=S(\overline{p}(n-p^{2}m),-pl;c).

Using the Weil bound, it follows that

(6.4) E1Np2p1p2NpεcNc32N34+ε.E_{1}\ll\frac{N}{p^{2}}\cdot p\cdot\frac{1}{p^{2}}\cdot Np^{\varepsilon}\cdot\sum_{c\leq\sqrt{N}}c^{-\frac{3}{2}}\ll N^{\frac{3}{4}+\varepsilon}.

The proofs for the corresponding bounds for E2E_{2} and E3E_{3} are very similar. Following the same steps, we can, in fact, get an even better bound of O(N34p1+ε)O(N^{\frac{3}{4}}p^{-1+\varepsilon}).

6.2. Proof of Lemma 4.2

We use Voronoi summation twice to prove Lemma 4.2.
Using Proposition 2.16 on T1(a,c,v)T_{1}(a,c,v), we have

(6.5) T1(a,c,v)=1cm1λf(m)e(a¯mc)w~c,v,N(m).T_{1}(a,c,v)=\frac{1}{c}\sum_{m\geq 1}\lambda_{f}(m)e\left(-\frac{\overline{a}m}{c}\right)\widetilde{w}_{c,v,N}(m).

Here,

w~c,v,N(m)=2πikoJk1(4πmxc)e(xv)wN(x)𝑑x.\widetilde{w}_{c,v,N}(m)=2\pi i^{k}\int_{o}^{\infty}J_{k-1}\left(\frac{4\pi\sqrt{mx}}{c}\right)e(xv)w_{N}(x)dx.

Using an integration by parts argument similar to the one in Lemma 4.6, we can show that the sum on the right in (6.5) is effectively for mc2Nm\ll\frac{c^{2}}{N}. As cC=Nc\leq C=\sqrt{N}, this is non-trivial only when c2N1εc^{2}\geq N^{1-\varepsilon}, in which case 1mpε1\leq m\ll p^{\varepsilon}.

T2(a,c,v)T_{2}(a,c,v) needs more attention. Since cc is divisible by pp, e(aαln¯p)e(anc)e\left(\frac{a_{\alpha}l\overline{n}}{p}\right)e\left(\frac{-an}{c}\right) is periodic modulo cc (whenever nn is coprime to pp), and hence can be written as a finite sum of additive characters. Using this and taking r=aαlr=a_{\alpha}l, we get

(6.6) T2(a,c,v)=Nn2N gcd(n,p)=1λf¯(n)wN(n)e(nv)1ct(modc)e(ntc)u(modc)gcd(u,p)=1e(ru¯p)e(au+tuc)=1ct(modc)u(modc)gcd(u,p)=1e(ru¯p)e(au+tuc)Nn2N gcd(n,p)=1λf¯(n)e(ntc)wN(n)e(nv).T_{2}(a,c,v)=\sum_{\begin{subarray}{c}N\leq n\leq 2N\\ \text{ gcd}(n,p)=1\end{subarray}}\overline{\lambda_{f}}(n)w_{N}(n)e(-nv)\frac{1}{c}\sum_{t\negthickspace\negthickspace\negthickspace\pmod{c}}e\left(\frac{nt}{c}\right)\sum_{\begin{subarray}{c}u\negthickspace\negthickspace\negthickspace\pmod{c}\\ \text{gcd}(u,p)=1\end{subarray}}e\left(\frac{r\overline{u}}{p}\right)e\left(-\frac{au+tu}{c}\right)\\ =\frac{1}{c}\sum_{t\negthickspace\negthickspace\negthickspace\pmod{c}}\sum_{\begin{subarray}{c}u\negthickspace\negthickspace\negthickspace\pmod{c}\\ \text{gcd}(u,p)=1\end{subarray}}e\left(\frac{r\overline{u}}{p}\right)e\left(-\frac{au+tu}{c}\right)\sum_{\begin{subarray}{c}N\leq n\leq 2N\\ \text{ gcd}(n,p)=1\end{subarray}}\overline{\lambda_{f}}(n)e\left(\frac{nt}{c}\right)w_{N}(n)e(-nv).

We want to use Voronoi summation for the last sum.Once again, because of an integration by parts argument, the length of the summation, post Voronoi summation, will effectively be up to n1N(cgcd(c,t))2n\ll\frac{1}{N}\left(\frac{c}{\text{gcd}(c,t)}\right)^{2}. As cNc\leq\sqrt{N}, this will have non-trivial contribution only when gcd(c,t)pε\text{gcd}(c,t)\ll p^{\varepsilon}, and c2N1εc^{2}\geq N^{1-\varepsilon}. We can in fact show that we can restrict to just gcd(c,t)=1\text{gcd}(c,t)=1 case.

Suppose gcd(c,t)=g>1\text{gcd}(c,t)=g>1. As gpε,pcgg\ll p^{\varepsilon},\ p\mid\frac{c}{g}. Now, we can write the contribution of these terms in (6.6) is

(6.7) T2(a,c,v)=1cgc1<gpεSa,c,v(g),T_{2}^{\prime}(a,c,v)=\frac{1}{c}\sum_{\begin{subarray}{c}g\mid c\\ 1<g\ll p^{\varepsilon}\end{subarray}}S_{a,c,v}(g),

where

(6.8) Sa,c,v(g)=t0(modc)gcd(t,c)=gNn2N gcd(n,p)=1λf¯(n)e(ntc)wN(n)e(nv)u(modc)gcd(u,p)=1e(ru¯p)e(au+tuc).S_{a,c,v}(g)=\sum_{\begin{subarray}{c}t\equiv 0\negthickspace\negthickspace\negthickspace\pmod{c}\\ \text{gcd}(t,c)=g\end{subarray}}\sum_{\begin{subarray}{c}N\leq n\leq 2N\\ \text{ gcd}(n,p)=1\end{subarray}}\overline{\lambda_{f}}(n)e\left(\frac{nt}{c}\right)w_{N}(n)e(-nv)\sum_{\begin{subarray}{c}u\negthickspace\negthickspace\negthickspace\pmod{c}\\ \text{gcd}(u,p)=1\end{subarray}}e\left(\frac{r\overline{u}}{p}\right)e\left(-\frac{au+tu}{c}\right).

Consider the uu-sum in (6.8). If we change variables with uu+cgu\rightarrow u+\frac{c}{g} in the uu-sum in (6.8) (this is well defined as pcgp\mid\frac{c}{g}), we notice that

(6.9) u(modc)gcd(u,p)=1e(ru¯p)e(au+tuc)=e(ag)u(modc)gcd(u,p)=1e(ru¯p)e(au+tuc).\sum_{\begin{subarray}{c}u\negthickspace\negthickspace\negthickspace\pmod{c}\\ \text{gcd}(u,p)=1\end{subarray}}e\left(\frac{r\overline{u}}{p}\right)e\left(-\frac{au+tu}{c}\right)=e\left(\frac{a}{g}\right)\sum_{\begin{subarray}{c}u\negthickspace\negthickspace\negthickspace\pmod{c}\\ \text{gcd}(u,p)=1\end{subarray}}e\left(\frac{r\overline{u}}{p}\right)e\left(-\frac{au+tu}{c}\right).

As gcd(a,c)=1(a,c)=1, and gcg\mid c, e(ag)1e\left(-\frac{a}{g}\right)\neq 1. So, the uu-sum in (6.8) must be zero. Thus, T2(a,c,v)=0T_{2}^{\prime}(a,c,v)=0, and it suffices to only consider the case when gcd(c,t)=1(c,t)=1.

Using Proposition 2.16 for that case, we have (up to a small error),

(6.10) T2(a,c,v)=1ct(modc)gcd(t,c)=1u(modc)gcd(u,p)=1e(ru¯p)e(au+tuc)1cnpελf¯(n)e(nt¯c)w~c,v,N(n).T_{2}(a,c,v)=\frac{1}{c}\sum_{\begin{subarray}{c}t\negthickspace\negthickspace\negthickspace\pmod{c}\\ \text{gcd}(t,c)=1\end{subarray}}\sum_{\begin{subarray}{c}u\negthickspace\negthickspace\negthickspace\pmod{c}\\ \text{gcd}(u,p)=1\end{subarray}}e\left(\frac{r\overline{u}}{p}\right)e\left(-\frac{au+tu}{c}\right)\frac{1}{c}\sum_{n\ll p^{\varepsilon}}\overline{\lambda_{f}}(n)e\left(-\frac{n\overline{t}}{c}\right)\widetilde{w}_{c,-v,N}(n).

So, using (6.5) and (6.10) in (4.9), we get that (up to a small error),

(6.11) S4(N,α)=lN/p2gcd(l,p)=1mpεnpεcgcd(c,p)=pλf(m)λf¯(n)c3I~N(c,l,m,n)B(c,l,m,n),S_{4}(N,\alpha)=\sum_{\begin{subarray}{c}l\leq N/p^{2}\\ \text{gcd}(l,p)=1\end{subarray}}\sum_{m\ll p^{\varepsilon}}\sum_{n\ll p^{\varepsilon}}\sum_{\begin{subarray}{c}c\\ \text{gcd}(c,p)=p\end{subarray}}\frac{\lambda_{f}(m)\overline{\lambda_{f}}(n)}{c^{3}}\tilde{I}_{N}(c,l,m,n)B(c,l,m,n),

where

(6.12) I~N(c,l,m,n)=gc(v)e(p2lv)w~c,v,N(m)w~c,v,N(n)𝑑v,\tilde{I}_{N}(c,l,m,n)=\int_{-\infty}^{\infty}g_{c}(v)e(-p^{2}lv)\widetilde{w}_{c,v,N}(m)\widetilde{w}_{c,-v,N}(n)dv,

and

(6.13) B(c,l,m,n)=a(modc)gcd(a,c)=1e(ap2lc)e(a¯mc)t(modc)gcd(t,c)=1u(modc)gcd(u,p)=1e(ru¯p)e(au+tuc)e(nt¯c).B(c,l,m,n)=\sum_{\begin{subarray}{c}a\negthickspace\negthickspace\negthickspace\pmod{c}\\ \text{gcd}(a,c)=1\end{subarray}}e\left(\frac{-ap^{2}l}{c}\right)e\left(-\frac{\overline{a}m}{c}\right)\sum_{\begin{subarray}{c}t\negthickspace\negthickspace\negthickspace\pmod{c}\\ \text{gcd}(t,c)=1\end{subarray}}\sum_{\begin{subarray}{c}u\negthickspace\negthickspace\negthickspace\pmod{c}\\ \text{gcd}(u,p)=1\end{subarray}}e\left(\frac{r\overline{u}}{p}\right)e\left(-\frac{au+tu}{c}\right)e\left(-\frac{n\overline{t}}{c}\right).

Let c=kpc=k\cdot p, and as cNp32+εc\leq\sqrt{N}\ll p^{\frac{3}{2}+\varepsilon}, kp12+εk\ll p^{\frac{1}{2}+\varepsilon} and gcd(k,p)=1(k,p)=1.
We can then rewrite (6.11) (up to small error) as

(6.14) S4(N,α)=lN/p2gcd(l,p)=1m,nm,npεkp12+ελf(m)λf¯(n)k3p3I~N(kp,l,m,n)B(kp,l,m,n).S_{4}(N,\alpha)=\sum_{\begin{subarray}{c}l\leq N/p^{2}\\ \text{gcd}(l,p)=1\end{subarray}}\sum_{\begin{subarray}{c}m,n\\ m,n\ll p^{\varepsilon}\end{subarray}}\sum_{k\ll p^{\frac{1}{2}+\varepsilon}}\frac{\lambda_{f}(m)\overline{\lambda_{f}}(n)}{k^{3}p^{3}}\tilde{I}_{N}(kp,l,m,n)B(kp,l,m,n).

We complete the proof by stating the following two lemmas.

Lemma 6.1.

Let ε>0\varepsilon>0 and let Np3+εN\ll p^{3+\varepsilon}. Let lNp2l\leq\frac{N}{p^{2}} with (l,p)=1(l,p)=1, and m,npεm,n\ll p^{\varepsilon}. We define I~N(c,l,m,n)\tilde{I}_{N}(c,l,m,n) as in (6.12). Now, if in addition cc satisfies N1εc2NN^{1-\varepsilon}\ll c^{2}\leq N, then

(6.15) I~N(c,l,m,n)εNpε.\tilde{I}_{N}(c,l,m,n)\ll_{\varepsilon}Np^{\varepsilon}.
Lemma 6.2.

Let ε,N,l,m,n\varepsilon,N,l,m,n be same as before, in Lemma 6.1. We define B(c,l,m,n)B(c,l,m,n) as in (6.13). Now, if in addition, c=kpc=k\cdot p, with (k,p)=1(k,p)=1, then

(6.16) B(kp,l,m,n)kkp2.B(kp,l,m,n)\ll k\sqrt{k}\ p^{2}.

We quickly show how Lemmas 6.1 and 6.2 imply Lemma 4.2. Using (6.15) and (6.16) in (6.14), we get (using trivial bounds)

(6.17) S4(N,α)lN/p2gcd(l,p)=1m,nm,npεkp12+ε|λf(m)λf¯(n)|k3p3Nk32p2+εN2p314+εNp14+ε.S_{4}(N,\alpha)\ll\sum_{\begin{subarray}{c}l\leq N/p^{2}\\ \text{gcd}(l,p)=1\end{subarray}}\sum_{\begin{subarray}{c}m,n\\ m,n\ll p^{\varepsilon}\end{subarray}}\sum_{k\ll p^{\frac{1}{2}+\varepsilon}}\frac{\lvert\lambda_{f}(m)\overline{\lambda_{f}}(n)\rvert}{k^{3}p^{3}}Nk^{\frac{3}{2}}p^{2+\varepsilon}\ll N^{2}p^{-3-\frac{1}{4}+\varepsilon}\ll Np^{-\frac{1}{4}+\varepsilon}.

Clearly, (6.17) implies (4.11).

6.2.1. Proof of Lemma 6.1

Using Proposition 2.6, we can restrict the integral in (6.12), up to a small error, to |v|1cC\lvert v\rvert\ll\frac{1}{cC}. Here, C=NC=\sqrt{N}. Using the trivial bounds w~c,v,N(m)N\widetilde{w}_{c,v,N}(m)\ll N, we get

(6.18) I~N(c,l,m,n)N2+ε1cC.\tilde{I}_{N}(c,l,m,n)\ll N^{2+\varepsilon}\frac{1}{cC}.

We have already noted that c2N1εc^{2}\geq N^{1-\varepsilon}. Using this, Lemma 6.1 follows.

6.2.2. Proof of Lemma 6.2

Using the Chinese Remainder theorem, we can rewrite B(kp,l,m,n)B(kp,l,m,n) as

(6.19) B(kp,l,m,n)=BkBpB(kp,l,m,n)=B_{k}\cdot B_{p}

where

(6.20) Bk=a1(modk)gcd(a1,k)=1e(a1plk)e(a1¯mp¯k)t1(modk)gcd(t1,k)=1e(nt1¯p¯k)u1(modk)e(a1u1p¯k)e(t1u1p¯k),B_{k}=\sum_{\begin{subarray}{c}a_{1}\negthickspace\negthickspace\negthickspace\pmod{k}\\ \text{gcd}(a_{1},k)=1\end{subarray}}e\left(\frac{-a_{1}pl}{k}\right)e\left(-\frac{\overline{a_{1}}m\overline{p}}{k}\right)\sum_{\begin{subarray}{c}t_{1}\negthickspace\negthickspace\negthickspace\pmod{k}\\ \text{gcd}(t_{1},k)=1\end{subarray}}e\left(-\frac{n\overline{t_{1}}\overline{p}}{k}\right)\sum_{u_{1}\negthickspace\negthickspace\negthickspace\pmod{k}}e\left(-\frac{a_{1}u_{1}\overline{p}}{k}\right)e\left(-\frac{t_{1}u_{1}\overline{p}}{k}\right),

and

(6.21) Bp=u2(modp)gcd(u2,p)=1e(ru2¯p)a2(modp)gcd(a2,p)=1e(a2¯mk¯p)e(a2u2k¯p)t2(modp)gcd(t2,p)=1e(nt2¯k¯p)e(t2u2k¯p).B_{p}=\sum_{\begin{subarray}{c}u_{2}\negthickspace\negthickspace\negthickspace\pmod{p}\\ \text{gcd}(u_{2},p)=1\end{subarray}}e\left(\frac{r\overline{u_{2}}}{p}\right)\sum_{\begin{subarray}{c}a_{2}\negthickspace\negthickspace\negthickspace\pmod{p}\\ \text{gcd}(a_{2},p)=1\end{subarray}}e\left(-\frac{\overline{a_{2}}m\overline{k}}{p}\right)e\left(-\frac{a_{2}u_{2}\overline{k}}{p}\right)\sum_{\begin{subarray}{c}t_{2}\negthickspace\negthickspace\negthickspace\pmod{p}\\ \text{gcd}(t_{2},p)=1\end{subarray}}e\left(-\frac{n\overline{t_{2}}\overline{k}}{p}\right)e\left(-\frac{t_{2}u_{2}\overline{k}}{p}\right).

Consider BkB_{k} first. Using orthogonality of characters modulo kk for the u1u_{1} sum in (6.20), and rearranging terms, we get

(6.22) Bk=ka1(modk)gcd(a1,k)=1e(a1plk)e(a1¯np¯k)e(a1¯mp¯k).B_{k}=k\sum_{\begin{subarray}{c}a_{1}\negthickspace\negthickspace\negthickspace\pmod{k}\\ \text{gcd}(a_{1},k)=1\end{subarray}}e\left(\frac{-a_{1}pl}{k}\right)e\left(\frac{\overline{a_{1}}n\overline{p}}{k}\right)e\left(-\frac{\overline{a_{1}}m\overline{p}}{k}\right).

We note that the a1a_{1}-sum is in fact a Kloosterman sum.

(6.23) Bk=kS(pl,(nm)p¯;k)=kS(l,(nm);k).B_{k}=k\cdot S(-pl,(n-m)\overline{p};k)=k\cdot S(-l,(n-m);k).

The Weil bound immediately gives,

(6.24) Bkkk.B_{k}\ll k\sqrt{k}.

Consider BpB_{p} now. The t2t_{2} and the u2u_{2} sums in (6.21) both define Kloosterman sums. This gives us,

(6.25) Bp=u2(modp)gcd(u2,p)=1e(ru2¯p)S(u2k¯,mk¯;p)S(u2k¯,nk¯;p).B_{p}=\sum_{\begin{subarray}{c}u_{2}\negthickspace\negthickspace\negthickspace\pmod{p}\\ \text{gcd}(u_{2},p)=1\end{subarray}}e\left(\frac{r\overline{u_{2}}}{p}\right)S(u_{2}\overline{k},m\overline{k};p)S(u_{2}\overline{k},n\overline{k};p).

Trivially bounding the u2u_{2} sum, and using the Weil bound twice gives us,

(6.26) Bpp2.B_{p}\ll p^{2}.

Now, (6.24) and (6.26) imply (6.16). This completes the proof of Lemma 6.2.

7. Application to Non-vanishing

We use the upper bound obtained in Theorem 1.4 to prove Theorem 1.5. We first state a lemma regarding the first moment for the family of LL-functions we have been working with so far.

Lemma 7.1.

Let p,q,f,αp,q,f,\alpha be as before. Then for ε>0\varepsilon>0,

(7.1) ψ(modp2)L(12,f(αψ))=p2+O(p74+ε).\sum_{\psi\negthickspace\negthickspace\negthickspace\pmod{p^{2}}}L\left(\tfrac{1}{2},f\otimes\left(\alpha\cdot\psi\right)\right)=p^{2}+O(p^{\frac{7}{4}+\varepsilon}).

Assume Lemma 7.1 for now. Using, Cauchy-Schwarz inequality, we have

|ψ(modp2)L(12,f(αψ))|2=|ψ(modp2)L(12,f(αψ))δ(L(12,f(αψ))0)|2ψ(modp2)|L(12,f(αψ))|2ψ(modp2)|δ(L(12,f(αψ))0)|2=ψ(modp2)|L(12,f(αψ))|2ψ(modp2)L(12,f(αψ))01.\left|\sum_{\psi\negthickspace\negthickspace\negthickspace\pmod{p^{2}}}L\left(\tfrac{1}{2},f\otimes\left(\alpha\cdot\psi\right)\right)\right|^{2}=\left|\sum_{\psi\negthickspace\negthickspace\negthickspace\pmod{p^{2}}}L\left(\tfrac{1}{2},f\otimes\left(\alpha\cdot\psi\right)\right)\cdot\delta\left(L\left(\tfrac{1}{2},f\otimes\left(\alpha\cdot\psi\right)\right)\neq 0\right)\right|^{2}\\ \leq\sum_{\psi\negthickspace\negthickspace\negthickspace\pmod{p^{2}}}\left|L\left(\tfrac{1}{2},f\otimes\left(\alpha\cdot\psi\right)\right)\right|^{2}\cdot\sum_{\psi\negthickspace\negthickspace\negthickspace\pmod{p^{2}}}\left|\delta\left(L\left(\tfrac{1}{2},f\otimes\left(\alpha\cdot\psi\right)\right)\neq 0\right)\right|^{2}\\ =\sum_{\psi\negthickspace\negthickspace\negthickspace\pmod{p^{2}}}\left|L\left(\tfrac{1}{2},f\otimes\left(\alpha\cdot\psi\right)\right)\right|^{2}\cdot\sum_{\begin{subarray}{c}\psi\negthickspace\negthickspace\negthickspace\pmod{p^{2}}\\ L\left(\tfrac{1}{2},f\otimes\left(\alpha\cdot\psi\right)\right)\neq 0\end{subarray}}1.

This implies,

(7.2) #{ψ(modp2);L(12,f(αψ))0}=ψ(modp2)L(12,f(αψ))01|ψ(modp2)L(12,f(αψ))|2ψ(modp2)|L(12,f(αψ))|2.\#\{\psi\negthickspace\negthickspace\negthickspace\pmod{p^{2}};\ L\left(\tfrac{1}{2},f\otimes\left(\alpha\cdot\psi\right)\right)\neq 0\}=\sum_{\begin{subarray}{c}\psi\negthickspace\negthickspace\negthickspace\pmod{p^{2}}\\ L\left(\tfrac{1}{2},f\otimes\left(\alpha\cdot\psi\right)\right)\neq 0\end{subarray}}1\geq\ \ \frac{\left|\sum_{\psi\negthickspace\negthickspace\negthickspace\pmod{p^{2}}}L\left(\tfrac{1}{2},f\otimes\left(\alpha\cdot\psi\right)\right)\right|^{2}}{\sum_{\psi\negthickspace\negthickspace\negthickspace\pmod{p^{2}}}\left|L\left(\tfrac{1}{2},f\otimes\left(\alpha\cdot\psi\right)\right)\right|^{2}}.

Theorem 1.5 now follows using the corresponding bounds in Lemma 7.1 and Theorem 1.4 in (7.2).

7.1. Proof of Lemma 7.1

Using the approximate functional equation in Proposition 2.1, we have

(7.3) ψ(modp2)L(12,f(αψ))=A+B,\sum_{\psi\negthickspace\negthickspace\negthickspace\pmod{p^{2}}}L\left(\tfrac{1}{2},f\otimes\left(\alpha\cdot\psi\right)\right)=A+B,

with

(7.4) A=ψ(modp2)n(Xq)1+ελf(n)α(n)ψ(n)nV(nXq),A=\sum_{\psi\negthickspace\negthickspace\negthickspace\pmod{p^{2}}}\sum_{n\ll(Xq)^{1+\varepsilon}}\frac{\lambda_{f}(n)\alpha(n)\psi(n)}{\sqrt{n}}V\left(\frac{n}{Xq}\right),

and

(7.5) B=ψ(modp2)ε(f(αψ),12)nqXqελf¯(n)α¯(n)ψ¯(n)nV(nXq).B=\sum_{\psi\negthickspace\negthickspace\negthickspace\pmod{p^{2}}}\varepsilon\left(f\otimes(\alpha\cdot\psi),\tfrac{1}{2}\right)\sum_{n\ll\frac{q}{X}\cdot q^{\varepsilon}}\frac{\overline{\lambda_{f}}(n)\overline{\alpha}(n)\overline{\psi}(n)}{\sqrt{n}}V\left(\frac{nX}{q}\right).

Here, V(y)V(y) is a smooth function satisfying V(y)f,A(1+y𝔮)AV(y)\ll_{f,A}(1+\tfrac{y}{\sqrt{\mathfrak{q}_{\infty}}})^{-A}, for any A>0A>0.

7.1.1. Bounds on A

Interchanging the order of summation in AA, we get, via the orthogonality of characters (and separating the diagonal term),

(7.6) A=p2V(1Xq)+p21lXqp2λf(1+p2l)α(1+p2l)ψ(1+p2l)1+p2lV(1+p2lXq).A=p^{2}\cdot V\left(\frac{1}{Xq}\right)+p^{2}\cdot\sum_{1\leq l\ll\frac{Xq}{p^{2}}}\frac{\lambda_{f}(1+p^{2}l)\alpha(1+p^{2}l)\psi(1+p^{2}l)}{\sqrt{1+p^{2}l}}V\left(\frac{1+p^{2}l}{Xq}\right).

Using

1lXqp2λf(1+p2l)α(1+p2l)ψ(1+p2l)1+p2lV(1+p2lXq)1p1lXqp21l(Xq)12p2,\sum_{1\leq l\ll\frac{Xq}{p^{2}}}\frac{\lambda_{f}(1+p^{2}l)\alpha(1+p^{2}l)\psi(1+p^{2}l)}{\sqrt{1+p^{2}l}}V\left(\frac{1+p^{2}l}{Xq}\right)\ll\frac{1}{p}\sum_{1\leq l\ll\frac{Xq}{p^{2}}}\frac{1}{\sqrt{l}}\ll\frac{(Xq)^{\frac{1}{2}}}{p^{2}},

we can get

(7.7) A=p2+O(X12q12)=p2+O(X12p32).A=p^{2}+O(X^{\frac{1}{2}}q^{\frac{1}{2}})=p^{2}+O(X^{\frac{1}{2}}p^{\frac{3}{2}}).

7.1.2. Bounds on B

Here, we prove that B=O(p2X12)B=O(p^{2}X^{-\tfrac{1}{2}}). Along with (7.6), this proves Lemma 7.1 once we set X=p12X=p^{\tfrac{1}{2}}.

The root number of L(f(αψ))L\left({f\otimes\left(\alpha\cdot\psi\right)}\right) is given by,

(7.8) ε(f(αψ),12)=τ(ψα)p3.\varepsilon\left(f\otimes(\alpha\cdot\psi),\tfrac{1}{2}\right)=\frac{\tau(\psi\cdot\alpha)}{p^{3}}.

Using this on (7.5), we get that

(7.9) B=ψ(modp2)τ(ψα)p3nqXλf¯(n)α¯(n)ψ¯(n)nV(nXq).B=\sum_{\psi\negthickspace\negthickspace\negthickspace\pmod{p^{2}}}\frac{\tau(\psi\cdot\alpha)}{p^{3}}\sideset{}{{}^{*}}{\sum}_{n\ll\frac{q}{X}}\frac{\overline{\lambda_{f}}(n)\overline{\alpha}(n)\overline{\psi}(n)}{\sqrt{n}}V\left(\frac{nX}{q}\right).

Expanding out the Gauss sum, and using orthogonality for the ψ\psi-sum, we get

(7.10) B=1p31a,bp3nqXλf¯(n)nV(nXq)e(a+bq)α(abn¯)p2δ(abn(modp2)).B=\frac{1}{p^{3}}\sideset{}{{}^{*}}{\sum}_{1\leq a,b\leq p^{3}}\sideset{}{{}^{*}}{\sum}_{n\ll\frac{q}{X}}\frac{\overline{\lambda_{f}}(n)}{\sqrt{n}}V\left(\frac{nX}{q}\right)e\left(\frac{a+b}{q}\right)\alpha(ab\overline{n})p^{2}\cdot\delta(ab\equiv n\negthickspace\negthickspace\negthickspace\pmod{p^{2}}).

Here, n¯n1(modp3)\overline{n}\cdot n\equiv 1(\mathrm{mod}\ p^{3}).
The δ(abn(modp2))\delta(ab\equiv n(\mathrm{mod}\ p^{2})) condition implies l, 1lp\exists\ l,\ 1\leq l\leq p with b=a¯n(1+p2l)b=\overline{a}n(1+p^{2}l), where a¯a1(modp3)\overline{a}\cdot a\equiv 1(\mathrm{mod}\ p^{3}).

Using this,

(7.11) B=1pnqXλf¯(n)nV(nXq)1ap31lpe(a+a¯n(1+p2l)p3)α(1+p2l).B=\frac{1}{p}\sideset{}{{}^{*}}{\sum}_{n\ll\frac{q}{X}}\frac{\overline{\lambda_{f}}(n)}{\sqrt{n}}V\left(\frac{nX}{q}\right)\sideset{}{{}^{*}}{\sum}_{1\leq a\leq p^{3}}\sum_{1\leq l\leq p}e\left(\frac{a+\overline{a}n(1+p^{2}l)}{p^{3}}\right)\alpha(1+p^{2}l).

Notice that α(1+p2())\alpha(1+p^{2}(\cdot)) is an additive character mod pp. So, cα(modp)\exists\ c_{\alpha}(\mathrm{mod}\ p) such that α(1+p2l)=e(cαlp)\alpha(1+p^{2}l)=e\left(\dfrac{c_{\alpha}\cdot l}{p}\right).

Using this, and orthogonality in the ll-sum,

(7.12) B=1pnqXλf¯(n)nV(nXq)1ap3e(a+a¯np3)pδ(cαa¯n(modp)).B=\frac{1}{p}\sideset{}{{}^{*}}{\sum}_{n\ll\frac{q}{X}}\frac{\overline{\lambda_{f}}(n)}{\sqrt{n}}V\left(\frac{nX}{q}\right)\sideset{}{{}^{*}}{\sum}_{1\leq a\leq p^{3}}e\left(\frac{a+\overline{a}n}{p^{3}}\right)p\cdot\delta\left(c_{\alpha}\equiv-\overline{a}n\negthickspace\negthickspace\negthickspace\pmod{p}\right).

Again, the δ(cαa¯n(modp))\delta\left(c_{\alpha}\equiv-\overline{a}n(\mathrm{mod}\ p)\right) condition implies y0,y1, 1y0,y1p\exists\ y_{0},y_{1},\ 1\leq y_{0},y_{1}\leq p with
a=ncα¯(1+py0+p2y1)a=-n\overline{c_{\alpha}}(1+py_{0}+p^{2}y_{1}), where cα¯cα1(modp3)\overline{c_{\alpha}}\cdot c_{\alpha}\equiv 1(\mathrm{mod}\ p^{3}). So,

(7.13) B=nqXλf¯(n)nV(nXq)e(cα+ncα¯p3)1y0pe(cαpy02+cαncα¯p2)1y1pe(y1(cαncα¯)p).B=\sideset{}{{}^{*}}{\sum}_{n\ll\frac{q}{X}}\frac{\overline{\lambda_{f}}(n)}{\sqrt{n}}V\left(\frac{nX}{q}\right)e\left(-\frac{c_{\alpha}+n\overline{c_{\alpha}}}{p^{3}}\right)\sum_{1\leq y_{0}\leq p}e\left(\frac{-c_{\alpha}py_{0}^{2}+c_{\alpha}-n\overline{c_{\alpha}}}{p^{2}}\right)\sum_{1\leq y_{1}\leq p}e\left(\frac{y_{1}(c_{\alpha}-n\overline{c_{\alpha}})}{p}\right).

Orthogonality in the y1y_{1} sum implies that we have ncα2(modp)n\equiv c_{\alpha}^{2}(\mathrm{mod}\ p). So,

(7.14) B=pnqXncα2(modp)λf¯(n)nV(nXq)e(cα+ncα¯p3)R,B=p\cdot\sum_{\begin{subarray}{c}n\ll\frac{q}{X}\\ n\equiv c_{\alpha}^{2}\negthickspace\negthickspace\negthickspace\pmod{p}\end{subarray}}\frac{\overline{\lambda_{f}}(n)}{\sqrt{n}}V\left(\frac{nX}{q}\right)e\left(-\frac{c_{\alpha}+n\overline{c_{\alpha}}}{p^{3}}\right)\cdot R,

with

(7.15) R=1y0pe(cαy02+kαy0p),R=\sum_{1\leq y_{0}\leq p}e\left(\frac{-c_{\alpha}y_{0}^{2}+k_{\alpha}y_{0}}{p}\right),

where kα=cαncα¯pk_{\alpha}=\frac{c_{\alpha}-n\overline{c_{\alpha}}}{p}. Completing the square in (7.15), we have

(7.16) R=e(kα24cα)1y0pe(cα(y0+kα2cα)2p)=e(kα2p4cα)εpp(cαp).R=e\left(\frac{k\alpha^{2}}{4c_{\alpha}}\right)\sum_{1\leq y_{0}\leq p}e\left(\frac{c_{\alpha}(y_{0}+\frac{k_{\alpha}}{2c_{\alpha}})^{2}}{p}\right)=e\left(\frac{k\alpha^{2}}{p4c_{\alpha}}\right)\varepsilon_{p}\sqrt{p}\cdot\left(\frac{c_{\alpha}}{p}\right).

Here εp\varepsilon_{p} is a constant of absolute value 11.

Using (7.16) in (7.14), we have

(7.17) B=pnqXncα2(modp)λf¯(n)nV(nXq)e(kα24pcαcα+ncα¯p3)εpp(cαp).B=p\cdot\sum_{\begin{subarray}{c}n\ll\frac{q}{X}\\ n\equiv c_{\alpha}^{2}\negthickspace\negthickspace\negthickspace\pmod{p}\end{subarray}}\frac{\overline{\lambda_{f}}(n)}{\sqrt{n}}V\left(\frac{nX}{q}\right)e\left(\frac{k\alpha^{2}}{4pc_{\alpha}}-\frac{c_{\alpha}+n\overline{c_{\alpha}}}{p^{3}}\right)\cdot\varepsilon_{p}\sqrt{p}\cdot\left(\frac{c_{\alpha}}{p}\right).

Using trivial bounds, we finally get

(7.18) Bp12+ε(qX)12=O(p2X12).B\ll p^{\frac{1}{2}+\varepsilon}\left(\frac{q}{X}\right)^{\frac{1}{2}}=O(p^{2}X^{-\frac{1}{2}}).

Using (7.7) and (7.18) in (7.3), we get that

(7.19) ψ(modp2)L(12,f(αψ))=p2+O(X12p32)+O(p2X12).\sum_{\psi\negthickspace\negthickspace\negthickspace\pmod{p^{2}}}L\left(\tfrac{1}{2},f\otimes\left(\alpha\cdot\psi\right)\right)=p^{2}+O(X^{\frac{1}{2}}p^{\frac{3}{2}})+O(p^{2}X^{-\frac{1}{2}}).

Choosing X=p12X=p^{\frac{1}{2}} now completes the proof of Lemma 7.1.

References

  • [Goo82] Anton Good. The square mean of Dirichlet series associated with cusp forms. Mathematika, 29(2):278–295, 1982.
  • [IK04] Henryk Iwaniec and Emmanuel Kowalski. Analytic number theory, volume 53 of American Mathematical Society Colloquium Publications. American Mathematical Society, Providence, RI, 2004.
  • [Iwa80] Henryk Iwaniec. Fourier coefficients of cusp forms and the Riemann zeta-function. In Seminar on Number Theory, 1979–1980 (French), pages Exp. No. 18, 36. Univ. Bordeaux I, Talence, 1980.
  • [KMN16] Rizwanur Khan, Djordje Milićević, and Hieu T. Ngo. Non-vanishing of Dirichlet LL-functions in Galois orbits. Int. Math. Res. Not. IMRN, 2016(22):6955–6978, 2016.
  • [KPY19] Eren Mehmet Kiral, Ian Petrow, and Matthew P. Young. Oscillatory integrals with uniformity in parameters. J. Théor. Nombres Bordeaux, 31(1):145–159, 2019.
  • [KrY21] Eren Mehmet Kı ral and Matthew Young. The fifth moment of modular LL-functions. J. Eur. Math. Soc. (JEMS), 23(1):237–314, 2021.
  • [L0̈9] Guangshi Lü. Average behavior of Fourier coefficients of cusp forms. Proc. Amer. Math. Soc., 137(6):1961–1969, 2009.
  • [Li75] Wen Ch’ing Winnie Li. Newforms and functional equations. Math. Ann., 212:285–315, 1975.
  • [LL11] Yuk-Kam Lau and Guangshi Lü. Sums of Fourier coefficients of cusp forms. Q. J. Math., 62(3):687–716, 2011.
  • [MS83] Carlos J. Moreno and Freydoon Shahidi. The fourth moment of Ramanujan τ\tau-function. Math. Ann., 266(2):233–239, 1983.
  • [MS19] Ritabrata Munshi and Saurabh Kumar Singh. Weyl bound for pp-power twist of GL(2)\rm GL(2) LL-functions. Algebra Number Theory, 13(6):1395–1413, 2019.
  • [Mun14] Ritabrata Munshi. The circle method and bounds for LL-functions—I. Math. Ann., 358(1-2):389–401, 2014.
  • [MW21] Djordje Milićević and Daniel White. Twelfth moment of Dirichlet LL-functions to prime power moduli. Ann. Sc. Norm. Super. Pisa Cl. Sci. (5), 22(4):1879–1898, 2021.
  • [Nun21] Ramon M. Nunes. The twelfth moment of Dirichlet LL-functions with smooth moduli. Int. Math. Res. Not. IMRN, 2021(12):9180–9202, 2021.
  • [PY20] Ian Petrow and Matthew P. Young. The Weyl bound for Dirichlet LL-functions of cube-free conductor. Ann. of Math. (2), 192(2):437–486, 2020.
  • [PY23] Ian Petrow and Matthew P. Young. The fourth moment of Dirichlet L-functions along a coset and the Weyl bound. Duke Math. J., 172(10):1879–1960, 2023.