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

Sums of cubes and the Ratios Conjectures

Victor Y. Wang Department of Mathematics, Princeton University, Princeton, NJ, USA Courant Institute of Mathematical Sciences, New York University, New York, NY, USA [email protected]
Abstract.

Works of Hooley and Heath-Brown imply a near-optimal bound on the number NN of integral solutions to x13++x63=0x_{1}^{3}+\dots+x_{6}^{3}=0 in expanding regions, conditional on automorphy and GRH for certain Hasse–Weil LL-functions; for regions of diameter X1X\geq 1, the bound takes the form NC(ε)X3+εN\leq C(\varepsilon)X^{3+\varepsilon} (ε>0\varepsilon>0). We attribute the ε\varepsilon to several subtly interacting proof factors; we then remove the ε\varepsilon assuming some standard number-theoretic hypotheses, mainly featuring the Ratios and Square-free Sieve Conjectures. In fact, our softest hypotheses imply conjectures of Hooley and Manin on NN, and show that almost all integers a±4mod9a\not\equiv\pm 4\bmod{9} are sums of three cubes. Our fullest hypotheses are capable of proving power-saving asymptotics for NN, and producing almost all primes p±4mod9p\not\equiv\pm 4\bmod{9}.

Key words and phrases:
Cubic form, circle method, rational points, Hasse–Weil LL-functions, correlations
1991 Mathematics Subject Classification:
Primary 11D45; Secondary 11D25, 11G40, 11M50, 11P55

1. Introduction

Let F0=F0(x,y,z)\colonequalsx3+y3+z3F_{0}=F_{0}(x,y,z)\colonequals x^{3}+y^{3}+z^{3}. For each aa\in\mathbb{Z}, the cubic surface F0=aF_{0}=a has a fairly rich set of rational points [segre1943note]. On the other hand, Mordell has suggested that producing large, general integer solutions to F0=aF_{0}=a for a=3a=3 (or for any other fixed aa\in\mathbb{Z}) could be as hard as “finding when an assigned sequence, e.g. 123456789123456789, occurs in the decimal expansion of π\pi[mordell1953integer]*p. 505. The recent work [booker2021question] of Booker and Sutherland resolves Mordell’s specific question for a=3a=3, but the spirit of Mordell’s suggestion certainly remains.

Heath-Brown has conjectured that F0=aF_{0}=a should have infinitely many solutions (x,y,z)3(x,y,z)\in\mathbb{Z}^{3} for any fixed a±4mod9a\not\equiv\pm 4\bmod{9} (see [heath1992density]*p. 623). To represent all a±4mod9a\not\equiv\pm 4\bmod{9} even once, one must allow both positive and negative values of x,y,zx,y,z. The set F0(03)F_{0}(\mathbb{Z}_{\geq 0}^{3}) has upper density Γ(4/3)3/6=0.1186788\leq\Gamma(4/3)^{3}/6=0.1186788\ldots in 0\mathbb{Z}_{\geq 0} [davenport1939waring]; Deshouillers, Hennecart, and Landreau have given a model and evidence suggesting a precise density of 0.09994250.0999425\ldots [deshouillers2006density]. Wooley has shown, unconditionally, that F0(03)F_{0}(\mathbb{Z}_{\geq 0}^{3}) contains A0.91709477\gg A^{0.91709477} integers a[0,A]a\in[0,A] for reals A1A\geq 1 [wooley1995breaking, wooley2000sums, wooley2015sums]. We now recall a result of Hooley [hooley1986HasseWeil, hooley_greaves_harman_huxley_1997] and Heath-Brown [heath1998circle], and state our main result building on it; we then give further details and background.

Theorem (Hooley; Heath-Brown).

For certain Hasse–Weil LL-functions, assume automorphy and GRH. Then ϵA1ϵ\gg_{\epsilon}A^{1-\epsilon} integers a[0,A]a\in[0,A] lie in F0(03)F_{0}(\mathbb{Z}_{\geq 0}^{3}), for any ϵ>0\epsilon>0.

Theorem.

For certain Hasse–Weil LL-functions, assume automorphy, GRH, and the Ratios Conjectures. For a certain polynomial Δ\Delta, assume the Square-free Sieve Conjecture. Then A\gg A integers a[0,A]a\in[0,A] lie in F0(03)F_{0}(\mathbb{Z}_{\geq 0}^{3}), and 100%100\% of integers a±4mod9a\not\equiv\pm 4\bmod{9} lie in F0(3)F_{0}(\mathbb{Z}^{3}).

Both results require estimating sums that roughly take the following form:

(1.1) 𝒄6tdtpΔ(𝒄)(local L-factors)(real analysis)pΔ(𝒄)(geometry/𝔽p + analysis/p).\sum_{\bm{c}\in\mathbb{Z}^{6}}\int_{t\in\mathbb{R}}\frac{dt}{\prod_{p\nmid\Delta(\bm{c})}(\textnormal{local $L$-factors})}\cdot(\textnormal{real analysis})\cdot\prod_{p\mid\Delta(\bm{c})}(\textnormal{geometry/$\mathbb{F}_{p}$ + analysis/$\mathbb{Z}_{p}$}).

The utility of GRH in this context has been highlighted by Bombieri; see [bombieri2006riemann]*p. 111. Also, in a function-field setting, [glas2022question] has made [heath1998circle] unconditional, and one could likely simplify our present hypotheses accordingly (since GRH and the Square-free Sieve Conjecture are known over function fields). We only focus on \mathbb{Q} for practical reasons.

For a typical aa, the integer solutions to F0=aF_{0}=a are expected to be at least exponentially sparse, if they in fact exist. Heath-Brown’s conjecture would imply that the only obstructions to solubility for F0=aF_{0}=a, for any aa, are local. The naive local-to-global analog of Heath-Brown’s conjecture for 5x3+12y3+9z35x^{3}+12y^{3}+9z^{3} is known to fail (see e.g. [ghosh2017integral]*p. 691, footnote 3), due to Brauer–Manin obstructions that do not apply to F0F_{0} [colliot2012groupe]*p. 1304.

We restrict ourselves to a statistical analysis of F0=aF_{0}=a over aa\in\mathbb{Z}, for (x,y,z)3(x,y,z)\in\mathbb{Z}^{3} lying in carefully chosen regions, conducted using second moments and the variance framework of [ghosh2017integral, diaconu2019admissible, wang2022thesis, wang2023prime]. This connects naturally to difficult open questions in 66 variables, e.g. [hooley1986some]*Conjecture 2, which lie beyond the square-root barrier in the classical Hardy–Littlewood circle method. We will attack these questions under standard number-theoretic hypotheses, primarily regarding LL-function statistics of Random Matrix Theory (RMT) type. Our work opens with the delta method of [duke1993bounds, heath1996new] (a clean modern form of the Kloosterman method of [kloosterman1926representation], more precise than the upper-bound variant used in [hooley1986HasseWeil, hooley_greaves_harman_huxley_1997]), whose harmonic analysis in principle allows for cancellation over the difficult classical minor arcs. We prove three levels of results, under three levels of hypotheses (the first two levels being relatively soft and qualitative; see Conjectures 1.4 and 1.8). We first recall a general weighted version of Hooley’s conjecture.

Fix a cubic form F(𝒙)=F(x1,,xm)[x1,,xm]F(\bm{x})=F(x_{1},\dots,x_{m})\in\mathbb{Z}[x_{1},\dots,x_{m}] in m4m\geq 4 variables with nonzero discriminant. Let VV be the hypersurface F=0F=0 in m1\mathbb{P}^{m-1}_{\mathbb{Q}}. Let Υ\Upsilon be the set of m/2\lfloor m/2\rfloor-dimensional vector spaces LmL\subseteq\mathbb{Q}^{m} over \mathbb{Q} on which FF vanishes. On a first reading, we suggest assuming that m=6m=6 and FF is diagonal, though we will often work generally.

Given a real X1X\geq 1, and a function wCc(m)w\in C^{\infty}_{c}(\mathbb{R}^{m}), let

(1.2) NF,w(X)\colonequals𝒙m:F(𝒙)=0w(𝒙/X),NF(X)\colonequals𝒙[X,X]m:F(𝒙)=01;N_{F,w}(X)\colonequals\sum_{\bm{x}\in\mathbb{Z}^{m}:\,F(\bm{x})=0}w(\bm{x}/X),\quad N_{F}(X)\colonequals\sum_{\bm{x}\in[-X,X]^{m}:\,F(\bm{x})=0}1;

and if m=6m=6 (our main case of interest, in which m/2=3\lfloor m/2\rfloor=3), let

(1.3) EF,w(X)\colonequalsNF,w(X)𝔖Fσ,F,wX3LΥ𝒙L6w(𝒙/X),E_{F,w}(X)\colonequals N_{F,w}(X)-\mathfrak{S}_{F}\cdot\sigma_{\infty,F,w}\cdot X^{3}-\sum_{L\in\Upsilon}\sum_{\bm{x}\in L\cap\mathbb{Z}^{6}}w(\bm{x}/X),

where 𝔖FF1\mathfrak{S}_{F}\ll_{F}1 is the familiar singular series defined in [wang2023_isolating_special_solutions]*§6, and where

(1.4) σ,F,w\colonequalslimϵ0(2ϵ)1|F(𝒙)|ϵ𝑑𝒙w(𝒙)F,w1.\sigma_{\infty,F,w}\colonequals\lim_{\epsilon\to 0}{(2\epsilon)^{-1}\int_{\lvert F(\bm{x})\rvert\leq\epsilon}d\bm{x}\,w(\bm{x})}\ll_{F,w}1.

One could attribute to Hooley [hooley1986some]*Conjecture 2, Manin (see e.g. [franke1989rational]), Vaughan–Wooley [vaughan1995certain]*Appendix, Peyre [peyre1995hauteurs], et al. the following conjecture:

(1.5) limXX3EF,w(X)=0.\lim_{X\to\infty}X^{-3}E_{F,w}(X)=0.

(The original [hooley1986some]*Conjecture 2 for l=3l=3 would follow from (1.5) for F=x13++x63F=x_{1}^{3}+\dots+x_{6}^{3}, applied to a suitable sequence of weights ww.) See [ding2020variance] for another related problem.

Unconditionally, NF(X)ϵX7/2/(logX)5/2ϵN_{F}(X)\ll_{\epsilon}X^{7/2}/(\log{X})^{5/2-\epsilon} for X2X\geq 2 [vaughan2020some], when m=6m=6 and FF is diagonal. Under standard hypotheses on the varieties V𝒄m1V_{\bm{c}}\subseteq\mathbb{P}^{m-1}_{\mathbb{Q}} cut out by F(𝒙)=𝒄𝒙=0F(\bm{x})=\bm{c}\cdot\bm{x}=0, one can prove the near-optimal Theorem 1.1 for the same FF. (Here 𝒄𝒙\colonequals1imcixi\bm{c}\cdot\bm{x}\colonequals\sum_{1\leq i\leq m}c_{i}x_{i}.) Let Δ(𝒄)[𝒄]=[c1,,cm]\Delta(\bm{c})\in\mathbb{Z}[\bm{c}]=\mathbb{Z}[c_{1},\dots,c_{m}] be the discriminant polynomial defined in §2. Let

(1.6) 𝒮0\colonequals{𝒄m:Δ(𝒄)=0},𝒮1\colonequals{𝒄m:Δ(𝒄)0}.\mathcal{S}_{0}\colonequals\{\bm{c}\in\mathbb{Z}^{m}:\Delta(\bm{c})=0\},\quad\mathcal{S}_{1}\colonequals\{\bm{c}\in\mathbb{Z}^{m}:\Delta(\bm{c})\neq 0\}.

For each 𝒄𝒮1\bm{c}\in\mathcal{S}_{1}, one can package local data on V𝒄V_{\bm{c}} into a Hasse–Weil LL-function L(s,V𝒄)L(s,V_{\bm{c}}), defined in §3 along with the rest of the list (1.7).

Theorem 1.1 ([hooley1986HasseWeil, hooley_greaves_harman_huxley_1997]; [heath1998circle]).

Assume m=6m=6 and FF is diagonal. For each 𝐜𝒮1\bm{c}\in\mathcal{S}_{1}, assume Conjecture 1.2 for L(s,V𝐜)L(s,V_{\bm{c}}). Then NF(X)ϵX3+ϵN_{F}(X)\ll_{\epsilon}X^{3+\epsilon} for all ϵ>0\epsilon>0.

Conjecture 1.2 concerns automorphy and the Grand Riemann Hypothesis (GRH). Strictly speaking, Hooley and Heath-Brown assume (in their “Hypothesis HW”) GRH plus certain Selberg-type axioms, e.g. analyticity, but it is natural to assume automorphy in place of such axioms. A nice reference bridging these two perspectives is [farmer2019analytic].

The proof of Theorem 1.1 involves GRH on 1/L(s,V𝒄)1/L(s,V_{\bm{c}}), surprisingly, coupled with subtle algebro-geometric factors of a different nature. See §2 for details. The LL-function ingredients extend directly to general FF, but some other ingredients have yet to be generalized.

GRH gives a pointwise upper bound on 1/L(s,V𝒄)1/L(s,V_{\bm{c}}) for Re(s)>1/2\operatorname{Re}(s)>1/2, sufficient for Theorem 1.1. Going past the critical line (in mean value over 𝒄\bm{c}) turns out to require much new work, both with LL-functions and with other factors. In terms of LL-functions, we mainly use L(s,V𝒄)L(s,V_{\bm{c}}) for Re(s)1/2δ\operatorname{Re}(s)\geq 1/2-\delta, as well as L(s,V𝒄,2)L(s,V_{\bm{c}},\bigwedge^{2}), ζ(s)\zeta(s), and L(s,V)L(s,V) for Re(s)1δ\operatorname{Re}(s)\geq 1-\delta. It will be convenient to assume Conjecture 1.2 in full, but average versions might also suffice.

Conjecture 1.2 (HW2).

Let 𝐜𝒮1\bm{c}\in\mathcal{S}_{1}. Let L(s)L(s) be one of the Hasse–Weil LL-functions

(1.7) L(s,V𝒄),L(s,V𝒄,2),L(s,V𝒄,Sym2),L(s,V𝒄,2),ζ(s),L(s,V).L(s,V_{\bm{c}}),\;L(s,V_{\bm{c}},{\textstyle\bigotimes^{2}}),\;L(s,V_{\bm{c}},\operatorname{Sym}^{2}),\;L(s,V_{\bm{c}},{\textstyle\bigwedge^{2}}),\;\zeta(s),\;L(s,V).

Let Lv(s)L_{v}(s) be the local factors (including L(s)L_{\infty}(s), the gamma factor) associated to LL.

  1. (1)

    There exists an integer d1d\geq 1, and an isobaric automorphic representation Π\Pi of GLd(𝐀)\operatorname{GL}_{d}(\mathbf{A}_{\mathbb{Q}}), such that Lv(s)=Lv(s,Π)L_{v}(s)=L_{v}(s,\Pi) at all places vv\leq\infty.

  2. (2)

    L(s,Π)L(s,\Pi) has no zeros in the half-plane Re(s)>1/2\operatorname{Re}(s)>1/2.

For the background needed to interpret Conjecture 1.2 (HW2), see §3.

Theorem 1.3.

Suppose F=x13++x63F=x_{1}^{3}+\dots+x_{6}^{3}. Assume Conjectures 1.2, 1.4, and 1.5. Then

(1.8) NF(X)X3.N_{F}(X)\ll X^{3}.

Let S0S\subseteq\mathbb{Z}_{\geq 0}. If SS has positive lower density in 0\mathbb{Z}_{\geq 0}, then so does F0(S3)F_{0}(S^{3}).

We use mean-value RMT-type predictions derived from the Moments and Ratios Conjectures of [conrey2005integral, conrey2007applications, conrey2008autocorrelation]. (See also [diaconu2003multiple, vcech2022ratios], and references within, for another important perspective on such conjectures.) The Hasse–Weil LL-functions L(s,V𝒄)L(s,V_{\bm{c}}) form a geometric family, in the sense of [sarnak2016families]. For each 𝒄𝒮1\bm{c}\in\mathcal{S}_{1}, define the Dirichlet series

(1.9) Φ𝒄,1(s)\colonequalsζ(2s)1L(s+1/2,V)1L(s,V𝒄)1=1/ζ(2s)L(s+1/2,V)L(s,V𝒄).\Phi^{\bm{c},1}(s)\colonequals\zeta(2s)^{-1}L(s+1/2,V)^{-1}L(s,V_{\bm{c}})^{-1}=1/{\zeta(2s)L(s+1/2,V)L(s,V_{\bm{c}})}.

Note that ζ(2s)\zeta(2s) and L(s+1/2,V)L(s+1/2,V) are independent of 𝒄\bm{c}.

Conjecture 1.4 (R2’).

Suppose 2m2\mid m. Let f:f\colon\mathbb{C}\to\mathbb{C} be entire, with f(s)f,b(1+|Im(s)|)bf(s)\ll_{f,b}(1+\lvert\operatorname{Im}(s)\rvert)^{-b} on the strip 0Re(s)20\leq\operatorname{Re}(s)\leq 2 for all b1b\in\mathbb{Z}_{\geq 1}. Let Z,N1Z,N\in\mathbb{R}_{\geq 1} with NZ3N\leq Z^{3}. If σ0(1,2)\sigma_{0}\in(1,2), then

(1.10) 𝒄𝒮1[Z,Z]m|(σ0)𝑑sΦ𝒄,1(s)f(s)Ns|2FZmNsup0σ2𝑑t(1+|t|)2|f(σ+it)|2.\sum_{\bm{c}\in\mathcal{S}_{1}\cap[-Z,Z]^{m}}\left\lvert\int_{(\sigma_{0})}ds\,\Phi^{\bm{c},1}(s)\cdot f(s)N^{s}\right\rvert^{2}\ll_{F}Z^{m}N\sup_{0\leq\sigma\leq 2}\int_{\mathbb{R}}dt\,(1+\lvert t\rvert)^{2}\lvert f(\sigma+it)\rvert^{2}.

The contour (σ0)(\sigma_{0}) in (1.10) runs from s=σ0is=\sigma_{0}-i\infty to s=σ0+is=\sigma_{0}+i\infty. The left-hand side of (1.10) is independent of σ0(1,2)\sigma_{0}\in(1,2) unconditionally, or further under (HW2). Other versions of Conjecture 1.4 might also suffice for our purposes; for instance, an 1+δ\ell^{1+\delta} analog of (1.10) (with some nontrivial adjustments) might suffice, the precise norm of ff on the right-hand side of (1.10) is not very important, and one might not need to allow such general ff.

There are no log factors on the right-hand side of (1.10); the factor ζ(2s)1L(s+1/2,V)1\zeta(2s)^{-1}L(s+1/2,V)^{-1} in (1.9), and the integral over (σ0)(\sigma_{0}) in (1.10), play a mollifying role. The statement (R2’) can be derived from (HW2) and the Ratios Conjecture 6.3 (R2oo); see Proposition 6.8. However, there may well be another route to (R2’) not passing through (R2oo). The statement (R2’) essentially concerns cancellation in the coefficients of Φ𝒄,1(s)\Phi^{\bm{c},1}(s) over moduli nn in a dyadic range; see Proposition 6.10. A similar log-free cancellation statement, [li2022moments]*(1.3), has recently played a crucial role in another context. Furthermore, over function fields (or under (HW2) over \mathbb{Q}), it should already be possible to obtain partial results towards (R2’), using ideas of [soundararajan2009moments, harper2013sharp, bui2021ratios, florea2021negative, bui2023negative] (after Cauchy–Schwarz over ss).

For our main results, we also need the Square-free Sieve Conjecture (cf. [miller2004one]*p. 956 and [granville1998abc, poonen2003squarefree, bhargava2014geometric]) for the polynomial Δ\Delta, restricted to a certain range 1PZ3/21\leq P\leq Z^{3/2}. This hypothesis concerns “unlikely divisors” of the outputs of Δ\Delta. Such hypotheses can be made unconditional over function fields; see e.g. [poonen2003squarefree]*Lemma 7.1.

Conjecture 1.5 (SFSCp,3).

There exists η0=η0(Δ)>0\eta_{0}=\eta_{0}(\Delta)\in\mathbb{R}_{>0} such that if Z,P1Z,P\in\mathbb{R}_{\geq 1} and PZ3/2P\leq Z^{3/2}, then #{𝐜m[Z,Z]m:a prime p[P,2P) with p2Δ(𝐜)}ΔZmPη0\#\{\bm{c}\in\mathbb{Z}^{m}\cap[-Z,Z]^{m}:\exists\;\textnormal{a prime $p\in[P,2P)$ with $p^{2}\mid\Delta(\bm{c})$}\}\ll_{\Delta}Z^{m}P^{-\eta_{0}}.

The Square-free Sieve Conjecture (SFSCp,3) is used to confront some novel algebro-geometric issues in our work. We use it to prove, for diagonal FF, a geometric relative (Conjecture 9.8) of the automorphic Sarnak–Xue Density Hypothesis ([sarnak1991bounds]*Conjecture 1, concerning the extent to which a naive generalization of the Ramanujan Conjecture can fail). This involves sieve-theoretic ideas weighted by somewhat dangerous factors. Even though we do not currently see how to eliminate the use of (SFSCp,3) over \mathbb{Q}, it is fortunate to be able to reduce Conjecture 9.8 to (SFSCp,3) when FF is diagonal. One can also reduce (SFSCp,3) for diagonal FF to the case m=4m=4 (but not to m=2m=2, it seems); see Proposition 9.13.

After Theorem 1.3, our next main result is the following:

Theorem 1.6.

Suppose m=6m=6 and FF is diagonal. Assume Conjectures 1.2, 1.4, 1.5, and 1.8. Then (1.5) holds for all functions wCc(m)w\in C^{\infty}_{c}(\mathbb{R}^{m}) for which

(1.11) {𝒙m:w(𝒙)0}¯{𝒙m:x1xm0}.\overline{\{\bm{x}\in\mathbb{R}^{m}:w(\bm{x})\neq 0\}}\subseteq\{\bm{x}\in\mathbb{R}^{m}:x_{1}\cdots x_{m}\neq 0\}.

Therefore, the Hasse principle holds for VV. Furthermore, if F=x13++x63F=x_{1}^{3}+\dots+x_{6}^{3}, then 100%100\% of integers a±4mod9a\not\equiv\pm 4\bmod{9} lie in F0(3)F_{0}(\mathbb{Z}^{3}).

For another conditional approach to the Hasse principle for VV when FF is diagonal, see [swinnerton2001solubility]. Over function fields of characteristic 7\geq 7, the Hasse principle for VV is already known in general when m=6m=6 [tian2017hasse]. But our approach has quantitative advantages, which become qualitative when applied to sums of three cubes.

We expect that the condition (1.11) could be removed with enough work. When F=x13++x63F=x_{1}^{3}+\dots+x_{6}^{3}, it is in fact possible to do this for free: Theorem 1.6 has the following corollary. (A similar but messier statement is possible for arbitrary diagonal FF when m=6m=6.)

Corollary 1.7.

Suppose F=x13++x63F=x_{1}^{3}+\dots+x_{6}^{3}. Assume Conjectures 1.2, 1.4, 1.5, and 1.8. Then (1.5) holds for all wCc(m)w\in C^{\infty}_{c}(\mathbb{R}^{m}). Consequently, [hooley1986some]*Conjecture 2 for l=3l=3 holds.

Theorem 1.6 makes use of a first-moment estimate for the quantity (1.9) over 𝒮1\mathcal{S}_{1}, and over some mildly localized pieces of 𝒮1\mathcal{S}_{1} (“adelic perturbations” of 𝒮1\mathcal{S}_{1} restricted by a parameter MM). For 𝒃=(b1,,bm)m\bm{b}=(b_{1},\dots,b_{m})\in\mathbb{Z}^{m} and M>0M\in\mathbb{R}_{>0}, consider the box

(1.12) M(𝒃)\colonequals[b1/M,(b1+1)/M)××[bm/M,(bm+1)/M)m.\mathcal{B}_{M}(\bm{b})\colonequals[b_{1}/M,(b_{1}+1)/M)\times\dots\times[b_{m}/M,(b_{m}+1)/M)\subseteq\mathbb{R}^{m}.

For Z>0Z\in\mathbb{R}_{>0}, let ZM(𝒃)Z\cdot\mathcal{B}_{M}(\bm{b}) denote the dilate {Z𝒓:𝒓M(𝒃)}m\{Z\bm{r}:\bm{r}\in\mathcal{B}_{M}(\bm{b})\}\subseteq\mathbb{R}^{m}. For each Z2Z\in\mathbb{R}_{\geq 2}, let

(1.13) σ(Z)\colonequals1/2+1/logZ.\sigma(Z)\colonequals 1/2+1/\log{Z}.
Conjecture 1.8 (RA1oo).

Suppose 2m2\mid m, and assume Conjecture 1.2. Let M1M\in\mathbb{R}_{\geq 1}; let n0[1,M]n_{0}\in\mathbb{Z}\cap[1,M] and 𝐚,𝐛m[M,M]m\bm{a},\bm{b}\in\mathbb{Z}^{m}\cap[-M,M]^{m}. Let AF,1𝐚,n0(s)A_{F,1}^{\bm{a},n_{0}}(s) be defined as in §6.3.1 (in terms of FF, 𝐚\bm{a}, n0n_{0}). If Z2Z\in\mathbb{R}_{\geq 2} and t[logZ,logZ]t\in[-\log{Z},\log{Z}], then for s=σ(Z)+its=\sigma(Z)+it, we have

(1.14) 𝒄𝒮1ZM(𝒃):𝒄𝒂modn0Φ𝒄,1(s)=𝒄𝒮1ZM(𝒃):𝒄𝒂modn0(1+oF,M;Z(1))AF,1𝒂,n0(s).\sum_{\begin{subarray}{c}\bm{c}\in\mathcal{S}_{1}\cap Z\cdot\mathcal{B}_{M}(\bm{b}):\\ \bm{c}\equiv\bm{a}\bmod{n_{0}}\end{subarray}}\Phi^{\bm{c},1}(s)=\sum_{\begin{subarray}{c}\bm{c}\in\mathcal{S}_{1}\cap Z\cdot\mathcal{B}_{M}(\bm{b}):\\ \bm{c}\equiv\bm{a}\bmod{n_{0}}\end{subarray}}(1+o_{F,M;Z\to\infty}(1))\cdot A_{F,1}^{\bm{a},n_{0}}(s).

Here AF,1𝒂,n0(s)A_{F,1}^{\bm{a},n_{0}}(s) is a Dirichlet series with an Euler product, absolutely convergent on the half-plane Re(s)>1/3\operatorname{Re}(s)>1/3. The terms oF,M;Z(1)o_{F,M;Z\to\infty}(1) are required to tend to 0 as ZZ\to\infty (when FF, MM are fixed). See Proposition 6.13 for the main use of Conjecture 1.8.

The Ratios Conjectures include (RA1oo), even with a power saving; see §6.3.1 for details. The choice (1.13) is permissible according to [conrey2007applications]*(2.11b). The essential feature of (1.13) for us is that (σ(Z)12)logZ(\sigma(Z)-\frac{1}{2})\cdot\log{Z} is positive and independent of ZZ (but its precise constant value is not important). We could get away with a larger choice of σ(Z)\sigma(Z) if we assumed a correspondingly stronger error term in (1.14); but the present formulation of (RA1oo) is clean, and easy to compare with other literature.111Note that if s=σ(Z)s=\sigma(Z), then in (1.9), we have ζ(2s)logZ\zeta(2s)\asymp\log{Z} and L(s+1/2,V)(logZ)rFL(s+1/2,V)\asymp(\log{Z})^{r_{F}}, where rF0r_{F}\in\mathbb{Z}_{\geq 0} is given explicitly for diagonal FF by [wang2022thesis]*Lemma 8.6.7. The moments of 1/L1/L we consider (over our orthogonal family of LL-functions L(s,V𝒄L(s,V_{\bm{c}})) are thus analogous to central moments over symplectic families. The paper [florea2021negative] seems to come close to proving (RA1oo) for a different family of LL-functions, over a function field.

We believe (RA1oo), like (R2’), represents a tantalizing research direction. There is another direction worth mentioning. In light of the log-free square-root cancellation in 2\ell^{2} conjectured in (1.10), one may hope that “better than square-root cancellation” occurs in 2ϵ\ell^{2-\epsilon}, by analogy with [harper2023typical]*(1.2) (an attractive conjecture based on random multiplicative functions and multiplicative chaos; see e.g. [gorodetsky2021magic, harper2023typical] for details and references). If true, this would provide additional cancellation in Proposition 7.5 (one of the key ingredients for Theorem 1.3), and thus provide an alternative approach to Theorem 1.6 (but not to Theorem 1.9).

Our final main result, Theorem 1.9, goes beyond Theorem 1.6.

Theorem 1.9.

Suppose m=6m=6 and FF is diagonal. Assume Conjectures 1.2, 1.5 with η0\eta_{0}, 1.10 with η1\eta_{1}, and 1.11 with HH. Then there exists a real δ=δ(η0,η1,degH)>0\delta=\delta(\eta_{0},\eta_{1},\deg{H})>0 such that EF,w(X)F,wX3δE_{F,w}(X)\ll_{F,w}X^{3-\delta} holds for all functions wCc(m)w\in C^{\infty}_{c}(\mathbb{R}^{m}) satisfying (1.11).

Our methods would allow one to prove a version of Theorem 1.9 uniform over small archimedean and non-archimedean perturbations to EF,w(X)E_{F,w}(X). By [wang2023prime]*Theorem 1.2, one could then show under the hypotheses of Theorem 1.9 that if F=x13++x63F=x_{1}^{3}+\dots+x_{6}^{3}, then 100%100\% of primes p±4mod9p\not\equiv\pm 4\bmod{9} lie in F0(3)F_{0}(\mathbb{Z}^{3}). One would also be able to give a power-saving analog of Corollary 1.7. But to give full details would obscure our exposition.

The power saving in Theorem 1.9 is small and complicated. (Egregious “exponent divisions” occur in Lemma 7.13 and Proposition 9.5, due to our use of (4.3); and similarly in (10.8), due to (4.4).) It would be very interesting to understand the limits of what one can hope for.

Conjecture 1.10 (RA1δ\delta).

Suppose 2m2\mid m, and assume Conjecture 1.2. Let Z2Z\in\mathbb{R}_{\geq 2} and M[1,Zη1]M\in[1,Z^{\eta_{1}}]. Let n0[1,M]n_{0}\in\mathbb{Z}\cap[1,M] and 𝐚,𝐛m[M,M]m\bm{a},\bm{b}\in\mathbb{Z}^{m}\cap[-M,M]^{m}. There exists a real η1=η1(F)>0\eta_{1}=\eta_{1}(F)>0, depending only on FF, such that if t[M,M]t\in[-M,M] and s=σ(Z)+its=\sigma(Z)+it, then

(1.15) 𝒄𝒮1ZM(𝒃):𝒄𝒂modn0Φ𝒄,1(s)=𝒄𝒮1ZM(𝒃):𝒄𝒂modn0(1+OF(Zη1))AF,1𝒂,n0(s).\sum_{\begin{subarray}{c}\bm{c}\in\mathcal{S}_{1}\cap Z\cdot\mathcal{B}_{M}(\bm{b}):\\ \bm{c}\equiv\bm{a}\bmod{n_{0}}\end{subarray}}\Phi^{\bm{c},1}(s)=\sum_{\begin{subarray}{c}\bm{c}\in\mathcal{S}_{1}\cap Z\cdot\mathcal{B}_{M}(\bm{b}):\\ \bm{c}\equiv\bm{a}\bmod{n_{0}}\end{subarray}}(1+O_{F}(Z^{-\eta_{1}}))\cdot A_{F,1}^{\bm{a},n_{0}}(s).

For Theorem 1.9, certain degenerate residue classes play a larger role in local calculations than for Theorem 1.6. To pacify these residue classes, we need effective control on the variation of an individual local factor Lp(s,V𝒄)L_{p}(s,V_{\bm{c}}) over 𝒄𝒮1\bm{c}\in\mathcal{S}_{1}. We work pp-adically for convenience, using Remark 3.1 to define Lp(s,V𝒄)L_{p}(s,V_{\bm{c}}) for each 𝒄pm\bm{c}\in\mathbb{Z}_{p}^{m} with Δ(𝒄)0\Delta(\bm{c})\neq 0.

Conjecture 1.11 (EKL).

There exists a nonzero homogeneous polynomial H[𝐜]H\in\mathbb{Z}[\bm{c}], with H/Δ[𝐜]H/\Delta\in\mathbb{Z}[\bm{c}], such that for all primes pp and tuples 𝐚,𝐛pm\bm{a},\bm{b}\in\mathbb{Z}_{p}^{m} with H(𝐛)0H(\bm{b})\neq 0 and 𝐚𝐛modpH(𝐛)\bm{a}\equiv\bm{b}\bmod{pH(\bm{b})}, we have H(𝐚)0H(\bm{a})\neq 0 and Lp(s,V𝐚)=Lp(s,V𝐛)L_{p}(s,V_{\bm{a}})=L_{p}(s,V_{\bm{b}}).

Conjecture 1.11 is an effective Krasner-type statement for Lp(s,V𝒄)L_{p}(s,V_{\bm{c}}). A soft version follows from [kisin1999local], and suffices for Theorem 1.6 but not for Theorem 1.9. When m=4m=4, it should be possible to prove Conjecture 1.11 (with H×ΔH\in\mathbb{Q}^{\times}\cdot\Delta) using a minimal model for the Jacobian of V𝒄V_{\bm{c}}. In general, one might hope to take HH to be a power of Δ\Delta (or perhaps Δ\Delta itself).

1.1. Proof overview

§2 gives background on discriminants and the delta method. The delta method (see (2.10)) connects the point count NF,w(X)N_{F,w}(X) (from (1.2)) to the local behavior of the intersections F(𝒙)=𝒄𝒙=0F(\bm{x})=\bm{c}\cdot\bm{x}=0 over 𝔽p\mathbb{F}_{p}, p\mathbb{Z}_{p}, \mathbb{R}, and other rings, as 𝒄m\bm{c}\in\mathbb{Z}^{m} and pp vary. Cf. (1.1). We highlight several distinct sources of epsilon in the Hooley–Heath-Brown Theorem 1.1, and state a result from [wang2023_isolating_special_solutions] (over 𝒮0\mathcal{S}_{0}) addressing one such source.

§3 provides background on Hasse–Weil LL-functions and automorphic LL-functions.

§4 gives some local control on polynomials and LL-factors (based in part on [kisin1999local]); we need this for some local estimates and calculations.

§5 gives a useful “smooth framework” for dyadic decomposition and separation of variables. This lets us break certain key sums throughout the paper into more manageable pieces.

§6 derives some LL-function statistics over 𝒄𝒮1\bm{c}\in\mathcal{S}_{1}, after first doing local calculations (in the spirit of the Deligne–Katz equidistribution theorem) connected to RMT Symmetry Types (cf. [sarnak2016families]*Universality Conjecture). In particular, we state, and build on, some cases of the Ratios Conjectures for L(s,V𝒄)L(s,V_{\bm{c}}). Importantly here, the Ratios Recipe (see §6.3) can only apply once we restrict to 𝒄𝒮1\bm{c}\in\mathcal{S}_{1}. The recipe, naively extended to all 𝒄m\bm{c}\in\mathbb{Z}^{m}, would give false results (failing to detect the special subvarieties LΥL\in\Upsilon on F=0F=0 isolated in [wang2023_isolating_special_solutions]).

§7 begins to connect the “pure” LL-function statistics from §6 to the delta method. We approximate certain Dirichlet series “past” the critical line, in a reasonably simple and uniform way over 𝒄𝒮1\bm{c}\in\mathcal{S}_{1}, and control the resulting “approximation errors” on average. Handling these “errors” demands careful use of Hölder and other ideas. For example, by algorithmic tree-like means, we construct in §7.3 a small exceptional set away from which one may apply the “pure” Conjectures 1.8 and 1.10; it is also here that Conjecture 1.11 plays some key role.

The “mollified” series Φ𝒄,1(s)\Phi^{\bm{c},1}(s) from (1.9) not only makes the formulas in §6 nicer, but (as we will explain in §7) also holds significance in (2.10); this double significance, though innocent at first glance, secretly reflects a randomness property (connected to Deligne–Katz) stemming from the fact that degF3\deg{F}\geq 3. Throughout the paper, we take much advantage of the structure of (1.9); this is essential in Conjecture 1.4 and Lemma 6.15, for instance.

§8 proves new integral bounds sensitive to some real geometry (involving discriminants). Our approach shares some important features with [huang2020density] (a beautiful recent paper on approximate integral points). In addition, we have several parameters of interest, and must obtain genuinely multivariate decay. Keeping track of uniformity is tricky.

§9, like §8, proves some new “discriminating” pointwise estimates, but on complete exponential sums instead of oscillatory integrals. We then apply these to formulate and address (under Conjecture 1.5) a geometric analog of the Sarnak–Xue Density Hypothesis. Finite-field geometry (see [wang2023dichotomous]) and the geometric sieve (see [ekedahl1991infinite, bhargava2014geometric]) both play an important role here, as does a nice result of Busé and Jouanolou on discriminants (see Theorem 2.2). It is crucial throughout §9 that degF3\deg{F}\geq 3 (see e.g. Remark 9.2); this reflects a “randomness” not present for quadrics.

§10 ties everything together to prove our main results. We also isolate “axioms” that—if true—would allow for non-diagonal FF; see Theorems 10.5, 10.7, and 10.8. Here we only consider 𝒄𝒮1\bm{c}\in\mathcal{S}_{1}; there are also separate issues for 𝒄𝒮0\bm{c}\in\mathcal{S}_{0} (see [wang2023_isolating_special_solutions]*Remark 1.6).

For Theorem 1.3, see §10.2. Here we use an “entirely positive” Hölder argument over 𝒄𝒮1\bm{c}\in\mathcal{S}_{1}: we do not detect any cancellation over 𝒄\bm{c} that would go beyond a log-free “Mertens-type heuristic on average” (cf. [ng2004distribution]*Theorem 1(iii)) over 𝒄\bm{c}. Despite the “decoupling” convenience and power of Hölder, we must therefore be careful in §10.2 to obtain ϵ\epsilon-free bounds. (The structure of §10.2 is inspired by our work with the large sieve in [wang2023_large_sieve_diagonal_cubic_forms].)

The proof of Theorem 1.3 tells us (conditionally) that even if one takes absolute values over 𝐜𝒮1\bm{c}\in\mathcal{S}_{1} in (2.10), the XϵX^{\epsilon} allowance of [hooley_greaves_harman_huxley_1997, heath1998circle] is unnecessary; see (10.14) in Theorem 10.5. Theorem 1.3 also highlights a nontrivial use of the log-free order of magnitude in Conjecture 1.4, a robust qualitative prediction that (if true) could perhaps be explained in other ways (not just following the rather arithmetic Ratios Recipe).

For Theorem 1.6, Corollary 1.7, and Theorem 1.9, see §10.3. In most ranges, the “entirely positive” moment estimates of §10.2 still suffice. But this time, in a few key ranges, we identify cancellation over 𝒄𝒮1\bm{c}\in\mathcal{S}_{1} (via §7). One critical step here is a reduction, via §8, to large moduli in (2.10) (over 𝒄𝒮1\bm{c}\in\mathcal{S}_{1}), over which certain “mollified” RMT-type main terms vanish (cf. Lemma 6.15). There is also an alternative approach to cancellation (which we do not pursue): instead of Lemma 6.15, we could use the fact that m𝑑𝒄J𝒄,X(n)=0\int_{\mathbb{R}^{m}}d\bm{c}\,J_{\bm{c},X}(n)=0 (provided w(𝟎)=0w(\bm{0})=0), where J𝒄,X(n)J_{\bm{c},X}(n) is defined as in (2.9); cf. [wang2021_HLH_vs_RMT]*Observation 10.7.

1.2. Conventions

We write fSgf\ll_{S}g, or gSfg\gg_{S}f, to mean |f|Cg\lvert f\rvert\leq Cg for some C=C(S)>0C=C(S)>0. We let OS(g)O_{S}(g) denote a quantity that is Sg\ll_{S}g. We write fSgf\asymp_{S}g if fSgSff\ll_{S}g\ll_{S}f. We let oS;X(g)o_{S;X\to\infty}(g) denote a quantity ff such that for every ϵ>0\epsilon>0, there exists X0=X0(ϵ,S)>0X_{0}=X_{0}(\epsilon,S)>0 such that |f|ϵg\lvert f\rvert\leq\epsilon g holds for all XX0X\geq X_{0}. When making estimates, we think of mm, FF, ww as fixed, but may still occasionally write F\ll_{F} (or similar) for emphasis.

We frequently use indicator notation, letting 𝟏E\colonequals1\bm{1}_{E}\colonequals 1 if EE holds, and 𝟏E\colonequals0\bm{1}_{E}\colonequals 0 otherwise. For any nonempty set SS with an obvious measure (e.g. the counting measure on a finite set, or the usual Haar measure on pm\mathbb{Z}_{p}^{m}), we let 𝔼bS[f(b)]\mathbb{E}_{b\in S}[f(b)] denote the average of f(b)f(b) over bSb\in S.

We let 0\colonequals{a:a0}\mathbb{Z}_{\geq 0}\colonequals\{a\in\mathbb{Z}:a\geq 0\}, and similarly define sets like 0\mathbb{Z}_{\neq 0}, >1\mathbb{R}_{>1}, 2\mathbb{R}_{\geq 2}. For c0c\in\mathbb{Z}_{\neq 0}, we let vp(c)v_{p}(c) denote the pp-adic valuation of cc. For n1n\in\mathbb{Z}_{\geq 1}, we let φ(n)\varphi(n) denote the totient function, ω(n)\omega(n) the number of distinct prime factors of nn, and rad(n)\operatorname{rad}(n) the radical of nn.

We let Cc(s)C^{\infty}_{c}(\mathbb{R}^{s}) (resp. Cc(s)C^{\infty}_{c}(\mathbb{R}^{s})\otimes\mathbb{C}) denote the set of smooth compactly supported functions s\mathbb{R}^{s}\to\mathbb{R} (resp. s\mathbb{R}^{s}\to\mathbb{C}). For any function f=f(𝒖)f=f(\bm{u}), we let Suppf\operatorname{Supp}{f} denote the closure of {𝒖:f(𝒖)0}\{\bm{u}:f(\bm{u})\neq 0\} in the domain of ff; so for instance, the left-hand side of (1.11) equals Suppw\operatorname{Supp}{w}.

We let e(t)\colonequalse2πite(t)\colonequals e^{2\pi it}, and er(t)\colonequalse(t/r)e_{r}(t)\colonequals e(t/r). In integrals, we use notation analogous to summation notation. For instance, we write X𝑑xf(x)\int_{X}dx\,f(x) to mean Xf(x)𝑑x\int_{X}f(x)\,dx (in conventional notation), and we then write X×Y𝑑x𝑑yf(x,y)\int_{X\times Y}dx\,dy\,f(x,y) to mean X𝑑x(Y𝑑yf(x,y))\int_{X}dx\,(\int_{Y}dy\,f(x,y)).

We need concise notation for LpL^{p}-norms and p\ell^{p}-norms. If ff is a quantity depending on a scalar or vector variable tt (and possibly also on other variables), we write fLtp(S)\colonequals(tS𝑑t|f|p)1/p\lVert f\rVert_{L^{p}_{t}(S)}\colonequals(\int_{t\in S}dt\,\lvert f\rvert^{p})^{1/p} or ftp(S)\colonequals(tS|f|p)1/p\lVert f\rVert_{\ell^{p}_{t}(S)}\colonequals(\sum_{t\in S}\lvert f\rvert^{p})^{1/p} to denote the pp-norm of ff over tSt\in S, according as tt is a continuous or discrete variable, respectively. If the variable tt is clear from context, we may omit it.

We let u\colonequals/u\partial_{u}\colonequals\partial/\partial u for uu\in\mathbb{R}. When doing calculus in dd dimensions, a multi-index is a tuple of dd nonnegative integers (where dd will always be clear from context). Given a multi-index 𝜶0\bm{\alpha}\geq 0, we let |𝜶|\lvert\bm{\alpha}\rvert denote the sum of the coordinates of 𝜶\bm{\alpha}. For a vector 𝒖s\bm{u}\in\mathbb{R}^{s}, we let 𝒖\colonequalsmaxi(|ui|)\lVert\bm{u}\rVert\colonequals\max_{i}(\lvert u_{i}\rvert), and write d𝒖\colonequalsdu1dusd\bm{u}\colonequals du_{1}\cdots du_{s} and 𝒖𝜶\colonequalsu1α1usαs\partial_{\bm{u}}^{\bm{\alpha}}\colonequals\partial_{u_{1}}^{\alpha_{1}}\cdots\partial_{u_{s}}^{\alpha_{s}}. For example, if f=f(𝒖)Cc(s)f=f(\bm{u})\in C^{\infty}_{c}(\mathbb{R}^{s})\otimes\mathbb{C} and k0k\in\mathbb{Z}_{\geq 0}, then under our conventions,

max|𝜶|kmax𝒖s|𝒖𝜶f|=maxα1,,αs0:α1++αskmaxu1,,us|u1α1usαsf|.\max_{\lvert\bm{\alpha}\rvert\leq k}\max_{\bm{u}\in\mathbb{R}^{s}}{\lvert\partial_{\bm{u}}^{\bm{\alpha}}{f}\rvert}=\max_{\alpha_{1},\dots,\alpha_{s}\geq 0:\,\alpha_{1}+\dots+\alpha_{s}\leq k}\,\max_{u_{1},\dots,u_{s}\in\mathbb{R}}{\lvert\partial_{u_{1}}^{\alpha_{1}}\cdots\partial_{u_{s}}^{\alpha_{s}}f\rvert}.

For (repeated) later use, we fix a function ν0Cc(m)\nu_{0}\in C^{\infty}_{c}(\mathbb{R}^{m}), supported on [2,2]m[-2,2]^{m}, with 0ν010\leq\nu_{0}\leq 1 everywhere, ν0=1\nu_{0}=1 on [1/2,1/2]m[-1/2,1/2]^{m}, and m𝑑𝒙ν0(𝒙)=1\int_{\mathbb{R}^{m}}d\bm{x}\,\nu_{0}(\bm{x})=1. We also fix a radial222in the Euclidean sense (so that the value of ν1(𝒙)\nu_{1}(\bm{x}) depends only on x12++xm2x_{1}^{2}+\dots+x_{m}^{2}) function ν1Cc(m)\nu_{1}\in C^{\infty}_{c}(\mathbb{R}^{m}), supported on the annulus mx12++xm2m2m\leq x_{1}^{2}+\dots+x_{m}^{2}\leq m^{2} (so that ν1|[1,1]m=0\nu_{1}|_{[-1,1]^{m}}=0 and Suppν1[m,m]m\operatorname{Supp}{\nu_{1}}\subseteq[-m,m]^{m}), such that λ>0dλλν1(𝒙/λ)=1\int_{\lambda>0}\frac{d\lambda}{\lambda}\,\nu_{1}(\bm{x}/\lambda)=1 for all 𝒙m{𝟎}\bm{x}\in\mathbb{R}^{m}\setminus\{\bm{0}\}.

2. Background on discriminants and delta

Let disc(F)0\operatorname{disc}(F)\neq 0 be the discriminant of FF. Let m\colonequalsm3m_{\ast}\colonequals m-3. Define the bi-homogeneous polynomial expression disc(F,𝒄)\operatorname{disc}(F,\bm{c}) as in [wang2023dichotomous]*§3, so that if 𝒄m\bm{c}\in\mathbb{C}^{m}, then disc(F,𝒄)0\operatorname{disc}(F,\bm{c})\neq 0 if and only if the variety F(𝒙)=𝒄𝒙=0F(\bm{x})=\bm{c}\cdot\bm{x}=0 in m1\mathbb{P}^{m-1}_{\mathbb{C}} is smooth of dimension mm_{\ast}. Let

(2.1) Δ(𝒄)\colonequalsdisc(F)disc(F,𝒄);\Delta(\bm{c})\colonequals\operatorname{disc}(F)\cdot\operatorname{disc}(F,\bm{c});

then for any 𝒄m\bm{c}\in\mathbb{Z}^{m} and prime pp with pΔ(𝒄)p\nmid\Delta(\bm{c}), the varieties F(𝒙)=0F(\bm{x})=0 and F(𝒙)=𝒄𝒙=0F(\bm{x})=\bm{c}\cdot\bm{x}=0 in 𝔽pm1\mathbb{P}^{m-1}_{\mathbb{F}_{p}} are both smooth. Now recall 𝒮0\mathcal{S}_{0}, 𝒮1\mathcal{S}_{1} from (1.6). For each 𝒄𝒮1\bm{c}\in\mathcal{S}_{1}, let

(2.2) 𝒩𝒄\colonequals{n1:pnpΔ(𝒄)},𝒩𝒄\colonequals{n1:pnpΔ(𝒄)}.\mathcal{N}^{\bm{c}}\colonequals\{n\geq 1:p\mid n\Rightarrow p\nmid\Delta(\bm{c})\},\quad\mathcal{N}_{\bm{c}}\colonequals\{n\geq 1:p\mid n\Rightarrow p\mid\Delta(\bm{c})\}.
Lemma 2.1.

Let N,R1N,R\geq 1 be integers. Then |{Nn<2N:nR}|ϵ(RN)ϵ\lvert\{N\leq n<2N:n\mid R^{\infty}\}\rvert\ll_{\epsilon}(RN)^{\epsilon}, where nRn\mid R^{\infty} means rad(n)R\operatorname{rad}(n)\mid R. In particular, if 𝐜𝒮1\bm{c}\in\mathcal{S}_{1}, then |{Nn<2N:n𝒩𝐜}|ϵ𝐜ϵNϵ\lvert\{N\leq n<2N:n\in\mathcal{N}_{\bm{c}}\}\rvert\ll_{\epsilon}\lVert\bm{c}\rVert^{\epsilon}N^{\epsilon}.

Proof.

See e.g. [heath1998circle]*antepenultimate display of p. 683. ∎

As usual, we write F=(F/x1,,F/xm)\nabla{F}=(\partial F/\partial x_{1},\dots,\partial F/\partial x_{m}). It is known that the equation Δ(𝒄)=0\Delta(\bm{c})=0 defines the projective dual variety of VV (see e.g. [wang2023dichotomous]*Proposition 4.4). So

(2.3) Δ(F(𝒙))/F(𝒙)[𝒙].\Delta(\nabla{F}(\bm{x}))/F(\bm{x})\in\mathbb{Q}[\bm{x}].

Also, Δ(𝒄)\Delta(\bm{c}) is irreducible in ¯[c1,,cm]\overline{\mathbb{Q}}[c_{1},\dots,c_{m}], and has total degree degΔ=32m2\deg{\Delta}=3\cdot 2^{m-2} in c1,,cmc_{1},\dots,c_{m}. In particular, degΔ12\deg\Delta\geq 12 (since m4m\geq 4), so by [castryck2020dimension]*Theorem 1, we have

(2.4) |𝒮0[Z,Z]m|mZm2\lvert\mathcal{S}_{0}\cap[-Z,Z]^{m}\rvert\ll_{m}Z^{m-2}

for all reals Z1Z\geq 1. (For diagonal FF, more is known, e.g. |𝒮0[Z,Z]m|m,ϵZm/2+ϵ\lvert\mathcal{S}_{0}\cap[-Z,Z]^{m}\rvert\ll_{m,\epsilon}Z^{m/2+\epsilon}.)

One can express disc(F,𝒄)\operatorname{disc}(F,\bm{c}) in terms of the discriminant of a cubic form in m1m-1 variables:

(2.5) disc(F,𝒄)=±disc(F(x1,,xm1,(c1x1++cm1xm1)/cm))cm32m2\operatorname{disc}(F,\bm{c})=\pm\operatorname{disc}(F(x_{1},\dots,x_{m-1},-(c_{1}x_{1}+\dots+c_{m-1}x_{m-1})/c_{m}))\cdot c_{m}^{3\cdot 2^{m-2}}

[wang2023dichotomous]*Proposition 3.2. This lets us bring into play a nice result of [buse2014discriminant] (though the weaker classical result [hooley1988nonary]*(84) on p. 62 would also suffice for most of our needs):

Theorem 2.2 ([buse2014discriminant]*Corollary 4.30).

Let RR be a ring and let fR[x1,,xm1]f\in R[x_{1},\dots,x_{m-1}] be a homogeneous polynomial. If 1am11\leq a\leq m-1, then there exists N0N\geq 0 such that xaNdisc(f)x_{a}^{N}\operatorname{disc}(f) lies in the homogeneous ideal of R[x1,,xm1]R[x_{1},\dots,x_{m-1}] generated by

f,{(xif)(xjf):1i,jm1}.f,\;\{(\partial_{x_{i}}{f})(\partial_{x_{j}}{f}):1\leq i,j\leq m-1\}.

Via (2.5), Theorem 2.2 has the following corollary, useful in §9:

Corollary 2.3.

If 1a,bm1\leq a,b\leq m, then there exists N0N\geq 0 such that xaNcbNΔ(𝐜)x_{a}^{N}c_{b}^{N}\Delta(\bm{c}) lies in the homogeneous ideal of [x1,,xm,c1,,cm]\mathbb{Z}[x_{1},\dots,x_{m},c_{1},\dots,c_{m}] generated by

F(𝒙),𝒄𝒙,{(ckxiFcixkF)(ckxjFcjxkF):1i,j,km}.F(\bm{x}),\;\bm{c}\cdot\bm{x},\;\{(c_{k}\cdot\partial_{x_{i}}{F}-c_{i}\cdot\partial_{x_{k}}{F})(c_{k}\cdot\partial_{x_{j}}{F}-c_{j}\cdot\partial_{x_{k}}{F}):1\leq i,j,k\leq m\}.
Proof.

Assume b=mb=m. Then use (2.1), (2.5), Theorem 2.2 with R=[c1/cm,,cm1/cm]R=\mathbb{Z}[c_{1}/c_{m},\dots,c_{m-1}/c_{m}], and the congruence (c1x1++cm1xm1)/cmxmmod(𝒄/cm)𝒙-(c_{1}x_{1}+\dots+c_{m-1}x_{m-1})/c_{m}\equiv x_{m}\bmod{(\bm{c}/c_{m})\cdot\bm{x}} in R[x1,,xm]R[x_{1},\dots,x_{m}]. ∎

Fix wCc(m)w\in C^{\infty}_{c}(\mathbb{R}^{m}) satisfying (1.11) if FF is diagonal, or satisfying

(2.6) Suppw{𝒙m:det(HessF(𝒙))0}\operatorname{Supp}{w}\subseteq\{\bm{x}\in\mathbb{R}^{m}:\det(\operatorname{Hess}{F}(\bm{x}))\neq 0\}

in general.333For convenience, we will maintain this hypothesis for the rest of the paper, except when specified otherwise. Recall NF,w(X)N_{F,w}(X) from (1.2). The delta method of [duke1993bounds, heath1996new] allows one to express NF,w(X)N_{F,w}(X), up to a negligible error, as a sum over 𝒄m\bm{c}\in\mathbb{Z}^{m} of “adelic” data. We use the precise setup from [wang2023_isolating_special_solutions]*§1 (based on [duke1993bounds, heath1996new, heath1998circle]).

Let h:(0,)×h\colon(0,\infty)\times\mathbb{R}\to\mathbb{R} be the smooth function given by [heath1998circle]*(2.3); we will need the full definition later, in §8. For each real X1X\geq 1, let Y\colonequalsX3/2Y\colonequals X^{3/2}. Let

(2.7) I𝒄(n)\displaystyle I_{\bm{c}}(n) \colonequals𝒙m𝑑𝒙w(𝒙/X)h(n/Y,F(𝒙)/Y2)e(𝒄𝒙/n),\displaystyle\colonequals\int_{\bm{x}\in\mathbb{R}^{m}}d\bm{x}\,w(\bm{x}/X)h(n/Y,F(\bm{x})/Y^{2})e(-\bm{c}\cdot\bm{x}/n),
(2.8) S𝒄(n)\displaystyle S_{\bm{c}}(n) \colonequals1an:gcd(a,n)=11𝒙nen(aF(𝒙)+𝒄𝒙),\displaystyle\colonequals\sum_{1\leq a\leq n:\,\gcd(a,n)=1}\,\sum_{1\leq\bm{x}\leq n}e_{n}(aF(\bm{x})+\bm{c}\cdot\bm{x}),
(2.9) S𝒄(n)\displaystyle S^{\natural}_{\bm{c}}(n) \colonequalsn(m+1)/2S𝒄(n),J𝒄,X(n)\colonequalsXmI𝒄(n).\displaystyle\colonequals n^{-(m+1)/2}S_{\bm{c}}(n),\quad J_{\bm{c},X}(n)\colonequals X^{-m}I_{\bm{c}}(n).

A simple rearrangement of [wang2023_isolating_special_solutions]*(1.3) gives (for all A>0A>0)

(2.10) (1+OA(YA))NF,w(X)/Xm3=n1𝒄mn(1m)/2S𝒄(n)J𝒄,X(n).(1+O_{A}(Y^{-A}))\cdot N_{F,w}(X)/X^{m-3}=\sum_{n\geq 1}\sum_{\bm{c}\in\mathbb{Z}^{m}}n^{(1-m)/2}S^{\natural}_{\bm{c}}(n)J_{\bm{c},X}(n).

The infinite sums in (2.10) are essentially finite, by the following standard result:

Proposition 2.4.

For some constant A0=A0(F,w)>0A_{0}=A_{0}(F,w)>0, we have J𝐜,X(n)=0J_{\bm{c},X}(n)=0 for all nA0Yn\geq A_{0}Y. Also, J𝐜,X(n)ϵ,AXAJ_{\bm{c},X}(n)\ll_{\epsilon,A}X^{-A} holds whenever ϵ,A>0\epsilon,A>0 and 𝐜X1/2+ϵ\lVert\bm{c}\rVert\geq X^{1/2+\epsilon}.

Proof.

See e.g. [wang2023_isolating_special_solutions]*Proposition 5.1. ∎

We now recall some background (cf. [hooley1986HasseWeil]*§§5–6) on the sums S𝒄(n)S_{\bm{c}}(n). It is known that S𝒄(n)S_{\bm{c}}(n), and thus S𝒄(n)S^{\natural}_{\bm{c}}(n) too, is multiplicative in nn. The Dirichlet series

(2.11) Φ(𝒄,s)\colonequalsn1nsS𝒄(n)\Phi(\bm{c},s)\colonequals\sum_{n\geq 1}n^{-s}S^{\natural}_{\bm{c}}(n)

thus has an Euler product. At prime powers, S𝒄S_{\bm{c}} is related to certain point counts in projective space. Given a prime power qq, let 𝒱(𝔽q)\mathcal{V}(\mathbb{F}_{q}) be the set of 𝔽q\mathbb{F}_{q}-points on the variety F(𝒙)=0F(\bm{x})=0 in 𝔽qm1\mathbb{P}^{m-1}_{\mathbb{F}_{q}}, let 𝒱𝒄(𝔽q)\mathcal{V}_{\bm{c}}(\mathbb{F}_{q}) be the set of 𝔽q\mathbb{F}_{q}-points on the variety F(𝒙)=𝒄𝒙=0F(\bm{x})=\bm{c}\cdot\bm{x}=0 in 𝔽qm1\mathbb{P}^{m-1}_{\mathbb{F}_{q}}, and let

EF(q)\colonequals|𝒱(𝔽q)||1+m(𝔽q)|,E𝒄(q)\colonequals|𝒱𝒄(𝔽q)||m(𝔽q)|,E_{F}(q)\colonequals\lvert\mathcal{V}(\mathbb{F}_{q})\rvert-\lvert\mathbb{P}^{1+m_{\ast}}(\mathbb{F}_{q})\rvert,\quad E_{\bm{c}}(q)\colonequals\lvert\mathcal{V}_{\bm{c}}(\mathbb{F}_{q})\rvert-\lvert\mathbb{P}^{m_{\ast}}(\mathbb{F}_{q})\rvert,

where |d(𝔽q)|=(qd+11)/(q1)\lvert\mathbb{P}^{d}(\mathbb{F}_{q})\rvert=(q^{d+1}-1)/(q-1). Then let

(2.12) EF(q)\colonequalsq(1+m)/2EF(q),E𝒄(q)\colonequalsqm/2E𝒄(q).E^{\natural}_{F}(q)\colonequals q^{-(1+m_{\ast})/2}E_{F}(q),\quad E^{\natural}_{\bm{c}}(q)\colonequals q^{-m_{\ast}/2}E_{\bm{c}}(q).

If p𝒄p\nmid\bm{c} (e.g. if pΔ(𝒄)p\nmid\Delta(\bm{c})), then by [wang2022thesis]*Proposition 3.2.4, we have

(2.13) S𝒄(p)=E𝒄(p)p1/2EF(p).S^{\natural}_{\bm{c}}(p)=E^{\natural}_{\bm{c}}(p)-p^{-1/2}E^{\natural}_{F}(p).

If pΔ(𝒄)p\nmid\Delta(\bm{c}) and l2l\geq 2, then by [wang2022thesis]*Proposition 3.2.6, we have

(2.14) S𝒄(pl)=0.S^{\natural}_{\bm{c}}(p^{l})=0.

Recall 𝒮0\mathcal{S}_{0}, 𝒮1\mathcal{S}_{1} from (1.6). For any set 𝒮m\mathcal{S}\subseteq\mathbb{Z}^{m}, let

(2.15) Σ(X,𝒮)\colonequalsn1𝒄𝒮n(1m)/2S𝒄(n)J𝒄,X(n),Σ(X,𝒮)\colonequalsXm3Σ(X,𝒮).\Sigma^{\natural}(X,\mathcal{S})\colonequals\sum_{n\geq 1}\sum_{\bm{c}\in\mathcal{S}}n^{(1-m)/2}S^{\natural}_{\bm{c}}(n)J_{\bm{c},X}(n),\quad\Sigma(X,\mathcal{S})\colonequals X^{m-3}\Sigma^{\natural}(X,\mathcal{S}).

If 𝒄𝒮0\bm{c}\in\mathcal{S}_{0}, then Φ(𝒄,s)\Phi(\bm{c},s) can resemble ζ(s1/2)\zeta(s-1/2) (in some sense), leading to the following result:

Theorem 2.5 ([wang2023_isolating_special_solutions]).

Suppose m=6m=6 and FF is diagonal. Then

(2.16) Σ(X,𝒮0)=OF,w,ϵ(X2.75+ϵ)+𝔖Fσ,F,wX3+LΥ𝒙L6w(𝒙/X),\Sigma(X,\mathcal{S}_{0})=O_{F,w,\epsilon}(X^{2.75+\epsilon})+\mathfrak{S}_{F}\cdot\sigma_{\infty,F,w}\cdot X^{3}+\sum_{L\in\Upsilon}\sum_{\bm{x}\in L\cap\mathbb{Z}^{6}}w(\bm{x}/X),

unconditionally. In particular, Σ(X,𝒮0)F,wX3\Sigma(X,\mathcal{S}_{0})\ll_{F,w}X^{3}.

Proof.

This follows from [wang2023_isolating_special_solutions]*Corollary 1.2 and (1.7), even if we relax the condition (2.6) to 𝟎Suppw\bm{0}\notin\operatorname{Supp}{w}. The earlier paper [heath1998circle] had proved Σ(X,𝒮0)F,w,ϵX3+ϵ\Sigma(X,\mathcal{S}_{0})\ll_{F,w,\epsilon}X^{3+\epsilon}. ∎

Since NF,w(X)=Σ(X,𝒮0)+Σ(X,𝒮1)+OA(XA)N_{F,w}(X)=\Sigma(X,\mathcal{S}_{0})+\Sigma(X,\mathcal{S}_{1})+O_{A}(X^{-A}), we may thus concentrate on Σ(X,𝒮1)\Sigma(X,\mathcal{S}_{1}). The rest of §2 provides some technical context for our work (in comparison with Theorem 1.1 due to [hooley_greaves_harman_huxley_1997, heath1998circle]), but is not logically necessary for the paper.

If 𝒄𝒮1\bm{c}\in\mathcal{S}_{1}, then by (2.13), (2.14), and (3.4), one might expect Φ(𝒄,s)\Phi(\bm{c},s) to resemble L(s,V𝒄)(1)mL(s,V_{\bm{c}})^{(-1)^{m_{\ast}}}, up to a factor absolutely convergent for Re(s)>1/2\operatorname{Re}(s)>1/2; cf. [wang2023_large_sieve_diagonal_cubic_forms]*(2.4). With this intuition in mind, let us now recall how one can prove (as is key for Theorem 1.1)

(2.17) Σ(X,𝒮1)ϵX3(m2)/4+ϵ,or equivalently Σ(X,𝒮1)ϵX(6m)/4+ϵ,\Sigma(X,\mathcal{S}_{1})\ll_{\epsilon}X^{3(m-2)/4+\epsilon},\quad\textnormal{or equivalently }\Sigma^{\natural}(X,\mathcal{S}_{1})\ll_{\epsilon}X^{(6-m)/4+\epsilon},

under Conjecture 1.2 (HW2) for L(s,V𝒄)L(s,V_{\bm{c}}), when mm is even and FF is diagonal. We find it illuminating to work in this generality, but key here is that 3(m2)/4=33(m-2)/4=3 when m=6m=6.

Conditional proof sketch for (2.17).

For this proof sketch only, let Z\colonequalsX1/2+ϵ0Z\colonequals X^{1/2+\epsilon_{0}}. Let

(2.18) Ψ𝒄,1(s)\colonequals1/L(s,V𝒄),Ψ𝒄,2(s)\colonequalsΦ(𝒄,s)L(s,V𝒄),\Psi^{\bm{c},1}(s)\colonequals 1/L(s,V_{\bm{c}}),\quad\Psi^{\bm{c},2}(s)\colonequals\Phi(\bm{c},s)L(s,V_{\bm{c}}),

for 𝒄𝒮1\bm{c}\in\mathcal{S}_{1}; then Φ=Ψ𝒄,1Ψ𝒄,2\Phi=\Psi^{\bm{c},1}\Psi^{\bm{c},2}. Let μ𝒄(n)\mu_{\bm{c}}(n), a𝒄!(n)a^{!}_{\bm{c}}(n) be the nnth coefficients of the Dirichlet series Ψ𝒄,1\Psi^{\bm{c},1}, Ψ𝒄,2\Psi^{\bm{c},2}, respectively. Then the following hold (see e.g. [wang2023_large_sieve_diagonal_cubic_forms]*Proposition 4.12):

  1. ]

  2. [B1’

    For NZ3N\leq Z^{3}, we have 𝒄𝒮1[Z,Z]mNn<2N|a𝒄!(n)|ϵZm+ϵN1/2\sum_{\bm{c}\in\mathcal{S}_{1}\cap[-Z,Z]^{m}}\sum_{N\leq n<2N}\lvert a^{!}_{\bm{c}}(n)\rvert\ll_{\epsilon}Z^{m+\epsilon}N^{1/2}.

  3. [B2’

    For NZ3N\leq Z^{3}, we have 𝒄𝒮1[Z,Z]m(Nn<2N|a𝒄!(n)|)2ϵZm+ϵN\sum_{\bm{c}\in\mathcal{S}_{1}\cap[-Z,Z]^{m}}(\sum_{N\leq n<2N}\lvert a^{!}_{\bm{c}}(n)\rvert)^{2}\ll_{\epsilon}Z^{m+\epsilon}N.

The arguments in [hooley1986HasseWeil, hooley_greaves_harman_huxley_1997] and [heath1998circle] can then loosely be interpreted as

  1. (H1)

    using partial summation over n[N,2N)n\in[N,2N) to “factor out” J𝒄,X(n)J_{\bm{c},X}(n) from the sum over nn in (2.15) (for 𝒮=𝒮1\mathcal{S}=\mathcal{S}_{1}), and then bounding the JJ-contribution in n([N,2N))\ell^{\infty}_{n}([N,2N));

  2. (H2)

    expanding S𝒄=μ𝒄a𝒄!S^{\natural}_{\bm{c}}=\mu_{\bm{c}}\ast a^{!}_{\bm{c}} using Φ=Ψ𝒄,1Ψ𝒄,2\Phi=\Psi^{\bm{c},1}\Psi^{\bm{c},2};

  3. (H3)

    using GRH to bound the Ψ𝒄,1\Psi^{\bm{c},1}-contribution in 𝒄(𝒮1[Z,Z]m)\ell^{\infty}_{\bm{c}}(\mathcal{S}_{1}\cap[-Z,Z]^{m}); and

  4. (H4)

    using [B1’] afterwards, to bound the Ψ𝒄,2\Psi^{\bm{c},2}-contribution in 𝒄1(𝒮1[Z,Z]m)\ell^{1}_{\bm{c}}(\mathcal{S}_{1}\cap[-Z,Z]^{m}).

Upon dyadic summation over 1NY1\ll N\ll Y, one gets Σ(X,𝒮1)ϵ0X(6m)/4+O(ϵ0)\Sigma^{\natural}(X,\mathcal{S}_{1})\ll_{\epsilon_{0}}X^{(6-m)/4+O(\epsilon_{0})}.

In place of GRH, one could use an elementary 2\ell^{2} statement in the spirit of a large sieve inequality. One would then use [B2’] instead of [B1’]. (See [wang2023_large_sieve_diagonal_cubic_forms].) ∎

We now diagnose (and sketch “cures” for) the key “sources of ϵ\epsilon” above:

  1. (1)

    In (H3), the pointwise GRH bound Nn<2Nμ𝒄(n)ϵ𝒄ϵN1/2+ϵ\sum_{N\leq n<2N}\mu_{\bm{c}}(n)\ll_{\epsilon}\lVert\bm{c}\rVert^{\epsilon}N^{1/2+\epsilon}. (Cure: RMT-type predictions such as Conjectures 1.4 and 1.8.)

  2. (2)

    In (H4), the bound [B1’]. (Or [B2’], for the argument of [wang2023_large_sieve_diagonal_cubic_forms].) In fact, upon closer inspection, the proofs of [B1’] and [B2’] each have two sources of ϵ\epsilon:

    1. (a)

      Good prime factors pΔ(𝒄)p\nmid\Delta(\bm{c}) of nn, via the “first-order error” present in Ψ𝒄,2\Psi^{\bm{c},2}. (Cure: Replacing Ψ𝒄,1\Psi^{\bm{c},1} with a “better approximation” of Φ\Phi.)

    2. (b)

      Bad prime factors pΔ(𝒄)p\mid\Delta(\bm{c}) of nn, via the sometimes large failure of square-root cancellation in individual sums of the form S𝒄(pl)S_{\bm{c}}(p^{l}). (Cure: New and old pointwise bounds on |S𝒄(pl)|\lvert S_{\bm{c}}(p^{l})\rvert, and the average-type Conjecture 1.5.)

  3. (3)

    The fact that over 1NY1\ll N\ll Y, each dyadic range Nn<2NN\leq n<2N contributes roughly equally to the final bound (2.17); cf. [wang2023_large_sieve_diagonal_cubic_forms]*Remark 5.3. If unaddressed, then the following sources of ϵ\epsilon would arise (in our work):

    1. (a)

      A fatal logX\log{X} factor in our proof of Theorem 1.3, via summation over NN.

    2. (b)

      A large contribution from relatively small nn in our proof of Theorems 1.6 and 1.9, when handling variation of J𝒄,X(n)J_{\bm{c},X}(n) over 𝒄\bm{c}—a step needed when applying Conjecture 1.8 or 1.10 to beat pointwise GRH.

    (Cure: New integral bounds that decay, as n0n\to 0, fairly uniformly over 𝒄\bm{c}.)

  4. (4)

    The lack of “ϵ\epsilon-care” in bounds and “decay cutoffs” for integrals. The best recorded integral estimates (valid at least for some ww) seem to be

    (2.19) J𝒄,X(n),nJ𝒄,X(n)/nA(1+𝒄/X1/2)Af(2+X𝒄/n),J_{\bm{c},X}(n),\,n\cdot\partial{J_{\bm{c},X}(n)}/\partial{n}\ll_{A}(1+\lVert\bm{c}\rVert/X^{1/2})^{-A}\cdot f(2+X\lVert\bm{c}\rVert/n),

    with f(r)=r1m/2(logr)mf(r)=r^{1-m/2}(\log{r})^{m}, from [hooley2014octonary]*p. 252, (31). Summing over n/X1ϵ𝒄X1/2n/X^{1-\epsilon}\leq\lVert\bm{c}\rVert\leq X^{1/2} carefully, or summing over X1/2𝒄X1/2+ϵX^{1/2}\leq\lVert\bm{c}\rVert\leq X^{1/2+\epsilon} carelessly, incurs ϵ\epsilon-losses. (Cure: Summing carefully over 𝒄\bm{c}, and using the “cure to (3)” over small nn.)

3. Background on individual LL-functions

3.1. Geometric background

We need to give a precise meaning to Conjecture 1.2. We first define the necessary Hasse–Weil LL-functions and their local factors, following [serre1969facteurs]. (Another option, not pursued here, would be to follow [taylor2004galois].) This is technical, but allows us to capture “variation in pp” in a representation-theoretic framework. At most primes, the data captured is very concrete; see e.g. (3.4).

For any perfect field KK, let GK\colonequalsGal(K¯/K)G_{K}\colonequals\operatorname{Gal}(\overline{K}/K). Let Γ(s)\colonequalsπs/2Γ(s/2)\Gamma_{\mathbb{R}}(s)\colonequals\pi^{-s/2}\Gamma(s/2) and Γ(s)\colonequals(2π)sΓ(s)\Gamma_{\mathbb{C}}(s)\colonequals(2\pi)^{-s}\Gamma(s). The case of the Riemann zeta function ζ(s)\zeta(s) in (1.7) is familiar (with ζp(s)=(1ps)1\zeta_{p}(s)=(1-p^{-s})^{-1} and ζ(s)=Γ(s)\zeta_{\infty}(s)=\Gamma_{\mathbb{R}}(s)), so we focus on the other cases. Let 𝒄𝒮1\bm{c}\in\mathcal{S}_{1}. Let \ell be a prime, and (viewing VV, V𝒄V_{\bm{c}} as subvarieties of m1\mathbb{P}^{m-1}) consider the \ell-adic Galois representations

ρV:GH1+m(Vׯ,)H1+m(¯m1,),ρV𝒄:GHm(V𝒄ׯ,)Hm(¯m1,).\rho_{V}\colon G_{\mathbb{Q}}\to\frac{H^{1+m_{\ast}}(V\times_{\mathbb{Q}}\overline{\mathbb{Q}},\mathbb{Q}_{\ell})}{H^{1+m_{\ast}}(\mathbb{P}^{m-1}_{\overline{\mathbb{Q}}},\mathbb{Q}_{\ell})},\quad\rho_{V_{\bm{c}}}\colon G_{\mathbb{Q}}\to\frac{H^{m_{\ast}}(V_{\bm{c}}\times_{\mathbb{Q}}\overline{\mathbb{Q}},\mathbb{Q}_{\ell})}{H^{m_{\ast}}(\mathbb{P}^{m-1}_{\overline{\mathbb{Q}}},\mathbb{Q}_{\ell})}.

It is known that the representations ρV\rho_{V}, ρV𝒄\rho_{V_{\bm{c}}}, 2ρV𝒄\bigotimes^{2}\rho_{V_{\bm{c}}}, Sym2ρV𝒄\operatorname{Sym}^{2}\rho_{V_{\bm{c}}}, 2ρV𝒄\bigwedge^{2}\rho_{V_{\bm{c}}} of GG_{\mathbb{Q}} have dimensions depending only on mm, and are pure of weight (1+m)/2(1+m_{\ast})/2, m/2m_{\ast}/2, mm_{\ast}, mm_{\ast}, mm_{\ast}, respectively.

Let ρ:GM\rho\colon G_{\mathbb{Q}}\to M be one of these five representations, and let dρd_{\rho}, wρw_{\rho} be the dimension and weight of ρ\rho, respectively. Define Γρ(s)\Gamma_{\rho}(s) using Hodge theory, following [serre1969facteurs]*§3.2 (after passing from MM to a singular cohomology group M/M^{\prime}/\mathbb{C} independent of \ell); then for certain integers hρ(+),hρ(),hρ(a,b)0h_{\rho}(+),h_{\rho}(-),h_{\rho}(a,b)\geq 0 with hρ(+)+hρ()+20a<b:a+b=wρhρ(a,b)=dρh_{\rho}(+)+h_{\rho}(-)+2\sum_{0\leq a<b:\,a+b=w_{\rho}}h_{\rho}(a,b)=d_{\rho}, we have

Γρ(s)=Γ(swρ/2)hρ(+)Γ(swρ/2+1)hρ()0a<b:a+b=wρΓ(sa)hρ(a,b).\Gamma_{\rho}(s)=\Gamma_{\mathbb{R}}(s-w_{\rho}/2)^{h_{\rho}(+)}\Gamma_{\mathbb{R}}(s-w_{\rho}/2+1)^{h_{\rho}(-)}\prod_{0\leq a<b:\,a+b=w_{\rho}}\Gamma_{\mathbb{C}}(s-a)^{h_{\rho}(a,b)}.

Let L(s,ρ)\colonequalsΓρ(s+wρ/2)L_{\infty}(s,\rho)\colonequals\Gamma_{\rho}(s+w_{\rho}/2); then L(s,ρ)L_{\infty}(s,\rho) is holomorphic on the half-plane Re(s)>0\operatorname{Re}(s)>0. Note that if we let 𝒄\bm{c}, ρ\rho vary, the number of possible functions L(s,ρ)L_{\infty}(s,\rho) could be is m1\ll_{m}1.

For each prime pp\neq\ell, we may restrict ρ\rho to GpG_{\mathbb{Q}_{p}}; let Pp(T)\colonequalsdet(1πpT)[T]P_{p}(T)\colonequals\det(1-\pi_{p}T)\in\mathbb{Q}_{\ell}[T] denote the reverse characteristic polynomial of geometric Frobenius on MIpM^{I_{\mathbb{Q}_{p}}} (the inertia invariants of MM), following [serre1969facteurs]*§2.2. Clearly degPpdρ\deg P_{p}\leq d_{\rho}, with equality if and only if ρ\rho is unramified at pp. Write Pp(T)=1jdegPp(1αρ,j(p)T)P_{p}(T)=\prod_{1\leq j\leq\deg P_{p}}(1-\alpha_{\rho,j}(p)T), and let

(3.1) α~ρ,j(p)\colonequalspwρ/2αρ,j(p),Lp(s,ρ)\colonequals1jdegPp(1α~ρ,j(p)ps)1=Pp(pswρ/2)1.\tilde{\alpha}_{\rho,j}(p)\colonequals p^{-w_{\rho}/2}\alpha_{\rho,j}(p),\quad L_{p}(s,\rho)\colonequals\prod_{1\leq j\leq\deg P_{p}}(1-\tilde{\alpha}_{\rho,j}(p)p^{-s})^{-1}=P_{p}(p^{-s-w_{\rho}/2})^{-1}.

Because VV, V𝒄V_{\bm{c}} are complete intersections in m1\mathbb{P}^{m-1}_{\mathbb{Q}}, it is now known444thanks to [laskar2017local]*Corollary 1.2 and its proof (cf. [saito2003weight]*Corollary 0.6), which builds on progress of [scholze2012perfectoid] on the weight-monodromy conjecture that the polynomial Pp(T)P_{p}(T) lies in [T]\mathbb{Q}[T] and is independent of \ell (so that {αρ,j(p)}j\{\alpha_{\rho,j}(p)\}_{j} is a multiset of algebraic numbers), and furthermore (for any embedding of αρ,j(p)\alpha_{\rho,j}(p) into the complex numbers) we have

(3.2) |αρ,j(p)|pwρ/2,|α~ρ,j(p)|1,\lvert\alpha_{\rho,j}(p)\rvert\leq p^{w_{\rho}/2},\quad\lvert\tilde{\alpha}_{\rho,j}(p)\rvert\leq 1,

for all pp, jj. One might also be able to directly (without automorphy) define a conductor and root number for ρ\rho independent of \ell (following [serre1969facteurs] and [deligne1969constantes]), but we need not do so.

Let L(s,V)\colonequalsL(s,ρV)L_{\infty}(s,V)\colonequals L_{\infty}(s,\rho_{V}) for any \ell. Given a prime pp, let Lp(s,V)\colonequalsLp(s,ρV)L_{p}(s,V)\colonequals L_{p}(s,\rho_{V}) for any p\ell\neq p, and let α~V,j(p)\colonequalsα~ρV,j(p)\tilde{\alpha}_{V,j}(p)\colonequals\tilde{\alpha}_{\rho_{V},j}(p) and dV,p\colonequalsdegPpd_{V,p}\colonequals\deg P_{p}. Let

L(s,V)\colonequalsp<Lp(s,V)\equalscolonn1λV(n)ns,L(s,V)\colonequals\prod_{p<\infty}L_{p}(s,V)\equalscolon\sum_{n\geq 1}\lambda^{\natural}_{V}(n)n^{-s},

so that λV(p)=1jdV,pα~V,j(p)\lambda^{\natural}_{V}(p)=\sum_{1\leq j\leq d_{V,p}}\tilde{\alpha}_{V,j}(p) for all pp. Let degL(s,V)\colonequalsmaxpdV,p=dρV\deg L(s,V)\colonequals\max_{p}d_{V,p}=d_{\rho_{V}}. Make analogous definitions for L(s,V𝒄)L(s,V_{\bm{c}}) and its tensor squares in (1.7), in terms of ρV𝒄\rho_{V_{\bm{c}}} and its tensor squares. If pΔ(𝒄)p\nmid\Delta(\bm{c}) and d=degL(s,V𝒄)d=\deg L(s,V_{\bm{c}}), then by (2.12), (3.1), smooth proper base change, and the Grothendieck–Lefschetz trace formula, we have (for instance)

(3.3) λV(p)\displaystyle\lambda^{\natural}_{V}(p) =1jdegL(s,V)α~V,j(p)=(1)1+mEF(p),\displaystyle=\sum_{1\leq j\leq\deg L(s,V)}\tilde{\alpha}_{V,j}(p)=(-1)^{1+m_{\ast}}E^{\natural}_{F}(p),
(3.4) λV𝒄(p)\displaystyle\lambda^{\natural}_{V_{\bm{c}}}(p) =1jdα~V𝒄,j(p)=(1)mE𝒄(p),\displaystyle=\sum_{1\leq j\leq d}\tilde{\alpha}_{V_{\bm{c}},j}(p)=(-1)^{m_{\ast}}E^{\natural}_{\bm{c}}(p),
(3.5) λV𝒄,Sym2(p)\displaystyle\lambda^{\natural}_{V_{\bm{c}},\operatorname{Sym}^{2}}(p) =1ijdα~V𝒄,i(p)α~V𝒄,j(p)=λV𝒄(p2),\displaystyle=\sum_{1\leq i\leq j\leq d}\tilde{\alpha}_{V_{\bm{c}},i}(p)\tilde{\alpha}_{V_{\bm{c}},j}(p)=\lambda^{\natural}_{V_{\bm{c}}}(p^{2}),
(3.6) λV𝒄,2(p)\displaystyle\lambda^{\natural}_{V_{\bm{c}},\bigwedge^{2}}(p) =1i<jdα~V𝒄,i(p)α~V𝒄,j(p)=λV𝒄(p)2λV𝒄(p2),\displaystyle=\sum_{1\leq i<j\leq d}\tilde{\alpha}_{V_{\bm{c}},i}(p)\tilde{\alpha}_{V_{\bm{c}},j}(p)=\lambda^{\natural}_{V_{\bm{c}}}(p)^{2}-\lambda^{\natural}_{V_{\bm{c}}}(p^{2}),
(3.7) (1)mE𝒄(p2)\displaystyle(-1)^{m_{\ast}}E^{\natural}_{\bm{c}}(p^{2}) =1jdα~V𝒄,j(p)2=E𝒄(p)22λV𝒄,2(p),\displaystyle=\sum_{1\leq j\leq d}\tilde{\alpha}_{V_{\bm{c}},j}(p)^{2}=E^{\natural}_{\bm{c}}(p)^{2}-2\lambda^{\natural}_{V_{\bm{c}},\bigwedge^{2}}(p),

where for all jj we have (by the Weil conjectures, since pΔ(𝒄)p\nmid\Delta(\bm{c}))

(3.8) |α~V,j(p)|=1,|α~V𝒄,j(p)|=1.\lvert\tilde{\alpha}_{V,j}(p)\rvert=1,\quad\lvert\tilde{\alpha}_{V_{\bm{c}},j}(p)\rvert=1.
Remark 3.1.

By working locally from the beginning, one can define (for any integer n1n\geq 1) the quantities pnLp(s,V𝒄)\prod_{p\mid n}L_{p}(s,V_{\bm{c}}) and λV𝒄(n)\lambda^{\natural}_{V_{\bm{c}}}(n) on all of {𝒄pnpm:Δ(𝒄)0}\{\bm{c}\in\prod_{p\mid n}\mathbb{Z}_{p}^{m}:\Delta(\bm{c})\neq 0\}. Extended definitions like this will be convenient for local calculations in §6.1 and §7.3.

3.2. Automorphic background

We need some background on automorphic representations Π\Pi of GLd(𝐀)\operatorname{GL}_{d}(\mathbf{A}_{\mathbb{Q}}) for d1d\geq 1. We will only work with cuspidal Π\Pi’s, or more generally, isobaric Π\Pi’s. These Π\Pi’s have well-defined LL-functions L(s,Π)L(s,\Pi), and good formal properties (due to Rankin, Selberg, Langlands, Godement, Jacquet, Shalika, and others):

  1. (1)

    If Π\Pi is cuspidal, then L(s,Π)L(s,\Pi) is primitive in the sense of [farmer2019analytic]*Lemma 2.4, and has certain familiar analytic properties [farmer2019analytic]*Theorem 3.1.

  2. (2)

    For each isobaric Π\Pi, there is a unique multiset {Π1,,Πr}\{\Pi_{1},\dots,\Pi_{r}\}, consisting of cuspidals, such that L(s,Π)=L(s,Π1)L(s,Πr)L(s,\Pi)=L(s,\Pi_{1})\cdots L(s,\Pi_{r}). We call the Πi\Pi_{i}’s cuspidal constituents of Π\Pi.

  3. (3)

    Strong multiplicity one: If Π\Pi, Π\Pi^{\prime} are isobaric, and Lp(s,Π)=Lp(s,Π)L_{p}(s,\Pi)=L_{p}(s,\Pi^{\prime}) for all but finitely many primes pp, then L(s,Π)=L(s,Π)L(s,\Pi)=L(s,\Pi^{\prime}).

Conjecture 1.2 has a host of standard consequences (which may be treated as a black box).

Proposition 3.2.

Let 𝐜𝒮1\bm{c}\in\mathcal{S}_{1}. Let L(s)L(s) be one of the Hasse–Weil LL-functions in (1.7). Assume Conjecture 1.2 for L(s)L(s) holds for some dd, Π\Pi. Then the following hold:

  1. (1)

    d=degL(s)d=\deg L(s); in particular, dm1d\ll_{m}1.

  2. (2)

    The conductor q(Π)1q(\Pi)\in\mathbb{Z}_{\geq 1} of Π\Pi satisfies q(Π)Δ(𝒄)Om(1)q(\Pi)\mid\Delta(\bm{c})^{O_{m}(1)}.

  3. (3)

    Each cuspidal constituent of Π\Pi has unitary, finite-order central character.

  4. (4)

    L(s)L(s) is holomorphic on \mathbb{C}, except possibly for poles at s=1s=1 corresponding to trivial constituents of Π\Pi.

  5. (5)

    (s1)dL(s)(s-1)^{d}L(s) is an entire function of order 11.

  6. (6)

    L(s)L(s) has a standard functional equation (with critical line Re(s)=1/2\operatorname{Re}(s)=1/2), involving q(Π)q(\Pi) and some root number ε(Π){z:|z|=1}\varepsilon(\Pi)\in\{z\in\mathbb{C}:\lvert z\rvert=1\}.

  7. (7)

    L(s)L(s) has real coefficients, Π\Pi is self-dual, and ε(Π){1,+1}\varepsilon(\Pi)\in\{-1,+1\}.

  8. (8)

    Let ψ\psi be a cuspidal or isobaric constituent of Π\Pi. Then 1/L(s,ψ)m,ϵ𝒄ϵ(1+|s|)ϵ1/L(s,\psi)\ll_{m,\epsilon}\lVert\bm{c}\rVert^{\epsilon}(1+\lvert s\rvert)^{\epsilon} for Re(s)1/2+ϵ\operatorname{Re}(s)\geq 1/2+\epsilon. If μψ(n)\mu_{\psi}(n) denotes the nnth coefficient of the Dirichlet series 1/L(s,ψ)1/L(s,\psi), then 1nNμψ(n)nitm,ϵ𝒄ϵ(1+|t|)ϵN1/2+ϵ\sum_{1\leq n\leq N}\mu_{\psi}(n)n^{-it}\ll_{m,\epsilon}\lVert\bm{c}\rVert^{\epsilon}(1+\lvert t\rvert)^{\epsilon}N^{1/2+\epsilon} for all tt\in\mathbb{R} and N>0N\in\mathbb{R}_{>0}.

Proof.

If pp is a prime, then Lp(s,Π)L_{p}(s,\Pi) has degree d\leq d, with equality if and only if pq(Π)p\nmid q(\Pi); cf. [farmer2019analytic]*Axiom 3(b) and (3.3).

(1): Compare the degrees of Lp(s)L_{p}(s), Lp(s,Π)L_{p}(s,\Pi) at a prime pq(Π)Δ(𝒄)p\nmid q(\Pi)\Delta(\bm{c}).

(2): For some ν\nu\in\mathbb{Z}, the local Dirichlet polynomial Lp(sν/2)L_{p}(s-\nu/2) has rational coefficients for all primes pp. So by [shin2014fields]*(3.2), there exists uu\in\mathbb{R} such that for all pq(Π)p\nmid q(\Pi), the representation Πp(||udet)\Pi_{p}\otimes(\lvert\cdot\rvert^{u}\circ\det) of GLd(p)\operatorname{GL}_{d}(\mathbb{Q}_{p}) has field of rationality \mathbb{Q}, in the sense of [shin2014fields]*Definition 2.2. So by strong multiplicity one, the field of rationality of Π(||udet)\Pi\otimes(\lvert\cdot\rvert^{u}\circ\det) is \mathbb{Q}. Now consider any pq(Π)p\mid q(\Pi). Then Lp(s,Π)L_{p}(s,\Pi) has degree <d<d, so Lp(s)L_{p}(s) has degree <d<d, whence pΔ(𝒄)p\mid\Delta(\bm{c}). By [shin2014fields]*Lemmas 3.11 and 3.13, then, vp(q(Π))d1v_{p}(q(\Pi))\ll_{d}1. So q(Π)Δ(𝒄)Od(1)Δ(𝒄)Om(1)q(\Pi)\mid\Delta(\bm{c})^{O_{d}(1)}\mid\Delta(\bm{c})^{O_{m}(1)}.

(3): Let Π\Pi^{\prime} be a cuspidal constituent of Π\Pi, so Π\Pi^{\prime} is a cuspidal automorphic representation of GLd(𝐀)\operatorname{GL}_{d^{\prime}}(\mathbf{A}_{\mathbb{Q}}) for some d1d^{\prime}\geq 1. Let ω:×\𝐀××\omega^{\prime}\colon\mathbb{Q}^{\times}\backslash\mathbf{A}_{\mathbb{Q}}^{\times}\to\mathbb{C}^{\times} be the central character of Π\Pi^{\prime}. Then ω\omega^{\prime} corresponds to a classical character n|n|zχ(n)n\mapsto\lvert n\rvert^{z}\chi(n) on \mathbb{Z}, where zz\in\mathbb{C} and χ\chi is a Dirichlet character of conductor dividing q(Π)q(\Pi), such that ωp(p)=pzχ(p)\omega^{\prime}_{p}(p)=p^{z}\chi(p) for all primes pq(Π)p\nmid q(\Pi). But at each prime pq(Π)p\nmid q(\Pi), if we write Lp(s,Π)=1jd(1αΠ,j(p)ps)1L_{p}(s,\Pi)=\prod_{1\leq j\leq d^{\prime}}(1-\alpha_{\Pi,j}(p)p^{-s})^{-1}, then ωp(p)=1jdαΠ,j(p)\omega^{\prime}_{p}(p)=\prod_{1\leq j\leq d^{\prime}}\alpha_{\Pi,j}(p); cf. [farmer2019analytic]*(3.3) and its proof. By (3.8) and the algebraicity of the eigenvalues α~j(p)\tilde{\alpha}_{j}(p), it follows that for infinitely many primes pp, we have |pz|=1\lvert p^{z}\rvert=1 and pz¯p^{z}\in\overline{\mathbb{Q}}. So Re(z)=0\operatorname{Re}(z)=0, i.e. ω\omega^{\prime} is unitary; and then Im(z)=0\operatorname{Im}(z)=0 by the six exponentials theorem (cf. [farmer2019analytic]*proof of Lemma 4.9), so ω\omega^{\prime} has finite order.

(4), (5), (6): Use (3) and results of Godement and Jacquet; cf. [farmer2019analytic]*Theorem 3.1.

(7): For some ν\nu\in\mathbb{Z}, the coefficients of L(sν/2)L(s-\nu/2) are all rational. Hence L(s)L(s) has real coefficients. So by (3) and strong multiplicity one, we have L(s,Π)=L(s,Π)L(s,\Pi)=L(s,\Pi^{\vee}) and thus Π\Pi is self-dual. The functional equation from (6) then implies ε(Π)\varepsilon(\Pi)\in\mathbb{R}, so ε(Π)=±1\varepsilon(\Pi)=\pm 1.

(8): By (3.2), (5), (6), and Conjecture 1.2(2), we have 1/L(s,ψ)d,ϵq(ψ)ϵ(1+|s|)ϵ1/L(s,\psi)\ll_{d,\epsilon}q(\psi)^{\epsilon}(1+\lvert s\rvert)^{\epsilon} for Re(s)1/2+ϵ\operatorname{Re}(s)\geq 1/2+\epsilon; see e.g. [iwaniec2004analytic]*Theorem 5.19 and the ensuing paragraph. But q(ψ)q(Π)Δ(𝒄)Om(1)q(\psi)\mid q(\Pi)\mid\Delta(\bm{c})^{O_{m}(1)}, so the desired bound on 1/L(s,ψ)1/L(s,\psi) follows. One can then prove 1nNμψ(n)nitm,ϵ𝒄ϵ(1+|t|)ϵN1/2+ϵ\sum_{1\leq n\leq N}\mu_{\psi}(n)n^{-it}\ll_{m,\epsilon}\lVert\bm{c}\rVert^{\epsilon}(1+\lvert t\rvert)^{\epsilon}N^{1/2+\epsilon} using Perron’s formula ([iwaniec2004analytic]*Proposition 5.54) and contour integration; cf. [hooley1986HasseWeil]*p. 75, proof of Lemma 10. ∎

4. Local control on polynomials and LL-functions

We first recall some standard bounds on the local near-zero loci of a fixed polynomial; cf. [serre1981quelques]*p. 146, (57) and [ganzburg2001polynomial]. Given f[x]f\in\mathbb{Z}[x], let N(f;q)N(f;q) be the number of solutions xmodqx\bmod{q} to f(x)0modqf(x)\equiv 0\bmod{q}, and let μ(f;λ)\mu_{\mathbb{R}}(f;\lambda) be the Lebesgue measure of the set {x[1,1]:|f(x)|λ}\{x\in[-1,1]:\lvert f(x)\rvert\leq\lambda\}. Similarly, for P[y1,,yn]P\in\mathbb{Z}[y_{1},\dots,y_{n}], define

N(P;q)\colonequals#{(y1,,yn)(/q)n:P(y1,,yn)0modq},μ(P;λ)\colonequalsvol{(y1,,yn)[1,1]n:|P(y1,,yn)|λ}.\begin{split}N(P;q)&\colonequals\#{\{(y_{1},\dots,y_{n})\in(\mathbb{Z}/q\mathbb{Z})^{n}:P(y_{1},\dots,y_{n})\equiv 0\bmod{q}\}},\\ \mu_{\mathbb{R}}(P;\lambda)&\colonequals\operatorname{vol}{\{(y_{1},\dots,y_{n})\in[-1,1]^{n}:\lvert P(y_{1},\dots,y_{n})\rvert\leq\lambda\}}.\end{split}
Proposition 4.1.

Suppose f[x]f\in\mathbb{Z}[x] has leading term axdax^{d} with a0a\neq 0 and d1d\geq 1. Then

(4.1) N(f;q)d|a|1/dq11/d,N(f;q)\ll_{d}\lvert a\rvert^{1/d}q^{1-1/d},

uniformly over integers q1q\geq 1. Also, uniformly over reals λ>0\lambda>0, we have

(4.2) μ(f;λ)d|a|1/dλ1/d.\mu_{\mathbb{R}}(f;\lambda)\ll_{d}\lvert a\rvert^{-1/d}\lambda^{1/d}.
Proof.

The bound (4.2) goes back to Pólya (see e.g. [ganzburg2001polynomial]*Theorem 1.1). Now let hh denote the greatest common divisor of qq and the coefficients of ff. The bound (4.1) follows from [konjagin1979number] if h=1h=1, and then in general from the inequality N(f;q)hN(f/h;q/h)N(f;q)\leq h\cdot N(f/h;q/h). ∎

Corollary 4.2.

Fix a nonconstant polynomial P[y1,,yn]P\in\mathbb{Z}[y_{1},\dots,y_{n}], where n1n\geq 1. Then

(4.3) N(P;q)Pqn1/degP,N(P;q)\ll_{P}q^{n-1/\deg{P}},

uniformly over integers q1q\geq 1. Also, uniformly over reals λ>0\lambda>0, we have

(4.4) μ(P;λ)Pλ1/degP.\mu_{\mathbb{R}}(P;\lambda)\ll_{P}\lambda^{1/\deg{P}}.
Proof.

Let d=degPd=\deg{P}. Given a matrix AGLn()A\in\operatorname{GL}_{n}(\mathbb{Q}) with integral entries, let Q=(detA)dPA1[y1,,yn]Q=(\det{A})^{d}\cdot P\circ A^{-1}\in\mathbb{Z}[y_{1},\dots,y_{n}]. Via the \mathbb{Z}-linear map 𝒚A𝒚\bm{y}\mapsto A\bm{y}, we have N(P;q)AN(Q;q)N(P;q)\ll_{A}N(Q;q) and μ(P;λ)Aμ(Q;λ)\mu_{\mathbb{R}}(P;\lambda)\ll_{A}\mu_{\mathbb{R}}(Q;\lambda). By [eisenbud1995commutative]*p. 283, proof of Lemma 13.2.c, we may choose AA so that QQ has y1y_{1}-leading term a1y1da_{1}y_{1}^{d}, with a10a_{1}\neq 0. Fix y2,,yny_{2},\dots,y_{n}; then #{y1:Q(𝒚)0}d,a1q11/d\#\{y_{1}:Q(\bm{y})\equiv 0\}\ll_{d,a_{1}}q^{1-1/d} by (4.1). Summing over y2,,yny_{2},\dots,y_{n} gives (4.3). Similarly, (4.2) implies (4.4). ∎

We now turn to local LL-factors. (For 𝒄pm\bm{c}\in\mathbb{Z}_{p}^{m} with Δ(𝒄)0\Delta(\bm{c})\neq 0, we define Lp(s,V𝒄)L_{p}(s,V_{\bm{c}}) using Remark 3.1.)

Proposition 4.3 ([kisin1999local]).

Fix a prime pp and a tuple 𝐛pm\bm{b}\in\mathbb{Z}_{p}^{m} with Δ(𝐛)0\Delta(\bm{b})\neq 0. Then there exists an integer l0l\geq 0, depending only on pp and 𝐛\bm{b}, such that for all tuples 𝐚pm\bm{a}\in\mathbb{Z}_{p}^{m} with 𝐚𝐛modp1+l\bm{a}\equiv\bm{b}\bmod{p^{1+l}}, we have Δ(𝐚)0\Delta(\bm{a})\neq 0 and Lp(s,V𝐚)=Lp(s,V𝐛)L_{p}(s,V_{\bm{a}})=L_{p}(s,V_{\bm{b}}).

Proof.

Let SS be the open subscheme Δ(𝒄)0\Delta(\bm{c})\neq 0 of 𝔸pm\mathbb{A}^{m}_{\mathbb{Q}_{p}}. Let WW be the closed subscheme F(𝒙)=𝒄𝒙=0F(\bm{x})=\bm{c}\cdot\bm{x}=0 of Sm1\mathbb{P}^{m-1}_{S}. Let p\ell\neq p be a prime. The maps f1:WSf_{1}\colon W\to S and f2:Sm1Sf_{2}\colon\mathbb{P}^{m-1}_{S}\to S induce local systems j\colonequalsRm(fj)()\mathcal{L}_{j}\colonequals R^{m_{\ast}}(f_{j})_{\ast}(\mathbb{Z}_{\ell}) on SS. By [kisin1999local]*Theorem 5.1, case (2), and its proof, there exists a pp-adic neighborhood UU of 𝒃\bm{b} in SS such that the Galois representations ρj,𝒂:GpAut(j,𝒂)\rho_{j,\bm{a}}\colon G_{\mathbb{Q}_{p}}\to\operatorname{Aut}(\mathcal{L}_{j,\bm{a}}) for 𝒂U(p)\bm{a}\in U(\mathbb{Q}_{p}) all factor through π1(U)\pi_{1}(U) in an appropriate sense. So the isomorphism class of the representation GpAut((1,𝒂/2,𝒂))G_{\mathbb{Q}_{p}}\to\operatorname{Aut}(\mathbb{Q}_{\ell}\otimes(\mathcal{L}_{1,\bm{a}}/\mathcal{L}_{2,\bm{a}})) is constant over 𝒂U(p)\bm{a}\in U(\mathbb{Q}_{p}), and thus Lp(s,V𝒂)L_{p}(s,V_{\bm{a}}) is too. ∎

Lemma 4.4.

Fix a prime pp. For each 𝐛pm\bm{b}\in\mathbb{Z}_{p}^{m} with Δ(𝐛)0\Delta(\bm{b})\neq 0, let l(p,𝐛)0l(p,\bm{b})\geq 0 be the smallest integer for which the conclusion of Proposition 4.3 holds. Then for each integer l0l\geq 0, the set

(4.5) {𝒃pm:Δ(𝒃)0,l(p,𝒃)l}\{\bm{b}\in\mathbb{Z}_{p}^{m}:\Delta(\bm{b})\neq 0,\;l(p,\bm{b})\leq l\}

is invariant under translation by any element of pl+1pmp^{l+1}\mathbb{Z}_{p}^{m}. Furthermore, the measure of (4.5) tends to 11 as ll\to\infty. (Here we use the usual Haar measure on pm\mathbb{Z}_{p}^{m}.)

Proof.

Suppose 𝒃,𝒄pm\bm{b},\bm{c}\in\mathbb{Z}_{p}^{m} with Δ(𝒃)0\Delta(\bm{b})\neq 0 and 𝒄𝒃modp1+l(p,𝒃)\bm{c}\equiv\bm{b}\bmod{p^{1+l(p,\bm{b})}}. Then Δ(𝒄)0\Delta(\bm{c})\neq 0 and Lp(s,V𝒄)=Lp(s,V𝒃)L_{p}(s,V_{\bm{c}})=L_{p}(s,V_{\bm{b}}), and thus l(p,𝒄)l(p,𝒃)l(p,\bm{c})\leq l(p,\bm{b}) (because for every 𝒂𝒄modp1+l(p,𝒃)\bm{a}\equiv\bm{c}\bmod{p^{1+l(p,\bm{b})}}, we have 𝒂𝒃modp1+l(p,𝒃)\bm{a}\equiv\bm{b}\bmod{p^{1+l(p,\bm{b})}} and thus Δ(𝒂)0\Delta(\bm{a})\neq 0 and Lp(s,V𝒂)=Lp(s,V𝒃)=Lp(s,V𝒄)L_{p}(s,V_{\bm{a}})=L_{p}(s,V_{\bm{b}})=L_{p}(s,V_{\bm{c}})). But then 𝒃𝒄modp1+l(p,𝒄)\bm{b}\equiv\bm{c}\bmod{p^{1+l(p,\bm{c})}}, so a similar argument gives l(p,𝒃)l(p,𝒄)l(p,\bm{b})\leq l(p,\bm{c}), whence

(4.6) l(p,𝒃)=l(p,𝒄).l(p,\bm{b})=l(p,\bm{c}).

Therefore, if 𝒃\bm{b} lies in (4.5) for some l0l\geq 0, then (4.5) indeed contains the set 𝒃+pl+1pm\bm{b}+p^{l+1}\mathbb{Z}_{p}^{m}.

Next, let A0A\in\mathbb{Z}_{\geq 0}. The set SA={𝒄pm:vp(Δ(𝒄))A}S_{A}=\{\bm{c}\in\mathbb{Z}_{p}^{m}:v_{p}(\Delta(\bm{c}))\leq A\} is closed in pm\mathbb{Z}_{p}^{m}, and thus compact. The function 𝒃l(p,𝒃)\bm{b}\mapsto l(p,\bm{b}) on SAS_{A} is locally constant (by (4.6)), and thus has a (finite) maximum value. Therefore, for all ll sufficiently large in terms of AA, the set (4.5) contains SAS_{A}. But by (4.3) (or by [serre1981quelques]*p. 146, Corollaire), the measure of SAS_{A} tends to 11 as AA\to\infty. ∎

Remark 4.5.

If Conjecture 1.11 holds, then l(p,𝒃)vp(H(𝒃))l(p,\bm{b})\leq v_{p}(H(\bm{b})) (whenever H(𝒃)0H(\bm{b})\neq 0).

5. General separation technique

At several points in the paper, we need to understand quantities that vary with 𝒄\bm{c} and nn. A key tool we use for this is smooth dyadic decomposition (minimizing convergence issues) followed by separation of variables (via Mellin inversion); see Lemma 5.2.

Let d×r\colonequalsdr/rd^{\times}{r}\colonequals dr/r and logr\colonequalsrr\partial_{\log{r}}\colonequals r\cdot\partial_{r} for r>0r\in\mathbb{R}_{>0}. For any k1k\in\mathbb{Z}_{\geq 1} and 𝒓>0k\bm{r}\in\mathbb{R}_{>0}^{k}, let d×𝒓=d×r1d×rkd^{\times}{\bm{r}}=d^{\times}{r_{1}}\cdots d^{\times}{r_{k}} and log𝒓𝜶\colonequalslogr1α1logrkαk\partial_{\log{\bm{r}}}^{\bm{\alpha}}\colonequals\partial_{\log{r_{1}}}^{\alpha_{1}}\cdots\partial_{\log{r_{k}}}^{\alpha_{k}}. Given gCc(>0k)g\in C^{\infty}_{c}(\mathbb{R}_{>0}^{k})\otimes\mathbb{C} and 𝒔k\bm{s}\in\mathbb{C}^{k}, let

(5.1) g(𝒔)\colonequals𝒓>0d×𝒓g(𝒓)𝒓𝒔=r1,,rk>0d×r1d×rkg(r1,,rk)r1s1rksk,g^{\vee}(\bm{s})\colonequals\int_{\bm{r}>0}d^{\times}\bm{r}\,g(\bm{r})\bm{r}^{\bm{s}}=\int_{r_{1},\dots,r_{k}>0}d^{\times}{r_{1}}\cdots d^{\times}{r_{k}}\,g(r_{1},\dots,r_{k})r_{1}^{s_{1}}\cdots r_{k}^{s_{k}},

so that Mellin inversion (see e.g. [iwaniec2004analytic]*p. 90, (4.106)) gives (for all 𝝈k\bm{\sigma}\in\mathbb{R}^{k})

(5.2) g(𝒓)=(2π)k𝒕k𝑑𝒕g(𝝈+i𝒕)𝒓𝝈i𝒕.g(\bm{r})=(2\pi)^{-k}\int_{\bm{t}\in\mathbb{R}^{k}}d\bm{t}\,g^{\vee}(\bm{\sigma}+i\bm{t})\cdot\bm{r}^{-\bm{\sigma}-i\bm{t}}.
Proposition 5.1 (Standard Mellin bound).

Fix a compact set I>0I\subseteq\mathbb{R}_{>0}. Let k1k\in\mathbb{Z}_{\geq 1} and (𝐌,𝐬)>0k×k(\bm{M},\bm{s})\in\mathbb{R}_{>0}^{k}\times\mathbb{C}^{k}. Let gCc(>0k)g\in C^{\infty}_{c}(\mathbb{R}_{>0}^{k})\otimes\mathbb{C} with Suppg1ik(MiI)\operatorname{Supp}{g}\subseteq\prod_{1\leq i\leq k}(M_{i}\cdot I). Then

(5.3) g(𝒔)k,bOI(1)kgexpb,(1+𝒔)b1ik(OI(1)|Re(si)|MiRe(si)),g^{\vee}(\bm{s})\ll_{k,b}O_{I}(1)^{k}\cdot\frac{\lVert g\circ\exp\rVert_{b,\infty}}{(1+\lVert\bm{s}\rVert)^{b}}\cdot\prod_{1\leq i\leq k}(O_{I}(1)^{\lvert\operatorname{Re}(s_{i})\rvert}M_{i}^{\operatorname{Re}(s_{i})}),

for all b0b\in\mathbb{Z}_{\geq 0}. Here gexpb,\colonequals|𝛂|bsup𝐫>0k|log𝐫𝛂g(𝐫)|\lVert g\circ\exp\rVert_{b,\infty}\colonequals\sum_{\lvert\bm{\alpha}\rvert\leq b}\sup_{\bm{r}\in\mathbb{R}_{>0}^{k}}{\lvert\partial_{\log\bm{r}}^{\bm{\alpha}}g(\bm{r})\rvert}.

Proof.

By (5.1), and the scale invariance of d×rid^{\times}{r_{i}}, logri\partial_{\log{r_{i}}} for each ii, we can reduce to the case where M1==Mk=1M_{1}=\cdots=M_{k}=1. Now let vol(logI)\colonequalsr>0d×r 1rI<\operatorname{vol}(\log{I})\colonequals\int_{r>0}d^{\times}{r}\,\bm{1}_{r\in I}<\infty. If 𝒔1\lVert\bm{s}\rVert\leq 1, we may assume b=0b=0. Now in general, choose ii with |si|=max(|s1|,,|sd|)\lvert s_{i}\rvert=\max(\lvert s_{1}\rvert,\dots,\lvert s_{d}\rvert). Then on the right-hand side of (5.1), integrate by parts bb times in logri\log{r_{i}}, to rewrite g(𝒔)g^{\vee}(\bm{s}) as

(1)b𝒓>0d×𝒓𝒓𝒔siblogribg(𝒓)vol(logI)kgexpb,|si|b1jkOSuppI(1)|Re(sj)|.(-1)^{b}\int_{\bm{r}>0}d^{\times}{\bm{r}}\,\frac{\bm{r}^{\bm{s}}}{s_{i}^{b}}\partial_{\log{r_{i}}}^{b}g(\bm{r})\ll\operatorname{vol}(\log{I})^{k}\frac{\lVert g\circ\exp\rVert_{b,\infty}}{\lvert s_{i}\rvert^{b}}\prod_{1\leq j\leq k}O_{\operatorname{Supp}{I}}(1)^{\lvert\operatorname{Re}(s_{j})\rvert}.

This suffices for (5.3). ∎

Fix a function ν2Cc(>0)\nu_{2}\in C^{\infty}_{c}(\mathbb{R}_{>0}), supported on [1,2][1,2], with

(5.4) r>0d×rν2(r)2=1.\int_{r>0}d^{\times}{r}\,\nu_{2}(r)^{2}=1.
Lemma 5.2 (“Dyadic partial Mellin summation”).

Let k1k\in\mathbb{Z}_{\geq 1}. Let a:1ka\colon\mathbb{Z}_{\geq 1}^{k}\to\mathbb{C} be a function. Let f:>0kf\colon\mathbb{R}_{>0}^{k}\to\mathbb{C} be a smooth function supported on 1jkrjA\prod_{1\leq j\leq k}r_{j}\leq A for some real A1A\geq 1. Let ν=ν2\nu=\nu_{2}; let ν(𝐫/𝐍)\colonequals1jkν(rj/Nj)\nu(\bm{r}/\bm{N})\colonequals\prod_{1\leq j\leq k}\nu(r_{j}/N_{j}) and g𝐍(𝐫)\colonequalsf(𝐫)ν(𝐫/𝐍)g_{\bm{N}}(\bm{r})\colonequals f(\bm{r})\nu(\bm{r}/\bm{N}). Then

(5.5) 𝒏1a(𝒏)f(𝒏)=(2π)k𝑵[1/2,)kd×𝑵𝒕k𝑑𝒕g𝑵(i𝒕)𝒏1ν(𝒏/𝑵)a(𝒏)𝒏i𝒕.\sum_{\bm{n}\geq 1}a(\bm{n})f(\bm{n})=(2\pi)^{-k}\int_{\bm{N}\in[1/2,\infty)^{k}}d^{\times}\bm{N}\int_{\bm{t}\in\mathbb{R}^{k}}d\bm{t}\,g_{\bm{N}}^{\vee}(i\bm{t})\sum_{\bm{n}\geq 1}\nu(\bm{n}/\bm{N})a(\bm{n})\bm{n}^{i\bm{t}}.
Proof.

For all 𝒏1\bm{n}\geq 1, we have f(𝒏)=𝑵1/2d×𝑵ν(𝒏/𝑵)g𝑵(𝒏)f(\bm{n})=\int_{\bm{N}\geq 1/2}d^{\times}{\bm{N}}\,\nu(\bm{n}/\bm{N})g_{\bm{N}}(\bm{n}) by (5.4) (since ν(𝒏/𝑵)=0\nu(\bm{n}/\bm{N})=0 for 𝑵<1/2\bm{N}<1/2). Also, g𝑵(𝒓)Cc(>0k)g_{\bm{N}}(\bm{r})\in C^{\infty}_{c}(\mathbb{R}_{>0}^{k})\otimes\mathbb{C} for each 𝑵1/2\bm{N}\geq 1/2. So by (5.2), we get

(5.6) 𝒏1a(𝒏)f(𝒏)=𝒏1a(𝒏)𝑵1/2d×𝑵ν(𝒏/𝑵)(2π)k𝒕k𝑑𝒕g𝑵(i𝒕)𝒏i𝒕.\sum_{\bm{n}\geq 1}a(\bm{n})f(\bm{n})=\sum_{\bm{n}\geq 1}a(\bm{n})\int_{\bm{N}\geq 1/2}d^{\times}\bm{N}\,\nu(\bm{n}/\bm{N})(2\pi)^{-k}\int_{\bm{t}\in\mathbb{R}^{k}}d\bm{t}\,g_{\bm{N}}^{\vee}(i\bm{t})\bm{n}^{i\bm{t}}.

If 𝒓Suppg𝑵\bm{r}\in\operatorname{Supp}{g_{\bm{N}}}, then jrjA\prod_{j}r_{j}\leq A and 𝒓j[Nj,2Nj]\bm{r}\in\prod_{j}[N_{j},2N_{j}], so jNjA\prod_{j}N_{j}\leq A. Therefore, there exist compact sets 𝒦1,𝒦2>0k\mathcal{K}_{1},\mathcal{K}_{2}\subseteq\mathbb{R}_{>0}^{k} such that if (𝒏,𝑵)𝒦1×𝒦2(\bm{n},\bm{N})\notin\mathcal{K}_{1}\times\mathcal{K}_{2} in (5.6), then ν(𝒏/𝑵)g𝑵(i𝒕)=0\nu(\bm{n}/\bm{N})\cdot g_{\bm{N}}^{\vee}(i\bm{t})=0. Furthermore, there exists a compact set 𝒦3>0k\mathcal{K}_{3}\subseteq\mathbb{R}_{>0}^{k} such that if 𝑵𝒦2\bm{N}\in\mathcal{K}_{2}, then Suppg𝑵𝒦3\operatorname{Supp}{g_{\bm{N}}}\subseteq\mathcal{K}_{3}. So Proposition 5.1 gives g𝑵(i𝒕)f,ν,b(1+𝒕)bg_{\bm{N}}^{\vee}(i\bm{t})\ll_{f,\nu,b}(1+\lVert\bm{t}\rVert)^{-b} for all b0b\geq 0, uniformly over 𝑵𝒦2\bm{N}\in\mathcal{K}_{2}. Thus 𝒏1𝑵1/2d×𝑵𝒕k𝑑𝒕|a(𝒏)||ν(𝒏/𝑵)||g𝑵(i𝒕)|\sum_{\bm{n}\geq 1}\int_{\bm{N}\geq 1/2}d^{\times}\bm{N}\int_{\bm{t}\in\mathbb{R}^{k}}d\bm{t}\,\lvert a(\bm{n})\rvert\cdot\lvert\nu(\bm{n}/\bm{N})\rvert\cdot\lvert g_{\bm{N}}^{\vee}(i\bm{t})\rvert is a,f,ν,b𝒏𝒦1𝑵𝒦2d×𝑵<\ll_{a,f,\nu,b}\sum_{\bm{n}\in\mathcal{K}_{1}\cap\mathbb{Z}}\int_{\bm{N}\in\mathcal{K}_{2}}d^{\times}\bm{N}<\infty, provided b>kb>k. Now (5.5) follows from (5.6) by Fubini. ∎

For later use, we now do a general dyadic calculation.

Lemma 5.3.

Let q,a,bq,a,b\in\mathbb{R} with aq<ba\leq q<b. Let r1,r2,τ>0r_{1},r_{2},\tau\in\mathbb{R}_{>0}. Then

r1rr2:log2(r)rqra+τbarb,r1rr2d×rrqra+τbarb,\sum_{r_{1}\leq r\leq r_{2}:\,\log_{2}(r)\in\mathbb{Z}}\frac{r^{q}}{r^{a}+\tau^{b-a}r^{b}},\quad\int_{r_{1}\leq r\leq r_{2}}d^{\times}{r}\,\frac{r^{q}}{r^{a}+\tau^{b-a}r^{b}},

are both q,a,bmin(τ1,r2)qa(log(1+r2/r1))𝟏q=a\ll_{q,a,b}\min(\tau^{-1},r_{2})^{q-a}\cdot(\log(1+r_{2}/r_{1}))^{\bm{1}_{q=a}}.

Proof.

We may assume r1r2r_{1}\leq r_{2}, or else the sum and integral both vanish. By subtracting qq, aa, bb by aa, we may also assume a=0a=0. If r2τ1r_{2}\leq\tau^{-1}, then the result follows from the bound 1+τbrb11+\tau^{b}r^{b}\geq 1 and a geometric series (or corresponding integral). If r2>τ1r_{2}>\tau^{-1}, then separately considering rτ1r\leq\tau^{-1} and r>τ1r>\tau^{-1} leads to the result. ∎

6. Statistics of families of LL-functions

Throughout §6, assume mm is even. Recall the eigenvalue and coefficient notation from §3.

6.1. Computing local averages

For convenience, let α~𝒄,j(p)\colonequalsα~V𝒄,j(p)\tilde{\alpha}_{\bm{c},j}(p)\colonequals\tilde{\alpha}_{V_{\bm{c}},j}(p) and λ𝒄(n)\colonequalsλV𝒄(n)\lambda^{\natural}_{\bm{c}}(n)\colonequals\lambda^{\natural}_{V_{\bm{c}}}(n). Let μ𝒄(n)\mu_{\bm{c}}(n) be the nnth coefficient of the Dirichlet series 1/L(s,V𝒄)1/L(s,V_{\bm{c}}). We have

(6.1) 1/Lp(s,V𝒄)=j(1α~𝒄,j(p)ps)1/L_{p}(s,V_{\bm{c}})={\textstyle\prod_{j}(1-\tilde{\alpha}_{\bm{c},j}(p)p^{-s})}

by (3.1). So if pΔ(𝒄)p\nmid\Delta(\bm{c}), then (3.4), (3.6), (3.7), and 2m2\nmid m_{\ast} imply

(6.2) μ𝒄(p)=λ𝒄(p)=E𝒄(p),μ𝒄(p2)=λV𝒄,2(p)=12(E𝒄(p)2+E𝒄(p2)).\mu_{\bm{c}}(p)=-\lambda^{\natural}_{\bm{c}}(p)=E^{\natural}_{\bm{c}}(p),\quad\mu_{\bm{c}}(p^{2})=\lambda^{\natural}_{V_{\bm{c}},\bigwedge^{2}}(p)=\tfrac{1}{2}(E^{\natural}_{\bm{c}}(p)^{2}+E^{\natural}_{\bm{c}}(p^{2})).

The local statistics of λ𝒄(n)\lambda^{\natural}_{\bm{c}}(n), μ𝒄(n)\mu_{\bm{c}}(n) over 𝒄\bm{c} play a basic role in the global statistics of L(s,V𝒄)L(s,V_{\bm{c}}) over 𝒄\bm{c}. To prove that certain averages exist, we will use (3.2) and Lemma 4.4. But to estimate said averages, we will take a point-counting approach (though one could use monodromy groups instead; see e.g. [sarnak2016families]*§2.11). The result is Proposition 6.1 below.

Let m\mathcal{B}\subseteq\mathbb{R}^{m} be a region of the form I1××ImI_{1}\times\cdots\times I_{m}, where I1,,ImI_{1},\dots,I_{m}\subseteq\mathbb{R} are compact intervals of positive length. Let 𝒂m\bm{a}\in\mathbb{Z}^{m} and n0,n,n1,n21n_{0},n,n_{1},n_{2}\in\mathbb{Z}_{\geq 1}. Let 𝔼𝒄S𝒂,n0[f]\mathbb{E}^{\bm{a},n_{0}}_{\bm{c}\in S}[f] be the average of ff over {𝒄S:𝒄𝒂modn0}\{\bm{c}\in S:\bm{c}\equiv\bm{a}\bmod{n_{0}}\} (assuming this set is nonempty). Let 𝔼1𝒄n𝒂,n0[f]\colonequals𝔼𝒄{1,2,,n}m𝒂,n0[f]\mathbb{E}^{\bm{a},n_{0}}_{1\leq\bm{c}\leq n}[f]\colonequals\mathbb{E}^{\bm{a},n_{0}}_{\bm{c}\in\{1,2,\dots,n\}^{m}}[f].

Proposition 6.1 (LocAv).

The following two limits exist, and are independent of \mathcal{B}:

μ¯F,1𝒂,n0(n)\colonequalslimZ𝔼𝒄𝒮1Z𝒂,n0[μ𝒄(n)],μ¯F,2𝒂,n0(n1,n2)\colonequalslimZ𝔼𝒄𝒮1Z𝒂,n0[μ𝒄(n1)μ𝒄(n2)].\begin{split}\bar{\mu}_{F,1}^{\bm{a},n_{0}}(n)&\colonequals\lim_{Z\to\infty}\mathbb{E}^{\bm{a},n_{0}}_{\bm{c}\in\mathcal{S}_{1}\cap Z\cdot\mathcal{B}}[\mu_{\bm{c}}(n)],\\ \bar{\mu}_{F,2}^{\bm{a},n_{0}}(n_{1},n_{2})&\colonequals\lim_{Z\to\infty}\mathbb{E}^{\bm{a},n_{0}}_{\bm{c}\in\mathcal{S}_{1}\cap Z\cdot\mathcal{B}}[\mu_{\bm{c}}(n_{1})\mu_{\bm{c}}(n_{2})].\end{split}

The quantity μ¯F,1𝐚,n0(n)\bar{\mu}_{F,1}^{\bm{a},n_{0}}(n) is multiplicative in nn: if gcd(n,n)=1\gcd(n,n^{\prime})=1, then

(6.3) μ¯F,1𝒂,n0(n)μ¯F,1𝒂,n0(n)=μ¯F,1𝒂,n0(nn).\bar{\mu}_{F,1}^{\bm{a},n_{0}}(n)\bar{\mu}_{F,1}^{\bm{a},n_{0}}(n^{\prime})=\bar{\mu}_{F,1}^{\bm{a},n_{0}}(nn^{\prime}).

The quantity μ¯F,2𝐚,n0(n1,n2)\bar{\mu}_{F,2}^{\bm{a},n_{0}}(n_{1},n_{2}) is multiplicative in (n1,n2)(n_{1},n_{2}): if gcd(n1n2,n1n2)=1\gcd(n_{1}n_{2},n^{\prime}_{1}n^{\prime}_{2})=1, then

(6.4) μ¯F,2𝒂,n0(n1,n2)μ¯F,2𝒂,n0(n1,n2)=μ¯F,2𝒂,n0(n1n1,n2n2).\bar{\mu}_{F,2}^{\bm{a},n_{0}}(n_{1},n_{2})\bar{\mu}_{F,2}^{\bm{a},n_{0}}(n^{\prime}_{1},n^{\prime}_{2})=\bar{\mu}_{F,2}^{\bm{a},n_{0}}(n_{1}n^{\prime}_{1},n_{2}n^{\prime}_{2}).

Now let pp be a prime, and let l,l1,l20l,l_{1},l_{2}\geq 0 be integers. Then (uniformly over p,l,l1,l2,𝐚,n0p,l,l_{1},l_{2},\bm{a},n_{0})

(6.5) μ¯F,1𝒂,n0(pl)ϵplϵ,μ¯F,2𝒂,n0(pl1,pl2)ϵp(l1+l2)ϵ.\bar{\mu}_{F,1}^{\bm{a},n_{0}}(p^{l})\ll_{\epsilon}p^{l\epsilon},\quad\bar{\mu}_{F,2}^{\bm{a},n_{0}}(p^{l_{1}},p^{l_{2}})\ll_{\epsilon}p^{(l_{1}+l_{2})\epsilon}.

Furthermore, if pn0p\nmid n_{0}, then

(6.6) μ¯F,1𝒂,n0(p)=λV(p)p1/2+O(p1),μ¯F,1𝒂,n0(p2)=1+O(p1),\displaystyle\bar{\mu}_{F,1}^{\bm{a},n_{0}}(p)=\lambda^{\natural}_{V}(p)p^{-1/2}+O(p^{-1}),\quad\bar{\mu}_{F,1}^{\bm{a},n_{0}}(p^{2})=1+O(p^{-1}),
(6.7) μ¯F,2𝒂,n0(pl,1)=μ¯F,2𝒂,n0(1,pl)=μ¯F,1𝒂,n0(pl),μ¯F,2𝒂,n0(p,p)=1+O(p1).\displaystyle\bar{\mu}_{F,2}^{\bm{a},n_{0}}(p^{l},1)=\bar{\mu}_{F,2}^{\bm{a},n_{0}}(1,p^{l})=\bar{\mu}_{F,1}^{\bm{a},n_{0}}(p^{l}),\quad\bar{\mu}_{F,2}^{\bm{a},n_{0}}(p,p)=1+O(p^{-1}).
Proof.

For convenience, define μ𝒄(n)𝟏Δ(𝒄)0\mu_{\bm{c}}(n)\bm{1}_{\Delta(\bm{c})\neq 0} to be μ𝒄(n)\mu_{\bm{c}}(n) if Δ(𝒄)0\Delta(\bm{c})\neq 0, and 0 if Δ(𝒄)=0\Delta(\bm{c})=0. (Here we allow 𝒄m\bm{c}\in\mathbb{Z}^{m}, or more generally, 𝒄pnpm\bm{c}\in\prod_{p\mid n}\mathbb{Z}_{p}^{m}.) By (3.2) and (6.1), we have

(6.8) μ𝒄(pl)𝟏Δ(𝒄)0m1,μ𝒄(n)𝟏Δ(𝒄)0ϵnϵ.\mu_{\bm{c}}(p^{l})\bm{1}_{\Delta(\bm{c})\neq 0}\ll_{m}1,\quad\mu_{\bm{c}}(n)\bm{1}_{\Delta(\bm{c})\neq 0}\ll_{\epsilon}n^{\epsilon}.

(In contrast, for λ𝒄\lambda^{\natural}_{\bm{c}}, we have λ𝒄(pl)𝟏Δ(𝒄)0(l+1)Om(1)m,ϵplϵ\lambda^{\natural}_{\bm{c}}(p^{l})\bm{1}_{\Delta(\bm{c})\neq 0}\ll(l+1)^{O_{m}(1)}\ll_{m,\epsilon}p^{l\epsilon} and λ𝒄(n)𝟏Δ(𝒄)0ϵnϵ\lambda^{\natural}_{\bm{c}}(n)\bm{1}_{\Delta(\bm{c})\neq 0}\ll_{\epsilon}n^{\epsilon}.)

Since μ𝒄(1)𝟏Δ(𝒄)0=𝟏Δ(𝒄)0\mu_{\bm{c}}(1)\bm{1}_{\Delta(\bm{c})\neq 0}=\bm{1}_{\Delta(\bm{c})\neq 0}, we have μ¯F,1𝒂,n0(1)=μ¯F,2𝒂,n0(1,1)=1\bar{\mu}_{F,1}^{\bm{a},n_{0}}(1)=\bar{\mu}_{F,2}^{\bm{a},n_{0}}(1,1)=1 (since |𝒮0Z|=o;Z(Zm)\lvert\mathcal{S}_{0}\cap Z\cdot\mathcal{B}\rvert=o_{\mathcal{B};Z\to\infty}(Z^{m}) by (2.4)). On the other hand, if n2n\geq 2 and k1k\geq 1, then by Lemma 4.4, the quantity μ𝒄(n)𝟏Δ(𝒄)0\mu_{\bm{c}}(n)\bm{1}_{\Delta(\bm{c})\neq 0} depends only on 𝒄modnk\bm{c}\bmod{n^{k}}, unless 𝒄\bm{c} lies in one of on;k(nkm)o_{n;k\to\infty}(n^{km}) exceptional residue classes of m\mathbb{Z}^{m} (or pnpm\prod_{p\mid n}\mathbb{Z}_{p}^{m}) modulo nkn^{k}. This, together with (6.8) and the Chinese remainder theorem, implies that (for any n1n\geq 1) the three quantities

μ¯F,1𝒂,n0(n),limk𝔼1𝒄nk𝒂,n0[μ𝒄(n)𝟏Δ(𝒄)0],pn𝔼𝒄pm𝒂,n0[μ𝒄(pvp(n))𝟏Δ(𝒄)0]\bar{\mu}_{F,1}^{\bm{a},n_{0}}(n),\quad\lim_{k\to\infty}\mathbb{E}^{\bm{a},n_{0}}_{1\leq\bm{c}\leq n^{k}}[\mu_{\bm{c}}(n)\bm{1}_{\Delta(\bm{c})\neq 0}],\quad\prod_{p\mid n}\mathbb{E}^{\bm{a},n_{0}}_{\bm{c}\in\mathbb{Z}_{p}^{m}}[\mu_{\bm{c}}(p^{v_{p}(n)})\bm{1}_{\Delta(\bm{c})\neq 0}]

all exist and equal one another. Similarly, the following exist and equal one another:

μ¯F,2𝒂,n0(n1,n2),limk𝔼1𝒄(n1n2)k𝒂,n0[μ𝒄(n1)μ𝒄(n2)𝟏Δ(𝒄)0],pn1n2𝔼𝒄pm𝒂,n0[μ𝒄(pvp(n1))μ𝒄(pvp(n2))𝟏Δ(𝒄)0].\bar{\mu}_{F,2}^{\bm{a},n_{0}}(n_{1},n_{2}),\quad\lim_{k\to\infty}\mathbb{E}^{\bm{a},n_{0}}_{1\leq\bm{c}\leq(n_{1}n_{2})^{k}}[\mu_{\bm{c}}(n_{1})\mu_{\bm{c}}(n_{2})\bm{1}_{\Delta(\bm{c})\neq 0}],\quad\prod_{p\mid n_{1}n_{2}}\mathbb{E}^{\bm{a},n_{0}}_{\bm{c}\in\mathbb{Z}_{p}^{m}}[\mu_{\bm{c}}(p^{v_{p}(n_{1})})\mu_{\bm{c}}(p^{v_{p}(n_{2})})\bm{1}_{\Delta(\bm{c})\neq 0}].

This establishes the required existence, independence, and multiplicativity of limits.

We now turn to the required estimates. First, (6.5) follows from (6.8). Now assume pn0p\nmid n_{0}; we must prove (6.6) and (6.7). But pn0p\nmid n_{0} implies {𝒄pm:𝒄𝒂modn0}=pm\{\bm{c}\in\mathbb{Z}_{p}^{m}:\bm{c}\equiv\bm{a}\bmod{n_{0}}\}=\mathbb{Z}_{p}^{m}, so 𝔼𝒄pm𝒂,n0[f]=𝔼𝒄pm[f]\mathbb{E}^{\bm{a},n_{0}}_{\bm{c}\in\mathbb{Z}_{p}^{m}}[f]=\mathbb{E}_{\bm{c}\in\mathbb{Z}_{p}^{m}}[f] for any quantity ff. So by our pp-adic interpretations (from the previous paragraph) of μ¯F,1𝒂,n0(pl)\bar{\mu}_{F,1}^{\bm{a},n_{0}}(p^{l}), μ¯F,1𝒂,n0(pl1,pl2)\bar{\mu}_{F,1}^{\bm{a},n_{0}}(p^{l_{1}},p^{l_{2}}), it remains to prove the following:

  1. (1)

    𝔼𝒄pm[μ𝒄(p)𝟏Δ(𝒄)0]=λV(p)p1/2+O(p1)\mathbb{E}_{\bm{c}\in\mathbb{Z}_{p}^{m}}[\mu_{\bm{c}}(p)\bm{1}_{\Delta(\bm{c})\neq 0}]=\lambda^{\natural}_{V}(p)p^{-1/2}+O(p^{-1}).

  2. (2)

    𝔼𝒄pm[μ𝒄(p2)𝟏Δ(𝒄)0]=1+O(p1)\mathbb{E}_{\bm{c}\in\mathbb{Z}_{p}^{m}}[\mu_{\bm{c}}(p^{2})\bm{1}_{\Delta(\bm{c})\neq 0}]=1+O(p^{-1}).

  3. (3)

    𝔼𝒄pm[μ𝒄(p)2𝟏Δ(𝒄)0]=1+O(p1)\mathbb{E}_{\bm{c}\in\mathbb{Z}_{p}^{m}}[\mu_{\bm{c}}(p)^{2}\bm{1}_{\Delta(\bm{c})\neq 0}]=1+O(p^{-1}).

We now prove (1)–(3). By [wang2023dichotomous]*Corollary 1.7 (and our definition (2.12)), we have

(6.9) 𝔼𝒄𝔽pm[E𝒄(p)𝟏pΔ(𝒄)]\displaystyle\mathbb{E}_{\bm{c}\in\mathbb{F}_{p}^{m}}[E^{\natural}_{\bm{c}}(p)\bm{1}_{p\nmid\Delta(\bm{c})}] =λV(p)p1/2+O(p1),\displaystyle=\lambda^{\natural}_{V}(p)p^{-1/2}+O(p^{-1}),
(6.10) 𝔼𝒄𝔽pm[E𝒄(p2)𝟏pΔ(𝒄)]\displaystyle\mathbb{E}_{\bm{c}\in\mathbb{F}_{p}^{m}}[E^{\natural}_{\bm{c}}(p^{2})\bm{1}_{p\nmid\Delta(\bm{c})}] =1+O(p1),\displaystyle=1+O(p^{-1}),
(6.11) 𝔼𝒄𝔽pm[E𝒄(p)2𝟏pΔ(𝒄)]\displaystyle\mathbb{E}_{\bm{c}\in\mathbb{F}_{p}^{m}}[E^{\natural}_{\bm{c}}(p)^{2}\bm{1}_{p\nmid\Delta(\bm{c})}] =1+O(p1).\displaystyle=1+O(p^{-1}).

But for each f{μ𝒄(p),μ𝒄(p2),μ𝒄(p)2}f\in\{\mu_{\bm{c}}(p),\mu_{\bm{c}}(p^{2}),\mu_{\bm{c}}(p)^{2}\}, we have 𝔼𝒄pm[f𝟏Δ(𝒄)0]=𝔼𝒄pm[f𝟏pΔ(𝒄)]+O(p1)\mathbb{E}_{\bm{c}\in\mathbb{Z}_{p}^{m}}[f\bm{1}_{\Delta(\bm{c})\neq 0}]=\mathbb{E}_{\bm{c}\in\mathbb{Z}_{p}^{m}}[f\bm{1}_{p\nmid\Delta(\bm{c})}]+O(p^{-1}), by (6.8) and Lang–Weil (for Δ(𝒄)0modp\Delta(\bm{c})\equiv 0\bmod{p}). After rewriting f𝟏pΔ(𝒄)f\bm{1}_{p\nmid\Delta(\bm{c})} using (6.2), we conclude that (6.9) implies (1), that (6.11) implies (3), and that (6.10)–(6.11) imply (2). ∎

For the rest of §6, assume that mm is even and that Conjecture 1.2 holds.

6.2. The Sarnak–Shin–Templier framework

Using Conjecture 1.2 and (6.9)–(6.11), we can obtain useful statistical information on L(s,V𝒄)L(s,V_{\bm{c}}) over 𝒄𝒮1\bm{c}\in\mathcal{S}_{1}.

Let Π𝒄\Pi_{\bm{c}} be an isobaric automorphic representation over \mathbb{Q} corresponding to L(s,V𝒄)L(s,V_{\bm{c}}) in Conjecture 1.2. Let 𝒮2\mathcal{S}_{2} be the set of 𝒄𝒮1\bm{c}\in\mathcal{S}_{1} for which Π𝒄\Pi_{\bm{c}} is cuspidal, self-dual, and symplectic, in the sense of [sarnak2016families]*p. 533. For each 𝒄𝒮2\bm{c}\in\mathcal{S}_{2}, the LL-function L(s,V𝒄,2)L(s,V_{\bm{c}},\bigwedge^{2}) has a pole at s=1s=1, whence there exists an isobaric automorphic representation ϕ𝒄,2\phi_{\bm{c},2} over \mathbb{Q} with

(6.12) L(s,V𝒄,2)=ζ(s)L(s,ϕ𝒄,2).L(s,V_{\bm{c}},{\textstyle\bigwedge^{2}})=\zeta(s)L(s,\phi_{\bm{c},2}).
Proposition 6.2.

Let Z1Z\in\mathbb{R}_{\geq 1}. Then |(𝒮1𝒮2)[Z,Z]m|m,ϵZm1/2+ϵ\lvert(\mathcal{S}_{1}\setminus\mathcal{S}_{2})\cap[-Z,Z]^{m}\rvert\ll_{m,\epsilon}Z^{m-1/2+\epsilon}.

Proof.

We want to show that the family 𝒄Π𝒄\bm{c}\mapsto\Pi_{\bm{c}} indexed by 𝒄𝒮1\bm{c}\in\mathcal{S}_{1} is essentially cuspidal, self-dual, and symplectic (in the sense of [sarnak2016families]*p. 538, (i)–(iii)), with a power-saving exceptional set. We follow the GRH strategy suggested in [sarnak2016families].

Let ν0\nu_{0} be as in §1.2. Fix ϵ(0,12)\epsilon\in(0,\frac{1}{2}) and let P=Z1ϵP=Z^{1-\epsilon}.

By Proposition 3.2(7), each Π𝒄\Pi_{\bm{c}} is self-dual. Let 𝒮1.5\mathcal{S}_{1.5} be the set of 𝒄𝒮1\bm{c}\in\mathcal{S}_{1} for which Π𝒄\Pi_{\bm{c}} is cuspidal. Then L(s,V𝒄,2)L(s,V_{\bm{c}},\bigotimes^{2}) has a pole at s=1s=1 of order exactly 11 if 𝒄𝒮1.5\bm{c}\in\mathcal{S}_{1.5}, and at least 22 if 𝒄𝒮1𝒮1.5\bm{c}\in\mathcal{S}_{1}\setminus\mathcal{S}_{1.5}; this follows from the theory of unramified Rankin–Selberg LL-functions (cf. [farmer2019analytic]*proof of Lemma 2.3). A calculation with LL(s,V𝒄,2)\frac{L^{\prime}}{L}(s,V_{\bm{c}},\bigotimes^{2}) (using (3.2), GRH, and [iwaniec2004analytic]*§5.6’s Exercise 6 and §5.7’s Theorem 5.15) then yields

(6.13) 𝒄𝒮1ν0(𝒄/Z)pP:pΔ(𝒄)(logp)(λV𝒄,2(p)1)P𝒄𝒮1ν0(𝒄/Z)(𝟏𝒄𝒮1.5+Oϵ(P1/2+ϵ)).\sum_{\bm{c}\in\mathcal{S}_{1}}\nu_{0}(\bm{c}/Z)\sum_{p\leq P:\,p\nmid\Delta(\bm{c})}(\log{p})\cdot(\lambda^{\natural}_{V_{\bm{c}},\bigotimes^{2}}(p)-1)\geq P\sum_{\bm{c}\in\mathcal{S}_{1}}\nu_{0}(\bm{c}/Z)(\bm{1}_{\bm{c}\notin\mathcal{S}_{1.5}}+O_{\epsilon}(P^{-1/2+\epsilon})).

On the other hand, L(s,V𝒄,Sym2)L(s,V_{\bm{c}},\operatorname{Sym}^{2}) has a pole at s=1s=1 if 𝒄𝒮1.5𝒮2\bm{c}\in\mathcal{S}_{1.5}\setminus\mathcal{S}_{2}; this follows from [sarnak2016families]*p. 533. So a calculation with LL(s,V𝒄,Sym2)\frac{L^{\prime}}{L}(s,V_{\bm{c}},\operatorname{Sym}^{2}) gives

(6.14) 𝒄𝒮1ν0(𝒄/Z)pP:pΔ(𝒄)(logp)λV𝒄,Sym2(p)P𝒄𝒮1ν0(𝒄/Z)(𝟏𝒄𝒮1.5𝒮2+Oϵ(P1/2+ϵ)).\sum_{\bm{c}\in\mathcal{S}_{1}}\nu_{0}(\bm{c}/Z)\sum_{p\leq P:\,p\nmid\Delta(\bm{c})}(\log{p})\cdot\lambda^{\natural}_{V_{\bm{c}},\operatorname{Sym}^{2}}(p)\geq P\sum_{\bm{c}\in\mathcal{S}_{1}}\nu_{0}(\bm{c}/Z)(\bm{1}_{\bm{c}\in\mathcal{S}_{1.5}\setminus\mathcal{S}_{2}}+O_{\epsilon}(P^{-1/2+\epsilon})).

But for all primes pp and tuples 𝒄m\bm{c}\in\mathbb{Z}^{m} with pΔ(𝒄)p\nmid\Delta(\bm{c}), we have 𝒄𝒮1\bm{c}\in\mathcal{S}_{1}, and we can use (3.4), (3.5), (3.6) to write λV𝒄,2(p)=λ𝒄(p)2=E𝒄(p)2\lambda^{\natural}_{V_{\bm{c}},\bigotimes^{2}}(p)=\lambda^{\natural}_{\bm{c}}(p)^{2}=E^{\natural}_{\bm{c}}(p)^{2} and λV𝒄,Sym2(p)=12(E𝒄(p)2E𝒄(p2))\lambda^{\natural}_{V_{\bm{c}},\operatorname{Sym}^{2}}(p)=\frac{1}{2}(E^{\natural}_{\bm{c}}(p)^{2}-E^{\natural}_{\bm{c}}(p^{2})) (since 2m2\nmid m_{\ast}). So the left-hand side of (6.13) equals

pP(logp)𝒄mν0(𝒄/Z)(E𝒄(p)21)𝟏pΔ(𝒄)=pP(logp)Zm(0+O(p1))ϵZmPϵ,\sum_{p\leq P}(\log{p})\sum_{\bm{c}\in\mathbb{Z}^{m}}\nu_{0}(\bm{c}/Z)(E^{\natural}_{\bm{c}}(p)^{2}-1)\bm{1}_{p\nmid\Delta(\bm{c})}=\sum_{p\leq P}(\log{p})Z^{m}(0+O(p^{-1}))\ll_{\epsilon}Z^{m}P^{\epsilon},

by Poisson summation and (6.11). Similarly, by (6.11) and (6.10), the left-hand side of (6.14) is ϵZmPϵ\ll_{\epsilon}Z^{m}P^{\epsilon}. Thus from (6.13) we get |(𝒮1𝒮1.5)[Z/2,Z/2]m|m,ϵZmP1/2+ϵ\lvert(\mathcal{S}_{1}\setminus\mathcal{S}_{1.5})\cap[-Z/2,Z/2]^{m}\rvert\ll_{m,\epsilon}Z^{m}P^{-1/2+\epsilon}, and from (6.14) we get |(𝒮1.5𝒮2)[Z/2,Z/2]m|m,ϵZmP1/2+ϵ\lvert(\mathcal{S}_{1.5}\setminus\mathcal{S}_{2})\cap[-Z/2,Z/2]^{m}\rvert\ll_{m,\epsilon}Z^{m}P^{-1/2+\epsilon}. Now let ϵ0\epsilon\to 0. ∎

6.3. The Ratios Recipe

By Proposition 6.1 and [sarnak2016families]*pp. 534–535, Geometric Families and Remark (i), the recipe [conrey2008autocorrelation]*§5.1 makes sense for the family 𝒄Π𝒄\bm{c}\mapsto\Pi_{\bm{c}}. We will soon derive Conjecture 1.8 accordingly, along with the following:

Conjecture 6.3 (R2oo).

Let AF,2(s1,s2)A_{F,2}(s_{1},s_{2}) be defined as in §6.3.2 (in terms of FF). For each real Z2Z\geq 2, let σ(Z)\sigma(Z) be as in (1.13), and write sj=σ(Z)+itjs_{j}=\sigma(Z)+it_{j}. Then there exists a real >0\hbar>0 such that uniformly over reals Z2Z\geq 2 and t1,t2[Z,Z]t_{1},t_{2}\in[-Z^{\hbar},Z^{\hbar}], we have

(6.15) 𝒄𝒮1[Z,Z]mΦ𝒄,1(s1)Φ𝒄,1(s2)=𝒄𝒮1[Z,Z]m(1+oZ(1))AF,2(s1,s2)ζ(s1+s2).\sum_{\bm{c}\in\mathcal{S}_{1}\cap[-Z,Z]^{m}}\Phi^{\bm{c},1}(s_{1})\Phi^{\bm{c},1}(s_{2})=\sum_{\bm{c}\in\mathcal{S}_{1}\cap[-Z,Z]^{m}}(1+o_{Z\to\infty}(1))\cdot A_{F,2}(s_{1},s_{2})\zeta(s_{1}+s_{2}).

Here AF,2(s1,s2)A_{F,2}(s_{1},s_{2}) is an Euler product absolutely convergent for Re(s1),Re(s2)>1/3\operatorname{Re}(s_{1}),\operatorname{Re}(s_{2})>1/3.

Before proceeding, we make some remarks on our specific Ratios Conjectures.

Remark 6.4.

The LL-functions L(s,V𝒄)L(s,V_{\bm{c}}) are not all primitive, as the recipe in [conrey2005integral, conrey2008autocorrelation] requires. But by Proposition 6.2 and GRH, there is no real difference (in Proposition 6.1 and in our Ratios Conjectures) between 𝒮1\mathcal{S}_{1} and 𝒮2\mathcal{S}_{2} (on which each L(s,V𝒄)L(s,V_{\bm{c}}) is primitive).

Remark 6.5.

We do not order our families by conductor (or by discriminant, for that matter). We are indexing by different level sets, as is natural for families like ours; cf. [sarnak2016families]*p. 535, Remark (i); and p. 560, second paragraph after (25).

Remark 6.6.

In Conjectures 1.8, 1.10, and 6.3, we restrict tt to (comfortably) respect the constraint [conrey2007applications]*(2.11c) on “vertical shifts”. But it would be reasonable to allow t[ZA,ZA]t\in[-Z^{A},Z^{A}] for arbitrarily large A>0A>0; cf. [bettin2020averages]*p. 4, the sentence before Conjecture 2.

6.3.1. Deriving (RA1)

To derive Conjecture 1.8, first use (1.9) to write Φ𝒄,1(s)\Phi^{\bm{c},1}(s) in terms of 1/L(s,V𝒄)1/L(s,V_{\bm{c}}), and then replace each term L(s,V𝒄)1=n1μ𝒄(n)nsL(s,V_{\bm{c}})^{-1}=\sum_{n\geq 1}\mu_{\bm{c}}(n)n^{-s} on the left-hand side of (1.14) with its “naive expected value over {𝒄𝒮1ZM(𝒃):𝒄𝒂modn0}\{\bm{c}\in\mathcal{S}_{1}\cap Z\cdot\mathcal{B}_{M}(\bm{b}):\bm{c}\equiv\bm{a}\bmod{n_{0}}\} as ZZ\to\infty” (computed using Proposition 6.1), i.e. the Dirichlet series

(6.16) n1μ¯F,1𝒂,n0(n)ns.\sum_{n\geq 1}\bar{\mu}_{F,1}^{\bm{a},n_{0}}(n)n^{-s}.

It turns out that the series (6.16) behaves much like ζ(2s)L(s+1/2,V)\zeta(2s)L(s+1/2,V), as we now explain.

Define AF,1𝒂,n0(s)A_{F,1}^{\bm{a},n_{0}}(s) to be the product of (6.16) and ζ(2s)1L(s+1/2,V)1\zeta(2s)^{-1}L(s+1/2,V)^{-1}, so that (6.16) factors as AF,1𝒂,n0(s)ζ(2s)L(s+1/2,V)A_{F,1}^{\bm{a},n_{0}}(s)\zeta(2s)L(s+1/2,V). Let a¯F,1𝒂,n0(n)\bar{a}_{F,1}^{\bm{a},n_{0}}(n) be the nnth coefficient of the Dirichlet series AF,1𝒂,n0(s)A_{F,1}^{\bm{a},n_{0}}(s). Then a¯F,1𝒂,n0(n)\bar{a}_{F,1}^{\bm{a},n_{0}}(n) is multiplicative in nn by (6.3); and by (6.5), (6.6) we have

(6.17) a¯F,1𝒂,n0(n)ϵnϵ,a¯F,1𝒂,n0(p)𝟏pn0p1,a¯F,1𝒂,n0(p2)𝟏pn0p1/2\bar{a}_{F,1}^{\bm{a},n_{0}}(n)\ll_{\epsilon}n^{\epsilon},\quad\bar{a}_{F,1}^{\bm{a},n_{0}}(p)\bm{1}_{p\nmid n_{0}}\ll p^{-1},\quad\bar{a}_{F,1}^{\bm{a},n_{0}}(p^{2})\bm{1}_{p\nmid n_{0}}\ll p^{-1/2}

(because a¯F,1𝒂,n0(p)=μ¯F,1𝒂,n0(p)λV(p)p1/2\bar{a}_{F,1}^{\bm{a},n_{0}}(p)=\bar{\mu}_{F,1}^{\bm{a},n_{0}}(p)-\lambda^{\natural}_{V}(p)p^{-1/2} and a¯F,1𝒂,n0(p2)=μ¯F,1𝒂,n0(p2)μ¯F,1𝒂,n0(p)λV(p)p1/21+O(p1)\bar{a}_{F,1}^{\bm{a},n_{0}}(p^{2})=\bar{\mu}_{F,1}^{\bm{a},n_{0}}(p^{2})-\bar{\mu}_{F,1}^{\bm{a},n_{0}}(p)\lambda^{\natural}_{V}(p)p^{-1/2}-1+O(p^{-1})). So AF,1𝒂,n0(s)A_{F,1}^{\bm{a},n_{0}}(s) has an Euler product, and satisfies

(6.18) |AF,1𝒂,n0(s)|n1|a¯F,1𝒂,n0(n)|n1/3+ϵp(1+O(𝟏pn0)+O(p1)p1/3+ϵ+O(p1/2)p2/3+2ϵ+Oϵ(pϵ)p1+3ϵ)ϵn0ϵ\lvert A_{F,1}^{\bm{a},n_{0}}(s)\rvert\leq\sum_{n\geq 1}\frac{\lvert\bar{a}_{F,1}^{\bm{a},n_{0}}(n)\rvert}{n^{1/3+\epsilon}}\leq\prod_{p}\biggl{(}1+O(\bm{1}_{p\mid n_{0}})+\frac{O(p^{-1})}{p^{1/3+\epsilon}}+\frac{O(p^{-1/2})}{p^{2/3+2\epsilon}}+\frac{O_{\epsilon}(p^{\epsilon})}{p^{1+3\epsilon}}\biggr{)}\ll_{\epsilon}n_{0}^{\epsilon}

for Re(s)1/3+ϵ\operatorname{Re}(s)\geq 1/3+\epsilon (for any ϵ>0\epsilon>0). On the other hand, ζ(2s)L(s+1/2,V)\zeta(2s)L(s+1/2,V) has a pole of order 1\geq 1 at s=1/2s=1/2, and thus is more dominant than AF,1𝒂,n0(s)A_{F,1}^{\bm{a},n_{0}}(s) in (6.16).

In view of the above, the Ratios Recipe [conrey2008autocorrelation]*§5.1 produces the conjecture

𝒄𝒮1ZM(𝒃):𝒄𝒂modn0L(s,V𝒄)1=𝒄𝒮1ZM(𝒃):𝒄𝒂modn0(1+oM;Z(1))AF,1𝒂,n0(s)ζ(2s)L(s+1/2,V)\sum_{\begin{subarray}{c}\bm{c}\in\mathcal{S}_{1}\cap Z\cdot\mathcal{B}_{M}(\bm{b}):\\ \bm{c}\equiv\bm{a}\bmod{n_{0}}\end{subarray}}L(s,V_{\bm{c}})^{-1}=\sum_{\begin{subarray}{c}\bm{c}\in\mathcal{S}_{1}\cap Z\cdot\mathcal{B}_{M}(\bm{b}):\\ \bm{c}\equiv\bm{a}\bmod{n_{0}}\end{subarray}}(1+o_{M;Z\to\infty}(1))\cdot A_{F,1}^{\bm{a},n_{0}}(s)\zeta(2s)L(s+1/2,V)

(for n0n_{0}, tt, ss as in Conjecture 1.8). This rearranges (upon division by ζ(2s)L(s+1/2,V)\zeta(2s)L(s+1/2,V)) to (1.14), giving Conjecture 1.8. Furthermore, the fullest Ratios Conjectures include a power-saving error term, leading naturally to Conjecture 1.10.

Here ζ(2s)\zeta(2s), L(s+1/2,V)L(s+1/2,V) are called polar factors (or polar terms). As we will see shortly, an additional polar factor, ζ(s1+s2)\zeta(s_{1}+s_{2}), arises in (R2).

Remark 6.7.

The Ratios Recipe involves the approximate functional equation for LL when there are LL’s in the numerator, but not when there are only LL’s in the denominator.

6.3.2. Deriving (R2)

For Conjecture 6.3, use (1.9) to write Φ𝒄,1(sj)\Phi^{\bm{c},1}(s_{j}) in terms of 1/L(sj,V𝒄)1/L(s_{j},V_{\bm{c}}), and then replace each term L(s1,V𝒄)1L(s2,V𝒄)1=n1,n21μ𝒄(n1)μ𝒄(n2)n1s1n2s2L(s_{1},V_{\bm{c}})^{-1}L(s_{2},V_{\bm{c}})^{-1}=\sum_{n_{1},n_{2}\geq 1}\mu_{\bm{c}}(n_{1})\mu_{\bm{c}}(n_{2})n_{1}^{-s_{1}}n_{2}^{-s_{2}} on the left-hand side of (6.15) with its “naive expected value over 𝒄𝒮1[Z,Z]m\bm{c}\in\mathcal{S}_{1}\cap[-Z,Z]^{m} as ZZ\to\infty” (computed using Proposition 6.1), i.e. the series

n1,n21μ¯F,2𝟎,1(n1,n2)n1s1n2s2,\sum_{n_{1},n_{2}\geq 1}\bar{\mu}_{F,2}^{\bm{0},1}(n_{1},n_{2})n_{1}^{-s_{1}}n_{2}^{-s_{2}},

which by (6.4) factors as AF,2(s1,s2)ζ(s1+s2)1j2(ζ(2sj)L(sj+1/2,V))A_{F,2}(s_{1},s_{2})\zeta(s_{1}+s_{2})\prod_{1\leq j\leq 2}(\zeta(2s_{j})L(s_{j}+1/2,V)) for some Euler product AF,2(s1,s2)A_{F,2}(s_{1},s_{2}). If Re(s1),Re(s2)1/3+ϵ\operatorname{Re}(s_{1}),\operatorname{Re}(s_{2})\geq 1/3+\epsilon, then by (6.5) and (6.7), we have

(6.19) |AF,2(s1,s2)|n1,n21|a¯F,2(n1,n2)|(n1n2)1/3ϵϵ1.\lvert A_{F,2}(s_{1},s_{2})\rvert\leq\sum_{n_{1},n_{2}\geq 1}\lvert\bar{a}_{F,2}(n_{1},n_{2})\rvert\,(n_{1}n_{2})^{-1/3-\epsilon}\ll_{\epsilon}1.

(The justification is similar to that for AF,1𝒂,n0(s)A_{F,1}^{\bm{a},n_{0}}(s). Note in particular that if a¯F,2(n1,n2)\bar{a}_{F,2}(n_{1},n_{2}) is the (n1,n2)(n_{1},n_{2})th coefficient of the double Dirichlet series AF,2(s1,s2)A_{F,2}(s_{1},s_{2}), then a¯F,2(pl,1)=a¯F,2(1,pl)=a¯F,1𝟎,1(pl)\bar{a}_{F,2}(p^{l},1)=\bar{a}_{F,2}(1,p^{l})=\bar{a}_{F,1}^{\bm{0},1}(p^{l}) and a¯F,2(p,p)=μ¯F,2𝟎,1(p,p)2μ¯F,1𝟎,1(p)λV(p)p1/21+λV(p)2p1\bar{a}_{F,2}(p,p)=\bar{\mu}_{F,2}^{\bm{0},1}(p,p)-2\bar{\mu}_{F,1}^{\bm{0},1}(p)\lambda^{\natural}_{V}(p)p^{-1/2}-1+\lambda^{\natural}_{V}(p)^{2}p^{-1}.)

Division by 1j2(ζ(2sj)L(sj+1/2,V))\prod_{1\leq j\leq 2}(\zeta(2s_{j})L(s_{j}+1/2,V)) (cf. §6.3.1) leads to Conjecture 6.3.

6.4. From (R2) to (R2’) and (R2’E)

Write AF,2(𝒔)=AF,2(s1,s2)A_{F,2}(\bm{s})=A_{F,2}(s_{1},s_{2}). Conjecture 6.3 implies, uniformly over Z2Z\geq 2 and 𝒕2\bm{t}\in\mathbb{R}^{2}, that for sj=σ(Z)+itjs_{j}=\sigma(Z)+it_{j}, we have (for all ϵ>0\epsilon>0)

(6.20) 𝒄𝒮1[Z,Z]mΦ𝒄,1(s1)Φ𝒄,1(s2)=oϵ;Z(Zm(1+𝒕)ϵ)+𝒄𝒮1[Z,Z]mAF,2(𝒔)ζ(s1+s2);\sum_{\bm{c}\in\mathcal{S}_{1}\cap[-Z,Z]^{m}}\Phi^{\bm{c},1}(s_{1})\Phi^{\bm{c},1}(s_{2})=o_{\epsilon;Z\to\infty}(Z^{m}(1+\lVert\bm{t}\rVert)^{\epsilon})+\sum_{\bm{c}\in\mathcal{S}_{1}\cap[-Z,Z]^{m}}A_{F,2}(\bm{s})\zeta(s_{1}+s_{2});

this follows for 𝒕Z\lVert\bm{t}\rVert\leq Z^{\hbar} by Conjecture 6.3, and for 𝒕>Z\lVert\bm{t}\rVert>Z^{\hbar} by GRH (see Proposition 3.2(8)). Using (6.20), we proceed to derive Conjecture 1.4 (R2’).

Proposition 6.8.

Assume Conjectures 1.2 and 6.3. Then Conjecture 1.4 holds.

Proof.

Let f¯(s)\colonequalsf(s¯)¯\overline{f}(s)\colonequals\overline{f(\overline{s})}. Let Z,N1Z,N\geq 1 with NZ3N\leq Z^{3}. The left-hand side of (1.10) is independent of σ0>1/2\sigma_{0}>1/2, since each integrand is holomorphic on Re(s)>1/2\operatorname{Re}(s)>1/2 by GRH. Now shift contours to Re(s)=σ(Z)\operatorname{Re}(s)=\sigma(Z), and expand squares using self-duality of the LL-functions in (1.9) (see Proposition 3.2(7)), to equate the left-hand side of (1.10) with the quantity

Σ0\colonequals𝒄𝒮1[Z,Z]mσ(Z)iσ(Z)+i𝑑s1σ(Z)+iσ(Z)i𝑑s2Φ𝒄,1(s1)Φ𝒄,1(s2)f(s1)f¯(s2)Ns1+s2.\Sigma_{0}\colonequals\sum_{\bm{c}\in\mathcal{S}_{1}\cap[-Z,Z]^{m}}\int_{\sigma(Z)-i\infty}^{\sigma(Z)+i\infty}ds_{1}\int_{\sigma(Z)+i\infty}^{\sigma(Z)-i\infty}ds_{2}\,\Phi^{\bm{c},1}(s_{1})\Phi^{\bm{c},1}(s_{2})\cdot f(s_{1})\overline{f}(s_{2})N^{s_{1}+s_{2}}.

After switching the order of 𝒄\bm{c} and 𝒔\bm{s} in Σ0\Sigma_{0} (using Fubini), and plugging in the estimate (6.20) and the bound 1+𝒕(1+|t1|)(1+|t2|)1+\lVert\bm{t}\rVert\leq(1+\lvert t_{1}\rvert)(1+\lvert t_{2}\rvert) for each 𝒔\bm{s}, we get the estimate

(6.21) Σ0=oϵ;Z(ZmΣ12)+Σ2,\Sigma_{0}=o_{\epsilon;Z\to\infty}(Z^{m}\Sigma_{1}^{2})+\Sigma_{2},

where Σ1\colonequalst𝑑t(1+|t|)ϵ|f(σ(Z)+it)|Nσ(Z)\Sigma_{1}\colonequals\int_{t\in\mathbb{R}}dt\,(1+\lvert t\rvert)^{\epsilon}\cdot\lvert f(\sigma(Z)+it)\rvert N^{\sigma(Z)} and

Σ2\colonequals𝒄𝒮1[Z,Z]mσ(Z)iσ(Z)+i𝑑s1σ(Z)+iσ(Z)i𝑑s2ζ(s1+s2)AF,2(𝒔)f(s1)f¯(s2)Ns1+s2.\Sigma_{2}\colonequals\sum_{\bm{c}\in\mathcal{S}_{1}\cap[-Z,Z]^{m}}\int_{\sigma(Z)-i\infty}^{\sigma(Z)+i\infty}ds_{1}\int_{\sigma(Z)+i\infty}^{\sigma(Z)-i\infty}ds_{2}\,\zeta(s_{1}+s_{2})A_{F,2}(\bm{s})\cdot f(s_{1})\overline{f}(s_{2})N^{s_{1}+s_{2}}.

Here NZO(1)N\leq Z^{O(1)}, so Σ1N1/2sup0σ2(1+|t|)ϵf(σ+it)Lt1()\Sigma_{1}\ll N^{1/2}\sup_{0\leq\sigma\leq 2}\lVert(1+\lvert t\rvert)^{\epsilon}f(\sigma+it)\rVert_{L^{1}_{t}(\mathbb{R})}. And by Cauchy–Schwarz,

(6.22) (1+|t|)ϵf(σ+it)Lt1()(1+|t|)1/2ϵLt2()(1+|t|)1/2+2ϵf(σ+it)Lt2().\lVert(1+\lvert t\rvert)^{\epsilon}f(\sigma+it)\rVert_{L^{1}_{t}(\mathbb{R})}\leq\lVert(1+\lvert t\rvert)^{-1/2-\epsilon}\rVert_{L^{2}_{t}(\mathbb{R})}\cdot\lVert(1+\lvert t\rvert)^{1/2+2\epsilon}f(\sigma+it)\rVert_{L^{2}_{t}(\mathbb{R})}.

Now let δ=1/20\delta=1/20, and assume ZZ is large enough that 1σ(Z)δ1/3+δ1-\sigma(Z)-\delta\geq 1/3+\delta. Shifting s2s_{2} (in Σ2\Sigma_{2}) from Re(s2)=σ(Z)\operatorname{Re}(s_{2})=\sigma(Z) to Re(s2)=1σ(Z)δ\operatorname{Re}(s_{2})=1-\sigma(Z)-\delta yields Σ2=Σ3+OF,ϵ(ZmΣ4)\Sigma_{2}=\Sigma_{3}+O_{F,\epsilon}(Z^{m}\Sigma_{4}), where

Σ3\colonequals𝒄𝒮1[Z,Z]mσ(Z)iσ(Z)+i𝑑s1(2πi)AF,2(s1,1s1)f(s1)f¯(1s1)N\Sigma_{3}\colonequals\sum_{\bm{c}\in\mathcal{S}_{1}\cap[-Z,Z]^{m}}\int_{\sigma(Z)-i\infty}^{\sigma(Z)+i\infty}ds_{1}\,(-2\pi i)A_{F,2}(s_{1},1-s_{1})\cdot f(s_{1})\overline{f}(1-s_{1})N

comes from the residue of ζ(s1+s2)\zeta(s_{1}+s_{2}) at s2=1s1s_{2}=1-s_{1}, and where on Re(s2)=1σ(Z)δ\operatorname{Re}(s_{2})=1-\sigma(Z)-\delta we use (6.19) and the consequence ζ(s1+s2)δ,ϵ(1+|t1+t2|)ϵ\zeta(s_{1}+s_{2})\ll_{\delta,\epsilon}(1+\lvert t_{1}+t_{2}\rvert)^{\epsilon} of RH to be able to take

Σ4\colonequals𝒕2𝑑t1𝑑t2(1+|t1+t2|)ϵ|f(σ(Z)+it1)f(1σ(Z)δit2)|N1δ.\Sigma_{4}\colonequals\int_{\bm{t}\in\mathbb{R}^{2}}dt_{1}\,dt_{2}\,(1+\lvert t_{1}+t_{2}\rvert)^{\epsilon}\lvert f(\sigma(Z)+it_{1})f(1-\sigma(Z)-\delta-it_{2})\rvert N^{1-\delta}.

Shifting s1s_{1} (in Σ3\Sigma_{3}) to Re(s1)=1/2\operatorname{Re}(s_{1})=1/2 yields Σ3FZmNf(1/2+it)Lt2()2\Sigma_{3}\ll_{F}Z^{m}N\lVert f(1/2+it)\rVert_{L^{2}_{t}(\mathbb{R})}^{2}. Also, Σ4N1δsup0σ2(1+|t|)ϵf(σ+it)Lt1()2\Sigma_{4}\ll N^{1-\delta}\sup_{0\leq\sigma\leq 2}\lVert(1+\lvert t\rvert)^{\epsilon}f(\sigma+it)\rVert_{L^{1}_{t}(\mathbb{R})}^{2}, since 1+|t1+t2|(1+|t1|)(1+|t2|)1+\lvert t_{1}+t_{2}\rvert\leq(1+\lvert t_{1}\rvert)(1+\lvert t_{2}\rvert). But Σ0O(ZmΣ12)+Σ3+OF,ϵ(ZmΣ4)\Sigma_{0}\leq O(Z^{m}\Sigma_{1}^{2})+\Sigma_{3}+O_{F,\epsilon}(Z^{m}\Sigma_{4}), by (6.21). By (6.22) with ϵ=1/4\epsilon=1/4, we get (1.10). ∎

As a stepping stone from Conjecture 1.4 (R2’) to Conjecture 7.4 (R2’E’), we now state (R2’E). Let a𝒄,1(n)a_{\bm{c},1}(n) be the nnth coefficient of the Dirichlet series Φ𝒄,1(s)\Phi^{\bm{c},1}(s).

Conjecture 6.9 (R2’E).

Fix a function DCc(>0)D\in C^{\infty}_{c}(\mathbb{R}_{>0}). Let t0t_{0}\in\mathbb{R}. Let Z,N>0Z,N\in\mathbb{R}_{>0} with NZ3N\leq Z^{3}. Then for some real A3>0A_{3}>0 depending only on FF, we have

(6.23) 𝒄𝒮1[Z,Z]m|n1D(n/N)nit0a𝒄,1(n)|2F,D(1+|t0|)A3ZmN.\sum_{\bm{c}\in\mathcal{S}_{1}\cap[-Z,Z]^{m}}\,\Bigl{\lvert}{\sum_{n\geq 1}D(n/N)\cdot n^{-it_{0}}a_{\bm{c},1}(n)}\Bigr{\rvert}^{2}\ll_{F,D}(1+\lvert t_{0}\rvert)^{A_{3}}Z^{m}N.
Proposition 6.10.

Assume Conjecture 1.4. Then Conjecture 6.9 holds.

Proof.

The case N<1N<1 is trivial, so assume N1N\geq 1. Let σ0=1.5\sigma_{0}=1.5. By (5.2), D(n/N)=(2π)1t𝑑tD(σ0+it)(N/n)σ0+itD(n/N)=(2\pi)^{-1}\int_{t\in\mathbb{R}}dt\,D^{\vee}(\sigma_{0}+it)(N/n)^{\sigma_{0}+it}. Apply Conjecture 1.4 with f(s)=D(sit0)f(s)=D^{\vee}(s-it_{0}), and bound ff using Proposition 5.1, to get (6.23) with A3=2A_{3}=2. ∎

6.5. From (RA1) to (RA1’E)

We now build on Conjectures 1.8 and 1.10, introducing flexible weights over 𝒄\bm{c}. We need some terminology on residue classes of m\mathbb{Z}^{m}.

Definition 6.11.

If =𝒂+qmm\mathcal{R}=\bm{a}+q\mathbb{Z}^{m}\subseteq\mathbb{Z}^{m} (where q1q\geq 1), let q\colonequalsqq_{\mathcal{R}}\colonequals q be the modulus of \mathcal{R}, and let AF,1(s)\colonequalsAF,1𝒂,q(s)A_{F,1}^{\mathcal{R}}(s)\colonequals A_{F,1}^{\bm{a},q}(s) and a¯F,1(n)\colonequalsa¯F,1𝒂,q(n)\bar{a}_{F,1}^{\mathcal{R}}(n)\colonequals\bar{a}_{F,1}^{\bm{a},q}(n) (where AF,1𝒂,q(s)A_{F,1}^{\bm{a},q}(s), a¯F,1𝒂,q(n)\bar{a}_{F,1}^{\bm{a},q}(n) are defined as in §6.3.1). Given a nonempty set 𝒮={}\mathscr{S}=\{\mathcal{R}\} of residue classes m\mathcal{R}\subseteq\mathbb{Z}^{m}, let Q(𝒮)\colonequalsmax𝒮(q)Q(\mathscr{S})\colonequals\max_{\mathcal{R}\in\mathscr{S}}(q_{\mathcal{R}}).

Let 𝒫={}\mathscr{P}=\{\mathcal{R}\} be a partition of m\mathbb{Z}^{m} into finitely many residue classes m\mathcal{R}\subseteq\mathbb{Z}^{m}. (In §7.3, we will construct the partitions needed for our main results.) Let I>0I\subseteq\mathbb{R}_{>0} be a compact set. Let ν=ν𝒄(r)\nu=\nu_{\bm{c}}(r) be a smooth function m×,(𝒄,r)ν𝒄(r)\mathbb{R}^{m}\times\mathbb{R}\to\mathbb{C},\,(\bm{c},r)\mapsto\nu_{\bm{c}}(r) supported on [1,1]m×I[-1,1]^{m}\times I. Let

(6.24) 1,k=1,k(ν)\colonequals|𝜶|10jksup(𝒄~,r)m×>0|𝒄~𝜶logrjν𝒄~(r)|.\mathcal{M}_{1,k}=\mathcal{M}_{1,k}(\nu)\colonequals\sum_{\lvert\bm{\alpha}\rvert\leq 1}\,\sum_{0\leq j\leq k}\,\sup_{(\tilde{\bm{c}},r)\in\mathbb{R}^{m}\times\mathbb{R}_{>0}}\left\lvert\partial_{\tilde{\bm{c}}}^{\bm{\alpha}}\partial_{\log r}^{j}{\nu_{\tilde{\bm{c}}}(r)}\right\rvert.
Conjecture 6.12 (RA1oo’E).

Let Z,N2Q(𝒫)Z,N\geq 2Q(\mathscr{P}) be reals with NZ3N\leq Z^{3}. Then the quantity

(6.25) 𝒫|𝒄𝒮1n1ν𝒄/Z(n/N)(a𝒄,1(n)a¯F,1(n))|\sum_{\mathcal{R}\in\mathscr{P}}\,\Bigl{\lvert}\sum_{\bm{c}\in\mathcal{S}_{1}\cap\mathcal{R}}\sum_{n\geq 1}\nu_{\bm{c}/Z}(n/N)\cdot(a_{\bm{c},1}(n)-\bar{a}_{F,1}^{\mathcal{R}}(n))\Bigr{\rvert}

is F,IZmN1/2oF,Q(𝒫);Z(1)1,A4\ll_{F,I}Z^{m}N^{1/2}\cdot o_{F,Q(\mathscr{P});Z\to\infty}(1)\cdot\mathcal{M}_{1,A_{4}}, for some real A4=A4(F)>0A_{4}=A_{4}(F)>0.

Proposition 6.13.

Assume Conjectures 1.2, 1.4, and 1.8. Then Conjecture 6.12 holds.

Proof.

(It would be nice to prove this only assuming (1.14) for t[M,M]t\in[-M,M], say, but it will be convenient to assume (1.14) for all t[logZ,logZ]t\in[-\log{Z},\log{Z}], as in Conjecture 1.8.)

By (5.2), we have ν𝒄/Z(n/N)=(2πi)1(σ(Z))𝑑sν𝒄/Z(s)(N/n)s\nu_{\bm{c}/Z}(n/N)=(2\pi i)^{-1}\int_{(\sigma(Z))}ds\,\nu_{\bm{c}/Z}^{\vee}(s)(N/n)^{s}, so that

n1ν𝒄/Z(n/N)(a𝒄,1(n)a¯F,1(n))=(2πi)1(σ(Z))𝑑sν𝒄/Z(s)Ns(Φ𝒄,1(s)AF,1(s)).\sum_{n\geq 1}\nu_{\bm{c}/Z}(n/N)(a_{\bm{c},1}(n)-\bar{a}_{F,1}^{\mathcal{R}}(n))=(2\pi i)^{-1}\int_{(\sigma(Z))}ds\,\nu_{\bm{c}/Z}^{\vee}(s)N^{s}(\Phi^{\bm{c},1}(s)-A_{F,1}^{\mathcal{R}}(s)).

Let MQ(𝒫)M\geq Q(\mathscr{P}) be a real parameter; soon below, we will let MM tend slowly to infinity as ZZ\to\infty. Recall M(𝒃)\mathcal{B}_{M}(\bm{b}) from (1.12). Weight ν\nu is supported on [1,1]m×I[-1,1]^{m}\times I, so the triangle inequality and the previous display imply that (6.25) is at most

Σ5\colonequals𝒫𝒃[M,M]m|𝒄𝒮1ZM(𝒃)(σ(Z))𝑑sν𝒄/Z(s)Ns(Φ𝒄,1(s)AF,1(s))|.\Sigma_{5}\colonequals\sum_{\mathcal{R}\in\mathscr{P}}\sum_{\bm{b}\in[-M,M]^{m}}\,\Bigl{\lvert}\sum_{\bm{c}\in\mathcal{S}_{1}\cap\mathcal{R}\cap Z\cdot\mathcal{B}_{M}(\bm{b})}\int_{(\sigma(Z))}ds\,\nu_{\bm{c}/Z}^{\vee}(s)N^{s}(\Phi^{\bm{c},1}(s)-A_{F,1}^{\mathcal{R}}(s))\Bigr{\rvert}.

Uniformly over 𝒄m\bm{c}\in\mathbb{R}^{m} and ss\in\mathbb{C}, Proposition 5.1 and (6.24) give (for all b0b\in\mathbb{Z}_{\geq 0})

(6.26) ν𝒄/Z(s)bOI(1)1+|Re(s)|1,b(ν)(1+|Im(s)|)b.\nu_{\bm{c}/Z}^{\vee}(s)\ll_{b}O_{I}(1)^{1+\lvert\operatorname{Re}(s)\rvert}\mathcal{M}_{1,b}(\nu)(1+\lvert\operatorname{Im}(s)\rvert)^{-b}.

For each pair (,𝒃)(\mathcal{R},\bm{b}), choose an element 𝒄(,𝒃)\bm{c}(\mathcal{R},\bm{b}) of 𝒮1ZM(𝒃)\mathcal{S}_{1}\cap\mathcal{R}\cap Z\cdot\mathcal{B}_{M}(\bm{b}), if such an element exists. Let Σ6\Sigma_{6} be Σ5\Sigma_{5} with ν𝒄/Z(s)ν𝒄(,𝒃)/Z(s)\nu_{\bm{c}/Z}^{\vee}(s)-\nu_{\bm{c}(\mathcal{R},\bm{b})/Z}^{\vee}(s) in place of ν𝒄/Z(s)\nu_{\bm{c}/Z}^{\vee}(s), and let Σ7\Sigma_{7} be Σ5\Sigma_{5} with ν𝒄(,𝒃)/Z(s)\nu_{\bm{c}(\mathcal{R},\bm{b})/Z}^{\vee}(s) in place of ν𝒄/Z(s)\nu_{\bm{c}/Z}^{\vee}(s). Clearly Σ5Σ6+Σ7\Sigma_{5}\leq\Sigma_{6}+\Sigma_{7}.

We need to split Σ7\Sigma_{7} further according to the size of tt, with some analytic care (keeping in mind the entireness hypothesis on ff in Conjecture 1.4). Let B=BM(s)\colonequalse(s/M)2B=B_{M}(s)\colonequals e^{(s/M)^{2}}. The following hold uniformly over σ\sigma in any fixed finite interval:

  1. (1)

    For all tt\in\mathbb{R}, we have B(σ+it)1B(\sigma+it)\ll 1.

  2. (2)

    If |t|>logZ\lvert t\rvert>\log{Z}, then B(σ+it)MZ1B(\sigma+it)\ll_{M}Z^{-1}.

  3. (3)

    If |t|M1/2\lvert t\rvert\leq M^{1/2}, then B(σ+it)=1+O(M1)B(\sigma+it)=1+O(M^{-1}).

Let Σ7(A)\Sigma_{7}(A) be Σ5\Sigma_{5} with ν𝒄(,𝒃)/Z(s)A(s)\nu_{\bm{c}(\mathcal{R},\bm{b})/Z}^{\vee}(s)\cdot A(s) in place of ν𝒄/Z(s)\nu_{\bm{c}/Z}^{\vee}(s), so that

(6.27) Σ7Σ7(B𝟏|t|logZ)+Σ7(B𝟏|t|>logZ)+Σ7(1B).\Sigma_{7}\leq\Sigma_{7}(B\cdot\bm{1}_{\lvert t\rvert\leq\log{Z}})+\Sigma_{7}(B\cdot\bm{1}_{\lvert t\rvert>\log{Z}})+\Sigma_{7}(1-B).

We first bound Σ7(B𝟏|t|logZ)\Sigma_{7}(B\cdot\bm{1}_{\lvert t\rvert\leq\log{Z}}) and Σ7(B𝟏|t|>logZ)\Sigma_{7}(B\cdot\bm{1}_{\lvert t\rvert>\log{Z}}). Plugging (1), (6.26) (with b=2b=2), and Conjecture 1.8 (for t[logZ,logZ]t\in[-\log{Z},\log{Z}]) into Σ7(B𝟏|t|logZ)\Sigma_{7}(B\cdot\bm{1}_{\lvert t\rvert\leq\log{Z}}), we find that

Σ7(B𝟏|t|logZ)𝒫𝒃[M,M]m𝒄𝒮1ZM(𝒃)OI(1,2(ν))Nσ(Z)oM;Z(1),\Sigma_{7}(B\cdot\bm{1}_{\lvert t\rvert\leq\log{Z}})\leq\sum_{\mathcal{R}\in\mathscr{P}}\sum_{\bm{b}\in[-M,M]^{m}}\sum_{\bm{c}\in\mathcal{S}_{1}\cap\mathcal{R}\cap Z\cdot\mathcal{B}_{M}(\bm{b})}O_{I}(\mathcal{M}_{1,2}(\nu))N^{\sigma(Z)}o_{M;Z\to\infty}(1),

which is in turn IZmN1/2oM;Z(1)1,2(ν)\ll_{I}Z^{m}N^{1/2}o_{M;Z\to\infty}(1)\mathcal{M}_{1,2}(\nu). Similarly, plugging (2), (6.26) (with b=2b=2), GRH (see Proposition 3.2(8)), and (6.18) into Σ7(B𝟏|t|>logZ)\Sigma_{7}(B\cdot\bm{1}_{\lvert t\rvert>\log{Z}}) reveals that

Σ7(B𝟏|t|>logZ)M,ϵ𝒄𝒮1[Z,Z]mZ1OI(1,2(ν))Nσ(Z)(Zϵ+Q(𝒫)ϵ),\Sigma_{7}(B\cdot\bm{1}_{\lvert t\rvert>\log{Z}})\ll_{M,\epsilon}\sum_{\bm{c}\in\mathcal{S}_{1}\cap[-Z,Z]^{m}}Z^{-1}O_{I}(\mathcal{M}_{1,2}(\nu))N^{\sigma(Z)}(Z^{\epsilon}+Q(\mathscr{P})^{\epsilon}),

which (if ϵ=1/2\epsilon=1/2, say) is IZmN1/2oM;Z(1)1,2(ν)\ll_{I}Z^{m}N^{1/2}o_{M;Z\to\infty}(1)\mathcal{M}_{1,2}(\nu) (since Q(𝒫)ZQ(\mathscr{P})\leq Z).

By choosing M=M(Z)Q(𝒫)M=M(Z)\geq Q(\mathscr{P}) appropriately, we may ensure both (i) that MM\to\infty as ZZ\to\infty, and (ii) that the two “oM;Z(1)o_{M;Z\to\infty}(1)” terms in the previous paragraph are oZ(1)o_{Z\to\infty}(1).

It remains to bound Σ6\Sigma_{6} and Σ7(1B)\Sigma_{7}(1-B). To handle both at once, we need a Fourier analog of Lemma 5.2. Let f:f\colon\mathbb{C}\to\mathbb{C} be one of the functions f𝒄,6:sν𝒄/Z(s)ν𝒄(,𝒃)/Z(s)f_{\bm{c},6}\colon s\mapsto\nu_{\bm{c}/Z}^{\vee}(s)-\nu_{\bm{c}(\mathcal{R},\bm{b})/Z}^{\vee}(s) (given (,𝒃,𝒄)(\mathcal{R},\bm{b},\bm{c}) with 𝒄𝒮1ZM(𝒃)\bm{c}\in\mathcal{S}_{1}\cap\mathcal{R}\cap Z\cdot\mathcal{B}_{M}(\bm{b})), f𝒄,7:sν𝒄/Z(s)(1B(s))f_{\bm{c},7}\colon s\mapsto\nu_{\bm{c}/Z}^{\vee}(s)\cdot(1-B(s)) (given 𝒄𝒮1\bm{c}\in\mathcal{S}_{1}). Let

(6.28) gt0(t)=f(it)ν2(i(tt0))/2π.g_{t_{0}}(t)=f(it)\nu_{2}^{\vee}(-i(t-t_{0}))/2\pi.

Note that ff, ν2\nu_{2}^{\vee} are entire (and rapidly decaying in vertical strips), so gt0g_{t_{0}} is entire (and rapidly decaying in horizontal strips). By Parseval’s theorem and (5.4), we have

𝑑tν2(it)ν2(it)=2πr>0d×rν2(r)2=2π.\int_{\mathbb{R}}dt\,\nu_{2}^{\vee}(it)\nu_{2}^{\vee}(-it)=2\pi\int_{r>0}d^{\times}{r}\,\nu_{2}(r)^{2}=2\pi.

So f(it)=𝑑t0gt0(t)ν2(i(tt0))f(it)=\int_{\mathbb{R}}dt_{0}\,g_{t_{0}}(t)\nu_{2}^{\vee}(i(t-t_{0})) by (6.28). By Fourier inversion applied to gt0g_{t_{0}}, we get

f(it)=2𝑑t0𝑑xg^t0(x)e(xt)ν2(i(tt0))=𝑑t0y>0d×yg^t0(logy2π)yitν2(i(tt0)).f(it)=\int_{\mathbb{R}^{2}}dt_{0}\,dx\,\hat{g}_{t_{0}}(x)e(xt)\nu_{2}^{\vee}(i(t-t_{0}))=\int_{\mathbb{R}}dt_{0}\int_{y>0}d^{\times}{y}\,\hat{g}_{t_{0}}(\tfrac{\log{y}}{2\pi})y^{it}\nu_{2}^{\vee}(i(t-t_{0})).

By analytic continuation, it follows that for any σ>1/2\sigma>1/2, the quantity

(6.29) (σ)𝑑sf(s)Ns(Φ𝒄,1(s)AF,1(s))\int_{(\sigma)}ds\,f(s)N^{s}(\Phi^{\bm{c},1}(s)-A_{F,1}^{\mathcal{R}}(s))

equals 𝑑t0y>0d×yg^t0(logy2π)(σ)𝑑sysν2(sit0)Ns(Φ𝒄,1(s)AF,1(s))\int_{\mathbb{R}}dt_{0}\int_{y>0}d^{\times}{y}\,\hat{g}_{t_{0}}(\tfrac{\log{y}}{2\pi})\int_{(\sigma)}ds\,y^{s}\nu_{2}^{\vee}(s-it_{0})N^{s}(\Phi^{\bm{c},1}(s)-A_{F,1}^{\mathcal{R}}(s)), and thus has absolute value at most

(6.30) 𝑑t0y>0d×y|g^t0(logy2π)||(σ)𝑑sν2(sit0)(Ny)s(Φ𝒄,1(s)AF,1(s))|.\int_{\mathbb{R}}dt_{0}\int_{y>0}d^{\times}{y}\,\lvert\hat{g}_{t_{0}}(\tfrac{\log{y}}{2\pi})\rvert\Bigl{\lvert}{\int_{(\sigma)}ds\,\nu_{2}^{\vee}(s-it_{0})(Ny)^{s}(\Phi^{\bm{c},1}(s)-A_{F,1}^{\mathcal{R}}(s))}\Bigr{\rvert}.

But for all zz\in\mathbb{C}, we have g^t0(logy2π)=𝑑tgt0(t)yit=𝑑tgt0(t+z)yi(t+z)\hat{g}_{t_{0}}(\tfrac{\log{y}}{2\pi})=\int_{\mathbb{R}}dt\,g_{t_{0}}(t)y^{-it}=\int_{\mathbb{R}}dt\,g_{t_{0}}(t+z)y^{-i(t+z)} (by the definition of g^t0\hat{g}_{t_{0}}, if z=0z=0; and then by analytic continuation, in general), and thus |g^t0(logy2π)|yIm(z)𝑑t|gt0(t+z)|\lvert\hat{g}_{t_{0}}(\tfrac{\log{y}}{2\pi})\rvert\leq y^{\operatorname{Im}(z)}\int_{\mathbb{R}}dt\,\lvert g_{t_{0}}(t+z)\rvert. Using (6.28), Proposition 5.1 (cf. (6.26)), and the definitions of 1,b\mathcal{M}_{1,b} (see (6.24)) to bound |gt0(t+z)|\lvert g_{t_{0}}(t+z)\rvert pointwise, we then get (by taking z=±iuiz=\pm iu\in i\mathbb{R})

(6.31) g^t0(logy2π)u,I,b,ν2,BM11,b(ν)(1+|t0|)b(yu+yu)1\hat{g}_{t_{0}}(\tfrac{\log{y}}{2\pi})\ll_{u,I,b,\nu_{2},B}M^{-1}\mathcal{M}_{1,b}(\nu)(1+\lvert t_{0}\rvert)^{-b}(y^{u}+y^{-u})^{-1}

for all reals u0u\geq 0 and integers b0b\geq 0. (If f=f𝒄,6f=f_{\bm{c},6}, the factor of M1M^{-1} in (6.31) arises from the bound 𝒄𝒄(,𝒃)Z/M\lVert\bm{c}-\bm{c}(\mathcal{R},\bm{b})\rVert\ll Z/M. If f=f𝒄,7f=f_{\bm{c},7}, the factor of M1M^{-1} comes from (3) over |t|M1/2\lvert t\rvert\leq M^{1/2}, and from the decay of gt0(t)g_{t_{0}}(t) in tt, tt0t-t_{0} if |t|>M1/2\lvert t\rvert>M^{1/2}.)

Note that for all σ1/3+ϵ\sigma\geq 1/3+\epsilon, we may apply (6.18) and Proposition 5.1 (after shifting contours to Re(s)=1/3+ϵ\operatorname{Re}(s)=1/3+\epsilon) to get (provided 0<ϵ1/120<\epsilon\leq 1/12)

(6.32) (σ)𝑑sν2(sit0)(Ny)sAF,1(s)ϵt𝑑t(Ny)1/3+ϵQ(𝒫)ϵ(1+|tt0|)2(Ny+Q(𝒫))1/2.\int_{(\sigma)}ds\,\nu_{2}^{\vee}(s-it_{0})(Ny)^{s}A_{F,1}^{\mathcal{R}}(s)\ll_{\epsilon}\int_{t\in\mathbb{R}}dt\,\frac{(Ny)^{1/3+\epsilon}Q(\mathscr{P})^{\epsilon}}{(1+\lvert t-t_{0}\rvert)^{2}}\ll(Ny+Q(\mathscr{P}))^{1/2}.

In view of the bound (6.30) for (6.29), we may now apply (6.31), (6.32), Conjecture 1.4 (with Z+(Ny)1/3Z+(Ny)^{1/3}, NyNy, ν2(sit0)\nu_{2}^{\vee}(s-it_{0}) in place of ZZ, NN, f(s)f(s)), and Cauchy–Schwarz to get

Σ6I𝑑t0y>0d×yM11,b(ν)(1+|t0|)b(yu+yu)(Z+(Ny)1/3)m(Ny+Q(𝒫))1/2(1+|t0|),\Sigma_{6}\ll_{I}\int_{\mathbb{R}}dt_{0}\int_{y>0}d^{\times}{y}\,\frac{M^{-1}\mathcal{M}_{1,b}(\nu)}{(1+\lvert t_{0}\rvert)^{b}(y^{u}+y^{-u})}(Z+(Ny)^{1/3})^{m}(Ny+Q(\mathscr{P}))^{1/2}(1+\lvert t_{0}\rvert),

which is IM11,b(ν)ZmN1/2\ll_{I}M^{-1}\mathcal{M}_{1,b}(\nu)Z^{m}N^{1/2} by Lemma 5.3 (provided um/3+1u\geq m/3+1 and b3b\geq 3, and NQ(𝒫)N\geq Q(\mathscr{P})). The same holds for Σ7(1B)\Sigma_{7}(1-B). Thus Conjecture 6.12 holds with A4=3A_{4}=3. ∎

Proposition 6.14 (RA1δ\delta’E).

Assume Conjectures 1.2 and 1.10. Suppose Z,N2Q(𝒫)Z,N\geq 2Q(\mathscr{P}) with NZ3N\leq Z^{3} and Q(𝒫)Zη2Q(\mathscr{P})\leq Z^{\eta_{2}}. Then the quantity (6.25) is F,IZmη2N1/21,A5\ll_{F,I}Z^{m-\eta_{2}}N^{1/2}\mathcal{M}_{1,A_{5}}. Here η2\eta_{2}, A5A_{5} are positive reals depending only on FF.

Proof.

Mimic the proof of Proposition 6.13, but take M=Zη1M=Z^{\eta_{1}}, replace Conjecture 1.8 with 1.10, replace every use of Conjecture 1.4 with GRH, and replace (6.27) with the bound Σ7Σ7(𝟏|t|Zη1)+Σ7(𝟏|t|>Zη1)\Sigma_{7}\leq\Sigma_{7}(\bm{1}_{\lvert t\rvert\leq Z^{\eta_{1}}})+\Sigma_{7}(\bm{1}_{\lvert t\rvert>Z^{\eta_{1}}}). The details simplify, since the desired final bound is not sensitive to losses of ZϵZ^{\epsilon}. (We can take η2=η1ϵ\eta_{2}=\eta_{1}-\epsilon and A5=3A_{5}=3.) ∎

When applying Propositions 6.13 and 6.14 (in §7.4), we need the following lemma:

Lemma 6.15.

Fix a real θ>0\theta>0. Suppose 𝒮m\mathcal{S}\subseteq\mathbb{Z}^{m} and |𝒮[C,C]m|Cθ\lvert\mathcal{S}\cap[-C,C]^{m}\rvert\ll C^{\theta} for all reals C1C\geq 1. Let Z,N1Z,N\geq 1 be reals. Then

(6.33) 𝒫|𝒄𝒮n1ν𝒄/Z(n/N)a¯F,1(n)|F,I,θ,ϵZθN1/3+ϵQ(𝒫)ϵ1,0.\sum_{\mathcal{R}\in\mathscr{P}}\,\Bigl{\lvert}\sum_{\bm{c}\in\mathcal{S}\cap\mathcal{R}}\sum_{n\geq 1}\nu_{\bm{c}/Z}(n/N)\bar{a}_{F,1}^{\mathcal{R}}(n)\Bigr{\rvert}\ll_{F,I,\theta,\epsilon}Z^{\theta}N^{1/3+\epsilon}Q(\mathscr{P})^{\epsilon}\mathcal{M}_{1,0}.
Proof.

By the triangle inequality, the left-hand side of (6.33) is at most

𝒫𝒄𝒮(supr>0|ν𝒄/Z(r)|)nNI|a¯F,1(n)|(sup𝒫nNI|a¯F,1(n)|)𝒄𝒮supr>0|ν𝒄/Z(r)|,\sum_{\mathcal{R}\in\mathscr{P}}\sum_{\bm{c}\in\mathcal{S}\cap\mathcal{R}}\left(\sup_{r>0}{\lvert\nu_{\bm{c}/Z}(r)\rvert}\right)\sum_{n\in N\cdot I}\lvert\bar{a}_{F,1}^{\mathcal{R}}(n)\rvert\leq\biggl{(}\,\sup_{\mathcal{R}\in\mathscr{P}}\sum_{n\in N\cdot I}\lvert\bar{a}_{F,1}^{\mathcal{R}}(n)\rvert\biggr{)}\sum_{\bm{c}\in\mathcal{S}}\sup_{r>0}{\lvert\nu_{\bm{c}/Z}(r)\rvert},

since ν\nu is supported on m×I\mathbb{R}^{m}\times I and 𝒫\mathscr{P} is a partition of m\mathbb{Z}^{m}. But sup𝒄msupr>0|ν𝒄/Z(r)|1,0(ν)\sup_{\bm{c}\in\mathbb{R}^{m}}\sup_{r>0}{\lvert\nu_{\bm{c}/Z}(r)\rvert}\leq\mathcal{M}_{1,0}(\nu) by (6.24); so 𝒄𝒮supr>0|ν𝒄/Z(r)|θZθ1,0(ν)\sum_{\bm{c}\in\mathcal{S}}\sup_{r>0}{\lvert\nu_{\bm{c}/Z}(r)\rvert}\ll_{\theta}Z^{\theta}\mathcal{M}_{1,0}(\nu) (since Suppν[1,1]m×I\operatorname{Supp}{\nu}\subseteq[-1,1]^{m}\times I). Also, sup𝒫nNI|a¯F,1(n)|I,ϵQ(𝒫)ϵN1/3+ϵ\sup_{\mathcal{R}\in\mathscr{P}}\sum_{n\in N\cdot I}\lvert\bar{a}_{F,1}^{\mathcal{R}}(n)\rvert\ll_{I,\epsilon}Q(\mathscr{P})^{\epsilon}N^{1/3+\epsilon} by (6.18). Multiplying gives (6.33). ∎

6.6. Bounding exterior squares

For each 𝒄𝒮1\bm{c}\in\mathcal{S}_{1}, let μ𝒄,2(n)\mu_{\bm{c},2}(n) denote the nnth coefficient of the Dirichlet series ζ(s)/L(s,V𝒄,2)\zeta(s)/L(s,V_{\bm{c}},\bigwedge^{2}). Recall 𝒩𝒄\mathcal{N}^{\bm{c}}, 𝒩𝒄\mathcal{N}_{\bm{c}} from (2.2). Assume Conjecture 1.2.

Proposition 6.16.

Let A2A\geq 2 be an even integer. Let Z,N,ϵ>0Z,N,\epsilon\in\mathbb{R}_{>0} with NZ3/2N\leq Z^{3/2}. Then

(6.34) 𝒄𝒮1[Z,Z]m|Nn<2N:n𝒩𝒄μ𝒄,2(n)|AA,ϵZmNA1/3+ϵ.\sum_{\bm{c}\in\mathcal{S}_{1}\cap[-Z,Z]^{m}}\,\Bigl{\lvert}{\sum_{N\leq n<2N:\,n\in\mathcal{N}^{\bm{c}}}\mu_{\bm{c},2}(n)}\Bigr{\rvert}^{A}\ll_{A,\epsilon}Z^{m}N^{A-1/3+\epsilon}.
Proof.

For N<2N<2, the left-hand side of (6.34) is AZm\ll_{A}Z^{m} by (3.2). Now assume N2N\geq 2.

Let ν0\nu_{0} be as in §1.2. Let f(Z,N)f(Z,N) denote the left-hand side of (6.34). Let g(Z,N)g(Z,N) denote f(Z,N)f(Z,N) with 𝒄𝒮1ν0(𝒄/Z)\sum_{\bm{c}\in\mathcal{S}_{1}}\nu_{0}(\bm{c}/Z)\cdots in place of 𝒄𝒮1[Z,Z]m\sum_{\bm{c}\in\mathcal{S}_{1}\cap[-Z,Z]^{m}}\cdots. Then by positivity,

(6.35) g(Z/2,N)f(Z,N)g(2Z,N)g(Z/2,N)\leq f(Z,N)\leq g(2Z,N)

for all Z>0Z>0. But if Z(2N)A+1Z\geq(2N)^{A+1}, then (if we let μ𝒄,2(𝒏)\colonequalsμ𝒄,2(n1)μ𝒄,2(nA)\mu_{\bm{c},2}(\bm{n})\colonequals\mu_{\bm{c},2}(n_{1})\cdots\mu_{\bm{c},2}(n_{A}) and P\colonequalsn1nAP\colonequals n_{1}\cdots n_{A}, and note that 𝟏gcd(P,Δ(𝒄))=1μ𝒄,2(𝒏)\bm{1}_{\gcd(P,\Delta(\bm{c}))=1}\cdot\mu_{\bm{c},2}(\bm{n}) is determined by PP and 𝒄modP\bm{c}\bmod{P})

g(Z,N)=Nn1,,nA<2N𝒄mν0(𝒄/Z)𝟏𝒄𝒮1𝟏gcd(P,Δ(𝒄))=1μ𝒄,2(𝒏)=OA(Zm)+Nn1,,nA<2N(Z/P)m𝒄(/P)m𝟏gcd(P,Δ(𝒄))=1μ𝒄,2(𝒏),\begin{split}g(Z,N)&=\sum_{N\leq n_{1},\dots,n_{A}<2N}\sum_{\bm{c}\in\mathbb{Z}^{m}}\nu_{0}(\bm{c}/Z)\bm{1}_{\bm{c}\in\mathcal{S}_{1}}\bm{1}_{\gcd(P,\Delta(\bm{c}))=1}\cdot\mu_{\bm{c},2}(\bm{n})\\ &=O_{A}(Z^{m})+\sum_{N\leq n_{1},\dots,n_{A}<2N}(Z/P)^{m}\sum_{\bm{c}\in(\mathbb{Z}/P\mathbb{Z})^{m}}\bm{1}_{\gcd(P,\Delta(\bm{c}))=1}\cdot\mu_{\bm{c},2}(\bm{n}),\end{split}

by (3.2) (or (3.8)) and Poisson summation in residue classes modulo PP (cf. the proof of Proposition 6.2); note that PNA>1P\geq N^{A}>1, so the condition gcd(P,Δ(𝒄))=1\gcd(P,\Delta(\bm{c}))=1 automatically implies 𝒄𝒮1\bm{c}\in\mathcal{S}_{1}. If g~(Z,N)\colonequalsg(Z,N)/Zm\tilde{g}(Z,N)\colonequals g(Z,N)/Z^{m}, then for Z(2N)A+1Z\geq(2N)^{A+1} we conclude that

(6.36) g~(Z,N)=g~((2N)A+1,N)+OA(1).\tilde{g}(Z,N)=\tilde{g}((2N)^{A+1},N)+O_{A}(1).

We now address Z(2N)A+1Z\leq(2N)^{A+1}. Recall 𝒮2\mathcal{S}_{2} from Proposition 6.2. If 𝒄𝒮2\bm{c}\in\mathcal{S}_{2}, then (6.12) and Proposition 3.2(8) (applied to L(s,ϕ𝒄,2)1L(s,\phi_{\bm{c},2})^{-1}), when combined with (3.2) at primes pΔ(𝒄)p\mid\Delta(\bm{c}), yield Nn<2N:n𝒩𝒄μ𝒄,2(n)ϵ𝒄ϵN1/2+ϵ\sum_{N\leq n<2N:\,n\in\mathcal{N}^{\bm{c}}}\mu_{\bm{c},2}(n)\ll_{\epsilon}\lVert\bm{c}\rVert^{\epsilon}N^{1/2+\epsilon}. If 𝒄𝒮1𝒮2\bm{c}\in\mathcal{S}_{1}\setminus\mathcal{S}_{2}, we still have the bound Nn<2N:n𝒩𝒄μ𝒄,2(n)ϵN1+ϵ\sum_{N\leq n<2N:\,n\in\mathcal{N}^{\bm{c}}}\mu_{\bm{c},2}(n)\ll_{\epsilon}N^{1+\epsilon} due to (3.2). Therefore, Proposition 6.2 yields

g(Z,N)A,ϵZm+ϵNA/2+ϵ+Zm1/2+ϵNA+ϵg(Z,N)\ll_{A,\epsilon}Z^{m+\epsilon}N^{A/2+\epsilon}+Z^{m-1/2+\epsilon}N^{A+\epsilon}

for all Z>0Z>0. It follows that g~(Z,N)A,ϵNA/2+ϵ+NA1/3+ϵ\tilde{g}(Z,N)\ll_{A,\epsilon}N^{A/2+\epsilon}+N^{A-1/3+\epsilon} for ZN2/3Z\geq N^{2/3} when Z(2N)A+1Z\leq(2N)^{A+1}, and thus (by (6.36)) for all ZN2/3Z\geq N^{2/3}. Now (6.34) follows from (6.35). ∎

To handle primes pΔ(𝒄)p\mid\Delta(\bm{c}), we prove the following (unconditional) result:

Lemma 6.17.

If Z,A,ϵ>0Z,A,\epsilon\in\mathbb{R}_{>0}, then 𝐜𝒮1[Z,Z]m|𝒩𝐜[1,N]|AA,ϵZmNϵ\sum_{\bm{c}\in\mathcal{S}_{1}\cap[-Z,Z]^{m}}\lvert\mathcal{N}_{\bm{c}}\cap[1,N]\rvert^{A}\ll_{A,\epsilon}Z^{m}N^{\epsilon}.

Proof.

Lemma 2.1 immediately suffices if Z<NAZ<N^{A}. Now suppose ZNAZ\geq N^{A}. By Hölder’s inequality, we may assume A1A\in\mathbb{Z}_{\geq 1}. The sum 𝒄𝒮1[Z,Z]m|𝒩𝒄[1,N]|A\sum_{\bm{c}\in\mathcal{S}_{1}\cap[-Z,Z]^{m}}\lvert\mathcal{N}_{\bm{c}}\cap[1,N]\rvert^{A} then equals

Σ8\colonequals𝒄𝒮1[Z,Z]mu1,,uAN:uiΔ(𝒄)1=u1,,uAN𝒄𝒮1[Z,Z]m𝟏rad(u1uA)Δ(𝒄).\Sigma_{8}\colonequals\sum_{\bm{c}\in\mathcal{S}_{1}\cap[-Z,Z]^{m}}\sum_{u_{1},\dots,u_{A}\leq N:\,u_{i}\mid\Delta(\bm{c})^{\infty}}1=\sum_{u_{1},\dots,u_{A}\leq N}\sum_{\bm{c}\in\mathcal{S}_{1}\cap[-Z,Z]^{m}}\bm{1}_{\operatorname{rad}(u_{1}\cdots u_{A})\mid\Delta(\bm{c})}.

In Σ8\Sigma_{8} we have u1uANAZu_{1}\cdots u_{A}\leq N^{A}\leq Z, so by Lang–Weil and the Chinese remainder theorem,

Σ8ϵu1,,uANZmrad(u1uA)ϵ1rNAZmrϵ1u1,,uAN:uir1.\Sigma_{8}\ll_{\epsilon}\sum_{u_{1},\dots,u_{A}\leq N}Z^{m}\operatorname{rad}(u_{1}\cdots u_{A})^{\epsilon-1}\leq\sum_{r\leq N^{A}}Z^{m}r^{\epsilon-1}\sum_{u_{1},\dots,u_{A}\leq N:\,u_{i}\mid r^{\infty}}1.

By Lemma 2.1, we conclude that Σ8A,ϵrNAZmrϵ1(Nr)ϵϵZmN(2A+1)ϵ\Sigma_{8}\ll_{A,\epsilon}\sum_{r\leq N^{A}}Z^{m}r^{\epsilon-1}(Nr)^{\epsilon}\ll_{\epsilon}Z^{m}N^{(2A+1)\epsilon}. ∎

In §7, we need the following technical complement to Proposition 6.10.

Proposition 6.18 (\bigwedge2E).

Assume Conjecture 1.2. Fix A>0A\in\mathbb{R}_{>0} and f(r){𝟏r1}Cc(>0)f(r)\in\{\bm{1}_{r\leq 1}\}\cup C^{\infty}_{c}(\mathbb{R}_{>0}). Let Z,N>0Z,N\in\mathbb{R}_{>0} with NZ3/2N\leq Z^{3/2}. Let t0t_{0}\in\mathbb{R}. For some η3=η3(A)>0\eta_{3}=\eta_{3}(A)>0, we have

(6.37) 𝒄𝒮1[Z,Z]m|n1f(n/N)nit0μ𝒄,2(n)|AA,f(1+|t0|)AZmN(1η3)A.\sum_{\bm{c}\in\mathcal{S}_{1}\cap[-Z,Z]^{m}}\,\Bigl{\lvert}{\sum_{n\geq 1}f(n/N)n^{-it_{0}}\cdot\mu_{\bm{c},2}(n)}\Bigr{\rvert}^{A}\ll_{A,f}(1+\lvert t_{0}\rvert)^{A}Z^{m}N^{(1-\eta_{3})A}.
Proof.

First suppose (f,t0)=(𝟏r1,0)(f,t_{0})=(\bm{1}_{r\leq 1},0) and A2A\in\mathbb{Z}_{\geq 2}. Let g(M)=rM:r𝒩𝒄μ𝒄,2(r)g(M)=\sum_{r\leq M:\,r\in\mathcal{N}^{\bm{c}}}\mu_{\bm{c},2}(r). By multiplicativity, nNμ𝒄,2(n)=dN:d𝒩𝒄μ𝒄,2(d)g(N/d)\sum_{n\leq N}\mu_{\bm{c},2}(n)=\sum_{d\leq N:\,d\in\mathcal{N}_{\bm{c}}}\mu_{\bm{c},2}(d)g(N/d), which is ϵNϵdN:d𝒩𝒄|g(N/d)|\ll_{\epsilon}N^{\epsilon}\sum_{d\leq N:\,d\in\mathcal{N}_{\bm{c}}}\lvert g(N/d)\rvert by (3.2). So by Hölder over dd, the left-hand side of (6.37) is at most OA,ϵ(Nϵ)O_{A,\epsilon}(N^{\epsilon}) times

Σ9\colonequals𝒄𝒮1[Z,Z]m|𝒩𝒄[1,N]|A1dN:d𝒩𝒄|g(N/d)|A.\Sigma_{9}\colonequals\sum_{\bm{c}\in\mathcal{S}_{1}\cap[-Z,Z]^{m}}\lvert\mathcal{N}_{\bm{c}}\cap[1,N]\rvert^{A-1}\sum_{d\leq N:\,d\in\mathcal{N}_{\bm{c}}}\lvert g(N/d)\rvert^{A}.

Switching 𝒄\bm{c}, dd yields Σ9dN𝒄𝒮1[Z,Z]m|𝒩𝒄[1,N]|A1|g(N/d)|A\Sigma_{9}\leq\sum_{d\leq N}\sum_{\bm{c}\in\mathcal{S}_{1}\cap[-Z,Z]^{m}}\lvert\mathcal{N}_{\bm{c}}\cap[1,N]\rvert^{A-1}\lvert g(N/d)\rvert^{A} (by positivity). Then Cauchy–Schwarz over 𝒄\bm{c}, followed by Lemma 6.17 and Proposition 6.16, gives

Σ9A,ϵdN(ZmNϵ)1/2(Zm(N/d)2A1/3+ϵ)1/2A,ϵZmNA1/6+ϵ,\Sigma_{9}\ll_{A,\epsilon}\sum_{d\leq N}(Z^{m}N^{\epsilon})^{1/2}(Z^{m}(N/d)^{2A-1/3+\epsilon})^{1/2}\ll_{A,\epsilon}Z^{m}N^{A-1/6+\epsilon},

where we use A1/6+ϵ/21+ϵ/2A-1/6+\epsilon/2\geq 1+\epsilon/2 to evaluate the sum over dd. By Hölder, then, (6.37) holds with η3=1/7A\eta_{3}=1/7A if A2A\geq 2, and with η3=1/14\eta_{3}=1/14 if 0<A20<A\leq 2. The general case follows from partial summation and Hölder, since nD(n/N)D1/N\frac{\partial}{\partial n}D(n/N)\ll_{D}1/N and nnit0|t0|/n\frac{\partial}{\partial n}n^{it_{0}}\ll\lvert t_{0}\rvert/n. ∎

(With more work, one could relax the GRH assumption. One could in fact unconditionally handle the case of very large ZZ, using (6.9)–(6.11); see [wang2021_HLH_vs_RMT]*§7.6.)

7. Adapting LL-function statistics to delta

7.1. Factorization

We need to mold the statistics from §6 into a form friendlier for the delta method. We first split the series Φ(𝒄,s)\Phi(\bm{c},s) (from (2.11)) into more manageable pieces. Recall 𝒩𝒄\mathcal{N}^{\bm{c}}, 𝒩𝒄\mathcal{N}_{\bm{c}} from (2.2). Given 𝒄𝒮1\bm{c}\in\mathcal{S}_{1}, consider the factorization Φ=ΦGΦB\Phi=\Phi^{\operatorname{G}}\Phi^{\operatorname{B}}, where

(7.1) ΦG(𝒄,s)\displaystyle\Phi^{\operatorname{G}}(\bm{c},s) \colonequalspΔ(𝒄)Φp(𝒄,s)=n𝒩𝒄S𝒄(n)ns,\displaystyle\colonequals\prod_{p\nmid\Delta(\bm{c})}\Phi_{p}(\bm{c},s)=\sum_{n\in\mathcal{N}^{\bm{c}}}S^{\natural}_{\bm{c}}(n)n^{-s},
(7.2) ΦB(𝒄,s)\displaystyle\Phi^{\operatorname{B}}(\bm{c},s) \colonequalspΔ(𝒄)Φp(𝒄,s)=n𝒩𝒄S𝒄(n)ns.\displaystyle\colonequals\prod_{p\mid\Delta(\bm{c})}\Phi_{p}(\bm{c},s)=\sum_{n\in\mathcal{N}_{\bm{c}}}S^{\natural}_{\bm{c}}(n)n^{-s}.

One can approximate ΦG\Phi^{\operatorname{G}} using Hasse–Weil LL-functions. It would be nice to also relate Φp\Phi_{p} for pΔ(𝒄)p\mid\Delta(\bm{c}) to LL-functions, even in special cases like when m=4m=4 and vp(Δ(𝒄))=1v_{p}(\Delta(\bm{c}))=1. For now, we study ΦB\Phi^{\operatorname{B}} by completely different means (see §9). In §7, we thus concentrate on ΦG\Phi^{\operatorname{G}}.

For the rest of §7, assume 2m2\mid m. We first factor ΦG\Phi^{\operatorname{G}} into three pieces: Φ𝒄,1\Phi^{\bm{c},1}, Φ𝒄,2\Phi^{\bm{c},2}, Φ𝒄,3\Phi^{\bm{c},3}.

Definition 7.1.

Let Φ𝒄,1(s)\colonequalsL(s,V𝒄)1L(1/2+s,V)1ζ(2s)1\Phi^{\bm{c},1}(s)\colonequals L(s,V_{\bm{c}})^{-1}L(1/2+s,V)^{-1}\zeta(2s)^{-1} as in (1.9), and let

(7.3) Φ𝒄,2(s)\displaystyle\Phi^{\bm{c},2}(s) \colonequalsζ(2s)/L(2s,V𝒄,2),\displaystyle\colonequals\zeta(2s)/L(2s,V_{\bm{c}},\textstyle{\bigwedge^{2}}),
(7.4) Φ𝒄,3(s)\displaystyle\Phi^{\bm{c},3}(s) \colonequalsΦG(𝒄,s)L(s,V𝒄)L(1/2+s,V)L(2s,V𝒄,2).\displaystyle\colonequals\Phi^{\operatorname{G}}(\bm{c},s)L(s,V_{\bm{c}})L(1/2+s,V)L(2s,V_{\bm{c}},\textstyle{\bigwedge^{2}}).

For each j{1,2,3}j\in\{1,2,3\}, let a𝒄,j(n)a_{\bm{c},j}(n) be the nnth coefficient of the Dirichlet series Φ𝒄,j(s)\Phi^{\bm{c},j}(s).

The factors Φ𝒄,2\Phi^{\bm{c},2}, Φ𝒄,3\Phi^{\bm{c},3} in ΦG=Φ𝒄,1Φ𝒄,2Φ𝒄,3\Phi^{\operatorname{G}}=\Phi^{\bm{c},1}\Phi^{\bm{c},2}\Phi^{\bm{c},3} hinder any attempt to apply statistics on Φ𝒄,1\Phi^{\bm{c},1} to ΦG\Phi^{\operatorname{G}}. Fortunately, Φ𝒄,2\Phi^{\bm{c},2}, Φ𝒄,3\Phi^{\bm{c},3} turn out to behave as “error factors” on average. Proposition 6.18 lets us handle large moduli in Φ𝒄,2\Phi^{\bm{c},2}; note that a𝒄,2(n)=0a_{\bm{c},2}(n)=0 if nn is not a square, and

(7.5) a𝒄,2(n)=μ𝒄,2(n1/2)a_{\bm{c},2}(n)=\mu_{\bm{c},2}(n^{1/2})

otherwise (where μ𝒄,2\mu_{\bm{c},2} is as in §6.6). We now prove results to handle large moduli in Φ𝒄,3\Phi^{\bm{c},3}.

The factor Φ𝒄,3\Phi^{\bm{c},3} measures the quality of Φ𝒄,1Φ𝒄,2\Phi^{\bm{c},1}\Phi^{\bm{c},2} as an approximation to ΦG\Phi^{\operatorname{G}}. Recall the “first-order approximation” Φ(𝒄,s)=Ψ𝒄,1(s)Ψ𝒄,2(s)\Phi(\bm{c},s)=\Psi^{\bm{c},1}(s)\Psi^{\bm{c},2}(s) given by Ψ𝒄,1\Psi^{\bm{c},1}, Ψ𝒄,2\Psi^{\bm{c},2} from (2.18); here Ψ𝒄,1(s)=1/L(s,V𝒄)\Psi^{\bm{c},1}(s)=1/L(s,V_{\bm{c}}). The “first-order error” Ψ𝒄,2(s)\Psi^{\bm{c},2}(s) is only expected to converge absolutely for Re(s)>1/2\operatorname{Re}(s)>1/2. As suggested in §2, this is a “source of ϵ\epsilon” in (2.17). On the other hand, the following result establishes absolute convergence for Φ𝒄,3(s)\Phi^{\bm{c},3}(s) past the critical line Re(s)=1/2\operatorname{Re}(s)=1/2.

Proposition 7.2.

Uniformly over 𝐜𝒮1\bm{c}\in\mathcal{S}_{1}, primes pp, and integers l1l\geq 1, we have

(7.6) a𝒄,3(p)𝟏pΔ(𝒄)=0,a𝒄,3(p2)𝟏pΔ(𝒄)p1/2,a𝒄,3(pl)ϵplϵ.a_{\bm{c},3}(p)\cdot\bm{1}_{p\nmid\Delta(\bm{c})}=0,\quad a_{\bm{c},3}(p^{2})\cdot\bm{1}_{p\nmid\Delta(\bm{c})}\ll p^{-1/2},\quad a_{\bm{c},3}(p^{l})\ll_{\epsilon}p^{l\epsilon}.

In particular, if 𝐜𝒮1\bm{c}\in\mathcal{S}_{1}, then Φ𝐜,3(s)\Phi^{\bm{c},3}(s) converges absolutely over Re(s)>1/3\operatorname{Re}(s)>1/3.

Proof.

Let 𝒄𝒮1\bm{c}\in\mathcal{S}_{1}. Let λ𝒄\colonequalsλV𝒄\lambda^{\natural}_{\bm{c}}\colonequals\lambda^{\natural}_{V_{\bm{c}}}, as in §6.1. Suppose first that pΔ(𝒄)p\mid\Delta(\bm{c}). Then ΦpG(𝒄,s)=1\Phi^{\operatorname{G}}_{p}(\bm{c},s)=1 by (7.1). So by (7.4) and (3.2), we have a𝒄,3(pl)ϵplϵa_{\bm{c},3}(p^{l})\ll_{\epsilon}p^{l\epsilon}. So (7.6) holds.

Now suppose pΔ(𝒄)p\nmid\Delta(\bm{c}). Then Φp(𝒄,s)=1+S𝒄(p)ps\Phi_{p}(\bm{c},s)=1+S^{\natural}_{\bm{c}}(p)p^{-s} by (2.14), and

S𝒄(p)=E𝒄(p)p1/2EF(p)=λ𝒄(p)p1/2λV(p)S^{\natural}_{\bm{c}}(p)=E^{\natural}_{\bm{c}}(p)-p^{-1/2}E^{\natural}_{F}(p)=-\lambda^{\natural}_{\bm{c}}(p)-p^{-1/2}\lambda^{\natural}_{V}(p)

by (2.13), (3.4), and (3.3). In particular, a𝒄,3(pl)ϵplϵa_{\bm{c},3}(p^{l})\ll_{\epsilon}p^{l\epsilon} by (7.4) and (3.2). Furthermore,

Φp(𝒄,s)Lp(s,V𝒄)=(1λ𝒄(p)psλV(p)p1/2s)(1+λ𝒄(p)ps+λ𝒄(p2)p2s+O(p3s))=1λV(p)p1/2s+[λ𝒄(p2)λ𝒄(p)2]p2s+O(p1/22s)+O(p3s).\begin{split}\Phi_{p}(\bm{c},s)L_{p}(s,V_{\bm{c}})&=(1-\lambda^{\natural}_{\bm{c}}(p)p^{-s}-\lambda^{\natural}_{V}(p)p^{-1/2-s})(1+\lambda^{\natural}_{\bm{c}}(p)p^{-s}+\lambda^{\natural}_{\bm{c}}(p^{2})p^{-2s}+O(p^{-3s}))\\ &=1-\lambda^{\natural}_{V}(p)p^{-1/2-s}+[\lambda^{\natural}_{\bm{c}}(p^{2})-\lambda^{\natural}_{\bm{c}}(p)^{2}]p^{-2s}+O(p^{-1/2-2s})+O(p^{-3s}).\end{split}

To get further cancellation, we multiply Φp(𝒄,s)Lp(s,V𝒄)\Phi_{p}(\bm{c},s)L_{p}(s,V_{\bm{c}}) by

Lp(1/2+s,V)=1+λV(p)p1/2s+O(p12s),Lp(2s,V𝒄,2)=1+λV𝒄,2(p)p2s+O(p4s),\begin{split}L_{p}(1/2+s,V)&=1+\lambda^{\natural}_{V}(p)p^{-1/2-s}+O(p^{-1-2s}),\\ L_{p}(2s,V_{\bm{c}},\textstyle{\bigwedge^{2}})&=1+\lambda^{\natural}_{V_{\bm{c}},\bigwedge^{2}}(p)p^{-2s}+O(p^{-4s}),\end{split}

to get (in view of λV𝒄,2(p)=λ𝒄(p)2λ𝒄(p2)\lambda^{\natural}_{V_{\bm{c}},\bigwedge^{2}}(p)=\lambda^{\natural}_{\bm{c}}(p)^{2}-\lambda^{\natural}_{\bm{c}}(p^{2}) from (3.6))

(7.7) Φp(𝒄,s)Lp(s,V𝒄)Lp(1/2+s,V)Lp(2s,V𝒄,2)=1+O(p1/22s)+O(p3s).\Phi_{p}(\bm{c},s)L_{p}(s,V_{\bm{c}})L_{p}(1/2+s,V)L_{p}(2s,V_{\bm{c}},\textstyle{\bigwedge^{2}})=1+O(p^{-1/2-2s})+O(p^{-3s}).

By (7.4), the left-hand side of (7.7) is precisely the local factor Φp𝒄,3(s)\Phi^{\bm{c},3}_{p}(s) of Φ𝒄,3(s)\Phi^{\bm{c},3}(s). Thus (7.7) completes the proof of (7.6). The convergence statement on Φ𝒄,3(s)\Phi^{\bm{c},3}(s) follows from (7.6). ∎

Corollary 7.3 (Φ\Phi3E).

Fix A>0A\in\mathbb{R}_{>0}. Then uniformly over Z,N>0Z,N\in\mathbb{R}_{>0}, we have

(7.8) 𝒄𝒮1[Z,Z]m(nN|a𝒄,3(n)|)AA,ϵZmN(1/3+ϵ)A.\sum_{\bm{c}\in\mathcal{S}_{1}\cap[-Z,Z]^{m}}\biggl{(}\,\sum_{n\leq N}\lvert a_{\bm{c},3}(n)\rvert\biggr{)}^{\!A}\ll_{A,\epsilon}Z^{m}N^{(1/3+\epsilon)A}.
Proof.

Recall 𝒩𝒄\mathcal{N}^{\bm{c}}, 𝒩𝒄\mathcal{N}_{\bm{c}} from (2.2). By multiplicativity (of |a𝒄,3|\lvert a_{\bm{c},3}\rvert) and positivity, we have

(7.9) nN|a𝒄,3(n)|dN:d𝒩𝒄|a𝒄,3(d)|rN:r𝒩𝒄|a𝒄,3(r)|.\sum_{n\leq N}\lvert a_{\bm{c},3}(n)\rvert\leq\sum_{d\leq N:\,d\in\mathcal{N}_{\bm{c}}}\lvert a_{\bm{c},3}(d)\rvert\sum_{r\leq N:\,r\in\mathcal{N}^{\bm{c}}}\lvert a_{\bm{c},3}(r)\rvert.

But for any ϵ>0\epsilon>0, the bounds in (7.6) imply a𝒄,3(d)ϵdϵa_{\bm{c},3}(d)\ll_{\epsilon}d^{\epsilon}, and that rN:r𝒩𝒄|a𝒄,3(r)|\sum_{r\leq N:\,r\in\mathcal{N}^{\bm{c}}}\lvert a_{\bm{c},3}(r)\rvert is N1/3+ϵr𝒩𝒄r1/3ϵ|a𝒄,3(r)|ϵN1/3+ϵ\leq N^{1/3+\epsilon}\sum_{r\in\mathcal{N}^{\bm{c}}}r^{-1/3-\epsilon}\lvert a_{\bm{c},3}(r)\rvert\ll_{\epsilon}N^{1/3+\epsilon} (cf. (6.18)). So by (7.9), the left-hand side of (7.8) is A,ϵN(1/3+2ϵ)A𝒄𝒮1[Z,Z]m|𝒩𝒄[1,N]|A1\ll_{A,\epsilon}N^{(1/3+2\epsilon)A}\sum_{\bm{c}\in\mathcal{S}_{1}\cap[-Z,Z]^{m}}\lvert\mathcal{N}_{\bm{c}}\cap[1,N]\rvert^{A-1}. Now (7.8) follows from Lemma 6.17. ∎

7.2. From (R2’E) to (R2’E’)

We now build on Conjecture 6.9 (R2’E).

Conjecture 7.4 (R2’E’).

Fix a function DCc(>0)D\in C^{\infty}_{c}(\mathbb{R}_{>0}). Let tt\in\mathbb{R}. Let Z,N,ϵ>0Z,N,\epsilon\in\mathbb{R}_{>0} with NZ3N\leq Z^{3} and ϵ1\epsilon\leq 1. Then for some real A6=A6(F,ϵ)>0A_{6}=A_{6}(F,\epsilon)>0, we have

(7.10) 𝒄𝒮1[Z,Z]m|n𝒩𝒄D(n/N)nitS𝒄(n)|2ϵF,D,ϵ(1+|t|)A6ZmN(2ϵ)/2.\sum_{\bm{c}\in\mathcal{S}_{1}\cap[-Z,Z]^{m}}\,\Bigl{\lvert}{\sum_{n\in\mathcal{N}^{\bm{c}}}D(n/N)\cdot n^{-it}S^{\natural}_{\bm{c}}(n)}\Bigr{\rvert}^{2-\epsilon}\ll_{F,D,\epsilon}(1+\lvert t\rvert)^{A_{6}}Z^{m}N^{(2-\epsilon)/2}.
Proposition 7.5.

Assume Conjectures 1.2 and 6.9. Then Conjecture 7.4 holds.

Proof.

Let ν2Cc(>0)\nu_{2}\in C^{\infty}_{c}(\mathbb{R}_{>0}) be as in §5, so that Suppν2[1,2]\operatorname{Supp}{\nu_{2}}\subseteq[1,2] and we have (5.4). In view of (7.1) and the factorization ΦG=Φ𝒄,1Φ𝒄,2Φ𝒄,3\Phi^{\operatorname{G}}=\Phi^{\bm{c},1}\Phi^{\bm{c},2}\Phi^{\bm{c},3}, we may use Lemma 5.2 (with k=3k=3 and a(𝒏)=1j3(njita𝒄,j(nj))a(\bm{n})=\prod_{1\leq j\leq 3}(n_{j}^{-it}a_{\bm{c},j}(n_{j})), and f(𝒓)=D(r1r2r3/N)f(\bm{r})=D(r_{1}r_{2}r_{3}/N)) to write

(7.11) n𝒩𝒄D(n/N)nitS𝒄(n)=(2π)3𝑵1/2d×𝑵𝒕3𝑑𝒕g𝑵(i𝒕)1j3Σ10,𝑵𝒄,j(𝒕),\sum_{n\in\mathcal{N}^{\bm{c}}}D(n/N)\cdot n^{-it}S^{\natural}_{\bm{c}}(n)=(2\pi)^{-3}\int_{\bm{N}\geq 1/2}d^{\times}\bm{N}\int_{\bm{t}\in\mathbb{R}^{3}}d\bm{t}\,g_{\bm{N}}^{\vee}(i\bm{t})\prod_{1\leq j\leq 3}\Sigma_{10,\bm{N}}^{\bm{c},j}(\bm{t}),

where 𝑵=(N1,N2,N3)\bm{N}=(N_{1},N_{2},N_{3}), where g𝑵(𝒓)\colonequalsD(r1r2r3/N)1j3ν2(rj/Nj)g_{\bm{N}}(\bm{r})\colonequals D(r_{1}r_{2}r_{3}/N)\prod_{1\leq j\leq 3}\nu_{2}(r_{j}/N_{j}), and where

(7.12) Σ10,𝑵𝒄,j(𝒕)=Σ10,𝑵𝒄,j(t1,t2,t3)\colonequalsnj1ν2(nj/Nj)nji(tj+t)a𝒄,j(nj).\Sigma_{10,\bm{N}}^{\bm{c},j}(\bm{t})=\Sigma_{10,\bm{N}}^{\bm{c},j}(t_{1},t_{2},t_{3})\colonequals\sum_{n_{j}\geq 1}\nu_{2}(n_{j}/N_{j})n_{j}^{-i(t_{j}+t)}a_{\bm{c},j}(n_{j}).

(Note that NN, tt are independent of 𝑵\bm{N}, 𝒕\bm{t}.) For all 𝑵1/2\bm{N}\geq 1/2 and b0b\geq 0, Proposition 5.1 gives

(7.13) g𝑵(i𝒕)b,D,ν2(1+𝒕)b.g_{\bm{N}}^{\vee}(i\bm{t})\ll_{b,D,\nu_{2}}(1+\lVert\bm{t}\rVert)^{-b}.

Fix an integer A1A\geq 1 for which SuppD[A1,A]\operatorname{Supp}{D}\subseteq[A^{-1},A]. If 𝒓Suppg𝑵\bm{r}\in\operatorname{Supp}{g_{\bm{N}}}, then N/Ar1r2r3ANN/A\leq r_{1}r_{2}r_{3}\leq AN and Njrj2NjN_{j}\leq r_{j}\leq 2N_{j} for all jj, so 𝑵\bm{N} lies in the set

(7.14) 10\colonequals{𝑵1/2:N1N2N3[N/8A,AN]}.\mathscr{R}_{10}\colonequals\{\bm{N}\geq 1/2:N_{1}N_{2}N_{3}\in[N/8A,AN]\}.

Thus the equality (7.11) remains true if we restrict the integral over 𝑵1/2\bm{N}\geq 1/2 to the region 10\mathscr{R}_{10}. Now set W1(𝑵,𝒕)\colonequalsN1η|g𝑵(i𝒕)|W_{1}(\bm{N},\bm{t})\colonequals N_{1}^{\eta}\cdot\lvert g_{\bm{N}}^{\vee}(i\bm{t})\rvert and W2(𝑵,𝒕)\colonequalsN1(β1)η|g𝑵(i𝒕)|W_{2}(\bm{N},\bm{t})\colonequals N_{1}^{-(\beta-1)\eta}\cdot\lvert g_{\bm{N}}^{\vee}(i\bm{t})\rvert, for a small constant η>0\eta>0 to be chosen later. Let 1\colonequals10×3d×𝑵𝑑𝒕W1(𝑵,𝒕)\mathcal{I}_{1}\colonequals\int_{\mathscr{R}_{10}\times\mathbb{R}^{3}}d^{\times}\bm{N}\,d\bm{t}\,W_{1}(\bm{N},\bm{t}). Letting β\colonequals2ϵ[1,2)\beta\colonequals 2-\epsilon\in[1,2), and using Hölder over (log𝑵,𝒕)3×3(\log\bm{N},\bm{t})\subseteq\mathbb{R}^{3}\times\mathbb{R}^{3} (restricted to 𝑵10\bm{N}\in\mathscr{R}_{10}), we obtain

(7.15) |n𝒩𝒄D(n/N)nitS𝒄(n)|β1β110×3d×𝑵𝑑𝒕W2(𝑵,𝒕)1j3|Σ10,𝑵𝒄,j(𝒕)|β,\Bigl{\lvert}{\sum_{n\in\mathcal{N}^{\bm{c}}}D(n/N)n^{-it}S^{\natural}_{\bm{c}}(n)}\Bigr{\rvert}^{\beta}\leq\mathcal{I}_{1}^{\beta-1}\cdot\int_{\mathscr{R}_{10}\times\mathbb{R}^{3}}d^{\times}\bm{N}\,d\bm{t}\,W_{2}(\bm{N},\bm{t})\prod_{1\leq j\leq 3}\lvert\Sigma_{10,\bm{N}}^{\bm{c},j}(\bm{t})\rvert^{\beta},

since β1\beta\geq 1 and W1(𝑵,𝒕)β1W2(𝑵,𝒕)=|g𝑵(i𝒕)|βW_{1}(\bm{N},\bm{t})^{\beta-1}\cdot W_{2}(\bm{N},\bm{t})=\lvert g_{\bm{N}}^{\vee}(i\bm{t})\rvert^{\beta}.

By (7.13) and (7.14), we have 1D,ν210d×𝑵N1ηA,ηNηN2,N31/2d×N2d×N3(N2N3)ηηNη\mathcal{I}_{1}\ll_{D,\nu_{2}}\int_{\mathscr{R}_{10}}d^{\times}\bm{N}\,N_{1}^{\eta}\ll_{A,\eta}N^{\eta}\int_{N_{2},N_{3}\geq 1/2}\frac{d^{\times}{N_{2}}\,d^{\times}{N_{3}}}{(N_{2}N_{3})^{\eta}}\ll_{\eta}N^{\eta}. Upon summing (7.15) over 𝒄𝒮1[Z,Z]m\bm{c}\in\mathcal{S}_{1}\cap[-Z,Z]^{m}, we thus find that the left-hand side of (7.10) is

(7.16) D,ν2,η10×3d×𝑵𝑑𝒕(N/N1)(β1)η|g𝑵(i𝒕)|𝒄𝒮1[Z,Z]m1j3|Σ10,𝑵𝒄,j(𝒕)|β.\ll_{D,\nu_{2},\eta}\int_{\mathscr{R}_{10}\times\mathbb{R}^{3}}d^{\times}\bm{N}\,d\bm{t}\,(N/N_{1})^{(\beta-1)\eta}\cdot\lvert g_{\bm{N}}^{\vee}(i\bm{t})\rvert\sum_{\bm{c}\in\mathcal{S}_{1}\cap[-Z,Z]^{m}}\prod_{1\leq j\leq 3}\lvert\Sigma_{10,\bm{N}}^{\bm{c},j}(\bm{t})\rvert^{\beta}.

Now let (γ1,γ2,γ3)\colonequals(2,4β/ϵ,4β/ϵ)(\gamma_{1},\gamma_{2},\gamma_{3})\colonequals(2,4\beta/\epsilon,4\beta/\epsilon). Then 1j3β/γj=1\sum_{1\leq j\leq 3}\beta/\gamma_{j}=1 (since β=2ϵ\beta=2-\epsilon), so by Hölder over 𝒄\bm{c} (writing 10=1j3Σ10,𝑵𝒄,j(𝒕)\heartsuit_{10}=\prod_{1\leq j\leq 3}\Sigma_{10,\bm{N}}^{\bm{c},j}(\bm{t}) and j=Σ10,𝑵𝒄,j(𝒕)𝒄γj(𝒮1[Z,Z]m)γj\mathscr{M}_{j}=\lVert\Sigma_{10,\bm{N}}^{\bm{c},j}(\bm{t})\rVert_{\ell^{\gamma_{j}}_{\bm{c}}(\mathcal{S}_{1}\cap[-Z,Z]^{m})}^{\gamma_{j}} for brevity),

(7.17) 𝒄𝒮1[Z,Z]m|10|β=10𝒄β(𝒮1[Z,Z]m)β1j3jβ/γj.\sum_{\bm{c}\in\mathcal{S}_{1}\cap[-Z,Z]^{m}}\lvert\heartsuit_{10}\rvert^{\beta}=\lVert\heartsuit_{10}\rVert_{\ell^{\beta}_{\bm{c}}(\mathcal{S}_{1}\cap[-Z,Z]^{m})}^{\beta}\leq\prod_{1\leq j\leq 3}\mathscr{M}_{j}^{\beta/\gamma_{j}}.

We now bound the necessary γj\ell^{\gamma_{j}}-norms. First, 1ν2(1+|t1+t|)A3(Z+N11/3)mN1γ1/2\mathscr{M}_{1}\ll_{\nu_{2}}(1+\lvert t_{1}+t\rvert)^{A_{3}}(Z+N_{1}^{1/3})^{m}N_{1}^{\gamma_{1}/2}, by (7.12) and Conjecture 6.9 (with Z+N11/3Z+N_{1}^{1/3}, N1N_{1} in place of ZZ, NN). Second, by (7.12), (7.5), and (6.37) (with A=γ2A=\gamma_{2}, with f=ν2f=\nu_{2}, and with Z+N21/3Z+N_{2}^{1/3}, N21/2N_{2}^{1/2} in place of ZZ, NN), we have 2γ2,ν2(1+|t2+t|)γ2(Z+N21/3)m(N21/2)(1η3(γ2))γ2\mathscr{M}_{2}\ll_{\gamma_{2},\nu_{2}}(1+\lvert t_{2}+t\rvert)^{\gamma_{2}}(Z+N_{2}^{1/3})^{m}(N_{2}^{1/2})^{(1-\eta_{3}(\gamma_{2}))\gamma_{2}}. Third, by (7.12), the bound ν2(n3/N3)n3i(t3+t)ν21\nu_{2}(n_{3}/N_{3})n_{3}^{-i(t_{3}+t)}\ll_{\nu_{2}}1, and Corollary 7.3, we have 3γ3,ν2ZmN311γ3/30\mathscr{M}_{3}\ll_{\gamma_{3},\nu_{2}}Z^{m}N_{3}^{11\gamma_{3}/30}.

Plugging into (7.17), and writing 1+|tj+t|(1+𝒕)(1+|t|)1+\lvert t_{j}+t\rvert\leq(1+\lVert\bm{t}\rVert)(1+\lvert t\rvert), we get that if 𝑵10\bm{N}\in\mathscr{R}_{10}, then the left-hand side of (7.17) is D,ν2,β(1+𝒕)A6(1+|t|)A6\ll_{D,\nu_{2},\beta}(1+\lVert\bm{t}\rVert)^{A_{6}}(1+\lvert t\rvert)^{A_{6}} times

Zjmβ/γjN1β/2N2(1η3(γ2))β/2N311β/30D,βZm(N1/2N2η3(γ2)/2N34/30)β,Z^{\sum_{j}m\beta/\gamma_{j}}N_{1}^{\beta/2}N_{2}^{(1-\eta_{3}(\gamma_{2}))\beta/2}N_{3}^{11\beta/30}\asymp_{D,\beta}Z^{m}(N^{1/2}N_{2}^{-\eta_{3}(\gamma_{2})/2}N_{3}^{-4/30})^{\beta},

where A6=(A3β/γ1+γ2β/γ2)=(A3/2+1)βA_{6}=(A_{3}\beta/\gamma_{1}+\gamma_{2}\beta/\gamma_{2})=(A_{3}/2+1)\beta. Upon plugging this result into (7.16), we get that the left-hand side of (7.10) is at most OD,ν2,β(1)O_{D,\nu_{2},\beta}(1) times

(7.18) 10×3d×𝑵𝑑𝒕(N/N1)(β1)η|g𝑵(i𝒕)|(1+𝒕)A6(1+|t|)A6ZmNβ/2(N2η3(γ2)/2N34/30)β.\int_{\mathscr{R}_{10}\times\mathbb{R}^{3}}d^{\times}\bm{N}\,d\bm{t}\,(N/N_{1})^{(\beta-1)\eta}\cdot\lvert g_{\bm{N}}^{\vee}(i\bm{t})\rvert\cdot(1+\lVert\bm{t}\rVert)^{A_{6}}\frac{(1+\lvert t\rvert)^{A_{6}}Z^{m}N^{\beta/2}}{(N_{2}^{\eta_{3}(\gamma_{2})/2}N_{3}^{4/30})^{\beta}}.

Finally, let η=min(η3(γ2)/2,4/30)\eta=\min(\eta_{3}(\gamma_{2})/2,4/30). Then for each 𝑵10\bm{N}\in\mathscr{R}_{10}, we have (N/N1)ηD,βN2η3(γ2)/2N34/30(N/N_{1})^{\eta}\ll_{D,\beta}N_{2}^{\eta_{3}(\gamma_{2})/2}N_{3}^{4/30}. Since β10\beta-1\geq 0, it follows that the quantity (7.18) is

D,β10×3d×𝑵𝑑𝒕|g𝑵(i𝒕)|(1+𝒕)A6(1+|t|)A6ZmNβ/2N2η3(γ2)/2N34/30D,β(1+|t|)A6ZmNβ/2,\begin{split}\ll_{D,\beta}\int_{\mathscr{R}_{10}\times\mathbb{R}^{3}}d^{\times}\bm{N}\,d\bm{t}\,\lvert g_{\bm{N}}^{\vee}(i\bm{t})\rvert\cdot(1+\lVert\bm{t}\rVert)^{A_{6}}\frac{(1+\lvert t\rvert)^{A_{6}}Z^{m}N^{\beta/2}}{N_{2}^{\eta_{3}(\gamma_{2})/2}N_{3}^{4/30}}\ll_{D,\beta}(1+\lvert t\rvert)^{A_{6}}Z^{m}N^{\beta/2},\end{split}

where we have used (7.13) to integrate over 𝒕\bm{t}, and (7.14) to integrate first over N1N_{1} (given N2N_{2}, N3N_{3}) and then over N2,N31/2N_{2},N_{3}\geq 1/2. Therefore, (7.10) holds (with A6=(A3/2+1)(2ϵ)A_{6}=(A_{3}/2+1)(2-\epsilon)). ∎

7.3. Handling variation of “error factors” for small fixed “error moduli”

We would like to build on Propositions 6.13 and 6.14, but we must first improve our understanding of certain local factors. For each 𝒄𝒮1\bm{c}\in\mathcal{S}_{1}, recall Φ𝒄,j(s)\Phi^{\bm{c},j}(s), a𝒄,j(n)a_{\bm{c},j}(n) from Definition 7.1, and let a𝒄(n)a^{\prime}_{\bm{c}}(n) denote the nnth coefficient of the Dirichlet series

(7.19) Φ𝒄,2Φ𝒄,3=ΦG/Φ𝒄,1=ΦG(𝒄,s)L(s,V𝒄)L(1/2+s,V)ζ(2s),\Phi^{\bm{c},2}\Phi^{\bm{c},3}=\Phi^{\operatorname{G}}/\Phi^{\bm{c},1}=\Phi^{\operatorname{G}}(\bm{c},s)L(s,V_{\bm{c}})L(1/2+s,V)\zeta(2s),

so that for all n1n\geq 1, we have

(7.20) S𝒄(n)𝟏n𝒩𝒄=n1d=na𝒄,1(n1)a𝒄(d).S^{\natural}_{\bm{c}}(n)\bm{1}_{n\in\mathcal{N}^{\bm{c}}}=\sum_{n_{1}d=n}a_{\bm{c},1}(n_{1})a^{\prime}_{\bm{c}}(d).

By Proposition 7.2, a𝒄,3(n)ϵnϵa_{\bm{c},3}(n)\ll_{\epsilon}n^{\epsilon}, and by (3.2) and Definition 7.1, a𝒄,2(n)ϵnϵa_{\bm{c},2}(n)\ll_{\epsilon}n^{\epsilon}; so certainly

(7.21) a𝒄(d)ϵdϵ.a^{\prime}_{\bm{c}}(d)\ll_{\epsilon}d^{\epsilon}.

Moreover, if dd is small (or fixed), we would like a𝒄(d)a^{\prime}_{\bm{c}}(d) to not vary too wildly with 𝒄\bm{c}.

Given n1n\in\mathbb{Z}_{\geq 1}, what data does a𝒄(n)a^{\prime}_{\bm{c}}(n) depend on? Note that L(1/2+s,V)ζ(2s)L(1/2+s,V)\zeta(2s) is fixed (in terms of FF). So by (7.19) and (7.1), the coefficient a𝒄(n)a^{\prime}_{\bm{c}}(n) is determined by the residue class 𝒄+nm\bm{c}+n\mathbb{Z}^{m} and the local factors Lp(s,V𝒄)L_{p}(s,V_{\bm{c}}) for pnp\mid n. Therefore, if we define l(p,𝒄)l(p,\bm{c}) as in Lemma 4.4, then a𝒄(n)a^{\prime}_{\bm{c}}(n) is determined by the residue class 𝒄+r(n,𝒄)m\bm{c}+r(n,\bm{c})\mathbb{Z}^{m}, where

(7.22) r(n,𝒄)\colonequalspnlcm(pvp(n),pl(p,𝒄)+1).r(n,\bm{c})\colonequals\prod_{p\mid n}\operatorname{lcm}(p^{v_{p}(n)},p^{l(p,\bm{c})+1}).
Proposition 7.6.

Let n,q1n,q\geq 1. Let 𝐚,𝐛𝒮1\bm{a},\bm{b}\in\mathcal{S}_{1} and suppose 𝐚𝐛modq\bm{a}\equiv\bm{b}\bmod{q}. Let pnp\mid n be a prime. Then vp(r(n,𝐚))vp(q)v_{p}(r(n,\bm{a}))\leq v_{p}(q) if and only if vp(r(n,𝐛))vp(q)v_{p}(r(n,\bm{b}))\leq v_{p}(q).

Proof.

By symmetry, it suffices to prove the “if” direction. So, say vp(q)vp(r(n,𝒃))v_{p}(q)\geq v_{p}(r(n,\bm{b})). Then vp(q)l(p,𝒃)+1v_{p}(q)\geq l(p,\bm{b})+1, so l(p,𝒂)=l(p,𝒃)l(p,\bm{a})=l(p,\bm{b}) by (4.6). Thus vp(r(n,𝒂))=vp(r(n,𝒃))vp(q)v_{p}(r(n,\bm{a}))=v_{p}(r(n,\bm{b}))\leq v_{p}(q). ∎

In what follows, recall the notation qq_{\mathcal{R}} from Definition 6.11.

Definition 7.7.

For every n1n\geq 1, let (n)\mathscr{E}(n) be the set of residue classes m\mathcal{R}\subseteq\mathbb{Z}^{m} for which there exists a tuple 𝒄𝒮1\bm{c}\in\mathcal{S}_{1}\cap\mathcal{R} with r(n,𝒄)qr(n,\bm{c})\nmid q_{\mathcal{R}}.

We now construct a partition 𝒫(n,k)={}\mathscr{P}(n,k)=\{\mathcal{R}\} of m\mathbb{Z}^{m} into residue classes m\mathcal{R}\subseteq\mathbb{Z}^{m}.

Definition 7.8.

Fix integers n,k1n,k\geq 1. We define 𝒫(n,k)\mathscr{P}(n,k) through a recursive decomposition process. Let 𝒮0\colonequals{𝒂+nm:1𝒂n}\mathscr{S}_{0}\colonequals\{\bm{a}+n\mathbb{Z}^{m}:1\leq\bm{a}\leq n\} denote the partition of m\mathbb{Z}^{m} into the nmn^{m} residue classes modulo nn. For each j0j\geq 0, define 𝒮j+1\mathscr{S}_{j+1} in terms of 𝒮j\mathscr{S}_{j} as follows:

  1. (1)

    If possible, choose a residue class 𝒮j(n)\mathcal{R}\in\mathscr{S}_{j}\cap\mathscr{E}(n) with qnk1q_{\mathcal{R}}\leq n^{k-1}. Otherwise, let 𝒮j+1\colonequals𝒮j\mathscr{S}_{j+1}\colonequals\mathscr{S}_{j}, and skip step (2).

  2. (2)

    Write =𝒂+qm\mathcal{R}=\bm{a}+q\mathbb{Z}^{m}, with q1q\geq 1. Choose 𝒄𝒮1\bm{c}\in\mathcal{S}_{1}\cap\mathcal{R} with r(n,𝒄)qr(n,\bm{c})\nmid q. Choose a prime pnp\mid n with vp(r(n,𝒄))>vp(q)v_{p}(r(n,\bm{c}))>v_{p}(q). Create 𝒮j+1\mathscr{S}_{j+1} by replacing the element 𝒮j\mathcal{R}\in\mathscr{S}_{j} with the pmp^{m} lifted residue classes (𝒂+q𝒊)+pqm(\bm{a}+q\bm{i})+pq\mathbb{Z}^{m} (with 1𝒊p1\leq\bm{i}\leq p, say). Formally,

    𝒮j+1\colonequals(𝒮j{}){(𝒂+q𝒊)+pqm:1𝒊p}.\mathscr{S}_{j+1}\colonequals(\mathscr{S}_{j}\setminus\{\mathcal{R}\})\cup\{(\bm{a}+q\bm{i})+pq\mathbb{Z}^{m}:1\leq\bm{i}\leq p\}.

Step (2) can only occur finitely many times (because we require qnk1q_{\mathcal{R}}\leq n^{k-1} in step (1)). Let j0\colonequalsmin{j0:𝒮j+1=𝒮j}j_{0}\colonequals\min{\{j\geq 0:\mathscr{S}_{j+1}=\mathscr{S}_{j}\}}. Let 𝒫(n,k)\colonequals𝒮j0\mathscr{P}(n,k)\colonequals\mathscr{S}_{j_{0}}.555The result may depend on the choices we make at each step, but all of our estimates based on 𝒫(n,k)\mathscr{P}(n,k) will apply uniformly over all possible outcomes. Let (n,k)\colonequals𝒫(n,k)(n)\mathscr{E}(n,k)\colonequals\mathscr{P}(n,k)\cap\mathscr{E}(n).

In Definition 7.8, we allow the initial 𝒮0\mathscr{S}_{0} to branch into many different moduli. If we did not do this, then to control a𝒄(n)a^{\prime}_{\bm{c}}(n) for nMn\leq M might require us to work with moduli exponentially large in MM (for some values of nn), which would be fatal to our approach to Theorem 1.9 (though perhaps OK for Theorem 1.6). We will eventually apply Propositions 6.13 and 6.14 in residue classes 𝒫(n,k)(n)\mathcal{R}\in\mathscr{P}(n,k)\setminus\mathscr{E}(n), for some values of n,k1n,k\geq 1. We first unravel the structure of 𝒫(n,k)\mathscr{P}(n,k), and provide some control on the exceptional set (n)\mathscr{E}(n).

Lemma 7.9.

Let n,k1n,k\geq 1. Suppose 𝒮j\mathcal{R}\in\mathscr{S}_{j} for some j0j\geq 0. Suppose q>nq_{\mathcal{R}}>n. Then j1j\geq 1. Furthermore, there exist an index i[0,j1]i\in[0,j-1], a prime pnp\mid n, and a residue class 𝒮i(n)\mathcal{R}^{\prime}\in\mathscr{S}_{i}\cap\mathscr{E}(n) of modulus qnk1q_{\mathcal{R}^{\prime}}\leq n^{k-1} with \mathcal{R}\subseteq\mathcal{R}^{\prime} and q/q=pq_{\mathcal{R}}/q_{\mathcal{R}^{\prime}}=p, such that vp(q)vp(r(n,𝐜))v_{p}(q_{\mathcal{R}})\leq v_{p}(r(n,\bm{c})) holds for all 𝐜𝒮1\bm{c}\in\mathcal{S}_{1}\cap\mathcal{R}^{\prime}.

Proof.

Each element of 𝒮0\mathscr{S}_{0} has modulus nn, so 𝒮0\mathcal{R}\notin\mathscr{S}_{0}; in particular, j0j\neq 0, so j1j\geq 1. Choose i[0,j1]i\in[0,j-1] maximal with 𝒮i\mathcal{R}\notin\mathscr{S}_{i}. Then by Definition 7.8, there exists a residue class 𝒮i(n)\mathcal{R}^{\prime}\in\mathscr{S}_{i}\cap\mathscr{E}(n), a tuple 𝒄𝒮1\bm{c}^{\prime}\in\mathcal{S}_{1}\cap\mathcal{R}^{\prime}, and a prime pnp\mid n, with qnk1q_{\mathcal{R}^{\prime}}\leq n^{k-1} and vp(r(n,𝒄))>vp(q)v_{p}(r(n,\bm{c}^{\prime}))>v_{p}(q_{\mathcal{R}^{\prime}}), such that \mathcal{R}\subseteq\mathcal{R}^{\prime} and q/q=pq_{\mathcal{R}}/q_{\mathcal{R}^{\prime}}=p. Now let 𝒄𝒮1\bm{c}\in\mathcal{S}_{1}\cap\mathcal{R}^{\prime}. By Proposition 7.6, vp(r(n,𝒄))>vp(q)v_{p}(r(n,\bm{c}))>v_{p}(q_{\mathcal{R}^{\prime}}) (since vp(r(n,𝒄))>vp(q)v_{p}(r(n,\bm{c}^{\prime}))>v_{p}(q_{\mathcal{R}^{\prime}})). So vp(r(n,𝒄))1+vp(q)=vp(q)v_{p}(r(n,\bm{c}))\geq 1+v_{p}(q_{\mathcal{R}^{\prime}})=v_{p}(q_{\mathcal{R}}). ∎

Proposition 7.10.

Let n,k1n,k\geq 1. Let 𝒫(n,k)\mathcal{R}\in\mathscr{P}(n,k). Then qnkq_{\mathcal{R}}\leq n^{k}. Furthermore, if n2n\geq 2 and 𝒫(n,k)(n)\mathcal{R}\in\mathscr{P}(n,k)\setminus\mathscr{E}(n), then 𝒮1\mathcal{R}\subseteq\mathcal{S}_{1}.

Proof.

By Lemma 7.9, we either have qnq_{\mathcal{R}}\leq n, or q=pqq_{\mathcal{R}}=pq_{\mathcal{R}^{\prime}} with pnp\mid n and qnk1q_{\mathcal{R}^{\prime}}\leq n^{k-1}. So qnkq_{\mathcal{R}}\leq n^{k}. Now suppose n2n\geq 2 and 𝒫(n,k)(n)\mathcal{R}\in\mathscr{P}(n,k)\setminus\mathscr{E}(n). Choose 𝒄𝒮1\bm{c}\in\mathcal{S}_{1}\cap\mathcal{R} (possible by (2.4)). Then r(n,𝒄)qr(n,\bm{c})\mid q_{\mathcal{R}}, by the definition of (n)\mathscr{E}(n). Now choose pnp\mid n (possible since n2n\geq 2). Then pl(p,𝒄)+1qp^{l(p,\bm{c})+1}\mid q_{\mathcal{R}}, by (7.22). But by the definition of l(p,𝒄)l(p,\bm{c}), we have Δ(𝒄)0\Delta(\bm{c}^{\prime})\neq 0 for all 𝒄𝒄modpl(p,𝒄)+1\bm{c}^{\prime}\equiv\bm{c}\bmod{p^{l(p,\bm{c})+1}}. Thus 𝒮1\mathcal{R}\subseteq\mathcal{S}_{1} (since 𝒄\bm{c}\in\mathcal{R}). ∎

Proposition 7.11.

Let n,k2n,k\geq 2. If (n,k)\mathcal{R}\in\mathscr{E}(n,k), then r(n,𝐜)nk2r(n,\bm{c})\geq n^{k-2} for all 𝐜𝒮1\bm{c}\in\mathcal{S}_{1}\cap\mathcal{R}.

Proof.

Suppose (n,k)\mathcal{R}\in\mathscr{E}(n,k). By Definition 7.8, (n,k)=𝒮j0(n)\mathscr{E}(n,k)=\mathscr{S}_{j_{0}}\cap\mathscr{E}(n). So 𝒮j0(n)\mathcal{R}\in\mathscr{S}_{j_{0}}\cap\mathscr{E}(n), whence q>nk1q_{\mathcal{R}}>n^{k-1} (or else we would have 𝒮j0+1𝒮j0\mathscr{S}_{j_{0}+1}\neq\mathscr{S}_{j_{0}} by the algorithm in Definition 7.8). By Lemma 7.9 (applied repeatedly), there exists a sequence of primes p1,,psnp_{1},\dots,p_{s}\mid n, with p1psqp_{1}\cdots p_{s}\mid q_{\mathcal{R}} and q/(p1ps1)>nq/(p1ps)q_{\mathcal{R}}/(p_{1}\cdots p_{s-1})>n\geq q_{\mathcal{R}}/(p_{1}\cdots p_{s}), such that

(7.23) vpi(q/(p1pi1))vpi(r(n,𝒄))v_{p_{i}}(q_{\mathcal{R}}/(p_{1}\cdots p_{i-1}))\leq v_{p_{i}}(r(n,\bm{c}))

holds for all i{1,2,,s}i\in\{1,2,\dots,s\} and 𝒄𝒮1\bm{c}\in\mathcal{S}_{1}\cap\mathcal{R}. If for each pp1psp\mid p_{1}\cdots p_{s}, we apply (7.23) with i=min{1us:pu=p}i=\min\{1\leq u\leq s:p_{u}=p\}, then we get vp(q)vp(r(n,𝒄))v_{p}(q_{\mathcal{R}})\leq v_{p}(r(n,\bm{c})). Since p1psqp_{1}\cdots p_{s}\mid q_{\mathcal{R}}, it follows that p1psr(n,𝒄)p_{1}\cdots p_{s}\mid r(n,\bm{c}), and thus r(n,𝒄)p1psq/n>nk2r(n,\bm{c})\geq p_{1}\cdots p_{s}\geq q_{\mathcal{R}}/n>n^{k-2}. ∎

Let μ()\colonequalsqm\mu(\mathcal{R})\colonequals q_{\mathcal{R}}^{-m} be the density of a residue class \mathcal{R} in m\mathbb{Z}^{m}.

Lemma 7.12 (KL’).

Let n2n\geq 2. Then limk(n,k)μ()=0\lim_{k\to\infty}\sum_{\mathcal{R}\in\mathscr{E}(n,k)}\mu(\mathcal{R})=0.

Proof.

Suppose (n,k)\mathcal{R}\in\mathscr{E}(n,k) and 𝒄𝒮1\bm{c}\in\mathcal{S}_{1}\cap\mathcal{R}, where k3k\geq 3. Then by Proposition 7.11, we have r(n,𝒄)nk2r(n,\bm{c})\geq n^{k-2}. But by (7.22), we have r(n,𝒄)pnpvp(n)+l(p,𝒄)=npnpl(p,𝒄)r(n,\bm{c})\mid\prod_{p\mid n}p^{v_{p}(n)+l(p,\bm{c})}=n\prod_{p\mid n}p^{l(p,\bm{c})}. Therefore (since n2n\geq 2), there exists a prime pnp\mid n with l(p,𝒄)k3l(p,\bm{c})\geq k-3. It follows that

(n,k)𝔼𝒄[Z,Z]m[𝟏𝒄𝒮1]pn𝔼𝒄[Z,Z]m[𝟏𝒄𝒮1𝟏l(p,𝒄)k3]\sum_{\mathcal{R}\in\mathscr{E}(n,k)}\mathbb{E}_{\bm{c}\in[-Z,Z]^{m}}[\bm{1}_{\bm{c}\in\mathcal{S}_{1}\cap\mathcal{R}}]\leq\sum_{p\mid n}\mathbb{E}_{\bm{c}\in[-Z,Z]^{m}}[\bm{1}_{\bm{c}\in\mathcal{S}_{1}}\bm{1}_{l(p,\bm{c})\geq k-3}]

for all reals Z1Z\geq 1. Taking ZZ\to\infty (using (2.4) on the left-hand side, and Lemma 4.4 on the right-hand side; cf. the proof of Proposition 6.1), we get

(7.24) (n,k)μ()pn𝔼𝒄pm[𝟏Δ(𝒄)0𝟏l(p,𝒄)k3].\sum_{\mathcal{R}\in\mathscr{E}(n,k)}\mu(\mathcal{R})\leq\sum_{p\mid n}\mathbb{E}_{\bm{c}\in\mathbb{Z}_{p}^{m}}[\bm{1}_{\Delta(\bm{c})\neq 0}\bm{1}_{l(p,\bm{c})\geq k-3}].

But by Lemma 4.4, the right-hand side of (7.24) tends to 0 as kk\to\infty. ∎

Lemma 7.13 (EKL’).

Assume Conjecture 1.11 (EKL). Let n2n\geq 2 and k3k\geq 3. Then (n,k)μ()H,ϵnkϵn(k3)/degH\sum_{\mathcal{R}\in\mathscr{E}(n,k)}\mu(\mathcal{R})\ll_{H,\epsilon}n^{k\epsilon}\cdot n^{-(k-3)/\deg{H}} (uniformly over nn and kk).

Proof.

Suppose (n,k)\mathcal{R}\in\mathscr{E}(n,k) and 𝒄𝒮1\bm{c}\in\mathcal{S}_{1}\cap\mathcal{R}. As in the proof of Lemma 7.12, we have r(n,𝒄)nk2r(n,\bm{c})\geq n^{k-2} and r(n,𝒄)npnpl(p,𝒄)r(n,\bm{c})\mid n\prod_{p\mid n}p^{l(p,\bm{c})}. By Remark 4.5, we have pnpl(p,𝒄)H(𝒄)\prod_{p\mid n}p^{l(p,\bm{c})}\mid H(\bm{c}). But nr(n,𝒄)nn\mid r(n,\bm{c})\mid n^{\infty} by (7.22). Hence H(𝒄)H(\bm{c}) is divisible by the integer q=r(n,𝒄)/nnk3q=r(n,\bm{c})/n\geq n^{k-3}, where qnq\mid n^{\infty}. Since every prime factor of qq is n\leq n, there exists uqu\mid q with nk3u<nk2n^{k-3}\leq u<n^{k-2}. On the other hand, if 𝒄𝒮0\bm{c}\in\mathcal{S}_{0}\cap\mathcal{R}, then H(𝒄)=0H(\bm{c})=0 (since ΔH\Delta\mid H), so nk3H(𝒄)n^{k-3}\mid H(\bm{c}). So for every 𝒄\bm{c}\in\mathcal{R} (if (n,k)\mathcal{R}\in\mathscr{E}(n,k)), there exists u[nk3,nk2)u\in[n^{k-3},n^{k-2}), with unu\mid n^{\infty}, such that uH(𝒄)u\mid H(\bm{c}). Thus

(n,k)𝔼𝒄[Z,Z]m[𝟏𝒄]nk3u<nk2:un𝔼𝒄[Z,Z]m[𝟏uH(𝒄)]\sum_{\mathcal{R}\in\mathscr{E}(n,k)}\mathbb{E}_{\bm{c}\in[-Z,Z]^{m}}[\bm{1}_{\bm{c}\in\mathcal{R}}]\leq\sum_{n^{k-3}\leq u<n^{k-2}:\,u\mid n^{\infty}}\mathbb{E}_{\bm{c}\in[-Z,Z]^{m}}[\bm{1}_{u\mid H(\bm{c})}]

for all reals Z1Z\geq 1. Taking ZZ\to\infty (using (4.3) on the right-hand side), we get

(n,k)μ()Hnk3u<nk2:unu1/degHunk3:unu1/degH.\sum_{\mathcal{R}\in\mathscr{E}(n,k)}\mu(\mathcal{R})\ll_{H}\sum_{n^{k-3}\leq u<n^{k-2}:\,u\mid n^{\infty}}u^{-1/\deg{H}}\leq\sum_{u\geq n^{k-3}:\,u\mid n^{\infty}}u^{-1/\deg{H}}.

But uN:unuβNϵβu1:unuϵϵNϵβnϵ\sum_{u\geq N:\,u\mid n^{\infty}}u^{-\beta}\leq N^{\epsilon-\beta}\sum_{u\geq 1:\,u\mid n^{\infty}}u^{-\epsilon}\ll_{\epsilon}N^{\epsilon-\beta}n^{\epsilon}, for all N,β,ϵ>0N,\beta,\epsilon\in\mathbb{R}_{>0} with ϵ<β\epsilon<\beta. ∎

7.4. From (RA1’E) to (RA1’E’)

We now build on Propositions 6.13 and 6.14. Let I>0I\subseteq\mathbb{R}_{>0} be a compact set. Let ν=ν𝒄(r)\nu=\nu_{\bm{c}}(r) be a smooth function m×,(𝒄,r)ν𝒄(r)\mathbb{R}^{m}\times\mathbb{R}\to\mathbb{C},\,(\bm{c},r)\mapsto\nu_{\bm{c}}(r), supported on [1,1]m×I[-1,1]^{m}\times I. Given (𝒂,n0)m×1(\bm{a},n_{0})\in\mathbb{Z}^{m}\times\mathbb{Z}_{\geq 1} and reals Z,N1Z,N\geq 1, let

(7.25) Σ11𝒂,n0(ν,Z,N)\colonequals𝒄𝒮1:𝒄𝒂modn0n1ν𝒄/Z(n/N)S𝒄(n)𝟏n𝒩𝒄\Sigma_{11}^{\bm{a},n_{0}}(\nu,Z,N)\colonequals\sum_{\bm{c}\in\mathcal{S}_{1}:\,\bm{c}\equiv\bm{a}\bmod{n_{0}}}\sum_{n\geq 1}\nu_{\bm{c}/Z}(n/N)S^{\natural}_{\bm{c}}(n)\bm{1}_{n\in\mathcal{N}^{\bm{c}}}
Conjecture 7.14 (RA1oo’E’).

Let M1M\in\mathbb{R}_{\geq 1}, and let n0M9/10n_{0}\leq M^{9/10} be a positive integer. Let Z,N2MZ,N\geq 2M be reals with NZ3N\leq Z^{3}. Then 1𝐚n0|Σ11𝐚,n0(ν,Z,N)|F,IZmN1/2(oF;M(1)+oF,M;Z(1))1,A7\sum_{1\leq\bm{a}\leq n_{0}}\lvert\Sigma_{11}^{\bm{a},n_{0}}(\nu,Z,N)\rvert\ll_{F,I}Z^{m}N^{1/2}\cdot(o_{F;M\to\infty}(1)+o_{F,M;Z\to\infty}(1))\cdot\mathcal{M}_{1,A_{7}} for some A7=A7(F)>0A_{7}=A_{7}(F)>0, where 1,k\mathcal{M}_{1,k} is defined as in (6.24).

The intermediate parameter MM here may seem strange, but it will ease our exposition.

Proposition 7.15.

Assume Conjectures 1.2, 6.9, and 6.12. Then Conjecture 7.14 holds.

Proof.

Plugging (7.20) into (7.25) reveals the equality

Σ11𝒂,n0(ν,Z,N)=𝒄𝒮1:𝒄𝒂modn0n1,d1ν𝒄/Z(n1d/N)a𝒄,1(n1)a𝒄(d).\Sigma_{11}^{\bm{a},n_{0}}(\nu,Z,N)=\sum_{\bm{c}\in\mathcal{S}_{1}:\,\bm{c}\equiv\bm{a}\bmod{n_{0}}}\sum_{n_{1},d\geq 1}\nu_{\bm{c}/Z}(n_{1}d/N)a_{\bm{c},1}(n_{1})a^{\prime}_{\bm{c}}(d).

Fix a function υ0Cc()\upsilon_{0}\in C^{\infty}_{c}(\mathbb{R}) with Suppυ0[1,1]\operatorname{Supp}{\upsilon_{0}}\subseteq[-1,1] and υ0|[1/2,1/2]=1\upsilon_{0}|_{[-1/2,1/2]}=1. Let L1L\geq 1 denote a real number to be chosen later. We first analyze the piece

Σ12𝒂,n0\colonequals𝒄𝒮1:𝒄𝒂modn0n1,d1(1υ0(d/L))ν𝒄/Z(n1d/N)a𝒄,1(n1)a𝒄(d)\Sigma_{12}^{\bm{a},n_{0}}\colonequals\sum_{\bm{c}\in\mathcal{S}_{1}:\,\bm{c}\equiv\bm{a}\bmod{n_{0}}}\sum_{n_{1},d\geq 1}(1-\upsilon_{0}(d/L))\nu_{\bm{c}/Z}(n_{1}d/N)a_{\bm{c},1}(n_{1})a^{\prime}_{\bm{c}}(d)

of Σ11𝒂,n0\Sigma_{11}^{\bm{a},n_{0}}. For later reference, note that (since Suppυ0[1,1]\operatorname{Supp}{\upsilon_{0}}\subseteq[-1,1])

(7.26) Σ11𝒂,n0Σ12𝒂,n0\displaystyle\Sigma_{11}^{\bm{a},n_{0}}-\Sigma_{12}^{\bm{a},n_{0}} =dLυ0(d/L)Σ13𝒂,n0(d),\displaystyle=\sum_{d\leq L}\upsilon_{0}(d/L)\Sigma_{13}^{\bm{a},n_{0}}(d),
(7.27) whereΣ13𝒂,n0(d)\displaystyle\textnormal{where}\quad\Sigma_{13}^{\bm{a},n_{0}}(d) \colonequals𝒄𝒮1:𝒄𝒂modn0a𝒄(d)n11ν𝒄/Z(n1d/N)a𝒄,1(n1).\displaystyle\colonequals\sum_{\bm{c}\in\mathcal{S}_{1}:\,\bm{c}\equiv\bm{a}\bmod{n_{0}}}a^{\prime}_{\bm{c}}(d)\sum_{n_{1}\geq 1}\nu_{\bm{c}/Z}(n_{1}d/N)a_{\bm{c},1}(n_{1}).

We bound Σ12𝒂,n0\Sigma_{12}^{\bm{a},n_{0}} using the Hölder technique behind (the simplest case, ϵ=1\epsilon=1, of) Proposition 7.5. Since a𝒄(d)a^{\prime}_{\bm{c}}(d) is the ddth coefficient of the Dirichlet series Φ𝒄,2(s)Φ𝒄,3(s)\Phi^{\bm{c},2}(s)\Phi^{\bm{c},3}(s) (see §7.3), we may write a𝒄a^{\prime}_{\bm{c}} in terms of a𝒄,2a_{\bm{c},2}, a𝒄,3a_{\bm{c},3}, and then apply Lemma 5.2 (with k=3k=3 and a(𝒏)=1j3a𝒄,j(nj)a(\bm{n})=\prod_{1\leq j\leq 3}a_{\bm{c},j}(n_{j}), and f(𝒓)=(1υ0(r2r3/L))ν𝒄/Z(r1r2r3/N)f(\bm{r})=(1-\upsilon_{0}(r_{2}r_{3}/L))\nu_{\bm{c}/Z}(r_{1}r_{2}r_{3}/N)), to get

(7.28) Σ12𝒂,n0=𝒄𝒮1:𝒄𝒂modn0(2π)3𝑵[1/2,)3d×𝑵𝒕3𝑑𝒕g𝒄,𝑵(i𝒕)1j3Σ12,𝑵𝒄,j(𝒕)\Sigma_{12}^{\bm{a},n_{0}}=\sum_{\bm{c}\in\mathcal{S}_{1}:\,\bm{c}\equiv\bm{a}\bmod{n_{0}}}(2\pi)^{-3}\int_{\bm{N}\in[1/2,\infty)^{3}}d^{\times}\bm{N}\int_{\bm{t}\in\mathbb{R}^{3}}d\bm{t}\,g_{\bm{c},\bm{N}}^{\vee}(i\bm{t})\prod_{1\leq j\leq 3}\Sigma_{12,\bm{N}}^{\bm{c},j}(\bm{t})

(cf. (7.11)), where g𝒄,𝑵(𝒓)\colonequals(1υ0(r2r3/L))ν𝒄/Z(r1r2r3/N)1j3ν2(rj/Nj)g_{\bm{c},\bm{N}}(\bm{r})\colonequals(1-\upsilon_{0}(r_{2}r_{3}/L))\nu_{\bm{c}/Z}(r_{1}r_{2}r_{3}/N)\prod_{1\leq j\leq 3}\nu_{2}(r_{j}/N_{j}) and

Σ12,𝑵𝒄,j(𝒕)\colonequals𝟏𝒄[Z,Z]mnj1ν2(nj/Nj)njitja𝒄,j(nj).\Sigma_{12,\bm{N}}^{\bm{c},j}(\bm{t})\colonequals\bm{1}_{\bm{c}\in[-Z,Z]^{m}}\sum_{n_{j}\geq 1}\nu_{2}(n_{j}/N_{j})n_{j}^{-it_{j}}a_{\bm{c},j}(n_{j}).

For every integer b0b\geq 0, Proposition 5.1 and (6.24) imply (uniformly over 𝒄\bm{c}, 𝑵\bm{N}, 𝒕\bm{t})

(7.29) g𝒄,𝑵(i𝒕)b1,b(ν)(1+𝒕)bg_{\bm{c},\bm{N}}^{\vee}(i\bm{t})\ll_{b}\mathcal{M}_{1,b}(\nu)(1+\lVert\bm{t}\rVert)^{-b}

(where the implied constant may depend on υ0\upsilon_{0}, ν2\nu_{2} as well as bb).

Since 1υ01-\upsilon_{0}, ν𝒄/Z\nu_{\bm{c}/Z}, ν2\nu_{2} are supported on [1/2,1/2]\mathbb{R}\setminus[-1/2,1/2], II, [1,2][1,2], respectively, we have g𝒄,𝑵(𝒓)=0g_{\bm{c},\bm{N}}(\bm{r})=0 unless r2r3L/2r_{2}r_{3}\geq L/2, r1r2r3NIr_{1}r_{2}r_{3}\in N\cdot I, and Njrj2NjN_{j}\leq r_{j}\leq 2N_{j} for all jj. Fix an integer A1A\geq 1 satisfying I[A1,A]I\subseteq[A^{-1},A]. Then g𝒄,𝑵=0g_{\bm{c},\bm{N}}=0 identically unless 𝑵\bm{N} lies in the set

(7.30) 12\colonequals{𝑵1/2:N2N3L/8,N1N2N3[N/8A,AN]}\mathscr{R}_{12}\colonequals\{\bm{N}\geq 1/2:N_{2}N_{3}\geq L/8,\;N_{1}N_{2}N_{3}\in[N/8A,AN]\}

(cf. the region 10\mathscr{R}_{10} from (7.14)). So (7.28) holds even if we restrict 𝑵\bm{N} to 12\mathscr{R}_{12}.

But if 𝑵12\bm{N}\in\mathscr{R}_{12} and 𝒕3\bm{t}\in\mathbb{R}^{3}, then (7.17) (with β=1\beta=1, (γ1,γ2,γ3)=(2,4,4)(\gamma_{1},\gamma_{2},\gamma_{3})=(2,4,4), t=0t=0) and the subsequent arguments up to (7.18) furnish (via Conjectures 1.2 and 6.9) the bound

(7.31) 𝑵12d×𝑵𝒄𝒮11j3|Σ12,𝑵𝒄,j(𝒕)|I𝑵12d×𝑵(1+𝒕)A6ZmN1/2N2η3(4)/2N34/30\int_{\bm{N}\in\mathscr{R}_{12}}d^{\times}\bm{N}\sum_{\bm{c}\in\mathcal{S}_{1}}\prod_{1\leq j\leq 3}\lvert\Sigma_{12,\bm{N}}^{\bm{c},j}(\bm{t})\rvert\ll_{I}\int_{\bm{N}\in\mathscr{R}_{12}}d^{\times}\bm{N}\,(1+\lVert\bm{t}\rVert)^{A_{6}}\frac{Z^{m}N^{1/2}}{N_{2}^{\eta_{3}(4)/2}N_{3}^{4/30}}

(where A6=(A3/2+1)β=A3/2+1A_{6}=(A_{3}/2+1)\beta=A_{3}/2+1, and η3(4)\eta_{3}(4) is as in Proposition 6.18). Upon taking absolute values in (7.28) (after restricting 𝑵\bm{N} to 12\mathscr{R}_{12}), summing over 1𝒂n01\leq\bm{a}\leq n_{0}, plugging in (7.29) (with b=A6+4b=\lceil A_{6}+4\rceil) and then (7.31), and integrating over 𝒕3\bm{t}\in\mathbb{R}^{3}, we conclude that

(7.32) 1𝒂n0|Σ12𝒂,n0|I1,b(ν)𝑵12d×𝑵ZmN1/2N2η3(4)/2N34/30I1,b(ν)ZmN1/2Lη4,\sum_{1\leq\bm{a}\leq n_{0}}\lvert\Sigma_{12}^{\bm{a},n_{0}}\rvert\ll_{I}\mathcal{M}_{1,b}(\nu)\int_{\bm{N}\in\mathscr{R}_{12}}d^{\times}\bm{N}\,\frac{Z^{m}N^{1/2}}{N_{2}^{\eta_{3}(4)/2}N_{3}^{4/30}}\ll_{I}\mathcal{M}_{1,b}(\nu)\frac{Z^{m}N^{1/2}}{L^{\eta_{4}}},

where b=A6+4b=\lceil A_{6}+4\rceil and η4=0.9min(η3(4)/2,4/30)\eta_{4}=0.9\min(\eta_{3}(4)/2,4/30).

We now turn to the sums Σ13𝒂,n0(d)\Sigma_{13}^{\bm{a},n_{0}}(d), for dLd\leq L. We first treat d=1d=1. By (7.27), we have Σ13𝒂,n0(1)=𝒄𝒮1:𝒄𝒂modn0n1ν𝒄/Z(n/N)a𝒄,1(n)\Sigma_{13}^{\bm{a},n_{0}}(1)=\sum_{\bm{c}\in\mathcal{S}_{1}:\,\bm{c}\equiv\bm{a}\bmod{n_{0}}}\sum_{n\geq 1}\nu_{\bm{c}/Z}(n/N)a_{\bm{c},1}(n). Therefore, by the triangle inequality, 1𝒂n0|Σ13𝒂,n0(1)|\sum_{1\leq\bm{a}\leq n_{0}}\lvert\Sigma_{13}^{\bm{a},n_{0}}(1)\rvert is at most the sum of the quantity (6.25) (with 𝒫={𝒂+n0m:1𝒂n0}\mathscr{P}=\{\bm{a}+n_{0}\mathbb{Z}^{m}:1\leq\bm{a}\leq n_{0}\}) and the left-hand side of (6.33) (with 𝒮=𝒮1\mathcal{S}=\mathcal{S}_{1}). So by Conjecture 6.12 (applicable since Z,N2M2n0=2Q(𝒫)Z,N\geq 2M\geq 2n_{0}=2Q(\mathscr{P})) and Lemma 6.15 (with θ=m\theta=m and Q(𝒫)=n0MQ(\mathscr{P})=n_{0}\leq M), the sum 1𝒂n0|Σ13𝒂,n0(1)|\sum_{1\leq\bm{a}\leq n_{0}}\lvert\Sigma_{13}^{\bm{a},n_{0}}(1)\rvert is at most OI,ϵ(𝔅0(N,ϵ))O_{I,\epsilon}(\mathfrak{B}_{0}(N,\epsilon)), where 𝔅0(N,ϵ)\mathfrak{B}_{0}(N,\epsilon) denotes the the expression666The replacement of n0n_{0} with the weaker (larger) MM here may seem strange, but it will be convenient later.

(7.33) ZmN1/2oM;Z(1)1,A4+ZmN1/3+ϵMϵ1,0.Z^{m}N^{1/2}o_{M;Z\to\infty}(1)\mathcal{M}_{1,A_{4}}+Z^{m}N^{1/3+\epsilon}M^{\epsilon}\mathcal{M}_{1,0}.

If ϵ\epsilon is sufficiently small, then (since MNM\leq N)

(7.34) 𝔅0(N,ϵ)ZmN1/2(oM;Z(1)+N1/6+2ϵ)1,A4.\mathfrak{B}_{0}(N,\epsilon)\leq Z^{m}N^{1/2}(o_{M;Z\to\infty}(1)+N^{-1/6+2\epsilon})\mathcal{M}_{1,A_{4}}.

Now suppose 2dL2\leq d\leq L. Let k=k(d)1k=k(d)\geq 1 denote an integer, with n0dk+1Mn_{0}d^{k+1}\leq M, to be chosen later. Using Definition 7.8, let

Σ13,0𝒂,n0(d)\colonequals(d,k)𝒄𝒮1(𝒂+n0m)a𝒄(d)n11ν𝒄/Z(n1d/N)a𝒄,1(n1),Σ13,1𝒂,n0(d)\colonequals𝒫(d,k)(d)𝒄𝒮1(𝒂+n0m)a𝒄(d)n11ν𝒄/Z(n1d/N)a𝒄,1(n1).\begin{split}\Sigma_{13,0}^{\bm{a},n_{0}}(d)&\colonequals\sum_{\mathcal{R}\in\mathscr{E}(d,k)}\sum_{\bm{c}\in\mathcal{S}_{1}\cap\mathcal{R}\cap(\bm{a}+n_{0}\mathbb{Z}^{m})}a^{\prime}_{\bm{c}}(d)\sum_{n_{1}\geq 1}\nu_{\bm{c}/Z}(n_{1}d/N)a_{\bm{c},1}(n_{1}),\\ \Sigma_{13,1}^{\bm{a},n_{0}}(d)&\colonequals\sum_{\mathcal{R}\in\mathscr{P}(d,k)\setminus\mathscr{E}(d)}\sum_{\bm{c}\in\mathcal{S}_{1}\cap\mathcal{R}\cap(\bm{a}+n_{0}\mathbb{Z}^{m})}a^{\prime}_{\bm{c}}(d)\sum_{n_{1}\geq 1}\nu_{\bm{c}/Z}(n_{1}d/N)a_{\bm{c},1}(n_{1}).\end{split}

Then Σ13𝒂,n0(d)=Σ13,0𝒂,n0(d)+Σ13,1𝒂,n0(d)\Sigma_{13}^{\bm{a},n_{0}}(d)=\Sigma_{13,0}^{\bm{a},n_{0}}(d)+\Sigma_{13,1}^{\bm{a},n_{0}}(d) by (7.27) (since 𝒫(d,k)\mathscr{P}(d,k) is a partition of m\mathbb{Z}^{m}). By Proposition 7.10, we have qdkq_{\mathcal{R}}\leq d^{k} for all 𝒫(d,k)\mathcal{R}\in\mathscr{P}(d,k), so Q(𝒫(d,k))dkQ(\mathscr{P}(d,k))\leq d^{k}.

By the triangle inequality,

(7.35) 1𝒂n0|Σ13,0𝒂,n0(d)|(d,k)𝒄𝒮1|a𝒄(d)||n11ν𝒄/Z(n1d/N)a𝒄,1(n1)|.\sum_{1\leq\bm{a}\leq n_{0}}\lvert\Sigma_{13,0}^{\bm{a},n_{0}}(d)\rvert\leq\sum_{\mathcal{R}\in\mathscr{E}(d,k)}\sum_{\bm{c}\in\mathcal{S}_{1}\cap\mathcal{R}}\lvert a^{\prime}_{\bm{c}}(d)\rvert\Bigl{\lvert}{\sum_{n_{1}\geq 1}\nu_{\bm{c}/Z}(n_{1}d/N)a_{\bm{c},1}(n_{1})}\Bigr{\rvert}.

By (7.21), a𝒄(d)ϵdϵa^{\prime}_{\bm{c}}(d)\ll_{\epsilon}d^{\epsilon}. However, by Lemma 5.2 (with k=1k=1 and a(n1)=a𝒄,1(n1)a(n_{1})=a_{\bm{c},1}(n_{1}), and f(r1)=ν𝒄/Z(r1d/N)f(r_{1})=\nu_{\bm{c}/Z}(r_{1}d/N)) and Proposition 5.1, we have (cf. (7.28), (7.29), and (7.30))

n11ν𝒄/Z(n1d/N)a𝒄,1(n1)b1/2AN/dd×Nt1𝑑t11,b(ν)|Σ12,(N1,1,1)𝒄,1(t1,0,0)|(1+|t1|)b\sum_{n_{1}\geq 1}\nu_{\bm{c}/Z}(n_{1}d/N)a_{\bm{c},1}(n_{1})\ll_{b}\int_{1/2}^{AN/d}d^{\times}{N}\int_{t_{1}\in\mathbb{R}}dt_{1}\,\mathcal{M}_{1,b}(\nu)\frac{\lvert\Sigma_{12,(N_{1},1,1)}^{\bm{c},1}(t_{1},0,0)\rvert}{(1+\lvert t_{1}\rvert)^{b}}

for all integers b0b\geq 0. Plugging this and a𝒄(d)ϵdϵa^{\prime}_{\bm{c}}(d)\ll_{\epsilon}d^{\epsilon} into (7.35), and then applying Cauchy–Schwarz over 𝒄(d,k)([Z,Z]m)\bm{c}\in\bigcup_{\mathcal{R}\in\mathscr{E}(d,k)}(\mathcal{R}\cap[-Z,Z]^{m}), we get (by Lemma 7.12 and Conjecture 6.9)

(7.36) 1𝒂n0|Σ13,0𝒂,n0(d)|I,ϵdϵ1,A3/2+2(ν)od;k(Zm)1/2(ZmN/d)1/2;\sum_{1\leq\bm{a}\leq n_{0}}\lvert\Sigma_{13,0}^{\bm{a},n_{0}}(d)\rvert\ll_{I,\epsilon}d^{\epsilon}\mathcal{M}_{1,\lceil A_{3}/2+2\rceil}(\nu)\cdot o_{d;k\to\infty}(Z^{m})^{1/2}\cdot(Z^{m}N/d)^{1/2};

cf. the numerics in (7.31) and (7.32). (Before applying Lemma 7.12, note that Q((d,k))dkMZQ(\mathscr{E}(d,k))\leq d^{k}\leq M\leq Z, so |(d,k)([Z,Z]m)|Zm(d,k)μ()\lvert\bigcup_{\mathcal{R}\in\mathscr{E}(d,k)}(\mathcal{R}\cap[-Z,Z]^{m})\rvert\ll Z^{m}\sum_{\mathcal{R}\in\mathscr{E}(d,k)}\mu(\mathcal{R}).)

By Definition 7.8, the quantity a𝒄(d)a^{\prime}_{\bm{c}}(d) is constant over 𝒮1\mathcal{S}_{1}\cap\mathcal{R} for each 𝒫(d,k)(d)\mathcal{R}\in\mathscr{P}(d,k)\setminus\mathscr{E}(d). But a𝒄(d)ϵdϵa^{\prime}_{\bm{c}}(d)\ll_{\epsilon}d^{\epsilon} by (7.21), so we get

Σ13,1𝒂,n0(d)ϵdϵ𝒫(d,k)(d)|𝒄𝒮1(𝒂+n0m)n11ν𝒄/Z(n1d/N)a𝒄,1(n1)|.\Sigma_{13,1}^{\bm{a},n_{0}}(d)\ll_{\epsilon}d^{\epsilon}\sum_{\mathcal{R}\in\mathscr{P}(d,k)\setminus\mathscr{E}(d)}\,\Bigl{\lvert}{\sum_{\bm{c}\in\mathcal{S}_{1}\cap\mathcal{R}\cap(\bm{a}+n_{0}\mathbb{Z}^{m})}\sum_{n_{1}\geq 1}\nu_{\bm{c}/Z}(n_{1}d/N)a_{\bm{c},1}(n_{1})}\Bigr{\rvert}.

We may apply Conjecture 6.12 (with N/dN/d in place of NN) and Lemma 6.15 (with 𝒮=𝒮1\mathcal{S}=\mathcal{S}_{1} and θ=m\theta=m) with 𝒫={(𝒂+n0m):𝒫(d,k), 1𝒂n0}\mathscr{P}=\{\mathcal{R}\cap(\bm{a}+n_{0}\mathbb{Z}^{m}):\mathcal{R}\in\mathscr{P}(d,k),\;1\leq\bm{a}\leq n_{0}\}, since Q(𝒫)n0dkM/dQ(\mathscr{P})\leq n_{0}d^{k}\leq M/d and Z,N2MZ,N\geq 2M (so that 2M,Z,N/d2Q(𝒫)2M,Z,N/d\geq 2Q(\mathscr{P}) and N/dNZ3N/d\leq N\leq Z^{3}). By the triangle inequality applied to the right-hand side of the previous display, we then obtain the bound

(7.37) 1𝒂n0|Σ13,1𝒂,n0(d)|I,ϵdϵ𝔅0(N/d,ϵ)(where 𝔅0(N,ϵ) denotes (7.33) as before).\sum_{1\leq\bm{a}\leq n_{0}}\lvert\Sigma_{13,1}^{\bm{a},n_{0}}(d)\rvert\ll_{I,\epsilon}d^{\epsilon}\mathfrak{B}_{0}(N/d,\epsilon)\quad\textnormal{(where $\mathfrak{B}_{0}(N,\epsilon)$ denotes \eqref{EXPR:frakB_0-soft-bound-for-Sigma_13(1)-total} as before)}.

Let K1K\geq 1 denote an integer to be chosen soon. Assembling (7.26), (7.32), and our work on Σ13𝒂,n0(d)\Sigma_{13}^{\bm{a},n_{0}}(d) for dLd\leq L (see (7.33) for d=1d=1, and (7.36), (7.37) for 2dL2\leq d\leq L), we get (by the triangle inequality) that the sum 1𝒂n0|Σ11𝒂,n0|\sum_{1\leq\bm{a}\leq n_{0}}\lvert\Sigma_{11}^{\bm{a},n_{0}}\rvert is at most

(7.38) I,ϵ1,A3/2+5(ν)(ZmN1/2Lη4+oL;K(ZmN1/2))+dLdϵ𝔅0(N/d,ϵ),\ll_{I,\epsilon}\mathcal{M}_{1,\lceil A_{3}/2+5\rceil}(\nu)(Z^{m}N^{1/2}L^{-\eta_{4}}+o_{L;K\to\infty}(Z^{m}N^{1/2}))+\sum_{d\leq L}d^{\epsilon}\mathfrak{B}_{0}(N/d,\epsilon),

provided k(d)Kk(d)\geq K and n0dk(d)+1Mn_{0}d^{k(d)+1}\leq M hold for all integers dd with 2dL2\leq d\leq L. But upon replacing NN in (7.33) with N/dN/d, and summing over dLd\leq L, we find that

(7.39) dLdϵ𝔅0(N/d,ϵ)L2/3𝔅0(N,ϵ).\sum_{d\leq L}d^{\epsilon}\mathfrak{B}_{0}(N/d,\epsilon)\ll L^{2/3}\mathfrak{B}_{0}(N,\epsilon).

It remains to carefully specify parameters. Choose K=K(L)1K=K(L)\geq 1 so that oL;K(ZmN1/2)Lη4ZmN1/2o_{L;K\to\infty}(Z^{m}N^{1/2})\leq L^{-\eta_{4}}Z^{m}N^{1/2}. Then let k(d)=K(L)k(d)=K(L) for all integers dd with 2dL2\leq d\leq L. Let L=L(M)1L=L(M)\geq 1 (in terms of MM) be the largest integer for which LK(L)+1M1/10L^{K(L)+1}\leq M^{1/10}; such an integer exists, because 1K(1)+1M1/101^{K(1)+1}\leq M^{1/10} and LK(L)+1L2L^{K(L)+1}\geq L^{2}. Crucially, we have limML(M)=\lim_{M\to\infty}{L(M)}=\infty, since the expression LK(L)+1L^{K(L)+1} is bounded on any finite set of integers LL.

Since n0M9/10n_{0}\leq M^{9/10}, we have n0LK(L)+1Mn_{0}L^{K(L)+1}\leq M. So the bound (7.38) on 1𝒂n0|Σ11𝒂,n0|\sum_{1\leq\bm{a}\leq n_{0}}\lvert\Sigma_{11}^{\bm{a},n_{0}}\rvert is valid, and therefore (by (7.39)) we get (letting A7=max(A4,A3/2+5)A_{7}=\max(A_{4},\lceil A_{3}/2+5\rceil))

1𝒂n0|Σ11𝒂,n0|I,ϵ1,A7(ν)ZmN1/2L(M)η4+L(M)2/3𝔅0(N,ϵ).\sum_{1\leq\bm{a}\leq n_{0}}\lvert\Sigma_{11}^{\bm{a},n_{0}}\rvert\ll_{I,\epsilon}\mathcal{M}_{1,A_{7}}(\nu)Z^{m}N^{1/2}L(M)^{-\eta_{4}}+L(M)^{2/3}\mathfrak{B}_{0}(N,\epsilon).

But L(M)2L(M)K(L(M))+1M1/10L(M)^{2}\leq L(M)^{K(L(M))+1}\leq M^{1/10}, and NMN\geq M, so by (7.34), we have

L(M)2/3𝔅0(N,ϵ)ZmN1/2(M1/301/6+2ϵ+oM;Z(1))1,A7,L(M)^{2/3}\mathfrak{B}_{0}(N,\epsilon)\ll Z^{m}N^{1/2}(M^{1/30-1/6+2\epsilon}+o_{M;Z\to\infty}(1))\mathcal{M}_{1,A_{7}},

provided ϵ\epsilon is sufficiently small. Both L(M)η4L(M)^{-\eta_{4}} and M1/301/6+2ϵM^{1/30-1/6+2\epsilon} are oM(1)o_{M\to\infty}(1). ∎

Proposition 7.16 (RA1δ\delta’E’).

Assume Conjectures 1.2, 1.10, and 1.11. Suppose Z,N2MZ,N\geq 2M and NZ3N\leq Z^{3}. Suppose 1MZη21\leq M\leq Z^{\eta_{2}} (with η2\eta_{2} as in Proposition 6.14), and let n0M1/2n_{0}\leq M^{1/2} be a positive integer. Then 1𝐚n0|Σ11𝐚,n0(ν,Z,N)|F,I,ϵZm+ϵN1/2(M1/6degHN1/6+Mη5/4degH)1,A8\sum_{1\leq\bm{a}\leq n_{0}}\lvert\Sigma_{11}^{\bm{a},n_{0}}(\nu,Z,N)\rvert\ll_{F,I,\epsilon}Z^{m+\epsilon}N^{1/2}\cdot(M^{1/6\deg{H}}N^{-1/6}+M^{-\eta_{5}/4\deg{H}})\cdot\mathcal{M}_{1,A_{8}}. Here η5\eta_{5}, A8A_{8} are positive reals depending only on FF.

Proof.

We adjust the proof of Proposition 7.15. Let L[1,M1/8]L\in[1,M^{1/8}] be a real to be chosen later. For each integer dd with 2dL2\leq d\leq L, let k(d)k(d) be the largest integer kk for which dk+1M1/2d^{k+1}\leq M^{1/2}; then k(d)3k(d)\geq 3 (since dLM1/8d\leq L\leq M^{1/8}), and dk(d)+2>M1/2d^{k(d)+2}>M^{1/2} (by the maximality of k(d)k(d)). Let

𝔅1(N,ϵ)\colonequalsZmη2N1/21,A5+ZmN1/3+ϵMϵ1,0.\mathfrak{B}_{1}(N,\epsilon)\colonequals Z^{m-\eta_{2}}N^{1/2}\mathcal{M}_{1,A_{5}}+Z^{m}N^{1/3+\epsilon}M^{\epsilon}\mathcal{M}_{1,0}.

Using Proposition 6.14 in place of Conjecture 6.12, we find that 1𝒂n0|Σ13𝒂,n0(1)|I,ϵ𝔅1(N,ϵ)\sum_{1\leq\bm{a}\leq n_{0}}\lvert\Sigma_{13}^{\bm{a},n_{0}}(1)\rvert\ll_{I,\epsilon}\mathfrak{B}_{1}(N,\epsilon) and (if 2dL2\leq d\leq L, then) 1𝒂n0|Σ13,1𝒂,n0(d)|I,ϵdϵ𝔅1(N/d,ϵ)\sum_{1\leq\bm{a}\leq n_{0}}\lvert\Sigma_{13,1}^{\bm{a},n_{0}}(d)\rvert\ll_{I,\epsilon}d^{\epsilon}\mathfrak{B}_{1}(N/d,\epsilon); cf. (7.33) and (7.37). Summing over 1dL1\leq d\leq L, and writing Σ13,1𝒂,n0(1)\colonequalsΣ13𝒂,n0(1)\Sigma_{13,1}^{\bm{a},n_{0}}(1)\colonequals\Sigma_{13}^{\bm{a},n_{0}}(1) for convenience, we obtain

(7.40) 1dL1𝒂n0|Σ13,1𝒂,n0(d)|I,ϵL2/3𝔅1(N,ϵ)L2/3Zm+4ϵN1/2(Zη2+N1/6)1,A5.\sum_{1\leq d\leq L}\sum_{1\leq\bm{a}\leq n_{0}}\lvert\Sigma_{13,1}^{\bm{a},n_{0}}(d)\rvert\ll_{I,\epsilon}L^{2/3}\mathfrak{B}_{1}(N,\epsilon)\leq L^{2/3}Z^{m+4\epsilon}N^{1/2}(Z^{-\eta_{2}}+N^{-1/6})\mathcal{M}_{1,A_{5}}.

On the other hand, if we plug GRH (Proposition 3.2(8)) into (7.35) (after partial summation over n1n_{1}, say, and recalling the definition of 1,1\mathcal{M}_{1,1} from (6.24)) and then use Lemma 7.13 (applicable since k(d)3k(d)\geq 3 and we assume Conjecture 1.11), then (if 2dL2\leq d\leq L) we get

(7.41) 1𝒂n0|Σ13,0𝒂,n0(d)|I,ϵdϵ1,1(ν)Zϵ(N/d)1/2+ϵZmdk(d)ϵd(k(d)3)/degH;\sum_{1\leq\bm{a}\leq n_{0}}\lvert\Sigma_{13,0}^{\bm{a},n_{0}}(d)\rvert\ll_{I,\epsilon}d^{\epsilon}\mathcal{M}_{1,1}(\nu)Z^{\epsilon}(N/d)^{1/2+\epsilon}\cdot Z^{m}d^{k(d)\epsilon}d^{-(k(d)-3)/\deg{H}};

cf. the use of Lemma 7.12 towards (7.36). On the right-hand side, dϵZϵ(N/d)ϵdk(d)ϵZ5ϵd^{\epsilon}Z^{\epsilon}(N/d)^{\epsilon}d^{k(d)\epsilon}\leq Z^{5\epsilon} (since NZ3N\leq Z^{3} and dk(d)Md^{k(d)}\leq M), and d(k(d)3)/degH>M1/2degHd5/degHM1/2degHd5/12d^{(k(d)-3)/\deg{H}}>M^{1/2\deg{H}}d^{-5/\deg{H}}\geq M^{1/2\deg{H}}d^{-5/12} (since dk(d)+2>M1/2d^{k(d)+2}>M^{1/2} and degH12\deg{H}\geq 12). Thus, summing (7.41) over 2dL2\leq d\leq L gives

(7.42) 2dL1𝒂n0|Σ13,0𝒂,n0(d)|I,ϵ1,1(ν)Zm+5ϵN1/2L11/12M1/2degH.\sum_{2\leq d\leq L}\sum_{1\leq\bm{a}\leq n_{0}}\lvert\Sigma_{13,0}^{\bm{a},n_{0}}(d)\rvert\ll_{I,\epsilon}\mathcal{M}_{1,1}(\nu)Z^{m+5\epsilon}N^{1/2}L^{11/12}\cdot M^{-1/2\deg{H}}.

As for Σ12𝒂,n0\Sigma_{12}^{\bm{a},n_{0}}, the bounds (7.31) and (7.32) must be adjusted slightly, since we do not assume Conjecture 6.9. However, by Proposition 3.2(8) and partial summation, the bound (6.23) in Conjecture 6.9 still holds up to a factor of ZϵZ^{\epsilon} (for any A3>0A_{3}>0). Therefore, (7.31) and (7.32) still hold if we replace ZmN1/2Z^{m}N^{1/2} with Zm+ϵN1/2Z^{m+\epsilon}N^{1/2}; so 1𝒂n0|Σ12𝒂,n0|I1,A3/2+5(ν)ZmN1/2/Lη4\sum_{1\leq\bm{a}\leq n_{0}}\lvert\Sigma_{12}^{\bm{a},n_{0}}\rvert\ll_{I}\mathcal{M}_{1,\lceil A_{3}/2+5\rceil}(\nu)Z^{m}N^{1/2}/L^{\eta_{4}}.

Let A8=max(A5,A3/2+5)A_{8}=\max(A_{5},\lceil A_{3}/2+5\rceil). Assembling (7.26), (7.40), (7.42), and our work on Σ12𝒂,n0\Sigma_{12}^{\bm{a},n_{0}}, we get (by the triangle inequality) that the sum 1𝒂n0|Σ11𝒂,n0|\sum_{1\leq\bm{a}\leq n_{0}}\lvert\Sigma_{11}^{\bm{a},n_{0}}\rvert is

I,ϵ1,A8(ν)Zm+5ϵN1/2(L2/3(Zη2+N1/6)+L11/12M1/2degH+Lη4).\ll_{I,\epsilon}\mathcal{M}_{1,A_{8}}(\nu)Z^{m+5\epsilon}N^{1/2}(L^{2/3}(Z^{-\eta_{2}}+N^{-1/6})+L^{11/12}M^{-1/2\deg{H}}+L^{-\eta_{4}}).

Let L=M1/4degHL=M^{1/4\deg{H}} and η5=min(1,η4)\eta_{5}=\min(1,\eta_{4}) to finish. ∎

8. New bounds on the integral factor

Recall J𝒄,X(n)J_{\bm{c},X}(n) from (2.9). As we explained in §2, we need to go beyond the integral estimates from standard sources like [duke1993bounds, heath1996new, heath1998circle, hooley2014octonary] (and the related estimates of [hooley1986HasseWeil, hooley_greaves_harman_huxley_1997]), such as (2.19). We will prove uniform bounds free of epsilons and logs, while also bringing discriminants into the picture via the following consequence of (2.3):

(8.1) Δ(F(𝒙))F|F(𝒙)|𝒙2deg(Δ)deg(F)for 𝒙m.\Delta(\nabla{F}(\bm{x}))\ll_{F}\lvert F(\bm{x})\rvert\cdot\lVert\bm{x}\rVert^{2\deg(\Delta)-\deg(F)}\quad\textnormal{for $\bm{x}\in\mathbb{R}^{m}$}.

To give clean, general bounds, we assume (2.6). We prove the following on J𝒄,X(n)J_{\bm{c},X}(n):

Proposition 8.1.

Assume (2.6). Let X,Z,n>0X,Z,n\in\mathbb{R}_{>0}. Let 𝐜m\bm{c}\in\mathbb{R}^{m} with 𝐜Z\lVert\bm{c}\rVert\leq Z. Then

lognj𝒄𝜶J𝒄,X(n)F,w,j,𝜶,b(X/n)|𝜶|(1+X𝒄/n)1m/2(1+𝒄/X1/2)b(1+XΔ(𝒄/Z)𝒄/n)b\partial_{\log{n}}^{j}\partial_{\bm{c}}^{\bm{\alpha}}J_{\bm{c},X}(n)\ll_{F,w,j,\bm{\alpha},b}\frac{(X/n)^{\lvert\bm{\alpha}\rvert}(1+X\lVert\bm{c}\rVert/n)^{1-m/2}}{(1+\lVert\bm{c}\rVert/X^{1/2})^{b}(1+X\lVert\Delta(\bm{c}/Z)\bm{c}\rVert/n)^{b}}

for all integers j,b0j,b\geq 0 and multi-indices 𝛂0\bm{\alpha}\geq 0.

The fact that increasing jj is harmless can be interpreted as an instance of “homogeneous dimensional analysis” (and ultimately arises from the homogeneity of FF, via a beautiful recursive structure due to [heath1996new]; see (8.5) below). The factor Δ(𝒄/Z)1\Delta(\bm{c}/Z)\ll 1 measures the “degeneracy” of the real hyperplane section F(𝒙)=𝒄𝒙=0F(\bm{x})=\bm{c}\cdot\bm{x}=0, and it arises in our proof for roughly the same reason that dual hypersurfaces arise in [huang2020density]*(5.4), (5.11)–(5.14).

Morally, in (2.10), Proposition 8.1 lets us “imagine that there are sharp cutoffs” 𝒄X1/2\lVert\bm{c}\rVert\ll X^{1/2} and nXΔ(𝒄/X1/2)𝒄n\gg X\lVert\Delta(\bm{c}/X^{1/2})\bm{c}\rVert. Since by (4.4) we typically have |Δ(𝒄/X1/2)|1\lvert\Delta(\bm{c}/X^{1/2})\rvert\asymp 1, one might thus expect (in view of Proposition 2.4, and our ϵ\epsilon-diagnosis at the end of §2) that nX3/2n\asymp X^{3/2} should be the “dominant range” on average, and there we have J𝒄,X(n)1J_{\bm{c},X}(n)\ll 1.

To prove Proposition 8.1, we must first dig into some technical aspects of [duke1993bounds, heath1996new]’s hh-function. Let ωHB,0(x)\colonequals𝟏|x|1exp((1x2)1)Cc()\omega_{\textnormal{HB},0}(x)\colonequals\bm{1}_{\lvert x\rvert\leq 1}\cdot\exp(-(1-x^{2})^{-1})\in C^{\infty}_{c}(\mathbb{R}) and cHB,0\colonequalsx𝑑xωHB,0(x)c_{\textnormal{HB},0}\colonequals\int_{x\in\mathbb{R}}dx\,\omega_{\textnormal{HB},0}(x); note that ωHB,0\omega_{\textnormal{HB},0} is supported on [1,1][-1,1]. Following [heath1996new]*p. 165, let

ωHB(x)\colonequals4cHB,01ωHB,0(4x3)Cc(),\omega_{\textnormal{HB}}(x)\colonequals 4c_{\textnormal{HB},0}^{-1}\cdot\omega_{\textnormal{HB},0}(4x-3)\in C^{\infty}_{c}(\mathbb{R}),

so that ωHB\omega_{\textnormal{HB}} is supported on [1/2,1][1/2,1]. For x>0x>0 and yy\in\mathbb{R}, let

h(x,y)\colonequalsj1(xj)1[ωHB(xj)ωHB(|y|/(xj))].h(x,y)\colonequals\sum_{j\geq 1}(xj)^{-1}[\omega_{\textnormal{HB}}(xj)-\omega_{\textnormal{HB}}(\lvert y\rvert/(xj))].

By (2.7) and (2.9), and a change of variables from 𝒙\bm{x} to 𝒙~=𝒙/X\tilde{\bm{x}}=\bm{x}/X, we have (since Y2=X3Y^{2}=X^{3})

(8.2) J𝒄,X(n)=𝒙~m𝑑𝒙~w(𝒙~)h(n/Y,F(𝒙~))e(X𝒄𝒙~/n).J_{\bm{c},X}(n)=\int_{\tilde{\bm{x}}\in\mathbb{R}^{m}}d\tilde{\bm{x}}\,w(\tilde{\bm{x}})h(n/Y,F(\tilde{\bm{x}}))e(-X\bm{c}\cdot\tilde{\bm{x}}/n).

The following shows that we may take A0=sup𝒙Suppwmax(1,2|F(𝒙)|)A_{0}=\sup_{\bm{x}\in\operatorname{Supp}{w}}\max(1,2\lvert F(\bm{x})\rvert) in Proposition 2.4.

Lemma 8.2 (See [heath1996new]*Lemma 4).

If ξ\xi\in\mathbb{R} and rmax(1,2|ξ|)r\geq\max(1,2\lvert\xi\rvert), then h(r,ξ)=0h(r,\xi)=0.

Following [heath1996new], we now build a Fourier transform. Fix υ1Cc()\upsilon_{1}\in C^{\infty}_{c}(\mathbb{R}) such that υ1(F(𝒙))1\upsilon_{1}(F(\bm{x}))\geq 1 for all 𝒙Suppw\bm{x}\in\operatorname{Supp}{w}. Let w0(𝒙)\colonequalsw(𝒙)/υ1(F(𝒙))Cc()w_{0}(\bm{x})\colonequals w(\bm{x})/\upsilon_{1}(F(\bm{x}))\in C^{\infty}_{c}(\mathbb{R}). For any r>0r>0, let

pr(u)\colonequalsξ𝑑ξυ1(ξ)h(r,ξ)e(uξ)p_{r}(u)\colonequals\int_{\xi\in\mathbb{R}}d\xi\,\upsilon_{1}(\xi)h(r,\xi)e(-u\xi)

be the Fourier transform of υ1(ξ)h(r,ξ)Cc()\upsilon_{1}(\xi)h(r,\xi)\in C^{\infty}_{c}(\mathbb{R}). Writing w(𝒙)=w0(𝒙)υ1(F(𝒙))w(\bm{x})=w_{0}(\bm{x})\upsilon_{1}(F(\bm{x})) and υ1(ξ)h(r,ξ)=u𝑑upr(u)e(uξ)\upsilon_{1}(\xi)h(r,\xi)=\int_{u\in\mathbb{R}}du\,p_{r}(u)e(u\xi) (for r=n/Yr=n/Y and ξ=F(𝒙)\xi=F(\bm{x})) in (8.2), we get

(8.3) J𝒄,X(n)=u𝑑upr(u)𝒙~m𝑑𝒙~w0(𝒙~)e(uF(𝒙~)𝒗𝒙~),J_{\bm{c},X}(n)=\int_{u\in\mathbb{R}}du\,p_{r}(u)\int_{\tilde{\bm{x}}\in\mathbb{R}^{m}}d\tilde{\bm{x}}\,w_{0}(\tilde{\bm{x}})e(uF(\tilde{\bm{x}})-\bm{v}\cdot\tilde{\bm{x}}),

where (r,𝒗)=(n/Y,X𝒄/n)(r,\bm{v})=(n/Y,X\bm{c}/n); cf. [heath1996new]*Lemma 17.

Let 𝔮r(t)\colonequalspr(t/r)\mathfrak{q}_{r}(t)\colonequals p_{r}(t/r), so that pr(u)=𝔮r(ru)p_{r}(u)=\mathfrak{q}_{r}(ru). It turns out (see Lemma 8.3) that 𝔮r\mathfrak{q}_{r} behaves somewhat like a “fixed” Schwartz function independent of rr. Thus uu may be compared with βY2\beta Y^{2} in the classical Dirichlet arc theory, where |β|1/(nY)=1/(rY2)\lvert\beta\rvert\leq 1/(nY)=1/(rY^{2}).

Lemma 8.3.

Let r>0r>0 and tt\in\mathbb{R}. Then logrj𝔮r(l)(t)l,j,k(1+|t|)k\partial_{\log{r}}^{j}\mathfrak{q}_{r}^{(l)}(t)\ll_{l,j,k}(1+\lvert t\rvert)^{-k} for all j,k,l0j,k,l\in\mathbb{Z}_{\geq 0}.

Proof.

We may assume k=0k=0 if |t|1\lvert t\rvert\leq 1; this lets us treat all tt simultaneously.

Since 𝔮r(t)=pr(t/r)=ξ𝑑ξυ1(ξ)h(r,ξ)e(tξ/r)\mathfrak{q}_{r}(t)=p_{r}(t/r)=\int_{\xi\in\mathbb{R}}d\xi\,\upsilon_{1}(\xi)h(r,\xi)e(-t\xi/r), we have

𝔮r(l)(t)=ξ𝑑ξυ1(ξ)h(r,ξ)(2πiξ/r)le(tξ/r).\mathfrak{q}_{r}^{(l)}(t)=\int_{\xi\in\mathbb{R}}d\xi\,\upsilon_{1}(\xi)h(r,\xi)\cdot(-2\pi i\xi/r)^{l}e(-t\xi/r).

Since logr((ξ/r)l)=l(ξ/r)l\partial_{\log{r}}((\xi/r)^{l})=-l(\xi/r)^{l} and logr(e(tξ/r))=(2πitξ/r)e(tξ/r)\partial_{\log{r}}(e(-t\xi/r))=(2\pi it\xi/r)e(-t\xi/r), we find by induction (and the Leibniz rule) that for some constants ca,bl,jc^{l,j}_{a,b}\in\mathbb{C}, we have

logrj𝔮r(l)(t)=a,b0:a+bjca,bl,jξ𝑑ξυ1(ξ)(lograh(r,ξ))(ξ/r)l(tξ/r)be(tξ/r).\partial_{\log{r}}^{j}\mathfrak{q}_{r}^{(l)}(t)=\sum_{a,b\geq 0:\,a+b\leq j}c^{l,j}_{a,b}\int_{\xi\in\mathbb{R}}d\xi\,\upsilon_{1}(\xi)\cdot(\partial_{\log{r}}^{a}h(r,\xi))\cdot(\xi/r)^{l}(t\xi/r)^{b}e(-t\xi/r).

Integrating by parts kk times in ξ\xi (repeatedly integrating e(tξ/r)e(-t\xi/r) and differentiating the complementary factor), and then taking absolute values, we get

logrj𝔮r(l)(t)l,j,k,υ1a,b,α,β0:a+bj,α+βk,βl+bξSuppυ1𝑑ξ(ξαlograh(r,ξ))|ξ/r|l|tξ/r|b|ξ|β1|t/r|k.\partial_{\log{r}}^{j}\mathfrak{q}_{r}^{(l)}(t)\ll_{l,j,k,\upsilon_{1}}\sum_{\begin{subarray}{c}a,b,\alpha,\beta\geq 0:\\ a+b\leq j,\;\alpha+\beta\leq k,\;\beta\leq l+b\end{subarray}}\int_{\xi\in\operatorname{Supp}{\upsilon_{1}}}d\xi\,(\partial_{\xi}^{\alpha}\partial_{\log{r}}^{a}h(r,\xi))\cdot\frac{\lvert\xi/r\rvert^{l}\lvert t\xi/r\rvert^{b}}{\lvert\xi\rvert^{\beta}}\frac{1}{\lvert t/r\rvert^{k}}.

By Lemma 8.2, we may assume rυ11r\ll_{\upsilon_{1}}1. Then rcξαh(r,ξ)c,α,Ar1cαmin(1,(r/|ξ|)A)\partial_{r}^{c}\partial_{\xi}^{\alpha}h(r,\xi)\ll_{c,\alpha,A}r^{-1-c-\alpha}\min(1,(r/\lvert\xi\rvert)^{A}) for c,α,A0c,\alpha,A\geq 0, by [heath1996new]*Lemma 5. Thus lograξαh(r,ξ)a0carc|rcξαh(r,ξ)|a,α,Ar1αmin(1,(r/|ξ|)A)\partial_{\log{r}}^{a}\partial_{\xi}^{\alpha}h(r,\xi)\ll_{a}\sum_{0\leq c\leq a}r^{c}\cdot\lvert\partial_{r}^{c}\partial_{\xi}^{\alpha}h(r,\xi)\rvert\ll_{a,\alpha,A}r^{-1-\alpha}\min(1,(r/\lvert\xi\rvert)^{A}), which when inserted in the previous display gives

logrj𝔮r(l)(t)l,j,k,Aa,b,α,β0:a+bj,α+βk,βl+bξ𝑑ξr1αlb+k|ξ|l+bβ|t|bk(1+|ξ/r|)A.\partial_{\log{r}}^{j}\mathfrak{q}_{r}^{(l)}(t)\ll_{l,j,k,A}\sum_{\begin{subarray}{c}a,b,\alpha,\beta\geq 0:\\ a+b\leq j,\;\alpha+\beta\leq k,\;\beta\leq l+b\end{subarray}}\int_{\xi\in\mathbb{R}}d\xi\,\frac{r^{-1-\alpha-l-b+k}\lvert\xi\rvert^{l+b-\beta}\lvert t\rvert^{b-k}}{(1+\lvert\xi/r\rvert)^{A}}.

Let A=l+j+2A=l+j+2; then each integral here over ξ\xi is l,j,kr1αlb+kr1+l+bβ|t|bk\ll_{l,j,k}r^{-1-\alpha-l-b+k}r^{1+l+b-\beta}\lvert t\rvert^{b-k} (by Lemma 5.3), and thus l,j,krkαβ|t|bk|t|bk\ll_{l,j,k}r^{k-\alpha-\beta}\lvert t\rvert^{b-k}\ll\lvert t\rvert^{b-k} (since kαβ0k-\alpha-\beta\geq 0 and r1r\ll 1). If |t|1\lvert t\rvert\leq 1, we are done (since k=0k=0). If |t|>1\lvert t\rvert>1, we may replace kk with k+jk+j to finish (since bjb\leq j). ∎

We now put (8.3) in a broader framework. For any Schwartz functions q:q\colon\mathbb{R}\to\mathbb{C} and ϕCc(m)\phi\in C^{\infty}_{c}(\mathbb{R}^{m})\otimes\mathbb{C} with SuppϕSuppw\operatorname{Supp}{\phi}\subseteq\operatorname{Supp}{w}, let

(8.4) 𝒥r,𝒗(q,ϕ)\colonequalsu𝑑uq(ru)𝒙m𝑑𝒙ϕ(𝒙)e(uF(𝒙)𝒗𝒙).\mathscr{J}_{r,\bm{v}}(q,\phi)\colonequals\int_{u\in\mathbb{R}}du\,q(ru)\int_{\bm{x}\in\mathbb{R}^{m}}d\bm{x}\,\phi(\bm{x})e(uF(\bm{x})-\bm{v}\cdot\bm{x}).

Then by (8.3), we have J𝒄,X(n)=𝒥r,𝒗(𝔮r,w0)J_{\bm{c},X}(n)=\mathscr{J}_{r,\bm{v}}(\mathfrak{q}_{r},w_{0}), where (r,𝒗)=(n/Y,X𝒄/n)(r,\bm{v})=(n/Y,X\bm{c}/n).

Proposition 8.4.

Assume (2.6). Let qq, ϕ\phi be as above (with qq, ϕ\phi Schwartz and SuppϕSuppw\operatorname{Supp}\phi\subseteq\operatorname{Supp}{w}). Then for all (k,𝐯)0×m(k,\bm{v})\in\mathbb{Z}_{\geq 0}\times\mathbb{R}^{m} and positive reals rA0r\leq A_{0} and M𝐯M\geq\lVert\bm{v}\rVert, we have

𝒥r,𝒗(q,ϕ)F,w,q,ϕ,k(1+𝒗)1m/2(1+r𝒗)k(1+Δ(𝒗/M)𝒗)k.\mathscr{J}_{r,\bm{v}}(q,\phi)\ll_{F,w,q,\phi,k}(1+\lVert\bm{v}\rVert)^{1-m/2}\cdot(1+\lVert r\bm{v}\rVert)^{-k}(1+\lVert\Delta(\bm{v}/M)\bm{v}\rVert)^{-k}.

Before proving Proposition 8.4, we first explain why it implies the desired Proposition 8.1. In order to handle logn1\partial_{\log{n}}^{\geq 1}, we need a recursion, (8.5), originally observed to first order by [heath1996new]. Without such a recursion, we might suffer for small moduli nn, as in [hooley1986HasseWeil]*§9’s analysis of “junior arcs” (repaired in [hooley_greaves_harman_huxley_1997] for some purposes, by a clever averaging argument).

Lemma 8.5.

Let (r,𝐯)=(n/Y,X𝐜/n)(r,\bm{v})=(n/Y,X\bm{c}/n). Let q1(t)\colonequalstq(t)q_{1}(t)\colonequals t\cdot q(t) and let ϕ1(𝐱)\colonequalsdiv(ϕ1(𝐱)𝐱)=𝐱ϕ1(𝐱)=1jmxjxjϕ1\phi_{1}(\bm{x})\colonequals\operatorname{div}(\phi_{1}(\bm{x})\bm{x})=\bm{x}\cdot\nabla{\phi_{1}}(\bm{x})=\sum_{1\leq j\leq m}x_{j}\cdot\partial_{x_{j}}{\phi_{1}}. Let ϕ2,j(𝐱)\colonequalsxjϕ(𝐱)\phi_{2,j}(\bm{x})\colonequals x_{j}\cdot\phi(\bm{x}). Then

(8.5) logn𝒥r,𝒗(q,ϕ)\displaystyle\partial_{\log{n}}\mathscr{J}_{r,\bm{v}}(q,\phi) =(m3)𝒥r,𝒗(q,ϕ)2𝒥r,𝒗(q1,ϕ)+𝒥r,𝒗(q,ϕ1),\displaystyle=(m-3)\mathscr{J}_{r,\bm{v}}(q,\phi)-2\mathscr{J}_{r,\bm{v}}(q_{1},\phi)+\mathscr{J}_{r,\bm{v}}(q,\phi_{1}),
(8.6) cj𝒥r,𝒗(q,ϕ)\displaystyle\partial_{c_{j}}\mathscr{J}_{r,\bm{v}}(q,\phi) =(2πiX/n)𝒥r,𝒗(q,ϕ2,j).\displaystyle=(-2\pi iX/n)\mathscr{J}_{r,\bm{v}}(q,\phi_{2,j}).
Proof.

The formula (8.6) immediately follows upon differentiating (8.4) by cjc_{j}. It is possible to prove (8.5) by a clever integration by parts (cf. [heath1996new]*p. 182, proof of Lemma 14). We give a slightly shorter argument. Write u=t/r3u=t/r^{3} and 𝒙=r𝒚\bm{x}=r\bm{y} in (8.4) to get

(8.7) 𝒥r,𝒗(q,ϕ)=rm3𝑑tq(t/r2)m𝑑𝒚ϕ(r𝒚)e(tF(𝒚)r𝒗𝒚).\mathscr{J}_{r,\bm{v}}(q,\phi)=r^{m-3}\int_{\mathbb{R}}dt\,q(t/r^{2})\int_{\mathbb{R}^{m}}d\bm{y}\,\phi(r\bm{y})e(tF(\bm{y})-r\bm{v}\cdot\bm{y}).

Differentiating both sides of (8.7) by logn\log{n}, using lognr=r\partial_{\log{n}}{r}=r and the fact that r𝒗=X𝒄/Yr\bm{v}=X\bm{c}/Y is independent of logn\log{n}, we get (8.5), by (8.7) applied to each of (q,ϕ)(q,\phi), (q1,ϕ)(q_{1},\phi), (q,ϕ1)(q,\phi_{1}). ∎

Proof of Proposition 8.1, assuming Proposition 8.4.

Recall that by (8.3), we have J𝒄,X(n)=𝒥r,𝒗(𝔮r,w0)J_{\bm{c},X}(n)=\mathscr{J}_{r,\bm{v}}(\mathfrak{q}_{r},w_{0}), where (r,𝒗)=(n/Y,X𝒄/n)(r,\bm{v})=(n/Y,X\bm{c}/n). By Lemmas 8.3 and 8.5 (first using the chain rule, (8.5), and Lemma 8.3 when differentiating by logn\log{n}, and then using (8.6) when differentiating 𝒄\bm{c}), we may thus write 𝒄𝜶lognjJ𝒄,X(n)\partial_{\bm{c}}^{\bm{\alpha}}\partial_{\log{n}}^{j}J_{\bm{c},X}(n) as a finite linear combination of integrals 𝒥r,𝒗(q,ϕ)\mathscr{J}_{r,\bm{v}}(q,\phi), with coefficients Om,j,𝜶((X/n)|𝜶|)O_{m,j,\bm{\alpha}}((X/n)^{\lvert\bm{\alpha}\rvert}), running over a set of at most 4j4^{j} pairs (q,ϕ)(q,\phi). (For example, for j=1j=1 and 𝜶=𝟎\bm{\alpha}=\bm{0}, we would use the chain rule and (8.5) to write logn𝒥r,𝒗(𝔮r,w0)\partial_{\log{n}}\mathscr{J}_{r,\bm{v}}(\mathfrak{q}_{r},w_{0}) as

𝒥r,𝒗(logr𝔮r,w0)+(m3)𝒥r,𝒗(𝔮r,w0)2𝒥r,𝒗(𝔮r,1,w0)+𝒥r,𝒗(q,w0,1),\mathscr{J}_{r,\bm{v}}(\partial_{\log{r}}{\mathfrak{q}_{r}},w_{0})+(m-3)\mathscr{J}_{r,\bm{v}}(\mathfrak{q}_{r},w_{0})-2\mathscr{J}_{r,\bm{v}}(\mathfrak{q}_{r,1},w_{0})+\mathscr{J}_{r,\bm{v}}(q,w_{0,1}),

where 𝔮r,1(t)=t𝔮r(t)\mathfrak{q}_{r,1}(t)=t\cdot\mathfrak{q}_{r}(t) and w0,1(𝒙)\colonequalsdiv(w0(𝒙)𝒙)w_{0,1}(\bm{x})\colonequals\operatorname{div}(w_{0}(\bm{x})\bm{x}).) Proposition 8.1 then immediately follows from Proposition 8.4 (applied to each individual 𝒥r,𝒗(q,ϕ)\mathscr{J}_{r,\bm{v}}(q,\phi)). ∎

To prove Proposition 8.4, we need the following lemma; the basic principle is familiar (see e.g. [hormander1990analysis]*Theorem 7.7.1) but the treatment of [heath1996new] is ideal for us.

Lemma 8.6 (Non-stationary phase).

Let λ,A>0\lambda,A\in\mathbb{R}_{>0} and d,k1d,k\in\mathbb{Z}_{\geq 1}. Suppose aCc(d)a\in C^{\infty}_{c}(\mathbb{R}^{d})\otimes\mathbb{C} is supported on [A,A]d[-A,A]^{d}, with |𝛂|k|𝐱𝛂a(𝐱)|A\sum_{\lvert\bm{\alpha}\rvert\leq k}\lvert\partial_{\bm{x}}^{\bm{\alpha}}{a}(\bm{x})\rvert\leq A for all 𝐱d\bm{x}\in\mathbb{R}^{d}. Suppose f:df\colon\mathbb{R}^{d}\to\mathbb{R} is smooth, with f(𝐱)λ/A\lVert\nabla{f}(\bm{x})\rVert\geq\lambda/A and 2|𝛂|k+1|𝐱𝛂f(𝐱)|Aλ\sum_{2\leq\lvert\bm{\alpha}\rvert\leq k+1}\lvert\partial_{\bm{x}}^{\bm{\alpha}}{f}(\bm{x})\rvert\leq A\lambda for all 𝐱Suppa\bm{x}\in\operatorname{Supp}{a}. Then

𝒙d𝑑𝒙a(𝒙)e(f(𝒙))d,A,kλk.\int_{\bm{x}\in\mathbb{R}^{d}}d\bm{x}\,a(\bm{x})e(f(\bm{x}))\ll_{d,A,k}\lambda^{-k}.
Proof.

See [heath1996new]*Lemma 10 and its proof (repeated integration by parts); see [heath1996new]*§2 for the definition of 𝒞(S)\mathscr{C}(S) (which allows for the required uniformity over weight functions). Note that we do not require any explicit upper bound on sup𝒙Suppa|f(𝒙)|\sup_{\bm{x}\in\operatorname{Supp}{a}}{\lvert f(\bm{x})\rvert}, or any control on the shape of the compact set Suppa[A,A]d\operatorname{Supp}{a}\subseteq[-A,A]^{d}. ∎

Proof of Proposition 8.4.

Certainly 𝟎Suppw\bm{0}\notin\operatorname{Supp}{w} by (2.6). Since VV is smooth, we thus have F(𝒙)>0\lVert\nabla{F}(\bm{x})\rVert>0 for all 𝒙Suppw\bm{x}\in\operatorname{Supp}{w}. Since Suppw\operatorname{Supp}{w} is compact, there thus exists A9=A9(F,w)2A_{9}=A_{9}(F,w)\geq 2 such that for all 𝒙Suppw\bm{x}\in\operatorname{Supp}{w}, we have 𝒙,F(𝒙)[A91,A9]\lVert\bm{x}\rVert,\lVert\nabla{F}(\bm{x})\rVert\in[A_{9}^{-1},A_{9}]. Let 𝒲\colonequals{a𝒙:(a,𝒙)[1/3A9,3A9]×Suppw}\mathscr{W}\colonequals\{a\bm{x}:(a,\bm{x})\in[1/3A_{9},3A_{9}]\times\operatorname{Supp}{w}\} be the union of all dilates of Suppw\operatorname{Supp}{w} by a scale factor a[1/3A9,3A9]a\in[1/3A_{9},3A_{9}]. The set 𝒲\mathscr{W} is compact, being the continuous image of a product of compact sets. Also, since the right-hand side of (2.6) is invariant under scaling, we have

(8.8) 𝒲{𝒘m:det(HessF(𝒘))0}.\mathscr{W}\subseteq\{\bm{w}\in\mathbb{R}^{m}:\det(\operatorname{Hess}{F}(\bm{w}))\neq 0\}.

Let A10=A10(F,w)2A_{10}=A_{10}(F,w)\geq 2 be a constant such that for all 𝒘𝒲\bm{w}\in\mathscr{W}, we have

(8.9) A101𝒘A10,A101F(𝒘)A10.A_{10}^{-1}\leq\lVert\bm{w}\rVert\leq A_{10},\quad A_{10}^{-1}\leq\lVert\nabla{F}(\bm{w})\rVert\leq A_{10}.

Our plan is to first consider the 𝒙\bm{x}-integral in 𝒥r,𝒗(q,ϕ)\mathscr{J}_{r,\bm{v}}(q,\phi) (see (8.4)) for a single value of uu at a time, and then integrate over uu. Given uu, 𝒗\bm{v}, let ψ0(𝒙)\colonequalsuF(𝒙)𝒗𝒙\psi_{0}(\bm{x})\colonequals uF(\bm{x})-\bm{v}\cdot\bm{x} and 𝒥u,𝒗(ϕ)\colonequals𝒙m𝑑𝒙ϕ(𝒙)e(ψ0(𝒙))\mathcal{J}_{u,\bm{v}}(\phi)\colonequals\int_{\bm{x}\in\mathbb{R}^{m}}d\bm{x}\,\phi(\bm{x})e(\psi_{0}(\bm{x})). Taking absolute values in 𝒥\mathcal{J} gives the trivial bound

(8.10) 𝒥u,𝒗(ϕ)ϕ1.\mathcal{J}_{u,\bm{v}}(\phi)\ll_{\phi}1.

On the other hand, ψ0(𝒙)=uF(𝒙)𝒗\nabla{\psi_{0}}(\bm{x})=u\nabla{F}(\bm{x})-\bm{v}. In particular, if |u|2A9𝒗\lvert u\rvert\geq 2A_{9}\lVert\bm{v}\rVert or |u|𝒗/2A9\lvert u\rvert\leq\lVert\bm{v}\rVert/2A_{9}, then we have ψ0(𝒙)max(|u/2A9|,𝒗/2)\lVert\nabla{\psi_{0}}(\bm{x})\rVert\geq\max(\lvert u/2A_{9}\rvert,\lVert\bm{v}\rVert/2) and 𝒙𝜶ψ0(𝒙)F,w,𝜶|u|+𝒗\partial_{\bm{x}}^{\bm{\alpha}}{\psi_{0}}(\bm{x})\ll_{F,w,\bm{\alpha}}\lvert u\rvert+\lVert\bm{v}\rVert for all 𝒙Suppw\bm{x}\in\operatorname{Supp}{w}, and thus 𝒥u,𝒗(ϕ)ϕ,k(|u|+𝒗)k\mathcal{J}_{u,\bm{v}}(\phi)\ll_{\phi,k}(\lvert u\rvert+\lVert\bm{v}\rVert)^{-k} by Lemma 8.6 (provided |u|+𝒗>0\lvert u\rvert+\lVert\bm{v}\rVert>0). This, together with (8.10) and the bound q(ru)q1q(ru)\ll_{q}1, gives

(8.11) |u|2A9𝒗𝑑u|q(ru)𝒥u,𝒗(ϕ)|\displaystyle\int_{\lvert u\rvert\geq 2A_{9}\lVert\bm{v}\rVert}du\,\lvert q(ru)\cdot\mathcal{J}_{u,\bm{v}}(\phi)\rvert q,ϕ,b|u|2A9𝒗du(1+|u|)b+1b(1+𝒗)b,\displaystyle\ll_{q,\phi,b}\int_{\lvert u\rvert\geq 2A_{9}\lVert\bm{v}\rVert}\frac{du}{(1+\lvert u\rvert)^{b+1}}\ll_{b}(1+\lVert\bm{v}\rVert)^{-b},
(8.12) |u|𝒗/2A9𝑑u|q(ru)𝒥u,𝒗(ϕ)|\displaystyle\int_{\lvert u\rvert\leq\lVert\bm{v}\rVert/2A_{9}}du\,\lvert q(ru)\cdot\mathcal{J}_{u,\bm{v}}(\phi)\rvert q,ϕ,b|u|𝒗/2A9du(1+𝒗)b+1b(1+𝒗)b,\displaystyle\ll_{q,\phi,b}\int_{\lvert u\rvert\leq\lVert\bm{v}\rVert/2A_{9}}\frac{du}{(1+\lVert\bm{v}\rVert)^{b+1}}\ll_{b}(1+\lVert\bm{v}\rVert)^{-b},

for all integers b1b\geq 1. Fix w1Cc()w_{1}\in C^{\infty}_{c}(\mathbb{R}) with

w1|[2A9,2A9][1/2A9,1/2A9]=1,Suppw1[3A9,3A9][1/3A9,1/3A9].w_{1}|_{[-2A_{9},2A_{9}]\setminus[-1/2A_{9},1/2A_{9}]}=1,\quad\operatorname{Supp}{w_{1}}\subseteq[-3A_{9},3A_{9}]\setminus[-1/3A_{9},1/3A_{9}].

Then by (8.11), (8.12), and the triangle inequality, we have (for integers b1b\geq 1)

(8.13) 𝒥r,𝒗(q,ϕ)u𝑑uq(ru)w1(u/𝒗)𝒥u,𝒗(ϕ)q,ϕ,b(1+𝒗)b.\mathscr{J}_{r,\bm{v}}(q,\phi)-\int_{u\in\mathbb{R}}du\,q(ru)w_{1}(u/\lVert\bm{v}\rVert)\mathcal{J}_{u,\bm{v}}(\phi)\ll_{q,\phi,b}(1+\lVert\bm{v}\rVert)^{-b}.

It remains to handle |u|/𝒗[1/3A9,3A9]\lvert u\rvert/\lVert\bm{v}\rVert\in[1/3A_{9},3A_{9}]. Suppose first that 𝒗1\lVert\bm{v}\rVert\leq 1. Plugging (8.10) and the bound q(ru)q1q(ru)\ll_{q}1 into (8.13), we get 𝒥r,𝒗(q,ϕ)q,ϕ(1+𝒗)1+|u|3A9𝑑u 11\mathscr{J}_{r,\bm{v}}(q,\phi)\ll_{q,\phi}(1+\lVert\bm{v}\rVert)^{-1}+\int_{\lvert u\rvert\leq 3A_{9}}du\,1\ll 1. This bound on 𝒥r,𝒗(q,ϕ)\mathscr{J}_{r,\bm{v}}(q,\phi) fits in Proposition 8.4, since r𝒗A0\lVert r\bm{v}\rVert\leq A_{0} and Δ(𝒗/M)𝒗F1\lVert\Delta(\bm{v}/M)\bm{v}\rVert\ll_{F}1.

For the rest of the proof, suppose 𝒗1\lVert\bm{v}\rVert\geq 1, let u~\colonequalsu/𝒗\tilde{u}\colonequals u/\lVert\bm{v}\rVert and 𝒗~\colonequals𝒗/𝒗\tilde{\bm{v}}\colonequals\bm{v}/\lVert\bm{v}\rVert, let 𝒛\colonequals|u~|1/2𝒙\bm{z}\colonequals\lvert\tilde{u}\rvert^{1/2}\bm{x}, let sgn(u)\colonequalsu/|u|{1,1}\operatorname{sgn}(u)\colonequals u/\lvert u\rvert\in\{-1,1\}, and let

ψ1(𝒛)\colonequals|u~|1/2ψ0(𝒙)/𝒗=|u~|1/2ψ0(𝒛/|u~|1/2)/𝒗=sgn(u)F(𝒛)𝒗~𝒛.\psi_{1}(\bm{z})\colonequals\lvert\tilde{u}\rvert^{1/2}\psi_{0}(\bm{x})/\lVert\bm{v}\rVert=\lvert\tilde{u}\rvert^{1/2}\psi_{0}(\bm{z}/\lvert\tilde{u}\rvert^{1/2})/\lVert\bm{v}\rVert=\operatorname{sgn}(u)F(\bm{z})-\tilde{\bm{v}}\cdot\bm{z}.

(The normalization by |u~|1/2\lvert\tilde{u}\rvert^{1/2} makes ψ1\psi_{1} nearly independent of uu; this is crucial later.) Note that ψ1(𝒛)=sgn(u)F(𝒛)𝒗~=u~F(𝒙)𝒗~\nabla{\psi_{1}}(\bm{z})=\operatorname{sgn}(u)\nabla{F}(\bm{z})-\tilde{\bm{v}}=\tilde{u}\nabla{F}(\bm{x})-\tilde{\bm{v}}. Now consider an individual u~\tilde{u}\in\mathbb{R} with

(8.14) 1/3A9|u~|3A9.1/3A_{9}\leq\lvert\tilde{u}\rvert\leq 3A_{9}.

For all 𝒙Suppw\bm{x}\in\operatorname{Supp}{w}, we have 𝒛𝒲\bm{z}\in\mathscr{W}, so 𝒛[1/A10,A10]\lVert\bm{z}\rVert\in[1/A_{10},A_{10}] and ψ1(𝒛)F(𝒛)+12A10\lVert\nabla{\psi_{1}}(\bm{z})\rVert\leq\lVert\nabla{F}(\bm{z})\rVert+1\leq 2A_{10}. Furthermore, the condition (8.8) implies

(8.15) vol{𝒛𝒲:ψ1(𝒛)mλ}F,wλm,\operatorname{vol}{\{\bm{z}\in\mathscr{W}:\lVert\nabla{\psi_{1}}(\bm{z})\rVert\leq m\lambda\}}\ll_{F,w}\lambda^{m},

uniformly over λ>0\lambda\in\mathbb{R}_{>0}, by calculus (see [heath1996new]*Lemma 21 and its proof). We will split 𝒥u,𝒗(ϕ)\mathcal{J}_{u,\bm{v}}(\phi) into pieces according to the size of ψ1(𝒛)\lVert\nabla{\psi_{1}}(\bm{z})\rVert; cf. [huang2020density]*𝒦i\mathscr{K}_{i} on p. 2061. To avoid the need for explicit stationary phase expansions (which can be messy after the leading term), we will also use a subdivision process inspired by [heath1996new]*§8.

Recall the weights ν0\nu_{0}, ν1\nu_{1} from §1.2. Let ν¯0\colonequals1ν0\overline{\nu}_{0}\colonequals 1-\nu_{0}. For reals λ>0\lambda>0, let

(8.16) 𝒥0,u,𝒗\displaystyle\mathcal{J}_{0,u,\bm{v}} \colonequalsmd𝒛|u~|m/2ν0(ψ1(𝒛)𝒗1/2)ϕ(𝒙)e(𝒗ψ1(𝒛)|u~|1/2),\displaystyle\colonequals\int_{\mathbb{R}^{m}}\frac{d\bm{z}}{\lvert\tilde{u}\rvert^{m/2}}\,\nu_{0}{\left(\frac{\nabla{\psi_{1}}(\bm{z})}{\lVert\bm{v}\rVert^{-1/2}}\right)}\phi(\bm{x})e{\left(\frac{\lVert\bm{v}\rVert\psi_{1}(\bm{z})}{\lvert\tilde{u}\rvert^{1/2}}\right)},
(8.17) 𝒥1,u,𝒗,λ\displaystyle\mathcal{J}_{1,u,\bm{v},\lambda} \colonequalsmd𝒛|u~|m/2ν¯0(ψ1(𝒛)𝒗1/2)ν1(ψ1(𝒛)λ)ϕ(𝒙)e(𝒗ψ1(𝒛)|u~|1/2),\displaystyle\colonequals\int_{\mathbb{R}^{m}}\frac{d\bm{z}}{\lvert\tilde{u}\rvert^{m/2}}\,\overline{\nu}_{0}{\left(\frac{\nabla{\psi_{1}}(\bm{z})}{\lVert\bm{v}\rVert^{-1/2}}\right)}\nu_{1}{\left(\frac{\nabla{\psi_{1}}(\bm{z})}{\lambda}\right)}\phi(\bm{x})e{\left(\frac{\lVert\bm{v}\rVert\psi_{1}(\bm{z})}{\lvert\tilde{u}\rvert^{1/2}}\right)},

where ϕ(𝒙)=ϕ(𝒛/|u~|1/2)\phi(\bm{x})=\phi(\bm{z}/\lvert\tilde{u}\rvert^{1/2}). By the definitions of ν0\nu_{0}, ν1\nu_{1}, we have 𝒥1,u,𝒗,λ=0\mathcal{J}_{1,u,\bm{v},\lambda}=0 unless there exists 𝒛𝒲\bm{z}\in\mathscr{W} (corresponding to 𝒙Suppw\bm{x}\in\operatorname{Supp}{w}) satisfying ψ1(𝒛)>𝒗1/2/2\lVert\nabla{\psi_{1}}(\bm{z})\rVert>\lVert\bm{v}\rVert^{-1/2}/2 and λψ1(𝒛)mλ\lambda\leq\lVert\nabla{\psi_{1}}(\bm{z})\rVert\leq m\lambda. Therefore, 𝒥1,u,𝒗,λ=0\mathcal{J}_{1,u,\bm{v},\lambda}=0 unless 𝒗1/2/2m<λ2A10\lVert\bm{v}\rVert^{-1/2}/2m<\lambda\leq 2A_{10}. Since λ>0d×λν1(𝒕/λ)=1\int_{\lambda>0}d^{\times}{\lambda}\,\nu_{1}(\bm{t}/\lambda)=1 for all 𝒕m{𝟎}\bm{t}\in\mathbb{R}^{m}\setminus\{\bm{0}\}, we obtain (under (8.14)) the decomposition

(8.18) 𝒥u,𝒗(ϕ)𝒥0,u,𝒗=𝒗1/2/2m2A10d×λ𝒥1,u,𝒗,λ.\mathcal{J}_{u,\bm{v}}(\phi)-\mathcal{J}_{0,u,\bm{v}}=\int_{\lVert\bm{v}\rVert^{-1/2}/2m}^{2A_{10}}d^{\times}{\lambda}\,\mathcal{J}_{1,u,\bm{v},\lambda}.

Fix functions ω1,,ωmCc(m)\omega_{1},\dots,\omega_{m}\in C^{\infty}_{c}(\mathbb{R}^{m}), all supported on Suppν1\operatorname{Supp}{\nu_{1}}, such that ω1++ωm=ν1\omega_{1}+\dots+\omega_{m}=\nu_{1} and we have 1/2|tj|2m1/2\leq\lvert t_{j}\rvert\leq 2m for all 𝒕Suppωj\bm{t}\in\operatorname{Supp}{\omega_{j}}. Let 𝒥1,j,u,𝒗,λ\mathcal{J}_{1,j,u,\bm{v},\lambda} denote 𝒥1,u,𝒗,λ\mathcal{J}_{1,u,\bm{v},\lambda} with ωj\omega_{j} in place of ν1\nu_{1}; clearly 𝒥1,u,𝒗,λ=1jm𝒥1,j,u,𝒗,λ\mathcal{J}_{1,u,\bm{v},\lambda}=\sum_{1\leq j\leq m}\mathcal{J}_{1,j,u,\bm{v},\lambda}.

We proceed based on the following rough idea: if ψ1(𝒛)𝒗1/2\lVert\nabla{\psi_{1}(\bm{z})}\rVert\gg\lVert\bm{v}\rVert^{-1/2} with a large constant, integration by parts over 𝒛\bm{z} is useful; and if ψ1(𝒛)|Δ(𝒗~)|1/2\lVert\nabla{\psi_{1}(\bm{z})}\rVert\ll\lvert\Delta(\tilde{\bm{v}})\rvert^{1/2} with a small constant, integration by parts over uu is useful (if Δ(𝒗~)𝒗1\lVert\Delta(\tilde{\bm{v}})\bm{v}\rVert\gg 1 with any constant).

Let λ[4𝒗1/2,2A10]\lambda\in[4\lVert\bm{v}\rVert^{-1/2},2A_{10}]. Since λ4𝒗1/2\lambda\geq 4\lVert\bm{v}\rVert^{-1/2}, the supports of ν0(ψ1(𝒛)/𝒗1/2)\nu_{0}(\nabla{\psi_{1}}(\bm{z})/\lVert\bm{v}\rVert^{-1/2}) and ν1(ψ1(𝒛)/λ)\nu_{1}(\nabla{\psi_{1}}(\bm{z})/\lambda) are disjoint, so ν¯0(ψ1(𝒛)/𝒗1/2)ωj(ψ1(𝒛)/λ)=ωj(ψ1(𝒛)/λ)\overline{\nu}_{0}(\nabla{\psi_{1}}(\bm{z})/\lVert\bm{v}\rVert^{-1/2})\omega_{j}(\nabla{\psi_{1}}(\bm{z})/\lambda)=\omega_{j}(\nabla{\psi_{1}}(\bm{z})/\lambda). Thus

(8.19) |u~|m/2𝒥1,j,u,𝒗,λ=m𝑑𝒛w2,j,0(𝒛)e(𝒗ψ1(𝒛)/|u~|1/2),\lvert\tilde{u}\rvert^{m/2}\mathcal{J}_{1,j,u,\bm{v},\lambda}=\int_{\mathbb{R}^{m}}d\bm{z}\,w_{2,j,0}(\bm{z})e(\lVert\bm{v}\rVert\psi_{1}(\bm{z})/\lvert\tilde{u}\rvert^{1/2}),

where w2,j,0=w2,j,0,u,𝒗,λ(𝒛)\colonequalsωj(ψ1(𝒛)/λ)ϕ(𝒛/|u~|1/2)w_{2,j,0}=w_{2,j,0,u,\bm{v},\lambda}(\bm{z})\colonequals\omega_{j}(\nabla{\psi_{1}}(\bm{z})/\lambda)\phi(\bm{z}/\lvert\tilde{u}\rvert^{1/2}).

Since Suppωj{𝒕m:1/2|tj|2m}Suppν1\operatorname{Supp}{\omega_{j}}\subseteq\{\bm{t}\in\mathbb{R}^{m}:1/2\leq\lvert t_{j}\rvert\leq 2m\}\cap\operatorname{Supp}{\nu_{1}}, we have

(8.20) zjψ1(𝒛)λ/2,ψ1(𝒛)mλ,\lVert\partial_{z_{j}}{\psi_{1}}(\bm{z})\rVert\geq\lambda/2,\quad\lVert\nabla{\psi_{1}}(\bm{z})\rVert\leq m\lambda,

for all 𝒛Suppw2,j,0\bm{z}\in\operatorname{Supp}{w_{2,j,0}}. For k1k\geq 1, recursively define w2,j,k=w2,j,k,u,𝒗,λ(𝒛)w_{2,j,k}=w_{2,j,k,u,\bm{v},\lambda}(\bm{z}) (a sequence of smooth functions m\mathbb{R}^{m}\to\mathbb{C} supported on Suppw2,j,0\operatorname{Supp}{w_{2,j,0}}) by setting

(8.21) w2,j,k\colonequals(zjψ1)kzj(w2,j,k1/(zjψ1)k)w_{2,j,k}\colonequals-(\partial_{z_{j}}{\psi_{1}})^{k}\cdot\partial_{z_{j}}(w_{2,j,k-1}/(\partial_{z_{j}}{\psi_{1}})^{k})

Integrating by parts kk times in (8.19), we get (cf. [heath1996new]*(5.4), from the proof of Lemma 10)

(8.22) |u~|m/2𝒥1,j,u,𝒗,λ=m𝑑𝒛w2,j,k,u,𝒗,λ(𝒛)(𝒗zjψ1(𝒛)/|u~|1/2)ke(𝒗ψ1(𝒛)/|u~|1/2).\lvert\tilde{u}\rvert^{m/2}\mathcal{J}_{1,j,u,\bm{v},\lambda}=\int_{\mathbb{R}^{m}}d\bm{z}\,\frac{w_{2,j,k,u,\bm{v},\lambda}(\bm{z})}{(\lVert\bm{v}\rVert\partial_{z_{j}}{\psi_{1}}(\bm{z})/\lvert\tilde{u}\rvert^{1/2})^{k}}\cdot e(\lVert\bm{v}\rVert\psi_{1}(\bm{z})/\lvert\tilde{u}\rvert^{1/2}).

We claim that for each k0k\geq 0, there exists a smooth function w3,j,k:m(k+2)+3w_{3,j,k}\colon\mathbb{R}^{m(k+2)+3}\to\mathbb{C} of the form w3,j,k=w3,j,k(𝒂1,0,,𝒂1,k,𝒂2,0,a2,1,a3,a4)w_{3,j,k}=w_{3,j,k}(\bm{a}_{1,0},\dots,\bm{a}_{1,k},\bm{a}_{2,0},a_{2,1},a_{3},a_{4}), defined in terms of FF, jj, kk (independently of uu, 𝒗\bm{v}, λ\lambda), such that for all uu, 𝒗\bm{v}, λ\lambda currently under consideration (namely, satisfying 𝒗1\lVert\bm{v}\rVert\geq 1, (8.14), and λ[4𝒗1/2,2A10]\lambda\in[4\lVert\bm{v}\rVert^{-1/2},2A_{10}]), and for all 𝒛Suppw2,j,0\bm{z}\in\operatorname{Supp}{w_{2,j,0}}, we have

(8.23) λkw2,j,k(𝒛)=w3,j,k(ψ1(𝒛)λ,zj1ψ1(𝒛),,zjkψ1(𝒛),𝒛|u~|1/2,1|u~|1/2,λzjψ1,λ).\lambda^{k}\cdot w_{2,j,k}(\bm{z})=w_{3,j,k}{\left(\frac{\nabla{\psi_{1}}(\bm{z})}{\lambda},\nabla{\partial_{z_{j}}^{1}{\psi_{1}}}(\bm{z}),\dots,\nabla{\partial_{z_{j}}^{k}{\psi_{1}}}(\bm{z}),\frac{\bm{z}}{\lvert\tilde{u}\rvert^{1/2}},\frac{1}{\lvert\tilde{u}\rvert^{1/2}},\frac{\lambda}{\partial_{z_{j}}{\psi_{1}}},\lambda\right)}.

This claim can be easily proven by induction on k0k\geq 0, where for k=0k=0 we take

w3,j,0(𝒂1,0,𝒂2,0,a2,1,a3,a4)\colonequalsωj(𝒂1,0)ϕ(𝒂2,0),w_{3,j,0}(\bm{a}_{1,0},\bm{a}_{2,0},a_{2,1},a_{3},a_{4})\colonequals\omega_{j}(\bm{a}_{1,0})\phi(\bm{a}_{2,0}),

and for k1k\geq 1 we use the product and chain rules in (8.21) (to compute λkw2,j,k=λλk1w2,j,k\lambda^{k}w_{2,j,k}=\lambda\cdot\lambda^{k-1}w_{2,j,k} in terms of λk1w2,j,k1\lambda^{k-1}w_{2,j,k-1} and its zjz_{j}-derivatives), and observe that (for l1l\geq 1)

λzj(ψ1λ)=zj1ψ1,λzj(zjl)=λzjl+1,λzj(𝒛)=λ,\displaystyle\lambda\cdot\partial_{z_{j}}{\left(\frac{\psi_{1}}{\lambda}\right)}=\partial_{z_{j}}^{1}{\psi_{1}},\quad\lambda\cdot\partial_{z_{j}}(\partial_{z_{j}}^{l})=\lambda\cdot\partial_{z_{j}}^{l+1},\quad\lambda\cdot\partial_{z_{j}}(\bm{z})=\lambda,
λzj(λzjψ1)=(zj2ψ1)(λzjψ1)2,(zjψ1)kzj(λ(zjψ1)k)=(zj2ψ1)kλzjψ1.\displaystyle\lambda\cdot\partial_{z_{j}}{\left(\frac{\lambda}{\partial_{z_{j}}{\psi_{1}}}\right)}=-(\partial_{z_{j}}^{2}{\psi_{1}})\cdot\left(\frac{\lambda}{\partial_{z_{j}}{\psi_{1}}}\right)^{2},\quad(\partial_{z_{j}}{\psi_{1}})^{k}\cdot\partial_{z_{j}}{\left(\frac{\lambda}{(\partial_{z_{j}}{\psi_{1}})^{k}}\right)}=(\partial_{z_{j}}^{2}{\psi_{1}})\cdot\frac{-k\lambda}{\partial_{z_{j}}{\psi_{1}}}.

(This proof uses the smoothness of ψ1\psi_{1}, and the nonvanishing of zjψ1\partial_{z_{j}}{\psi_{1}} on Suppw2,j,0\operatorname{Supp}{w_{2,j,0}}.)

By (8.9), we have 𝜶ψ1(𝒛)𝜶1\partial^{\bm{\alpha}}{\psi_{1}}(\bm{z})\ll_{\bm{\alpha}}1 whenever |𝜶|2\lvert\bm{\alpha}\rvert\geq 2 and 𝒛𝒲\bm{z}\in\mathscr{W}. Inserting this and the bounds (8.20), (8.14), (8.9), and λ1\lambda\ll 1 into (8.23), we find that λkw2,j,k(𝒛)w3,j,k1\lambda^{k}w_{2,j,k}(\bm{z})\ll_{w_{3,j,k}}1 for all 𝒛m\bm{z}\in\mathbb{R}^{m}. Thus (8.22) and (8.20) immediately give (for λ[4𝒗1/2,2A10]\lambda\in[4\lVert\bm{v}\rVert^{-1/2},2A_{10}])

(8.24) |u~|m/2𝒥1,j,u,𝒗,λk𝒛𝒲𝑑𝒛𝟏ψ1(𝒛)mλ(λ2𝒗/|u~|1/2)k,\lvert\tilde{u}\rvert^{m/2}\mathcal{J}_{1,j,u,\bm{v},\lambda}\ll_{k}\int_{\bm{z}\in\mathscr{W}}d\bm{z}\,\frac{\bm{1}_{\lVert\nabla{\psi_{1}}(\bm{z})\rVert\leq m\lambda}}{(\lambda^{2}\lVert\bm{v}\rVert/\lvert\tilde{u}\rvert^{1/2})^{k}},

under (8.14). Our derivation of (8.24) has essentially followed the proof of Lemma 8.6 (with a small but important twist: we use Suppw2,j,kSuppw2,j,0\operatorname{Supp}{w_{2,j,k}}\subseteq\operatorname{Supp}{w_{2,j,0}} to get the factor 𝟏ψ1(𝒛)mλ\bm{1}_{\lVert\nabla{\psi_{1}}(\bm{z})\rVert\leq m\lambda}), but in some key ranges we need to go further. We need to decide when to integrate by parts over u~\tilde{u}\in\mathbb{R}; this will be informed by the next two paragraphs.

By (8.8) and the inverse function theorem, we know that for each 𝒛𝒲\bm{z}\in\mathscr{W}, there exists an open neighborhood UmU\subseteq\mathbb{R}^{m} of 𝒛\bm{z} such that the gradient map F:mm\nabla{F}\colon\mathbb{R}^{m}\to\mathbb{R}^{m} maps UU diffeomorphically onto F(U)\nabla{F}(U). But F(𝒲)\nabla{F}(\mathscr{W}) is compact (since 𝒲\mathscr{W} is compact), so there exists a constant η6=η6(F,w)>0\eta_{6}=\eta_{6}(F,w)>0 such that for any 𝒛𝒲\bm{z}\in\mathscr{W} and 𝒃[1,1]m\bm{b}\in[-1,1]^{m} with F(𝒛)𝒃η6\lVert\nabla{F}(\bm{z})-\bm{b}\rVert\leq\eta_{6}, there exists 𝒔(F)1(𝒃)\bm{s}\in(\nabla{F})^{-1}(\bm{b}) with 𝒛𝒔η61F(𝒛)𝒃\lVert\bm{z}-\bm{s}\rVert\leq\eta_{6}^{-1}\lVert\nabla{F}(\bm{z})-\bm{b}\rVert.

We apply this as follows. Let 𝒃=sgn(u)𝒗~\bm{b}=\operatorname{sgn}(u)\tilde{\bm{v}} and 𝒛𝒲\bm{z}\in\mathscr{W}, and suppose ψ1(𝒛)η6\lVert\nabla{\psi_{1}(\bm{z})}\rVert\leq\eta_{6}. Then there exists 𝒔m\bm{s}\in\mathbb{R}^{m} with ψ1(𝒔)=𝟎\nabla{\psi_{1}}(\bm{s})=\bm{0} and 𝒛𝒔η61ψ1(𝒛)\lVert\bm{z}-\bm{s}\rVert\leq\eta_{6}^{-1}\lVert\nabla{\psi_{1}(\bm{z})}\rVert. Since ψ1(𝒔)=𝟎\nabla{\psi_{1}}(\bm{s})=\bm{0}, Taylor expansion of ψ1\psi_{1} at 𝒔\bm{s} gives ψ1(𝒛)=ψ1(𝒔)+OF,w(𝒛𝒔2)\psi_{1}(\bm{z})=\psi_{1}(\bm{s})+O_{F,w}(\lVert\bm{z}-\bm{s}\rVert^{2}). Furthermore, ψ1(𝒔)=𝟎\nabla{\psi_{1}}(\bm{s})=\bm{0} implies 𝒃𝒔=F(𝒔)𝒔=3F(𝒔)\bm{b}\cdot\bm{s}=\nabla{F}(\bm{s})\cdot\bm{s}=3F(\bm{s}), so |ψ1(𝒔)|=|2F(𝒔)|\lvert\psi_{1}(\bm{s})\rvert=\lvert 2F(\bm{s})\rvert. But Δ(𝒃)=Δ(F(𝒔))F,w|F(𝒔)|\Delta(\bm{b})=\Delta(\nabla{F}(\bm{s}))\ll_{F,w}\lvert F(\bm{s})\rvert by (8.1). Combining the above, we get

(8.25) |ψ1(𝒛)|η7|Δ(𝒗~)|η71ψ1(𝒛)2(for some η7=η7(F,w)>0).\lvert\psi_{1}(\bm{z})\rvert\geq\eta_{7}\lvert\Delta(\tilde{\bm{v}})\rvert-\eta_{7}^{-1}\lVert\nabla{\psi_{1}(\bm{z})}\rVert^{2}\quad\textnormal{(for some $\eta_{7}=\eta_{7}(F,w)>0$)}.

Let A11\colonequals1+sup𝒂[1,1]m|Δ(𝒂)|A_{11}\colonequals 1+\sup_{\bm{a}\in[-1,1]^{m}}{\lvert\Delta(\bm{a})\rvert}, and let η8\colonequalsm1min(η6/A111/2,η7/2)\eta_{8}\colonequals m^{-1}\min(\eta_{6}/A_{11}^{1/2},\eta_{7}/2).

Our remaining analysis breaks into two cases: Δ(𝒗~)𝒗(4/η8)2\lVert\Delta(\tilde{\bm{v}})\bm{v}\rVert\geq(4/\eta_{8})^{2} and Δ(𝒗~)𝒗<(4/η8)2\lVert\Delta(\tilde{\bm{v}})\bm{v}\rVert<(4/\eta_{8})^{2}.

Suppose first that Δ(𝒗~)𝒗<(4/η8)2\lVert\Delta(\tilde{\bm{v}})\bm{v}\rVert<(4/\eta_{8})^{2}. Consider a u~\tilde{u}\in\mathbb{R} satisfying (8.14). By (8.15), the right-hand side of (8.24) is kλm(λ2𝒗)k=𝒗m/2(λ𝒗1/2)m2k\ll_{k}\lambda^{m}(\lambda^{2}\lVert\bm{v}\rVert)^{-k}=\lVert\bm{v}\rVert^{-m/2}(\lambda\lVert\bm{v}\rVert^{1/2})^{m-2k}. Choosing k=m/2+1k=\lceil m/2+1\rceil and inserting (8.24) into (8.18), we then get

𝒥u,𝒗(ϕ)𝒥0,u,𝒗𝒗1/2/2m4𝒗1/2d×λ𝒥1,u,𝒗,λm𝒗m/2,\mathcal{J}_{u,\bm{v}}(\phi)-\mathcal{J}_{0,u,\bm{v}}-\int_{\lVert\bm{v}\rVert^{-1/2}/2m}^{4\lVert\bm{v}\rVert^{-1/2}}d^{\times}{\lambda}\,\mathcal{J}_{1,u,\bm{v},\lambda}\ll_{m}\lVert\bm{v}\rVert^{-m/2},

since 4𝒗1/2d×λ(λ𝒗1/2)m2k=4d×aam2k1\int_{4\lVert\bm{v}\rVert^{-1/2}}^{\infty}d^{\times}{\lambda}\,(\lambda\lVert\bm{v}\rVert^{1/2})^{m-2k}=\int_{4}^{\infty}d^{\times}{a}\,a^{m-2k}\ll 1 (via the substitution a=λ𝒗1/2a=\lambda\lVert\bm{v}\rVert^{1/2}). But if we simply take absolute values in 𝒥0,u,𝒗\mathcal{J}_{0,u,\bm{v}} (see (8.16)) and in 𝒥1,u,𝒗,λ\mathcal{J}_{1,u,\bm{v},\lambda} (see (8.17)), and then apply (8.15), we get 𝒥0,u,𝒗𝒗m/2\mathcal{J}_{0,u,\bm{v}}\ll\lVert\bm{v}\rVert^{-m/2} and 𝒥1,u,𝒗,λ𝒗m/2\mathcal{J}_{1,u,\bm{v},\lambda}\ll\lVert\bm{v}\rVert^{-m/2} for λ[𝒗1/2/2m,4𝒗1/2]\lambda\in[\lVert\bm{v}\rVert^{-1/2}/2m,4\lVert\bm{v}\rVert^{-1/2}]. Integrating over λ\lambda in the previous display, we conclude that 𝒥u,𝒗(ϕ)m𝒗m/2\mathcal{J}_{u,\bm{v}}(\phi)\ll_{m}\lVert\bm{v}\rVert^{-m/2} (under (8.14)). Hence by (8.13) (after writing u=𝒗u~u=\lVert\bm{v}\rVert\tilde{u}) we have

𝒥r,𝒗(q,ϕ)b(1+𝒗)b+𝒗1/3A9|u~|3A9𝑑u~|q(r𝒗u~)|𝒗m/2.\mathscr{J}_{r,\bm{v}}(q,\phi)\ll_{b}(1+\lVert\bm{v}\rVert)^{-b}+\lVert\bm{v}\rVert\int_{1/3A_{9}\leq\lvert\tilde{u}\rvert\leq 3A_{9}}d\tilde{u}\,\lvert q(r\lVert\bm{v}\rVert\tilde{u})\rvert\cdot\lVert\bm{v}\rVert^{-m/2}.

But q(r𝒗u~)b(1+r𝒗)bq(r\lVert\bm{v}\rVert\tilde{u})\ll_{b}(1+r\lVert\bm{v}\rVert)^{-b} (under (8.14)). Thus 𝒥r,𝒗(q,ϕ)b𝒗1m/2(1+r𝒗)b\mathscr{J}_{r,\bm{v}}(q,\phi)\ll_{b}\lVert\bm{v}\rVert^{1-m/2}(1+r\lVert\bm{v}\rVert)^{-b}, which suffices for Proposition 8.4 (since 𝒗1\lVert\bm{v}\rVert\geq 1 and Δ(𝒗/M)𝒗Δ(𝒗~)𝒗1\lVert\Delta(\bm{v}/M)\bm{v}\rVert\leq\lVert\Delta(\tilde{\bm{v}})\bm{v}\rVert\ll 1).

For the rest of the proof, assume Δ(𝒗~)𝒗(4/η8)2\lVert\Delta(\tilde{\bm{v}})\bm{v}\rVert\geq(4/\eta_{8})^{2}, so that 4𝒗1/2η8|Δ(𝒗~)|1/24\lVert\bm{v}\rVert^{-1/2}\leq\eta_{8}\lvert\Delta(\tilde{\bm{v}})\rvert^{1/2}. We will integrate by parts over u~\tilde{u} in the integrals

𝒥0\colonequals𝑑u~q(r𝒗u~)w1(u~)𝒥0,u,𝒗,𝒥1,λ\colonequals𝑑u~q(r𝒗u~)w1(u~)𝒥1,u,𝒗,λ\mathscr{J}_{0}\colonequals\int_{\mathbb{R}}d\tilde{u}\,q(r\lVert\bm{v}\rVert\tilde{u})w_{1}(\tilde{u})\mathcal{J}_{0,u,\bm{v}},\quad\mathscr{J}_{1,\lambda}\colonequals\int_{\mathbb{R}}d\tilde{u}\,q(r\lVert\bm{v}\rVert\tilde{u})w_{1}(\tilde{u})\mathcal{J}_{1,u,\bm{v},\lambda}

for λη8|Δ(𝒗~)|1/2\lambda\leq\eta_{8}\lvert\Delta(\tilde{\bm{v}})\rvert^{1/2}. It is crucial to work in terms of 𝒛\bm{z} rather than 𝒙=𝒛/|u~|1/2\bm{x}=\bm{z}/\lvert\tilde{u}\rvert^{1/2}. Before proceeding, note that inserting (8.18) into (8.13) (and writing u=𝒗u~u=\lVert\bm{v}\rVert\tilde{u}) gives

(8.26) 𝒥r,𝒗𝒗𝒥0𝒗𝒗1/2/2m2A10d×λ𝒥1,λb(1+𝒗)b.\mathscr{J}_{r,\bm{v}}-\lVert\bm{v}\rVert\cdot\mathscr{J}_{0}-\lVert\bm{v}\rVert\cdot\int_{\lVert\bm{v}\rVert^{-1/2}/2m}^{2A_{10}}d^{\times}{\lambda}\,\mathscr{J}_{1,\lambda}\ll_{b}(1+\lVert\bm{v}\rVert)^{-b}.

Let 𝒛𝒲\bm{z}\in\mathscr{W}, and suppose ψ1(𝒛)mη8|Δ(𝒗~)|1/2\lVert\nabla{\psi_{1}}(\bm{z})\rVert\leq m\eta_{8}\lvert\Delta(\tilde{\bm{v}})\rvert^{1/2}. By the definition of η8\eta_{8}, we then have ψ1(𝒛)η6\lVert\nabla{\psi_{1}}(\bm{z})\rVert\leq\eta_{6} and ψ1(𝒛)η7|Δ(𝒗~)|1/2/2\lVert\nabla{\psi_{1}}(\bm{z})\rVert\leq\eta_{7}\lvert\Delta(\tilde{\bm{v}})\rvert^{1/2}/2, so by (8.25), we have

(8.27) |ψ1(𝒛)|3η7|Δ(𝒗~)|/4.\lvert\psi_{1}(\bm{z})\rvert\geq 3\eta_{7}\lvert\Delta(\tilde{\bm{v}})\rvert/4.

For convenience, let 𝒲(λ)\colonequals{𝒛𝒲:ψ1(𝒛)mλ}\mathscr{W}(\lambda)\colonequals\{\bm{z}\in\mathscr{W}:\lVert\nabla{\psi_{1}}(\bm{z})\rVert\leq m\lambda\} for each λ>0\lambda>0.

We first bound 𝒥0\mathscr{J}_{0}. For each u~Suppw1\tilde{u}\in\operatorname{Supp}{w_{1}}, the condition (8.14) holds, so if we plug in the definition of 𝒥0,u,𝒗\mathcal{J}_{0,u,\bm{v}} (see (8.16)), and then switch the order of uu, 𝒛\bm{z}, we may rewrite 𝒥0\mathscr{J}_{0} as

𝒲(𝒗1/2)𝑑𝒛ν0(ψ1(𝒛)𝒗1/2)du~|u~|m/2q(r𝒗u~)w1(u~)ϕ(𝒛|u~|1/2)e(𝒗ψ1(𝒛)|u~|1/2).\int_{\mathscr{W}(\lVert\bm{v}\rVert^{-1/2})}d\bm{z}\,\nu_{0}{\left(\frac{\nabla{\psi_{1}}(\bm{z})}{\lVert\bm{v}\rVert^{-1/2}}\right)}\int_{\mathbb{R}}\frac{d\tilde{u}}{\lvert\tilde{u}\rvert^{m/2}}\,q(r\lVert\bm{v}\rVert\tilde{u})w_{1}(\tilde{u})\phi{\left(\frac{\bm{z}}{\lvert\tilde{u}\rvert^{1/2}}\right)}e{\left(\frac{\lVert\bm{v}\rVert\psi_{1}(\bm{z})}{\lvert\tilde{u}\rvert^{1/2}}\right)}.

Here ψ1\psi_{1} is independent of |u~|\lvert\tilde{u}\rvert (when sgn(u)\operatorname{sgn}(u) is fixed). Therefore, for each 𝒛𝒲(𝒗1/2)\bm{z}\in\mathscr{W}(\lVert\bm{v}\rVert^{-1/2}), the inner integral (over u~\tilde{u}) is k(1+r𝒗)kΔ(𝒗~)𝒗k\ll_{k}(1+r\lVert\bm{v}\rVert)^{-k}\lVert\Delta(\tilde{\bm{v}})\bm{v}\rVert^{-k} by (8.27) and Lemma 8.6, because |u~|1\lvert\tilde{u}\rvert\asymp 1 for u~Suppw1\tilde{u}\in\operatorname{Supp}{w_{1}} and we have u~lq(r𝒗u~)=(r𝒗)lq(l)(r𝒗u~)l,k(1+r𝒗)k\partial_{\tilde{u}}^{l}q(r\lVert\bm{v}\rVert\tilde{u})=(r\lVert\bm{v}\rVert)^{l}q^{(l)}(r\lVert\bm{v}\rVert\tilde{u})\ll_{l,k}(1+r\lVert\bm{v}\rVert)^{-k} for all l,k0l,k\geq 0 (since qq is Schwartz). Applying this inner integral estimate for each 𝒛𝒲(𝒗1/2)\bm{z}\in\mathscr{W}(\lVert\bm{v}\rVert^{-1/2}), and then using (8.15), we get 𝒥0k𝒗m/2(1+r𝒗)kΔ(𝒗~)𝒗k\mathscr{J}_{0}\ll_{k}\lVert\bm{v}\rVert^{-m/2}(1+r\lVert\bm{v}\rVert)^{-k}\lVert\Delta(\tilde{\bm{v}})\bm{v}\rVert^{-k}.

One can similarly prove 𝒥1,λk𝒗m/2(1+r𝒗)k\mathscr{J}_{1,\lambda}\ll_{k}\lVert\bm{v}\rVert^{-m/2}(1+r\lVert\bm{v}\rVert)^{-k} for λ[𝒗1/2/2m,4𝒗1/2]\lambda\in[\lVert\bm{v}\rVert^{-1/2}/2m,4\lVert\bm{v}\rVert^{-1/2}].

Now suppose 4𝒗1/2λmin(η8|Δ(𝒗~)|1/2,2A10)4\lVert\bm{v}\rVert^{-1/2}\leq\lambda\leq\min(\eta_{8}\lvert\Delta(\tilde{\bm{v}})\rvert^{1/2},2A_{10}). Let 𝒥1,j,λ\mathscr{J}_{1,j,\lambda} denote 𝒥1,λ\mathscr{J}_{1,\lambda} with 𝒥1,j,u,𝒗,λ\mathcal{J}_{1,j,u,\bm{v},\lambda} in place of 𝒥1,u,𝒗,λ\mathcal{J}_{1,u,\bm{v},\lambda}. Then 𝒥1,λ=1jm𝒥1,j,λ\mathscr{J}_{1,\lambda}=\sum_{1\leq j\leq m}\mathscr{J}_{1,j,\lambda}. By (8.22), we have (for all integers l0l\geq 0)

𝒥1,j,λ=𝒛𝒲(λ)𝑑𝒛du~|u~|m/2q(r𝒗u~)w1(u~)w2,j,l,u,𝒗,λ(𝒛)e(𝒗ψ1(𝒛)/|u~|1/2)(𝒗zjψ1(𝒛)/|u~|1/2)l.\mathscr{J}_{1,j,\lambda}=\int_{\bm{z}\in\mathscr{W}(\lambda)}d\bm{z}\int_{\mathbb{R}}\frac{d\tilde{u}}{\lvert\tilde{u}\rvert^{m/2}}\,q(r\lVert\bm{v}\rVert\tilde{u})w_{1}(\tilde{u})\frac{w_{2,j,l,u,\bm{v},\lambda}(\bm{z})e(\lVert\bm{v}\rVert\psi_{1}(\bm{z})/\lvert\tilde{u}\rvert^{1/2})}{(\lVert\bm{v}\rVert\partial_{z_{j}}{\psi_{1}}(\bm{z})/\lvert\tilde{u}\rvert^{1/2})^{l}}.

Since ψ1\psi_{1} is independent of |u~|\lvert\tilde{u}\rvert, and we have 𝒛1\lVert\bm{z}\rVert\ll 1 and |u~|1\lvert\tilde{u}\rvert\asymp 1 for all (𝒛,u~)𝒲×Suppw1(\bm{z},\tilde{u})\in\mathscr{W}\times\operatorname{Supp}{w_{1}}, we find by (8.20), (8.23), (8.27), and Lemma 8.6 (applied to the inner integral over u~\tilde{u}, for each 𝒛𝒲(λ)\bm{z}\in\mathscr{W}(\lambda) for which there exists uu\in\mathbb{R} with w1(u~)w2,j,l,u,𝒗,λ(𝒛)0w_{1}(\tilde{u})w_{2,j,l,u,\bm{v},\lambda}(\bm{z})\neq 0) that

𝒥1,j,λk,l𝒛𝒲(λ)𝑑𝒛(1+r𝒗)kΔ(𝒗~)𝒗k(λ2𝒗)lλm(1+r𝒗)kΔ(𝒗~)𝒗k(λ2𝒗)l\mathscr{J}_{1,j,\lambda}\ll_{k,l}\int_{\bm{z}\in\mathscr{W}(\lambda)}d\bm{z}\,\frac{(1+r\lVert\bm{v}\rVert)^{-k}\lVert\Delta(\tilde{\bm{v}})\bm{v}\rVert^{-k}}{(\lambda^{2}\lVert\bm{v}\rVert)^{l}}\ll\frac{\lambda^{m}(1+r\lVert\bm{v}\rVert)^{-k}\lVert\Delta(\tilde{\bm{v}})\bm{v}\rVert^{-k}}{(\lambda^{2}\lVert\bm{v}\rVert)^{l}}

for all integers k,l0k,l\geq 0, where in the final step we use (8.15).

Finally, suppose η8|Δ(𝒗~)|1/2λ2A10\eta_{8}\lvert\Delta(\tilde{\bm{v}})\rvert^{1/2}\leq\lambda\leq 2A_{10}. Then (8.24) and (8.15) directly give (for k,l0k,l\geq 0)

𝒥1,λl𝑑u~|q(r𝒗u~)w1(u~)|λm(λ2𝒗)lk(1+r𝒗)k𝒗m/2(λ𝒗1/2)2lm.\mathscr{J}_{1,\lambda}\ll_{l}\int_{\mathbb{R}}d\tilde{u}\,\frac{\lvert q(r\lVert\bm{v}\rVert\tilde{u})w_{1}(\tilde{u})\rvert\cdot\lambda^{m}}{(\lambda^{2}\lVert\bm{v}\rVert)^{l}}\ll_{k}\frac{(1+r\lVert\bm{v}\rVert)^{-k}\lVert\bm{v}\rVert^{-m/2}}{(\lambda\lVert\bm{v}\rVert^{1/2})^{2l-m}}.

Inserting our work from the last four paragraphs (ignoring the last one if η8|Δ(𝒗~)|1/2>2A10\eta_{8}\lvert\Delta(\tilde{\bm{v}})\rvert^{1/2}>2A_{10}) into the left-hand side of (8.26), and applying (8.26) with b=m/2+2kb=\lceil m/2+2k\rceil, we get the bound

𝒥r,𝒗𝒗k,l1+λ4𝒗1/2d×λ(λ𝒗1/2)m2l𝒗m/2(1+r𝒗)kΔ(𝒗~)𝒗k+λη8|Δ(𝒗~)|1/2d×λ(λ𝒗1/2)m2l+2k𝒗m/2(1+r𝒗)kΔ(𝒗~)𝒗k.\frac{\mathscr{J}_{r,\bm{v}}}{\lVert\bm{v}\rVert}\ll_{k,l}\frac{1+\int_{\lambda\geq 4\lVert\bm{v}\rVert^{-1/2}}d^{\times}{\lambda}\,(\lambda\lVert\bm{v}\rVert^{1/2})^{m-2l}}{\lVert\bm{v}\rVert^{m/2}(1+r\lVert\bm{v}\rVert)^{k}\lVert\Delta(\tilde{\bm{v}})\bm{v}\rVert^{k}}+\frac{\int_{\lambda\geq\eta_{8}\lvert\Delta(\tilde{\bm{v}})\rvert^{1/2}}d^{\times}{\lambda}\,(\lambda\lVert\bm{v}\rVert^{1/2})^{m-2l+2k}}{\lVert\bm{v}\rVert^{m/2}(1+r\lVert\bm{v}\rVert)^{k}\lVert\Delta(\tilde{\bm{v}})\bm{v}\rVert^{k}}.

Taking l=m/2+k+1l=\lceil m/2+k+1\rceil, and evaluating both integrals over λ\lambda (the first being 1\ll 1 since 4𝒗1/2𝒗1/2=414\lVert\bm{v}\rVert^{-1/2}\cdot\lVert\bm{v}\rVert^{1/2}=4\gg 1, and the second being 1\ll 1 since η8Δ(𝒗~)𝒗1/241\eta_{8}\lVert\Delta(\tilde{\bm{v}})\bm{v}\rVert^{1/2}\geq 4\gg 1), we get 𝒥r,𝒗k𝒗1m/2(1+r𝒗)kΔ(𝒗~)𝒗k\mathscr{J}_{r,\bm{v}}\ll_{k}\lVert\bm{v}\rVert^{1-m/2}(1+r\lVert\bm{v}\rVert)^{-k}\lVert\Delta(\tilde{\bm{v}})\bm{v}\rVert^{-k}, which suffices for Proposition 8.4. ∎

9. New bounds on bad exponential sums

We first provide some new, general, vanishing and boundedness criteria for S𝒄(pl)S_{\bm{c}}(p^{l}), which when combined with classical estimates from [hooley1986HasseWeil, heath1998circle] will (under Conjecture 1.5) allow us to break a critical ϵ\epsilon-barrier behind (2.17). Recall S𝒄(n)S_{\bm{c}}(n), S𝒄(n)S^{\natural}_{\bm{c}}(n) from (2.8), (2.9).

Lemma 9.1.

Let 𝐜𝒮1\bm{c}\in\mathcal{S}_{1}, and let pp be a prime.

  1. (1)

    If pgcd(Δ(𝒄),c1Δ(𝒄),,cmΔ(𝒄))p\nmid\gcd(\Delta(\bm{c}),\partial_{c_{1}}{\Delta}(\bm{c}),\dots,\partial_{c_{m}}{\Delta}(\bm{c})), then S𝒄(p)m1S^{\natural}_{\bm{c}}(p)\ll_{m}1.

  2. (2)

    If vp(Δ(𝒄))1v_{p}(\Delta(\bm{c}))\leq 1, then S𝒄(p2)F1S^{\natural}_{\bm{c}}(p^{2})\ll_{F}1.

  3. (3)

    We have S𝒄(pl)=0S_{\bm{c}}(p^{l})=0 for all integers l2+vp(Δ(𝒄))l\geq 2+v_{p}(\Delta(\bm{c})).

Proof.

(1): If p6p\mid 6, then |S𝒄(p)|p1+m\lvert S_{\bm{c}}(p)\rvert\leq p^{1+m} trivially, so |S𝒄(p)|p(1+m)/2m1\lvert S^{\natural}_{\bm{c}}(p)\rvert\leq p^{(1+m)/2}\ll_{m}1. Now suppose p6gcd(Δ(𝒄),c1Δ(𝒄),,cmΔ(𝒄))p\nmid 6\gcd(\Delta(\bm{c}),\partial_{c_{1}}{\Delta}(\bm{c}),\dots,\partial_{c_{m}}{\Delta}(\bm{c})). Then by (2.1), we have

pdisc(F),p6,pgcd(disc(F,𝒄),c1disc(F,𝒄),,cmdisc(F,𝒄)).p\nmid\operatorname{disc}(F),\quad p\nmid 6,\quad p\nmid\gcd(\operatorname{disc}(F,\bm{c}),\partial_{c_{1}}{\operatorname{disc}(F,\bm{c})},\dots,\partial_{c_{m}}{\operatorname{disc}(F,\bm{c})}).

Therefore, by (2.12) and [wang2023dichotomous]*Theorem 1.1, and Proposition 2.7(2)\Rightarrow(1) with (n,r,d)=(m1,2,3)(n,r,d)=(m-1,2,3), we have |E𝒄(p)|729m\lvert E^{\natural}_{\bm{c}}(p)\rvert\leq 72\cdot 9^{m}. Yet EF(p)m1E^{\natural}_{F}(p)\ll_{m}1, by (2.12) and the Weil conjectures (since pdisc(F)p\nmid\operatorname{disc}(F)). Plugging these two estimates into (2.13), we get S𝒄(p)m1S^{\natural}_{\bm{c}}(p)\ll_{m}1.

(If m=6m=6 and FF is diagonal, one can also improve (1) to a “codimension-three” statement, by using [wang2023dichotomous]*Theorem 1.3 in place of [wang2023dichotomous]*Theorem 1.1.)

We now turn to (2)–(3). For any vector 𝒖m\bm{u}\in\mathbb{Z}^{m}, let vp(𝒖)\colonequalsvp(gcd(u1,,um))v_{p}(\bm{u})\colonequals v_{p}(\gcd(u_{1},\dots,u_{m})). Write 𝒄=pg𝒄~\bm{c}=p^{g}\tilde{\bm{c}}, where g=vp(𝒄)<g=v_{p}(\bm{c})<\infty and 𝒄~𝒮1\tilde{\bm{c}}\in\mathcal{S}_{1}. For integers u,d0u,d\geq 0, let

(d)(𝒄;p,u)\colonequalsλ:pλ{𝒙m:p𝒙,puF(𝒙),pu𝒄~𝒙,pdF(𝒙)λ𝒄}.\mathscr{B}^{(d)}(\bm{c};p,u)\colonequals\bigcup_{\lambda\in\mathbb{Z}:\,p\nmid\lambda}\{\bm{x}\in\mathbb{Z}^{m}:p\nmid\bm{x},\;p^{u}\mid F(\bm{x}),\;p^{u}\mid\tilde{\bm{c}}\cdot\bm{x},\;p^{d}\mid\nabla{F}(\bm{x})-\lambda\bm{c}\}.

For any u,d0u,d\geq 0 and 𝒙(d)(𝒄;p,u)\bm{x}\in\mathscr{B}^{(d)}(\bm{c};p,u), Corollary 2.3 (with 𝒄~\tilde{\bm{c}} in place of 𝒄\bm{c}) implies

(9.1) gcd(pu,p2d)Δ(𝒄~).\gcd(p^{u},p^{2d})\mid\Delta(\tilde{\bm{c}}).

For any set AmA\subseteq\mathbb{Z}^{m} that can be written as a finite union of residue classes m\mathcal{R}\subseteq\mathbb{Z}^{m}, let μ(A)\mu(A) be the density of AA in m\mathbb{Z}^{m}. Now let l2+2gl\geq 2+2g be an integer, and let d=l/21+gd=\lfloor l/2\rfloor\geq 1+g; then by [wang2023_isolating_special_solutions]*Proposition 7.4, we have

(9.2) plmφ(pl)S𝒄(pl)=p2l+gμ((d)(𝒄;p,l))p2l2+gμ((d)(𝒄;p,l1)),p^{-lm}\varphi(p^{l})S^{\prime}_{\bm{c}}(p^{l})=p^{2l+g}\mu(\mathscr{B}^{(d)}(\bm{c};p,l))-p^{2l-2+g}\mu(\mathscr{B}^{(d)}(\bm{c};p,l-1)),

where S𝒄(pl)S^{\prime}_{\bm{c}}(p^{l}) (the “restriction to p𝒙p\nmid\bm{x}” of S𝒄(pl)S_{\bm{c}}(p^{l})) is defined as in [wang2023_isolating_special_solutions]*(7.5).

(2): Suppose vp(Δ(𝒄))1v_{p}(\Delta(\bm{c}))\leq 1. Then p𝒄p\nmid\bm{c} (since degΔ2\deg{\Delta}\geq 2), so g=0g=0 and 𝒄~=𝒄\tilde{\bm{c}}=\bm{c}. In particular, S𝒄(p2)=S𝒄(p2)S_{\bm{c}}(p^{2})=S^{\prime}_{\bm{c}}(p^{2}) by [wang2023_isolating_special_solutions]*Lemma 7.2. By (9.2) (with l=2l=2 and d=l/2=1d=\lfloor l/2\rfloor=1), it remains to analyze (1)(𝒄;p,u)\mathscr{B}^{(1)}(\bm{c};p,u) for u{1,2}u\in\{1,2\}. By (9.1), we have (1)(𝒄;p,2)=\mathscr{B}^{(1)}(\bm{c};p,2)=\emptyset (since p2Δ(𝒄~)p^{2}\nmid\Delta(\tilde{\bm{c}})). We now analyze (1)(𝒄;p,1)\mathscr{B}^{(1)}(\bm{c};p,1).

For each 𝒂(1)(𝒄;p,1)\bm{a}\in\mathscr{B}^{(1)}(\bm{c};p,1), the variety F(𝒙)=𝒄𝒙=0F(\bm{x})=\bm{c}\cdot\bm{x}=0 in 𝔽pm1\mathbb{P}^{m-1}_{\mathbb{F}_{p}} is singular at the point [a1::am]m1(𝔽p)[a_{1}:\cdots:a_{m}]\in\mathbb{P}^{m-1}(\mathbb{F}_{p}). Since p𝒄p\nmid\bm{c}, this variety is isomorphic to a cubic hypersurface f(y1,,ym1)=0f(y_{1},\dots,y_{m-1})=0 in 𝔽pm2\mathbb{P}^{m-2}_{\mathbb{F}_{p}}, and has a singular locus of dimension 0\leq 0 if pdisc(F)p\nmid\operatorname{disc}(F) (by [wang2023dichotomous]*Theorem 2.3, due to Zak). So if p3disc(F)p\nmid 3\operatorname{disc}(F), then pmμ((1)(𝒄;p,1))(p1)2m1p^{m}\mu(\mathscr{B}^{(1)}(\bm{c};p,1))\leq(p-1)\cdot 2^{m-1} by Bézout’s theorem (applied to the system f=0\nabla{f}=0 in 𝔽pm2\mathbb{P}^{m-2}_{\mathbb{F}_{p}}; note that 3f=𝒚f3f=\bm{y}\cdot\nabla{f}).

But if p3disc(F)p\mid 3\operatorname{disc}(F), then pmμ((1)(𝒄;p,1))pm1p^{m}\mu(\mathscr{B}^{(1)}(\bm{c};p,1))\leq p^{m}-1 trivially. So in every case, pmμ((1)(𝒄;p,1))Fp1p^{m}\mu(\mathscr{B}^{(1)}(\bm{c};p,1))\ll_{F}p-1. Now by (2.9) and (9.2), we have

S𝒄(p2)=S𝒄(p2)p1+m=S𝒄(p2)p1+m=pm1φ(p2)(p40p2μ((1)(𝒄;p,1)))Fp(p1)φ(p2)=1.S^{\natural}_{\bm{c}}(p^{2})=\frac{S_{\bm{c}}(p^{2})}{p^{1+m}}=\frac{S^{\prime}_{\bm{c}}(p^{2})}{p^{1+m}}=\frac{p^{m-1}}{\varphi(p^{2})}\cdot\left(p^{4}\cdot 0-p^{2}\cdot\mu(\mathscr{B}^{(1)}(\bm{c};p,1))\right)\ll_{F}\frac{p(p-1)}{\varphi(p^{2})}=1.

(3): Take a counterexample (𝒄,l)𝒮1×0(\bm{c},l)\in\mathcal{S}_{1}\times\mathbb{Z}_{\geq 0} with ll minimal. Then l2+vp(Δ(𝒄))l\geq 2+v_{p}(\Delta(\bm{c})) and S𝒄(pl)0S_{\bm{c}}(p^{l})\neq 0. So l2+2gl\geq 2+2g, because degΔ2\deg{\Delta}\geq 2. Furthermore, d=l/2(l1)/2d=\lfloor l/2\rfloor\geq(l-1)/2. By (9.1), we have (d)(𝒄;p,u)=\mathscr{B}^{(d)}(\bm{c};p,u)=\emptyset for all ul1u\geq l-1, since 2dl12d\geq l-1 and pl1Δ(𝒄)p^{l-1}\nmid\Delta(\bm{c}). So (9.2) gives S𝒄(pl)=0S^{\prime}_{\bm{c}}(p^{l})=0, whence S𝒄(pl)S𝒄(pl)S_{\bm{c}}(p^{l})\neq S^{\prime}_{\bm{c}}(p^{l}). If g=0g=0, this contradicts [wang2023_isolating_special_solutions]*Lemma 7.2. If g=1g=1, then l4l\geq 4, so S𝒄(pl)S𝒄(pl)=0S_{\bm{c}}(p^{l})-S^{\prime}_{\bm{c}}(p^{l})=0 by [wang2023_isolating_special_solutions]*Lemma 7.2(2); again, a contradiction. Now suppose g2g\geq 2; then p2𝒄p^{2}\mid\bm{c} and l32+vp(Δ(𝒄/p2))l-3\geq 2+v_{p}(\Delta(\bm{c}/p^{2})) (since degΔ3/2\deg{\Delta}\geq 3/2), so S𝒄/p2(pl3)=0S_{\bm{c}/p^{2}}(p^{l-3})=0 by the minimality hypothesis. But [wang2023_isolating_special_solutions]*Lemma 7.2(2) then gives S𝒄(pl)S𝒄(pl)=0S_{\bm{c}}(p^{l})-S^{\prime}_{\bm{c}}(p^{l})=0; another contradiction. Therefore, no counterexample (𝒄,l)𝒮1×0(\bm{c},l)\in\mathcal{S}_{1}\times\mathbb{Z}_{\geq 0} to (3) in fact exists.

For an alternative, more algorithmic and computational approach to (2)–(3) (at least when FF is diagonal), see [wang2022thesis]*§7.2 and [wang2021_HLH_vs_RMT]*Appendix D. ∎

Remark 9.2.

If FF were quadratic (rather than cubic), then Lemma 9.1(1) would be false whenever 2m2\mid m. See [wang2022thesis]*Remark 7.2.4 or [wang2023dichotomous]*sentence after Theorem 1.1.

Let 𝒩(t)\colonequals{n1:pnvp(n)t}\mathcal{N}_{\leq}(t)\colonequals\{n\geq 1:p\mid n\Rightarrow v_{p}(n)\leq t\} and 𝒩(t)\colonequals{n1:pnvp(n)t}\mathcal{N}_{\geq}(t)\colonequals\{n\geq 1:p\mid n\Rightarrow v_{p}(n)\geq t\} for each integer t1t\geq 1. For any integers N,t1N,t\geq 1, we have (see e.g. [bateman1958theorem])

(9.3) |{Nn<2N:n𝒩(t)}|tN1/t.\lvert\{N\leq n<2N:n\in\mathcal{N}_{\geq}(t)\}\rvert\ll_{t}N^{1/t}.

Recall 𝒩𝒄\mathcal{N}_{\bm{c}} from (2.2). Lemma 9.1(1) has a useful unconditional consequence:

Proposition 9.3.

Let A1A\in\mathbb{R}_{\geq 1}. Uniformly over reals Z,N>0Z,N>0 with NZ3N\leq Z^{3}, we have

(9.4) 𝒄𝒮1[Z,Z]m(n𝒩𝒄𝒩(1)[N,2N)n1/2|S𝒄(n)|)2AF,A,ϵZmmin(N,Z)1+ϵ.\sum_{\bm{c}\in\mathcal{S}_{1}\cap[-Z,Z]^{m}}\biggl{(}\,\sum_{n\in\mathcal{N}_{\bm{c}}\cap\mathcal{N}_{\leq}(1)\cap[N,2N)}n^{-1/2}\lvert S^{\natural}_{\bm{c}}(n)\rvert\biggr{)}^{\!2A}\ll_{F,A,\epsilon}Z^{m}\min(N,Z)^{-1+\epsilon}.
Proof.

The case N<1N<1 is trivial, so assume N1N\geq 1. Let 𝒟={1,2,4,8,}\mathcal{D}=\{1,2,4,8,\ldots\}. Let

(9.5) 𝒩𝒄,0\colonequals{n𝒩(1):pnpgcd(Δ(𝒄),c1Δ(𝒄),,cmΔ(𝒄))}\mathcal{N}_{\bm{c},0}\colonequals\{n\in\mathcal{N}_{\leq}(1):p\mid n\Rightarrow p\mid\gcd(\Delta(\bm{c}),\partial_{c_{1}}{\Delta}(\bm{c}),\dots,\partial_{c_{m}}{\Delta}(\bm{c}))\}

for each 𝒄m\bm{c}\in\mathbb{Z}^{m}. For each Q𝒟Q\in\mathcal{D}, let 𝒮3(Q)\mathcal{S}_{3}(Q) be the set of tuples 𝒄𝒮1\bm{c}\in\mathcal{S}_{1} for which 𝒩𝒄,0[Q,2Q)\mathcal{N}_{\bm{c},0}\cap[Q,2Q)\neq\emptyset. Note that 1𝒩𝒄,01\in\mathcal{N}_{\bm{c},0}, so 𝒮3(1)=𝒮1\mathcal{S}_{3}(1)=\mathcal{S}_{1}.

It is known (e.g. by [heath1983cubic]*Lemma 11) that S𝒄(p)mp1/2S^{\natural}_{\bm{c}}(p)\ll_{m}p^{1/2} for all 𝒄m\bm{c}\in\mathbb{Z}^{m} and primes pp. So for all 𝒄𝒮1\bm{c}\in\mathcal{S}_{1} and n𝒩(1)n\in\mathcal{N}_{\leq}(1), Lemma 9.1(1) implies

S𝒄(n)m,ϵnϵmaxqn:q𝒩𝒄,0q1/2nϵ(max{Q𝒟[1,n]:𝒄𝒮3(Q)})1/2.S^{\natural}_{\bm{c}}(n)\ll_{m,\epsilon}n^{\epsilon}\cdot\max_{q\mid n:\,q\in\mathcal{N}_{\bm{c},0}}{q^{1/2}}\ll n^{\epsilon}\cdot\left(\max{\{Q\in\mathcal{D}\cap[1,n]:\bm{c}\in\mathcal{S}_{3}(Q)\}}\right)^{1/2}.

Therefore, the left-hand side of (9.4) is at most Om,A,ϵ(1)O_{m,A,\epsilon}(1) times the quantity

(9.6) 𝒄𝒮1:𝒄Z|𝒩𝒄[N,2N)|2AQ𝒟[1,2N):𝒄𝒮3(Q)NϵQANA=Q𝒟[1,2N)NϵQANA𝒄𝒮3(Q):𝒄Z|𝒩𝒄[N,2N)|2A.\sum_{\begin{subarray}{c}\bm{c}\in\mathcal{S}_{1}:\\ \lVert\bm{c}\rVert\leq Z\end{subarray}}\lvert\mathcal{N}_{\bm{c}}\cap[N,2N)\rvert^{2A}\sum_{\begin{subarray}{c}Q\in\mathcal{D}\cap[1,2N):\\ \bm{c}\in\mathcal{S}_{3}(Q)\end{subarray}}\frac{N^{\epsilon}Q^{A}}{N^{A}}=\sum_{Q\in\mathcal{D}\cap[1,2N)}\frac{N^{\epsilon}Q^{A}}{N^{A}}\sum_{\begin{subarray}{c}\bm{c}\in\mathcal{S}_{3}(Q):\\ \lVert\bm{c}\rVert\leq Z\end{subarray}}\lvert\mathcal{N}_{\bm{c}}\cap[N,2N)\rvert^{2A}.

Let ϵ(0,1)\epsilon\in(0,1), and consider an individual Q𝒟Q\in\mathcal{D}. Lemma 6.17, and Hölder over 𝒄\bm{c}, give

(9.7) 𝒄𝒮3(Q)[Z,Z]m|𝒩𝒄[N,2N)|2AA,ϵ|𝒮3(Q)[Z,Z]m|1ϵ(ZmNϵ)ϵ.\sum_{\bm{c}\in\mathcal{S}_{3}(Q)\cap[-Z,Z]^{m}}\lvert\mathcal{N}_{\bm{c}}\cap[N,2N)\rvert^{2A}\ll_{A,\epsilon}\lvert\mathcal{S}_{3}(Q)\cap[-Z,Z]^{m}\rvert^{1-\epsilon}\cdot(Z^{m}N^{\epsilon})^{\epsilon}.

On the other hand, the scheme Δ=c1Δ==cmΔ=0\Delta=\partial_{c_{1}}{\Delta}=\dots=\partial_{c_{m}}{\Delta}=0 in 𝔸m\mathbb{A}^{m}_{\mathbb{Q}} has dimension m2\leq m-2 (since Δ=0\Delta=0 is generically smooth, due to the absolute irreducibility of Δ\Delta). So by a quantitative form of [ekedahl1991infinite]’s geometric sieve (see [bhargava2014geometric]*Theorem 3.3 for the case of prime moduli, which extends to square-free moduli as in [bhargava2021galois]*§5, Case III), we have

(9.8) |𝒮3(Q)[Z,Z]m|Δ,ϵZmQ1+ϵ+Zm1+ϵ;\lvert\mathcal{S}_{3}(Q)\cap[-Z,Z]^{m}\rvert\ll_{\Delta,\epsilon}Z^{m}Q^{-1+\epsilon}+Z^{m-1+\epsilon};

this bound follows from Lang–Weil if QZQ\leq Z (since n𝒩(1)[Q,2Q)Oϵ(Zmn2ϵ)ϵZmQ1+ϵ\sum_{n\in\mathcal{N}_{\leq}(1)\cap[Q,2Q)}O_{\epsilon}(\frac{Z^{m}}{n^{2-\epsilon}})\ll_{\epsilon}Z^{m}Q^{-1+\epsilon}; cf. [bhargava2014geometric]*(16)), and from elimination theory if Q>ZQ>Z (cf. [bhargava2014geometric]*(17) and [bhargava2021galois]*the three paragraphs after Proposition 33).

Upon inserting (9.8) into (9.7), we find that the right-hand side of (9.6) is

A,ϵQ𝒟[1,2N)NϵQANAZmNϵ2(Q(1ϵ)2+Z(1ϵ)2)A,ϵN2ϵZm(N(1ϵ)2+Z(1ϵ)2),\ll_{A,\epsilon}\sum_{Q\in\mathcal{D}\cap[1,2N)}\frac{N^{\epsilon}Q^{A}}{N^{A}}\cdot Z^{m}N^{\epsilon^{2}}(Q^{-(1-\epsilon)^{2}}+Z^{-(1-\epsilon)^{2}})\ll_{A,\epsilon}N^{2\epsilon}\cdot Z^{m}(N^{-(1-\epsilon)^{2}}+Z^{-(1-\epsilon)^{2}}),

since A1A\geq 1. Since Nϵmin(N,Z)3ϵN^{\epsilon}\leq\min(N,Z)^{3\epsilon}, the bound (9.4) follows immediately. ∎

We now build on Conjecture 1.5.

Conjecture 9.4 (SFSCq,3).

There exists a real η9=η9(Δ)>0\eta_{9}=\eta_{9}(\Delta)>0 such that

(9.9) #{𝒄[Z,Z]m:q𝒩(2)[Q,2Q) with qΔ(𝒄)}ΔZmQη9\#\{\bm{c}\in[-Z,Z]^{m}:\exists\;\textnormal{$q\in\mathcal{N}_{\geq}(2)\cap[Q,2Q)$ with $q\mid\Delta(\bm{c})$}\}\ll_{\Delta}Z^{m}Q^{-\eta_{9}}

holds uniformly over reals Z,Q1Z,Q\geq 1 with QZ3Q\leq Z^{3}.

Proposition 9.5.

Assume Conjecture 1.5. Then Conjecture 9.4 holds.

Proof.

Suppose Z,Q1Z,Q\geq 1 are reals with QZ3Q\leq Z^{3}.

Case 1: QZQ\leq Z. Recall the notation N(P;q)N(P;q) from §4. Let δ=(degΔ)1\delta=(\deg{\Delta})^{-1} and ϵ(0,δ)\epsilon\in(0,\delta). The left-hand side of (9.9) is at most

(9.10) q𝒩(2)[Q,2Q)N(Δ;q)O(Zmqm)ϵZmQδ/2ϵq𝒩(2)qδ/2ϵN(Δ;q)qm.\sum_{q\in\mathcal{N}_{\geq}(2)\cap[Q,2Q)}N(\Delta;q)\cdot O{\left(\frac{Z^{m}}{q^{m}}\right)}\ll_{\epsilon}\frac{Z^{m}}{Q^{\delta/2-\epsilon}}\sum_{q\in\mathcal{N}_{\geq}(2)}q^{\delta/2-\epsilon}\cdot\frac{N(\Delta;q)}{q^{m}}.

Let pp be a prime and l2l\geq 2 an integer. By (4.3), we have N(Δ;pl)Δpl(mδ)N(\Delta;p^{l})\ll_{\Delta}p^{l(m-\delta)}. On the other hand, by Hensel’s lemma (over the smooth locus of Δ=0\Delta=0) and Lang–Weil (over the singular locus of Δ=0\Delta=0), we have N(Δ;p2)Δp2m2N(\Delta;p^{2})\ll_{\Delta}p^{2m-2}, and thus N(Δ;pl)p(l2)mN(Δ;p2)Δplm2N(\Delta;p^{l})\leq p^{(l-2)m}N(\Delta;p^{2})\ll_{\Delta}p^{lm-2}. So by the Chinese remainder theorem, the right-hand side of (9.10) is

ZmQδ/2ϵp(1+l2p(δ/2ϵ)lOΔ(pmax(δl,2)))ZmQδ/2ϵ(1+l2OΔ(p1ϵl))\leq\frac{Z^{m}}{Q^{\delta/2-\epsilon}}\prod_{p}\biggl{(}1+\sum_{l\geq 2}p^{(\delta/2-\epsilon)l}\cdot O_{\Delta}(p^{-\max(\delta l,2)})\biggr{)}\leq\frac{Z^{m}}{Q^{\delta/2-\epsilon}}\biggl{(}1+\sum_{l\geq 2}O_{\Delta}(p^{-1-\epsilon l})\biggr{)}

(since (δ/2ϵ)lmax(δl,2)1ϵl(\delta/2-\epsilon)l-\max(\delta l,2)\leq-1-\epsilon l). Since ϵ>0\epsilon>0, it follows that the right-hand side of (9.10) is Δ,ϵZmQδ/2+ϵ\ll_{\Delta,\epsilon}Z^{m}Q^{-\delta/2+\epsilon}. Hence the left-hand side of (9.9) is Δ,ϵZmQδ/2+ϵ\ll_{\Delta,\epsilon}Z^{m}Q^{-\delta/2+\epsilon}.

Case 2: Q>ZQ>Z. Suppose q𝒩(2)[Q,2Q)q\in\mathcal{N}_{\geq}(2)\cap[Q,2Q). Let ss be the largest square divisor of qq; then sq2/3Q2/3>Z2/3s\geq q^{2/3}\geq Q^{2/3}>Z^{2/3}. The integer s1/2s^{1/2} thus lies in (Z1/4,2Q1/2)(Z^{1/4},2Q^{1/2}), and hence either has a prime factor p>Z1/4p>Z^{1/4} (so that p[P,2P)p\in[P,2P) for some P[Z1/4,Q1/2]P\in[Z^{1/4},Q^{1/2}]), or an integer factor d(Z1/4,Z1/2]d\in(Z^{1/4},Z^{1/2}] (so that d2[D,2D)d^{2}\in[D,2D) for some D[Z1/2,Z]D\in[Z^{1/2},Z]). So by Conjecture 1.5 (with P[Z1/4,Q1/2]P\in[Z^{1/4},Q^{1/2}]) and Case 1 (with D[Z1/2,Z]D\in[Z^{1/2},Z] in place of QQ), the left-hand side of (9.9) is

Δ,ϵZm(Z1/4)η0+Zm(Z1/2)δ/2+ϵ.\ll_{\Delta,\epsilon}Z^{m}(Z^{1/4})^{-\eta_{0}}+Z^{m}(Z^{1/2})^{-\delta/2+\epsilon}.

Since ZQ1/3Z\geq Q^{1/3}, it follows that (9.9) holds with η9=112min(η0,910(degΔ)1)\eta_{9}=\frac{1}{12}\min(\eta_{0},\frac{9}{10}(\deg{\Delta})^{-1}). ∎

For all W[𝒄]W\in\mathbb{Z}[\bm{c}] and Z,N,A>0Z,N,A\in\mathbb{R}_{>0}, let

(9.11) Σ14W,A(Z,N)\colonequals𝒄𝒮1[Z,Z]m:W(𝒄)0(n𝒩𝒄[N,2N)n1/2|S𝒄(n)|)A.\Sigma_{14}^{W,A}(Z,N)\colonequals\sum_{\bm{c}\in\mathcal{S}_{1}\cap[-Z,Z]^{m}:\,W(\bm{c})\neq 0}\biggl{(}\,\sum_{n\in\mathcal{N}_{\bm{c}}\cap[N,2N)}n^{-1/2}\lvert S^{\natural}_{\bm{c}}(n)\rvert\biggr{)}^{\!A}.

Close analogs of the moment Σ141,A(Z,N)\Sigma_{14}^{1,A}(Z,N) have been considered for A=1A=1 classically (e.g. in [heath1983cubic, hooley1986HasseWeil, heath1998circle]), and for A=2A=2 in [wang2023_large_sieve_diagonal_cubic_forms]*Propositions 4.12 and A.1.

Conjecture 9.6 (B2).

Let Z,N1Z,N\in\mathbb{R}_{\geq 1} with NZ3N\leq Z^{3}. Then Σ141,2(Z,N)F,ϵZm+ϵ\Sigma_{14}^{1,2}(Z,N)\ll_{F,\epsilon}Z^{m+\epsilon}.

Proposition 9.7.

Suppose FF is diagonal. Then Conjecture 9.6 holds. Also, for W=c1cmW=c_{1}\cdots c_{m} and 1NZ31\leq N\leq Z^{3}, we have Σ141,2(Z,N)Σ14W,2(Z,N)F,ϵZm1+ϵN1/3\Sigma_{14}^{1,2}(Z,N)-\Sigma_{14}^{W,2}(Z,N)\ll_{F,\epsilon}Z^{m-1+\epsilon}N^{1/3}.

Proposition 9.7 is essentially due to [wang2023_large_sieve_diagonal_cubic_forms]; but we need to go one step further.

Conjecture 9.8 (B3G).

Let 1A21\leq A\leq 2. There exists a nonzero polynomial W=WF,A[𝐜]W=W_{F,A}\in\mathbb{Z}[\bm{c}], and a real η10=η10(F,A)>0\eta_{10}=\eta_{10}(F,A)>0, such that if 1NZ31\leq N\leq Z^{3}, then Σ14W,A(Z,N)F,AZmNη10\Sigma_{14}^{W,A}(Z,N)\ll_{F,A}Z^{m}N^{-\eta_{10}}.

One can extend Conjecture 9.8 to A=2+δA=2+\delta, almost for free (see Proposition 10.3), but it is not clear to us what the limit is. Also, for 1A21\leq A\leq 2, one might hope for η10(F,A)A/2\eta_{10}(F,A)\approx A/2 to be admissible, by comparing with Proposition 9.3 or [sarnak1991bounds]*Conjecture 1.

Proposition 9.9.

Suppose FF is diagonal. Assume Conjecture 1.5. Then Conjecture 9.8 holds with W=c1cmW=c_{1}\cdots c_{m} (for all A[1,2]A\in[1,2]).

Remark 9.10.

If for each m{2,3,,m}m^{\prime}\in\{2,3,\dots,m\}, one assumed an analog of Conjecture 1.5 with mm^{\prime} in place of mm, then one could prove Proposition 9.9 with W=1W=1. See [wang2022thesis]*Conjecture 7.3.7 (B3) and Remark 7.3.13 and [wang2021_HLH_vs_RMT]*Lemma 7.43 for details.

We need a classical pointwise bound on S𝒄(n)S^{\natural}_{\bm{c}}(n). For integers c0c\neq 0, let sq(c)\colonequalsp2cpvp(c)\operatorname{sq}(c)\colonequals\prod_{p^{2}\mid c}p^{v_{p}(c)} and cub(c)\colonequalsp3cpvp(c)\operatorname{cub}(c)\colonequals\prod_{p^{3}\mid c}p^{v_{p}(c)}. Also let sq(0)\colonequals0\operatorname{sq}(0)\colonequals 0 and cub(0)\colonequals0\operatorname{cub}(0)\colonequals 0.

Proposition 9.11 ([hooley1986HasseWeil, heath1998circle]; see e.g. [wang2023_large_sieve_diagonal_cubic_forms]*Proposition 4.9).

Assume FF is diagonal. Let ϵ>0\epsilon>0. There exists A12(F,ϵ)1A_{12}(F,\epsilon)\in\mathbb{R}_{\geq 1} such that if 𝐜m\bm{c}\in\mathbb{Z}^{m} and n1n\in\mathbb{Z}_{\geq 1}, then

n1/2|S𝒄(n)|A12(F,ϵ)nϵ1jmgcd(cub(n)2,gcd(cub(n),sq(cj))3)1/12.n^{-1/2}\lvert S^{\natural}_{\bm{c}}(n)\rvert\leq A_{12}(F,\epsilon)n^{\epsilon}\prod_{1\leq j\leq m}{\gcd}{\left(\operatorname{cub}(n)^{2},\gcd(\operatorname{cub}(n),\operatorname{sq}(c_{j}))^{3}\right)}^{1/12}.

One could replace the nϵn^{\epsilon} in Proposition 9.11 with O(1)ω(n)O(1)^{\omega(n)}. This would be important if one wanted to try to prove a softer version of Proposition 9.9 (conditional on a softer version of Conjecture 1.5). We do not explore this direction in the present paper, since it would seem to involve complicated divisor-type sums with square-full parts and discriminant divisibility conditions (which might require complicated parameterizations to handle).

Proposition 9.11 is an explicit stratification result for S𝒄(n)S_{\bm{c}}(n), based on vp(cj)v_{p}(c_{j}) for pnp\mid n. It would be very nice to have a usable (perhaps less explicit) replacement for Proposition 9.11 when FF is no longer diagonal. Work such as [denef1984rationality, pas1989uniform] could conceivably help.

We now prove Propositions 9.7 and 9.9. For convenience, let 𝒮4\colonequals{𝒄𝒮1:c1cm0}\mathcal{S}_{4}\colonequals\{\bm{c}\in\mathcal{S}_{1}:c_{1}\cdots c_{m}\neq 0\}.

Proof of Proposition 9.7.

Let W=c1cmW=c_{1}\cdots c_{m}. Suppose Z,N1Z,N\in\mathbb{R}_{\geq 1} with NZ3N\leq Z^{3}. For each 𝒄𝒮4\bm{c}\in\mathcal{S}_{4}, Proposition 9.11 implies n1/2S𝒄(n)F,ϵnϵ1jmsq(cj)1/4n^{-1/2}S^{\sharp}_{\bm{c}}(n)\ll_{F,\epsilon}n^{\epsilon}\prod_{1\leq j\leq m}\operatorname{sq}(c_{j})^{1/4}. Plugging this into Σ14W,2(Z,N)\Sigma_{14}^{W,2}(Z,N) (see (9.11)), and using Lemma 2.1 to bound |𝒩𝒄[N,2N)|\lvert\mathcal{N}_{\bm{c}}\cap[N,2N)\rvert, we get Σ14W,2(Z,N)F,ϵZϵ1jm1|c|Zsq(cj)1/2\Sigma_{14}^{W,2}(Z,N)\ll_{F,\epsilon}Z^{\epsilon}\prod_{1\leq j\leq m}\sum_{1\leq\lvert c\rvert\leq Z}\operatorname{sq}(c_{j})^{1/2}. However, by (9.3), we have

(9.12) 1|c|Zsq(c)1/2e𝒩(2)e1/2(Z/e)Zlog(1+Z).\sum_{1\leq\lvert c\rvert\leq Z}\operatorname{sq}(c)^{1/2}\ll\sum_{e\in\mathcal{N}_{\geq}(2)}e^{1/2}\cdot(Z/e)\ll Z\log(1+Z).

Therefore, Σ14W,2(Z,N)F,ϵZm+ϵ\Sigma_{14}^{W,2}(Z,N)\ll_{F,\epsilon}Z^{m+\epsilon}. The statement Σ141,2(Z,N)Σ14W,2(Z,N)F,ϵZm1+ϵN1/3\Sigma_{14}^{1,2}(Z,N)-\Sigma_{14}^{W,2}(Z,N)\ll_{F,\epsilon}Z^{m-1+\epsilon}N^{1/3} holds by a similar calculation with Proposition 9.11, Lemma 2.1, and (9.12), ending with

cub(n)(mt)/31it1|ci|Zsq(ci)1/2t,ϵN(mt)/3Zt+ϵN1/3Zm1+ϵ\operatorname{cub}(n)^{(m-t)/3}\prod_{1\leq i\leq t}\sum_{1\leq\lvert c_{i}\rvert\leq Z}\operatorname{sq}(c_{i})^{1/2}\ll_{t,\epsilon}N^{(m-t)/3}Z^{t+\epsilon}\leq N^{1/3}Z^{m-1+\epsilon}

for n[N,2N)n\in[N,2N), t[1,m1]t\in[1,m-1]; cf. [wang2023_large_sieve_diagonal_cubic_forms]*(4.11) in the proof of Proposition 4.12. Since NZ3N\leq Z^{3}, we obtain Σ141,2(Z,N)F,ϵZm+ϵ\Sigma_{14}^{1,2}(Z,N)\ll_{F,\epsilon}Z^{m+\epsilon}, proving Conjecture 9.6. ∎

Proof of Proposition 9.9.

Suppose 1NZ31\leq N\leq Z^{3}. For each 𝒄m\bm{c}\in\mathbb{Z}^{m} and n1n\in\mathbb{Z}_{\geq 1}, let

S𝒄(n)\colonequalspnmax(1,|S𝒄(pvp(n))|pvp(n)/2).S^{\sharp}_{\bm{c}}(n)\colonequals\prod_{p\mid n}\max\left(1,\frac{\lvert S^{\natural}_{\bm{c}}(p^{v_{p}(n)})\rvert}{p^{v_{p}(n)/2}}\right).

Let P1P\geq 1 be a real parameter to be specified later. Let

Σ15\colonequals𝒄𝒮1:𝒄Z(n𝒩𝒄[N,2N):S𝒄(n)<P|S𝒄(n)|n1/2)2,Σ16\colonequals𝒄𝒮4:𝒄Z(n𝒩𝒄[N,2N):S𝒄(n)P|S𝒄(n)|n1/2)2.\Sigma_{15}\colonequals\sum_{\begin{subarray}{c}\bm{c}\in\mathcal{S}_{1}:\\ \lVert\bm{c}\rVert\leq Z\end{subarray}}\,\biggl{(}\,\sum_{\begin{subarray}{c}n\in\mathcal{N}_{\bm{c}}\cap[N,2N):\\ S^{\sharp}_{\bm{c}}(n)<P\end{subarray}}\frac{\lvert S^{\natural}_{\bm{c}}(n)\rvert}{n^{1/2}}\biggr{)}^{\!2},\quad\Sigma_{16}\colonequals\sum_{\begin{subarray}{c}\bm{c}\in\mathcal{S}_{4}:\\ \lVert\bm{c}\rVert\leq Z\end{subarray}}\,\biggl{(}\,\sum_{\begin{subarray}{c}n\in\mathcal{N}_{\bm{c}}\cap[N,2N):\\ S^{\sharp}_{\bm{c}}(n)\geq P\end{subarray}}\frac{\lvert S^{\natural}_{\bm{c}}(n)\rvert}{n^{1/2}}\biggr{)}^{\!2}.

Let W=c1cmW=c_{1}\cdots c_{m}; then Σ14W,2(Z,N)2(Σ15+Σ16)\Sigma_{14}^{W,2}(Z,N)\leq 2(\Sigma_{15}+\Sigma_{16}), since (a+b)22(a2+b2)(a+b)^{2}\leq 2(a^{2}+b^{2}) for a,ba,b\in\mathbb{R}. We will bound Σ15\Sigma_{15} conditionally (using Conjecture 1.5), and Σ16\Sigma_{16} unconditionally (using the diagonality of FF). We will lose factors of ZϵZ^{\epsilon} at first, and then remove ZϵZ^{\epsilon} later.

We first handle Σ15\Sigma_{15}. Given 𝒄𝒮1[Z,Z]m\bm{c}\in\mathcal{S}_{1}\cap[-Z,Z]^{m}, let

𝒩𝒄,1\colonequals{n𝒩(2):pnlcm(p2,pvp(n)1)Δ(𝒄)},𝒩𝒄,2\colonequals{n𝒩(2):pnlcm(p2,pvp(n)1)Δ(𝒄)}.\begin{split}\mathcal{N}_{\bm{c},1}&\colonequals\{n\in\mathcal{N}_{\geq}(2):p\mid n\Rightarrow\operatorname{lcm}(p^{2},p^{v_{p}(n)-1})\nmid\Delta(\bm{c})\},\\ \mathcal{N}_{\bm{c},2}&\colonequals\{n\in\mathcal{N}_{\geq}(2):p\mid n\Rightarrow\operatorname{lcm}(p^{2},p^{v_{p}(n)-1})\mid\Delta(\bm{c})\}.\end{split}

Any integer n1n\geq 1 can be written uniquely as qn1n2qn_{1}n_{2}, where qq, n1n_{1}, n2n_{2} are pairwise coprime integers satisfying q𝒩(1)q\in\mathcal{N}_{\leq}(1) and nj𝒩𝒄,jn_{j}\in\mathcal{N}_{\bm{c},j} (for j{1,2}j\in\{1,2\}). We then have

n1/2|S𝒄(n)|mindn(S𝒄(n/d)d1/2|S𝒄(d)|)S𝒄(n)min(1,n11/2|S𝒄(n1)|,q1/2|S𝒄(q)|)n^{-1/2}\lvert S^{\natural}_{\bm{c}}(n)\rvert\leq\min_{d\mid n}(S^{\sharp}_{\bm{c}}(n/d)\cdot d^{-1/2}\lvert S^{\natural}_{\bm{c}}(d)\rvert)\leq S^{\sharp}_{\bm{c}}(n)\min(1,n_{1}^{-1/2}\lvert S^{\natural}_{\bm{c}}(n_{1})\rvert,q^{-1/2}\lvert S^{\natural}_{\bm{c}}(q)\rvert)

by the definition of SS^{\sharp}. Here S𝒄(n1)F1S^{\natural}_{\bm{c}}(n_{1})\ll_{F}1 by Lemma 9.1(2)–(3). Also, the integer pn2lcm(p2,pvp(n2)1)𝒩2[n22/3,n2]\prod_{p\mid n_{2}}\operatorname{lcm}(p^{2},p^{v_{p}(n_{2})-1})\in\mathcal{N}_{\geq 2}\cap[n_{2}^{2/3},n_{2}] divides Δ(𝒄)\Delta(\bm{c}). Since |𝒩𝒄[1,2N)|ϵ(ZN)ϵ\lvert\mathcal{N}_{\bm{c}}\cap[1,2N)\rvert\ll_{\epsilon}(ZN)^{\epsilon} (by Lemma 2.1) and min(q,n1,n2)n1/3\min(q,n_{1},n_{2})\geq n^{1/3}, we conclude that

Σ15F,ϵP2Zϵ𝒄𝒮1[Z,Z]m( 1𝒄𝒮5+1(N1/3)1/2+q𝒩𝒄𝒩(1)[N1/3,2N)|S𝒄(q)|q1/2)2,\Sigma_{15}\ll_{F,\epsilon}P^{2}Z^{\epsilon}\sum_{\bm{c}\in\mathcal{S}_{1}\cap[-Z,Z]^{m}}\biggl{(}\,\bm{1}_{\bm{c}\in\mathcal{S}_{5}}+\frac{1}{(N^{1/3})^{1/2}}+\sum_{q\in\mathcal{N}_{\bm{c}}\cap\mathcal{N}_{\leq}(1)\cap[N^{1/3},2N)}\frac{\lvert S^{\natural}_{\bm{c}}(q)\rvert}{q^{1/2}}\biggr{)}^{\!2},

where 𝒮5\mathcal{S}_{5} denotes the set of tuples 𝒄𝒮1\bm{c}\in\mathcal{S}_{1} for which there exists d𝒩2[(N1/3)2/3,2N)d\in\mathcal{N}_{\geq 2}\cap[(N^{1/3})^{2/3},2N) with dΔ(𝒄)d\mid\Delta(\bm{c}). Since (r+s+t)23(r2+s2+t2)(r+s+t)^{2}\leq 3(r^{2}+s^{2}+t^{2}) (for r,s,tr,s,t\in\mathbb{R}), the sum over 𝒄\bm{c} on the right above is F,ϵZm(N2/9)η9+ZmN1/3+ZmNϵmin(N1/3,Z)1+ϵ\ll_{F,\epsilon}Z^{m}(N^{2/9})^{-\eta_{9}}+Z^{m}N^{-1/3}+Z^{m}N^{\epsilon}\min(N^{1/3},Z)^{-1+\epsilon}, by Propositions 9.5 and 9.3. Since NZ3N\leq Z^{3}, it follows that

(9.13) Σ15F,ϵP2Zm+ϵ(N2η9/9+N1/3).\Sigma_{15}\ll_{F,\epsilon}P^{2}Z^{m+\epsilon}(N^{-2\eta_{9}/9}+N^{-1/3}).

We next handle Σ16\Sigma_{16}, by introducing a new Ekedahl-type idea. For each 𝒄𝒮1\bm{c}\in\mathcal{S}_{1}, let

𝒩𝒄,3(P)\colonequals{n𝒩𝒄:S𝒄(n)0,S𝒄(n)P}.\mathcal{N}_{\bm{c},3}(P)\colonequals\{n\in\mathcal{N}_{\bm{c}}:S^{\natural}_{\bm{c}}(n)\neq 0,\;S^{\sharp}_{\bm{c}}(n)\geq P\}.

Let ϵ(0,1)\epsilon\in(0,1) (to be specified); suppose P1/10A12(F,ϵ)(2N)ϵP^{1/10}\geq A_{12}(F,\epsilon)\cdot(2N)^{\epsilon}. Now consider an individual 𝒄𝒮4[Z,Z]m\bm{c}\in\mathcal{S}_{4}\cap[-Z,Z]^{m} and n𝒩𝒄,3(P)[N,2N)n\in\mathcal{N}_{\bm{c},3}(P)\cap[N,2N). Here S𝒄(n)0S^{\natural}_{\bm{c}}(n)\neq 0, so pcub(n)pvp(n)1Δ(𝒄)\prod_{p\mid\operatorname{cub}(n)}p^{v_{p}(n)-1}\mid\Delta(\bm{c}) by Lemma 9.1(3). And S𝒄(n)PS^{\sharp}_{\bm{c}}(n)\geq P, so

P9/10S𝒄(n)/P1/10S𝒄(n)/A12(F,ϵ)nϵ1jmgcd(cub(n),sq(cj))1/4,P^{9/10}\leq S^{\sharp}_{\bm{c}}(n)/P^{1/10}\leq S^{\sharp}_{\bm{c}}(n)/A_{12}(F,\epsilon)n^{\epsilon}\leq\prod_{1\leq j\leq m}\gcd(\operatorname{cub}(n),\operatorname{sq}(c_{j}))^{1/4},

by Proposition 9.11. Thus there exists jj with d\colonequalsgcd(cub(n),sq(cj))P3.6/md\colonequals\gcd(\operatorname{cub}(n),\operatorname{sq}(c_{j}))\geq P^{3.6/m}. Clearly d𝒩(2)d\in\mathcal{N}_{\geq}(2). Also, dcjd\mid c_{j}, so dZd\leq Z (since W(𝒄)0W(\bm{c})\neq 0 implies cj0c_{j}\neq 0); and dcub(n)pcub(n)p2(vp(n)1)Δ(𝒄)2d\mid\operatorname{cub}(n)\mid\prod_{p\mid\operatorname{cub}(n)}p^{2(v_{p}(n)-1)}\mid\Delta(\bm{c})^{2}. Letting 𝒮6(j,d)\colonequals{𝒄𝒮4:dgcd(cj,Δ(𝒄)2)}\mathcal{S}_{6}(j,d)\colonequals\{\bm{c}\in\mathcal{S}_{4}:d\mid\gcd(c_{j},\Delta(\bm{c})^{2})\}, it follows (by taking n𝒩𝒄,3(P)[N,2N)n\in\mathcal{N}_{\bm{c},3}(P)\cap[N,2N) with n1/2|S𝒄(n)|n^{-1/2}\lvert S^{\natural}_{\bm{c}}(n)\rvert maximal, if such an nn exists) that

Σ161jmd𝒩(2):P3.6/mdZ𝒄𝒮6(j,d):𝒄Z|𝒩𝒄[N,2N)|2d1/21im:ijsq(ci)1/2.\Sigma_{16}\leq\sum_{1\leq j\leq m}\sum_{\begin{subarray}{c}d\in\mathcal{N}_{\geq}(2):\\ P^{3.6/m}\leq d\leq Z\end{subarray}}\,\sum_{\begin{subarray}{c}\bm{c}\in\mathcal{S}_{6}(j,d):\\ \lVert\bm{c}\rVert\leq Z\end{subarray}}\,\lvert\mathcal{N}_{\bm{c}}\cap[N,2N)\rvert^{2}\cdot d^{1/2}\prod_{\begin{subarray}{c}1\leq i\leq m:\\ i\neq j\end{subarray}}\operatorname{sq}(c_{i})^{1/2}.

Using Lemma 2.1 to bound |𝒩𝒄[N,2N)|\lvert\mathcal{N}_{\bm{c}}\cap[N,2N)\rvert, we get

(9.14) Σ16F,ϵZϵ1jmd𝒩(2):P3.6/mdZ𝒄𝒮6(j,d):𝒄Zd1/21im:ijsq(ci)1/2.\Sigma_{16}\ll_{F,\epsilon}Z^{\epsilon}\sum_{1\leq j\leq m}\sum_{\begin{subarray}{c}d\in\mathcal{N}_{\geq}(2):\\ P^{3.6/m}\leq d\leq Z\end{subarray}}\,\sum_{\begin{subarray}{c}\bm{c}\in\mathcal{S}_{6}(j,d):\\ \lVert\bm{c}\rVert\leq Z\end{subarray}}d^{1/2}\prod_{\begin{subarray}{c}1\leq i\leq m:\\ i\neq j\end{subarray}}\operatorname{sq}(c_{i})^{1/2}.

For notational simplicity, suppose j=mj=m. For k{0,1}k\in\{0,1\}, let

𝒮6,k(m)\colonequals{(c1,,cm1)({0})m1:𝟏Δ(c1,,cm1,0)0=k}.\mathcal{S}_{6,k}(m)\colonequals\{(c_{1},\dots,c_{m-1})\in(\mathbb{Z}\setminus\{0\})^{m-1}:\bm{1}_{\Delta(c_{1},\dots,c_{m-1},0)\neq 0}=k\}.

For each (c1,,cm1)𝒮6,1(m)[Z,Z]m1(c_{1},\dots,c_{m-1})\in\mathcal{S}_{6,1}(m)\cap[-Z,Z]^{m-1}, we have

d𝒩(2):P3.6/mdZ1|cm|Z:(c1,,cm)𝒮6(m,d)d1/2dΔ(c1,,cm1,0)2:P3.6/mdZ1|cm|Z:dcmd1/2F,ϵZϵZ/(P3.6/m)1/2,\sum_{\begin{subarray}{c}d\in\mathcal{N}_{\geq}(2):\\ P^{3.6/m}\leq d\leq Z\end{subarray}}\,\sum_{\begin{subarray}{c}1\leq\lvert c_{m}\rvert\leq Z:\\ (c_{1},\dots,c_{m})\in\mathcal{S}_{6}(m,d)\end{subarray}}d^{1/2}\leq\sum_{\begin{subarray}{c}d\mid\Delta(c_{1},\dots,c_{m-1},0)^{2}:\\ P^{3.6/m}\leq d\leq Z\end{subarray}}\,\sum_{\begin{subarray}{c}1\leq\lvert c_{m}\rvert\leq Z:\\ d\mid c_{m}\end{subarray}}d^{1/2}\ll_{F,\epsilon}Z^{\epsilon}\cdot Z/(P^{3.6/m})^{1/2},

by the divisor bound for Δ(c1,,cm1,0)20\Delta(c_{1},\dots,c_{m-1},0)^{2}\neq 0. Therefore, the contribution to the right-hand side of (9.14) from 𝒄\bm{c} with (c1,,cm1)𝒮6,1(m)(c_{1},\dots,c_{m-1})\in\mathcal{S}_{6,1}(m) (or the analogous condition if jmj\neq m) is

F,ϵZ1+2ϵP1.8/m1im11|ci|Zsq(ci)1/2m,ϵZm+3ϵP1.8/m,\ll_{F,\epsilon}\frac{Z^{1+2\epsilon}}{P^{1.8/m}}\prod_{1\leq i\leq m-1}\sum_{1\leq\lvert c_{i}\rvert\leq Z}\operatorname{sq}(c_{i})^{1/2}\ll_{m,\epsilon}\frac{Z^{m+3\epsilon}}{P^{1.8/m}},

where we use (9.12) in the final step.

On the other hand, for any c1,,cm2c_{1},\dots,c_{m-2}\in\mathbb{Z}, there are at most degΔ\deg{\Delta} integers cm1c_{m-1} for which (c1,,cm1)𝒮6,0(m)(c_{1},\dots,c_{m-1})\in\mathcal{S}_{6,0}(m). (This is because for diagonal FF, the cidegΔc_{i}^{\deg{\Delta}} coefficient of Δ\Delta is nonzero for all ii; see e.g. [heath1998circle]*(4.2).) Therefore, the contribution to the right-hand side of (9.14) from 𝒄\bm{c} with (c1,,cm1)𝒮6,0(m)(c_{1},\dots,c_{m-1})\in\mathcal{S}_{6,0}(m) (or the analogous condition if jmj\neq m) is

F,ϵZϵd𝒩(2):P3.6/mdZ1|cm|Z:dcmd1/2max1|cm1|Zsq(cm1)1/21im21|ci|Zsq(ci)1/2,\ll_{F,\epsilon}Z^{\epsilon}\sum_{\begin{subarray}{c}d\in\mathcal{N}_{\geq}(2):\\ P^{3.6/m}\leq d\leq Z\end{subarray}}\,\sum_{\begin{subarray}{c}1\leq\lvert c_{m}\rvert\leq Z:\\ d\mid c_{m}\end{subarray}}d^{1/2}\max_{1\leq\lvert c_{m-1}\rvert\leq Z}{\operatorname{sq}(c_{m-1})^{1/2}}\prod_{1\leq i\leq m-2}\sum_{1\leq\lvert c_{i}\rvert\leq Z}\operatorname{sq}(c_{i})^{1/2},

and thus (by (9.12)) F,ϵZ2ϵZZ1/2Zm2=Zm1/2+2ϵ\ll_{F,\epsilon}Z^{2\epsilon}\cdot Z\cdot Z^{1/2}\cdot Z^{m-2}=Z^{m-1/2+2\epsilon}.

The previous two paragraphs imply that the right-hand side of (9.14) is F,ϵZm+3ϵ(P1.8/m+Z1/2)\ll_{F,\epsilon}Z^{m+3\epsilon}(P^{-1.8/m}+Z^{-1/2}). This, combined with (9.13), gives

Σ14W,2(Z,N)2(Σ15+Σ16)F,ϵP2Zm+ϵ(N2η9/9+N1/3)+Zm+3ϵ(P1.8/m+Z1/2).\Sigma_{14}^{W,2}(Z,N)\leq 2(\Sigma_{15}+\Sigma_{16})\ll_{F,\epsilon}P^{2}Z^{m+\epsilon}(N^{-2\eta_{9}/9}+N^{-1/3})+Z^{m+3\epsilon}(P^{-1.8/m}+Z^{-1/2}).

Suppose ϵ2min(1,η9)/270\epsilon\leq 2\min(1,\eta_{9})/270, and let P=A12(F,ϵ)10210ϵN2min(1,η9)/271P=A_{12}(F,\epsilon)^{10}2^{10\epsilon}N^{2\min(1,\eta_{9})/27}\geq 1; then we get

Σ14W,2(Z,N)F,ϵZm+3ϵ(N3.6min(1,η9)/27m+Z1/2)Zm+3ϵNmin(1,η9)/9m,\Sigma_{14}^{W,2}(Z,N)\ll_{F,\epsilon}Z^{m+3\epsilon}(N^{-3.6\min(1,\eta_{9})/27m}+Z^{-1/2})\ll Z^{m+3\epsilon}N^{-\min(1,\eta_{9})/9m},

since m4m\geq 4 and ZN1/3Z\geq N^{1/3}. It follows (upon redefining ϵ\epsilon, now that we can forget about PP) that Σ14W,2(Z,N)F,ϵZm+ϵNmin(1,η9)/9m\Sigma_{14}^{W,2}(Z,N)\ll_{F,\epsilon}Z^{m+\epsilon}N^{-\min(1,\eta_{9})/9m} for all ϵ>0\epsilon>0.

It remains to replace ZϵZ^{\epsilon} with NϵN^{\epsilon}. This is trivial unless NN is very small relative to ZZ. Expanding the square in (9.11), and noting that S𝒄(n)S_{\bm{c}}(n) depends only on 𝒄modn\bm{c}\bmod{n}, gives

Σ14W,2(Z,N)n1,n2[N,2N)(1+2Zn1n2)m1𝒂n1n2:rad(n1n2)Δ(𝒂)(n1n2)1/2|S𝒂(n1)S𝒂(n2)|\Sigma_{14}^{W,2}(Z,N)\leq\sum_{n_{1},n_{2}\in[N,2N)}\left(1+\frac{2Z}{n_{1}n_{2}}\right)^{m}\sum_{\begin{subarray}{c}1\leq\bm{a}\leq n_{1}n_{2}:\\ \operatorname{rad}(n_{1}n_{2})\mid\Delta(\bm{a})\end{subarray}}(n_{1}n_{2})^{-1/2}\lvert S^{\natural}_{\bm{a}}(n_{1})S^{\natural}_{\bm{a}}(n_{2})\rvert

for all Z,N1Z,N\geq 1. Yet for any n1,n21n_{1},n_{2}\geq 1 and 𝒂[1,n1n2]m\bm{a}\in[1,n_{1}n_{2}]^{m}, there exists 𝒄𝒮4[1,Bn1n2]m\bm{c}\in\mathcal{S}_{4}\cap[1,Bn_{1}n_{2}]^{m} with 𝒄𝒂modn1n2\bm{c}\equiv\bm{a}\bmod{n_{1}n_{2}}, where B=1+deg(c1cmΔ(𝒄))B=1+\deg(c_{1}\cdots c_{m}\Delta(\bm{c})). So

Σ14W,2(Z,N)n1,n2[N,2N)1𝒂n1n2:rad(n1n2)Δ(𝒂)(n1n2)1/2|S𝒂(n1)S𝒂(n2)|\Sigma_{14}^{W,2}(Z,N)\geq\sum_{n_{1},n_{2}\in[N,2N)}\sum_{\begin{subarray}{c}1\leq\bm{a}\leq n_{1}n_{2}:\\ \operatorname{rad}(n_{1}n_{2})\mid\Delta(\bm{a})\end{subarray}}(n_{1}n_{2})^{-1/2}\lvert S^{\natural}_{\bm{a}}(n_{1})S^{\natural}_{\bm{a}}(n_{2})\rvert

for all Z4BN2Z\geq 4BN^{2}. Hence for all ZN2Z\geq N^{2} we have

Σ14W,2(Z,N)(3Z/N2)mΣ14W,2(4BN2,N)F,ϵ(Z/N2)m(4BN2)m+ϵNmin(1,η9)/9m,\Sigma_{14}^{W,2}(Z,N)\leq(3Z/N^{2})^{m}\cdot\Sigma_{14}^{W,2}(4BN^{2},N)\ll_{F,\epsilon}(Z/N^{2})^{m}(4BN^{2})^{m+\epsilon}N^{-\min(1,\eta_{9})/9m},

since 1N(4BN2)31\leq N\leq(4BN^{2})^{3}. For all ZN1/3Z\geq N^{1/3} (including Z<N2Z<N^{2}, where Zϵ<N2ϵZ^{\epsilon}<N^{2\epsilon}), then,

Σ14W,2(Z,N)F,ϵZmNϵmin(1,η9)/9m.\Sigma_{14}^{W,2}(Z,N)\ll_{F,\epsilon}Z^{m}N^{\epsilon-\min(1,\eta_{9})/9m}.

Therefore, Conjecture 9.8 holds with W=c1cmW=c_{1}\cdots c_{m} and η10(F,2r)=rmin(1,η9)/10m\eta_{10}(F,2r)=r\min(1,\eta_{9})/10m (for r=1r=1 at first, and then for r[1/2,1]r\in[1/2,1] by Hölder over 𝒄\bm{c}). ∎

Remark 9.12.

Handling sq(ci)1/2\operatorname{sq}(c_{i})^{1/2} takes care, because 1|c|Zsq(c)1/2\sum_{1\leq\lvert c\rvert\leq Z}\operatorname{sq}(c)^{1/2} remains roughly the same size even if we restrict |c|\lvert c\rvert to 𝒩(2)\mathcal{N}_{\geq}(2) (a rather sparse set). One could take W=1W=1 in Proposition 9.9 if we were only interested in Σ14W,A(Z,N)\Sigma_{14}^{W,A}(Z,N) for A[1,2)A\in[1,2), or for NZ3δN\leq Z^{3-\delta}.

To end §9, we prove a result clarifying the nature of Conjecture 1.5 for diagonal FF:

Proposition 9.13.

Assume Conjecture 1.5 holds when (m,F)=(4,x13+x23+x33+x43)(m,F)=(4,x_{1}^{3}+x_{2}^{3}+x_{3}^{3}+x_{4}^{3}). Then Conjecture 1.5 holds whenever FF is diagonal and m4m\geq 4.

Proof.

(This result is not used in the rest of the paper, so we confine ourselves to a sketch.)

Assume 1PZ3/21\leq P\leq Z^{3/2}. Recall 𝒩𝒄,0\mathcal{N}_{\bm{c},0} from (9.5). By [bhargava2014geometric]*Theorem 3.3, we have

(9.15) #{𝒄[Z,Z]m:p[P,2P)𝒩𝒄,0}ΔZmPΘ(1).\#\{\bm{c}\in[-Z,Z]^{m}:\exists\;\textnormal{$p\in[P,2P)\cap\mathcal{N}_{\bm{c},0}$}\}\ll_{\Delta}Z^{m}P^{-\Theta(1)}.

By (9.15), we see that for any given FF, Conjecture 1.5 is equivalent to the statement

(9.16) #{𝒄[Z,Z]m:p[P,2P)𝒩𝒄,0 with p2Δ(𝒄)}ZmPΘ(1).\#\{\bm{c}\in[-Z,Z]^{m}:\exists\;\textnormal{$p\in[P,2P)\setminus\mathcal{N}_{\bm{c},0}$ with $p^{2}\mid\Delta(\bm{c})$}\}\ll Z^{m}P^{-\Theta(1)}.

Let 𝒰1,𝒰2m1\mathcal{U}_{1},\mathcal{U}_{2}\subseteq\mathbb{P}^{m-1}_{\mathbb{Z}} be the smooth loci of the hypersurfaces F=0F=0, Δ=0\Delta=0, respectively, over \mathbb{Z}. The gradient F\nabla{F} defines a Gauss map [F]:𝒰1m1[\nabla{F}]\colon\mathcal{U}_{1}\to\mathbb{P}^{m-1}_{\mathbb{Z}}; let 𝒰3𝒰1\mathcal{U}_{3}\subseteq\mathcal{U}_{1} be the inverse image of 𝒰2\mathcal{U}_{2} under this map. The map [F]:𝒰3𝒰2[\nabla{F}]\colon\mathcal{U}_{3}\to\mathcal{U}_{2} is an isomorphism over \mathbb{Q} (by the biduality theorem), and thus an isomorphism over \mathbb{Z} (since 𝒰2\mathcal{U}_{2}, 𝒰3\mathcal{U}_{3} are flat over \mathbb{Z}). In particular, 𝒰2(/p2)=[F](𝒰3(/p2))\mathcal{U}_{2}(\mathbb{Z}/p^{2}\mathbb{Z})=[\nabla{F}](\mathcal{U}_{3}(\mathbb{Z}/p^{2}\mathbb{Z})). Furthermore, it is known that 𝒰3\mathcal{U}_{3} lies in the open subscheme det(HessF(𝒙))0\det(\operatorname{Hess}{F}(\bm{x}))\neq 0 of m1\mathbb{P}^{m-1}_{\mathbb{Z}}. These two facts, plus (9.15), imply that (9.16) is equivalent to

(9.17) #{𝒄[Z,Z]m:p[P,2P),𝒙m,λ[1,p1] with pdet(HessF(𝒙)),p2F(𝒙),F(𝒙)λ𝒄}ZmPΘ(1).\begin{split}\#\{\bm{c}\in[-Z,Z]^{m}:\exists\;&p\in[P,2P),\;\bm{x}\in\mathbb{Z}^{m},\;\lambda\in[1,p-1]\textnormal{ with }\\ &p\nmid\det(\operatorname{Hess}{F}(\bm{x})),\;p^{2}\mid F(\bm{x}),\nabla{F}(\bm{x})-\lambda\bm{c}\}\ll Z^{m}P^{-\Theta(1)}.\end{split}

We would like to convert each “exists” into a sum. Each pp on the left-hand side of (9.17) satisfies pΔ(𝒄)p\mid\Delta(\bm{c}), so there are at most finitely many possibilities for pp if 𝒄𝒮1\bm{c}\in\mathcal{S}_{1}. Also, |𝒮0[Z,Z]m|PmZm1/2\lvert\mathcal{S}_{0}\cap[-Z,Z]^{m}\rvert\cdot P\ll_{m}Z^{m-1/2} by (2.4). Therefore, the statement (9.17) is equivalent to

(9.18) 𝒄[Z,Z]mp[P,2P)1𝒙p2𝟏pdet(HessF(𝒙))𝟏p2F(𝒙)𝔼1λp1[𝟏p2F(𝒙)λ𝒄]ZmPΘ(1).\sum_{\bm{c}\in[-Z,Z]^{m}}\sum_{p\in[P,2P)}\sum_{1\leq\bm{x}\leq p^{2}}\bm{1}_{p\nmid\det(\operatorname{Hess}{F}(\bm{x}))}\bm{1}_{p^{2}\mid F(\bm{x})}\mathbb{E}_{1\leq\lambda\leq p-1}[\bm{1}_{p^{2}\mid\nabla{F}(\bm{x})-\lambda\bm{c}}]\ll Z^{m}P^{-\Theta(1)}.

Now suppose F=F1x13++Fmxm3F=F_{1}x_{1}^{3}+\dots+F_{m}x_{m}^{3} (where Fj{0}F_{j}\in\mathbb{Z}\setminus\{0\}), and for each G,λG,\lambda\in\mathbb{Z} let

T3/2,G,λ(a,p2)\colonequalsc[Z,Z], 1xp2:px,p23Gx2λcep2(aGx3).T_{3/2,G,\lambda}(a,p^{2})\colonequals\sum_{\begin{subarray}{c}c\in[-Z,Z],\;1\leq x\leq p^{2}:\\ p\nmid x,\;p^{2}\mid 3Gx^{2}-\lambda c\end{subarray}}e_{p^{2}}(aGx^{3})\in\mathbb{R}.

Then the left-hand side of (9.18) equals 1/(p1)1/(p-1) times

(9.19) p[P,2P),1ap2, 1λp11jmT3/2,Fj,λ(a,p2)FZm41j4p[P,2P),1ap2, 1λp1T3/2,Fj,λ(a,p2)4,\sum_{\begin{subarray}{c}p\in[P,2P),\\ 1\leq a\leq p^{2},\;1\leq\lambda\leq p-1\end{subarray}}\prod_{1\leq j\leq m}T_{3/2,F_{j},\lambda}(a,p^{2})\ll_{F}Z^{m-4}\sum_{1\leq j\leq 4}\sum_{\begin{subarray}{c}p\in[P,2P),\\ 1\leq a\leq p^{2},\;1\leq\lambda\leq p-1\end{subarray}}T_{3/2,F_{j},\lambda}(a,p^{2})^{4},

since T3/2,G,λ(a,p2)GZT_{3/2,G,\lambda}(a,p^{2})\ll_{G}Z trivially (by considering the cases p3Gp\nmid 3G and p3Gp\mid 3G separately) and 1j4|T3/2,Fj,λ(a,p2)|1j4T3/2,Fj,λ(a,p2)4\prod_{1\leq j\leq 4}\lvert T_{3/2,F_{j},\lambda}(a,p^{2})\rvert\leq\sum_{1\leq j\leq 4}T_{3/2,F_{j},\lambda}(a,p^{2})^{4} (since T3/2,G,λT_{3/2,G,\lambda}\in\mathbb{R}).

Since Conjecture 1.5 holds by assumption when F=x13+x23+x33+x43F=x_{1}^{3}+x_{2}^{3}+x_{3}^{3}+x_{4}^{3}, it also holds when F=Fj(x13+x23+x33+x43)F=F_{j}\cdot(x_{1}^{3}+x_{2}^{3}+x_{3}^{3}+x_{4}^{3}) for any fixed jj (since scaling Δ\Delta does not affect the truth of Conjecture 1.5). So by (9.19), the statement (9.18) holds for all diagonal FF (assuming m4m\geq 4). Hence (9.17), (9.16), and Conjecture 1.5 hold too. ∎

10. Delta endgame

Throughout §10, assume 2m2\mid m. Recall Σ(X,𝒮)\Sigma(X,\mathcal{S}), Σ(X,𝒮)\Sigma^{\natural}(X,\mathcal{S}) from (2.15). Explicitly, we have

(10.1) Σ(X,𝒮)=n1𝒄𝒮n(1m)/2S𝒄(n)J𝒄,X(n)=𝒄𝒮n1n(1m)/2S𝒄(n)J𝒄,X(n)\Sigma^{\natural}(X,\mathcal{S})=\sum_{n\geq 1}\sum_{\bm{c}\in\mathcal{S}}n^{(1-m)/2}S^{\natural}_{\bm{c}}(n)J_{\bm{c},X}(n)=\sum_{\bm{c}\in\mathcal{S}}\sum_{n\geq 1}n^{(1-m)/2}S^{\natural}_{\bm{c}}(n)J_{\bm{c},X}(n)

(by Proposition 2.4 and Fubini). We are finally prepared to analyze these sums for 𝒮𝒮1\mathcal{S}\subseteq\mathcal{S}_{1}.

10.1. Delta decomposition

Consider an individual 𝒄𝒮1\bm{c}\in\mathcal{S}_{1}. In view of (7.1), (7.2) and the factorization Φ=ΦGΦB\Phi=\Phi^{\operatorname{G}}\Phi^{\operatorname{B}}, we may use Lemma 5.2 with k=2k=2,

a(𝒏)=a(n0,n1)=(n0it0S𝒄(n0)𝟏n0𝒩𝒄)(n1it1S𝒄(n1)𝟏n1𝒩𝒄),a(\bm{n})=a(n_{0},n_{1})=(n_{0}^{-it_{0}}S^{\natural}_{\bm{c}}(n_{0})\bm{1}_{n_{0}\in\mathcal{N}_{\bm{c}}})\cdot(n_{1}^{-it_{1}}S^{\natural}_{\bm{c}}(n_{1})\bm{1}_{n_{1}\in\mathcal{N}^{\bm{c}}}),

and f(𝒓)=f(r0,r1)=(r0r1)(1m)/2J𝒄,X(r0r1)f(\bm{r})=f(r_{0},r_{1})=(r_{0}r_{1})^{(1-m)/2}J_{\bm{c},X}(r_{0}r_{1}), to write

(10.2) n1n(1m)/2S𝒄(n)J𝒄,X(n)=(2π)2𝑵1/2d×𝑵𝒕2𝑑𝒕g𝒄,X,𝑵(i𝒕)0j1Σ17,𝑵𝒄,j(𝒕),\sum_{n\geq 1}n^{(1-m)/2}S^{\natural}_{\bm{c}}(n)J_{\bm{c},X}(n)=(2\pi)^{-2}\int_{\bm{N}\geq 1/2}d^{\times}\bm{N}\int_{\bm{t}\in\mathbb{R}^{2}}d\bm{t}\,g_{\bm{c},X,\bm{N}}^{\vee}(i\bm{t})\prod_{0\leq j\leq 1}\Sigma_{17,\bm{N}}^{\bm{c},j}(\bm{t}),

where 𝑵=(N0,N1)\bm{N}=(N_{0},N_{1}) runs over [1/2,)2[1/2,\infty)^{2}, where 𝒕=(t0,t1)2\bm{t}=(t_{0},t_{1})\in\mathbb{R}^{2}, and where

(10.3) g𝒄,X,𝑵(𝒓)\displaystyle g_{\bm{c},X,\bm{N}}(\bm{r}) \colonequals(r0r1)(1m)/2J𝒄,X(r0r1)ν2(r0/N0)ν2(r1/N1),\displaystyle\colonequals(r_{0}r_{1})^{(1-m)/2}J_{\bm{c},X}(r_{0}r_{1})\cdot\nu_{2}(r_{0}/N_{0})\nu_{2}(r_{1}/N_{1}),
(10.4) Σ17,𝑵𝒄,0(𝒕)\displaystyle\Sigma_{17,\bm{N}}^{\bm{c},0}(\bm{t}) \colonequalsn0𝒩𝒄ν2(n0/N0)n0it0S𝒄(n0),Σ17,𝑵𝒄,1(𝒕)\colonequalsn1𝒩𝒄ν2(n1/N1)n1it1S𝒄(n1).\displaystyle\colonequals\sum_{n_{0}\in\mathcal{N}_{\bm{c}}}\nu_{2}(n_{0}/N_{0})n_{0}^{-it_{0}}S^{\natural}_{\bm{c}}(n_{0}),\quad\Sigma_{17,\bm{N}}^{\bm{c},1}(\bm{t})\colonequals\sum_{n_{1}\in\mathcal{N}^{\bm{c}}}\nu_{2}(n_{1}/N_{1})n_{1}^{-it_{1}}S^{\natural}_{\bm{c}}(n_{1}).

Since J𝒄,X(r)J_{\bm{c},X}(r) is supported on rA0Yr\leq A_{0}Y (by Proposition 2.4) and ν2\nu_{2} is supported on [1,2][1,2], we have g𝒄,X,𝑵(𝒓)=0g_{\bm{c},X,\bm{N}}(\bm{r})=0 unless r0r1A0Yr_{0}r_{1}\leq A_{0}Y and Njrj2NjN_{j}\leq r_{j}\leq 2N_{j} for all jj. Thus g𝒄,X,𝑵=0g_{\bm{c},X,\bm{N}}=0 identically unless N0N1A0YN_{0}N_{1}\leq A_{0}Y. So g𝒄,X,𝑵(i𝒕)=0g_{\bm{c},X,\bm{N}}^{\vee}(i\bm{t})=0 unless 𝑵\bm{N} lies in the set

(10.5) 17(X)\colonequals{𝑵1/2:N0N1A0Y}\mathscr{R}_{17}(X)\colonequals\{\bm{N}\geq 1/2:N_{0}N_{1}\leq A_{0}Y\}

(cf. the region 10\mathscr{R}_{10} from (7.14)). So (10.2) holds even if we restrict 𝑵\bm{N} to 17(X)\mathscr{R}_{17}(X).

Given 𝑵=(N0,N1)\bm{N}=(N_{0},N_{1}), let N\colonequalsN0N1N\colonequals N_{0}N_{1}. For integers C1C\geq 1 and reals λ>0\lambda>0, let

𝒮1(C,λ)\colonequals{𝒄𝒮1:𝒄[C,2C),|Δ(𝒄/C)|(λ/2,λ]}.\mathcal{S}_{1}(C,\lambda)\colonequals\{\bm{c}\in\mathcal{S}_{1}:\lVert\bm{c}\rVert\in[C,2C),\;\lvert\Delta(\bm{c}/C)\rvert\in(\lambda/2,\lambda]\}.

Assume (2.6). The definition (10.3) and Propositions 5.1 and 8.1 imply, for 𝒄𝒮1(C,λ)\bm{c}\in\mathcal{S}_{1}(C,\lambda) and b0b\in\mathbb{Z}_{\geq 0} (and multi-indices 𝜶0\bm{\alpha}\geq 0), that

(10.6) N1/2𝒄𝜶g𝒄,X,𝑵(i𝒕)F,w,ν2,b(X/N)|𝜶|(N+XC)1m/2(1+𝒕)b(1+C/X1/2)b(1+XCλ/N)b.N^{1/2}\partial_{\bm{c}}^{\bm{\alpha}}{g_{\bm{c},X,\bm{N}}^{\vee}(i\bm{t})}\ll_{F,w,\nu_{2},b}\frac{(X/N)^{\lvert\bm{\alpha}\rvert}(N+XC)^{1-m/2}}{(1+\lVert\bm{t}\rVert)^{b}(1+C/X^{1/2})^{b}(1+XC\lambda/N)^{b}}.

For each 𝒄𝒮1\bm{c}\in\mathcal{S}_{1}, we have 𝒄m\bm{c}\in\mathbb{Z}^{m} and Δ(𝒄)0\Delta(\bm{c})\neq 0, so 𝒄1\lVert\bm{c}\rVert\geq 1 and 1|Δ(𝒄)|𝒄degΔ1\leq\lvert\Delta(\bm{c})\rvert\ll\lVert\bm{c}\rVert^{\deg{\Delta}}. Let 𝒟={1,2,4,8,}\mathcal{D}=\{1,2,4,8,\ldots\} and 𝒟2=𝒟{1/d:d𝒟}\mathcal{D}_{2}=\mathcal{D}\cup\{1/d:d\in\mathcal{D}\}; then we deduce that

(10.7) 𝒮1C𝒟,λ𝒟2: 1/CdegΔλA13𝒮1(C,λ)\mathcal{S}_{1}\subseteq\bigcup_{C\in\mathcal{D},\;\lambda\in\mathcal{D}_{2}:\,1/C^{\deg{\Delta}}\leq\lambda\leq A_{13}}\mathcal{S}_{1}(C,\lambda)

for some A13=A13(F)1A_{13}=A_{13}(F)\geq 1. Also, the volume bound (4.4) implies (for all C1C\geq 1 and λA13\lambda\leq A_{13})

(10.8) |𝒮1(C,λ)|FCm(λ+C1)1/degΔ,\lvert\mathcal{S}_{1}(C,\lambda)\rvert\ll_{F}C^{m}(\lambda+C^{-1})^{1/\deg{\Delta}},

because |Δ(𝒛)|F|Δ(𝒚)|+C1\lvert\Delta(\bm{z})\rvert\ll_{F}\lvert\Delta(\bm{y})\rvert+C^{-1} for all 𝒚[2,2]m\bm{y}\in[-2,2]^{m} and 𝒛𝒚+[C1,C1]m\bm{z}\in\bm{y}+[-C^{-1},C^{-1}]^{m}.

10.2. Sharp delta bounds

In [wang2023_large_sieve_diagonal_cubic_forms], we used Cauchy–Schwarz on 1/L1/L, ΦL\Phi L over 𝒄\bm{c} to conditionally prove (2.17) (for diagonal FF with 2m2\mid m) under a large-sieve hypothesis, based on the first-order approximation Φ1/L\Phi\approx 1/L (see (2.18)). Now, in §10.2, we will use Hölder in a similar spirit to prove Theorem 1.3. This relies crucially on several new features, including the more precise Φ\Phi-data captured in Σ17,𝑵𝒄,1(𝒕)\Sigma_{17,\bm{N}}^{\bm{c},1}(\bm{t}) (compared to 1/L1/L in [wang2023_large_sieve_diagonal_cubic_forms]).

For the next four results, let C𝒟C\in\mathcal{D}, let 𝑵17(X)\bm{N}\in\mathscr{R}_{17}(X), let 𝒕2\bm{t}\in\mathbb{R}^{2}, and let 𝒮𝒮1[C,C]m\mathcal{S}\subseteq\mathcal{S}_{1}\cap[-C,C]^{m}. Roughly speaking, (10.11) will be useful when CX1/2δC\leq X^{1/2-\delta}; and otherwise, (10.9) will be useful for small N0N_{0} (relative to XC/NXC/N), and (10.13) useful for larger N0N_{0}.

Lemma 10.1.

Assume Conjecture 7.4. Then for some A14=A14(F,ϵ)>0A_{14}=A_{14}(F,\epsilon)>0, we have

(10.9) Σ17,𝑵𝒄,0(𝒕)Σ17,𝑵𝒄,1(𝒕)𝒄1(𝒮)F,ϵ(1+|t1|)A14|𝒮|1/2ϵ(Cm+N1m/3)1/2+ϵN0m/2+ϵN1/2.\lVert\Sigma_{17,\bm{N}}^{\bm{c},0}(\bm{t})\cdot\Sigma_{17,\bm{N}}^{\bm{c},1}(\bm{t})\rVert_{\ell^{1}_{\bm{c}}(\mathcal{S})}\ll_{F,\epsilon}(1+\lvert t_{1}\rvert)^{A_{14}}\cdot\lvert\mathcal{S}\rvert^{1/2-\epsilon}(C^{m}+N_{1}^{m/3})^{1/2+\epsilon}N_{0}^{m/2+\epsilon}N^{1/2}.
Proof.

Plugging the trivial bound |S𝒄(n0)|n0(1+m)/2\lvert S^{\natural}_{\bm{c}}(n_{0})\rvert\leq n_{0}^{(1+m)/2} into (10.4) gives

(10.10) Σ17,𝑵𝒄,0(𝒕)ν2N0(1+m)/2|𝒩𝒄[N0,2N0)|\Sigma_{17,\bm{N}}^{\bm{c},0}(\bm{t})\ll_{\nu_{2}}N_{0}^{(1+m)/2}\cdot\lvert\mathcal{N}_{\bm{c}}\cap[N_{0},2N_{0})\rvert

for all 𝒄𝒮\bm{c}\in\mathcal{S}. So by Lemma 6.17 (with A=(2ϵ)/ϵA=(2-\epsilon)/\epsilon), Conjecture 7.4 (with C+N11/3C+N_{1}^{1/3}, N1N_{1} in place of ZZ, NN), and Hölder over 𝒄𝒮\bm{c}\in\mathcal{S}, the left-hand of (10.9) is

F,ϵN0(1+m)/2(CmN0ϵ)ϵ/(2ϵ)|𝒮|(12ϵ)/(2ϵ)[(1+|t1|)A6(F,ϵ)(Cm+N1m/3)N1(2ϵ)/2]1/(2ϵ)\ll_{F,\epsilon}N_{0}^{(1+m)/2}\cdot(C^{m}N_{0}^{\epsilon})^{\epsilon/(2-\epsilon)}\cdot\lvert\mathcal{S}\rvert^{(1-2\epsilon)/(2-\epsilon)}\cdot[(1+\lvert t_{1}\rvert)^{A_{6}(F,\epsilon)}(C^{m}+N_{1}^{m/3})N_{1}^{(2-\epsilon)/2}]^{1/(2-\epsilon)}

for all ϵ(0,12)\epsilon\in(0,\frac{1}{2}), since 1=ϵ2ϵ+12ϵ2ϵ+12ϵ1=\frac{\epsilon}{2-\epsilon}+\frac{1-2\epsilon}{2-\epsilon}+\frac{1}{2-\epsilon}. Writing N0(1+m)/2N11/2=N0m/2N1/2N_{0}^{(1+m)/2}N_{1}^{1/2}=N_{0}^{m/2}N^{1/2} gives (10.9), since 12ϵ2ϵ12O(ϵ)\frac{1-2\epsilon}{2-\epsilon}\geq\frac{1}{2}-O(\epsilon) and |𝒮|CmCm+N1m/3\lvert\mathcal{S}\rvert\leq C^{m}\leq C^{m}+N_{1}^{m/3}. ∎

Lemma 10.2.

Assume Conjectures 1.2 and 9.6. Then

(10.11) Σ17,𝑵𝒄,0(𝒕)Σ17,𝑵𝒄,1(𝒕)𝒄1(𝒮)F,ϵ(1+|t1|)ϵ|𝒮|1/2(Cm+N0m/3)1/2+ϵN1/2+ϵ.\lVert\Sigma_{17,\bm{N}}^{\bm{c},0}(\bm{t})\cdot\Sigma_{17,\bm{N}}^{\bm{c},1}(\bm{t})\rVert_{\ell^{1}_{\bm{c}}(\mathcal{S})}\ll_{F,\epsilon}(1+\lvert t_{1}\rvert)^{\epsilon}\cdot\lvert\mathcal{S}\rvert^{1/2}(C^{m}+N_{0}^{m/3})^{1/2+\epsilon}N^{1/2+\epsilon}.
Proof.

By Proposition 3.2(8) and partial summation, |Σ17,𝑵𝒄,1(𝒕)|ν2,ϵCϵ(1+|t1|)ϵN11/2+ϵ\lvert\Sigma_{17,\bm{N}}^{\bm{c},1}(\bm{t})\rvert\ll_{\nu_{2},\epsilon}C^{\epsilon}(1+\lvert t_{1}\rvert)^{\epsilon}N_{1}^{1/2+\epsilon}. By Cauchy–Schwarz and (9.11), the left-hand side of (10.11) is F,ϵ[N0Σ141,2(C,N0)]1/2|𝒮|1/2Cϵ(1+|t1|)ϵN11/2+ϵ\ll_{F,\epsilon}[N_{0}\cdot\Sigma_{14}^{1,2}(C,N_{0})]^{1/2}\cdot\lvert\mathcal{S}\rvert^{1/2}\cdot C^{\epsilon}(1+\lvert t_{1}\rvert)^{\epsilon}N_{1}^{1/2+\epsilon}. Now use Conjecture 9.6 (with C+N01/3C+N_{0}^{1/3}, N0N_{0} in place of ZZ, NN). ∎

Proposition 10.3.

Assume Conjecture 9.8 with some WW. Then for any real δ0\delta\geq 0, we have

(10.12) 𝒄𝒮:W(𝒄)0|N01/2Σ17,𝑵𝒄,0(𝒕)|2+δF,δ(Cm+N0m/3)N0mδ/29η10(F,2)/10.\sum_{\bm{c}\in\mathcal{S}:\,W(\bm{c})\neq 0}\lvert N_{0}^{-1/2}\cdot\Sigma_{17,\bm{N}}^{\bm{c},0}(\bm{t})\rvert^{2+\delta}\ll_{F,\delta}(C^{m}+N_{0}^{m/3})N_{0}^{m\delta/2-9\eta_{10}(F,2)/10}.
Proof.

We may assume 𝒮{W0}\mathcal{S}\subseteq\{W\neq 0\}. By Hölder over 𝒮\mathcal{S}, the left-hand side of (10.12) is

(N01/2Σ17,𝑵𝒄,0(𝒕))δ+ϵ𝒄2/ϵ(𝒮)(N01/2Σ17,𝑵𝒄,0(𝒕))2ϵ𝒄2/(2ϵ)(𝒮),\leq\lVert(N_{0}^{-1/2}\cdot\Sigma_{17,\bm{N}}^{\bm{c},0}(\bm{t}))^{\delta+\epsilon}\rVert_{\ell^{2/\epsilon}_{\bm{c}}(\mathcal{S})}\cdot\lVert(N_{0}^{-1/2}\cdot\Sigma_{17,\bm{N}}^{\bm{c},0}(\bm{t}))^{2-\epsilon}\rVert_{\ell^{2/(2-\epsilon)}_{\bm{c}}(\mathcal{S})},

for any ϵ[0,δ]\epsilon\in[0,\delta], since ϵ2+2ϵ2=1\frac{\epsilon}{2}+\frac{2-\epsilon}{2}=1. The first factor here is δ,ϵ(N0m/2)δ+ϵ(CmN0ϵ)ϵ/2\ll_{\delta,\epsilon}(N_{0}^{m/2})^{\delta+\epsilon}(C^{m}N_{0}^{\epsilon})^{\epsilon/2} by (10.10) and Lemma 6.17; the second factor is ϵ[(C+N01/3)mN0η10(F,2)](2ϵ)/2\ll_{\epsilon}[(C+N_{0}^{1/3})^{m}N_{0}^{-\eta_{10}(F,2)}]^{(2-\epsilon)/2} by Conjecture 9.8. Taking ϵ\epsilon sufficiently small in terms of mm, η10(F,2)\eta_{10}(F,2) gives (10.12). ∎

Lemma 10.4.

Fix ξ0\xi\in\mathbb{R}_{\geq 0}. Assume Conjectures 1.2, 9.6, and 9.8. Also assume Conjecture 7.4 if ξ=0\xi=0. Assume NC6N\ll C^{6}. Then for some A15=A15(F,ϵ)>0A_{15}=A_{15}(F,\epsilon)>0, we have

(10.13) Σ17,𝑵𝒄,0(𝒕)Σ17,𝑵𝒄,1(𝒕)𝒄1(𝒮)F,ϵ(1+|t1|)A15(CN1)ξ(Cm)1/2ϵ(Cm+Nm/3)1/2+ϵN1/2min(C9/20,N04η10(F,2)/5).\lVert\Sigma_{17,\bm{N}}^{\bm{c},0}(\bm{t})\cdot\Sigma_{17,\bm{N}}^{\bm{c},1}(\bm{t})\rVert_{\ell^{1}_{\bm{c}}(\mathcal{S})}\ll_{F,\epsilon}(1+\lvert t_{1}\rvert)^{A_{15}}\cdot\frac{(CN_{1})^{\xi}(C^{m})^{1/2-\epsilon}(C^{m}+N^{m/3})^{1/2+\epsilon}N^{1/2}}{\min(C^{9/20},N_{0}^{4\eta_{10}(F,2)/5})}.
Proof.

Case 1: 𝒮{W=0}\mathcal{S}\subseteq\{W=0\}. Then |𝒮|WCm1\lvert\mathcal{S}\rvert\ll_{W}C^{m-1} (see e.g. [bhargava2014geometric]*Lemma 3.1). Inserting this into (10.11), we get (10.13) upon writing C(m1)/2NϵϵC(m1)/2+6ϵC^{(m-1)/2}N^{\epsilon}\ll_{\epsilon}C^{(m-1)/2+6\epsilon} (for a small ϵ\epsilon).

Case 2: 𝒮{W0}\mathcal{S}\subseteq\{W\neq 0\}. Since 1=1ϵ2ϵ+12ϵ1=\frac{1-\epsilon}{2-\epsilon}+\frac{1}{2-\epsilon}, we may use (10.12) (with δ=ϵ1ϵ\delta=\frac{\epsilon}{1-\epsilon}), Conjecture 7.4 (if ξ=0\xi=0) or GRH (if ξ>0\xi>0, using (7.17) and Proposition 3.2(8) to prove Conjecture 7.4 up to a factor of ZϵZ^{\epsilon}), and Hölder over 𝒄𝒮\bm{c}\in\mathcal{S} to bound the left-hand side of (10.13) by OF,ϵ(N1/2(CN1)ξ)[(Cm+N0m/3)N0mδ/29η10(F,2)/10](1ϵ)/(2ϵ)[(1+|t1|)A6(F,ϵ)(Cm+N1m/3)]1/(2ϵ)O_{F,\epsilon}(N^{1/2}(CN_{1})^{\xi})\cdot[(C^{m}+N_{0}^{m/3})N_{0}^{m\delta/2-9\eta_{10}(F,2)/10}]^{(1-\epsilon)/(2-\epsilon)}\cdot[(1+\lvert t_{1}\rvert)^{A_{6}(F,\epsilon)}(C^{m}+N_{1}^{m/3})]^{1/(2-\epsilon)}. Since Cm+Njm/3CmC^{m}+N_{j}^{m/3}\ll C^{m} for some j{0,1}j\in\{0,1\}, we get (10.13) if ϵ\epsilon is small.

Case 3: The general case. Decompose 𝒮\mathcal{S} as (𝒮{W=0})(𝒮{W=0})(\mathcal{S}\cap\{W=0\})\cup(\mathcal{S}\setminus\{W=0\}). ∎

Theorem 10.5.

Fix ξ0\xi\in\mathbb{R}_{\geq 0}. Assume (2.6) and 2m2\mid m. Assume Conjectures 1.2, 9.6, and 9.8. Also assume Conjecture 1.4 if ξ=0\xi=0. Then Σ(X,𝒮1)F,wX(6m)/4+ξ\Sigma^{\natural}(X,\mathcal{S}_{1})\ll_{F,w}X^{(6-m)/4+\xi}; in fact,

(10.14) 𝒄𝒮1|n1n(1m)/2S𝒄(n)J𝒄,X(n)|F,wX(6m)/4+ξ.\sum_{\bm{c}\in\mathcal{S}_{1}}\,\Bigl{\lvert}{\sum_{n\geq 1}n^{(1-m)/2}S^{\natural}_{\bm{c}}(n)J_{\bm{c},X}(n)}\Bigr{\rvert}\ll_{F,w}X^{(6-m)/4+\xi}.

Moreover, for any X1X\in\mathbb{R}_{\geq 1} and P0,P>0P_{0},P\in\mathbb{R}_{>0}, we have (for some constant η11=η11(F)>0\eta_{11}=\eta_{11}(F)>0)

(10.15) 𝑵17,P0,P(X)d×𝑵𝒕2𝑑𝒕𝒄𝒮1|g𝒄,X,𝑵(i𝒕)|0j1|Σ17,𝑵𝒄,j(𝒕)|F,wX(6m)/4+ξmin(P0P,X)η11,\int_{\bm{N}\in\mathscr{R}_{17,P_{0},P}(X)}d^{\times}\bm{N}\int_{\bm{t}\in\mathbb{R}^{2}}d\bm{t}\sum_{\bm{c}\in\mathcal{S}_{1}}\lvert g_{\bm{c},X,\bm{N}}^{\vee}(i\bm{t})\rvert\prod_{0\leq j\leq 1}\lvert\Sigma_{17,\bm{N}}^{\bm{c},j}(\bm{t})\rvert\ll_{F,w}\frac{X^{(6-m)/4+\xi}}{\min(P_{0}P,X)^{\eta_{11}}},

where 17,P0,P(X)\colonequals{𝐍17(X):N0P0/2,NA0Y/P}\mathscr{R}_{17,P_{0},P}(X)\colonequals\{\bm{N}\in\mathscr{R}_{17}(X):N_{0}\geq P_{0}/2,\;N\leq A_{0}Y/P\}.

Proof.

By (10.2) and (10.5), the bound (10.15) (with P0=1P_{0}=1 and P=1P=1) implies (10.14). And by (10.1), the bound (10.14) implies Σ(X,𝒮1)F,wX(6m)/4+ξ\Sigma^{\natural}(X,\mathcal{S}_{1})\ll_{F,w}X^{(6-m)/4+\xi}. It remains to prove (10.15); for this, we may assume P01P_{0}\geq 1 and P1P\geq 1 (since if P0<1P_{0}<1 or P<1P<1, we may increase P0P_{0} or PP while keeping the set 17,P0,P(X)\mathscr{R}_{17,P_{0},P}(X) the same). To begin, we decompose 𝒮1\mathcal{S}_{1} using (10.7), and then plug in (10.6) (with 𝜶=𝟎\bm{\alpha}=\bm{0}); this bounds the left-hand side of (10.15) by OF,w,b(1)O_{F,w,b}(1) times

(10.16) 17,P0,P(X)d×𝑵2𝑑𝒕C𝒟,λ𝒟2(C)f(𝑵,𝒕,C,λ),\int_{\mathscr{R}_{17,P_{0},P}(X)}d^{\times}\bm{N}\int_{\mathbb{R}^{2}}d\bm{t}\sum_{C\in\mathcal{D},\;\lambda\in\mathcal{D}_{2}(C)}f(\bm{N},\bm{t},C,\lambda),

where we write (for convenience) 𝒟2(C)={λ𝒟2:1/CdegΔλA13}\mathcal{D}_{2}(C)=\{\lambda\in\mathcal{D}_{2}:1/C^{\deg{\Delta}}\leq\lambda\leq A_{13}\} and

(10.17) f(𝑵,𝒕,C,λ)=Σ17,𝑵𝒄,0(𝒕)Σ17,𝑵𝒄,1(𝒕)𝒄1(𝒮1(C,λ))N1/2(N+XC)1m/2(1+𝒕)b(1+C/X1/2)b(1+XCλ/N)b.f(\bm{N},\bm{t},C,\lambda)=\frac{\lVert\Sigma_{17,\bm{N}}^{\bm{c},0}(\bm{t})\cdot\Sigma_{17,\bm{N}}^{\bm{c},1}(\bm{t})\rVert_{\ell^{1}_{\bm{c}}(\mathcal{S}_{1}(C,\lambda))}\cdot N^{-1/2}(N+XC)^{1-m/2}}{(1+\lVert\bm{t}\rVert)^{b}(1+C/X^{1/2})^{b}(1+XC\lambda/N)^{b}}.

Next, note that if ξ=0\xi=0, then Conjecture 7.4 holds by Propositions 6.10 and 7.5. So (10.9), (10.11), (10.13) are all at our disposal, no matter what ξ\xi is. Let θ=110m\theta=\frac{1}{10m} and ϵ=θ100m\epsilon=\frac{\theta}{100m}, say. Suppose 𝑵17(X)\bm{N}\in\mathscr{R}_{17}(X) and b3+max(A14(F,ϵ),ϵ,A15(F,ϵ))b\geq 3+\max(A_{14}(F,\epsilon),\epsilon,A_{15}(F,\epsilon)). Let C𝒟C\in\mathcal{D} and λ𝒟2(C)\lambda\in\mathcal{D}_{2}(C).

Plugging (10.8) into (10.9) (for 𝒮=𝒮1(C,λ)\mathcal{S}=\mathcal{S}_{1}(C,\lambda)) and integrating (10.17) over 𝒕\bm{t} gives

2𝑑𝒕f(𝑵,𝒕,C,λ)[Cm(λ+C1)1/degΔ]1/2ϵ(Cm+N1m/3)1/2+ϵN0m/2+ϵ(N+XC)m/21(1+C/X1/2)b(1+XCλ/N)b.\int_{\mathbb{R}^{2}}d\bm{t}\,f(\bm{N},\bm{t},C,\lambda)\ll\frac{[C^{m}(\lambda+C^{-1})^{1/\deg{\Delta}}]^{1/2-\epsilon}(C^{m}+N_{1}^{m/3})^{1/2+\epsilon}N_{0}^{m/2+\epsilon}}{(N+XC)^{m/2-1}(1+C/X^{1/2})^{b}(1+XC\lambda/N)^{b}}.

Let α=(1/2ϵ)/degΔ\alpha=(1/2-\epsilon)/\deg{\Delta}. Summing over λ𝒟2(C)\lambda\in\mathcal{D}_{2}(C) (using Lemma 5.3 with r=λr=\lambda, τ=XCN\tau=\frac{XC}{N}, a=0a=0, and q{α,0}q\in\{\alpha,0\}, noting that 0α<b0\leq\alpha<b and (λ+C1)αλα+Cα(\lambda+C^{-1})^{\alpha}\ll\lambda^{\alpha}+C^{-\alpha}), we get

λ𝒟2(C)2𝑑𝒕fC(1/2ϵ)m[min(NXC,1)α+C(1ϵ)α](Cm+N1m/3)1/2+ϵN0m/2+ϵ(N+XC)m/21(1+C/X1/2)b.\sum_{\lambda\in\mathcal{D}_{2}(C)}\int_{\mathbb{R}^{2}}d\bm{t}\,f\ll\frac{C^{(1/2-\epsilon)m}[\min(\frac{N}{XC},1)^{\alpha}+C^{-(1-\epsilon)\alpha}](C^{m}+N_{1}^{m/3})^{1/2+\epsilon}N_{0}^{m/2+\epsilon}}{(N+XC)^{m/2-1}(1+C/X^{1/2})^{b}}.

Writing min(NXC,1)NXC\min(\frac{N}{XC},1)\leq\frac{N}{XC} and (N+XC)m/21(XC)m/21(N+XC)^{m/2-1}\geq(XC)^{m/2-1}, and then summing over C𝒟C\in\mathcal{D} using Lemma 5.3 (with r=Cr=C and τ=X1/2\tau=X^{-1/2}, noting that (12ϵ)m>α+(m21)(\frac{1}{2}-\epsilon)m>\alpha+(\frac{m}{2}-1)), we get

C𝒟,λ𝒟2(C)2𝑑𝒕f[(NY)α+X(1ϵ)α/2](Xm/2+X(1/2ϵ)m/2N1(1/2+ϵ)m/3)N0m/2+ϵYm/21,\sum_{C\in\mathcal{D},\;\lambda\in\mathcal{D}_{2}(C)}\int_{\mathbb{R}^{2}}d\bm{t}\,f\ll\frac{[(\frac{N}{Y})^{\alpha}+X^{-(1-\epsilon)\alpha/2}](X^{m/2}+X^{(1/2-\epsilon)m/2}N_{1}^{(1/2+\epsilon)m/3})N_{0}^{m/2+\epsilon}}{Y^{m/2-1}},

since Y=XX1/2Y=X\cdot X^{1/2}. But N1X3/2N_{1}\ll X^{3/2} (by (10.5)), so the right-hand side is [(NY)α+X(1ϵ)α/2]N0m/2+ϵX(6m)/4\ll[(\tfrac{N}{Y})^{\alpha}+X^{-(1-\epsilon)\alpha/2}]\cdot N_{0}^{m/2+\epsilon}\cdot X^{(6-m)/4}. Letting 18\colonequals{𝑵17,P0,P(X):N0m+ϵmin(YN,X(1ϵ)/2)α/2}\mathscr{R}_{18}\colonequals\{\bm{N}\in\mathscr{R}_{17,P_{0},P}(X):N_{0}^{m+\epsilon}\leq\min(\tfrac{Y}{N},X^{(1-\epsilon)/2})^{\alpha/2}\}, and writing N0m/2+ϵN0m/2min(YN,X(1ϵ)/2)α/2N_{0}^{m/2+\epsilon}\leq N_{0}^{-m/2}\cdot\min(\tfrac{Y}{N},X^{(1-\epsilon)/2})^{\alpha/2} for each 𝑵18\bm{N}\in\mathscr{R}_{18}, we get

18d×𝑵C𝒟,λ𝒟2(C)2𝑑𝒕f1/4A0Y/Pd×N[(NY)α/2+X(1ϵ)α/4]P0m/2X(6m)/4,\int_{\mathscr{R}_{18}}d^{\times}\bm{N}\sum_{C\in\mathcal{D},\;\lambda\in\mathcal{D}_{2}(C)}\int_{\mathbb{R}^{2}}d\bm{t}\,f\ll\int_{1/4}^{A_{0}Y/P}d^{\times}{N}\,[(\tfrac{N}{Y})^{\alpha/2}+X^{-(1-\epsilon)\alpha/4}]P_{0}^{-m/2}X^{(6-m)/4},

by integrating over N0P0/2N_{0}\geq P_{0}/2 for each fixed valued of N0N1=NN_{0}N_{1}=N. Thus

(10.18) 18d×𝑵C𝒟,λ𝒟2(C)2𝑑𝒕f[Pα/2+X(12ϵ)α/4]P0m/2X(6m)/4.\int_{\mathscr{R}_{18}}d^{\times}\bm{N}\sum_{C\in\mathcal{D},\;\lambda\in\mathcal{D}_{2}(C)}\int_{\mathbb{R}^{2}}d\bm{t}\,f\ll[P^{-\alpha/2}+X^{-(1-2\epsilon)\alpha/4}]P_{0}^{-m/2}X^{(6-m)/4}.

On the other hand, discarding the factor (1+XCλ/N)b(1+XC\lambda/N)^{b} in (10.17) (using 1+XCλ/N11+XC\lambda/N\geq 1), plugging |𝒮|Cm\lvert\mathcal{S}\rvert\ll C^{m} into (10.11) (for 𝒮=λ𝒟2(C)𝒮1(C,λ)\mathcal{S}=\bigcup_{\lambda\in\mathcal{D}_{2}(C)}\mathcal{S}_{1}(C,\lambda)), and integrating over 𝒕\bm{t} gives

λ𝒟2(C)2𝑑𝒕f(𝑵,𝒕,C,λ)Cm/2(Cm+N0m/3)1/2+ϵNϵ(N+XC)m/21(1+C/X1/2)b.\sum_{\lambda\in\mathcal{D}_{2}(C)}\int_{\mathbb{R}^{2}}d\bm{t}\,f(\bm{N},\bm{t},C,\lambda)\ll\frac{C^{m/2}(C^{m}+N_{0}^{m/3})^{1/2+\epsilon}N^{\epsilon}}{(N+XC)^{m/2-1}(1+C/X^{1/2})^{b}}.

Writing (N+XC)m/21(XC)m/21(N+XC)^{m/2-1}\geq(XC)^{m/2-1}, and then summing over 1CX1/2θ1\leq C\leq X^{1/2-\theta}, we get

C𝒟: 1CX1/2θλ𝒟2(C)2𝑑𝒕f(X1/2θ)m/2((X1/2θ)m+N0m/3)1/2+ϵNϵ(X3/2θ)m/21.\sum_{C\in\mathcal{D}:\,1\leq C\leq X^{1/2-\theta}}\sum_{\lambda\in\mathcal{D}_{2}(C)}\int_{\mathbb{R}^{2}}d\bm{t}\,f\ll\frac{(X^{1/2-\theta})^{m/2}((X^{1/2-\theta})^{m}+N_{0}^{m/3})^{1/2+\epsilon}N^{\epsilon}}{(X^{3/2-\theta})^{m/2-1}}.

But N0m/3Xm/2N_{0}^{m/3}\ll X^{m/2} by (10.5); integrating the previous display over 17(X)\mathscr{R}_{17}(X) thus gives

(10.19) 17(X)d×𝑵C𝒟: 1CX1/2θλ𝒟2(C)2𝑑𝒕fX(6m)/4X(m/2+2)ϵXθX(6m)/4X9θ/10.\int_{\mathscr{R}_{17}(X)}d^{\times}\bm{N}\sum_{C\in\mathcal{D}:\,1\leq C\leq X^{1/2-\theta}}\sum_{\lambda\in\mathcal{D}_{2}(C)}\int_{\mathbb{R}^{2}}d\bm{t}\,f\ll\frac{X^{(6-m)/4}X^{(m/2+2)\epsilon}}{X^{\theta}}\leq\frac{X^{(6-m)/4}}{X^{9\theta/10}}.

Now suppose C>X1/2θC>X^{1/2-\theta}; then C>X1/21/10=X2/5N1/6C>X^{1/2-1/10}=X^{2/5}\gg N^{1/6}. Discarding (1+XCλ/N)b(1+XC\lambda/N)^{b} in (10.17), taking 𝒮=λ𝒟2(C)𝒮1(C,λ)\mathcal{S}=\bigcup_{\lambda\in\mathcal{D}_{2}(C)}\mathcal{S}_{1}(C,\lambda) in (10.13), and integrating over 𝒕\bm{t} gives

λ𝒟2(C)2𝑑𝒕f(𝑵,𝒕,C,λ)(CN1)ξC(1/2ϵ)m(Cm+Nm/3)1/2+ϵ(N+XC)1m/2min(C9/20,N04η10(F,2)/5)(1+C/X1/2)b.\sum_{\lambda\in\mathcal{D}_{2}(C)}\int_{\mathbb{R}^{2}}d\bm{t}\,f(\bm{N},\bm{t},C,\lambda)\ll\frac{(CN_{1})^{\xi}\cdot C^{(1/2-\epsilon)m}(C^{m}+N^{m/3})^{1/2+\epsilon}\cdot(N+XC)^{1-m/2}}{\min(C^{9/20},N_{0}^{4\eta_{10}(F,2)/5})(1+C/X^{1/2})^{b}}.

Writing (N+XC)1m/2(XC)1m/2(N+XC)^{1-m/2}\leq(XC)^{1-m/2}, and summing over CC using Lemma 5.3, gives

C𝒟:C>X1/2θλ𝒟2(C)2𝑑𝒕f(X1/2N1)ξX(1/2ϵ)m/2(Xm/2+Nm/3)1/2+ϵY1m/2min(X9/50,N04η10(F,2)/5).\sum_{C\in\mathcal{D}:\,C>X^{1/2-\theta}}\sum_{\lambda\in\mathcal{D}_{2}(C)}\int_{\mathbb{R}^{2}}d\bm{t}\,f\ll\frac{(X^{1/2}N_{1})^{\xi}\cdot X^{(1/2-\epsilon)m/2}(X^{m/2}+N^{m/3})^{1/2+\epsilon}\cdot Y^{1-m/2}}{\min(X^{9/50},N_{0}^{4\eta_{10}(F,2)/5})}.

Here N1,NX3/2N_{1},N\ll X^{3/2}, so the right-hand side is X2ξX(6m)/4(X9/50+N04η10(F,2)/5)\ll X^{2\xi}\cdot X^{(6-m)/4}\cdot(X^{-9/50}+N_{0}^{-4\eta_{10}(F,2)/5}). Let 19\colonequals17,P0,P(X)18\mathscr{R}_{19}\colonequals\mathscr{R}_{17,P_{0},P}(X)\setminus\mathscr{R}_{18}. Each 𝑵19\bm{N}\in\mathscr{R}_{19} satisfies N0m+ϵ>min(YN,X(1ϵ)/2)α/2N_{0}^{m+\epsilon}>\min(\tfrac{Y}{N},X^{(1-\epsilon)/2})^{\alpha/2}, and thus N0η10(F,2)<min(YN,X(1ϵ)/2)β(NY)β+X(1ϵ)β/2N_{0}^{-\eta_{10}(F,2)}<\min(\tfrac{Y}{N},X^{(1-\epsilon)/2})^{-\beta}\leq(\tfrac{N}{Y})^{\beta}+X^{-(1-\epsilon)\beta/2}, where β=η10(F,2)α2(m+ϵ)\beta=\frac{\eta_{10}(F,2)\alpha}{2(m+\epsilon)}. Thus the integral 19d×𝑵\int_{\mathscr{R}_{19}}d^{\times}\bm{N} of (the left-hand side of) the previous display is

17,P0,P(X)d×𝑵X2ξX(6m)/4(X9/50+N02η10(F,2)/5[(NY)2β/5+X(1ϵ)β/5]),\ll\int_{\mathscr{R}_{17,P_{0},P}(X)}d^{\times}\bm{N}\,X^{2\xi}\cdot X^{(6-m)/4}\cdot(X^{-9/50}+N_{0}^{-2\eta_{10}(F,2)/5}[(\tfrac{N}{Y})^{2\beta/5}+X^{-(1-\epsilon)\beta/5}]),

which is (Xϵ9/50+P02η10(F,2)/5[P2β/5+X(12ϵ)β/5])X(6m)/4+2ξ\ll(X^{\epsilon-9/50}+P_{0}^{-2\eta_{10}(F,2)/5}[P^{-2\beta/5}+X^{-(1-2\epsilon)\beta/5}])\cdot X^{(6-m)/4+2\xi} (by integrating first over N0P0/2N_{0}\geq P_{0}/2 when NN is fixed, and then integrating over NA0Y/PN\leq A_{0}Y/P). This, when combined with (10.18) and (10.19), establishes (10.15) with η11=min(α2,0.9α4,9θ10,17100,2η10(F,2)5,2β5,0.9β5)\eta_{11}=\min(\frac{\alpha}{2},\frac{0.9\alpha}{4},\frac{9\theta}{10},\frac{17}{100},\frac{2\eta_{10}(F,2)}{5},\frac{2\beta}{5},\frac{0.9\beta}{5}) (where θ=110m\theta=\frac{1}{10m}, α0.4degΔ\alpha\geq\frac{0.4}{\deg{\Delta}}, and βη10(F,2)10m\beta\geq\frac{\eta_{10}(F,2)}{10m}), after replacing ξ\xi with ξ/2\xi/2 if ξ>0\xi>0. ∎

Remark 10.6.

Our use of Hölder above is fairly uniform (each of (10.9), (10.11), (10.13) being based on “approximately Cauchy–Schwarz”), but the input required could maybe be slightly relaxed by applying Hölder with more varied exponents. For instance, if CX1/2C\approx X^{1/2} and N0XδN_{0}\geq X^{\delta}, one could work with Σ17,𝑵𝒄,0(𝒕)\Sigma_{17,\bm{N}}^{\bm{c},0}(\bm{t}) in 1\ell^{1} and Σ17,𝑵𝒄,1(𝒕)\Sigma_{17,\bm{N}}^{\bm{c},1}(\bm{t}) in \ell^{\infty} (the N0δN_{0}^{\delta} saving from the former drowning out the XϵX^{\epsilon} loss from the latter); cf. (10.13). And if N0XδN_{0}\leq X^{\delta}, one might hope to work with Σ17,𝑵𝒄,1(𝒕)\Sigma_{17,\bm{N}}^{\bm{c},1}(\bm{t}) in 1\ell^{1} over 𝒄\bm{c} in a residue class to modulus n0N0n_{0}\asymp N_{0} (with some nontrivial archimedean restrictions on 𝒄\bm{c}), and then work with Σ17,𝑵𝒄,0(𝒕)\Sigma_{17,\bm{N}}^{\bm{c},0}(\bm{t}) in 1\ell^{1} afterwards.

Proof of Theorem 1.3.

Let 𝒟={1,2,4,8,}\mathcal{D}=\{1,2,4,8,\ldots\}. Let T0,X(θ)\colonequals|x|Xe(θx3)T_{0,X}(\theta)\colonequals\sum_{\lvert x\rvert\leq X}e(\theta x^{3}) for θ\theta\in\mathbb{R}; then

(10.20) NF(X)=[0,1]𝑑θ|T0,X(θ)|6=T0,X(θ)6Lθ1([0,1]).N_{F}(X)=\int_{[0,1]}d\theta\,\lvert T_{0,X}(\theta)\rvert^{6}=\lVert T_{0,X}(\theta)^{6}\rVert_{L^{1}_{\theta}([0,1])}.

Now let T1,X(θ)\colonequals|x|(X/2,X]e(θx3)T_{1,X}(\theta)\colonequals\sum_{\lvert x\rvert\in(X/2,X]}e(\theta x^{3}). Let f(λ)λq(𝒟)\colonequals(λ𝒟|f(λ)|q)1/q\lVert f(\lambda)\rVert_{\ell^{q}_{\lambda}(\mathcal{D})}\colonequals(\sum_{\lambda\in\mathcal{D}}\lvert f(\lambda)\rvert^{q})^{1/q} for q1q\geq 1; then

(10.21) T0,X(θ)=1+λ𝒟T1,X/λ(θ)1+(X/λ)ρλ6/5(𝒟)(X/λ)ρT1,X/λ(θ)λ6(𝒟)T_{0,X}(\theta)=1+\sum_{\lambda\in\mathcal{D}}T_{1,X/\lambda}(\theta)\ll 1+\lVert(X/\lambda)^{\rho}\rVert_{\ell^{6/5}_{\lambda}(\mathcal{D})}\cdot\lVert(X/\lambda)^{-\rho}T_{1,X/\lambda}(\theta)\rVert_{\ell^{6}_{\lambda}(\mathcal{D})}

(by Hölder), where ρ=1/100\rho=1/100, say. Since (X/λ)ρλ6/5(𝒟)Xρ\lVert(X/\lambda)^{\rho}\rVert_{\ell^{6/5}_{\lambda}(\mathcal{D})}\ll X^{\rho}, we deduce (upon inserting (10.21) into (10.20)) that if T1,X/λ(θ)6Lθ1([0,1])(X/λ)3\lVert T_{1,X/\lambda}(\theta)^{6}\rVert_{L^{1}_{\theta}([0,1])}\ll(X/\lambda)^{3} holds, then (since 6ρ<36\rho<3)

(10.22) NF(X)1+X6ρ(X/λ)6ρT1,X/λ(θ)6L(λ,θ)1(𝒟×[0,1])X3.N_{F}(X)\ll 1+X^{6\rho}\cdot\lVert(X/\lambda)^{-6\rho}T_{1,X/\lambda}(\theta)^{6}\rVert_{L^{1}_{(\lambda,\theta)}(\mathcal{D}\times[0,1])}\ll X^{3}.

Now let w(𝒙)\colonequals1j6𝟏|xj|(1/2,1]w_{\star}(\bm{x})\colonequals\prod_{1\leq j\leq 6}\bm{1}_{\lvert x_{j}\rvert\in(1/2,1]}, and choose wCc(({0})6)w\in C^{\infty}_{c}((\mathbb{R}\setminus\{0\})^{6}) with www\geq w_{\star}. Then

(10.23) T1,X(θ)6Lθ1([0,1])=NF,w(X)NF,w(X).\lVert T_{1,X}(\theta)^{6}\rVert_{L^{1}_{\theta}([0,1])}=N_{F,w_{\star}}(X)\leq N_{F,w}(X).

But ww satisfies (1.11) (and thus (2.6)). And we are assuming (for Theorem 1.3) Conjectures 1.2, 1.4, and 1.5; in particular, Conjectures 9.6 and 9.8 hold by Propositions 9.7 and 9.9, respectively. So Theorem 10.5 (with ξ=0\xi=0) gives Σ(X,𝒮1)1\Sigma^{\natural}(X,\mathcal{S}_{1})\ll 1. But Σ(X,𝒮0)1\Sigma^{\natural}(X,\mathcal{S}_{0})\ll 1 by Theorem 2.5. So by (2.10) (and the definitions (2.15), (1.6)), we have

NF,w(X)/X3=OA(YA)+Σ(X,𝒮0)+Σ(X,𝒮1)1.N_{F,w}(X)/X^{3}=O_{A}(Y^{-A})+\Sigma^{\natural}(X,\mathcal{S}_{0})+\Sigma^{\natural}(X,\mathcal{S}_{1})\ll 1.

Hence NF(X)X3N_{F}(X)\ll X^{3} by (10.23), (10.22). Given (1.8), a standard Cauchy–Schwarz argument then leads to the desired application to F0(S3)F_{0}(S^{3}) (producing a “positive lower density”). ∎

10.3. Delta cancellation

To prove Theorems 1.6 and 1.9, we will complement (10.15) using Propositions 7.15 and 7.16, identifying 𝒄\bm{c}-cancellation in some pieces of Σ(X,𝒮1)\Sigma^{\natural}(X,\mathcal{S}_{1}). Since J𝒄,X(n)J_{\bm{c},X}(n) is not compactly supported in 𝒄\bm{c}, we begin with a decomposition resembling (8.18).

Recall ν0\nu_{0}, ν1\nu_{1} from §1.2. Let ν¯0\colonequals1ν0\overline{\nu}_{0}\colonequals 1-\nu_{0}. For all κ>0\kappa\in\mathbb{R}_{>0}, let (cf. (8.16), (8.17))

(10.24) Σ18,0(X,𝑵,𝒕)\displaystyle\Sigma_{18,0}(X,\bm{N},\bm{t}) \colonequals𝒄𝒮1ν0(𝒄/X1/2)g𝒄,X,𝑵(i𝒕)0j1Σ17,𝑵𝒄,j(𝒕),\displaystyle\colonequals\sum_{\bm{c}\in\mathcal{S}_{1}}\nu_{0}(\bm{c}/X^{1/2})g_{\bm{c},X,\bm{N}}^{\vee}(i\bm{t})\prod_{0\leq j\leq 1}\Sigma_{17,\bm{N}}^{\bm{c},j}(\bm{t}),
(10.25) Σ18,1,κ(X,𝑵,𝒕)\displaystyle\Sigma_{18,1,\kappa}(X,\bm{N},\bm{t}) \colonequals𝒄𝒮1ν¯0(𝒄/X1/2)ν1(𝒄/κ)g𝒄,X,𝑵(i𝒕)0j1Σ17,𝑵𝒄,j(𝒕).\displaystyle\colonequals\sum_{\bm{c}\in\mathcal{S}_{1}}\overline{\nu}_{0}(\bm{c}/X^{1/2})\nu_{1}(\bm{c}/\kappa)g_{\bm{c},X,\bm{N}}^{\vee}(i\bm{t})\prod_{0\leq j\leq 1}\Sigma_{17,\bm{N}}^{\bm{c},j}(\bm{t}).

We have Σ18,1,κ(X,𝑵,𝒕)=0\Sigma_{18,1,\kappa}(X,\bm{N},\bm{t})=0 unless there exists 𝒄𝒮1\bm{c}\in\mathcal{S}_{1} satisfying X1/2/2<𝒄mκX^{1/2}/2<\lVert\bm{c}\rVert\leq m\kappa. Using κ>0d×κν1(𝒄/κ)=1\int_{\kappa>0}d^{\times}{\kappa}\,\nu_{1}(\bm{c}/\kappa)=1 (valid for 𝒄m{𝟎}\bm{c}\in\mathbb{R}^{m}\setminus\{\bm{0}\}), we may thus write (cf. (8.18))

(10.26) 𝒄𝒮1g𝒄,X,𝑵(i𝒕)0j1Σ17,𝑵𝒄,j(𝒕)=Σ18,0(X,𝑵,𝒕)+X1/2/2md×κΣ18,1,κ(X,𝑵,𝒕).\sum_{\bm{c}\in\mathcal{S}_{1}}g_{\bm{c},X,\bm{N}}^{\vee}(i\bm{t})\prod_{0\leq j\leq 1}\Sigma_{17,\bm{N}}^{\bm{c},j}(\bm{t})=\Sigma_{18,0}(X,\bm{N},\bm{t})+\int_{X^{1/2}/2m}^{\infty}d^{\times}{\kappa}\,\Sigma_{18,1,\kappa}(X,\bm{N},\bm{t}).

Recall 17(X)\mathscr{R}_{17}(X) from (10.5). Let 17P0,P(X)\colonequals{𝑵17(X):N0<P0/2,N>A0Y/P}\mathscr{R}_{17}^{P_{0},P}(X)\colonequals\{\bm{N}\in\mathscr{R}_{17}(X):N_{0}<P_{0}/2,\;N>A_{0}Y/P\} for reals P0,P>0P_{0},P>0. In terms of 17,P0,P(X)\mathscr{R}_{17,P_{0},P}(X) from Theorem 10.5, we have

(10.27) 17P0,P(X)17(X)17P0,P(X)17,1,P(X)17,P0,1(X).\mathscr{R}_{17}^{P_{0},P}(X)\subseteq\mathscr{R}_{17}(X)\subseteq\mathscr{R}_{17}^{P_{0},P}(X)\cup\mathscr{R}_{17,1,P}(X)\cup\mathscr{R}_{17,P_{0},1}(X).

By (10.1), (10.2), (10.5), and (10.27), the bound (10.15) (when applicable) implies

(10.28) Σ(X,𝒮1)17P0,P(X)d×𝑵2𝑑𝒕𝒄𝒮1g𝒄,X,𝑵(i𝒕)0j1Σ17,𝑵𝒄,j(𝒕)X(6m)/4+ξmin(P0,P,X)η11.\Sigma^{\natural}(X,\mathcal{S}_{1})-\int_{\mathscr{R}_{17}^{P_{0},P}(X)}d^{\times}\bm{N}\int_{\mathbb{R}^{2}}d\bm{t}\sum_{\bm{c}\in\mathcal{S}_{1}}g_{\bm{c},X,\bm{N}}^{\vee}(i\bm{t})\prod_{0\leq j\leq 1}\Sigma_{17,\bm{N}}^{\bm{c},j}(\bm{t})\ll\frac{X^{(6-m)/4+\xi}}{\min(P_{0},P,X)^{\eta_{11}}}.

This leads to the following results.

Theorem 10.7.

Assume (2.6) and 2m2\mid m. Assume Conjectures 1.2, 1.4, 1.8, 9.6, and 9.8. Then for X1X\geq 1, we have Σ(X,𝒮1)=oF,w;X(X(6m)/4)\Sigma^{\natural}(X,\mathcal{S}_{1})=o_{F,w;X\to\infty}(X^{(6-m)/4}).

Proof.

Before proceeding, note that Conjecture 7.14 holds by Propositions 6.10, 6.13, and 7.15, since we assume Conjectures 1.2, 1.4, and 1.8. Let MM be a real number to be chosen later.

Let 𝑵17P0,P(X)\bm{N}\in\mathscr{R}_{17}^{P_{0},P}(X). For each real Z2X1/2Z\geq 2X^{1/2}, let (in the context of (7.25))

ν18,0=ν18,0,X,𝑵,𝒕,Z=ν0(Z𝒄/X1/2)gZ𝒄,X,𝑵(i𝒕)ν2(r)\nu_{18,0}=\nu_{18,0,X,\bm{N},\bm{t},Z}=\nu_{0}(Z\bm{c}/X^{1/2})g_{Z\bm{c},X,\bm{N}}^{\vee}(i\bm{t})\nu_{2}(r)

(noting that Suppν0[2,2]m\operatorname{Supp}{\nu_{0}}\subseteq[-2,2]^{m}, so ν18,0\nu_{18,0} is supported on [1,1]m×[1,2][-1,1]^{m}\times[1,2]). Then after plugging (10.4) into (10.24) and decomposing 𝒮1\mathcal{S}_{1} into residue classes modulo n0n_{0}, we get

Σ18,0(X,𝑵,𝒕)=n01ν2(n0/N0)n0it01𝒂n0:n0𝒩𝒂S𝒂(n0)Σ11𝒂,n0(ν18,0,Z,N1)\Sigma_{18,0}(X,\bm{N},\bm{t})=\sum_{n_{0}\geq 1}\nu_{2}(n_{0}/N_{0})n_{0}^{-it_{0}}\sum_{1\leq\bm{a}\leq n_{0}:\,n_{0}\in\mathcal{N}_{\bm{a}}}S^{\natural}_{\bm{a}}(n_{0})\cdot\Sigma_{11}^{\bm{a},n_{0}}(\nu_{18,0},Z,N_{1})

(in terms of Σ11𝒂,n0\Sigma_{11}^{\bm{a},n_{0}} from (7.25)). By Conjecture 7.14 (with ν=ν18,0\nu=\nu_{18,0}, and with 2X1/2+N11/32X^{1/2}+N_{1}^{1/3}, N1N_{1} in place of ZZ, NN) and the trivial bound |S𝒂(n0)|n0(1+m)/2\lvert S^{\natural}_{\bm{a}}(n_{0})\rvert\leq n_{0}^{(1+m)/2}, we get

Σ18,0(X,𝑵,𝒕)N0(3+m)/2Xm/2N11/2(oM(1)+oM;X(1))1,A7(ν18,0),\Sigma_{18,0}(X,\bm{N},\bm{t})\ll N_{0}^{(3+m)/2}X^{m/2}N_{1}^{1/2}(o_{M\to\infty}(1)+o_{M;X\to\infty}(1))\cdot\mathcal{M}_{1,A_{7}}(\nu_{18,0}),

provided min(2X1/2,N1)2M2(2N0)10/9\min(2X^{1/2},N_{1})\geq 2M\geq 2(2N_{0})^{10/9}. Additionally, by (10.6) (with Z𝒄=(2X1/2+N11/3)𝒄Z\bm{c}=(2X^{1/2}+N_{1}^{1/3})\bm{c} in place of 𝒄\bm{c}) and the definition (6.24) of 1,k\mathcal{M}_{1,k} (with ν=ν18,0\nu=\nu_{18,0}, k=A7k=A_{7}), we have

1,A7(N1/2ν18,0)b(X1/2/X1/2+XX1/2/N)1(N+0)1m/2(1+𝒕)b(1+0)b(1+0)b(Y/N)N1m/2(1+𝒕)b.\mathcal{M}_{1,A_{7}}(N^{1/2}\nu_{18,0})\ll_{b}\frac{(X^{1/2}/X^{1/2}+X\cdot X^{1/2}/N)^{1}\cdot(N+0)^{1-m/2}}{(1+\lVert\bm{t}\rVert)^{b}(1+0)^{b}(1+0)^{b}}\ll\frac{(Y/N)\cdot N^{1-m/2}}{(1+\lVert\bm{t}\rVert)^{b}}.

Now, for each κX1/2/2m\kappa\geq X^{1/2}/2m and ZmκZ\geq m\kappa (noting that Suppν1[m,m]m\operatorname{Supp}{\nu_{1}}\subseteq[-m,m]^{m}), let

ν18,1,κ=ν18,1,κ,X,𝑵,𝒕,Z=ν¯0(Z𝒄/X1/2)ν1(Z𝒄/κ)gZ𝒄,X,𝑵(i𝒕)ν2(r).\nu_{18,1,\kappa}=\nu_{18,1,\kappa,X,\bm{N},\bm{t},Z}=\overline{\nu}_{0}(Z\bm{c}/X^{1/2})\nu_{1}(Z\bm{c}/\kappa)g_{Z\bm{c},X,\bm{N}}^{\vee}(i\bm{t})\nu_{2}(r).

Applying Conjecture 7.14 with mκ+N11/3m\kappa+N_{1}^{1/3}, N1N_{1} in place of ZZ, NN, we get

Σ18,1,κ(X,𝑵,𝒕)N0(3+m)/2κmN11/2(oM(1)+oM;X(1))1,A7(ν18,1,κ),\Sigma_{18,1,\kappa}(X,\bm{N},\bm{t})\ll N_{0}^{(3+m)/2}\kappa^{m}N_{1}^{1/2}(o_{M\to\infty}(1)+o_{M;X\to\infty}(1))\cdot\mathcal{M}_{1,A_{7}}(\nu_{18,1,\kappa}),

provided min(X1/2/2,N1)2M(2N0)10/9\min(X^{1/2}/2,N_{1})\geq 2M\geq(2N_{0})^{10/9}. This time, since 𝟎Suppν1\bm{0}\notin\operatorname{Supp}{\nu_{1}}, the bound (10.6) (with (mκ+N11/3)𝒄(m\kappa+N_{1}^{1/3})\bm{c} in place of 𝒄\bm{c}) and (6.24) (with ν=ν18,1,κ\nu=\nu_{18,1,\kappa}, k=A7k=A_{7}) give

1,A7(N1/2ν18,1,κ)b(κ/X1/2+Xκ/N)1(N+Xκ)1m/2(1+𝒕)b(1+κ/X1/2)b(1+0)b(Y/N)(Xκ)1m/2(1+𝒕)b(κ/X1/2)b1.\mathcal{M}_{1,A_{7}}(N^{1/2}\nu_{18,1,\kappa})\ll_{b}\frac{(\kappa/X^{1/2}+X\kappa/N)^{1}\cdot(N+X\kappa)^{1-m/2}}{(1+\lVert\bm{t}\rVert)^{b}(1+\kappa/X^{1/2})^{b}(1+0)^{b}}\ll\frac{(Y/N)\cdot(X\kappa)^{1-m/2}}{(1+\lVert\bm{t}\rVert)^{b}(\kappa/X^{1/2})^{b-1}}.

Inserting our bounds on Σ18,0\Sigma_{18,0}, Σ18,1,κ\Sigma_{18,1,\kappa} into (10.26), assuming b3+m/2b\geq 3+m/2, we find that if min(X1/2/2,N1)2M(2N0)10/9\min(X^{1/2}/2,N_{1})\geq 2M\geq(2N_{0})^{10/9}, then the left-hand side of (10.26) is

(10.29) N0(2+m)/2Xm/2(oM(1)+oM;X(1))(Y/N)N1m/2/(1+𝒕)b.\ll N_{0}^{(2+m)/2}X^{m/2}(o_{M\to\infty}(1)+o_{M;X\to\infty}(1))\cdot(Y/N)\cdot N^{1-m/2}/(1+\lVert\bm{t}\rVert)^{b}.

Since N0<P0/2N_{0}<P_{0}/2 and N1=N/N0>A0Y/P0PN_{1}=N/N_{0}>A_{0}Y/P_{0}P, we conclude (upon integrating over 𝒕2\bm{t}\in\mathbb{R}^{2}, then over N0<P0/2N_{0}<P_{0}/2 with N0N1=NN_{0}N_{1}=N fixed, and finally over A0Y/P<NA0YA_{0}Y/P<N\leq A_{0}Y) that

17P0,P(X)d×𝑵2𝑑𝒕𝒄𝒮1g𝒄,X,𝑵(i𝒕)0j1Σ17,𝑵𝒄,j(𝒕)P0(2+m)/2Xm/2oM(1)Y(Y/P)m/2,\int_{\mathscr{R}_{17}^{P_{0},P}(X)}d^{\times}\bm{N}\int_{\mathbb{R}^{2}}d\bm{t}\sum_{\bm{c}\in\mathcal{S}_{1}}g_{\bm{c},X,\bm{N}}^{\vee}(i\bm{t})\prod_{0\leq j\leq 1}\Sigma_{17,\bm{N}}^{\bm{c},j}(\bm{t})\ll\frac{P_{0}^{(2+m)/2}X^{m/2}o_{M\to\infty}(1)\cdot Y}{(Y/P)^{m/2}},

provided XM,P0,P1X\gg_{M,P_{0},P}1 and MP01M\gg_{P_{0}}1 hold with large enough implied constants. Therefore, there exist functions f3,f4:>01f_{3},f_{4}\colon\mathbb{R}_{>0}\to\mathbb{Z}_{\geq 1} such that if Mf3(P0+P)M\geq f_{3}(P_{0}+P) and Xf4(M)X\geq f_{4}(M), then the left-hand side of the previous display is oP0+P(Y1m/2Xm/2)o_{P_{0}+P\to\infty}(Y^{1-m/2}X^{m/2}). It follows from (10.28) (with ξ=0\xi=0) that if Xf4(f3(P0+P))X\geq f_{4}(f_{3}(P_{0}+P)), then Σ(X,𝒮1)=omin(P0,P,X)(X(6m)/4)\Sigma^{\natural}(X,\mathcal{S}_{1})=o_{\min(P_{0},P,X)\to\infty}(X^{(6-m)/4}). Finally, suppose Xf4(f3(2))X\geq f_{4}(f_{3}(2)), let PP be the largest integer in [1,X][1,X] with f4(f3(2P))Xf_{4}(f_{3}(2P))\leq X (so that PP\to\infty as XX\to\infty), and take P0=PP_{0}=P, to get Σ(X,𝒮1)=oX(X(6m)/4)\Sigma^{\natural}(X,\mathcal{S}_{1})=o_{X\to\infty}(X^{(6-m)/4}). ∎

Proof of Theorem 1.6.

Unconditionally, by (1.3), (2.10), and (2.16), we have

(10.30) EF,w(X)/X3=O(X2.75+ϵ)/X3+Σ(X,𝒮1).E_{F,w}(X)/X^{3}=O(X^{2.75+\epsilon})/X^{3}+\Sigma^{\natural}(X,\mathcal{S}_{1}).

Now assume (1.11) (so (2.6) holds). Then Theorem 10.7 gives Σ(X,𝒮1)=oX(X(6m)/4)\Sigma^{\natural}(X,\mathcal{S}_{1})=o_{X\to\infty}(X^{(6-m)/4}), since Conjectures 9.6 and 9.8 hold by Propositions 9.7 and 9.9, respectively (since we assume Conjecture 1.5). Upon plugging this bound into (10.30), we get (1.5).

The Hasse principle for VV follows from (1.5), upon choosing ww with σ,F,w>0\sigma_{\infty,F,w}>0 (possible since V()V(\mathbb{R}) contains a point with x1x60x_{1}\cdots x_{6}\neq 0). Now suppose F=x13++x63F=x_{1}^{3}+\dots+x_{6}^{3}. Then by [wang2023prime]*Theorem 1.1 (or [wang2022thesis]*Theorem 2.1.8), we find (from (1.5)) that F0(3)F_{0}(\mathbb{Z}^{3}) has density 11 in {a:a±4mod9}\{a\in\mathbb{Z}:a\not\equiv\pm 4\bmod{9}\}. (See [wang2023prime] for details on this last deduction, which is based on [diaconu2019admissible]. Diaconu assumes an analog of (1.5) over rather quantitatively deformed regions RR^{\ast}, whereas we work with fixed weights ww. It would be very interesting to see if there could be any miraculous cancellation or symmetries in the analog of J𝒄,X(n)J_{\bm{c},X}(n) over RR^{\ast}, but at the moment it seems easier to handle minimally deformed regions.) ∎

Proof of Corollary 1.7.

(Here we drop the assumption (2.6).) Let ρ>0\rho>0 be a parameter tending to 0 slowly as XX\to\infty. Use (1.8) and Hölder’s inequality to upper bound the contribution to NF,w(X)N_{F,w}(X) from points 𝒙m\bm{x}\in\mathbb{Z}^{m} with min(|x1|,,|xm|)ρX\min(\lvert x_{1}\rvert,\dots,\lvert x_{m}\rvert)\leq\rho\cdot X. Then use Theorem 1.6 to estimate the remaining contribution to NF,w(X)N_{F,w}(X). This gives (1.5) for arbitrary wCc(m)w\in C^{\infty}_{c}(\mathbb{R}^{m}). Hooley’s conjecture follows upon taking a suitable sequence of weights ww. ∎

Theorem 10.8.

Assume (2.6) and 2m2\mid m. Assume Conjectures 1.2, 1.10, 1.11, 9.6, and 9.8. Then for some constant η12=η12(F)>0\eta_{12}=\eta_{12}(F)>0, we have Σ(X,𝒮1)F,wX(6m)/4η12\Sigma^{\natural}(X,\mathcal{S}_{1})\ll_{F,w}X^{(6-m)/4-\eta_{12}}.

Proof.

Proposition 7.16 applies, since we assume Conjectures 1.2, 1.10, 1.11. We now mimic the proof of Theorem 10.7, while replacing each use of Conjecture 7.14 with Proposition 7.16.

Let 𝑵17P0,P(X)\bm{N}\in\mathscr{R}_{17}^{P_{0},P}(X). Proposition 7.16 implies that for any real M[1,Xη2/2]M\in[1,X^{\eta_{2}/2}] satisfying min(2X1/2,N1)2M2(2N0)2\min(2X^{1/2},N_{1})\geq 2M\geq 2(2N_{0})^{2}, we have

Σ18,0(X,𝑵,𝒕)ϵN0(3+m)/2Xm/2+ϵN11/2(M1/6degHN11/6+Mη5/4degH)1,A8(ν18,0).\Sigma_{18,0}(X,\bm{N},\bm{t})\ll_{\epsilon}N_{0}^{(3+m)/2}X^{m/2+\epsilon}N_{1}^{1/2}(M^{1/6\deg{H}}N_{1}^{-1/6}+M^{-\eta_{5}/4\deg{H}})\mathcal{M}_{1,A_{8}}(\nu_{18,0}).

Furthermore, for each κX1/2/2m\kappa\geq X^{1/2}/2m, Proposition 7.16 implies that

Σ18,1,κ(X,𝑵,𝒕)ϵN0(3+m)/2κm+ϵN11/2(M1/6degHN11/6+Mη5/4degH)1,A8(ν18,1,κ).\Sigma_{18,1,\kappa}(X,\bm{N},\bm{t})\ll_{\epsilon}N_{0}^{(3+m)/2}\kappa^{m+\epsilon}N_{1}^{1/2}(M^{1/6\deg{H}}N_{1}^{-1/6}+M^{-\eta_{5}/4\deg{H}})\mathcal{M}_{1,A_{8}}(\nu_{18,1,\kappa}).

for any real M[1,(X1/2/2)η2]M\in[1,(X^{1/2}/2)^{\eta_{2}}] satisfying min(X1/2/2,N1)2M2(2N0)2\min(X^{1/2}/2,N_{1})\geq 2M\geq 2(2N_{0})^{2}.

Now let M=Xmin(1,η2)/2/(4+2η2)M=X^{\min(1,\eta_{2})/2}/(4+2^{\eta_{2}}), and suppose

(10.31) (A0Y/P0P)1/22M2P02,(A_{0}Y/P_{0}P)^{1/2}\geq 2M\geq 2P_{0}^{2},

so that N1M2N_{1}\gg M^{2} and (thus) M1/6degHN11/6+Mη5/4degHMmin(1,η5)/6degHM^{1/6\deg{H}}N_{1}^{-1/6}+M^{-\eta_{5}/4\deg{H}}\ll M^{-\min(1,\eta_{5})/6\deg{H}}. Arguing as we did for (10.29), we find (assuming b3+m/2b\geq 3+m/2) that the left-hand side of (10.26) is

ϵN0(2+m)/2Xm/2+ϵ(Mmin(1,η5)/6degH)(Y/N)N1m/2/(1+𝒕)b.\ll_{\epsilon}N_{0}^{(2+m)/2}X^{m/2+\epsilon}(M^{-\min(1,\eta_{5})/6\deg{H}})\cdot(Y/N)\cdot N^{1-m/2}/(1+\lVert\bm{t}\rVert)^{b}.

So (upon integrating over 𝒕\bm{t}, then over N0N_{0}, and finally over NN)

17P0,P(X)d×𝑵2𝑑𝒕𝒄𝒮1g𝒄,X,𝑵(i𝒕)0j1Σ17,𝑵𝒄,j(𝒕)ϵP0(2+m)/2Xm/2+ϵYMmin(1,η5)/6degH(Y/P)m/2.\int_{\mathscr{R}_{17}^{P_{0},P}(X)}d^{\times}\bm{N}\int_{\mathbb{R}^{2}}d\bm{t}\sum_{\bm{c}\in\mathcal{S}_{1}}g_{\bm{c},X,\bm{N}}^{\vee}(i\bm{t})\prod_{0\leq j\leq 1}\Sigma_{17,\bm{N}}^{\bm{c},j}(\bm{t})\ll_{\epsilon}\frac{P_{0}^{(2+m)/2}X^{m/2+\epsilon}\cdot Y}{M^{\min(1,\eta_{5})/6\deg{H}}\cdot(Y/P)^{m/2}}.

Now let γ=min(1,η5)/12degH(0,1/12]\gamma=\min(1,\eta_{5})/12\deg{H}\in(0,1/12] and P0=Mγ/(2+m)P_{0}=M^{\gamma/(2+m)}, P=A0Mγ/m/4P=A_{0}M^{\gamma/m}/4. Then (10.31) holds (since γ1/2\gamma\leq 1/2 and YM3Y\geq M^{3}), and the right-hand side of the previous display is Mγ/2+γ/22γXm/2+ϵY1m/2=MγX(6m)/4+ϵ\ll M^{\gamma/2+\gamma/2-2\gamma}X^{m/2+\epsilon}Y^{1-m/2}=M^{-\gamma}X^{(6-m)/4+\epsilon}. So by (10.28) (with ξ=ϵ>0\xi=\epsilon>0) we have

Σ(X,𝒮1)ϵX(6m)/4+ϵ(Mγ+P0η11)X(6m)/4η12,\Sigma^{\natural}(X,\mathcal{S}_{1})\ll_{\epsilon}X^{(6-m)/4+\epsilon}(M^{-\gamma}+P_{0}^{-\eta_{11}})\ll X^{(6-m)/4-\eta_{12}},

where η12=910min(1,η11/(2+m))γmin(1,η2)/2\eta_{12}=\frac{9}{10}\cdot\min(1,\eta_{11}/(2+m))\cdot\gamma\cdot\min(1,\eta_{2})/2, say. ∎

Proof of Theorem 1.9.

Proceed as in the proof of Theorem 10.7, but use Theorem 10.8 instead of Theorem 10.7, to get (from (10.30)) the bound EF,w(X)/X3ϵX0.25+ϵ+Xη12E_{F,w}(X)/X^{3}\ll_{\epsilon}X^{-0.25+\epsilon}+X^{-\eta_{12}}. ∎

Acknowledgements

Many thanks to Amit Ghosh and Peter Sarnak for suggesting the main problem addressed here. I also thank my advisor, Peter Sarnak, for his guidance and encouragement.This work was partially supported by NSF grant DMS-1802211. I thank Calvin Deng and Yotam Hendel for sharing references for (4.1) and (4.2), and Yotam for discussions related to §9. For early comments, I thank Manjul Bhargava, Andy Booker, Tim Browning, Brian Conrey, Simona Diaconu, Bill Duke, Roger Heath-Brown, Nick Katz, Will Sawin, Bob Vaughan, and Trevor Wooley. Thanks also to everyone listed in [wang2022thesis], and to people from seminars, conferences, and other events. For more recent interactions, I thank Louis-Pierre Arguin, Emma Bailey, Paul Bourgade, Alex Gamburd, Jayce Getz, Jeff Hoffstein, Junehyuk Jung, Victor Kolyvagin, Valeriya Kovaleva, Lillian Pierce, Kannan Soundararajan, Yuri Tschinkel, Katy Woo, Max Xu, Liyang Yang, and Peter Zenz. Finally, I thank my family for their exceptional support.

References