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

Special cubic zeros and the dual variety

Victor Y. Wang Fine Hall, 304 Washington Road, Princeton, NJ 08540, USA Courant Institute, 251 Mercer Street, New York, NY 10012, USA IST Austria, Am Campus 1, 3400 Klosterneuburg, Austria [email protected]
Abstract.

Let FF be a diagonal cubic form over \mathbb{Z} in 66 variables. From the dual variety in the delta method of Duke–Friedlander–Iwaniec and Heath-Brown, we unconditionally extract a weighted count of certain special integral zeros of FF in regions of diameter XX\to\infty. Heath-Brown did the same in 44 variables, but our analysis differs and captures some novel features. We also put forth an axiomatic framework for more general FF.

Key words and phrases:
Cubic form, circle method, special subvarieties, biases, dual variety
1991 Mathematics Subject Classification:
Primary 11D45; Secondary 11D25, 11G25, 11P55, 14B05

1. Introduction

Let F[𝒙]=[x1,,xm]F\in\mathbb{Z}[\bm{x}]=\mathbb{Z}[x_{1},\dots,x_{m}] be a cubic form in m4m\geq 4 variables. Let VV be the hypersurface F=0F=0 in m1\mathbb{P}^{m-1}_{\mathbb{Q}}. Let w:mw\colon\mathbb{R}^{m}\to\mathbb{R} be a smooth, compactly supported function. Let Suppw\operatorname{Supp}{w} be the closure of the set {𝒙m:w(𝒙)0}\{\bm{x}\in\mathbb{R}^{m}:w(\bm{x})\neq 0\} in m\mathbb{R}^{m}. Assume that VV is smooth and (for convenience) that 𝟎Suppw\bm{0}\notin\operatorname{Supp}w. For reals X1X\geq 1, let

(1.1) NF,w(X)\colonequals𝒙m:F(𝒙)=0w(𝒙/X).N_{F,w}(X)\colonequals\sum_{\bm{x}\in\mathbb{Z}^{m}:\,F(\bm{x})=0}w(\bm{x}/X).

Let Υ\Upsilon be the set of vector spaces LmL\subseteq\mathbb{Q}^{m} over \mathbb{Q} of dimension m/2\lfloor m/2\rfloor with F|L=0F|_{L}=0. The “Hardy–Littlewood model” for NF,w(X)N_{F,w}(X) can fail via Υ\Upsilon if m6m\leq 6.111For singular VV, a similar failure can occur for larger mm; see [brudern2019instance] for a nice example with m=8m=8. But based on [hooley1986some]*Conjecture 2, [franke1989rational], [vaughan1995certain]*Appendix, [peyre1995hauteurs], et al., one may conjecture

(1.2) NF,w(X)𝒙m:𝒙LΥLw(𝒙/X)=(cF,w+oX(1))Xm3(logX)r1+𝟏m=4N_{F,w}(X)-\sum_{\bm{x}\in\mathbb{Z}^{m}:\,\bm{x}\in\bigcup_{L\in\Upsilon}L}w(\bm{x}/X)=(c_{F,w}+o_{X\to\infty}(1))\cdot X^{m-3}(\log{X})^{r-1+\bm{1}_{m=4}}

for a certain predicted constant cF,wc_{F,w}\in\mathbb{R} and integer r1r\geq 1, where r=1r=1 if m5m\geq 5. Here rr is the rank of the Picard group of VV.

For further discussion and references on (1.2), see [bombieri2009problems]*§2 or [wang2022thesis]*§6.5. The natural analog of (1.2) for rational points is closely related to (1.2), and is in fact equivalent when m5m\geq 5, but the precise sieve-theoretic relationship is delicate when m=4m=4; cf. [heath1996new]*paragraph preceding Corollary 2 and [heath1996new]*Theorem 8 versus Corollary 2 on quadrics.

In the present paper, we seek to organically detect, within the delta method of [duke1993bounds, heath1996new], the locus LΥL\bigcup_{L\in\Upsilon}L featured in (1.2). In the delta method, as in [kloosterman1926representation]’s circle method, one averages over numerators to a given denominator (modulus), and often then invokes Poisson summation. Set Y\colonequalsX3/2Y\colonequals X^{3/2} as on [heath1998circle]*p. 676; then

(1.3) (1+OA(YA))NF,w(X)=Y2n1𝒄mnmS𝒄(n)I𝒄(n)(1+O_{A}(Y^{-A}))\cdot N_{F,w}(X)=Y^{-2}\sum_{n\geq 1}\sum_{\bm{c}\in\mathbb{Z}^{m}}n^{-m}S_{\bm{c}}(n)I_{\bm{c}}(n)

(for all A>0A>0) by [heath1996new]*Theorem 2, (1.2), up to easy manipulations from §3, where

(1.4) I𝒄(n)\colonequals𝒙m𝑑𝒙w(𝒙/X)h(n/Y,F(𝒙)/Y2)e(𝒄𝒙/n)I_{\bm{c}}(n)\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)

for a certain fixed smooth function h:(0,)×h\colon(0,\infty)\times\mathbb{R}\to\mathbb{R} usually left in the background (see [heath1998circle]*(2.3) for the precise definition of hh), and where

(1.5) S𝒄(n)\colonequals1an:gcd(a,n)=11𝒙nen(aF(𝒙)+𝒄𝒙).S_{\bm{c}}(n)\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}).

Here 𝒄𝒙\colonequalsc1x1++cmxm\bm{c}\cdot\bm{x}\colonequals c_{1}x_{1}+\dots+c_{m}x_{m}, and in S𝒄(n)S_{\bm{c}}(n) the variable 𝒙\bm{x} runs over 1x1,,xmn1\leq x_{1},\dots,x_{m}\leq n. See Proposition 5.1 for the basic analytic properties of I𝒄(n)I_{\bm{c}}(n).

Given 𝒄m\bm{c}\in\mathbb{Z}^{m}, the sum S𝒄(n)S_{\bm{c}}(n) is multiplicative in nn, and roughly governed by the solutions to F(𝒙)=𝒄𝒙=0F(\bm{x})=\bm{c}\cdot\bm{x}=0 over /n\mathbb{Z}/n\mathbb{Z}. Let V𝒄V_{\bm{c}} be the intersection F(𝒙)=𝒄𝒙=0F(\bm{x})=\bm{c}\cdot\bm{x}=0 in m1\mathbb{P}^{m-1}_{\mathbb{Q}}. There is a classical discriminant polynomial F[𝒄]F^{\vee}\in\mathbb{Z}[\bm{c}] associated to the family 𝒄V𝒄\bm{c}\mapsto V_{\bm{c}}. The equation F(𝒄)=0F^{\vee}(\bm{c})=0 cuts out the projective dual variety VV^{\vee} of VV. See §2 for details on FF^{\vee}, VV^{\vee}.

It is natural to analyze the right-hand side of (1.3) separately over F(𝒄)0F^{\vee}(\bm{c})\neq 0 and F(𝒄)=0F^{\vee}(\bm{c})=0. In another paper ([wang2023_HLH_vs_RMT]), we conditionally address F(𝒄)0F^{\vee}(\bm{c})\neq 0 for certain FF, ww. In the present (self-contained) paper, we focus on F(𝒄)=0F^{\vee}(\bm{c})=0. From this locus in (1.3), our main theorem unconditionally isolates LΥL\bigcup_{L\in\Upsilon}L, for certain FF:

Theorem 1.1.

Suppose FF is diagonal and m{4,6}m\in\{4,6\}. For Y=X3/2Y=X^{3/2} as above, we have

(1.6) Y2𝒄m:F(𝒄)=0,𝒄𝟎n1nmS𝒄(n)I𝒄(n)=OF,w,ϵ(Xm/21/4+ϵ)+LΥ𝒙Lmw(𝒙/X).Y^{-2}\sum_{\bm{c}\in\mathbb{Z}^{m}:\,F^{\vee}(\bm{c})=0,\;\bm{c}\neq\bm{0}}\,\sum_{n\geq 1}n^{-m}S_{\bm{c}}(n)I_{\bm{c}}(n)=O_{F,w,\epsilon}(X^{m/2-1/4+\epsilon})+\sum_{L\in\Upsilon}\sum_{\bm{x}\in L\cap\mathbb{Z}^{m}}w(\bm{x}/X).

For m=4m=4, Theorem 1.1 follows from [heath1998circle]*Lemmas 7.2 and 8.1, with a better error of OF,w,ϵ(X3/2+ϵ)O_{F,w,\epsilon}(X^{3/2+\epsilon}). As we will explain in §8, the case m=6m=6 seems to present new difficulties, perhaps most easily resolved using our new, more geometric, strategy. But [heath1998circle] does bound the left-hand side of (1.6) by OF,w,ϵ(X3+ϵ)O_{F,w,\epsilon}(X^{3+\epsilon}) for m=6m=6, giving us a good foundation to build on. In view of [glas2022question], one might also hope to adapt our results to function fields.222Very recently, our work has largely been extended to function fields [BGW2024forthcoming].

Our main results assume FF is diagonal, but we do also discuss non-singular FF to a nontrivial extent. In particular, we will isolate explicit technical ingredients (listed in Remark 1.6) that—if true more generally—would allow one to generalize Theorem 1.1.

If 2m2\mid m (as in Theorem 1.1), the set Υ\Upsilon is known to be finite for general reasons (recalled in §3 below). Furthermore, we can relate (1.6) to (1.2) by inclusion-exclusion:

LΥ𝒙Lmw(𝒙/X)=OF,w(Xm/21)+𝒙m:𝒙LΥLw(𝒙/X).\sum_{L\in\Upsilon}\sum_{\bm{x}\in L\cap\mathbb{Z}^{m}}w(\bm{x}/X)=O_{F,w}(X^{m/2-1})+\sum_{\bm{x}\in\mathbb{Z}^{m}:\,\bm{x}\in\bigcup_{L\in\Upsilon}L}w(\bm{x}/X).

For each LΥL\in\Upsilon, if we define σ,L,w\sigma_{\infty,L^{\perp},w} as in (4.1), we also have (for all A>0A>0)

(1.7) 𝒙Lmw(𝒙/X)=σ,L,wXm/2+OL,w,A(XA).\sum_{\bm{x}\in L\cap\mathbb{Z}^{m}}w(\bm{x}/X)=\sigma_{\infty,L^{\perp},w}X^{m/2}+O_{L,w,A}(X^{-A}).

If FF is diagonal, Υ\Upsilon has an explicit classical description: see Proposition 3.6. In general, linear subvarieties of cubics are closely tied to the hh-invariant introduced by [davenport1962exponential, davenport1964non]. (See [dietmann2017h]*Lemma 1.1 for a more general relationship.) In fact, the present §7 relies on a convenient choice of “hh-decompositions of FF” corresponding to the elements of Υ\Upsilon.

We believe that with enough work, the power saving 1/41/4 in Theorem 1.1 could be improved. It would be very interesting to go past 1/21/2 for m=4m=4.

Theorem 1.1 has the following corollary, which we need for subsequent work ([wang2023_HLH_vs_RMT]).

Corollary 1.2.

Let m=6m=6. Define σ,F,w\sigma_{\infty,F,w}, 𝔖F\mathfrak{S}_{F} as in §6. Then in the setting of Theorem 1.1,

Y2𝒄m:F(𝒄)=0n1nmS𝒄(n)I𝒄(n)=OF,w,ϵ(X2.75+ϵ)+σ,F,w𝔖FX3+LΥσ,L,wX3.Y^{-2}\sum_{\bm{c}\in\mathbb{Z}^{m}:\,F^{\vee}(\bm{c})=0}\,\sum_{n\geq 1}n^{-m}S_{\bm{c}}(n)I_{\bm{c}}(n)=O_{F,w,\epsilon}(X^{2.75+\epsilon})+\sigma_{\infty,F,w}\mathfrak{S}_{F}X^{3}+\sum_{L\in\Upsilon}\sigma_{\infty,L^{\perp},w}X^{3}.
Proof.

Combine (1.6) (on 𝒄𝟎\bm{c}\neq\bm{0}) with (6.5) (on 𝒄=𝟎\bm{c}=\bm{0}). Note that m/2=m3=3m/2=m-3=3. ∎

Theorem 1.1 and Corollary 1.2 are completely unconditional. When m=6m=6 and FF is diagonal, these results let us reformulate the conjecture (1.2) as a statement purely about cancellation in the sum 𝒄m:F(𝒄)0n1nmS𝒄(n)I𝒄(n)\sum_{\bm{c}\in\mathbb{Z}^{m}:\,F^{\vee}(\bm{c})\neq 0}\sum_{n\geq 1}n^{-m}S_{\bm{c}}(n)I_{\bm{c}}(n). A similar reformulation might be possible much more generally. But at least when m=4m=4, subtleties in the constant cF,wc_{F,w} in (1.2) would likely demand a recipe beyond “restriction to F=0F^{\vee}=0” in (1.3).

Before proceeding, we make some convenient definitions. Let

(1.8) S𝒄(n)\colonequalsn(1+m)/2S𝒄(n);S^{\natural}_{\bm{c}}(n)\colonequals n^{-(1+m)/2}S_{\bm{c}}(n);

then in particular, nmS𝒄(n)I𝒄(n)=n(1m)/2S𝒄(n)I𝒄(n)n^{-m}S_{\bm{c}}(n)I_{\bm{c}}(n)=n^{(1-m)/2}S^{\natural}_{\bm{c}}(n)I_{\bm{c}}(n). The following is also convenient:

Definition 1.3.

Given a vector space LmL\subseteq\mathbb{Q}^{m} over \mathbb{Q}, let LL^{\perp} be the orthogonal complement of LL with respect to 𝒄𝒙\bm{c}\cdot\bm{x}. Then let Λ\colonequalsLm\Lambda\colonequals L\cap\mathbb{Z}^{m} and Λ\colonequalsLm\Lambda^{\perp}\colonequals L^{\perp}\cap\mathbb{Z}^{m}.

We now sketch the proof of Theorem 1.1. The proof starts generally, observing that F|L=0F^{\vee}|_{L^{\perp}}=0 for all LΥL\in\Upsilon (see Proposition 3.1). Conversely, at least for diagonal FF, most 𝒄\bm{c}’s on the left-hand side of (1.6) are in fact linear, in the sense of the following definition:

Definition 1.4.

Call a solution 𝒄m\bm{c}\in\mathbb{Z}^{m} to F(𝒄)=0F^{\vee}(\bm{c})=0 linear if 𝒄LΥΛ\bm{c}\in\bigcup_{L\in\Upsilon}\Lambda^{\perp}.

We expect “typical” linear 𝒄\bm{c}’s to be the simplest. Proposition 3.1(2), which establishes a vanishing baseline for the jets jFj^{\bullet}{F^{\vee}} over LΥL\bigcup_{L\in\Upsilon}L^{\perp}, thus inspires the following definition:

Definition 1.5.

Call FF^{\vee} unsurprising if uniformly over reals C1C\geq 1, the box [C,C]m[-C,C]^{m} contains at most Oϵ(Cm/21+ϵ)O_{\epsilon}(C^{m/2-1+\epsilon}) points in the union of the following two sets:

(1.9) 1\colonequals{𝒄m:F(𝒄)=0}LΥΛ,2\colonequals{𝒄LΥΛ:j2m/21F(𝒄)=𝟎}.{\textstyle\mathcal{E}_{1}\colonequals\{\bm{c}\in\mathbb{Z}^{m}:F^{\vee}(\bm{c})=0\}\setminus\bigcup_{L\in\Upsilon}\Lambda^{\perp},\quad\mathcal{E}_{2}\colonequals\{\bm{c}\in\bigcup_{L\in\Upsilon}\Lambda^{\perp}:j^{2^{m/2-1}}{F^{\vee}}(\bm{c})=\bm{0}\}.}

We prove (in §8) that when FF is diagonal, FF^{\vee} is unsurprising, so that if Υ\Upsilon\neq\emptyset, then almost all solutions to F=0F^{\vee}=0 are linear with nonzero 2m/212^{m/2-1}-jet.333An amusing corollary is that the Gauss map γ:VV\gamma\colon V\to V^{\vee} introduced in §2 can be far from surjective on \mathbb{Q}-points, since γ1([𝒄])V()\gamma^{-1}([\bm{c}])\cap V(\mathbb{Q}) may be empty for a typical nonzero 𝒄LΥL\bm{c}\in\bigcup_{L\in\Upsilon}L^{\perp}. A weaker version of Definition 1.5, with Cm/2δC^{m/2-\delta} in place of Cm/21+ϵC^{m/2-1+\epsilon}, would suffice for qualitative purposes.

For the “least degenerate” linear 𝒄\bm{c}’s, Lemma 7.11 isolates an explicit positive bias

(1.10) S𝒄(pl)=[Apl(𝒄)+O(pl/2)](1p1)pl/2S^{\natural}_{\bm{c}}(p^{l})=[A_{p^{l}}(\bm{c})+O(p^{-l/2})]\cdot(1-p^{-1})\cdot p^{l/2}

for most primes pp, with Ap(𝒄)=1A_{p}(\bm{c})=1 and Apl(𝒄)1A_{p^{l}}(\bm{c})\ll 1. The resulting reduction in arithmetic complexity of S𝒄(n)S_{\bm{c}}(n) lets us dramatically simplify 𝒄ΛS𝒄(n)I𝒄(n)\sum_{\bm{c}\in\Lambda^{\perp}}S_{\bm{c}}(n)I_{\bm{c}}(n) by Poisson summation over various individual residue classes 𝒄𝒃modn0Λ\bm{c}\equiv\bm{b}\bmod{n_{0}\Lambda^{\perp}} with n0n/Xn_{0}\ll n/X dividing nn. When m=6m=6, we must also carefully separate 𝒄=𝟎\bm{c}=\bm{0} from 𝒄𝟎\bm{c}\neq\bm{0}; we use Lemma 6.1 (decay of the singular series over large moduli). Eventually, σ,L,wXm/2\sigma_{\infty,L^{\perp},w}X^{m/2} appears, yielding (1.6).

As the positivity of S𝒄(n)S_{\bm{c}}(n) for linear 𝒄\bm{c}’s might suggest, we do not need cancellation over nn in (1.6). The deepest result we use on LL-functions (when m=6m=6) is the (purely local) Weil bound for hyperelliptic curves of genus 2\leq 2. However, it might be possible to improve the error term in Theorem 1.1 using deeper results on LL-functions.

The following remark proposes axioms that would let us go beyond diagonal FF.

Remark 1.6.

In general, (1.6) holds, provided m{4,6,8}m\in\{4,6,8\} and (1)–(4) hold:

  1. (1)

    FF^{\vee} is unsurprising (in the sense of Definition 1.5).

  2. (2)

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

  3. (3)

    There exists a homogeneous polynomial W[𝒄]W\in\mathbb{Z}[\bm{c}], satisfying

    (1.11) {𝒄LΥΛ:W(𝒄)=0}{𝒄m:j2m/21F(𝒄)=𝟎},{\textstyle\{\bm{c}\in\bigcup_{L\in\Upsilon}\Lambda^{\perp}:W(\bm{c})=0\}}\subseteq\{\bm{c}\in\mathbb{Z}^{m}:j^{2^{m/2-1}}{F^{\vee}}(\bm{c})=\bm{0}\},

    such that uniformly over integers C,n1C,n\geq 1, we have

    (1.12) 𝒄[C,C]m:F(𝒄)=0,W(𝒄)0n1|S𝒄(n)|2\displaystyle\sum_{\bm{c}\in[-C,C]^{m}:\,F^{\vee}(\bm{c})=0,\;W(\bm{c})\neq 0}n^{-1}\lvert S^{\natural}_{\bm{c}}(n)\rvert^{2} ϵ(Cn)ϵ(Cm/2+nm/6),\displaystyle\ll_{\epsilon}(Cn)^{\epsilon}(C^{m/2}+n^{m/6}),
    (1.13) 𝒄[C,C]m:F(𝒄)=0,W(𝒄)=0n1/2|S𝒄(n)|\displaystyle\sum_{\bm{c}\in[-C,C]^{m}:\,F^{\vee}(\bm{c})=0,\;W(\bm{c})=0}n^{-1/2}\lvert S^{\natural}_{\bm{c}}(n)\rvert ϵ(Cn)ϵ(C(m1)/2n1/6+nm/6).\displaystyle\ll_{\epsilon}(Cn)^{\epsilon}(C^{(m-1)/2}n^{1/6}+n^{m/6}).
  4. (4)

    In Lemma 7.11, the formula for S𝒄(p)S_{\bm{c}}(p) and upper bound for S𝒄(p2)S_{\bm{c}}(p^{\geq 2}) remain true, provided pp exceeds some constant depending on mm and FF.

(There could be alternatives to (3), but (3) would be convenient. See §5 and §8 for details.)

In Remark 1.6, we expect (4) to be the most tractable in general out of (1), (3), and (4). In fact, [wang2023dichotomous, BGW2024forthcoming] already make progress on (4). Also, [salberger2023counting] could be helpful for (1), at least if m=4m=4. Furthermore, the case where m=8m=8 and FF is diagonal might be fully accessible, and might provide insight on secondary terms in the circle method. This would be similar in spirit to [getz2018secondary], where secondary terms were obtained for quadrics in an even number of variables. The most mysterious axiom might be (3), but it does at least hold for diagonal FF, and could plausibly follow from some undiscovered geometric stratification.

Proposition 1.7.

Suppose FF is diagonal, and m4m\geq 4 is arbitrary. Then 1.6(3) holds.

Proof sketch.

Let W\colonequalsc1cmW\colonequals c_{1}\cdots c_{m}. One can prove (1.12) and (1.13) using (5.2). (See [wang2023_large_sieve_diagonal_cubic_forms]*§7; (1.12) holds with Cm/2C^{m/2} in place of Cm/2+nm/6C^{m/2}+n^{m/6}.) Also, Observation 3.10 implies (1.11). ∎

Remark 1.8.

It would be interesting to extend our analysis to m=5m=5, even just for diagonal FF. See [bombieri2009problems]*§3 for a discussion of the potentially infinite family of lines on V4V\subseteq\mathbb{P}^{4} if m=5m=5. Can one see these lines via VV^{\vee} (cf. Proposition 3.1)? It is conceivable that for m=5m=5, one might have to look past the locus F=0F^{\vee}=0 in (1.3), and perhaps include all 𝒄\bm{c} for which V𝒄V_{\bm{c}} has Picard rank 2\geq 2. The “bias” philosophy behind (1.10) may shed some light here.

We now outline the rest of the paper.

§2 collects some classical algebraic geometry used as a black box in parts of §3.

§3 suggests a geometric backbone in (1.3) for the eventual harmonic detection of Υ\Upsilon.

§4 proves some identities to be used in §8 to obtain the expected “main terms” in (1.6).

§5 provides general upper bounds to be used at several points in §8.

§6 analyzes the 𝒄=𝟎\bm{c}=\bm{0} contribution in (1.3), while also proving Lemma 6.1.

§7 proves some new asymptotic formulas for S𝒄(pl)S_{\bm{c}}(p^{l}) (see (1.10) and Lemma 7.11).

§8 uses Lemmas 6.1 and 7.11, plus results from §§35 and [heath1998circle], to prove Theorem 1.1.

1.1. Conventions

We let disc(F)\operatorname{disc}(F) denote the discriminant of FF, so that disc(F)0\operatorname{disc}(F)\neq 0 if and only if VV is smooth. We always assume VV is smooth, unless specified otherwise.

We define 𝟏E\bm{1}_{E} to be 11 if EE holds, and 0 if EE does not hold. We let e(t)\colonequalse2πite(t)\colonequals e^{2\pi it}, and er(t)\colonequalse(t/r)e_{r}(t)\colonequals e(t/r). Also, we use the following notation for integrals:

X𝑑xf(x)\colonequalsXf(x)𝑑x,X×Y𝑑x𝑑yf(x,y)\colonequalsX𝑑x(Y𝑑yf(x,y)).\int_{X}dx\,f(x)\colonequals\int_{X}f(x)\,dx,\quad\int_{X\times Y}dx\,dy\,f(x,y)\colonequals\int_{X}dx\left(\int_{Y}dy\,f(x,y)\right).

We let 𝔼bS[f(b)]\colonequals|S|1bSf(b)\mathbb{E}_{b\in S}[f(b)]\colonequals\lvert S\rvert^{-1}\sum_{b\in S}f(b) (if SS is finite and nonempty).

If F=F1x13++Fmxm3F=F_{1}x_{1}^{3}+\dots+F_{m}x_{m}^{3}, then we assume (after scaling FF^{\vee} if necessary) that

(1.14) F[𝒄]andgcd(coefficients of F)=(6m)!F1Fm.F^{\vee}\in\mathbb{Z}[\bm{c}]\qquad\textnormal{and}\qquad\gcd(\textnormal{coefficients of }F^{\vee})=(6^{m})!F_{1}\cdots F_{m}.

For a vector space UU, we let U\mathbb{P}{U} denote the projectivization of UU.

We let HessF=HessF(𝒙)\operatorname{Hess}{F}=\operatorname{Hess}{F}(\bm{x}) denote the usual m×mm\times m Hessian matrix of FF.

We let 0\colonequals{n:n0}\mathbb{Z}_{\geq 0}\colonequals\{n\in\mathbb{Z}:n\geq 0\}. We let x\colonequals/x\partial_{x}\colonequals\partial/\partial x; we let 𝒄𝒓\colonequalsc1r1cmrm\partial_{\bm{c}}^{\bm{r}}\colonequals\partial_{c_{1}}^{r_{1}}\cdots\partial_{c_{m}}^{r_{m}} (for 𝒓0m\bm{r}\in\mathbb{Z}_{\geq 0}^{m}) and |𝒃|\colonequalsiSbi\lvert\bm{b}\rvert\colonequals\sum_{i\in S}b_{i} (for 𝒃S\bm{b}\in\mathbb{Z}^{S}). We let

(1.15) jrF\colonequals(𝒄𝒓F)𝒓0m:|𝒓|rj^{r}{F^{\vee}}\colonequals(\partial_{\bm{c}}^{\bm{r}}{F^{\vee}})_{\bm{r}\in\mathbb{Z}_{\geq 0}^{m}:\,\lvert\bm{r}\rvert\leq r}

denote the rr-jet of FF^{\vee}, recording all partial derivatives of FF^{\vee} of order r\leq r.

We write fSgf\ll_{S}g, or gSfg\gg_{S}f, to mean “|f|Bg\lvert f\rvert\leq Bg for some real B=B(S)>0B=B(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.

When making estimates, we think of mm, FF, ww as fixed constants.

2. Algebraic geometry background

We do algebraic geometry in the language of schemes, with the symbols == and \subseteq interpreted scheme-theoretically unless specified otherwise. We call a reduced, locally closed subscheme of projective space over a field a variety. Ultimately, geometry lets us cut out subsets of m\mathbb{Z}^{m}, and count points over finite rings, in rigorous ways (see e.g. the key Lemma 7.11). Most of our work can be interpreted classically, in terms of points over algebraically closed fields.

The rest of this section reviews the geometry of the gradient map

F=(F/x1,,F/xm).\nabla{F}=(\partial F/\partial x_{1},\dots,\partial F/\partial x_{m}).

For diagonal FF, a more explicit analysis is possible (see §3.2).

Since VV is smooth (and 3F=𝒙F3F=\bm{x}\cdot\nabla{F}), the gradient F\nabla{F} defines a morphism [F]:m1(m1)[\nabla{F}]\colon\mathbb{P}^{m-1}\to(\mathbb{P}^{m-1})^{\vee}. (We will often identify the dual projective space (m1)(\mathbb{P}^{m-1})^{\vee} with m1\mathbb{P}^{m-1}, using c1,,cmc_{1},\dots,c_{m} as projective coordinates.) The map [F][\nabla{F}], known as the polar map of VV, is finite surjective of degree 2m12^{m-1}, by dimension and intersection theory; cf. [dolgachev2012classical]*p. 29. Since m1\mathbb{P}^{m-1} is smooth, [F][\nabla{F}] must then be flat (by “miracle flatness”).

Definition 2.1.

Let V(m1)V^{\vee}\subseteq(\mathbb{P}^{m-1}_{\mathbb{Q}})^{\vee} be the scheme-theoretic image of VV under [F][\nabla{F}].

Since VV is a smooth projective hypersurface, VV^{\vee} is the dual variety of VV. Upon restricting [F][\nabla{F}] to VV, we get the finite surjective Gauss map γ:VV\gamma\colon V\to V^{\vee}.

Here VV is geometrically irreducible, so V=imγV^{\vee}=\operatorname{im}\gamma must be too. Since γ\gamma is finite, VV^{\vee} must therefore be a geometrically integral hypersurface, i.e. the zero scheme of some absolutely irreducible form F[𝒄]F^{\vee}\in\mathbb{Q}[\bm{c}]. The definition of VV^{\vee} then implies that if 𝒄m{𝟎}\bm{c}\in\mathbb{C}^{m}\setminus\{\bm{0}\}, then F(𝒄)=0F^{\vee}(\bm{c})=0 if and only if the projective scheme F(𝒙)=𝒄𝒙=0F(\bm{x})=\bm{c}\cdot\bm{x}=0 over \mathbb{C} is singular.

(At least for diagonal FF, one can explicitly compute FF^{\vee}; see (3.3) in §3.2.)

Remark 2.2.

There are many equivalent ways to define VV^{\vee}. For example, VV^{\vee} is the locus of zeros 𝒄\bm{c} of a certain polynomial disc(F,𝒄)\operatorname{disc}(F,\bm{c}) in c1,,cmc_{1},\dots,c_{m} and the coefficients of FF; see e.g. [wang2023dichotomous]*Proposition 4.4. It should also be possible to interpret VV^{\vee} as a norm of VV under [F][\nabla{F}], in the sense of [stacks-project]*Tag 0BD2.

The polar map [F]:m1m1[\nabla{F^{\vee}}]\colon\mathbb{P}^{m-1}\dashrightarrow\mathbb{P}^{m-1} of VV^{\vee} is a rational map defined away from Sing(V)\operatorname{Sing}(V^{\vee}), the set of singular points of VV^{\vee}. (Here Sing(V)\operatorname{Sing}(V^{\vee}) is a proper closed subset of VV^{\vee}. It is known that VV^{\vee} is singular, since degF3\deg{F}\geq 3; see e.g. [wang2023dichotomous]*Proposition 4.4.)

The reflexivity theorem says that (V)=V(V^{\vee})^{\vee}=V. The biduality theorem says that if [𝒙]V[\bm{x}]\in V and [𝒄]V[\bm{c}]\in V^{\vee} are smooth points, then [F(𝒙)]=[𝒄][\nabla{F}(\bm{x})]=[\bm{c}] if and only if [F(𝒄)]=[𝒙][\nabla{F^{\vee}}(\bm{c})]=[\bm{x}]. (Both facts are on [dolgachev2012classical]*p. 30.) For us, VV is smooth, so biduality implies that the polar maps [F][\nabla{F}], [F][\nabla{F^{\vee}}] restrict to inverse morphisms between V[F]1(Sing(V))V\setminus[\nabla{F}]^{-1}(\operatorname{Sing}(V^{\vee})) and VSing(V)V^{\vee}\setminus\operatorname{Sing}(V^{\vee}).

It is known that degF=32m2\deg{F^{\vee}}=3\cdot 2^{m-2} [dolgachev2012classical]*p. 33, (1.47).

Since [F]:m1(m1)[\nabla{F}]\colon\mathbb{P}^{m-1}\to(\mathbb{P}^{m-1})^{\vee} is a finite surjective morphism of smooth varieties (and in particular, is generically étale), its ramification theory is well-behaved. Following [stacks-project]*Tag 0BWJ, let R[F]R_{[\nabla{F}]} be the closed subscheme of m1\mathbb{P}^{m-1} cut out by the different ideal 𝔇[F]𝒪m1\mathfrak{D}_{[\nabla{F}]}\subseteq\mathcal{O}_{\mathbb{P}^{m-1}} of [F][\nabla{F}]. Following [stacks-project]*Tag 0BW8 and Tag 0BWA, let B[F]B_{[\nabla{F}]} be the norm of R[F]R_{[\nabla{F}]} (or equivalently, the discriminant of [F][\nabla{F}]). In our setting, R[F]R_{[\nabla{F}]} is an effective Cartier divisor in m1\mathbb{P}^{m-1}, and B[F]B_{[\nabla{F}]} is thus an effective Cartier divisor in (m1)(\mathbb{P}^{m-1})^{\vee}.

The points of R[F]R_{[\nabla{F}]} are precisely those xm1x\in\mathbb{P}^{m-1} at which [F][\nabla{F}] is unramified. Furthermore, we have a set-theoretic equality B[F]=[F](R[F])B_{[\nabla{F}]}=[\nabla{F}](R_{[\nabla{F}]}).

Definition 2.3.

Call R[F]R_{[\nabla{F}]} and B[F]B_{[\nabla{F}]} the ramification divisor and branch divisor of [F][\nabla{F}], respectively. Let HR[𝒙]H_{R}\in\mathbb{Z}[\bm{x}] and HB[𝒄]H_{B}\in\mathbb{Z}[\bm{c}] be homogeneous polynomials defining R[F]R_{[\nabla{F}]} in m1\mathbb{P}^{m-1} and B[F]B_{[\nabla{F}]} in (m1)(\mathbb{P}^{m-1})^{\vee}, respectively.

Since degF3\deg{F}\geq 3, one can show that R[F]R_{[\nabla{F}]} and B[F]B_{[\nabla{F}]} are nonempty, and thus hypersurfaces. In fact, by [dolgachev2012classical]*p. 29, Proposition 1.2.1, R[F]=hess(V)R_{[\nabla{F}]}=\operatorname{hess}(V), where hess(V)\operatorname{hess}(V) denotes the subscheme det(HessF(𝒙))=0\det(\operatorname{Hess}{F}(\bm{x}))=0 of m1\mathbb{P}^{m-1}.

Proposition 2.4.

Say we let FF vary over the locus disc(F)0\operatorname{disc}(F)\neq 0. Then one can choose FF^{\vee} and HBH_{B} to be polynomials in c1,,cmc_{1},\dots,c_{m} and the coefficients of FF.

Proof.

This is possible by standard “universal” constructions compatible with our definitions of VV^{\vee} and B[F]B_{[\nabla{F}]}. For B[F]B_{[\nabla{F}]}, one can appeal to [stacks-project]*Tag 0BD2; norms respect base change (and thus vary nicely in families). For VV^{\vee}, see e.g. Remark 2.2. ∎

It is known that Vhess(V)V\not\subseteq\operatorname{hess}(V) [hooley1988nonary]*Lemma 1. How about after applying [F][\nabla{F}]?

Question 2.5.

Is it necessarily true that VB[F]V^{\vee}\not\subseteq B_{[\nabla{F}]}?

Question 2.5 comes up in Proposition 3.1, but we happen to be able to sidestep it there.

3. Maximal linear subvarieties under duality

Fix a smooth cubic V/V/\mathbb{Q} as in §1. The reader only interested in diagonal FF can skim forwards to §3.2, which explicitly analyzes Υ\Upsilon through the lens of FF^{\vee}.

3.1. A preliminary general analysis

If LΥL\in\Upsilon, then differentiating FF along LL implies LF(𝒙)L\perp\nabla{F}(\bm{x}) for all 𝒙L\bm{x}\in L. So the restriction γ|L=[F]|L\gamma|_{\mathbb{P}{L}}=[\nabla{F}]|_{\mathbb{P}{L}} maps into L\mathbb{P}{L^{\perp}}.

Since degF3\deg{F}\geq 3, it is also known by [debarre2003lines]*Lemma 3 (or Starr [starr2005fact_in_browning2006density]*Appendix) that if mm is even, then Υ\Upsilon is finite. We would like to understand Υ\Upsilon in terms of (1.3). Proposition 3.1 suggests one plausible starting route: duality (i.e. detecting LL^{\perp} through FF^{\vee}).

Proposition 3.1.

Suppose 2m42\mid m\geq 4, and fix LL in Υ\Upsilon.

  1. (1)

    γ|L\gamma|_{\mathbb{P}{L}} is a finite flat surjective morphism LL\mathbb{P}{L}\to\mathbb{P}{L^{\perp}} of degree 2m/2122^{m/2-1}\geq 2.

  2. (2)

    The jet j2m/211Fj^{2^{m/2-1}-1}{F^{\vee}}, defined as in (1.15), vanishes over L\mathbb{P}{L^{\perp}}.

For diagonal FF, we provide an explicit proof of Proposition 3.1 in §3.2. For more general FF, we instead rely on §2, plus some progress on Question 3.2. We would not be surprised if better technique allowed one to answer Question 3.2 in general.

Question 3.2.

Let y=[𝒄](m1)y=[\bm{c}]\in(\mathbb{P}^{m-1})^{\vee}. If the scheme-theoretic fiber m1×[F]y\mathbb{P}^{m-1}\times_{[\nabla{F}]}y of [F][\nabla{F}] over yy has degree dd, is it necessarily true that jd1F(𝒄)=𝟎j^{d-1}{F^{\vee}}(\bm{c})=\bm{0}?

We now begin the proof of Proposition 3.1. Since [F][\nabla{F}] is finite, the restriction γ|L:LL\gamma|_{\mathbb{P}{L}}\colon\mathbb{P}{L}\to\mathbb{P}{L^{\perp}} is itself finite, and thus surjective by dimension theory. So Limγ|LimγV\mathbb{P}{L^{\perp}}\subseteq\operatorname{im}\gamma|_{\mathbb{P}{L}}\subseteq\operatorname{im}\gamma\subseteq V^{\vee}.

Proof of (1).

Since γ|L\gamma|_{\mathbb{P}{L}} is finite surjective and L\mathbb{P}{L}, L\mathbb{P}{L^{\perp}} are smooth, “miracle flatness” implies flatness of γ|L\gamma|_{\mathbb{P}{L}}. Also, γ|L\gamma|_{\mathbb{P}{L}} has degree 2m/212^{m/2-1} (cf. [dolgachev2012classical]*top of p. 29), since it is a morphism given by quadratic polynomials, between projective spaces of dimension m/21m/2-1. ∎

Proposition 3.1(2) is inspired by the factorization of FF^{\vee} over ¯[𝒄1/2]\overline{\mathbb{Q}}[\bm{c}^{1/2}] when FF is diagonal (see (3.3) in §3.2 below). However, giving a rigorous “factorization” of FF^{\vee} seems to require a bit of setup, since the map [F][\nabla{F}] presumably need not be Galois in general. Furthermore, our factorization will only be useful for (2) if

(3.1) LB[F](or equivalently, HB|L0).\mathbb{P}{L^{\perp}}\not\subseteq B_{[\nabla{F}]}\quad(\textnormal{or equivalently, $H_{B}|_{L^{\perp}}\neq 0$}).
Question 3.3.

Is LB[F]\mathbb{P}{L^{\perp}}\subseteq B_{[\nabla{F}]} possible (if VV is smooth and LΥL\in\Upsilon)?

We do not know the answer to Question 3.3. Fortunately, if we fix mm and LL, then the set 𝒰\mathscr{U} of all mm-variable cubic forms P/P/\mathbb{Q} with disc(P)0\operatorname{disc}(P)\neq 0 and P|L=0P|_{L}=0 is a dense open set in a copy of 𝔸N(m)\mathbb{A}^{N(m)}_{\mathbb{Q}}, where N(m)=(m+23)(m2+23)N(m)=\binom{m+2}{3}-\binom{\frac{m}{2}+2}{3}. Furthermore, Proposition 2.4 implies that 3.1(2) is a closed condition on F𝒰F\in\mathscr{U}, and that (3.1) is an open condition on F𝒰F\in\mathscr{U}. Since (3.1) holds when LL is x1+x2=x3+x4=x5+x6=0x_{1}+x_{2}=x_{3}+x_{4}=x_{5}+x_{6}=0 (as we may assume, by a linear change of variables) and FF is x13++x63x_{1}^{3}+\dots+x_{6}^{3}, it thus suffices to prove 3.1(2) assuming (3.1).

Hence, for the rest of §3.1, we assume (3.1), though (3.1) will only come into play after some initial work. Consider the hypersurface complements S\colonequalsm1B[F]S\colonequals\mathbb{P}^{m-1}\setminus B_{[\nabla{F}]} and X\colonequals[F]1Sm1X\colonequals[\nabla{F}]^{-1}S\subseteq\mathbb{P}^{m-1}. Then [F]|X:XS[\nabla{F}]|_{X}\colon X\to S is finite étale of degree 2m12^{m-1}. Write ϕ\colonequals[F]|X\phi\colonequals[\nabla{F}]|_{X}. By Grothendieck’s Galois theory, there exists a finite étale Galois cover π:XX\pi\colon X^{\prime}\to X with XX^{\prime} connected and ϕπ:XS\phi\circ\pi\colon X^{\prime}\to S (finite étale) Galois. Let G\colonequalsAutS(X)G\colonequals\operatorname{Aut}_{S}(X^{\prime}) and H\colonequalsAutX(X)H\colonequals\operatorname{Aut}_{X}(X^{\prime}).

The group GG acts transitively on XX^{\prime}. So for any geometric points [𝒄]S(¯)[\bm{c}]\in S(\overline{\mathbb{Q}}) and pX[𝒄]p\in X^{\prime}_{[\bm{c}]}, we can characterize the fiber X[𝒄]X_{[\bm{c}]} as the set {π(gp):gH\G}X(¯)\{\pi(gp):g\in H\backslash G\}\subseteq X(\overline{\mathbb{Q}}).

3.1.1. Constructing a product “divisible” by FF^{\vee}

View F=F(𝒙)F=F(\bm{x}) as a section of 𝒪X(3)\mathcal{O}_{X}(3). Consider the GG-equivariant line bundle \colonequalsgH\Gg(π𝒪X(3))\mathcal{L}\colonequals\bigotimes_{g\in H\backslash G}g^{\ast}(\pi^{\ast}\mathcal{O}_{X}(3)) on XX^{\prime}. The product

(3.2) α\colonequalsgH\G(gπF)\alpha\colonequals\prod_{g\in H\backslash G}(g^{\ast}\pi^{\ast}{F})

defines a GG-invariant section of \mathcal{L} on XX^{\prime}. (A GG-invariant section αΓ(X,)\alpha\in\Gamma(X^{\prime},\mathcal{L}) is equivalent in data to a GG-equivariant morphism α:𝒪X\alpha\colon\mathcal{O}_{X^{\prime}}\to\mathcal{L}.)

For every geometric point pX(¯)p\in X^{\prime}(\overline{\mathbb{Q}}) with ϕπ(p)SV\phi\pi(p)\in S\cap V^{\vee}, there exists gH\Gg\in H\backslash G with π(gp)XV\pi(gp)\in X\cap V, so that (gπF)(p)=F(πgp)=0(g^{\ast}\pi^{\ast}{F})(p)=F(\pi gp)=0. So

α|(ϕπ)1(SV)=0.\alpha|_{(\phi\pi)^{-1}(S\cap V^{\vee})}=0.

Therefore, by Galois descent (in the form of an equivalence of categories), there exist a line bundle 𝒟\mathcal{D} on SS with (ϕπ)𝒟\mathcal{L}\cong(\phi\pi)^{\ast}\mathcal{D}, and a section δΓ(S,𝒟)\delta\in\Gamma(S,\mathcal{D}) vanishing along SVS\cap V^{\vee}, with α=(ϕπ)δ\alpha=(\phi\pi)^{\ast}\delta. Let \colonequalsϕ𝒟\mathcal{F}\colonequals\phi^{\ast}\mathcal{D} and β\colonequalsϕδΓ(X,)\beta\colonequals\phi^{\ast}\delta\in\Gamma(X,\mathcal{F}); then π\mathcal{L}\cong\pi^{\ast}\mathcal{F} and α=πβ\alpha=\pi^{\ast}\beta.

But SS, XX are hypersurface complements in m1\mathbb{P}^{m-1}, so Pic(m1)Pic(S)\operatorname{Pic}(\mathbb{P}^{m-1})\to\operatorname{Pic}(S) and Pic(m1)Pic(X)\operatorname{Pic}(\mathbb{P}^{m-1})\to\operatorname{Pic}(X) are surjective and we may identify \mathcal{F}, 𝒟\mathcal{D} with suitable powers of 𝒪X(1)\mathcal{O}_{X}(1), 𝒪S(1)\mathcal{O}_{S}(1), respectively. Then up to a choice of nonzero constant factors, we may view β\beta, δ\delta as homogeneous rational functions (i.e. ratios of homogeneous mm-variable polynomials) satisfying F(𝒄)δF^{\vee}(\bm{c})\mid\delta on SS and F(F(𝒙))=ϕFβF^{\vee}(\nabla{F}(\bm{x}))=\phi^{\ast}{F^{\vee}}\mid\beta on XX. Here we interpret divisibility of two sections on a scheme to mean their ratio is a global section of the obvious “tensor-quotient” line bundle.

3.1.2. “Factoring” FF^{\vee}

By the definition of VV^{\vee} and FF^{\vee} in §2, we have F(𝒙)F(F(𝒙))F(\bm{x})\mid F^{\vee}(\nabla{F}(\bm{x})) in [𝒙]\mathbb{Q}[\bm{x}]. So FϕFF\mid\phi^{\ast}{F^{\vee}} on XX. By (3.2), it follows that α(πϕF)|H\G|\alpha\mid(\pi^{\ast}\phi^{\ast}{F^{\vee}})^{\lvert H\backslash G\rvert} on XX^{\prime}, since gπϕF=πϕFg^{\ast}\pi^{\ast}\phi^{\ast}{F^{\vee}}=\pi^{\ast}\phi^{\ast}{F^{\vee}} for all gGg\in G. Since α=πϕδ\alpha=\pi^{\ast}\phi^{\ast}\delta, it follows (by Galois descent) that δ(F)|H\G|\delta\mid(F^{\vee})^{\lvert H\backslash G\rvert} on SS. Since FδF^{\vee}\mid\delta on SS, and FF^{\vee} is prime in [𝒄]\mathbb{Q}[\bm{c}], we conclude that there exists an integer e1e\geq 1 satisfying (F)eδ(F^{\vee})^{e}\mid\delta and δ(F)e\delta\mid(F^{\vee})^{e} on SS.

We need to restrict to VV^{\vee}. Luckily, (3.1) and LV\mathbb{P}{L^{\perp}}\subseteq V^{\vee} imply that SVS\cap V^{\vee}\neq\emptyset (but see Question 2.5). Furthermore, VV^{\vee} is geometrically reduced, so VSing(V)V^{\vee}\setminus\operatorname{Sing}(V^{\vee}) is a dense open subvariety of VV^{\vee} (by “generic smoothness”). Thus (SV)Sing(V)(S\cap V^{\vee})\setminus\operatorname{Sing}(V^{\vee})\neq\emptyset.

Choose a geometric point [𝒄][\bm{c}] of (SV)Sing(V)(S\cap V^{\vee})\setminus\operatorname{Sing}(V^{\vee}). Biduality furnishes a unique point [𝒙]X[𝒄][\bm{x}]\in X_{[\bm{c}]} with F(𝒙)=0F(\bm{x})=0. So if pX[𝒙]p\in X^{\prime}_{[\bm{x}]} and gGg\in G, then the section gπFg^{\ast}\pi^{\ast}{F} on XX^{\prime} evaluates to 0 at pp if and only if gHg\in H. Thus πF[1]gH\G(gπF)\pi^{\ast}{F}\nmid\prod_{[1]\neq g\in H\backslash G}(g^{\ast}\pi^{\ast}{F}) on XX^{\prime}. It follows that (πF)2α(\pi^{\ast}{F})^{2}\nmid\alpha, whence F2βF^{2}\nmid\beta; whence (F)2δ(F^{\vee})^{2}\nmid\delta.

Thus e=1e=1. In particular, δF\delta\mid F^{\vee} on SS, so α(ϕπ)F\alpha\mid(\phi\pi)^{\ast}{F^{\vee}} on XX^{\prime}.

3.1.3. Differentiating the product

Using (3.1) one last time (more seriously than before), we will now complete the proof of the second part of Proposition 3.1.

Proof of (2).

By (3.1), SLS\cap\mathbb{P}{L^{\perp}}\neq\emptyset. Yet ϕ|XL=γ|XL:XLSL\phi|_{X\cap\mathbb{P}{L}}=\gamma|_{X\cap\mathbb{P}{L}}\colon X\cap\mathbb{P}{L}\to S\cap\mathbb{P}{L^{\perp}} is finite étale of degree 2m/212^{m/2-1}, by part (1) and the definition of SS. Let [𝒄](SL)(¯)[\bm{c}]\in(S\cap\mathbb{P}{L^{\perp}})(\overline{\mathbb{Q}}), and fix pX[𝒄]p\in X^{\prime}_{[\bm{c}]}. Then there exist at least 2m/212^{m/2-1} cosets gH\Gg\in H\backslash G with πgp(XL)(¯)V(¯)\pi gp\in(X\cap\mathbb{P}{L})(\overline{\mathbb{Q}})\subseteq V(\overline{\mathbb{Q}}). Applying the Leibniz rule to (3.2), after restricting to a small affine neighborhood of pp, we thus get jprα(p)=𝟎j_{p}^{r}{\alpha}(p)=\bm{0} for r\colonequals2m/211r\colonequals 2^{m/2-1}-1, where jr:Jrj^{r}\colon\mathcal{L}\to J^{r}\mathcal{L} denotes the rrth-order jet map “along” \mathcal{L} (from \mathcal{L} to its rrth jet bundle JrJ^{r}\mathcal{L}).

Since α(ϕπ)F\alpha\mid(\phi\pi)^{\ast}{F^{\vee}}, Leibniz then implies jpr(ϕπ)F(p)=𝟎j_{p}^{r}{(\phi\pi)^{\ast}{F^{\vee}}}(p)=\bm{0} “along” the pullback line bundle (ϕπ)𝒪S(degF)(\phi\pi)^{\ast}\mathcal{O}_{S}(\deg{F^{\vee}}). But ϕπ:XS\phi\pi\colon X^{\prime}\to S is étale at pXp\in X^{\prime}, so j[𝒄]rF([𝒄])=𝟎j_{[\bm{c}]}^{r}{F^{\vee}}([\bm{c}])=\bm{0} “along” 𝒪S(degF)\mathcal{O}_{S}(\deg{F^{\vee}}) itself, over all points [𝒄](SL)(¯)[\bm{c}]\in(S\cap\mathbb{P}{L^{\perp}})(\overline{\mathbb{Q}}). Finally, SLS\cap\mathbb{P}{L^{\perp}} is dense in L\mathbb{P}{L^{\perp}}, so the vanishing of the rrth-order jet section jrFj^{r}F^{\vee} extends to all points [𝒄]L[\bm{c}]\in\mathbb{P}{L^{\perp}}, as desired. ∎

Remark 3.4.

In the friendly setting above, our étale morphisms (such as ϕπ:XS\phi\pi\colon X^{\prime}\to S), after base change to an algebraically closed field, always induce isomorphisms on completed local rings. So we could do calculus purely in terms of formal power series.

3.2. The diagonal case

Say mm is even and FF is diagonal, and write F=F1x13++Fmxm3F=F_{1}x_{1}^{3}+\dots+F_{m}x_{m}^{3}. Then we can explicitly verify all the theory above. Here [F]:[𝒙][3F1x12,,3Fmxm2][\nabla{F}]\colon[\bm{x}]\mapsto[3F_{1}x_{1}^{2},\dots,3F_{m}x_{m}^{2}] is multi-quadratic Galois of degree 2m12^{m-1}. We proceed by studying Υ\Upsilon and FF^{\vee} combinatorially.

For combinatorial purposes, let [n]\colonequals{1,2,,n}[n]\colonequals\{1,2,\dots,n\} for each integer n1n\geq 1.

Definition 3.5.

Let 𝒥=(𝒥(k))k𝒦\mathcal{J}=(\mathcal{J}(k))_{k\in\mathcal{K}} denote an ordered set partition of [m][m]: a list of pairwise disjoint nonempty sets 𝒥(k)[m]\mathcal{J}(k)\subseteq[m] covering [m][m], indexed by a set 𝒦{[1],[2],[3],}\mathcal{K}\in\{[1],[2],[3],\ldots\}.

  1. (1)

    Call 𝒥\mathcal{J}, 𝒥\mathcal{J}^{\prime} equivalent if they define the same unordered partition of [m][m] (i.e. if 𝒦=𝒦\mathcal{K}=\mathcal{K}^{\prime} and {𝒥(k):k𝒦}={𝒥(k):k𝒦}\{\mathcal{J}(k):k\in\mathcal{K}\}=\{\mathcal{J}^{\prime}(k):k\in\mathcal{K}^{\prime}\}).

  2. (2)

    Call 𝒥\mathcal{J} a pairing if |𝒥(k)|=2\lvert\mathcal{J}(k)\rvert=2 for all k𝒦k\in\mathcal{K}.

  3. (3)

    Call 𝒥\mathcal{J} permissible if for all k𝒦k\in\mathcal{K} and i,j𝒥(k)i,j\in\mathcal{J}(k), we have Fj/Fi(×)3F_{j}/F_{i}\in(\mathbb{Q}^{\times})^{3}. For a permissible 𝒥\mathcal{J}, let 𝒥\colonequals{𝒄m:if k𝒦 and i,j𝒥(k), then ci/Fi1/3=cj/Fj1/3}\mathcal{R}_{\mathcal{J}}\colonequals\{\bm{c}\in\mathbb{Z}^{m}:\textnormal{if $k\in\mathcal{K}$ and $i,j\in\mathcal{J}(k)$, then $c_{i}/F_{i}^{1/3}=c_{j}/F_{j}^{1/3}$}\}; and given 𝒄𝒥\bm{c}\in\mathcal{R}_{\mathcal{J}}, define c:𝒦c\colon\mathcal{K}\to\mathbb{R} so that if k𝒦k\in\mathcal{K} and i𝒥(k)i\in\mathcal{J}(k), then ci/Fi1/3=c(k)c_{i}/F_{i}^{1/3}=c(k).

We now recall the well-known construction of the m2\frac{m}{2}-dimensional vector spaces L/L/\mathbb{C} with F|L=0F|_{L}=0. Each equivalence class of pairings 𝒥\mathcal{J} yields 3m/23^{m/2} distinct L/L/\mathbb{C}, obtained by setting Fixi3+Fjxj3=0F_{i}x_{i}^{3}+F_{j}x_{j}^{3}=0 for each part 𝒥(k)={i,j}\mathcal{J}(k)=\{i,j\}. Over \mathbb{Q}, we must set xi+(Fj/Fi)1/3xj=0x_{i}+(F_{j}/F_{i})^{1/3}x_{j}=0—which is valid when FiFjmod(×)3F_{i}\equiv F_{j}\bmod{(\mathbb{Q}^{\times})^{3}}. The vectors spaces L/L/\mathbb{C} and L/L/\mathbb{Q} thus constructed are known to be the only possibilities [wang2022thesis]*Remark 6.3.8. Hence the following holds:

Proposition 3.6.

There is a canonical bijection, between Υ\Upsilon and the set of equivalence classes of permissible pairings 𝒥\mathcal{J}, characterized by Lm=𝒥L\cap\mathbb{Z}^{m}=\mathcal{R}_{\mathcal{J}}^{\perp} (an equality of sublattices of m\mathbb{Z}^{m}).

Next, we turn to FF^{\vee}. For convenience, fix square roots Fi1/2¯×F_{i}^{1/2}\in\overline{\mathbb{Q}}^{\times}. For some constant βF1,,Fm×\beta_{F_{1},\dots,F_{m}}\in\mathbb{Q}^{\times}, the polynomial F(𝒄)F^{\vee}(\bm{c}) factors in ¯[𝒄1/2]\overline{\mathbb{Q}}[\bm{c}^{1/2}] as follows:

(3.3) F(𝒄)=βF1,,Fmϵ(ϵ1F11/2c13/2+ϵ2F21/2c23/2++ϵmFm1/2cm3/2)[𝒄],F^{\vee}(\bm{c})=\beta_{F_{1},\dots,F_{m}}\cdot\prod_{\bm{\epsilon}}(\epsilon_{1}F_{1}^{-1/2}c_{1}^{3/2}+\epsilon_{2}F_{2}^{-1/2}c_{2}^{3/2}+\dots+\epsilon_{m}F_{m}^{-1/2}c_{m}^{3/2})\in\mathbb{Q}[\bm{c}],

with the product taken over ϵ=(ϵ1,,ϵm)\bm{\epsilon}=(\epsilon_{1},\dots,\epsilon_{m}) with ϵ1=1\epsilon_{1}=1 and ϵ2,,ϵm=±1\epsilon_{2},\dots,\epsilon_{m}=\pm 1. (This classical formula is a simple consequence of Definition 2.1 and the Jacobian criterion.)

Let ϵ\colonequalsϵ1F11/2c13/2++ϵmFm1/2cm3/2¯[𝒄1/2]\mathscr{L}_{\bm{\epsilon}}\colonequals\epsilon_{1}F_{1}^{-1/2}c_{1}^{3/2}+\dots+\epsilon_{m}F_{m}^{-1/2}c_{m}^{3/2}\in\overline{\mathbb{Q}}[\bm{c}^{1/2}]. For each 𝒄m{𝟎}\bm{c}\in\mathbb{Q}^{m}\setminus\{\bm{0}\}, fix square roots ci1/2¯c_{i}^{1/2}\in\overline{\mathbb{Q}}. Using formal power series calculus over variables ci0c_{i}\neq 0 (by Remark 3.4, adapted to 𝔸¯1𝔸¯1,tt2\mathbb{A}^{1}_{\overline{\mathbb{Q}}}\to\mathbb{A}^{1}_{\overline{\mathbb{Q}}},\;t\mapsto t^{2} away from the origin), we will prove the following result, which precisely characterizes the order of vanishing of FF^{\vee} at 𝒄\bm{c}:

Proposition 3.7.

Fix r0r\geq 0 and 𝐜m{𝟎}\bm{c}\in\mathbb{Q}^{m}\setminus\{\bm{0}\}. Then the following are equivalent: (1) jrFj^{r}{F^{\vee}} vanishes at 𝐜\bm{c}, and (2) there exist at least r+1r+1 distinct ϵ\bm{\epsilon} with ϵ=0\mathscr{L}_{\bm{\epsilon}}=0.

Proof.

We first prove (1)\Rightarrow(2), by induction on r0r\geq 0. The base case r=0r=0 follows directly from (3.3). Now fix r1r\geq 1, and assume the implication (1)\Rightarrow(2) holds for r1r-1. Suppose (1) holds; we wish to prove that (2) holds.

By the inductive hypothesis, #{ϵ:ϵ=0}r\#\{\bm{\epsilon}:\mathscr{L}_{\bm{\epsilon}}=0\}\geq r. To rule out the possibility that #{ϵ:ϵ=0}=r\#\{\bm{\epsilon}:\mathscr{L}_{\bm{\epsilon}}=0\}=r, we work with “pure” derivatives cir\partial_{c_{i}}^{\leq r}, for just a single index ii with ci0c_{i}\neq 0. For example, if c10c_{1}\neq 0, and #{ϵ:ϵ=0}=r\#\{\bm{\epsilon}:\mathscr{L}_{\bm{\epsilon}}=0\}=r, then the product rule and (3.3) imply

c1rF(𝒄)=βF1,,Fmr!(32F11/2c11/2)rϵ:ϵ0ϵ0,\partial_{c_{1}}^{r}{F^{\vee}}(\bm{c})=\beta_{F_{1},\dots,F_{m}}\cdot r!\cdot(\tfrac{3}{2}F_{1}^{-1/2}c_{1}^{1/2})^{r}\prod_{\bm{\epsilon}:\,\mathscr{L}_{\bm{\epsilon}}\neq 0}\mathscr{L}_{\bm{\epsilon}}\neq 0,

contradicting (1). Thus #{ϵ:ϵ=0}r+1\#\{\bm{\epsilon}:\mathscr{L}_{\bm{\epsilon}}=0\}\geq r+1. This completes the induction.

It remains to prove (2)\Rightarrow(1). To avoid confusion, rename our given 𝒄\bm{c} to 𝝃\bm{\xi}. Say ξi=0\xi_{i}=0 for iIi\in I, and ξi0\xi_{i}\neq 0 for i[m]Ii\in[m]\setminus I. We must take extra care over iIi\in I.

Let 𝒯I\mathscr{T}_{I} be the set of triples (𝒂,𝒃,E)(\bm{a},\bm{b},E), with (𝒂,𝒃)0I×[m]I(\bm{a},\bm{b})\in\mathbb{Z}_{\geq 0}^{I}\times\mathbb{Z}^{[m]\setminus I} and E{ϵ{±1}m:ϵ1=1}E\subseteq\{\bm{\epsilon}\in\{\pm 1\}^{m}:\epsilon_{1}=1\}, such that for each iIi\in I, the set E(E)E\cup(-E) is invariant under the flip ϵiϵi\epsilon_{i}\mapsto-\epsilon_{i}. Consider the following “formal analytic functions” (inspired by (3.3)), indexed by (𝒂,𝒃,E)𝒯I(\bm{a},\bm{b},E)\in\mathscr{T}_{I}:

(3.4) (iIciai)(iIcibi/2)(ϵEϵ)¯[ci]iI[ci1/2,ci1/2]iI.\biggl{(}\,\prod_{i\in I}c_{i}^{a_{i}}\biggr{)}\biggl{(}\,\prod_{i\notin I}c_{i}^{b_{i}/2}\biggr{)}\biggl{(}\,\prod_{\bm{\epsilon}\in E}\mathscr{L}_{\bm{\epsilon}}\biggr{)}\in\overline{\mathbb{Q}}[c_{i}]_{i\in I}[c_{i}^{1/2},c_{i}^{-1/2}]_{i\notin I}.

The functions (3.4) span a vector space over ¯\overline{\mathbb{Q}} that contains FF^{\vee} and is closed under differentiation in 𝒄\bm{c}. In fact, differentiating (3.4) in cic_{i} leads to terms with aiai1a_{i}\mapsto a_{i}-1 or (ai,|E|)(ai+2,|E|2)(a_{i},\lvert E\rvert)\mapsto(a_{i}+2,\lvert E\rvert-2) if iIi\in I, and to terms with bibi2b_{i}\mapsto b_{i}-2 or (bi,|E|)(bi+1,|E|1)(b_{i},\lvert E\rvert)\mapsto(b_{i}+1,\lvert E\rvert-1) if iIi\notin I. In each case, applying ci\partial_{c_{i}} decreases min𝒂,𝒃,E(|𝒂|+|E|)\min_{\bm{a},\bm{b},E}(\lvert\bm{a}\rvert+\lvert E\rvert) by at most 11.

Now suppose (2) holds, and fix 𝒓0m\bm{r}\in\mathbb{Z}_{\geq 0}^{m} with |𝒓|r\lvert\bm{r}\rvert\leq r. Then 𝒄𝒓F\partial_{\bm{c}}^{\bm{r}}{F^{\vee}} is a ¯\overline{\mathbb{Q}}-linear combination of functions (3.4) indexed by triples (𝒂,𝒃,E)𝒯I(\bm{a},\bm{b},E)\in\mathscr{T}_{I} with |𝒂|+|E|2m1r\lvert\bm{a}\rvert+\lvert E\rvert\geq 2^{m-1}-r (and thus |E|2m1r\lvert E\rvert\geq 2^{m-1}-r or |𝒂|1\lvert\bm{a}\rvert\geq 1). By (2), each such function must vanish at our original given point 𝒄=𝝃\bm{c}=\bm{\xi}. Thus 𝒄𝒓F(𝝃)=0\partial_{\bm{c}}^{\bm{r}}{F^{\vee}}(\bm{\xi})=0. So (1) holds. ∎

Remark 3.8.

A short computation yields the equality

(3.5) #{ϵ:ϵ=0}=[𝒙]γ(¯)1([𝒄])2#{i[m]:xi=0},\#\{\bm{\epsilon}:\mathscr{L}_{\bm{\epsilon}}=0\}=\sum_{[\bm{x}]\in\gamma(\overline{\mathbb{Q}})^{-1}([\bm{c}])}2^{\#\{i\in[m]:\,x_{i}=0\}},

where γ(¯)1([𝒄])\colonequals{[𝒙]V(¯):[F(𝒙)]=[𝒄]}={singular ¯-points of V𝒄}\gamma(\overline{\mathbb{Q}})^{-1}([\bm{c}])\colonequals\{[\bm{x}]\in V(\overline{\mathbb{Q}}):[\nabla{F}(\bm{x})]=[\bm{c}]\}=\{\textnormal{singular $\overline{\mathbb{Q}}$-points of $V_{\bm{c}}$}\}. (Here xix_{i} corresponds to ϵiFi1/2ci1/2\epsilon_{i}F_{i}^{-1/2}c_{i}^{1/2}, with some ambiguity or “multiplicity” in ϵi\epsilon_{i} when xi=0x_{i}=0.) Using (3.5), one can formulate Proposition 3.7 more geometrically, without reference to ϵ\bm{\epsilon}’s. Does this geometric formulation extend somehow to more general FF?

Finally, we analyze the interaction between Υ\Upsilon, FF^{\vee}. Fix LΥL\in\Upsilon. By Proposition 3.6, LL corresponds to some permissible pairing 𝒥\mathcal{J}. Proposition 3.7 has the following corollary:

Corollary 3.9.

For LL as above, we have (j2m/211F)|L=𝟎(j^{2^{m/2-1}-1}{F^{\vee}})|_{L^{\perp}}=\bm{0}.

Proof.

For each part 𝒥(k)={i,j}\mathcal{J}(k)=\{i,j\}, there are exactly two choices of signs (ϵi,ϵj){±1}2(\epsilon_{i},\epsilon_{j})\in\{\pm 1\}^{2}—or only one choice if 1𝒥(k)1\in\mathcal{J}(k)—such that ϵiFi1/2ci3/2+ϵjFj1/2cj3/2\epsilon_{i}F_{i}^{-1/2}c_{i}^{3/2}+\epsilon_{j}F_{j}^{-1/2}c_{j}^{3/2} vanishes over all 𝒄Lm=𝒥\bm{c}\in L^{\perp}\cap\mathbb{Z}^{m}=\mathcal{R}_{\mathcal{J}} lying in a given orthant of m\mathbb{R}^{m}. So given 𝒄L{𝟎}\bm{c}\in L^{\perp}\setminus\{\bm{0}\}, we can apply Proposition 3.7(2)\Rightarrow(1) with r\colonequals2m/211r\colonequals 2^{m/2-1}-1. Since L{𝟎}L^{\perp}\setminus\{\bm{0}\} is dense in LL^{\perp}, the vanishing of jrFj^{r}{F^{\vee}} then extends to all of LL^{\perp}. ∎

Thus we have explicitly verified the conclusion of Proposition 3.1. The next result shows that in fact, FF^{\vee} generally does not vanish to higher order along LL^{\perp} (and furthermore, 3.10(1) gives us a simple description of when higher vanishing occurs). Let s:,qq2s\colon\mathbb{Q}\to\mathbb{Q},\;q\mapsto q^{2}.

Observation 3.10.

Given LL, 𝒥\mathcal{J} as above, fix 𝒄Lm=𝒥\bm{c}\in L^{\perp}\cap\mathbb{Z}^{m}=\mathcal{R}_{\mathcal{J}}. Define c(k)c(k) as in Definition 3.5. For each k𝒦k\in\mathcal{K}, fix a square root c(k)3/2¯c(k)^{3/2}\in\overline{\mathbb{Q}} of c(k)3c(k)^{3}. Then the following hold:

  1. (1)

    j2m/21F(𝒄)=𝟎j^{2^{m/2-1}}{F^{\vee}}(\bm{c})=\bm{0} if and only if there exist l1l\geq 1 distinct indices k1,,kl𝒦k_{1},\dots,k_{l}\in\mathcal{K} such that c(k1)3/2±±c(kl)3/2=0c(k_{1})^{3/2}\pm\cdots\pm c(k_{l})^{3/2}=0 holds for some choice of signs.

  2. (2)

    If j2m/21F(𝒄)=𝟎j^{2^{m/2-1}}{F^{\vee}}(\bm{c})=\bm{0}, then c(k1)3c(k2)3s()c(k_{1})^{3}c(k_{2})^{3}\in s(\mathbb{Q}) for some distinct k1,k2𝒦k_{1},k_{2}\in\mathcal{K}.

Proof.

(1): If 𝒄=𝟎\bm{c}=\bm{0}, then j2m/21F(𝒄)=𝟎j^{2^{m/2-1}}{F^{\vee}}(\bm{c})=\bm{0} (since degF=32m21+2m/21\deg{F^{\vee}}=3\cdot 2^{m-2}\geq 1+2^{m/2-1}), and c(k)3=0c(k)^{3}=0 for all k𝒦k\in\mathcal{K}. If 𝒄𝟎\bm{c}\neq\bm{0}, apply Proposition 3.7(1)\Leftrightarrow(2) with r\colonequals2m/21r\colonequals 2^{m/2-1} (and then “simplify” the condition 3.7(2) using the fact that 𝒥\mathcal{J} is a pairing).

(2): Suppose j2m/21F(𝒄)=𝟎j^{2^{m/2-1}}{F^{\vee}}(\bm{c})=\bm{0}. Apply (1); choose k1,,kl𝒦k_{1},\dots,k_{l}\in\mathcal{K} with ll minimal. We now do casework on ll, making use of minimality if l2l\geq 2.

  • If l=1l=1, then c(k)3=0c(k)^{3}=0 for some k𝒦k\in\mathcal{K}.

  • If l=2l=2, then c(k1)3=c(k2)3×c(k_{1})^{3}=c(k_{2})^{3}\in\mathbb{Q}^{\times} for some distinct k1,k2𝒦k_{1},k_{2}\in\mathcal{K}.

  • If l3l\geq 3, then c(kt)3×c(k_{t})^{3}\in\mathbb{Q}^{\times} for all t[l]t\in[l], and by multi-quadratic field theory in characteristic 0, the square classes c(k1)3,,c(kl)3mod(×)2c(k_{1})^{3},\dots,c(k_{l})^{3}\bmod{(\mathbb{Q}^{\times})^{2}} must all coincide. (In fact, say we fix it𝒥(kt)i_{t}\in\mathcal{J}(k_{t}) for t[l]t\in[l]. Then cit/Fit=xit2dc_{i_{t}}/F_{i_{t}}=x_{i_{t}}^{2}d for some d,xit×d,x_{i_{t}}\in\mathbb{Q}^{\times} such that Fi1xi13++Filxil3=0F_{i_{1}}x_{i_{1}}^{3}+\dots+F_{i_{l}}x_{i_{l}}^{3}=0. So c(kt)3=Fit2xit6d3c(k_{t})^{3}=F_{i_{t}}^{2}x_{i_{t}}^{6}d^{3}.)

In each case, c(k1)3c(k2)3s()c(k_{1})^{3}c(k_{2})^{3}\in s(\mathbb{Q}) holds for some distinct k1,k2𝒦k_{1},k_{2}\in\mathcal{K}. ∎

Remark 3.11.

Though written in characteristic 0, the main results of §3.2 carry over to arbitrary fields of characteristic p(6m)!F1Fmp\nmid(6^{m})!F_{1}\cdots F_{m}, under (1.14). Such extensions of Proposition 3.7, (3.5), Corollary 3.9, and Observation 3.10(1) to 𝔽p\mathbb{F}_{p} will prove useful in §7.

4. Poisson summation for the endgame

Suppose 2m2\mid m. Fix LΥL\in\Upsilon, and recall Definition 1.3. Then F|Λ=0F|_{\Lambda}=0, and Λ\Lambda, Λ\Lambda^{\perp} are primitive rank-m2\frac{m}{2} sublattices of m\mathbb{Z}^{m}. Now choose bases 𝚲\bm{\Lambda}, 𝚲\bm{\Lambda^{\perp}} of Λ\Lambda, Λ\Lambda^{\perp}, respectively. Identify 𝚲\bm{\Lambda}, 𝚲\bm{\Lambda^{\perp}} with m×m2m\times\frac{m}{2} and m2×m\frac{m}{2}\times m integer matrices, respectively, so that Λ=𝚲m/2\Lambda=\bm{\Lambda}\mathbb{Z}^{m/2} and Λ=m/2𝚲\Lambda^{\perp}=\mathbb{Z}^{m/2}\bm{\Lambda^{\perp}} (where we view Λ\Lambda as a “column space” and Λ\Lambda^{\perp} as a “row space”).

We seek to prove Lemma 4.6 and Proposition 4.7 below, about certain averages over a given shifted dilate of Λ\Lambda^{\perp} (for the “endgame” of §8).

For the rest of §4, let 𝒙\bm{x}, 𝒉\bm{h}, 𝒙\bm{x}^{\prime} denote column vectors and 𝒄\bm{c}, 𝒃\bm{b}, 𝒗\bm{v} row vectors. In particular, the dot product 𝒄𝒙\bm{c}\cdot\bm{x} then coincides with standard matrix multiplication.

4.1. Preliminaries

We have Λ=(Λ)\Lambda=(\Lambda^{\perp})^{\perp}, i.e. Λ={𝒙m:𝚲𝒙=𝟎}\Lambda=\{\bm{x}\in\mathbb{Z}^{m}:\bm{\Lambda^{\perp}}\bm{x}=\bm{0}\}. Let

(4.1) σ,L,w\colonequalslimϵ0(2ϵ)m/2𝚲𝒙~[ϵ,ϵ]m/2𝑑𝒙~w(𝒙~).\sigma_{\infty,L^{\perp},w}\colonequals\lim_{\epsilon\to 0}{(2\epsilon)^{-m/2}\int_{\bm{\Lambda^{\perp}}\tilde{\bm{x}}\in[-\epsilon,\epsilon]^{m/2}}d\tilde{\bm{x}}\,w(\tilde{\bm{x}})}.

Then the formula (1.7) holds, by Poisson summation over Lm=ΛL\cap\mathbb{Z}^{m}=\Lambda (or, at least morally, by the circle method applied to 𝚲𝒙=𝟎\bm{\Lambda^{\perp}}\bm{x}=\bm{0}). In particular, σ,L,w\sigma_{\infty,L^{\perp},w} does not depend on the choice of 𝚲\bm{\Lambda^{\perp}}. For an alternative interpretation of σ,L,w\sigma_{\infty,L^{\perp},w}, see the second part of Lemma 4.6.

For calculations to come, it will help to extend 𝚲\bm{\Lambda^{\perp}} to a basis of m\mathbb{Z}^{m}.

Definition 4.1.

By primitivity of Λ\Lambda^{\perp}, choose Γ\Gamma (itself primitive) such that m=ΛΓ\mathbb{Z}^{m}=\Lambda^{\perp}\oplus\Gamma. Then choose a m2×m\frac{m}{2}\times m basis matrix 𝚪\mathbf{\Gamma} so that Γ=m/2𝚪\Gamma=\mathbb{Z}^{m/2}\bm{\Gamma}.

The rows of the m×mm\times m matrix [𝚲𝚪]\left[\begin{smallmatrix}\bm{\Lambda^{\perp}}\\ \bm{\Gamma}\end{smallmatrix}\right] form a basis of m\mathbb{Z}^{m}. Therefore,

(4.2) det[𝚲𝚪]=±1.\det\left[\begin{smallmatrix}\bm{\Lambda^{\perp}}\\ \bm{\Gamma}\end{smallmatrix}\right]=\pm 1.

Let RR denote a ring, e.g. \mathbb{R} or /n\mathbb{Z}/n\mathbb{Z}. Given a \mathbb{Z}-module AA, let AR\colonequalsARA_{R}\colonequals A\otimes R. To study Fourier transforms over 𝒙Rm\bm{x}\in R^{m} at 𝒄Λ\bm{c}\in\Lambda^{\perp}, it will help to rewrite 𝒄𝒙\bm{c}\cdot\bm{x}.

Definition 4.2.

Using the definition of 𝚲\bm{\Lambda^{\perp}} as a basis matrix, let ψ:Λm/2\psi\colon\Lambda^{\perp}\to\mathbb{Z}^{m/2} be the \mathbb{Z}-linear isomorphism 𝒄𝒄\bm{c}\mapsto\bm{c}^{\star} defined by the equation 𝒄=𝒄𝚲\bm{c}=\bm{c}^{\star}\bm{\Lambda^{\perp}}. Let ψR\colonequalsψR\psi_{R}\colonequals\psi\otimes R be the RR-linear isomorphism (Λ)RRm/2(\Lambda^{\perp})_{R}\to R^{m/2} induced by ψ\psi.

For each 𝒙Rm\bm{x}\in R^{m}, let [𝒉𝒙]\colonequals[𝚲𝚪]𝒙\left[\begin{smallmatrix}\bm{h}\\ \bm{x}^{\prime}\end{smallmatrix}\right]\colonequals\left[\begin{smallmatrix}\bm{\Lambda^{\perp}}\\ \bm{\Gamma}\end{smallmatrix}\right]\bm{x}. Equivalently, let

(4.3) 𝒉\colonequals𝚲𝒙Rm/2,𝒙\colonequals𝚪𝒙Rm/2.\bm{h}\colonequals\bm{\Lambda^{\perp}}\bm{x}\in R^{m/2},\qquad\bm{x}^{\prime}\colonequals\bm{\Gamma}\bm{x}\in R^{m/2}.
Example 4.3.

If F=x13++xm3F=x_{1}^{3}+\dots+x_{m}^{3} and Λ=𝒥\Lambda^{\perp}=\mathcal{R}_{\mathcal{J}} (in the notation of Proposition 3.6), then we can choose 𝚲\bm{\Lambda^{\perp}} so that hk=i𝒥(k)xih_{k}=\sum_{i\in\mathcal{J}(k)}x_{i} for all k𝒦k\in\mathcal{K}.

We need two more algebraic propositions. Over RR, let 𝒄𝒙:(Λ)R×RmR\bm{c}\cdot\bm{x}\colon(\Lambda^{\perp})_{R}\times R^{m}\to R denote the “obvious” RR-bilinear map induced by the usual dot product 𝒄𝒙:Λ×m\bm{c}\cdot\bm{x}\colon\Lambda^{\perp}\times\mathbb{Z}^{m}\to\mathbb{Z}.

Proposition 4.4.

If 𝐜(Λ)R\bm{c}\in(\Lambda^{\perp})_{R} and 𝐱Rm\bm{x}\in R^{m}, then 𝐜𝐱=ψR(𝐜)𝐡\bm{c}\cdot\bm{x}=\psi_{R}(\bm{c})\cdot\bm{h}.

Proof.

By RR-linearity properties, reduce to the case R=R=\mathbb{Z}, which is trivial. ∎

Let ΛRRm\Lambda\cdot R\subseteq R^{m} be the RR-module generated by Λ\Lambda (via the composite map ΛmRm\Lambda\to\mathbb{Z}^{m}\to R^{m}).

Proposition 4.5.

Let 𝐱Rm\bm{x}\in R^{m}. Then 𝐡=𝟎\bm{h}=\bm{0} if and only if 𝐱ΛR\bm{x}\in\Lambda\cdot R.

Proof.

The multiplication map 𝚲:mm/2\bm{\Lambda^{\perp}}\colon\mathbb{Z}^{m}\to\mathbb{Z}^{m/2} (given by 𝒙𝚲𝒙\bm{x}\mapsto\bm{\Lambda^{\perp}}\bm{x}) is surjective by (4.2), and thus defines an exact sequence Λmm/20\Lambda\to\mathbb{Z}^{m}\to\mathbb{Z}^{m/2}\to 0. The right exact functor R\otimes R therefore gives an exact sequence ΛRRmRm/20\Lambda_{R}\to R^{m}\to R^{m/2}\to 0. So

ker(𝚲:RmRm/2)=im(ΛRRm).\ker(\bm{\Lambda^{\perp}}\colon R^{m}\to R^{m/2})=\operatorname{im}(\Lambda_{R}\to R^{m}).

But im(ΛRRm)=ΛR\operatorname{im}(\Lambda_{R}\to R^{m})=\Lambda\cdot R, since ΛR\Lambda_{R} is RR-linearly generated by Λ\Lambda. ∎

4.2. Averaging the oscillatory integrals

Recall I𝒄(n)I_{\bm{c}}(n) from (1.4). Let R\colonequalsR\colonequals\mathbb{R} in (4.3). Since 𝒙(𝒉,𝒙)\bm{x}\mapsto(\bm{h},\bm{x}^{\prime}) is unimodular by (4.2), we have

(4.4) I𝒄(n)=(𝒉,𝒙)m𝑑𝒉𝑑𝒙w(𝒙/X)h(n/Y,F(𝒙)/Y2)e(𝒄𝒙/n),I_{\bm{c}}(n)=\int_{(\bm{h},\bm{x}^{\prime})\in\mathbb{R}^{m}}d\bm{h}\,d\bm{x}^{\prime}\,w(\bm{x}/X)h(n/Y,F(\bm{x})/Y^{2})e(-\bm{c}\cdot\bm{x}/n),

an integral in which we view 𝒙\bm{x} as a function of (𝒉,𝒙)(\bm{h},\bm{x}^{\prime}). For each 𝒉m/2\bm{h}\in\mathbb{Z}^{m/2}, let

J(𝒉;n)\colonequals𝒙m/2𝑑𝒙w(𝒙/X)h(n/Y,F(𝒙)/Y2).J(\bm{h};n)\colonequals\int_{\bm{x}^{\prime}\in\mathbb{R}^{m/2}}d\bm{x}^{\prime}\,w(\bm{x}/X)h(n/Y,F(\bm{x})/Y^{2}).

(The integral J(𝒉;n)J(\bm{h};n) corresponds to the integral Jq(𝒋)J_{q}(\bm{j}) on [heath1998circle]*p. 692. But in the present paper, we do not make any serious use of J(𝒉;n)J(\bm{h};n) for 𝒉𝟎\bm{h}\neq\bm{0}.)

Lemma 4.6.

Suppose n=n0n1n=n_{0}n_{1} (where n0,n11n_{0},n_{1}\geq 1 are integers) and 𝐛Λ\bm{b}\in\Lambda^{\perp}. Then

(4.5) n1m/2𝒄𝒃+n0ΛI𝒄(n)=𝒉n1m/2en(𝒃𝒉)J(𝒉;n).n_{1}^{-m/2}\sum_{\bm{c}\in\bm{b}+n_{0}\Lambda^{\perp}}I_{\bm{c}}(n)=\sum_{\bm{h}\in n_{1}\mathbb{Z}^{m/2}}e_{n}(-\bm{b}^{\star}\cdot\bm{h})J(\bm{h};n).

Here J(𝟎;n)=σ,L,wXm/2h(n/Y,0)J(\bm{0};n)=\sigma_{\infty,L^{\perp},w}X^{m/2}\cdot h(n/Y,0). Furthermore, if n1M1Xn_{1}\geq M_{1}X for a sufficiently large positive real M1𝚲,w1M_{1}\ll_{\bm{\Lambda^{\perp}},w}1, then J(𝐡;n)=0J(\bm{h};n)=0 for all nonzero 𝐡n1m/2\bm{h}\in n_{1}\mathbb{Z}^{m/2}.

Proof.

Write 𝒄=𝒃+n0𝒗\bm{c}=\bm{b}+n_{0}\bm{v} for 𝒗Λ\bm{v}\in\Lambda^{\perp}. Define 𝒄,𝒃,𝒗m/2\bm{c}^{\star},\bm{b}^{\star},\bm{v}^{\star}\in\mathbb{Z}^{m/2} using Definition 4.2. Proposition 4.4 then delivers the equality 𝒄𝒙/n=𝒄𝒉/n=𝒃𝒉/n+𝒗𝒉/n1\bm{c}\cdot\bm{x}/n=\bm{c}^{\star}\cdot\bm{h}/n=\bm{b}^{\star}\cdot\bm{h}/n+\bm{v}^{\star}\cdot\bm{h}/n_{1}. So by (4.4), the integral I𝒄(n)=I𝒃+n0𝒗(n)I_{\bm{c}}(n)=I_{\bm{b}+n_{0}\bm{v}}(n) is the Fourier transform at 𝒗/n1m/2\bm{v}^{\star}/n_{1}\in\mathbb{R}^{m/2} of the function

m/2,𝒉e(𝒃𝒉/n)J(𝒉;n).\mathbb{R}^{m/2}\to\mathbb{R},\quad\bm{h}\mapsto e(-\bm{b}^{\star}\cdot\bm{h}/n)J(\bm{h};n).

Poisson summation over 𝒉n1m/2\bm{h}\in n_{1}\mathbb{Z}^{m/2} hence yields

𝒉n1m/2e(𝒃𝒉/n)J(𝒉;n)=n1m/2𝒗m/2I𝒃+n0𝒗(n),\sum_{\bm{h}\in n_{1}\mathbb{Z}^{m/2}}e(-\bm{b}^{\star}\cdot\bm{h}/n)J(\bm{h};n)=n_{1}^{-m/2}\sum_{\bm{v}^{\star}\in\mathbb{Z}^{m/2}}I_{\bm{b}+n_{0}\bm{v}}(n),

which implies (4.5) (upon recalling the correspondence between 𝒗m/2\bm{v}^{\star}\in\mathbb{Z}^{m/2} and 𝒗Λ\bm{v}\in\Lambda^{\perp}).

We now compute J(𝟎;n)J(\bm{0};n). Let 𝒙~\colonequals𝒙/X\tilde{\bm{x}}\colonequals\bm{x}/X. Since F|Λ=0F|_{\Lambda\cdot\mathbb{R}}=0, we have

J(𝟎;n)=Xm/2h(n/Y,0)(𝒉~,𝒙~){𝟎}×m/2𝑑𝒙~w(𝒙~).J(\bm{0};n)=X^{m/2}\cdot h(n/Y,0)\cdot\int_{(\tilde{\bm{h}},\tilde{\bm{x}}^{\prime})\in\{\bm{0}\}\times\mathbb{R}^{m/2}}d\tilde{\bm{x}}^{\prime}\,w(\tilde{\bm{x}}).

But an ϵ\epsilon-thickening in 𝒉~\tilde{\bm{h}}, followed by an application of (4.1), yields

(𝒉~,𝒙~){𝟎}×m/2𝑑𝒙~w(𝒙~)=limϵ0(𝒉~,𝒙~)m𝑑𝒉~𝑑𝒙~w(𝒙~)𝟏𝒉~[ϵ,ϵ]m/2(2ϵ)m/2=σ,L,w.\int_{(\tilde{\bm{h}},\tilde{\bm{x}}^{\prime})\in\{\bm{0}\}\times\mathbb{R}^{m/2}}d\tilde{\bm{x}}^{\prime}\,w(\tilde{\bm{x}})=\lim_{\epsilon\to 0}{\int_{(\tilde{\bm{h}},\tilde{\bm{x}}^{\prime})\in\mathbb{R}^{m}}d\tilde{\bm{h}}\,d\tilde{\bm{x}}^{\prime}\,w(\tilde{\bm{x}})\cdot\frac{\bm{1}_{\tilde{\bm{h}}\in[-\epsilon,\epsilon]^{m/2}}}{(2\epsilon)^{m/2}}}=\sigma_{\infty,L^{\perp},w}.

Finally, suppose 𝒉n1m/2\bm{h}\in n_{1}\mathbb{Z}^{m/2} and J(𝒉;n)0J(\bm{h};n)\neq 0. Take 𝒙m/2\bm{x}^{\prime}\in\mathbb{R}^{m/2} with w(𝒙/X)0w(\bm{x}/X)\neq 0. If n1M1Xn_{1}\geq M_{1}X, then n1𝒉=𝚲𝒙Xn_{1}\mid\bm{h}=\bm{\Lambda^{\perp}}\bm{x}\ll X, so 𝒉=𝟎\bm{h}=\bm{0} if M1M_{1} is sufficiently large. ∎

4.3. Vertically averaging the exponential sums

Recall S𝒄(n)S_{\bm{c}}(n) from (1.5), for each tuple 𝒄m\bm{c}\in\mathbb{Z}^{m} and integer n1n\geq 1. Let ϕ(n)\phi(n) denote Euler’s totient function.

Proposition 4.7.

The quantity S𝐜(n)S_{\bm{c}}(n) is a function of nn and 𝐜modn\bm{c}\bmod{n}. If 𝐣m/2\bm{j}\in\mathbb{Z}^{m/2}, then

(4.6) 𝔼𝒄Λ/nΛ[S𝒄(n)en(𝒄𝒋)]=1an:gcd(a,n)=11𝒙nen(aF(𝒙))𝟏n𝒉𝒋\mathbb{E}_{\bm{c}\in\Lambda^{\perp}/n\Lambda^{\perp}}[S_{\bm{c}}(n)e_{n}(-\bm{c}^{\star}\cdot\bm{j})]=\sum_{1\leq a\leq n:\,\gcd(a,n)=1}\,\sum_{1\leq\bm{x}\leq n}e_{n}(aF(\bm{x}))\cdot\bm{1}_{n\mid\bm{h}-\bm{j}}

(where the left-hand side is well-defined). In particular, 𝔼𝐜Λ/nΛ[S𝐜(n)]=ϕ(n)nm/2\mathbb{E}_{\bm{c}\in\Lambda^{\perp}/n\Lambda^{\perp}}[S_{\bm{c}}(n)]=\phi(n)n^{m/2}.

Proof.

The first sentence is clear by definition of S𝒄(n)S_{\bm{c}}(n). (Here 𝒄modn\bm{c}\bmod{n} refers to 𝒄modnm\bm{c}\bmod{n\mathbb{Z}^{m}}.)

A priori, if 𝒄Λ\bm{c}\in\Lambda^{\perp}, then 𝒄modnΛ\bm{c}\bmod{n\Lambda^{\perp}} determines 𝒄modn\bm{c}\bmod{n}. So (4.6) makes sense. Since Λ\Lambda^{\perp} is primitive, one can say more about “congruence modulo nn”, but we need not do so.

Now fix 𝒋m/2\bm{j}\in\mathbb{Z}^{m/2}. Let R\colonequals/nR\colonequals\mathbb{Z}/n\mathbb{Z}. Via the map ψR\psi_{R} from Definition 4.2, elements 𝒄Λ/nΛ=(Λ)R\bm{c}\in\Lambda^{\perp}/n\Lambda^{\perp}=(\Lambda^{\perp})_{R} correspond isomorphically to 𝒄Rm/2=(/n)m/2\bm{c}^{\star}\in R^{m/2}=(\mathbb{Z}/n\mathbb{Z})^{m/2}, so that if 𝒙Rm\bm{x}\in R^{m}, then 𝒄𝒙=𝒄𝒉\bm{c}\cdot\bm{x}=\bm{c}^{\star}\cdot\bm{h} by Proposition 4.4. By (1.5), it follows that

𝔼𝒄Λ/nΛ[S𝒄(n)en(𝒄𝒋)]=aR×𝒙Rmen(aF(𝒙))𝔼𝒄Rm/2[en(𝒄𝒉𝒄𝒋)].\mathbb{E}_{\bm{c}\in\Lambda^{\perp}/n\Lambda^{\perp}}[S_{\bm{c}}(n)e_{n}(-\bm{c}^{\star}\cdot\bm{j})]=\sum_{a\in R^{\times}}\sum_{\bm{x}\in R^{m}}e_{n}(aF(\bm{x}))\cdot\mathbb{E}_{\bm{c}^{\star}\in R^{m/2}}[e_{n}(\bm{c}^{\star}\cdot\bm{h}-\bm{c}^{\star}\cdot\bm{j})].

Summing over 𝒄\bm{c}^{\star} gives (4.6). Furthermore, if we take 𝒋=𝟎\bm{j}=\bm{0}, then by Proposition 4.5 and the identity F|ΛR=0F|_{\Lambda\cdot R}=0, the right-hand side of (4.6) simplifies to ϕ(n)nm/2\phi(n)n^{m/2}. ∎

Let T(𝒋;n)T(\bm{j};n) denote the right-hand side of (4.6). Then T(𝒋;n)T(\bm{j};n) corresponds to the sum Tq(𝒋)T_{q}(\bm{j}) on [heath1998circle]*p. 692. It would be interesting to better understand T(𝒋;n)T(\bm{j};n) in general. The sums T(𝒋;n)T(\bm{j};n) are multiplicative in nn, and related to the singular series of certain m2\frac{m}{2}-variable affine quadrics varying with 𝒋\bm{j}. Below, however, we only make use of the 𝒋=𝟎\bm{j}=\bm{0} case of (4.6).

5. General upper bounds

We first recall some background on (1.3). Let 𝒄\colonequalsmax(|c1|,,|cm|)\lVert\bm{c}\rVert\colonequals\max(\lvert c_{1}\rvert,\dots,\lvert c_{m}\rvert). Recall (1.8).

Proposition 5.1 (See e.g. [wang2022thesis]*§3.1, or [heath1998circle]*(2.3)–(3.7)).

The following hold:

  1. (1)

    I𝒄(n)I_{\bm{c}}(n) vanishes over nM2Yn\geq M_{2}Y, for some positive M2F,w1M_{2}\ll_{F,w}1 independent of 𝒄m\bm{c}\in\mathbb{Z}^{m}.

  2. (2)

    For any fixed ϵ,A>0\epsilon,A>0, we have I𝒄(n)F,w,ϵ,AXAI_{\bm{c}}(n)\ll_{F,w,\epsilon,A}X^{-A} over 𝒄X1/2+ϵ\lVert\bm{c}\rVert\geq X^{1/2+\epsilon}.

  3. (3)

    We have

    n1𝒄mn(1m)/2|S𝒄(n)I𝒄(n)|=n1𝒄mnm|S𝒄(n)I𝒄(n)|<.\sum_{n\geq 1}\sum_{\bm{c}\in\mathbb{Z}^{m}}n^{(1-m)/2}\lvert S^{\natural}_{\bm{c}}(n)I_{\bm{c}}(n)\rvert=\sum_{n\geq 1}\sum_{\bm{c}\in\mathbb{Z}^{m}}n^{-m}\lvert S_{\bm{c}}(n)I_{\bm{c}}(n)\rvert<\infty.

Now fix a set 𝒮{𝒄m:F(𝒄)=0}\mathcal{S}\subseteq\{\bm{c}\in\mathbb{Z}^{m}:F^{\vee}(\bm{c})=0\} and a real δ0\delta\geq 0. Suppose 𝒮[C,C]m\mathcal{S}\cap[-C,C]^{m} has size O(Cm/2δ)O(C^{m/2-\delta}) for all reals C1C\geq 1. Let

(5.1) f(𝒮)\colonequalsY2𝒄𝒮{𝟎}n1|I𝒄(n)|n1m/2nnn1/2|S𝒄(n)|.f(\mathcal{S})\colonequals Y^{-2}\sum_{\bm{c}\in\mathcal{S}\setminus\{\bm{0}\}}\sum_{n\geq 1}\lvert I_{\bm{c}}(n)\rvert\cdot n^{1-m/2}\cdot\sum_{n_{\star}\mid n}n_{\star}^{-1/2}\lvert S^{\natural}_{\bm{c}}(n_{\star})\rvert.

At several points in §8, Lemma 5.2 will let us cleanly discard f(𝒮)f(\mathcal{S}) for various choices of 𝒮\mathcal{S}.

Lemma 5.2.

Assume FF is diagonal. Suppose {q𝐜:(q,𝐜)××𝒮}m=𝒮\{q\bm{c}:(q,\bm{c})\in\mathbb{Q}^{\times}\times\mathcal{S}\}\cap\mathbb{Z}^{m}=\mathcal{S}. Assume m6m\leq 6 and δ12(m2)\delta\leq\frac{1}{2}(m-2). Then f(𝒮)ϵX(mδ)/2+ϵf(\mathcal{S})\ll_{\epsilon}X^{(m-\delta)/2+\epsilon}.

The work [heath1998circle]*pp. 688–689 proves a simplified version of Lemma 5.2 with (m,δ){(4,1),(6,0)}(m,\delta)\in\{(4,1),(6,0)\}, and with 𝟏n=n\bm{1}_{n_{\star}=n} instead of nn\sum_{n_{\star}\mid n}. The method of [heath1998circle] should directly extend to Lemma 5.2. However, we transpose Heath-Brown’s argument a bit in order to highlight an intermediate result that one may hope to generalize: Proposition 5.3. Proposition 5.3 offers a potential partial alternative to axiom (3) in Remark 1.6.

Let vp()v_{p}(-) denote the usual pp-adic valuation. 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)}; and for convenience, let sq(0)\colonequals0\operatorname{sq}(0)\colonequals 0 and cub(0)\colonequals0\operatorname{cub}(0)\colonequals 0. A positive integer nn is said to be square-full if n=sq(n)n=\operatorname{sq}(n), and cube-full if n=cub(n)n=\operatorname{cub}(n). In the absence of a deeper algebro-geometric understanding of S𝒄S_{\bm{c}}, one relies heavily on the following bound of Hooley and Heath-Brown, valid for diagonal FF (for all n1n\geq 1 and 𝒄m\bm{c}\in\mathbb{Z}^{m}):

(5.2) n1/2S𝒄(n)F,ϵnϵ1jmgcd(cub(n)2,gcd(cub(n),sq(cj))3)1/12.n^{-1/2}S^{\natural}_{\bm{c}}(n)\ll_{F,\epsilon}n^{\epsilon}\prod_{1\leq j\leq m}\gcd\bigl{(}\operatorname{cub}(n)^{2},\gcd(\operatorname{cub}(n),\operatorname{sq}(c_{j}))^{3}\bigr{)}^{1/12}.

(See [wang2023_large_sieve_diagonal_cubic_forms]*Proposition 4.9 or [wang2022thesis]*Proposition 3.3.3 for precise references.)

Suppose m6m\leq 6, and assume FF is diagonal. Let C,N{1,2,4,8,}C,N\in\{1,2,4,8,\ldots\}. For each i[m]i\in[m], the cidegFc_{i}^{\deg F^{\vee}} coefficient of the homogeneous polynomial F[𝒄]F^{\vee}\in\mathbb{Z}[\bm{c}] is nonzero, by (3.3). Suppose 𝒄𝒮\bm{c}\in\mathcal{S} with C𝒄<2CC\leq\lVert\bm{c}\rVert<2C; then |ci|C\lvert c_{i}\rvert\gg C for at least two indices i[m]i\in[m]. In particular,

(5.3) I𝒄(n)ϵXm+ϵ(XC/N)(2m)/4,I_{\bm{c}}(n)\ll_{\epsilon}X^{m+\epsilon}(XC/N)^{(2-m)/4},

by [heath1998circle]*p. 688, (7.3). We also have the following result (for 𝒄\bm{c}, NN as above):

Proposition 5.3.

Here Nn<2Nnnn1/2|S𝐜(n)|ϵ(CN)ϵcub(gcd(𝐜))1/6N\sum_{N\leq n<2N}\sum_{n_{\star}\mid n}n_{\star}^{-1/2}\lvert S^{\natural}_{\bm{c}}(n_{\star})\rvert\ll_{\epsilon}(CN)^{\epsilon}\operatorname{cub}(\gcd(\bm{c}))^{1/6}N.

Proof.

Let MF\colonequals1imFiM_{F}\colonequals\prod_{1\leq i\leq m}F_{i}. Write 𝒄=g𝒄\bm{c}=g\bm{c}^{\prime} with g\colonequalsgcd(𝒄)>0g\colonequals\gcd(\bm{c})>0, so that 𝒄\bm{c}^{\prime} is primitive. For each prime pp, the equation F(𝒄)=0F^{\vee}(\bm{c}^{\prime})=0 implies, via (3.3), that #{i[m]:vp(ci)vp(MF)}2\#\{i\in[m]:v_{p}(c^{\prime}_{i})\leq v_{p}(M_{F})\}\geq 2. The bound (5.2) for nnn_{\star}\mid n now implies

nnn1/2|S𝒄(n)|ϵnϵcub(n)(m2)/6gcd(cub(n),sq(MFg))2/4\sum_{n_{\star}\mid n}n_{\star}^{-1/2}\lvert S^{\natural}_{\bm{c}}(n_{\star})\rvert\ll_{\epsilon}n^{\epsilon}\operatorname{cub}(n)^{(m-2)/6}\gcd(\operatorname{cub}(n),\operatorname{sq}(M_{F}\cdot g))^{2/4}

(since cub(n)cub(n)\operatorname{cub}(n_{\star})\mid\operatorname{cub}(n), and vp(sq(ci))vp(sq(MFg))v_{p}(\operatorname{sq}(c_{i}))\leq v_{p}(\operatorname{sq}(M_{F}\cdot g)) whenever vp(ci)vp(MF)v_{p}(c^{\prime}_{i})\leq v_{p}(M_{F})). Thus

(5.4) Nn<2Nnnn1/2|S𝒄(n)|ϵdsq(MFg)q<2N:q=cub(q)(N/q)q(m2)/6d1/2𝟏dq.\sum_{N\leq n<2N}\,\sum_{n_{\star}\mid n}n_{\star}^{-1/2}\lvert S^{\natural}_{\bm{c}}(n_{\star})\rvert\ll_{\epsilon}\sum_{d\mid\operatorname{sq}(M_{F}\cdot g)}\,\sum_{\begin{subarray}{c}q<2N:\,q=\operatorname{cub}(q)\end{subarray}}(N/q)\cdot q^{(m-2)/6}d^{1/2}\bm{1}_{d\mid q}.

However, for any integer d1d\geq 1 and real ϵ>0\epsilon>0, the sum q<2N:q=cub(q)(2N)ϵq1/3d1/2𝟏dq\sum_{\begin{subarray}{c}q<2N:\,q=\operatorname{cub}(q)\end{subarray}}(2N)^{-\epsilon}q^{-1/3}d^{1/2}\bm{1}_{d\mid q} is

q1:q=cub(q)d1/2𝟏dqq1/3+ϵ=(pdO(pvp(d)/2)pmax(1,vp(d)/3))pd(1+O(p13ϵ))ϵdϵcub(d)1/6.\leq\sum_{q\geq 1:\,q=\operatorname{cub}(q)}\frac{d^{1/2}\bm{1}_{d\mid q}}{q^{1/3+\epsilon}}=\biggl{(}\,\prod_{p\mid d}\frac{O(p^{v_{p}(d)/2})}{p^{\max(1,v_{p}(d)/3)}}\biggr{)}\prod_{p\nmid d}(1+O(p^{-1-3\epsilon}))\ll_{\epsilon}d^{\epsilon}\operatorname{cub}(d)^{1/6}.

Since (m2)/62/3(m-2)/6\leq 2/3, the right-hand side of (5.4) is therefore ϵ(CN)ϵcub(MFg)1/6N\ll_{\epsilon}(CN)^{\epsilon}\operatorname{cub}(M_{F}\cdot g)^{1/6}N. But cub(MFg)MF3cub(g)\operatorname{cub}(M_{F}\cdot g)\mid M_{F}^{3}\operatorname{cub}(g), so Proposition 5.3 follows. ∎

Proof of Lemma 5.2.

By Proposition 5.1, (5.3), Proposition 5.3, and dyadic decomposition of 𝒄,n\lVert\bm{c}\rVert,n, the quantity f(𝒮)f(\mathcal{S}) (see (5.1)) is

ϵ,AXA+Y2Xm+ϵC,N{1,2,4,8,}:1CX1/2+ϵ, 1NM2Y(XC/N)(2m)/4N2m/2𝒄𝒮:C𝒄<2Ccub(gcd(𝒄))1/6.\ll_{\epsilon,A}X^{-A}+Y^{-2}X^{m+\epsilon}\sum_{\begin{subarray}{c}C,N\in\{1,2,4,8,\ldots\}:\\ 1\leq C\leq X^{1/2+\epsilon},\;1\leq N\leq M_{2}Y\end{subarray}}(XC/N)^{(2-m)/4}N^{2-m/2}\sum_{\begin{subarray}{c}\bm{c}\in\mathcal{S}:\\ C\leq\lVert\bm{c}\rVert<2C\end{subarray}}\operatorname{cub}(\gcd(\bm{c}))^{1/6}.

Our hypotheses on 𝒮\mathcal{S} imply

(5.5) 𝒄𝒮:C𝒄<2Ccub(gcd(𝒄))1/6g<2Ccub(g)1/6(2C/g)m/2δ.\sum_{\bm{c}\in\mathcal{S}:\,C\leq\lVert\bm{c}\rVert<2C}\operatorname{cub}(\gcd(\bm{c}))^{1/6}\leq\sum_{g<2C}\operatorname{cub}(g)^{1/6}\cdot(2C/g)^{m/2-\delta}.

But the Dirichlet series g1cub(g)1/6gs\sum_{g\geq 1}\operatorname{cub}(g)^{1/6}g^{-s} converges absolutely for (s)>1\Re(s)>1. Since m/2δ1m/2-\delta\geq 1, the right-hand side of (5.5) is therefore ϵCm/2δ+ϵ\ll_{\epsilon}C^{m/2-\delta+\epsilon}. So

(5.6) f(𝒮)ϵ,AXA+Y2Xm+2ϵC,N{1,2,4,8,}:1CX1/2+ϵ, 1NM2Y(XC/N)(2m)/4N2m/2Cm/2δ.f(\mathcal{S})\ll_{\epsilon,A}X^{-A}+Y^{-2}X^{m+2\epsilon}\sum_{\begin{subarray}{c}C,N\in\{1,2,4,8,\ldots\}:\\ 1\leq C\leq X^{1/2+\epsilon},\;1\leq N\leq M_{2}Y\end{subarray}}(XC/N)^{(2-m)/4}N^{2-m/2}C^{m/2-\delta}.

The total exponent of NN in (5.6) is m24+2m2=6m40\frac{m-2}{4}+2-\frac{m}{2}=\frac{6-m}{4}\geq 0, and the total exponent of CC in (5.6) is 2m4+m2δ=m+24δ0\frac{2-m}{4}+\frac{m}{2}-\delta=\frac{m+2}{4}-\delta\geq 0 (since m+24m22\frac{m+2}{4}\geq\frac{m-2}{2} for m6m\leq 6). It follows that

f(𝒮)ϵY2Xm+3ϵ(X3/2+ϵ/Y)(2m)/4Y2m/2(X1/2+ϵ)m/2δ.f(\mathcal{S})\ll_{\epsilon}Y^{-2}X^{m+3\epsilon}(X^{3/2+\epsilon}/Y)^{(2-m)/4}Y^{2-m/2}(X^{1/2+\epsilon})^{m/2-\delta}.

Since Y=X3/2Y=X^{3/2}, we get f(𝒮)ϵXm+O(mϵ)X3m/4Xm/4δ/2=Xm/2δ/2+O(mϵ)f(\mathcal{S})\ll_{\epsilon}X^{m+O(m\epsilon)}X^{-3m/4}X^{m/4-\delta/2}=X^{m/2-\delta/2+O(m\epsilon)}. ∎

Now drop the earlier assumptions “m6m\leq 6” and “FF is diagonal”. The axioms in Remark 1.6 would allow us to prove a version of Lemma 5.2 without assuming FF is diagonal.

Lemma 5.4.

Assume 1.6(2)–(3). Say δmin(1,23(10m))\delta\leq\min(1,\frac{2}{3}(10-m)). Then f(𝒮)ϵXm/2δ/4+ϵf(\mathcal{S})\ll_{\epsilon}X^{m/2-\delta/4+\epsilon}.

When m9m\leq 9, this beats the square-root threshold Xm/2X^{m/2}, corresponding to linear subspaces. (Some degenerate ranges of 𝒄\bm{c}, nn seem to prevent us from handling m10m\geq 10.)

Proof sketch for Lemma 5.4.

Under 1.6(2), it is known that

(5.7) I𝒄(n)ϵXm+ϵ(1+X𝒄/n)1m/2;I_{\bm{c}}(n)\ll_{\epsilon}X^{m+\epsilon}(1+X\lVert\bm{c}\rVert/n)^{1-m/2};

see e.g. [hooley2014octonary]*p. 252, (31). It is also known that if nn is cube-free, then

(5.8) n1/2S𝒄(n)ϵnϵ;n^{-1/2}S^{\natural}_{\bm{c}}(n)\ll_{\epsilon}n^{\epsilon};

see e.g. [hooley2014octonary]*Lemmas 8–9. Fix ϵ>0\epsilon>0. For integers C,N1C,N\geq 1, let

B(C,N)\colonequals(Cm/2δ)1/2(Cm/2+Nm/6)1/2+(C(m1)/2N1/6+Nm/6).B(C,N)\colonequals(C^{m/2-\delta})^{1/2}(C^{m/2}+N^{m/6})^{1/2}+(C^{(m-1)/2}N^{1/6}+N^{m/6}).

By (1.13), plus Cauchy on (1.12) over 𝒮\mathcal{S}, it follows that if XX is sufficiently large, then

f(𝒮)ϵY2Xm+ϵ(1+XC/N)1m/2N1m/2(N/N)N1/3B(C,N)f(\mathcal{S})\ll_{\epsilon}Y^{-2}X^{m+\epsilon}(1+XC/N)^{1-m/2}N^{1-m/2}(N/N_{\star})\cdot N_{\star}^{1/3}B(C,N_{\star})

for some choice of C,N,N{1,2,4,8,}C,N,N_{\star}\in\{1,2,4,8,\ldots\} with CX1/2+ϵC\leq X^{1/2+\epsilon} and NNYN_{\star}\leq N\leq Y. Optimizing NN_{\star} over 1NN1\leq N_{\star}\leq N yields f(𝒮)ϵXm3+ϵ(N+XC)1m/2(NB(C,1)+N1/3B(C,N))f(\mathcal{S})\ll_{\epsilon}X^{m-3+\epsilon}(N+XC)^{1-m/2}(N\cdot B(C,1)+N^{1/3}\cdot B(C,N)). But B(C,1)Cm/2δ/2B(C,1)\ll C^{m/2-\delta/2} (since δ1\delta\leq 1) and B(C,N)(C1/2+N1/6)mB(C,N)\ll(C^{1/2}+N^{1/6})^{m}, so

f(𝒮)ϵXm3+ϵ(N+XC)1m/2(NCm/2δ/2+N1/3Cm/2+N1/3+m/6).f(\mathcal{S})\ll_{\epsilon}X^{m-3+\epsilon}(N+XC)^{1-m/2}(NC^{m/2-\delta/2}+N^{1/3}C^{m/2}+N^{1/3+m/6}).

Let M\colonequalsN+XCM\colonequals N+XC, and plug in the bounds NMN\leq M and CM/XC\leq M/X, to get

f(𝒮)/Xm3+ϵϵM2δ/2/Xm/2δ/2+M4/3/Xm/2+M4/3m/3.f(\mathcal{S})/X^{m-3+\epsilon}\ll_{\epsilon}M^{2-\delta/2}/X^{m/2-\delta/2}+M^{4/3}/X^{m/2}+M^{4/3-m/3}.

But XM2X3/2+ϵX\leq M\leq 2X^{3/2+\epsilon}, so f(𝒮)ϵX3ϵ(Xm/2δ/4+Xm/21+X2m/35/3)f(\mathcal{S})\ll_{\epsilon}X^{3\epsilon}(X^{m/2-\delta/4}+X^{m/2-1}+X^{2m/3-5/3}). Here m21m2δ4\frac{m}{2}-1\leq\frac{m}{2}-\frac{\delta}{4} and 2m353m2δ4\frac{2m}{3}-\frac{5}{3}\leq\frac{m}{2}-\frac{\delta}{4}, since δmin(1,23(10m))\delta\leq\min(1,\frac{2}{3}(10-m)). ∎

6. Estimates at the center

In this section, we collect some standard facts we need about the quantities I𝟎(n)I_{\bm{0}}(n), S𝟎(n)S_{\bm{0}}(n) defined in (1.4) and (1.5). By [heath1996new]*Lemma 16, we have

(6.1) I𝟎(n)Xm.I_{\bm{0}}(n)\ll X^{m}.

By (6.1) and [heath1996new]*Lemma 13, we have (uniformly over X,n1X,n\geq 1)

(6.2) XmI𝟎(n)=σ,F,w+OA((n/Y)A)X^{-m}I_{\bm{0}}(n)=\sigma_{\infty,F,w}+O_{A}((n/Y)^{A})

(for all A>0A>0), where

(6.3) 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.

Aside from the real density σ,F,w\sigma_{\infty,F,w}, we also have the singular series

(6.4) 𝔖F\colonequalsn1nmS𝟎(n)=n1n(1m)/2S𝟎(n),\mathfrak{S}_{F}\colonequals\sum_{n\geq 1}n^{-m}S_{\bm{0}}(n)=\sum_{n\geq 1}n^{(1-m)/2}S^{\natural}_{\bm{0}}(n),

which converges absolutely for m5m\geq 5 (as one can show using Lemma 6.1, for instance).

Lemma 6.1.

Assume m4m\geq 4. If N{1,2,4,8,}N\in\{1,2,4,8,\ldots\}, then

Nn<2Nn1m/2dnd1/2|S𝟎(d)|ϵN(4m)/3+ϵ.\sum_{N\leq n<2N}n^{1-m/2}\sum_{d\mid n}d^{-1/2}\lvert S^{\natural}_{\bm{0}}(d)\rvert\ll_{\epsilon}N^{(4-m)/3+\epsilon}.
Proof.

We have S𝟎(d)ϵd1/2+ϵcub(d)m/6S^{\natural}_{\bm{0}}(d)\ll_{\epsilon}d^{1/2+\epsilon}\operatorname{cub}(d)^{m/6} by (5.2) for diagonal FF, and by [hooley1988nonary]*p. 95, (170) in general. Thus dnd1/2|S𝟎(d)|ϵnϵcub(n)m/6\sum_{d\mid n}d^{-1/2}\lvert S^{\natural}_{\bm{0}}(d)\rvert\ll_{\epsilon}n^{\epsilon}\operatorname{cub}(n)^{m/6}. But

Nn<2Ncub(n)m/6n3<2N:n3=cub(n3)n3m/6(N/n3)ϵNNm/62/3+ϵn3=cub(n3)n31/3ϵ,\sum_{N\leq n<2N}\operatorname{cub}(n)^{m/6}\ll\sum_{n_{3}<2N:\,n_{3}=\operatorname{cub}(n_{3})}n_{3}^{m/6}\cdot(N/n_{3})\ll_{\epsilon}N\cdot N^{m/6-2/3+\epsilon}\sum_{n_{3}=\operatorname{cub}(n_{3})}n_{3}^{-1/3-\epsilon},

since m/62/30m/6-2/3\geq 0. Observe that n3=cub(n3)n31/3ϵϵ1\sum_{n_{3}=\operatorname{cub}(n_{3})}n_{3}^{-1/3-\epsilon}\ll_{\epsilon}1, and multiply by N1m/2+ϵN^{1-m/2+\epsilon}. ∎

Since I𝟎(n)=0I_{\bm{0}}(n)=0 for nM2Yn\geq M_{2}Y (by Proposition 5.1), a routine calculation444with the same numerics as the diagonal treatment in [wang2023_large_sieve_diagonal_cubic_forms]*§6 using (6.2), Lemma 6.1, and (6.4) gives, for m5m\geq 5, the equality

(6.5) Y2n1nmS𝟎(n)I𝟎(n)=σ,F,w𝔖FXm3+Oϵ(Xm/21+ϵ).Y^{-2}\sum_{n\geq 1}n^{-m}S_{\bm{0}}(n)I_{\bm{0}}(n)=\sigma_{\infty,F,w}\mathfrak{S}_{F}X^{m-3}+O_{\epsilon}(X^{m/2-1+\epsilon}).
Remark 6.2.

The presence of dn\sum_{d\mid n} in Lemma 6.1 is not important for (6.5) (where 𝟏d=n\bm{1}_{d=n} would suffice in place of dn\sum_{d\mid n}), but rather for Lemma 8.12 (to help separate 𝒄=𝟎\bm{c}=\bm{0} from 𝒄𝟎\bm{c}\neq\bm{0}).

7. Establishing bias in exponential sums

In this section, we will realize (1.10) from §1. We will start with general theory, and gradually impose restrictions. It will be convenient to work over the pp-adic integers, p\mathbb{Z}_{p}.

Let pp be a prime. Certain hyperplane sections govern the behavior of S𝒄(pl)S_{\bm{c}}(p^{l}) for l1l\geq 1, as 𝒄\bm{c} ranges over m\mathbb{Z}^{m} or more generally, pm\mathbb{Z}_{p}^{m}.555The formula for S𝒄(pl)S_{\bm{c}}(p^{l}) in (1.5) still makes sense for 𝒄pm\bm{c}\in\mathbb{Z}_{p}^{m}, because 𝒄modpl/pl\bm{c}\bmod{p^{l}}\in\mathbb{Z}/p^{l}\mathbb{Z}. Let 𝒱\mathcal{V} and 𝒱𝒄\mathcal{V}_{\bm{c}} denote the closed subschemes of pm1\mathbb{P}^{m-1}_{\mathbb{Z}_{p}} defined by the equations F(𝒙)=0F(\bm{x})=0 and F(𝒙)=𝒄𝒙=0F(\bm{x})=\bm{c}\cdot\bm{x}=0, respectively. Throughout the present §7 only, let V\colonequals𝒱𝔽pV\colonequals\mathcal{V}_{\mathbb{F}_{p}} and V𝒄\colonequals(𝒱𝒄)𝔽pV_{\bm{c}}\colonequals(\mathcal{V}_{\bm{c}})_{\mathbb{F}_{p}}, so that VV and V𝒄V_{\bm{c}} live over 𝔽p\mathbb{F}_{p} (not \mathbb{Q}).

We now recall some standard background recorded (with references) in [wang2022thesis]*§3.2. Let m\colonequalsm3m_{\ast}\colonequals m-3. In terms of the point counts |V(𝔽p)|\lvert V(\mathbb{F}_{p})\rvert and |V𝒄(𝔽p)|\lvert V_{\bm{c}}(\mathbb{F}_{p})\rvert over 𝔽p\mathbb{F}_{p}, let

E(p)\colonequals|V(𝔽p)|(pm11)/(p1),E𝒄(p)\colonequals|V𝒄(𝔽p)|(pm21)/(p1).E(p)\colonequals\lvert V(\mathbb{F}_{p})\rvert-(p^{m-1}-1)/(p-1),\quad E_{\bm{c}}(p)\colonequals\lvert V_{\bm{c}}(\mathbb{F}_{p})\rvert-(p^{m-2}-1)/(p-1).

Then let E(p)\colonequalsp(m+1)/2E(p)E^{\natural}(p)\colonequals p^{-(m_{\ast}+1)/2}E(p) and E𝒄(p)\colonequalspm/2E𝒄(p)E^{\natural}_{\bm{c}}(p)\colonequals p^{-m_{\ast}/2}E_{\bm{c}}(p). By Deligne’s resolution of the Weil conjectures,

(7.1) E(p)F1.E^{\natural}(p)\ll_{F}1.

Furthermore, whenever p𝒄p\nmid\bm{c}, we have S𝒄(p)=p2E𝒄(p)pE(p)S_{\bm{c}}(p)=p^{2}E_{\bm{c}}(p)-pE(p), or equivalently,

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

For prime powers plp^{l} with l2l\geq 2, a different flavor of geometry, based on Hensel lifting, comes into play; cf. [hooley1986HasseWeil]*pp. 65–66. For the sake of other work ([wang2023_HLH_vs_RMT]), we prove more than we presently need; we encourage the reader to skip ahead to Corollary 7.5 on a first reading. We start by identifying a clean source of cancellation in (1.5). For each integer d0d\geq 0, let

(7.3) 𝒮(𝒄,pd)\colonequalsλp×{𝒙pm:pdF(𝒙)λ𝒄}.\mathscr{S}(\bm{c},p^{d})\colonequals\bigcup_{\lambda\in\mathbb{Z}_{p}^{\times}}\{\bm{x}\in\mathbb{Z}_{p}^{m}:p^{d}\mid\nabla{F}(\bm{x})-\lambda\bm{c}\}.
Lemma 7.1.

Let 𝐜,𝐱0pm\bm{c},\bm{x}_{0}\in\mathbb{Z}_{p}^{m}. Let l2l\geq 2 and d[0,(l+𝟏p𝐱0)/2]d\in[0,(l+\bm{1}_{p\mid\bm{x}_{0}})/2]. Then dl1d\leq l-1, and

(7.4) 1apl:pa1𝒙pl𝟏𝒙𝒙0modpld𝟏𝒙𝒮(𝒄,pd)epl(aF(𝒙)+𝒄𝒙)=0.\sum_{1\leq a\leq p^{l}:\,p\nmid a}\,\sum_{1\leq\bm{x}\leq p^{l}}\bm{1}_{\bm{x}\equiv\bm{x}_{0}\bmod{p^{l-d}}}\cdot\bm{1}_{\bm{x}\notin\mathscr{S}(\bm{c},p^{d})}\cdot e_{p^{l}}(aF(\bm{x})+\bm{c}\cdot\bm{x})=0.
Proof.

If l=2l=2, then d12l/3d\leq 1\leq 2l/3; if l3l\geq 3, then d(l+1)/22l/3d\leq(l+1)/2\leq 2l/3. Thus d2l/3l1d\leq\lfloor 2l/3\rfloor\leq l-1.

On the left-hand side of (7.4), write 𝒙=𝒙0+pld𝒓\bm{x}=\bm{x}_{0}+p^{l-d}\bm{r} (with 𝒓pm\bm{r}\in\mathbb{Z}_{p}^{m} running over a complete set of residues modulo pdp^{d}). Trivially, 𝒄𝒙𝒄𝒙0+𝒄pld𝒓modpl\bm{c}\cdot\bm{x}\equiv\bm{c}\cdot\bm{x}_{0}+\bm{c}\cdot p^{l-d}\bm{r}\bmod{p^{l}}. Also, since FF is homogeneous of degree 33, Taylor expansion (using min(𝟏p𝒙0+2(ld),3(ld))l\min(\bm{1}_{p\mid\bm{x}_{0}}+2(l-d),3(l-d))\geq l) gives

F(𝒙)F(𝒙0)+F(𝒙0)pld𝒓modpl.F(\bm{x})\equiv F(\bm{x}_{0})+\nabla{F}(\bm{x}_{0})\cdot p^{l-d}\bm{r}\bmod{p^{l}}.

Furthermore, F(𝒙)F(𝒙0)modpd\nabla{F}(\bm{x})\equiv\nabla{F}(\bm{x}_{0})\bmod{p^{d}} (since F\nabla{F} is “homogeneous of degree 22”, and min(𝟏p𝒙0+(ld),2(ld))d\min(\bm{1}_{p\mid\bm{x}_{0}}+(l-d),2(l-d))\geq d), whence 𝟏𝒙𝒮(𝒄,pd)=𝟏𝒙0𝒮(𝒄,pd)\bm{1}_{\bm{x}\notin\mathscr{S}(\bm{c},p^{d})}=\bm{1}_{\bm{x}_{0}\notin\mathscr{S}(\bm{c},p^{d})} (by (7.3)).

If 𝒙0𝒮(𝒄,pd)\bm{x}_{0}\in\mathscr{S}(\bm{c},p^{d}), then the left-hand side of (7.4) directly vanishes. Now suppose 𝒙0𝒮(𝒄,pd)\bm{x}_{0}\notin\mathscr{S}(\bm{c},p^{d}). Then for each aa in (7.4), we have pdaF(𝒙0)+𝒄p^{d}\nmid a\nabla{F}(\bm{x}_{0})+\bm{c}, whence the sum of epl(aF(𝒙)+𝒄𝒙)e_{p^{l}}(aF(\bm{x})+\bm{c}\cdot\bm{x}) over 𝒙𝒙0modpld\bm{x}\equiv\bm{x}_{0}\bmod{p^{l-d}} (i.e. over 𝒓modpd\bm{r}\bmod{p^{d}}) vanishes. Summing over aa gives (7.4). ∎

We now show that to understand S𝒄(pl)S_{\bm{c}}(p^{l}) for l2l\geq 2, it suffices to understand

(7.5) S𝒄(pl)\colonequals1apl:pa1𝒙pl:p𝒙epl(aF(𝒙)+𝒄𝒙).S^{\prime}_{\bm{c}}(p^{l})\colonequals\sum_{1\leq a\leq p^{l}:\,p\nmid a}\,\sum_{1\leq\bm{x}\leq p^{l}:\,p\nmid\bm{x}}e_{p^{l}}(aF(\bm{x})+\bm{c}\cdot\bm{x}).

Before proceeding, note that by (1.5), (7.5), we have

(7.6) S𝒄(pl)S𝒄(pl)=1apl:pa1𝒙pl:p𝒙epl(aF(𝒙)+𝒄𝒙).S_{\bm{c}}(p^{l})-S^{\prime}_{\bm{c}}(p^{l})=\sum_{1\leq a\leq p^{l}:\,p\nmid a}\,\sum_{1\leq\bm{x}\leq p^{l}:\,p\mid\bm{x}}e_{p^{l}}(aF(\bm{x})+\bm{c}\cdot\bm{x}).
Lemma 7.2.

Fix a tuple 𝐜pm\bm{c}\in\mathbb{Z}_{p}^{m} and an integer l2l\geq 2. Then S𝐜(pl)S𝐜(pl)S_{\bm{c}}(p^{l})-S^{\prime}_{\bm{c}}(p^{l}) equals

  1. (1)

    𝟏p𝒄ϕ(p2)pm\bm{1}_{p\mid\bm{c}}\cdot\phi(p^{2})p^{m} if l=2l=2, and

  2. (2)

    𝟏p2𝒄[ϕ(pl)/ϕ(pl3)]p2mS𝒄/p2(pl3)\bm{1}_{p^{2}\mid\bm{c}}\cdot[\phi(p^{l})/\phi(p^{l-3})]\cdot p^{2m}\cdot S_{\bm{c}/p^{2}}(p^{l-3}) if l3l\geq 3.

In particular, if p𝐜p\nmid\bm{c}, then S𝐜(pl)=S𝐜(pl)S_{\bm{c}}(p^{l})=S^{\prime}_{\bm{c}}(p^{l}).

Proof.

Let d\colonequalsmin(2,(l+1)/2)d\colonequals\min(2,\lfloor(l+1)/2\rfloor); then Lemma 7.1 applies whenever p𝒙0p\mid\bm{x}_{0}. Summing (7.4) over {1𝒙0pld:p𝒙0}\{1\leq\bm{x}_{0}\leq p^{l-d}:p\mid\bm{x}_{0}\}, we get (by (7.6))

S𝒄(pl)S𝒄(pl)=1apl:pa1𝒙pl:p𝒙𝟏𝒙𝒮(𝒄,pd)epl(aF(𝒙)+𝒄𝒙).S_{\bm{c}}(p^{l})-S^{\prime}_{\bm{c}}(p^{l})=\sum_{1\leq a\leq p^{l}:\,p\nmid a}\,\sum_{1\leq\bm{x}\leq p^{l}:\,p\mid\bm{x}}\bm{1}_{\bm{x}\in\mathscr{S}(\bm{c},p^{d})}\cdot e_{p^{l}}(aF(\bm{x})+\bm{c}\cdot\bm{x}).

Here p2F(𝒙)p^{2}\mid\nabla{F}(\bm{x}), and d2d\leq 2, so 𝟏𝒙𝒮(𝒄,pd)=𝟏pd𝒄\bm{1}_{\bm{x}\in\mathscr{S}(\bm{c},p^{d})}=\bm{1}_{p^{d}\mid\bm{c}} (by (7.3)). So if pd𝒄p^{d}\nmid\bm{c}, then S𝒄(pl)S𝒄(pl)=0S_{\bm{c}}(p^{l})-S^{\prime}_{\bm{c}}(p^{l})=0, which suffices (since d=1d=1 if l=2l=2, and d=2d=2 if l3l\geq 3). Now suppose pd𝒄p^{d}\mid\bm{c}.

If l=2l=2, then plF(𝒙),𝒄𝒙p^{l}\mid F(\bm{x}),\bm{c}\cdot\bm{x}, so S𝒄(pl)S𝒄(pl)=ϕ(pl)1𝒙pl:p𝒙1=ϕ(p2)pmS_{\bm{c}}(p^{l})-S^{\prime}_{\bm{c}}(p^{l})=\phi(p^{l})\sum_{1\leq\bm{x}\leq p^{l}:\,p\mid\bm{x}}1=\phi(p^{2})p^{m}.

Now suppose l3l\geq 3. Then d=2d=2, and 𝒄\colonequals𝒄/p2pm\bm{c}^{\prime}\colonequals\bm{c}/p^{2}\in\mathbb{Z}_{p}^{m}. Now write 𝒙=p𝒙\bm{x}=p\bm{x}^{\prime} to get

S𝒄(pl)S𝒄(pl)=1apl:pa1𝒙pl1epl(ap3F(𝒙)+p3𝒄𝒙).S_{\bm{c}}(p^{l})-S^{\prime}_{\bm{c}}(p^{l})=\sum_{1\leq a\leq p^{l}:\,p\nmid a}\,\sum_{1\leq\bm{x}^{\prime}\leq p^{l-1}}e_{p^{l}}(ap^{3}F(\bm{x}^{\prime})+p^{3}\bm{c}^{\prime}\cdot\bm{x}^{\prime}).

Now let l\colonequalsl30l^{\prime}\colonequals l-3\geq 0. Then epl()e_{p^{l}}(-) is determined by (a,𝒙)modpl(a,\bm{x}^{\prime})\bmod{p^{l^{\prime}}}. Therefore

S𝒄(pl)S𝒄(pl)=[ϕ(pl)/ϕ(pl)]p2m1apl:pa1𝒙plepl(aF(𝒙)+𝒄𝒙),S_{\bm{c}}(p^{l})-S^{\prime}_{\bm{c}}(p^{l})=[\phi(p^{l})/\phi(p^{l^{\prime}})]\cdot p^{2m}\cdot\sum_{1\leq a\leq p^{l^{\prime}}:\,p\nmid a}\,\sum_{1\leq\bm{x}^{\prime}\leq p^{l^{\prime}}}e_{p^{l^{\prime}}}(aF(\bm{x}^{\prime})+\bm{c}^{\prime}\cdot\bm{x}^{\prime}),

which equals [ϕ(pl)/ϕ(pl)]p2mS𝒄(pl)[\phi(p^{l})/\phi(p^{l^{\prime}})]\cdot p^{2m}\cdot S_{\bm{c}^{\prime}}(p^{l^{\prime}}) by (1.5). This completes the proof. ∎

The general study of S𝒄(pl)S^{\prime}_{\bm{c}}(p^{l}) needs some setup. For any vector 𝒃pm\bm{b}\in\mathbb{Z}_{p}^{m}, let vp(𝒃)\colonequalsvp(gcd(b1,,bm))[0,]v_{p}(\bm{b})\colonequals v_{p}(\gcd(b_{1},\dots,b_{m}))\in[0,\infty]. Given 𝒄pm{𝟎}\bm{c}\in\mathbb{Z}_{p}^{m}\setminus\{\bm{0}\} and integers r,s,d0r,s,d\geq 0, let

(7.7) 𝒮r,s(𝒄,d)\colonequals{𝒙pm𝒮(𝒄,pd):p𝒙,prF(𝒙),ps+vp(𝒄)𝒄𝒙},𝒮r,s(𝒄,d)\colonequals{𝒙𝒮r,s(𝒄,d):vp(F(𝒙))=vp(𝒄)}.\begin{split}\mathcal{S}_{r,s}(\bm{c},d)&\colonequals\{\bm{x}\in\mathbb{Z}_{p}^{m}\cap\mathscr{S}(\bm{c},p^{d}):p\nmid\bm{x},\;p^{r}\mid F(\bm{x}),\;p^{s+v_{p}(\bm{c})}\mid\bm{c}\cdot\bm{x}\},\\ \mathcal{S}^{\ast}_{r,s}(\bm{c},d)&\colonequals\{\bm{x}\in\mathcal{S}_{r,s}(\bm{c},d):v_{p}(\nabla{F}(\bm{x}))=v_{p}(\bm{c})\}.\end{split}

Let μp\mu_{p} denote the usual Haar measure on pm\mathbb{Z}_{p}^{m}, so that for all lmax(1,r,s,d)l\geq\max(1,r,s,d), we have

(7.8) μp(𝒮r,s(𝒄,d))=plm|{𝒙𝒮(𝒄,pd):1𝒙pl,p𝒙,prF(𝒙),ps+vp(𝒄)𝒄𝒙}|.\mu_{p}(\mathcal{S}_{r,s}(\bm{c},d))=p^{-lm}\cdot\lvert\{\bm{x}\in\mathscr{S}(\bm{c},p^{d}):1\leq\bm{x}\leq p^{l},\;p\nmid\bm{x},\;p^{r}\mid F(\bm{x}),\;p^{s+v_{p}(\bm{c})}\mid\bm{c}\cdot\bm{x}\}\rvert.
Lemma 7.3.

Suppose 𝐜pm{𝟎}\bm{c}\in\mathbb{Z}_{p}^{m}\setminus\{\bm{0}\} and g=vp(𝐜)g=v_{p}(\bm{c}). Let u,v1u,v\geq 1 and dgd\geq g be integers. Then

(7.9) μp(𝒮u,v(𝒄,d))\displaystyle\mu_{p}(\mathcal{S}_{u,v}(\bm{c},d)) =puvμp(𝒮u,u(𝒄,d)), if vd and dug+v;\displaystyle=p^{u-v}\cdot\mu_{p}(\mathcal{S}_{u,u}(\bm{c},d)),\textnormal{ if }v\geq d\textnormal{ and }d\leq u\leq g+v;
(7.10) μp(𝒮u,v(𝒄,d))\displaystyle\mu_{p}(\mathcal{S}^{\ast}_{u,v}(\bm{c},d)) =p1μp(𝒮u1,v(𝒄,d)), if u1gmax(1+g,d,v).\displaystyle=p^{-1}\cdot\mu_{p}(\mathcal{S}^{\ast}_{u-1,v}(\bm{c},d)),\textnormal{ if }u-1-g\geq\max(1+g,d,v).
Proof.

Let r,s1r,s\geq 1 be integers. Using (7.7) and the congruence

F(𝒙+pmax(d,rg)𝒉)F(𝒙)+F(𝒙)pmax(d,rg)𝒉modprF(\bm{x}+p^{\max(d,r-g)}\bm{h})\equiv F(\bm{x})+\nabla{F}(\bm{x})\cdot p^{\max(d,r-g)}\bm{h}\bmod{p^{r}}

(valid since 2max(d,rg)2max(g,rg)r2\max(d,r-g)\geq 2\max(g,r-g)\geq r), we find that

  1. (1)

    𝒮r,s(𝒄,d)\mathcal{S}_{r,s}(\bm{c},d) is invariant under addition by any element of pmax(d,rg,s)pmp^{\max(d,r-g,s)}\mathbb{Z}_{p}^{m} (since pgF(𝒙)p^{g}\mid\nabla{F}(\bm{x}) for all 𝒙𝒮r,s(𝒄,d)\bm{x}\in\mathcal{S}_{r,s}(\bm{c},d), by (7.3)), and therefore

  2. (2)

    𝒮r,s(𝒄,d)\mathcal{S}^{\ast}_{r,s}(\bm{c},d) is invariant under pmax(1+g,d,rg,s)pmp^{\max(1+g,d,r-g,s)}\mathbb{Z}_{p}^{m}.

Case 1: vdv\geq d and dug+vd\leq u\leq g+v. Then max(d,ug,v)=v\max(d,u-g,v)=v and max(d,ug,u)=u\max(d,u-g,u)=u.

  • If uvu\leq v, then the inclusion 𝒮u,v(𝒄,d)𝒮u,u(𝒄,d)\mathcal{S}_{u,v}(\bm{c},d)\to\mathcal{S}_{u,u}(\bm{c},d) descends to a map 𝒮u,v(𝒄,d)/pvpm𝒮u,u(𝒄,d)/pupm\mathcal{S}_{u,v}(\bm{c},d)/p^{v}\mathbb{Z}_{p}^{m}\to\mathcal{S}_{u,u}(\bm{c},d)/p^{u}\mathbb{Z}_{p}^{m}; and this map has fibers of size p(m1)(vu)p^{(m-1)(v-u)}, so (7.9) holds.

  • If uvu\geq v, then the inclusion 𝒮u,u(𝒄,d)𝒮u,v(𝒄,d)\mathcal{S}_{u,u}(\bm{c},d)\to\mathcal{S}_{u,v}(\bm{c},d) descends to a map 𝒮u,u(𝒄,d)/pupm𝒮u,v(𝒄,d)/pvpm\mathcal{S}_{u,u}(\bm{c},d)/p^{u}\mathbb{Z}_{p}^{m}\to\mathcal{S}_{u,v}(\bm{c},d)/p^{v}\mathbb{Z}_{p}^{m}; and this map has fibers of size p(m1)(uv)p^{(m-1)(u-v)}, so (7.9) holds.

Case 2: u1gmax(1+g,d,v)u-1-g\geq\max(1+g,d,v). Then u2u\geq 2, and max(1+g,d,rg,v)=rg\max(1+g,d,r-g,v)=r-g for all ru1r\geq u-1. So the inclusion 𝒮u,v(𝒄,d)𝒮u1,v(𝒄,d)\mathcal{S}^{\ast}_{u,v}(\bm{c},d)\to\mathcal{S}^{\ast}_{u-1,v}(\bm{c},d) descends to a map

𝒮u,v(𝒄,d)/pugpm𝒮u1,v(𝒄,d)/pug1pm.\mathcal{S}^{\ast}_{u,v}(\bm{c},d)/p^{u-g}\mathbb{Z}_{p}^{m}\to\mathcal{S}^{\ast}_{u-1,v}(\bm{c},d)/p^{u-g-1}\mathbb{Z}_{p}^{m}.

This map has fibers of size pm1p^{m-1}, since for all 𝒙0𝒮u1,v(𝒄,d)\bm{x}_{0}\in\mathcal{S}^{\ast}_{u-1,v}(\bm{c},d) and 𝒓pm\bm{r}\in\mathbb{Z}_{p}^{m}, we have vp(F(𝒙0))=gv_{p}(\nabla{F}(\bm{x}_{0}))=g and the “lifting congruence”

F(𝒙0+pug1𝒓)F(𝒙0)+F(𝒙0)pug1𝒓modpu.F(\bm{x}_{0}+p^{u-g-1}\bm{r})\equiv F(\bm{x}_{0})+\nabla{F}(\bm{x}_{0})\cdot p^{u-g-1}\bm{r}\bmod{p^{u}}.

(This congruence holds because 2(ug1)u2(u-g-1)\geq u.) Thus (7.10) holds. ∎

The next result synthesizes a lot of old and new Hensel work.

Proposition 7.4.

Let 𝐜pm{𝟎}\bm{c}\in\mathbb{Z}_{p}^{m}\setminus\{\bm{0}\}, and let g=vp(𝐜)g=v_{p}(\bm{c}). Let dgd\geq g if VV is smooth, and let d1+gd\geq 1+g if VV is singular. Let lmax(2+2g,2d)l\geq\max(2+2g,2d). Then

(7.11) plmϕ(pl)S𝒄(pl)=p2l+gμp(𝒮l,l(𝒄,d))p2l2+gμp(𝒮l1,l1(𝒄,d)).p^{-lm}\phi(p^{l})S^{\prime}_{\bm{c}}(p^{l})=p^{2l+g}\mu_{p}(\mathcal{S}_{l,l}(\bm{c},d))-p^{2l-2+g}\mu_{p}(\mathcal{S}_{l-1,l-1}(\bm{c},d)).
Proof.

Lemma 7.1 applies, since dl/2d\leq l/2. Summing (7.4) over {1𝒙0pld:p𝒙0}\{1\leq\bm{x}_{0}\leq p^{l-d}:p\nmid\bm{x}_{0}\} gives

S𝒄(pl)=1apl:pa1𝒙pl:p𝒙𝟏𝒙𝒮(𝒄,pd)epl(aF(𝒙)+𝒄𝒙).S^{\prime}_{\bm{c}}(p^{l})=\sum_{1\leq a\leq p^{l}:\,p\nmid a}\,\sum_{1\leq\bm{x}\leq p^{l}:\,p\nmid\bm{x}}\bm{1}_{\bm{x}\in\mathscr{S}(\bm{c},p^{d})}\cdot e_{p^{l}}(aF(\bm{x})+\bm{c}\cdot\bm{x}).

Replacing 𝒄\bm{c} with λ𝒄\lambda\bm{c} for λp×\lambda\in\mathbb{Z}_{p}^{\times}, and summing over λ\lambda, we get (via the scalar symmetries S𝒄(pl)=Sλ𝒄(pl)S^{\prime}_{\bm{c}}(p^{l})=S^{\prime}_{\lambda\bm{c}}(p^{l}) and 𝒮(𝒄,pd)=𝒮(λ𝒄,pd)\mathscr{S}(\bm{c},p^{d})=\mathscr{S}(\lambda\bm{c},p^{d}) that follow from (7.5) and (7.3), respectively)

(7.12) plmϕ(pl)S𝒄(pl)=1λpl:pλplmSλ𝒄(pl)=l1u,vl(p)u+vμp(𝒮u,vg(𝒄,d)),p^{-lm}\phi(p^{l})S^{\prime}_{\bm{c}}(p^{l})=\sum_{1\leq\lambda\leq p^{l}:\,p\nmid\lambda}p^{-lm}S^{\prime}_{\lambda\bm{c}}(p^{l})=\sum_{l-1\leq u,v\leq l}(-p)^{u+v}\mu_{p}(\mathcal{S}_{u,v-g}(\bm{c},d)),

by a short calculation using 1bpl:pb=1bpl1bpl:pb\sum_{1\leq b\leq p^{l}:\,p\nmid b}=\sum_{1\leq b\leq p^{l}}-\sum_{1\leq b\leq p^{l}:\,p\mid b} (for b=a,λb=a,\lambda) and (7.8).

Let r,s1r,s\geq 1. Before proceeding, we prove (by casework) that

(7.13) 𝒮r,s(𝒄,d)=𝒮r,s(𝒄,d).\mathcal{S}^{\ast}_{r,s}(\bm{c},d)=\mathcal{S}_{r,s}(\bm{c},d).

Case 1: d=g1d=g\geq 1. Then VV is smooth, so {𝒙pm:p𝒙,pF(𝒙),pF(𝒙)}=\{\bm{x}\in\mathbb{Z}_{p}^{m}:p\nmid\bm{x},\;p\mid F(\bm{x}),\;p\mid\nabla{F}(\bm{x})\}=\emptyset. Since p𝒄p\mid\bm{c}, we conclude by (7.3), (7.7) that 𝒮1,1(𝒄,1)=\mathcal{S}_{1,1}(\bm{c},1)=\emptyset. So both sides of (7.13) are empty.

Case 2: d=g=0d=g=0. Then VV is smooth, and p𝒄p\nmid\bm{c}. So by (7.7), we have (7.13).

Case 3: dg+1d\geq g+1. Then vp(F(𝒙))=gv_{p}(\nabla{F}(\bm{x}))=g for all 𝒙𝒮(𝒄,pd)\bm{x}\in\mathscr{S}(\bm{c},p^{d}). So by (7.7), we have (7.13).

Having established (7.13) in all cases, we now return to (7.12). Since dgd\geq g and lmax(2+2g,1+g+d,l)l\geq\max(2+2g,1+g+d,l), we may apply (7.10) (with u=lu=l and v=l1gv=l-1-g) and (7.13) to get

μp(𝒮l,l1g(𝒄,d))=p1μp(𝒮l1,l1g(𝒄,d)).\mu_{p}(\mathcal{S}_{l,l-1-g}(\bm{c},d))=p^{-1}\cdot\mu_{p}(\mathcal{S}_{l-1,l-1-g}(\bm{c},d)).

But since dgd\geq g and l1gmax(1,d)l-1-g\geq\max(1,d), we may use (7.9) to get

μp(𝒮u,v(𝒄,d))=puvμp(𝒮u,u(𝒄,d))\mu_{p}(\mathcal{S}_{u,v}(\bm{c},d))=p^{u-v}\cdot\mu_{p}(\mathcal{S}_{u,u}(\bm{c},d))

for all (u,v){l1,l}×{l1g,lg}(u,v)\in\{l-1,l\}\times\{l-1-g,l-g\} such that ug+vu\leq g+v. Thus the right-hand side of (7.12) equals pl+l+gμp(𝒮l,l(𝒄,d))(pl+(l1)1+g+p(l1)+l+(g1)p(l1)+(l1)+g)μp(𝒮l1,l1(𝒄,d))p^{l+l+g}\cdot\mu_{p}(\mathcal{S}_{l,l}(\bm{c},d))-(p^{l+(l-1)-1+g}+p^{(l-1)+l+(g-1)}-p^{(l-1)+(l-1)+g})\cdot\mu_{p}(\mathcal{S}_{l-1,l-1}(\bm{c},d)), which simplifies to the right-hand side of (7.11). So (7.12) implies (7.11). ∎

Corollary 7.5.

Suppose 𝐜pm{𝟎}\bm{c}\in\mathbb{Z}_{p}^{m}\setminus\{\bm{0}\} is primitive and VV is smooth. Let l2l\geq 2. Then

S𝒄(pl)=p2l|𝒱𝒄(/pl)|p2l+m|𝒱𝒄(/pl1)|.S_{\bm{c}}(p^{l})=p^{2l}\lvert\mathcal{V}_{\bm{c}}(\mathbb{Z}/p^{l}\mathbb{Z})\rvert-p^{2l+m_{\ast}}\lvert\mathcal{V}_{\bm{c}}(\mathbb{Z}/p^{l-1}\mathbb{Z})\rvert.
Proof.

Here vp(𝒄)=0v_{p}(\bm{c})=0, so (7.8) implies μp(𝒮u,u(𝒄,0))=pumϕ(pu)|𝒱𝒄(/pu)|\mu_{p}(\mathcal{S}_{u,u}(\bm{c},0))=p^{-um}\phi(p^{u})\lvert\mathcal{V}_{\bm{c}}(\mathbb{Z}/p^{u}\mathbb{Z})\rvert for u1u\geq 1. Plug this into Proposition 7.4 (with d=g=0d=g=0); then note that S𝒄(pl)=S𝒄(pl)S_{\bm{c}}(p^{l})=S^{\prime}_{\bm{c}}(p^{l}) by Lemma 7.2. ∎

Now fix LΥL\in\Upsilon. We will build up to Lemma 7.11 (realizing (1.10) from §1). As Remark 1.6 suggests, Lemma 7.11 might extend to more general FF. But to maximize the accessibility of §7, we focus on the diagonal case. We will use an ad hoc change of coordinates, highlighting specific features (of diagonal forms) that may be of independent interest.

So for the rest of §7, assume FF is diagonal with m{4,6}m\in\{4,6\}. Let 𝒥\mathcal{J} be a permissible pairing corresponding to LL in Proposition 3.6. Since 𝒥\mathcal{J} is permissible, there exist unique cube-free integers F(k)F_{(k)} (for k𝒦k\in\mathcal{K}) such that Fi/F(k)F_{i}/F_{(k)} is an integer cube for all k𝒦k\in\mathcal{K} and i𝒥(k)i\in\mathcal{J}(k).

Suppose 𝒄\bm{c} lies in Λ\Lambda^{\perp} or more generally, Λp\Lambda^{\perp}\otimes\mathbb{Z}_{p}. Assume

(7.14) pj2m/21F(𝒄).p\nmid j^{2^{m/2-1}}{F^{\vee}}(\bm{c}).

In particular, by (1.14), we have

(7.15) p(6m)!F1Fm.p\nmid(6^{m})!F_{1}\cdots F_{m}.
Proposition 7.6.

Under (7.14) and (7.15), the following hold:

  1. (1)

    Each c(k)3c(k)^{3} lies in p\mathbb{Z}_{p}.

  2. (2)

    If k1<<ktk_{1}<\dots<k_{t}, then p(c(k1)3/2±±c(kt)3/2)p\nmid\prod(c(k_{1})^{3/2}\pm\cdots\pm c(k_{t})^{3/2}). In particular, pc(k)3p\nmid c(k)^{3}, hence c(k)3p×c(k)^{3}\in\mathbb{Z}_{p}^{\times}, for each kk. Also, pc(i)3c(j)3p\nmid c(i)^{3}-c(j)^{3} when iji\neq j.

  3. (3)

    V𝒄V_{\bm{c}} has exactly 2m/212^{m/2-1} singular 𝔽¯p\overline{\mathbb{F}}_{p}-points.

Proof.

(1): By Definition 3.5, c(k)3=ci3/Fic(k)^{3}=c_{i}^{3}/F_{i} for all i𝒥(k)i\in\mathcal{J}(k). Here pFip\nmid F_{i} by (7.15).

Next, we use some results of §3.2, carried over from \mathbb{Q} to 𝔽p\mathbb{F}_{p} via Remark 3.11.

(2): Use (7.14) and the 𝔽p\mathbb{F}_{p}-analog of Observation 3.10(1).

(3): pj2m/211F(𝒄)p\mid j^{2^{m/2-1}-1}{F^{\vee}}(\bm{c}) by Corollary 3.9, carried over to 𝔽p\mathbb{F}_{p}. But pj2m/21F(𝒄)p\nmid j^{2^{m/2-1}}{F^{\vee}}(\bm{c}) by (7.14). So by the 𝔽p\mathbb{F}_{p}-analogs of Proposition 3.7 and (3.5), the scheme V𝒄V_{\bm{c}} has at least, but also at most, 2m/212^{m/2-1} singular 𝔽¯p\overline{\mathbb{F}}_{p}-points. (Note that each singular 𝔽¯p\overline{\mathbb{F}}_{p}-point of V𝒄V_{\bm{c}} has all coordinates nonzero, by the Jacobian criterion, since pc1cmp\nmid c_{1}\cdots c_{m} by (2).) ∎

Assume, until further notice, that

(7.16) Fi=F(k)F_{i}=F_{(k)}

for all k𝒦k\in\mathcal{K} and i𝒥(k)i\in\mathcal{J}(k). For convenience, assume 𝒥(k)={k,k+m/2}\mathcal{J}(k)=\{k,k+m/2\} for each k𝒦=[m/2]k\in\mathcal{K}=[m/2]. Then 𝒄Λp\bm{c}\in\Lambda^{\perp}\otimes\mathbb{Z}_{p} implies ck=ck+m/2c_{k}=c_{k+m/2}. Let ck\colonequalsck=ck+m/2c^{\star}_{k}\colonequals c_{k}=c_{k+m/2}, so that

(7.17) c(k)3=(ck)3/F(k).c(k)^{3}=(c^{\star}_{k})^{3}/F_{(k)}.

Now consider the equations F(𝒙)=0F(\bm{x})=0 and 𝒄𝒙=0\bm{c}\cdot\bm{x}=0 defining 𝒱𝒄\mathcal{V}_{\bm{c}}. These equations become

(7.18) 1km/2F(k)h[k]y[k]2=31km/2F(k)h[k]3and1km/2ckh[k]=0\sum_{1\leq k\leq m/2}F_{(k)}\cdot h[k]y[k]^{2}=-3\sum_{1\leq k\leq m/2}F_{(k)}\cdot h[k]^{3}\qquad\textnormal{and}\qquad\sum_{1\leq k\leq m/2}c^{\star}_{k}\cdot h[k]=0

after a linear change of variables over [1/6]\mathbb{Z}[1/6]. Explicitly, if 𝒥(k)={i,j}\mathcal{J}(k)=\{i,j\} with i<ji<j, then we take h[k]\colonequalsxi+xjh[k]\colonequals x_{i}+x_{j} and y[k]\colonequals3(xixj)y[k]\colonequals 3(x_{i}-x_{j}), so that the equation h[1]==h[m/2]=0h[1]=\cdots=h[m/2]=0 cuts out Λp\Lambda\otimes\mathbb{Z}_{p}. (We use the letter “hh” in analogy with van der Corput or Weyl differencing. The definition of 𝒉\colonequals(h[k])1km/2\bm{h}\colonequals(h[k])_{1\leq k\leq m/2} is compatible with that in §4.)

Geometrically, over K\colonequals𝔽pK\colonequals\mathbb{F}_{p}, the space {[𝒉]m/21:𝒄𝒉=0}m/22\{[\bm{h}]\in\mathbb{P}^{m/2-1}:\bm{c}^{\star}\cdot\bm{h}=0\}\cong\mathbb{P}^{m/2-2} parameterizes projective m2\frac{m}{2}-planes H𝒄m2\mathbb{P}{H}\subseteq\mathbb{P}\bm{c}^{\perp}\cong\mathbb{P}^{m-2} containing the fixed (m21)(\frac{m}{2}-1)-plane ΛK\mathbb{P}\Lambda_{K}. Over this [𝒉][\bm{h}]-space, we have the “quadratic fibration” (related to the blow-up of V𝒄V_{\bm{c}} along ΛK\mathbb{P}\Lambda_{K})

V𝒄ΛK[𝒄]m/22,[𝒙][𝒉].V_{\bm{c}}\setminus\mathbb{P}\Lambda_{K}\to\mathbb{P}[\bm{c}^{\star}]^{\perp}\cong\mathbb{P}^{m/2-2},\quad[\bm{x}]\mapsto[\bm{h}].

Concretely, each slice V𝒄HV_{\bm{c}}\cap\mathbb{P}{H} consists of ΛK\mathbb{P}\Lambda_{K} and a (possibly singular) quadric hypersurface QHHQ_{H}\subseteq\mathbb{P}{H} of dimension m/21m/2-1, where QHΛKQ_{H}\setminus\mathbb{P}\Lambda_{K} is the fiber of V𝒄ΛKV_{\bm{c}}\setminus\mathbb{P}\Lambda_{K} over H\mathbb{P}{H}.

Below, let (rp)(\frac{r}{p}) denote the Legendre symbol (if pp is odd), and write χ(r)\colonequals(rp)\chi(r)\colonequals(\frac{r}{p}).

Lemma 7.7.

Under (7.14), (7.15), and (7.16), we have E𝐜(p)=p1/2+O(1)E^{\natural}_{\bm{c}}(p)=p^{1/2}+O(1).

Proof.

Let C(V𝒄)C(V_{\bm{c}}) denote the affine cone of V𝒄V_{\bm{c}}: the subscheme F(𝒙)=𝒄𝒙=0F(\bm{x})=\bm{c}\cdot\bm{x}=0 of 𝔸m\mathbb{A}^{m}. Then |C(V𝒄)(𝔽p)|=1+(p1)|V𝒄(𝔽p)|\lvert C(V_{\bm{c}})(\mathbb{F}_{p})\rvert=1+(p-1)\lvert V_{\bm{c}}(\mathbb{F}_{p})\rvert. We must show that |C(V𝒄)(𝔽p)|=pm2+pm/2+O(p(m1)/2)\lvert C(V_{\bm{c}})(\mathbb{F}_{p})\rvert=p^{m-2}+p^{m/2}+O(p^{(m-1)/2}).

We count solutions to F(𝒙)=𝒄𝒙=0F(\bm{x})=\bm{c}\cdot\bm{x}=0 using the (𝒉,𝒚)(\bm{h},\bm{y}) coordinates in (7.18). The locus 𝒉=𝟎\bm{h}=\bm{0} contributes |Λ/pΛ|=pm/2\lvert\Lambda/p\Lambda\rvert=p^{m/2} solutions to (7.18). Let

U=V𝒄{𝒉𝟎},U=V𝒄{h[k]0},Zk=U{h[k]=0},{\textstyle U=V_{\bm{c}}\cap\{\bm{h}\neq\bm{0}\},\quad U^{\prime}=V_{\bm{c}}\cap\{\prod h[k]\neq 0\},\quad Z_{k}=U\cap\{h[k]=0\},}

and by slight abuse of notation, define the corresponding cones with origins removed:

C(U)=C(V𝒄){𝒉𝟎},C(U)=C(V𝒄){h[k]0},C(Zk)=C(U){h[k]=0}.{\textstyle C(U)=C(V_{\bm{c}})\cap\{\bm{h}\neq\bm{0}\},\quad C(U^{\prime})=C(V_{\bm{c}})\cap\{\prod h[k]\neq 0\},\quad C(Z_{k})=C(U)\cap\{h[k]=0\}.}

Recall (7.17). Since pckp\nmid c^{\star}_{k} for all k[m/2]k\in[m/2] (by Proposition 7.6(2)), the equation 𝒄𝒉=0\bm{c}^{\star}\cdot\bm{h}=0 implies that U=UU=U^{\prime} if m=4m=4, and that Z1Z_{1}, Z2Z_{2}, Z3Z_{3} are pairwise disjoint if m=6m=6.

Suppose first that m=4m=4. Then U=UU=U^{\prime} is covered by a single affine chart, say with h[2]=c1h[2]=c^{\star}_{1} and h[1]=c2h[1]=-c^{\star}_{2}. The (remaining) defining equation F(𝒙)=0F(\bm{x})=0 becomes

F(1)c2y[1]2+F(2)c1y[2]2=3[F(1)(c2)3F(2)(c1)3].-F_{(1)}\cdot c^{\star}_{2}\cdot y[1]^{2}+F_{(2)}\cdot c^{\star}_{1}\cdot y[2]^{2}=3[F_{(1)}\cdot(c^{\star}_{2})^{3}-F_{(2)}\cdot(c^{\star}_{1})^{3}].

Since pc(2)3c(1)3p\nmid c(2)^{3}-c(1)^{3} (by Proposition 7.6(2)), we get |U(𝔽p)|=pχ(c(1)3c(2)3)\lvert U(\mathbb{F}_{p})\rvert=p-\chi(c(1)^{3}c(2)^{3}) by comparing UU with 1\mathbb{P}^{1}. Thus |C(U)(𝔽p)|=(p1)|U(𝔽p)|=p2+O(p)\lvert C(U)(\mathbb{F}_{p})\rvert=(p-1)\lvert U(\mathbb{F}_{p})\rvert=p^{2}+O(p). So |C(V𝒄)(𝔽p)|=|Λ/pΛ|+|C(U)(𝔽p)|=pm2+pm/2+O(p)\lvert C(V_{\bm{c}})(\mathbb{F}_{p})\rvert=\lvert\Lambda/p\Lambda\rvert+\lvert C(U)(\mathbb{F}_{p})\rvert=p^{m-2}+p^{m/2}+O(p), which is better than satisfactory.

Suppose next that m=6m=6. Plugging h[k]=0h[k]=0 into (7.18) identifies C(Zk)C(Z_{k}) as the product of 𝔸1\mathbb{A}^{1} (with coordinate y[k]y[k]) with a cone of the shape “C(U)C(U) for m=4m=4” (which has p2+O(p)p^{2}+O(p) points by the previous paragraph). Thus |C(Zk)(𝔽p)|=p3+O(p2)\lvert C(Z_{k})(\mathbb{F}_{p})\rvert=p^{3}+O(p^{2}) for each kk, and |kC(Zk)(𝔽p)|=k|C(Zk)(𝔽p)|=3p3+O(p2)\lvert\bigcup_{k}C(Z_{k})(\mathbb{F}_{p})\rvert=\sum_{k}\lvert C(Z_{k})(\mathbb{F}_{p})\rvert=3p^{3}+O(p^{2}) in total over 1km/21\leq k\leq m/2.

For U(𝔽p)U^{\prime}(\mathbb{F}_{p}), first consider an individual 𝒉𝔽pm/2\bm{h}\in\mathbb{F}_{p}^{m/2} with h[k]0\prod h[k]\neq 0. The equation F(𝒙)=0F(\bm{x})=0 (a ternary quadratic in 𝒚\bm{y}) has 𝒩𝒉=p2+pχ(F(1)h[1]F(3)h[3]3F(k)h[k]3)\mathscr{N}_{\bm{h}}=p^{2}+p\cdot\chi{\left(F_{(1)}h[1]\cdots F_{(3)}h[3]\cdot 3\sum F_{(k)}h[k]^{3}\right)} solutions 𝒚𝔽pm/2\bm{y}\in\mathbb{F}_{p}^{m/2}. Indeed, if F(k)h[k]3=0\sum F_{(k)}h[k]^{3}=0, then {𝒚𝔸3:F(𝒙)=0}\{\bm{y}\in\mathbb{A}^{3}:F(\bm{x})=0\} is an affine cone over a smooth conic Q1Q\cong\mathbb{P}^{1}; otherwise, {𝒚𝔸3:F(𝒙)=0}\{\bm{y}\in\mathbb{A}^{3}:F(\bm{x})=0\} is a non-degenerate affine quadric (and is thus the complement of a smooth conic in a smooth quadric in 3\mathbb{P}^{3}).

Now identify UU^{\prime} with its affine chart h[3]=1h[3]=1. The equation 𝒄𝒉=0\bm{c}^{\star}\cdot\bm{h}=0 becomes h[2]=(c2)1(c1h[1]+c3)h[2]=-(c^{\star}_{2})^{-1}(c^{\star}_{1}\cdot h[1]+c^{\star}_{3}). Thus |U(𝔽p)|\lvert U^{\prime}(\mathbb{F}_{p})\rvert equals the sum of 𝒩𝒉\mathscr{N}_{\bm{h}} over t\colonequalsh[1]𝔽p×{c3/c1}t\colonequals h[1]\in\mathbb{F}_{p}^{\times}\setminus\{-c^{\star}_{3}/c^{\star}_{1}\} (where we restrict tt so that h[1]h[2]0h[1]\cdot h[2]\neq 0):

(7.19) t{0,c3/c1}𝒩𝒉+t𝔽p𝒩𝒉=2p2+p3+p(#{(z,t)𝔽p2:z2=P𝒄(t)}p),-\sum_{t\in\{0,-c^{\star}_{3}/c^{\star}_{1}\}}\mathscr{N}_{\bm{h}}+\sum_{t\in\mathbb{F}_{p}}\mathscr{N}_{\bm{h}}=-2p^{2}+p^{3}+p\cdot\left(\#\{(z,t)\in\mathbb{F}_{p}^{2}:z^{2}=P_{\bm{c}}(t)\}-p\right),

where P𝒄(t)\colonequals3F(1)F(2)F(3)(c2)1t(c1t+c3)[F(1)t3F(2)(c2)3(c1t+c3)3+F(3)]P_{\bm{c}}(t)\colonequals-3F_{(1)}F_{(2)}F_{(3)}(c^{\star}_{2})^{-1}\cdot t\cdot(c^{\star}_{1}\cdot t+c^{\star}_{3})\cdot[F_{(1)}t^{3}-F_{(2)}(c^{\star}_{2})^{-3}(c^{\star}_{1}\cdot t+c^{\star}_{3})^{3}+F_{(3)}]. The count |C(U)(𝔽p)|\lvert C(U^{\prime})(\mathbb{F}_{p})\rvert is p1p-1 times the right-hand side of (7.19).

Here degtP𝒄=5\deg_{t}P_{\bm{c}}=5, since pc(1)3c(2)3p\nmid c(1)^{3}-c(2)^{3} by Proposition 7.6(2). By a routine computer calculation, the discriminant of the quintic polynomial P𝒄(t)P_{\bm{c}}(t) simplifies—up to a harmless “unit monomial” in 3±13^{\pm 1}, F(k)±1F_{(k)}^{\pm 1}, (ck)±1(c^{\star}_{k})^{\pm 1}—to [c(1)3c(3)3]2[c(2)3c(3)3]2[c(1)3/2±c(2)3/2±c(3)3/2][c(1)^{3}-c(3)^{3}]^{2}\cdot[c(2)^{3}-c(3)^{3}]^{2}\cdot\prod[c(1)^{3/2}\pm c(2)^{3/2}\pm c(3)^{3/2}], which lies in p×\mathbb{Z}_{p}^{\times} (by Proposition 7.6(2)). Thus z2=P𝒄(t)z^{2}=P_{\bm{c}}(t) defines an affine hyperelliptic curve over 𝔽p\mathbb{F}_{p} of genus 22. All in all, we have (by the Weil bound for z2=P𝒄(t)z^{2}=P_{\bm{c}}(t))

|C(V𝒄)(𝔽p)|=|Λ/pΛ|+(3p3+O(p2))2(p3p2)+(p4p3)+(p2p)O(p1/2),\lvert C(V_{\bm{c}})(\mathbb{F}_{p})\rvert=\lvert\Lambda/p\Lambda\rvert+(3p^{3}+O(p^{2}))-2(p^{3}-p^{2})+(p^{4}-p^{3})+(p^{2}-p)\cdot O(p^{1/2}),

which simplifies to |Λ/pΛ|+p4+O(p5/2)=pm2+pm/2+O(p(m1)/2)\lvert\Lambda/p\Lambda\rvert+p^{4}+O(p^{5/2})=p^{m-2}+p^{m/2}+O(p^{(m-1)/2}), as desired. ∎

Remark 7.8.

By Lang–Weil for curves, we only need z2=P𝒄(t)z^{2}=P_{\bm{c}}(t) to be absolutely irreducible over 𝔽p\mathbb{F}_{p}—not necessarily smooth. However, pc(k)3p\nmid c(k)^{3} and pc(i)3c(j)3p\nmid c(i)^{3}-c(j)^{3} remain essential throughout the proof of Lemma 7.7; without them, the bias could increase.

Lemma 7.9.

Let l2l\geq 2 be an integer. Under (7.14), (7.15), and (7.16), we have

(7.20) S𝒄(pl)=𝟏χ(c(1)3)==χ(c(m/2)3)2m/21ϕ(pl)pl/2.S^{\natural}_{\bm{c}}(p^{l})=\bm{1}_{\chi(c(1)^{3})=\cdots=\chi(c(m/2)^{3})}\cdot 2^{m/2-1}\phi(p^{l})p^{-l/2}.
Proof.

By (1.8), the desired formula (7.20) is equivalent to

(7.21) p2lS𝒄(pl)=𝟏χ(c(1)3)==χ(c(m/2)3)2m/21(p1)pl(m2)/21.p^{-2l}S_{\bm{c}}(p^{l})=\bm{1}_{\chi(c(1)^{3})=\cdots=\chi(c(m/2)^{3})}\cdot 2^{m/2-1}(p-1)p^{l(m-2)/2-1}.

By Proposition 7.6 and (7.17), we have pc(k)3=(ck)3/F(k)p\nmid c(k)^{3}=(c^{\star}_{k})^{3}/F_{(k)} for all k[m/2]k\in[m/2]. In particular, p𝒄p\nmid\bm{c}, so Corollary 7.5 applies, since VV is smooth by (7.15). Consider the map

(7.22) 𝒱𝒄(/pl)𝒱𝒄(/pl1).\mathcal{V}_{\bm{c}}(\mathbb{Z}/p^{l}\mathbb{Z})\to\mathcal{V}_{\bm{c}}(\mathbb{Z}/p^{l-1}\mathbb{Z}).

Whenever [𝒙]𝒱𝒄(/pl1)[\bm{x}]\in\mathcal{V}_{\bm{c}}(\mathbb{Z}/p^{l-1}\mathbb{Z}) lies over a smooth point of V𝒄V_{\bm{c}}, the fiber of (7.22) over [𝒙][\bm{x}] has size exactly pmp^{m_{\ast}} (by Hensel’s lemma). Therefore, Corollary 7.5 simplifies to

(7.23) p2lS𝒄(pl)=|(l)|pm|(l1)|,p^{-2l}S_{\bm{c}}(p^{l})=\lvert\mathcal{B}^{\prime}(l)\rvert-p^{m_{\ast}}\lvert\mathcal{B}^{\prime}(l-1)\rvert,

where (v)\mathcal{B}^{\prime}(v) denotes the subset of 𝒱𝒄(/pv)\mathcal{V}_{\bm{c}}(\mathbb{Z}/p^{v}\mathbb{Z}) lying over the singular locus of V𝒄V_{\bm{c}}.

If χ(c(i)3)χ(c(j)3)\chi(c(i)^{3})\neq\chi(c(j)^{3}), i.e. χ(ci/3F(i))χ(cj/3F(j))\chi(c^{\star}_{i}/3F_{(i)})\neq\chi(c^{\star}_{j}/3F_{(j)}), for some i,j𝒦i,j\in\mathcal{K}, then (1)=\mathcal{B}^{\prime}(1)=\emptyset: in fact, there are no 𝒙𝔽pm{𝟎}\bm{x}\in\mathbb{F}_{p}^{m}\setminus\{\bm{0}\} with F(𝒙)\nabla{F}(\bm{x}), 𝒄\bm{c} linearly dependent over 𝔽p\mathbb{F}_{p}. So unless

(7.24) χ(c(1)3)==χ(c(m/2)3)\chi(c(1)^{3})=\cdots=\chi(c(m/2)^{3})

holds, we have S𝒄(pl)=0S_{\bm{c}}(p^{l})=0 by (7.23). So, from now on, assume (7.24) holds.

By the conditions (7.24) and p2p\neq 2, there exists λp×\lambda\in\mathbb{Z}_{p}^{\times} such that λck/F(k)(p×)2\lambda\cdot c^{\star}_{k}/F_{(k)}\in(\mathbb{Z}_{p}^{\times})^{2} for all k[m/2]k\in[m/2]. Say λck=F(k)d(k)2\lambda\cdot c^{\star}_{k}=F_{(k)}d(k)^{2} for some choices d(k)p×d(k)\in\mathbb{Z}_{p}^{\times}; write di=d(k)d_{i}=d(k) when i𝒥(k)i\in\mathcal{J}(k). Then (1)\mathcal{B}^{\prime}(1) is the set of 𝔽p\mathbb{F}_{p}-points [𝒙]=[±di¯]i[m][\bm{x}]=[\pm\overline{d_{i}}]_{i\in[m]} with F(𝒙)=0F(\bm{x})=0. But by Proposition 7.6(3), the scheme V𝒄V_{\bm{c}} has exactly 2m/212^{m/2-1} singular points [𝒙]V𝒄(𝔽¯p)[\bm{x}]\in V_{\bm{c}}(\overline{\mathbb{F}}_{p}), which—by the Jacobian criterion, and the fact that ±di¯𝔽p\pm\overline{d_{i}}\in\mathbb{F}_{p} for all i[m]i\in[m]—must all lie in (1)\mathcal{B}^{\prime}(1). Explicitly, these 2m/212^{m/2-1} points [𝒙][\bm{x}] arise from the sign choices for which 𝒙Λ𝔽p\bm{x}\in\Lambda\otimes\mathbb{F}_{p}.

We now seek to count (l)\mathcal{B}^{\prime}(l) for l2l\geq 2. Recall (7.18), expressing 𝒱𝒄\mathcal{V}_{\bm{c}} in terms of the (𝒉,𝒚)(\bm{h},\bm{y}) coordinates. Fix a point [𝒙](1)[\bm{x}]\in\mathcal{B}^{\prime}(1), say given in (𝒉,𝒚)(\bm{h},\bm{y}) coordinates by 𝒉𝟎modp\bm{h}\equiv\bm{0}\bmod{p} and y[k]d(k)modpy[k]\equiv d(k)\bmod{p}. Upon writing ck=F(k)d(k)2/λc^{\star}_{k}=F_{(k)}d(k)^{2}/\lambda, the system (7.18) modulo plp^{l} becomes

(7.25) 1km/2F(k)h[k]y[k]2pl31km/2F(k)h[k]3,1km/2F(k)d(k)2h[k]pl0.\sum_{1\leq k\leq m/2}F_{(k)}h[k]y[k]^{2}\equiv_{p^{l}}-3\sum_{1\leq k\leq m/2}F_{(k)}h[k]^{3},\qquad\sum_{1\leq k\leq m/2}F_{(k)}d(k)^{2}h[k]\equiv_{p^{l}}0.

Let 𝒅(l)\mathcal{B}^{\prime}_{\bm{d}}(l) be the set of solutions (𝒉,𝒚)(\bm{h},\bm{y}) to (7.25) lying over our fixed [𝒙](1)[\bm{x}]\in\mathcal{B}^{\prime}(1).

Fix an affine chart (i.e. representatives in 𝒅(l)\mathcal{B}^{\prime}_{\bm{d}}(l)) by setting y[m/2]=d(m/2)y[m/2]=d(m/2) identically over p\mathbb{Z}_{p}. Write h[k]=pshs[k]h[k]=p^{s}h_{s}[k] and y[k]=d(k)+psys[k]y[k]=d(k)+p^{s}y_{s}[k] (with s=1s=1 for now, but all s1s\geq 1 to be relevant below), so ys[m/2]=0y_{s}[m/2]=0. Then (7.25) becomes

km/2F(k)hs[k](2d(k)ys[k]+psys[k]2)pl2s3pskF(k)hs[k]3,kF(k)d(k)2hs[k]pls0.\sum_{k\neq m/2}F_{(k)}h_{s}[k](2d(k)y_{s}[k]+p^{s}y_{s}[k]^{2})\equiv_{p^{l-2s}}-3p^{s}\sum_{k}F_{(k)}h_{s}[k]^{3},\;\sum_{k}F_{(k)}d(k)^{2}h_{s}[k]\equiv_{p^{l-s}}0.

So |𝒅(l)|=pm2|𝒜1(l2)|\lvert\mathcal{B}^{\prime}_{\bm{d}}(l)\rvert=p^{m-2}\lvert\mathcal{A}_{1}(l-2)\rvert, where 𝒜s(l)\mathcal{A}_{s}(l) is the (non-homogeneous, affine) system

km/2F(k)hs[k](2d(k)ys[k]+psys[k]2)pl3pskF(k)hs[k]3,kF(k)d(k)2hs[k]pl0.\sum_{k\neq m/2}F_{(k)}h_{s}[k](2d(k)y_{s}[k]+p^{s}y_{s}[k]^{2})\equiv_{p^{l}}-3p^{s}\sum_{k}F_{(k)}h_{s}[k]^{3},\quad\sum_{k}F_{(k)}d(k)^{2}h_{s}[k]\equiv_{p^{l}}0.

Fix s1s\geq 1. Clearly |𝒜s(0)|=1\lvert\mathcal{A}_{s}(0)\rvert=1, while 𝒜s(1)\mathcal{A}_{s}(1) is isomorphic to a cone over a smooth666hs[m/2]h_{s}[m/2] is determined by the remaining hs[k]h_{s}[k], and km/2F(k)d(k)hs[k]ys[k]=0\sum_{k\neq m/2}F_{(k)}d(k)\cdot h_{s}[k]y_{s}[k]=0 is smooth. quadric in m2m-2 variables (i.e. in m\mathbb{P}^{m_{\ast}}, of even dimension m1m_{\ast}-1) with discriminant in (1)m/21(𝔽p×)2(-1)^{m/2-1}(\mathbb{F}_{p}^{\times})^{2}, so |𝒜s(1)|=pm+(p1)p(m1)/2\lvert\mathcal{A}_{s}(1)\rvert=p^{m_{\ast}}+(p-1)p^{(m_{\ast}-1)/2}. For l2l\geq 2, the origin of the cone 𝒜s(1)\mathcal{A}_{s}(1) contributes pm2|𝒜s+1(l2)|p^{m-2}\lvert\mathcal{A}_{s+1}(l-2)\rvert points to 𝒜s(l)\mathcal{A}_{s}(l), while points away from the origin (i.e. smooth points!) lift uniformly to a total of (|𝒜s(1)|1)p(l1)m(\lvert\mathcal{A}_{s}(1)\rvert-1)\cdot p^{(l-1)m_{\ast}} points of 𝒜s(l)\mathcal{A}_{s}(l). Thus

(7.26) |𝒜s(l)|=pm+1|𝒜s+1(l2)|+(|𝒜s(1)|1)p(l1)m\lvert\mathcal{A}_{s}(l)\rvert=p^{m_{\ast}+1}\lvert\mathcal{A}_{s+1}(l-2)\rvert+(\lvert\mathcal{A}_{s}(1)\rvert-1)\cdot p^{(l-1)m_{\ast}}

for s1s\geq 1 and l2l\geq 2. (The same holds for l=1l=1, provided we interpret |𝒜s(1)|\colonequalsp(m+1)\lvert\mathcal{A}_{s}(-1)\rvert\colonequals p^{-(m_{\ast}+1)}.) By induction on l0l\geq 0 (with base cases l=0,1l=0,1), we immediately find that |𝒜s(l)|\lvert\mathcal{A}_{s}(l)\rvert is independent of the choice of λ\lambda and the pp-adic square roots d(k)d(k); furthermore, |𝒜s(l)|=|𝒜1(l)|\lvert\mathcal{A}_{s}(l)\rvert=\lvert\mathcal{A}_{1}(l)\rvert for all s1s\geq 1, i.e. there is no 𝒅\bm{d}-dependence or ss-dependence!

Finally, by symmetry, |(l)|=2m/21|𝒅(l)|=2m/21pm2|𝒜1(l2)|\lvert\mathcal{B}^{\prime}(l)\rvert=2^{m/2-1}\lvert\mathcal{B}^{\prime}_{\bm{d}}(l)\rvert=2^{m/2-1}p^{m-2}\lvert\mathcal{A}_{1}(l-2)\rvert for all l1l\geq 1. (For l=1l=1, recall |𝒜1(1)|\colonequalsp(m+1)=p(m2)\lvert\mathcal{A}_{1}(-1)\rvert\colonequals p^{-(m_{\ast}+1)}=p^{-(m-2)}.) Thus (7.23) gives

p2lS𝒄(pl)=|(l)|pm|(l1)|=2m/21pm2(|𝒜1(l2)|pm|𝒜1(l3)|)p^{-2l}S_{\bm{c}}(p^{l})=\lvert\mathcal{B}^{\prime}(l)\rvert-p^{m_{\ast}}\lvert\mathcal{B}^{\prime}(l-1)\rvert=2^{m/2-1}p^{m-2}(\lvert\mathcal{A}_{1}(l-2)\rvert-p^{m_{\ast}}\lvert\mathcal{A}_{1}(l-3)\rvert)

for l2l\geq 2. To prove (7.21), it remains to show that

|𝒜1(l)|pm|𝒜1(l1)|=(p1)pl(m2)/21=(p1)pl(m+1)/21\lvert\mathcal{A}_{1}(l)\rvert-p^{m_{\ast}}\lvert\mathcal{A}_{1}(l-1)\rvert=(p-1)p^{l(m-2)/2-1}=(p-1)p^{l(m_{\ast}+1)/2-1}

for l0l\geq 0. To this end, we compute |𝒜1(l)|pm|𝒜1(l1)|\lvert\mathcal{A}_{1}(l)\rvert-p^{m_{\ast}}\lvert\mathcal{A}_{1}(l-1)\rvert (using (7.26) if l2l\geq 2) to get

  1. (1)

    1pmp(m+1)=1p11-p^{m_{\ast}}\cdot p^{-(m_{\ast}+1)}=1-p^{-1}, i.e. (p1)p1(p-1)p^{-1}, for l=0l=0;

  2. (2)

    [pm+(p1)p(m1)/2]pm1=(p1)p(m1)/2[p^{m_{\ast}}+(p-1)p^{(m_{\ast}-1)/2}]-p^{m_{\ast}}\cdot 1=(p-1)p^{(m_{\ast}-1)/2}, i.e. (p1)p(m+1)/21(p-1)p^{(m_{\ast}+1)/2-1}, for l=1l=1;

  3. (3)

    pm+1(|𝒜1(l2)|pm|𝒜1(l3)|)p^{m_{\ast}+1}(\lvert\mathcal{A}_{1}(l-2)\rvert-p^{m_{\ast}}\lvert\mathcal{A}_{1}(l-3)\rvert) for l2l\geq 2, since p(l1)m=pmp(l2)mp^{(l-1)m_{\ast}}=p^{m_{\ast}}\cdot p^{(l-2)m_{\ast}}.

By induction on l0l\geq 0, we are done, since pm+1p(l2)(m+1)/21=pl(m+1)/21p^{m_{\ast}+1}\cdot p^{(l-2)(m_{\ast}+1)/2-1}=p^{l(m_{\ast}+1)/2-1}. ∎

Remark 7.10.

By induction on l0l\geq 0 (with base cases l=0,1l=0,1), one could prove the explicit formula |𝒜s(l)|=plm+(p1)pl(m+1)/21(pl(m1)/21)/(p(m1)/21)\lvert\mathcal{A}_{s}(l)\rvert=p^{lm_{\ast}}+(p-1)p^{l(m_{\ast}+1)/2-1}(p^{l(m_{\ast}-1)/2}-1)/(p^{(m_{\ast}-1)/2}-1) (also valid for l=1l=-1). One could then explicitly compute |𝒅(l)|=pm+1|𝒜1(l2)|\lvert\mathcal{B}^{\prime}_{\bm{d}}(l)\rvert=p^{m_{\ast}+1}\lvert\mathcal{A}_{1}(l-2)\rvert for l1l\geq 1.

For the rest of the paper, drop the assumption (7.16). We can finally state and prove the main result of §7. Let (p)×\colonequals{q×:vp(q)=0}=p×\mathbb{Z}_{(p)}^{\times}\colonequals\{q\in\mathbb{Q}^{\times}:v_{p}(q)=0\}=\mathbb{Z}_{p}^{\times}\cap\mathbb{Q}.

Lemma 7.11.

Assume FF is diagonal, with m{4,6}m\in\{4,6\}. Let 𝐜Λ\bm{c}\in\Lambda^{\perp}. Assume (7.14). Then

(7.27) S𝒄(p)=ϕ(p)p1/2+O(1).S^{\natural}_{\bm{c}}(p)=\phi(p)p^{-1/2}+O(1).

Also, c(k)3(p)×c(k)^{3}\in\mathbb{Z}_{(p)}^{\times} for all k𝒦k\in\mathcal{K}. Finally,

(7.28) S𝒄(pl)=ϕ(pl)pl/21km/21(1+χ(c(k)3c(k+1)3))ϕ(pl)pl/2S^{\natural}_{\bm{c}}(p^{l})=\phi(p^{l})p^{-l/2}\cdot\prod_{1\leq k\leq m/2-1}\left(1+\chi{\left(c(k)^{3}c(k+1)^{3}\right)}\right)\ll\phi(p^{l})p^{-l/2}

for all integers l2l\geq 2. The implied constants in (7.27) and (7.28) depend only on mm.

Proof.

As we noted earlier, (7.14) implies (7.15). Now consider the unique invertible [1/F1Fm]\mathbb{Z}[1/F_{1}\cdots F_{m}]-linear map 𝒙𝒙\bm{x}\mapsto\bm{x}^{\prime} such that Fixi3=F(k)(xi)3F_{i}x_{i}^{3}=F_{(k)}(x^{\prime}_{i})^{3} for all k𝒦k\in\mathcal{K} and i𝒥(k)i\in\mathcal{J}(k). This map transforms F(𝒙)F(\bm{x}) into F(𝒙)=F1(x1)3++Fm(xm)3F^{\prime}(\bm{x}^{\prime})=F^{\prime}_{1}(x^{\prime}_{1})^{3}+\dots+F^{\prime}_{m}(x^{\prime}_{m})^{3}, where Fi=F(k)F^{\prime}_{i}=F_{(k)} for all k𝒦k\in\mathcal{K} and i𝒥(k)i\in\mathcal{J}(k). If we let 𝒄𝒄\bm{c}\mapsto\bm{c}^{\prime} denote the dual linear map, then the following hold:

  • if pF1Fmp\nmid F_{1}\cdots F_{m}, then S𝒄(pl)S_{\bm{c}}(p^{l}), defined using FF as in (1.5), equals S𝒄(pl)S_{\bm{c}^{\prime}}(p^{l}), defined using FF^{\prime} in place of FF;

  • one can define the polynomial (F)(𝒄)(F^{\prime})^{\vee}(\bm{c}^{\prime}) to be F(𝒄)F^{\vee}(\bm{c}) times a power of F1FmF_{1}\cdots F_{m};

  • the vector space LL^{\prime} corresponding to LL is still associated to 𝒥\mathcal{J}; and

  • we have ci3/Fi=(ci)3/Fic_{i}^{3}/F_{i}=(c^{\prime}_{i})^{3}/F^{\prime}_{i} for all i[m]i\in[m].

By (7.15), we may thus assume (7.16) (when proving Lemma 7.11).

The claim (7.27) now follows upon plugging Lemma 7.7 and (7.1) into (7.2). For the claim c(k)3(p)×c(k)^{3}\in\mathbb{Z}_{(p)}^{\times}, see Proposition 7.6(1)–(2). Finally, Lemma 7.9 implies (7.28) for l2l\geq 2. ∎

Remark 7.12.

Since 𝔼𝒄Λ/nΛ[S𝒄(n)]=ϕ(n)n1/2\mathbb{E}_{\bm{c}\in\Lambda^{\perp}/n\Lambda^{\perp}}[S^{\natural}_{\bm{c}}(n)]=\phi(n)n^{-1/2} (for all n1n\geq 1) by Proposition 4.7, we have chosen to formulate Lemma 7.11 using ϕ(n)n1/2\phi(n)n^{-1/2}, not n1/2n^{1/2}.

8. Main delta-method analysis

Fix FF in Theorem 1.1. For each LΥL\in\Upsilon, recall Λ\Lambda, Λ\Lambda^{\perp} from Definition 1.3. By Proposition 3.1, F(𝒄)=0F^{\vee}(\bm{c})=0 for all 𝒄LΥΛ\bm{c}\in\bigcup_{L\in\Upsilon}\Lambda^{\perp}. (Recall, from Definition 1.4, that we call such 𝒄\bm{c}’s linear.)

Since FF is diagonal, Proposition 3.6 characterizes the linear 𝒄\bm{c}’s via certain pairings introduced in Definition 3.5. More precisely, the identification Λ=𝒥\Lambda^{\perp}=\mathcal{R}_{\mathcal{J}} defines a bijection between Υ\Upsilon and the set of equivalence classes of permissible pairings 𝒥\mathcal{J} of [m][m].

Consider the left-hand side of (1.6). Recall the sets 1\mathcal{E}_{1}, 2\mathcal{E}_{2} from (1.9).

Proposition 8.1.

For reals C1C\geq 1, we have |1[C,C]m|ϵCm/21+ϵ\lvert\mathcal{E}_{1}\cap[-C,C]^{m}\rvert\ll_{\epsilon}C^{m/2-1+\epsilon}.

Proof.

This follows from the combinatorics of [heath1998circle]*p. 687. (In Heath-Brown’s notation, any exponent kek2\sum_{k}\frac{e_{k}}{2} coming from 1\mathcal{E}_{1} must lie in {22}\{\frac{2}{2}\} if m=4m=4, and in {42,32,2+22,22}\{\frac{4}{2},\frac{3}{2},\frac{2+2}{2},\frac{2}{2}\} if m=6m=6.) ∎

Proposition 8.2.

For reals C1C\geq 1, we have |2[C,C]m|ϵCm/21+ϵ\lvert\mathcal{E}_{2}\cap[-C,C]^{m}\rvert\ll_{\epsilon}C^{m/2-1+\epsilon}.

Proof.

Let s:,qq2s\colon\mathbb{Q}\to\mathbb{Q},\;q\mapsto q^{2}. Let M\colonequalsmax(|F1|,,|Fm|)M\colonequals\max(\lvert F_{1}\rvert,\dots,\lvert F_{m}\rvert). Suppose LΥL\in\Upsilon and 𝒄2Λ\bm{c}\in\mathcal{E}_{2}\cap\Lambda^{\perp}. Then by Observation 3.10(2), there exist distinct k1,k2𝒦k_{1},k_{2}\in\mathcal{K} with c(k1)3c(k2)3s()c(k_{1})^{3}c(k_{2})^{3}\in s(\mathbb{Q}). Fix i𝒥(k1)i\in\mathcal{J}(k_{1}) and j𝒥(k2)j\in\mathcal{J}(k_{2}); then (ci3/Fi)(cj3/Fj)s()(c_{i}^{3}/F_{i})(c_{j}^{3}/F_{j})\in s(\mathbb{Q}). So (Fici)(Fjcj)s()(F_{i}c_{i})(F_{j}c_{j})\in s(\mathbb{Z}), whence there exists aa\in\mathbb{Z} with 0<|a|MC0<\lvert a\rvert\leq MC such that Fici/as()F_{i}c_{i}/a\in s(\mathbb{Z}) and Fjcj/as()F_{j}c_{j}/a\in s(\mathbb{Z}). Since |𝒦|=m/2\lvert\mathcal{K}\rvert=m/2, it follows (upon summing over all possibilities for 𝒥\mathcal{J}, k1k_{1}, k2k_{2}, ii, jj) that 2[C,C]m\mathcal{E}_{2}\cap[-C,C]^{m} has size mCm/220<|a|MC(MC/|a|)1/2(MC/|a|)1/2MCm/21log(2+C)\ll_{m}C^{m/2-2}\sum_{0<\lvert a\rvert\leq MC}(MC/\lvert a\rvert)^{1/2}\cdot(MC/\lvert a\rvert)^{1/2}\ll_{M}C^{m/2-1}\log(2+C). ∎

Propositions 8.1 and 8.2 imply that FF^{\vee} is unsurprising (in the sense of Definition 1.5). Lemma 5.2 then gives the useful bound f(12)ϵX(m1)/2+ϵf(\mathcal{E}_{1}\cup\mathcal{E}_{2})\ll_{\epsilon}X^{(m-1)/2+\epsilon}. (Recall f(𝒮)f(\mathcal{S}) from (5.1).)

Corollary 8.3.

The equality (1.6) holds, provided that for each LΥL\in\Upsilon, we have

(8.1) Y2𝒄Λ2n1n(1m)/2S𝒄(n)I𝒄(n)=Oϵ(Xm/21/4+ϵ)+σ,L,wXm/2.Y^{-2}\sum_{\bm{c}\in\Lambda^{\perp}\setminus\mathcal{E}_{2}}\sum_{n\geq 1}n^{(1-m)/2}S^{\natural}_{\bm{c}}(n)I_{\bm{c}}(n)=O_{\epsilon}(X^{m/2-1/4+\epsilon})+\sigma_{\infty,L^{\perp},w}X^{m/2}.
Proof.

Assume (8.1) for LΥL\in\Upsilon. Since f(2)ϵX(m1)/2+ϵf(\mathcal{E}_{2})\ll_{\epsilon}X^{(m-1)/2+\epsilon}, the relation (8.1) implies

Y2𝒄Λ{𝟎}n1n(1m)/2S𝒄(n)I𝒄(n)=Oϵ(Xm/21/4+ϵ)+σ,L,wXm/2.Y^{-2}\sum_{\bm{c}\in\Lambda^{\perp}\setminus\{\bm{0}\}}\sum_{n\geq 1}n^{(1-m)/2}S^{\natural}_{\bm{c}}(n)I_{\bm{c}}(n)=O_{\epsilon}(X^{m/2-1/4+\epsilon})+\sigma_{\infty,L^{\perp},w}X^{m/2}.

Upon summing over the finite set Υ\Upsilon (handling intersections using Lemma 5.2), we obtain

Y2𝒄LΥΛ{𝟎}n1n(1m)/2S𝒄(n)I𝒄(n)=Oϵ(Xm/21/4+ϵ)+LΥσ,L,wXm/2.Y^{-2}\sum_{\bm{c}\in\bigcup_{L\in\Upsilon}\Lambda^{\perp}\setminus\{\bm{0}\}}\sum_{n\geq 1}n^{(1-m)/2}S^{\natural}_{\bm{c}}(n)I_{\bm{c}}(n)=O_{\epsilon}(X^{m/2-1/4+\epsilon})+\sum_{L\in\Upsilon}\sigma_{\infty,L^{\perp},w}X^{m/2}.

By (1.7), (1.8), and the bound f(1)ϵX(m1)/2+ϵf(\mathcal{E}_{1})\ll_{\epsilon}X^{(m-1)/2+\epsilon}, the desired (1.6) follows. ∎

So (8.1) would imply Theorem 1.1. The rest of §8 is devoted to the proof of (8.1). Fix LΥL\in\Upsilon, and recall Proposition 3.6. We first explain why Heath-Brown’s approach for m=4m=4 in [heath1998circle] does not seem to directly extend to m=6m=6; we then describe our approach.

Using Lemma 4.6 (with n0=nn_{0}=n and n1=1n_{1}=1) and Proposition 4.7, one can show that (in terms of certain quantities T(𝒋;n),J(𝒋;n)T(\bm{j};n),J(\bm{j};n) we briefly discussed in §4)

(8.2) X3𝒄𝒥n1nmS𝒄(n)I𝒄(n)=X3n1nm/2𝒋m/2T(𝒋;n)J(𝒋;n);X^{-3}\sum_{\bm{c}\in\mathcal{R}_{\mathcal{J}}}\sum_{n\geq 1}n^{-m}S_{\bm{c}}(n)I_{\bm{c}}(n)=X^{-3}\sum_{n\geq 1}n^{-m/2}\sum_{\bm{j}\in\mathbb{Z}^{m/2}}T(\bm{j};n)J(\bm{j};n);

cf. [heath1998circle]*p. 692, Poisson summation underlying Lemma 8.2. When m=4m=4, Heath-Brown proves that 𝒋=𝟎\bm{j}=\bm{0} in (8.2) captures the “𝒥\mathcal{J}-diagonal” contribution to (1.1), and that the locus 𝒋𝟎\bm{j}\neq\bm{0} in (8.2) forms an “error term” of ϵX3/2+ϵ\ll_{\epsilon}X^{3/2+\epsilon}.

When m=4m=4 and σ,L,w0\sigma_{\infty,L^{\perp},w}\neq 0, the 𝒥\mathcal{J}-diagonal in (1.1) strictly dominates the 𝒄=𝟎\bm{c}=\bm{0} contribution to (8.2). When m=6m=6 and σ,F,wσ,L,w0\sigma_{\infty,F,w}\cdot\sigma_{\infty,L^{\perp},w}\neq 0, however, 𝒄=𝟎\bm{c}=\bm{0} in (8.2) is comparable in size to the 𝒥\mathcal{J}-diagonal in (1.1), so that 𝒋𝟎\bm{j}\neq\bm{0} in (8.2) is likely no longer an error term. Perhaps for typical 𝒋𝟎\bm{j}\neq\bm{0}, the sums T(𝒋;n)T(\bm{j};n) can be analyzed in terms of LL-functions, but it is not clear where the 𝒄=𝟎\bm{c}=\bm{0} contribution to (8.2) would arise for m=6m=6. To push (8.2) further—perhaps by considering small and large nn separately—thus seems technical and possibly delicate, though it could be enlightening.777Heath-Brown’s argument for m=4m=4 is already challenging, and the geometry involved might become even more complicated as mm grows (though the parity of m/2m/2 may also play some role).

Our approach to (8.1) delays Poisson summation to the “endgame”, thus sidestepping (8.2). Consider the left-hand side of (8.1). We open not with Poisson summation over 𝒄Λ\bm{c}\in\Lambda^{\perp}, but with local geometry (Lemma 7.11). Lemma 7.11 exposes a uniform bias in S𝒄(pl)S_{\bm{c}}(p^{l}) over 𝒄Λ2\bm{c}\in\Lambda^{\perp}\setminus\mathcal{E}_{2}, allowing us to decompose S𝒄(n)S_{\bm{c}}(n) into simpler pieces (see (8.11)). Even then, tricky issues remain (especially regarding the excised loci {𝟎}\{\bm{0}\} and 2{𝟎}\mathcal{E}_{2}\setminus\{\bm{0}\}), but Lemma 7.11 is undoubtedly the driving force in our argument. (However, if we could compute I𝒄(n)I_{\bm{c}}(n) to greater precision when 𝒄Λ2\bm{c}\in\Lambda^{\perp}\setminus\mathcal{E}_{2}, that might reduce our reliance on Lemma 7.11.)

For 𝒄Λ\bm{c}\in\Lambda^{\perp}, consider the Dirichlet series Φ(𝒄,s)\colonequalsn1nsS𝒄(n)\Phi(\bm{c},s)\colonequals\sum_{n\geq 1}n^{-s}S^{\natural}_{\bm{c}}(n). Let

Ψ(s)\colonequalsn1nsϕ(n)n1/2=ζ(s+1/2)1ζ(s1/2)\Psi(s)\colonequals\sum_{n\geq 1}n^{-s}\phi(n)n^{-1/2}=\zeta(s+1/2)^{-1}\cdot\zeta(s-1/2)

be the Dirichlet series for ϕ(n)n1/2\phi(n)n^{-1/2}. By Lemma 7.11, Φ(𝒄,s)\Phi(\bm{c},s) should typically resemble Ψ(s)\Psi(s), to leading order. So divide Φ(𝒄,s)\Phi(\bm{c},s) by Ψ(s)\Psi(s) to define the “error series”

(8.3) n1nsS𝒄,0(n)\colonequalsΦ(𝒄,s)/Ψ(s)=ζ(s1/2)1ζ(s+1/2)Φ(𝒄,s).\sum_{n\geq 1}n^{-s}S^{\natural}_{\bm{c},0}(n)\colonequals\Phi(\bm{c},s)/\Psi(s)=\zeta(s-1/2)^{-1}\cdot\zeta(s+1/2)\cdot\Phi(\bm{c},s).

Since ζ(s)1=n1nsμ(n)\zeta(s)^{-1}=\sum_{n\geq 1}n^{-s}\mu(n) and ζ(s)=n1ns\zeta(s)=\sum_{n\geq 1}n^{-s}, it follows from (8.3) that

(8.4) S𝒄,0(n)=d0d1d2=nμ(d0)d01/2d11/2S𝒄(d2).S^{\natural}_{\bm{c},0}(n)=\sum_{d_{0}d_{1}d_{2}=n}\mu(d_{0})d_{0}^{1/2}\cdot d_{1}^{-1/2}\cdot S^{\natural}_{\bm{c}}(d_{2}).

For convenience, let S𝒄,0(n)\colonequalsn(1+m)/2S𝒄,0(n)S_{\bm{c},0}(n)\colonequals n^{(1+m)/2}S^{\natural}_{\bm{c},0}(n). The multiplicativity of S𝒄(n)S_{\bm{c}}(n), ϕ(n)\phi(n) in nn leads to Euler products for Φ\Phi, Ψ\Psi, and then to multiplicativity of S𝒄,0(n)S_{\bm{c},0}(n).

We will need some basic properties of S𝒄,0(n)S_{\bm{c},0}(n) as a function of 𝒄Λ\bm{c}\in\Lambda^{\perp} and n1n\geq 1.

Proposition 8.4.

The quantity S𝐜,0(n)S_{\bm{c},0}(n) is a function of nn and 𝐜modn\bm{c}\bmod{n}. Also,

(8.5) 𝔼𝒄Λ/nΛ[S𝒄,0(n)]=𝟏n=1.\mathbb{E}_{\bm{c}\in\Lambda^{\perp}/n\Lambda^{\perp}}[S^{\natural}_{\bm{c},0}(n)]=\bm{1}_{n=1}.
Proof.

By the first sentence of Proposition 4.7, S𝒄(n)S_{\bm{c}}(n) depends at most on nn and 𝒄modn\bm{c}\bmod{n}. But S𝒄,0(n)S_{\bm{c},0}(n) depends at most on the list of values (S𝒄(d2))d2n(S_{\bm{c}}(d_{2}))_{d_{2}\mid n}, hence at most on nn and 𝒄modn\bm{c}\bmod{n}. Yet by (8.3), the Dirichlet series identity n1nsS𝒄(n)=Ψ(s)n1nsS𝒄,0(n)\sum_{n\geq 1}n^{-s}S^{\natural}_{\bm{c}}(n)=\Psi(s)\sum_{n\geq 1}n^{-s}S^{\natural}_{\bm{c},0}(n) holds for any 𝒄Λ\bm{c}\in\Lambda^{\perp}. Averaging formally (coefficient-wise) over 𝒄Λ\bm{c}\in\Lambda^{\perp}, we get

(8.6) n1ns𝔼𝒄Λ/nΛ[S𝒄(n)]=Ψ(s)n1ns𝔼𝒄Λ/nΛ[S𝒄,0(n)].\sum_{n\geq 1}n^{-s}\mathbb{E}_{\bm{c}\in\Lambda^{\perp}/n\Lambda^{\perp}}[S^{\natural}_{\bm{c}}(n)]=\Psi(s)\sum_{n\geq 1}n^{-s}\mathbb{E}_{\bm{c}\in\Lambda^{\perp}/n\Lambda^{\perp}}[S^{\natural}_{\bm{c},0}(n)].

But by the final sentence of Proposition 4.7, the left-hand side of (8.6) equals Ψ(s)\Psi(s). So (8.5) follows formally by division. ∎

We now provide some bounds on S𝒄,0(n)S_{\bm{c},0}(n) for 𝒄Λ\bm{c}\in\Lambda^{\perp}. By (8.4), we have

(8.7) n1/2S𝒄,0(n)=d0d1d2=nμ(d0)d11d21/2S𝒄(d2).n^{-1/2}S^{\natural}_{\bm{c},0}(n)=\sum_{d_{0}d_{1}d_{2}=n}\mu(d_{0})\cdot d_{1}^{-1}\cdot d_{2}^{-1/2}S^{\natural}_{\bm{c}}(d_{2}).

Let τ3(n)\colonequalsd0d1d2=n1\tau_{3}(n)\colonequals\sum_{d_{0}d_{1}d_{2}=n}1. For any n1n\geq 1, the triangle inequality on (8.7), followed by an application of (5.8) to cube-free divisors of nn, yields

(8.8) |S𝒄,0(n)|τ3(n)n1/2maxdnd1/2|S𝒄(d)|ϵτ3(n)n1/2+ϵmaxd𝒩(3):dnd1/2|S𝒄(d)|.\lvert S^{\natural}_{\bm{c},0}(n)\rvert\leq\tau_{3}(n)n^{1/2}\cdot\max_{d\mid n}d^{-1/2}\lvert S^{\natural}_{\bm{c}}(d)\rvert\ll_{\epsilon}\tau_{3}(n)n^{1/2+\epsilon}\cdot\max_{d\in\mathcal{N}_{\geq}(3):\,d\mid n}d^{-1/2}\lvert S^{\natural}_{\bm{c}}(d)\rvert.

If pj2m/21F(𝒄)p\nmid j^{2^{m/2-1}}{F^{\vee}}(\bm{c}) and l2l\geq 2, then (8.8) and (7.28) imply

(8.9) |S𝒄,0(pl)|τ3(pl)pl/2.\lvert S^{\natural}_{\bm{c},0}(p^{l})\rvert\ll\tau_{3}(p^{l})p^{l/2}.

If pj2m/21F(𝒄)p\nmid j^{2^{m/2-1}}{F^{\vee}}(\bm{c}), then (8.4) and (7.27) imply

(8.10) S𝒄,0(p)=p1/2+p1/2+S𝒄(p)=O(1).S^{\natural}_{\bm{c},0}(p)=-p^{1/2}+p^{-1/2}+S^{\natural}_{\bm{c}}(p)=O(1).

The bounds (8.8), (8.9), (8.10) are most useful in conjunction with multiplicativity. Let

𝒩𝒄\colonequals{n1:pnpj2m/21F(𝒄)},𝒩𝒄\colonequals{n1:pnpj2m/21F(𝒄)}\mathcal{N}^{\bm{c}}\colonequals\{n\geq 1:p\mid n\Rightarrow p\nmid j^{2^{m/2-1}}{F^{\vee}}(\bm{c})\},\quad\mathcal{N}_{\bm{c}}\colonequals\{n\geq 1:p\mid n\Rightarrow p\mid j^{2^{m/2-1}}{F^{\vee}}(\bm{c})\}

for each 𝒄Λ\bm{c}\in\Lambda^{\perp}. For each integer t1t\geq 1, let

𝒩(t)\colonequals{n1:pnvp(n)t},𝒩(t)\colonequals{n1:pnvp(n)t}.\mathcal{N}_{\leq}(t)\colonequals\{n\geq 1:p\mid n\Rightarrow v_{p}(n)\leq t\},\quad\mathcal{N}_{\geq}(t)\colonequals\{n\geq 1:p\mid n\Rightarrow v_{p}(n)\geq t\}.
Lemma 8.5 (Cf. [wang2023_large_sieve_diagonal_cubic_forms]*Lemma 3.4).

Let N,t1N,t\geq 1 be integers. Then |{Nn<2N:n𝒩(t)}|tN1/t\lvert\{N\leq n<2N:n\in\mathcal{N}_{\geq}(t)\}\rvert\ll_{t}N^{1/t}. Also, if 𝐜Λ2\bm{c}\in\Lambda^{\perp}\setminus\mathcal{E}_{2}, then |{Nn<2N:n𝒩𝐜}|ϵNϵ𝐜ϵ\lvert\{N\leq n<2N:n\in\mathcal{N}_{\bm{c}}\}\rvert\ll_{\epsilon}N^{\epsilon}\lVert\bm{c}\rVert^{\epsilon}.

Proof.

The first bound is familiar. It remains to prove the second. Say 𝒄Λ2\bm{c}\in\Lambda^{\perp}\setminus\mathcal{E}_{2}. Then the vector j2m/21F(𝒄)j^{2^{m/2-1}}{F^{\vee}}(\bm{c}) is nonzero, and thus has a nonzero coordinate RR. But #{Nn<2N:nR}ϵ(RN)ϵ\#\{N\leq n<2N:n\mid R^{\infty}\}\ll_{\epsilon}(RN)^{\epsilon} (by Rankin’s trick), and R𝒄degFR\ll\lVert\bm{c}\rVert^{\deg F^{\vee}}. ∎

Lemma 8.6.

Let 𝐜Λ2\bm{c}\in\Lambda^{\perp}\setminus\mathcal{E}_{2} and N{1,2,4,8,}N\in\{1,2,4,8,\ldots\}. Suppose 𝐜X10\lVert\bm{c}\rVert\leq X^{10}. Then

Nn<2N|S𝒄,0(n)|ϵ(XN)ϵNd<2N:d𝒩(3)𝒩𝒄d1|S𝒄(d)|.\sum_{N\leq n<2N}\lvert S^{\natural}_{\bm{c},0}(n)\rvert\ll_{\epsilon}(XN)^{\epsilon}\cdot N\sum_{d<2N:\,d\in\mathcal{N}_{\geq}(3)\cap\mathcal{N}_{\bm{c}}}d^{-1}\lvert S^{\natural}_{\bm{c}}(d)\rvert.
Proof.

Any integer n1n\geq 1 can be written (uniquely) as n1n2n3n_{1}n_{2}n_{3}, where n1n_{1}, n2n_{2}, n3n_{3} are pairwise coprime integers satisfying n1𝒩𝒄𝒩(1)n_{1}\in\mathcal{N}^{\bm{c}}\cap\mathcal{N}_{\leq}(1), n2𝒩𝒄𝒩(2)n_{2}\in\mathcal{N}^{\bm{c}}\cap\mathcal{N}_{\geq}(2), n3𝒩𝒄n_{3}\in\mathcal{N}_{\bm{c}}. Upon writing S𝒄,0(n)=1i3S𝒄,0(ni)S_{\bm{c},0}(n)=\prod_{1\leq i\leq 3}S_{\bm{c},0}(n_{i}), and applying (8.10) to primes pn1p\mid n_{1}, (8.9) to primes pn2p\mid n_{2}, and (8.8) to n3n_{3}, we get (by dyadic summation over n1n_{1}, n2n_{2}, n3n_{3})

Nn<2N|S𝒄,0(n)|ϵN1,N2,N32N:N/4<N1N2N3<2NNϵn1,n21:N1n1<2N1,N2n2<2N2,n2𝒩(2)n21/2n3,d1:N3n3<2N3,n3𝒩𝒄,d𝒩(3),dn3(n3/d)1/2|S𝒄(d)|.\sum_{N\leq n<2N}\lvert S^{\natural}_{\bm{c},0}(n)\rvert\ll_{\epsilon}\sum_{\begin{subarray}{c}N_{1},N_{2},N_{3}\mid 2N:\\ N/4<N_{1}N_{2}N_{3}<2N\end{subarray}}N^{\epsilon}\sum_{\begin{subarray}{c}n_{1},n_{2}\geq 1:\\ N_{1}\leq n_{1}<2N_{1},\\ N_{2}\leq n_{2}<2N_{2},\;n_{2}\in\mathcal{N}_{\geq}(2)\end{subarray}}n_{2}^{1/2}\sum_{\begin{subarray}{c}n_{3},d\geq 1:\\ N_{3}\leq n_{3}<2N_{3},\;n_{3}\in\mathcal{N}_{\bm{c}},\\ d\in\mathcal{N}_{\geq}(3),\;d\mid n_{3}\end{subarray}}(n_{3}/d)^{1/2}\lvert S^{\natural}_{\bm{c}}(d)\rvert.

Upon summing over n1n_{1}, n2n_{2}, n3n_{3} (for each fixed dd), we get, by Lemma 8.5,

Nn<2N|S𝒄,0(n)|ϵN1,N2,N32N:N1N2N3<2NNϵN1N2d<2N3:d𝒩𝒄𝒩(3)(N3𝒄)ϵ(N3/d)1/2|S𝒄(d)|.\sum_{N\leq n<2N}\lvert S^{\natural}_{\bm{c},0}(n)\rvert\ll_{\epsilon}\sum_{\begin{subarray}{c}N_{1},N_{2},N_{3}\mid 2N:\\ N_{1}N_{2}N_{3}<2N\end{subarray}}N^{\epsilon}N_{1}N_{2}\sum_{\begin{subarray}{c}d<2N_{3}:\\ d\in\mathcal{N}_{\bm{c}}\cap\mathcal{N}_{\geq}(3)\end{subarray}}(N_{3}\lVert\bm{c}\rVert)^{\epsilon}(N_{3}/d)^{1/2}\lvert S^{\natural}_{\bm{c}}(d)\rvert.

But (N3/d)1/2N3/d(N_{3}/d)^{1/2}\ll N_{3}/d for d<2N3d<2N_{3}. And 𝒄X10\lVert\bm{c}\rVert\leq X^{10}. So Lemma 8.6 follows. ∎

Given nn in (8.1), we may use (8.3) to decompose S𝒄(n)S^{\natural}_{\bm{c}}(n) as a Dirichlet convolution:

(8.11) S𝒄(n)=n0n1=nS𝒄,0(n0)ϕ(n1)n11/2.S^{\natural}_{\bm{c}}(n)=\sum_{n_{0}n_{1}=n}S^{\natural}_{\bm{c},0}(n_{0})\cdot\phi(n_{1})n_{1}^{-1/2}.

We will study the ranges n1Y/Pn_{1}\geq Y/P and n1<Y/Pn_{1}<Y/P separately, for a parameter PP to be chosen in (8.16). We first handle the range n1<Y/Pn_{1}<Y/P, using Lemma 8.6. It turns out we will not need the full strength of Lemma 8.6 (which might however still be useful in the future).

Definition 8.7.

For a real number K>0K>0 and a set 𝒯m\mathcal{T}\subseteq\mathbb{Z}^{m}, let

Σ<K(X,𝒯)\colonequalsY2𝒄𝒯n1n(1m)/2I𝒄(n)n0n1=n:n1<KS𝒄,0(n0)ϕ(n1)n11/2.\Sigma_{<K}(X,\mathcal{T})\colonequals Y^{-2}\sum_{\bm{c}\in\mathcal{T}}\sum_{n\geq 1}n^{(1-m)/2}I_{\bm{c}}(n)\sum_{\begin{subarray}{c}n_{0}n_{1}=n:\\ n_{1}<K\end{subarray}}S^{\natural}_{\bm{c},0}(n_{0})\cdot\phi(n_{1})n_{1}^{-1/2}.

Similarly define ΣK(X,𝒯)\Sigma_{\geq K}(X,\mathcal{T}) (by replacing n1<Kn_{1}<K with n1Kn_{1}\geq K).

Using (8.11), one may rewrite the left-hand side of (8.1) as

(8.12) Σ<Y/P(X,Λ2)+ΣY/P(X,Λ)ΣY/P(X,2).\Sigma_{<Y/P}(X,\Lambda^{\perp}\setminus\mathcal{E}_{2})+\Sigma_{\geq Y/P}(X,\Lambda^{\perp})-\Sigma_{\geq Y/P}(X,\mathcal{E}_{2}).
Lemma 8.8.

Uniformly over reals X,P1X,P\geq 1, we have

(8.13) Σ<Y/P(X,Λ2)ϵXm/2+ϵP1/2.\Sigma_{<Y/P}(X,\Lambda^{\perp}\setminus\mathcal{E}_{2})\ll_{\epsilon}X^{m/2+\epsilon}P^{-1/2}.
Proof.

We proceed somewhat crudely. By Proposition 5.1, the absolute value of the left-hand side of (8.13) is at most Oϵ,A(XA)O_{\epsilon,A}(X^{-A}) plus the quantity

(8.14) Y2𝒄Λ2:𝒄X1/2+ϵn0n1<M2Y:n1<Y/P(n0n1)(1m)/2|I𝒄(n0n1)|n11/2|S𝒄,0(n0)|.Y^{-2}\sum_{\begin{subarray}{c}\bm{c}\in\Lambda^{\perp}\setminus\mathcal{E}_{2}:\\ \lVert\bm{c}\rVert\leq X^{1/2+\epsilon}\end{subarray}}\,\sum_{\begin{subarray}{c}n_{0}n_{1}<M_{2}Y:\\ n_{1}<Y/P\end{subarray}}(n_{0}n_{1})^{(1-m)/2}\lvert I_{\bm{c}}(n_{0}n_{1})\rvert\cdot n_{1}^{1/2}\lvert S^{\natural}_{\bm{c},0}(n_{0})\rvert.

We now examine an individual 𝒄Λ2\bm{c}\in\Lambda^{\perp}\setminus\mathcal{E}_{2}. By Observation 3.10(1), c1cm0c_{1}\cdots c_{m}\neq 0. So

(8.15) I𝒄(n)ϵXm+ϵ(X𝒄/n)1im(X|ci|/n)1/2I_{\bm{c}}(n)\ll_{\epsilon}X^{m+\epsilon}(X\lVert\bm{c}\rVert/n)\prod_{1\leq i\leq m}(X\lvert c_{i}\rvert/n)^{-1/2}

by [heath1998circle]*Lemma 3.2. Upon inserting (8.15) into (8.14), dyadically decomposing n0n_{0}, and then applying Lemma 8.6, we find that the quantity (8.14) is

ϵ𝒄Λ2:𝒄X1/2+ϵN0n1<M2Y:N0{1,2,4,8,},n1<Y/PY2Xm+ϵX1m/2𝒄(N0n1)1/2|c1cm|1/2n11/2N0d<2N0:d𝒩(3)d1|S𝒄(d)|.\ll_{\epsilon}\sum_{\begin{subarray}{c}\bm{c}\in\Lambda^{\perp}\setminus\mathcal{E}_{2}:\\ \lVert\bm{c}\rVert\leq X^{1/2+\epsilon}\end{subarray}}\,\sum_{\begin{subarray}{c}N_{0}n_{1}<M_{2}Y:\\ N_{0}\in\{1,2,4,8,\ldots\},\;n_{1}<Y/P\end{subarray}}\frac{Y^{-2}X^{m+\epsilon}X^{1-m/2}\lVert\bm{c}\rVert}{(N_{0}n_{1})^{1/2}\lvert c_{1}\cdots c_{m}\rvert^{1/2}}\cdot n_{1}^{1/2}\cdot N_{0}\sum_{\begin{subarray}{c}d<2N_{0}:\\ d\in\mathcal{N}_{\geq}(3)\end{subarray}}d^{-1}\lvert S^{\natural}_{\bm{c}}(d)\rvert.

But (5.2) yields S𝒄(d)ϵd1/2+ϵ1imsq(ci)1/4S^{\natural}_{\bm{c}}(d)\ll_{\epsilon}d^{1/2+\epsilon}\prod_{1\leq i\leq m}\operatorname{sq}(c_{i})^{1/4} (when c1cm0c_{1}\cdots c_{m}\neq 0). Plugging this in, and noting that d𝒩(3)d1/21\sum_{d\in\mathcal{N}_{\geq}(3)}d^{-1/2}\ll 1 by Lemma 8.5, we find that (8.14) is

ϵ𝒄Λ2:𝒄X1/2+ϵn1<Y/PY2X1+m/2+ϵ𝒄(M2Y/n1)1/2k𝒦sq(cmin𝒥(k))1/2|cmin𝒥(k)|,\ll_{\epsilon}\sum_{\bm{c}\in\Lambda^{\perp}\setminus\mathcal{E}_{2}:\,\lVert\bm{c}\rVert\leq X^{1/2+\epsilon}}\,\sum_{n_{1}<Y/P}Y^{-2}X^{1+m/2+\epsilon}\lVert\bm{c}\rVert\cdot(M_{2}Y/n_{1})^{1/2}\prod_{k\in\mathcal{K}}\frac{\operatorname{sq}(c_{\min\mathcal{J}(k)})^{1/2}}{\lvert c_{\min\mathcal{J}(k)}\rvert},

in the notation of Definition 3.5. But 0<|c|Csq(c)1/2/|c|ϵCϵ\sum_{0<\lvert c\rvert\leq C}\operatorname{sq}(c)^{1/2}/\lvert c\rvert\ll_{\epsilon}C^{\epsilon} by Lemma 8.5 (or a calculation with Euler products). So (8.14) is

ϵn1<Y/PY2X1+m/2+ϵX1/2+ϵ(Y/n1)1/2Y2X3/2+m/2+2ϵY1/2(Y/P)1/2.\ll_{\epsilon}\sum_{n_{1}<Y/P}Y^{-2}X^{1+m/2+\epsilon}X^{1/2+\epsilon}\cdot(Y/n_{1})^{1/2}\ll Y^{-2}X^{3/2+m/2+2\epsilon}Y^{1/2}(Y/P)^{1/2}.

Plugging in Y=X3/2Y=X^{3/2} leads to (8.13). ∎

Remark 8.9.

We have used diagonality of FF. In general (leaving details to the interested reader), one can prove (8.13) for m11m\leq 11 under the axioms 1.6(2)–(3), by replacing (8.15) with (5.7), then using (1.11), (1.12) instead of (5.2), and then verifying the inequality

Y2Xm+ϵN01/2N1P1/2(XC+N0N1)m/21d<2N0:d𝒩(3)Cm/4(Cm/4+dm/12)d1/2ϵXm/2+O(ϵ)\frac{Y^{-2}X^{m+\epsilon}N_{0}^{1/2}N_{1}P^{1/2}}{(XC+N_{0}N_{1})^{m/2-1}}\sum_{d<2N_{0}:\,d\in\mathcal{N}_{\geq}(3)}\frac{C^{m/4}(C^{m/4}+d^{m/12})}{d^{1/2}}\ll_{\epsilon}X^{m/2+O(\epsilon)}

over reals C,P,N0,N11C,P,N_{0},N_{1}\geq 1 with CX1/2+ϵC\leq X^{1/2+\epsilon} and N0,P100(1+M2)Y/N1N_{0},P\leq 100(1+M_{2})Y/N_{1}.

In terms of M1M_{1} from Lemma 4.6, let

(8.16) P=PX\colonequalsM11X1/2.P=P_{X}\colonequals M_{1}^{-1}X^{1/2}.

We now turn to the range n1Y/Pn_{1}\geq Y/P in (8.12). Recall ΣK(X,𝒯)\Sigma_{\geq K}(X,\mathcal{T}) from Definition 8.7.

Lemma 8.10.

Uniformly over reals X1X\geq 1, we have

(8.17) ΣY/P(X,Λ)=σ,L,wXm/2(1+O(P1)).\Sigma_{\geq Y/P}(X,\Lambda^{\perp})=\sigma_{\infty,L^{\perp},w}X^{m/2}(1+O(P^{-1})).
Proof.

Suppose n,n0,n11n,n_{0},n_{1}\geq 1 are integers with n1Y/Pn_{1}\geq Y/P. By the first part of Proposition 8.4, S𝒄,0(n0)S_{\bm{c},0}(n_{0}) only depends on 𝒄modn0\bm{c}\bmod{n_{0}}. Because n1Y/P=M1Xn_{1}\geq Y/P=M_{1}X, Lemma 4.6 therefore implies

𝒄ΛS𝒄,0(n0)I𝒄(n)=𝒃Λ/n0ΛS𝒃,0(n0)n1m/2σ,L,wXm/2h(n/Y,0).\sum_{\bm{c}\in\Lambda^{\perp}}S^{\natural}_{\bm{c},0}(n_{0})\cdot I_{\bm{c}}(n)=\sum_{\bm{b}\in\Lambda^{\perp}/n_{0}\Lambda^{\perp}}S^{\natural}_{\bm{b},0}(n_{0})\cdot n_{1}^{m/2}\cdot\sigma_{\infty,L^{\perp},w}X^{m/2}h(n/Y,0).

By (8.5) and the equality |Λ/n0Λ|=n0m/2\lvert\Lambda^{\perp}/n_{0}\Lambda^{\perp}\rvert=n_{0}^{m/2}, we conclude that

𝒄ΛS𝒄,0(n0)I𝒄(n)=nm/2σ,L,wXm/2h(n/Y,0)𝟏n0=1.\sum_{\bm{c}\in\Lambda^{\perp}}S^{\natural}_{\bm{c},0}(n_{0})\cdot I_{\bm{c}}(n)=n^{m/2}\cdot\sigma_{\infty,L^{\perp},w}X^{m/2}h(n/Y,0)\cdot\bm{1}_{n_{0}=1}.

The left-hand side of (8.17) (see Definition 8.7) thus simplifies to

Y2nY/Pσ,L,wXm/2h(n/Y,0)ϕ(n),Y^{-2}\sum_{n\geq Y/P}\sigma_{\infty,L^{\perp},w}X^{m/2}h(n/Y,0)\cdot\phi(n),

which equals σ,L,wXm/2(1+O(P1))\sigma_{\infty,L^{\perp},w}X^{m/2}(1+O(P^{-1})) by Proposition 8.11 (below). ∎

Proposition 8.11.

nY/Pϕ(n)h(n/Y,0)=Y2(1+O(P1))\sum_{n\geq Y/P}\phi(n)h(n/Y,0)=Y^{2}(1+O(P^{-1})).

Proof.

By [heath1998circle]*final paragraph on p. 692, and second paragraph on p. 676, we have (in terms of the function ω\omega defined on [heath1998circle]*p. 676)

n1ϕ(n)h(n/Y,0)=Yn1ω(n/Y)=Y2(1+OA(YA)).\sum_{n\geq 1}\phi(n)h(n/Y,0)=Y\sum_{n\geq 1}\omega(n/Y)=Y^{2}(1+O_{A}(Y^{-A})).

Furthermore, h(x,0)x1h(x,0)\ll x^{-1} by [heath1996new]*Lemma 4, so

n<Y/Pϕ(n)h(n/Y,0)n<Y/Pϕ(n)Y/nn<Y/PYY2/P.\sum_{n<Y/P}\phi(n)h(n/Y,0)\ll\sum_{n<Y/P}\phi(n)\cdot Y/n\leq\sum_{n<Y/P}Y\ll Y^{2}/P.

Proposition 8.11 follows upon writing nY/P=n1n<Y/P\sum_{n\geq Y/P}=\sum_{n\geq 1}-\sum_{n<Y/P}. ∎

Lemma 8.12.

Uniformly over reals X1X\geq 1, we have

(8.18) ΣY/P(X,2)ϵX(m1)/2+ϵ.\Sigma_{\geq Y/P}(X,\mathcal{E}_{2})\ll_{\epsilon}X^{(m-1)/2+\epsilon}.
Proof.

The main subtlety here is that we must treat 𝒄𝟎\bm{c}\neq\bm{0} and 𝒄=𝟎\bm{c}=\bm{0} separately.

First, given nn, the bound (8.8), applied directly to S𝒄,0(n0)S_{\bm{c},0}(n_{0}) for each n0nn_{0}\mid n, implies

(8.19) n0n1=n:n1Y/P|S𝒄,0(n0)|ϕ(n1)n11/2ϵ𝟏nY/Pn1/2+ϵmaxdnd1/2|S𝒄(d)|.\sum_{\begin{subarray}{c}n_{0}n_{1}=n:\\ n_{1}\geq Y/P\end{subarray}}\,\lvert S^{\natural}_{\bm{c},0}(n_{0})\rvert\cdot\phi(n_{1})n_{1}^{-1/2}\ll_{\epsilon}\bm{1}_{n\geq Y/P}\cdot n^{1/2+\epsilon}\cdot\max_{d\mid n}d^{-1/2}\lvert S^{\natural}_{\bm{c}}(d)\rvert.

Inserting (8.19) into ΣY/P(X,2{𝟎})\Sigma_{\geq Y/P}(X,\mathcal{E}_{2}\setminus\{\bm{0}\}) (see Definition 8.7), and recalling (5.1), we get

ΣY/P(X,2{𝟎})ϵY2𝒄2{𝟎}n1n1m/2+ϵ|I𝒄(n)|maxdnd1/2|S𝒄(d)|ϵYϵf(2),\Sigma_{\geq Y/P}(X,\mathcal{E}_{2}\setminus\{\bm{0}\})\ll_{\epsilon}Y^{-2}\sum_{\bm{c}\in\mathcal{E}_{2}\setminus\{\bm{0}\}}\sum_{n\geq 1}n^{1-m/2+\epsilon}\lvert I_{\bm{c}}(n)\rvert\cdot\max_{d\mid n}d^{-1/2}\lvert S^{\natural}_{\bm{c}}(d)\rvert\ll_{\epsilon}Y^{\epsilon}f(\mathcal{E}_{2}),

where f(2)ϵX(m1)/2+ϵf(\mathcal{E}_{2})\ll_{\epsilon}X^{(m-1)/2+\epsilon} by Proposition 8.2 and Lemma 5.2. Similarly, (8.19) gives

ΣY/P(X,{𝟎})ϵY2nY/Pn1m/2+ϵ|I𝟎(n)|maxdnd1/2|S𝟎(d)|,\Sigma_{\geq Y/P}(X,\{\bm{0}\})\ll_{\epsilon}Y^{-2}\sum_{n\geq Y/P}n^{1-m/2+\epsilon}\lvert I_{\bm{0}}(n)\rvert\cdot\max_{d\mid n}d^{-1/2}\lvert S^{\natural}_{\bm{0}}(d)\rvert,

which is ϵY2Xm+ϵ(Y/P)(4m)/3Xm3+(4m)/3+ϵ\ll_{\epsilon}Y^{-2}X^{m+\epsilon}(Y/P)^{(4-m)/3}\asymp X^{m-3+(4-m)/3+\epsilon} by (6.1) and Lemma 6.1 (summed over N{1,2,4,8,}N\in\{1,2,4,8,\ldots\}). But m3+(4m)/3(m1)/2m-3+(4-m)/3\leq(m-1)/2, since 4m74\leq m\leq 7. ∎

Remark 8.13.

We have used diagonality of FF. In general, under 1.6(1)–(3), one can prove ΣY/P(X,2)ϵXm/21/4+ϵ\Sigma_{\geq Y/P}(X,\mathcal{E}_{2})\ll_{\epsilon}X^{m/2-1/4+\epsilon} (by using Lemma 5.4 in place of Lemma 5.2).

By (8.13), (8.16), (8.17), and (8.18), the quantity (8.12) simplifies to

Oϵ(Xm/21/4+ϵ)+σ,L,wXm/2,O_{\epsilon}(X^{m/2-1/4+\epsilon})+\sigma_{\infty,L^{\perp},w}X^{m/2},

matching the right-hand side of (8.1). So (8.1) holds, thus concluding §8.

Acknowledgements

This paper is an important component of the thesis work described in [wang2022thesis]; many of my acknowledgements there apply here as well. I also thank my advisor, Peter Sarnak, for many helpful suggestions and questions on the exposition, references, assumptions, and scope of (various drafts of) the present work.This work was partially supported by NSF grant DMS-1802211, and the European Union’s Horizon 2020 research and innovation program under the Marie Skłodowska-Curie Grant Agreement No. 101034413. I am also grateful to Trevor Wooley for providing some helpful general comments on special subvarieties and the reference [vaughan1995certain]. I thank Tim Browning for inspiring part of the current title of the paper. Finally, thanks are due to the referee for providing many helpful suggestions.

References