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

Geometric generalizations of the square sieve,
with an application to cyclic covers

Alina Bucur - Department of Mathematics, University of California at San Diego, 9500 Gilman Dr 0112, La Jolla, California 92093, USA alina@math.ucsd.edu Alina Carmen Cojocaru - Department of Mathematics, Statistics and Computer Science, University of Illinois at Chicago, 851 S Morgan St, 322 SEO, Chicago, 60607, IL, USA; - Institute of Mathematics “Simion Stoilow” of the Romanian Academy, 21 Calea Grivitei St, Bucharest, 010702, Sector 1, Romania cojocaru@uic.edu Matilde N. Lalín - Département de Mathématiques et de Statistique, Université de Montréal, CP 6128, Succ. Centre-Ville, Montreal, Quebec H3C 3J7, Canada mlalin@gmail.com  and  Lillian B. Pierce
with an appendix by Joseph Rabinoff
- Department of Mathematics, Duke University, 120 Science Drive, Durham, North Carolina 27708, USA; pierce@math.duke.edu
Abstract.

We formulate a general problem: given projective schemes 𝕐\mathbb{Y} and 𝕏\mathbb{X} over a global field KK and a KK-morphism η\eta from 𝕐\mathbb{Y} to 𝕏\mathbb{X} of finite degree, how many points in 𝕏(K)\mathbb{X}(K) of height at most BB have a pre-image under η\eta in 𝕐(K)\mathbb{Y}(K)? This problem is inspired by a well-known conjecture of Serre on quantitative upper bounds for the number of points of bounded height on an irreducible projective variety defined over a number field. We give a non-trivial answer to the general problem when K=𝔽q(T)K=\mathbb{F}_{q}(T) and 𝕐\mathbb{Y} is a prime degree cyclic cover of 𝕏=Kn\mathbb{X}=\mathbb{P}_{K}^{n}. Our tool is a new geometric sieve, which generalizes the polynomial sieve to a geometric setting over global function fields.

1. Introduction

We consider the following general problem: given a morphism between two projective schemes defined over a global field, how many points in the domain yield points with bounded height in the image? As we will outline, this problem is related to a well-known conjecture of Serre formulated over number fields; in our proposed formulation, it may be viewed as a general version of a question that arises in a wide array of problems in analytic number theory.

To be precise, let KK be a global field of arbitrary characteristic and let 𝒪\mathcal{O} be a ring of integers in KK. We denote by VKV_{K} the set of places of KK. For each vVKv\in V_{K}, we denote by ||v|\cdot|_{v} the associated valuation, normalized such that the product formula holds, namely, for every xKx\in K^{\ast},

vVK|x|v=1.\prod_{v\in V_{K}}|x|_{v}=1. (1.1)

Using the notation

[x]:=[x0:x1::xn]Kn[x]:=[x_{0}:x_{1}:\ldots:x_{n}]\in\mathbb{P}^{n}_{K}

for projective points, we consider the height function

htK:Kn(0,),\operatorname{ht}_{K}:\mathbb{P}^{n}_{K}\rightarrow(0,\infty),
htK([x]):=vVKmax{|x0|v,,|xn|v},\operatorname{ht}_{K}([x]):=\prod_{v\in V_{K}}\max\{|x_{0}|_{v},\dots,|x_{n}|_{v}\}, (1.2)

and note that it gives rise to the height function

htK:K[0,),\operatorname{ht}_{K}:K\rightarrow[0,\infty),
htK(x):={htK([x:1])ifx0,0ifx=0.\operatorname{ht}_{K}(x):=\left\{\begin{array}[]{cc}\operatorname{ht}_{K}([x:1])&\text{if}\ x\neq 0,\\ 0&\text{if}\ x=0.\end{array}\right. (1.3)

General Problem

Given a global field KK of arbitrary characteristic, a ring of integers 𝒪\mathcal{O} in KK, projective schemes 𝕏/K\mathbb{X}/K, 𝕐/K\mathbb{Y}/K over KK with fixed models over 𝒪\mathcal{O}, and a KK-morphism η:𝕐𝕏\eta:\mathbb{Y}\longrightarrow\mathbb{X}, defined over 𝒪\mathcal{O} and of finite degree, find an upper bound for the cardinality of the set

{[x]𝕏(𝒪):htK([x])<B,[y]𝕐(𝒪)such thatη([y])=[x]}\left\{[x]\in\mathbb{X}(\mathcal{O}):\operatorname{ht}_{K}([x])<B,\exists[y]\in\mathbb{Y}(\mathcal{O})\;\text{such that}\;\eta([y])=[x]\right\}

that holds for every B1B\geq 1.

Note that upper bounds for the above cardinality are always given by one of the two cardinalities below,

#{[x]𝕏(K):htK([x])<B}#{[x]Kn:htK([x])<B}.\#\{[x]\in\mathbb{X}(K):\operatorname{ht}_{K}([x])<B\}\leq\#\{[x]\in\mathbb{P}_{K}^{n}:\operatorname{ht}_{K}([x])<B\}.

For example, when KK is a number field of degree dd over \mathbb{Q}, by Schanuel’s theorem (e.g. [Ser97, §2.5 p. 17]), there exists an explicit positive constant C(K,n)C(K,n) such that, as BB\rightarrow\infty,

#{[x]Kn:htK([x])<B}C(K,n)Bd(n+1).\#\left\{[x]\in\mathbb{P}_{K}^{n}:\operatorname{ht}_{K}([x])<B\right\}\sim C(K,n)B^{d(n+1)}. (1.4)

As a second example, when KK is the function field of an absolutely irreducible projective curve over 𝔽q\mathbb{F}_{q}, of genus gg, as an immediate consequence of [Ser97, §2.5, Thm. p. 19], there exists an explicit positive constant C(K,n)C(K,n) such that, as bb\rightarrow\infty,

#{[x]Kn:htK([x])<qb}C(K,n)q(bg)(n+1).\#\left\{[x]\in\mathbb{P}_{K}^{n}:\operatorname{ht}_{K}([x])<q^{b}\right\}\sim C(K,n)q^{(b-g)(n+1)}. (1.5)

As usual when navigating between the number field and the function field settings, the parameter BB in (1.4) was replaced by qbq^{b} in (1.5).

In our general problem, for nontrivial choices of 𝕏,𝕐\mathbb{X},\mathbb{Y}, we seek a nontrivial upper bound, namely a bound that grows more slowly than the trivial bound, as BB\rightarrow\infty (respectively, as qbq^{b}\rightarrow\infty as a function of bb, or of qq, or of both bb and qq).

1.1. Serre’s question

Our General Problem has an antecedent in a well-known question of Serre [Ser97, §13.1 (4) p. 178], which we now recall.

Let K/K/\mathbb{Q} be a number field of degree dd, let n1n\geq 1 be an integer, and let VV be an irreducible (non-linear) projective variety in Kn\mathbb{P}_{K}^{n}. Serre seeks an upper bound in BB for the cardinality of the set

{[x]V(K):htK([x])B}.\left\{[x]\in V(K):\operatorname{ht}_{K}([x])\leq B\right\}. (1.6)

The trivial upper bound for (1.6) is C(K,n)Bd(n+1)C(K,n)B^{d(n+1)}, as mentioned in (1.4). In [Ser97, §13.1, Thm. 4 p. 178], Serre improves upon the trivial bound by showing that there exists a constant 0<γ<10<\gamma<1 such that, for all BB,

#{[x]V(K):htK([x])B}n,K,V(Bd)(n+1)12(logB)γ.\#\left\{[x]\in V(K):\operatorname{ht}_{K}([x])\leq B\right\}\ll_{n,K,V}(B^{d})^{(n+1)-\frac{1}{2}}(\operatorname{log}B)^{\gamma}. (1.7)

Serre deduces (1.7) from a result counting integral points on affine thin sets, which he proves using the large sieve. A variant due to Cohen [Coh81] of the result counting integral points on affine thin sets may also be used; however, Cohen’s result leads to γ=1\gamma=1. Serre then poses the question of whether (1.7) can be improved to

#{[x]V(K):htK([x])B}(Bd)(n+1)1(logB)c\#\left\{[x]\in V(K):\operatorname{ht}_{K}([x])\leq B\right\}\ll(B^{d})^{(n+1)-1}(\operatorname{log}B)^{c} (1.8)

for some c0c\geq 0, without specifying whether the implied \ll-constant might depend on any of n,K,Vn,K,V; see [Ser97, §13.1.3, p. 178]. Additionally, Serre notes that the logarithmic factor is necessarily present in certain cases.

Our General Problem is a generalization of Serre’s question and, as a special case, encompasses a global function field version of Serre’s question (1.8). The specific case of K=𝔽q(T)K=\mathbb{F}_{q}(T) has been studied recently by Browning and Vishe, who proved an analogue of (1.7) by adapting Serre’s argument, using a version of the large sieve inequality over function fields developed by Hsu [Hsu96]; see [BV15, Lemma 2.9] where their result is stated in an affine formulation. In particular, Browning and Vishe commented on the scarcity of results counting points of bounded height on geometrically irreducible (non-linear) varieties in the function field setting [BV15, p. 675]; this paper explores a particular class of such problems.

1.2. Main goals

The purpose of the present paper is to investigate the General Problem in a particular function field setting, and to go beyond the analogue of (1.7) in the case of prime degree cyclic covers. Precisely, our goals are two-fold:

  1. (I)

    to provide a nontrivial upper bound for the General Problem when K=𝔽q(T)K=\mathbb{F}_{q}(T), 𝒪=𝔽q[T]\mathcal{O}=\mathbb{F}_{q}[T], 𝕏=Kn\mathbb{X}=\mathbb{P}^{n}_{K}, 𝕐\mathbb{Y} is a prime degree cyclic cover of 𝕏\mathbb{X}, and η\eta is the natural projection;

  2. (II)

    to accomplish (I) by developing a geometric sieve method which generalizes recent sieve methods (such as the square sieve of Heath-Brown and the polynomial sieve of Browning) that have been used to improve on Serre’s bound (1.7) in the setting over \mathbb{Q}.

We will present our main results in the next two sections, according to the above two goals.

1.3. Main results I: counting rational points

We treat the General Problem in the following concrete case: K=𝔽q(T)K=\mathbb{F}_{q}(T), 𝒪=𝔽q[T]\mathcal{O}=\mathbb{F}_{q}[T], 𝕏=Kn\mathbb{X}=\mathbb{P}^{n}_{K}, 𝕐\mathbb{Y} a prime degree cyclic cover of 𝕏\mathbb{X}, and η\eta the natural projection.

To be precise, let qq be an odd rational prime power, n1n\geq 1 an integer, m2m\geq 2 an integer, and 2\ell\geq 2 a rational prime such that m\ell\mid m. We set

𝕏:=𝔽q(T)n\mathbb{X}:=\mathbb{P}^{n}_{\mathbb{F}_{q}(T)}

and take 𝕐\mathbb{Y} as the projective scheme in the weighted projective space 𝔽q(T)n+1(1,,1,m)\mathbb{P}_{\mathbb{F}_{q}(T)}^{n+1}\left(1,\ldots,1,\frac{m}{\ell}\right) defined by the weighted projective model

𝕐:Xn+1=F(X0,,Xn)\mathbb{Y}:\qquad X_{n+1}^{\ell}=F(X_{0},\ldots,X_{n}) (1.9)

for some polynomial F𝔽q[T][X0,,Xn]F\in\mathbb{F}_{q}[T][X_{0},\ldots,X_{n}] of total degree mm in X0,,XnX_{0},\ldots,X_{n}. We take

η:𝕐𝔽q(T)n\eta:\mathbb{Y}\longrightarrow\mathbb{P}^{n}_{\mathbb{F}_{q}(T)}

as the natural projection defined by

η([x0:x1::xn:xn+1]):=[x0:x1::xn].\eta([x_{0}:x_{1}:\ldots:x_{n}:x_{n+1}]):=[x_{0}:x_{1}:\ldots:x_{n}]. (1.10)

Our interest is in estimating, from above and as a function of qbq^{b} (qq fixed, bb\rightarrow\infty), the counting function

N(𝕐,𝔽q(T),η;b):=#{[x]𝔽q(T)n:ht𝔽q(T)[x]<qb,[y]𝕐(𝔽q(T))such thatη([y])=[x]},N(\mathbb{Y},\mathbb{F}_{q}(T),\eta;b):=\#\left\{[x]\in\mathbb{P}^{n}_{\mathbb{F}_{q}(T)}:\operatorname{ht}_{\mathbb{F}_{q}(T)}[x]<q^{b},\ \exists[y]\in\mathbb{Y}(\mathbb{F}_{q}(T))\;\text{such that}\;\eta([y])=[x]\right\}, (1.11)

or, equivalently, the counting function

#{(x0,,xn)𝔽q[T]n+1:degT(xi)<b0in,xn+1𝔽q[T]such thatxn+1=F(x0,,xn)},\#\left\{(x_{0},\ldots,x_{n})\in\mathbb{F}_{q}[T]^{n+1}:\;\operatorname{deg}_{T}(x_{i})<b\;\forall 0\leq i\leq n,\exists x_{n+1}\in\mathbb{F}_{q}[T]\;\text{such that}\;x_{n+1}^{\ell}=F(x_{0},\ldots,x_{n})\right\},

where degT(xi)\operatorname{deg}_{T}(x_{i}) denotes the degree of xix_{i} as a polynomial in TT.

Note that, in this setting, the trivial bound is

N(𝕐,𝔽q(T),η;b)#{[x]𝔽q(T)n:ht𝔽q(T)([x])<qb}(qb)(n+1)N(\mathbb{Y},\mathbb{F}_{q}(T),\eta;b)\leq\#\left\{[x]\in\mathbb{P}^{n}_{\mathbb{F}_{q}(T)}:\operatorname{ht}_{\mathbb{F}_{q}(T)}([x])<q^{b}\right\}\leq\left(q^{b}\right)^{(n+1)} (1.12)

(see (1.5) and the comment in equation (5.1) of §5).

In contrast, the function field analogue of Serre’s conjecture (1.8) suggests that it might be possible to prove, under appropriate conditions on FF, that there exists some constant cc for which

N(𝕐,𝔽q(T),η;b),m,n,q,F(qb)(n+1)1bc,N(\mathbb{Y},\mathbb{F}_{q}(T),\eta;b)\ll_{\ell,m,n,q,F}\left(q^{b}\right)^{(n+1)-1}b^{c}, (1.13)

with the implicit constant possibly depending on ,m,n,q,F\ell,m,n,q,F.

If FF of degree m2m\geq 2 is such that F(X0,,Xn)=0F(X_{0},\ldots,X_{n})=0 defines a nonsingular projective hypersurface, Browning and Vishe’s work in [BV15] implies that

N(𝕐,𝔽q(T),η;b),m,n(qb)(n+1)1/2blogq.N(\mathbb{Y},\mathbb{F}_{q}(T),\eta;b)\ll_{\ell,m,n}\left(q^{b}\right)^{(n+1)-1/2}b\operatorname{log}q. (1.14)

This establishes a benchmark of roughly analogous strength to (1.7), and is the first improvement of the trivial bound (1.12). This can be derived by applying [BV15, Lemma 2.9] to count solutions on the affine model Xn+1=F(X0,,Xn)X_{n+1}^{\ell}=F(X_{0},\ldots,X_{n}) which is irreducible under the condition on FF; see (3.1) to interpret the height function when applying this result. (While we focus exclusively on cyclic covers, we remark that Browning and Vishe’s work applies more generally; see [BV15, p. 674] and [BV15, Lemmas 2.9, 2.10] for more general results counting points of bounded height on absolutely irreducible (non-linear) varieties in affine and projective settings, of equivalent strength to (1.7).)

Our first main theorem improves upon the trivial bound (1.12) as well as (1.14), and approaches, in the limit as nn\rightarrow\infty (upon omitting an analysis of how the limit impacts the dependence of the \ll-constant on nn), the upper bound (qb)(n+1)1\left(q^{b}\right)^{(n+1)-1} appearing in (1.13), as long as the defining polynomial FF is such that F(X0,,Xn)=0F(X_{0},\ldots,X_{n})=0 defines a nonsingular projective hypersurface.

Our first main theorem is:

Theorem 1.1 (Counting Rational Points on a Prime Degree Cyclic Cover of 𝔽q(T)n\mathbb{P}^{n}_{\mathbb{F}_{q}(T)}).

Let qq be an odd rational prime power, n2n\geq 2 an integer, 2\ell\geq 2 a rational prime, and F𝔽q[T][X0,,Xn]F\in\mathbb{F}_{q}[T][X_{0},\ldots,X_{n}] a homogeneous polynomial of degree m2m\geq 2 in X0,,XnX_{0},\ldots,X_{n}, with char𝔽qm\operatorname{char}\mathbb{F}_{q}\nmid m. Assume that:

  1. (i)

    gcd(m,q1)\ell\mid\operatorname{gcd}(m,q-1);

  2. (ii)

    F(X0,,Xn)=0F(X_{0},\ldots,X_{n})=0 defines a nonsingular projective hypersurface in 𝔽q(T)¯n\mathbb{P}^{n}_{\overline{\mathbb{F}_{q}(T)}}.

Let 𝕐\mathbb{Y} be the projective scheme in the weighted projective space 𝔽q(T)n+1(1,,1,m)\mathbb{P}_{\mathbb{F}_{q}(T)}^{n+1}\left(1,\ldots,1,\frac{m}{\ell}\right) defined by the weighted projective model (1.9). Let η\eta be the projection (1.10). Then for all b1b\geq 1 the quantity N(𝕏,𝕐,𝔽q(T),η;b)N(\mathbb{X},\mathbb{Y},\mathbb{F}_{q}(T),\eta;b) defined in (1.11) satisfies the bound

N(𝕐,𝔽q(T),η;b),m,n,q,F(qb)(n+1)n+1n+2bn+1n+2,N(\mathbb{Y},\mathbb{F}_{q}(T),\eta;b)\ll_{\ell,m,n,q,F}\;\left(q^{b}\right)^{(n+1)-\frac{n+1}{n+2}}b^{\frac{n+1}{n+2}},

where the implicit constant depends on ,m,n,q\ell,m,n,q, and FF.

Later on in Theorem 5.1, we will use the function mentioned in the displayed equation below (1.11) to state a version of Theorem 1.1 in terms of counting perfect \ell-th power values of a homogeneous polynomial F𝔽q[T][X0,,Xn]F\in\mathbb{F}_{q}[T][X_{0},\ldots,X_{n}]. For more information on the way in which the implicit constant depends on FF, see §9.

To put Theorem 1.1 in context, let us recall the current state of knowledge toward Serre’s conjecture (1.8) when K=K=\mathbb{Q}. For n=1,2n=1,2, Broberg [Bro09] proved a weak form of Serre’s conjecture (with BϵB^{\epsilon} in place of a logarithmic factor) via the determinant method. For n3n\geq 3 and in the case of cyclic covers of degree \ell, the power sieve argument presented by Munshi in [Mun09] leads to the upper bound

,m,n,FB(n+1)nn+1(logB)nn+1,\ll_{\ell,m,n,F}B^{(n+1)-\frac{n}{n+1}}(\operatorname{log}B)^{\frac{n}{n+1}}, (1.15)

where FF is the defining polynomial of the cover and mm is its degree. Recently, Bonolis [Bon21] refined the argument given in [Mun09] and obtained the upper bound

,m,n,FB(n+1)n+1n+2(logB)n+1n+2.\ll_{\ell,m,n,F}B^{(n+1)-\frac{n+1}{n+2}}(\operatorname{log}B)^{\frac{n+1}{n+2}}. (1.16)

(For clarity, we remark that Theorem 1.1 of [Mun09] states a bound of the strength (1.16), but the argument as written therein proves a result of the strength (1.15). At the suggestion of Munshi, Bonolis [Bon21] implemented nontrivial averaging in the relevant sieve inequality in order to prove (1.16) (as well as a more general result over \mathbb{Q}).) Our result in Theorem 1.1 is thus an analogue over 𝔽q(T)\mathbb{F}_{q}(T) of the result (1.16) over \mathbb{Q}. Note that, in the limit nn\rightarrow\infty and aside from an analysis of how the limit impacts the dependence of the \ll-constant on nn, the upper bound (1.15) or (1.16) approaches one of the form conjectured in (1.8).

Over \mathbb{Q}, the best known result is due to Heath-Brown and the third author [HBP12], who proved Serre’s conjecture (1.8) for all n9n\geq 9 in the case of cyclic covers, by combining a sieve method with the qq-analogue of Van der Corput’s method. It would be interesting to to adapt Heath-Brown and Pierce’s qq-analogue of Van der Corput method to the function field setting of Theorem 1.1.

1.4. Main results II: geometric sieve inequalities

Our approach to the General Problem proceeds via a sieve, formulated in the setting of the General Problem, with KK a global field of arbitrary characteristic, 𝒪\mathcal{O} a ring of integers in KK, htK:K[0,)\operatorname{ht}_{K}:K\longrightarrow[0,\infty) the height function (1.3) constructed from valuations that satisfy the product formula (1.1), 𝕏/K\mathbb{X}/K and 𝕐/K\mathbb{Y}/K projective schemes over KK with fixed models over 𝒪{\mathcal{O}}, and η:𝕐𝕏\eta:\mathbb{Y}\longrightarrow\mathbb{X} a KK-morphism, defined over 𝒪{\mathcal{O}} and of finite degree.

As it will not result in any significant loss in sharpness of the results, we will, for convenience, count points on 𝕏\mathbb{X} in the affine sense. To clarify, using the notation

x¯:=(x0,x1,,xn)𝔸Kn+1\underline{x}:=(x_{0},x_{1},\ldots,x_{n})\in\mathbb{A}^{n+1}_{K}

for affine points, we will work with the height function in the affine space 𝔸Kn+1\mathbb{A}_{K}^{n+1} given by

htK(x¯):=max{htK(xi):x¯=(x0,,xn)},\operatorname{ht}_{K}(\underline{x}):=\max\left\{\operatorname{ht}_{K}(x_{i}):\underline{x}=(x_{0},\ldots,x_{n})\right\}, (1.17)

focus on the set

𝒜=𝒜(B):={x¯𝕏(𝒪):htK(x¯)<B},{\mathcal{A}}=\mathcal{A}(B):=\left\{\underline{x}\in\mathbb{X}({\mathcal{O}}):\operatorname{ht}_{K}(\underline{x})<B\right\}, (1.18)

and seek an upper bound, in terms of BB, for the cardinality of the set

𝒮(𝒜)=𝒮(𝒜(B)):={x¯𝒜:y¯𝕐(𝒪)such thatη(y¯)=x¯}.{\mathcal{S}}({\mathcal{A}})={\mathcal{S}}({\mathcal{A}(B)}):=\left\{\underline{x}\in{\mathcal{A}}:\;\exists\underline{y}\in\mathbb{Y}({\mathcal{O}})\;\text{such that}\;\eta(\underline{y})=\underline{x}\right\}. (1.19)

The typical sieve approach is to derive information about 𝒮(𝒜)\mathcal{S}({\mathcal{A}}) from reductions of η\eta modulo primes. Towards this goal, for each finite place vVKv\in V_{K}, we denote by (𝒪v,Mv)({\mathcal{O}}_{v},M_{v}) the associated discrete valuation ring and by kv:=𝒪v/Mvk_{v}:={\mathcal{O}}_{v}/M_{v} the associated residue field. For all but finitely many finite places vv (in which case we refer to vv as a place of good reduction for η\eta), we denote by ηv:𝕐𝕏\eta_{v}:\mathbb{Y}\rightarrow\mathbb{X} the reduction of η\eta modulo vv. We call ηv\eta_{v} ramified at some x¯𝕏(𝒪v)\underline{x}\in\mathbb{X}({\mathcal{O}}_{v}) if x¯(modv)𝕏(kv)\underline{x}\,(\operatorname{mod}v)\in\mathbb{X}(k_{v}) is a branch point for the function ηv:𝕐𝕏\eta_{v}:\mathbb{Y}\rightarrow\mathbb{X}. For each x¯𝕏(𝒪)\underline{x}\in\mathbb{X}({\mathcal{O}}), we define the set

VKram(x¯,η):={vVK:vof good reduction forηandηvis ramified at x¯(modv)}.V_{K}^{\mathrm{ram}}(\underline{x},\eta):=\left\{v\in V_{K}:v\ \text{of good reduction for}\ \eta\ \text{and}\ \eta_{v}\;\text{is ramified at $\underline{x}\,(\operatorname{mod}v)$}\right\}.

Moreover, for any nonempty finite set 𝒫VK\mathcal{P}\subseteq V_{K} of finite places of good reduction for η\eta, we define the subset

V𝒫ram(x¯):=VKram(x¯,η)𝒫.V_{\mathcal{P}}^{\mathrm{ram}}(\underline{x}):=V_{K}^{\mathrm{ram}}(\underline{x},\eta)\cap{\mathcal{P}}. (1.20)

With this notation, we prove the following sieve inequality:

Theorem 1.2.

(Geometric Sieve)
Let KK be a global field of arbitrary characteristic, 𝒪\mathcal{O} be the ring of integers in KK, and htK:K[0,)\operatorname{ht}_{K}:K\longrightarrow[0,\infty) the height function (1.3) constructed from valuations that satisfy the product formula (1.1). Let 𝕏/K\mathbb{X}/K, 𝕐/K\mathbb{Y}/K be projective schemes over KK with fixed models over 𝒪\mathcal{O} and let η:𝕐𝕏\eta:\mathbb{Y}\longrightarrow\mathbb{X} be a KK-morphism, defined over 𝒪\mathcal{O} and of finite degree. For an arbitrary B>0B>0, define the sets 𝒜\mathcal{A} and 𝒮(𝒜)\mathcal{S}(\mathcal{A}) as in (1.18), (1.19). Then, for any real number α1\alpha\geq 1 and for any nonempty finite set 𝒫VK{\mathcal{P}}\subseteq V_{K} of finite places of good reduction for η\eta,

|𝒮(𝒜)|2|𝒫|x¯𝒜|V𝒫ram(x¯)|+1|𝒫|2x¯𝒜Iα(x¯)2,\left|{\mathcal{S}}({\mathcal{A}})\right|\leq\frac{2}{|{\mathcal{P}}|}\displaystyle\sum_{\underline{x}\in{\mathcal{A}}}|V_{\mathcal{P}}^{\mathrm{ram}}(\underline{x})|+\frac{1}{|{\mathcal{P}}|^{2}}\displaystyle\sum_{\underline{x}\in{\mathcal{A}}}I_{\alpha}(\underline{x})^{2}, (1.21)

where, for any x¯𝕏(𝒪)\underline{x}\in\mathbb{X}({\mathcal{O}}),

Iα(x¯):=v𝒫\V𝒫ram(x¯)(α+(|ηv1(x¯(modv))|1)(degη|ηv1(x¯(modv))|)).I_{\alpha}(\underline{x}):=\displaystyle\sum_{v\in{\mathcal{P}}\backslash V_{\mathcal{P}}^{\mathrm{ram}}(\underline{x})}\left(\alpha+\left(\left|\eta_{v}^{-1}(\underline{x}\,(\operatorname{mod}v))\right|-1\right)\cdot\left(\operatorname{deg}\eta-\left|\eta_{v}^{-1}(\underline{x}\,(\operatorname{mod}v))\right|\right)\right).

Impetus for inequality (1.21) comes from a sequence of papers on sieve inequalities, beginning with the square sieve over \mathbb{Q} of [HB84] (itself inspired by [Hoo78]), which was later developed into the power sieve over \mathbb{Q} in [Mun09] and [Bra15], and into the square sieve over 𝔽q(T)\mathbb{F}_{q}(T) in [CD08]. The square and power sieves over \mathbb{Q} were strengthened by being combined with the qq-analogue of Van der Corput’s method in [Pie06] and [HBP12]. Most recently, Browning [Bro15] expanded the square sieve over \mathbb{Q} into a polynomial sieve over \mathbb{Q}, while Bonolis [Bon21] developed a related polynomial sieve over \mathbb{Q} involving expansions via trace functions. It is Browning’s work [Bro15] that motivates our approach towards generalizing the existing versions of the square sieve (over \mathbb{Q} and over 𝔽q(T)\mathbb{F}_{q}(T)) to a geometric sieve over global fields. We will make an explicit comparison to the square sieve, polynomial sieve, and other relatives in §2.4.

Since we will deduce Theorem 1.1 from the above sieve inequality, we now state the relevant consequence of Theorem 1.2 in the case of cyclic covers of prime degree that we derive by optimizing the choice of α\alpha.

Theorem 1.3.

(Geometric Sieve for Prime Degree Cyclic Covers of Kn\mathbb{P}^{n}_{K})

Let KK be a global field of arbitrary characteristic, let 𝒪\mathcal{O} be the ring of integers in KK, and let htK:K[0,)\operatorname{ht}_{K}:K\longrightarrow[0,\infty) be the height function (1.3) constructed from valuations that satisfy the product formula (1.1). Let n,m1n,m\geq 1 be integers, take 𝕏:=Kn\mathbb{X}:=\mathbb{P}^{n}_{K}, and let 𝕐\mathbb{Y} be the projective scheme in the weighted projective space Kn+1(1,,1,m)\mathbb{P}_{K}^{n+1}\left(1,\dots,1,\frac{m}{\ell}\right), defined by a model

Xn+1=F(X0,,Xn)X_{n+1}^{\ell}=F(X_{0},\ldots,X_{n}) (1.22)

with \ell prime such that m\ell\mid m, and with F𝒪[X0,,Xn]F\in{\mathcal{O}}[X_{0},\ldots,X_{n}] homogeneous of degree mm and having the property that the hypersurface in K¯n\mathbb{P}^{n}_{\overline{K}} defined by

F(X0,,Xn)=0F(X_{0},\ldots,X_{n})=0

is nonsingular. Let η:𝕐𝕏\eta:\mathbb{Y}\longrightarrow\mathbb{X} be the cyclic map of degree \ell defined over 𝔸K¯n\mathbb{A}_{\overline{K}}^{n} by

η(x0,x1,,xn,xn+1):=(x0,x1,,xn).\eta(x_{0},x_{1},\ldots,x_{n},x_{n+1}):=(x_{0},x_{1},\ldots,x_{n}). (1.23)

For an arbitrary B>0B>0, define 𝒜{\mathcal{A}} and 𝒮(𝒜){\mathcal{S}}({\mathcal{A}}) as in (1.18), (1.19). Then, for any nonempty finite set 𝒫VK{\mathcal{P}}\subseteq V_{K} of finite places of good reduction for η\eta,

|𝒮(𝒜)|\displaystyle\left|{\mathcal{S}}({\mathcal{A}})\right|\leq |𝒜||𝒫|(degη1)2+2|𝒫|x¯𝒜|V𝒫ram(x¯)|\displaystyle\frac{\left|{\mathcal{A}}\right|}{|{\mathcal{P}}|}(\operatorname{deg}\eta-1)^{2}+\frac{2}{|{\mathcal{P}}|}\displaystyle\sum_{\underline{x}\in{\mathcal{A}}}|V_{\mathcal{P}}^{\mathrm{ram}}(\underline{x})|
+1|𝒫|2v1,v2𝒫v1v2|x¯𝒜x¯𝕏𝒪ram(v1)𝕏𝒪ram(v2)(|ηv11(x¯(modv1))|1)(|ηv21(x¯(modv2))|1)|,\displaystyle+\frac{1}{\left|{\mathcal{P}}\right|^{2}}\displaystyle\sum_{\begin{subarray}{c}v_{1},v_{2}\in{\mathcal{P}}\\ v_{1}\neq v_{2}\end{subarray}}\left|\displaystyle\sum_{\begin{subarray}{c}\underline{x}\in{\mathcal{A}}\\ \underline{x}\not\in\mathbb{X}_{\mathcal{O}}^{\mathrm{ram}}(v_{1})\cup\mathbb{X}_{\mathcal{O}}^{\mathrm{ram}}(v_{2})\end{subarray}}\left(\left|\eta_{v_{1}}^{-1}(\underline{x}\,(\operatorname{mod}v_{1}))\right|-1\right)\cdot\left(\left|\eta_{v_{2}}^{-1}(\underline{x}\,(\operatorname{mod}v_{2}))\right|-1\right)\right|, (1.24)

where, for each finite place vVKv\in V_{K} of good reduction for η\eta,

𝕏𝒪ram(v):={x¯𝕏(𝒪):ηvis ramified atx¯(modv)}.\mathbb{X}_{{\mathcal{O}}}^{\mathrm{ram}}(v):=\left\{\underline{x}\in\mathbb{X}({\mathcal{O}}):\eta_{v}\;\text{is ramified at}\;\underline{x}\,(\operatorname{mod}v)\right\}. (1.25)

On the right-hand side of inequality (1.24), the first term may be regarded as a main term for which a trivial upper bound suffices. We refer to the second term as the ramified sieve term and remark that it is similar in size to the main term. We refer to the third term as the unramified sieve term and remark that, in applications, the primary difficulty is to bound it nontrivially, and then to choose the sieving set 𝒫\mathcal{P} appropriately to balance the third term’s contribution with that of the first term.

1.5. Outline of the paper

In §2, we prove Theorem 1.2 and Theorem 1.3. To prove Theorem 1.3, the essential point is to compute the optimal choice of α\alpha for which to apply Theorem 1.2. The rest of the paper focuses on proving Theorem 1.1, as follows. In §3, we recall notation and basic results related to the function field setting of Theorem 1.1. In §4, we present results on duals and reductions necessary in our analysis of the unramified sieve term. In §5, we reformulate Theorem 1.1 as an affine statement (Theorem 5.1) and bound all but the unramified sieve term in the inequality (1.24). (Here we apply a simple Schwartz-Zippel counting bound, for which we provide a proof in §10, for completeness.) The remainder of the work is focused on treating the unramified sieve term. In §6, we introduce background material on Fourier analysis on function fields and prove that the unramified sieve term can be stated as a mixed character sum. In §7, we state and verify the Weil-Deligne bounds we require for the unramified sieve term. In §8, we finally bound the unramified sieve term. In §9, we optimize the choice of the sieving set, thus completing the proof of Theorem 1.1. Finally, Appendix §11, written by Joseph Rabinoff, provides a self-contained account of certain standard facts about duals and reductions employed in §4.

Acknowledgements

We thank Dante Bonolis, Mihran Papikian, Joseph Rabinoff, and Melanie Wood for helpful conversations related to parts of this work. In addition, we thank the anonymous referee for exceptionally helpful recommendations with two significant effects. First, these recommendations simplified and strengthened the proof of Proposition 7.1 (i), so that our initial restriction that FF in Theorem 1.1 be Dwork-regular could be weakened to assume that F=0F=0 is a nonsingular projective hypersurface. Second, these recommendations suggested the strategy to prove Lemma 8.2; this enabled averaging over prime elements in the sieving set, hence upgrading the main theorem from a result analogous (1.15) to the present result. We thank the referee for these astute and generous recommendations.

2. Derivation of the fundamental sieve inequalities

2.1. Proof of Theorem 1.2

We begin by proving the most general version of the sieve, as stated in Theorem 1.2. The proof follows the general framework of [HB84] and [Bro15].

For any fixed real number α1\alpha\geq 1, we consider the sum

Σ:=x¯𝒜(v𝒫\V𝒫ram(x¯)(α+(|ηv1(x¯(modv))|1)(degη|ηv1(x¯(modv))|)))2\Sigma:=\displaystyle\sum_{\underline{x}\in{\mathcal{A}}}\left(\displaystyle\sum_{v\in{\mathcal{P}}\backslash V_{\mathcal{P}}^{\mathrm{ram}}(\underline{x})}\left(\alpha+\left(|\eta_{v}^{-1}(\underline{x}(\bmod v))|-1\right)\cdot\left(\operatorname{deg}\eta-|\eta_{v}^{-1}(\underline{x}(\bmod v))|\right)\right)\right)^{2}

and note that each x¯\underline{x} therein is added with a non-negative weight. Moreover, observe that, if x¯𝒮(𝒜)\underline{x}\in\mathcal{S}(\mathcal{A}), then the fiber above x¯\underline{x} has at least one element and at most degη\operatorname{deg}\eta elements, that is,

1|η1(x¯)|degη.1\leq|\eta^{-1}(\underline{x})|\leq\operatorname{deg}\eta.

Therefore, in this case, for each v𝒫\V𝒫ram(x¯)v\in{\mathcal{P}}\backslash V_{\mathcal{P}}^{\mathrm{ram}}(\underline{x}), we have

α+(|ηv1(x¯(modv))|1)(degη|ηv1(x¯(modv))|)1.\alpha+\left(|\eta_{v}^{-1}(\underline{x}(\bmod v))|-1\right)\cdot\left(\operatorname{deg}\eta-|\eta_{v}^{-1}(\underline{x}(\bmod v))|\right)\geq 1.

We deduce that, for every x¯𝒮(𝒜)\underline{x}\in\mathcal{S}(\mathcal{A}),

v𝒫\V𝒫ram(x¯)\displaystyle\displaystyle\sum_{v\in{\mathcal{P}}\backslash V_{\mathcal{P}}^{\mathrm{ram}}(\underline{x})} (α+(|ηv1(x¯(modv))|1)(degη|ηv1(x¯(modv))|))\displaystyle\left(\alpha+\left(|\eta_{v}^{-1}(\underline{x}(\bmod v))|-1\right)\cdot\left(\operatorname{deg}\eta-|\eta_{v}^{-1}(\underline{x}(\bmod v))|\right)\right) (2.1)
v𝒫\V𝒫ram(x¯)1=|𝒫||V𝒫ram(x¯)|,\displaystyle\geq\sum_{v\in{\mathcal{P}}\backslash V_{\mathcal{P}}^{\mathrm{ram}}(\underline{x})}1=|\mathcal{P}|-|V_{\mathcal{P}}^{\mathrm{ram}}(\underline{x})|, (2.2)

which implies that

Σ\displaystyle\Sigma \displaystyle\geq x¯𝒮(𝒜)(|𝒫||V𝒫ram(x¯)|)2\displaystyle\displaystyle\sum_{\underline{x}\in\mathcal{S}(\mathcal{A})}(|\mathcal{P}|-|V_{\mathcal{P}}^{\mathrm{ram}}(\underline{x})|)^{2} (2.3)
=\displaystyle= |𝒫|2|𝒮(𝒜)|+x¯𝒮(𝒜)(2|𝒫||V𝒫ram(x¯)|+|V𝒫ram(x¯)|2)\displaystyle|\mathcal{P}|^{2}\cdot|\mathcal{S}(\mathcal{A})|+\sum_{\underline{x}\in\mathcal{S}(\mathcal{A})}(-2|\mathcal{P}|\cdot|V_{\mathcal{P}}^{\mathrm{ram}}(\underline{x})|+|V_{\mathcal{P}}^{\mathrm{ram}}(\underline{x})|^{2})
\displaystyle\geq |𝒫|2|𝒮(𝒜)|2x¯𝒮(𝒜)|𝒫||V𝒫ram(x¯)|.\displaystyle|\mathcal{P}|^{2}\cdot|\mathcal{S}(\mathcal{A})|-2\sum_{\underline{x}\in\mathcal{S}(\mathcal{A})}|\mathcal{P}|\cdot|V_{\mathcal{P}}^{\mathrm{ram}}(\underline{x})|.

Rearranging the terms and using the non-negativity of |V𝒫ram(x¯)||V_{\mathcal{P}}^{\mathrm{ram}}(\underline{x})| to enlarge the sum over 𝒮(𝒜)\mathcal{S}(\mathcal{A}) to 𝒜\mathcal{A}, we deduce that

|𝒮(𝒜)||𝒫|2Σ+2|𝒫|1x¯𝒜|V𝒫ram(x¯)|,|\mathcal{S}(\mathcal{A})|\leq|\mathcal{P}|^{-2}\Sigma+2|\mathcal{P}|^{-1}\sum_{\underline{x}\in\mathcal{A}}|V_{\mathcal{P}}^{\mathrm{ram}}(\underline{x})|,

which completes the proof of (1.21).

2.2. Equivalent formulation of Theorem 1.2

Recalling the notation 𝕏𝒪ram(v)\mathbb{X}_{{\mathcal{O}}}^{\mathrm{ram}}(v) in (1.25), we see that the bound for |𝒮(𝒜)||\mathcal{S}(\mathcal{A})| may be rewritten by expanding the square inside Σ\Sigma. This shows that

|𝒮(𝒜)|1|𝒫|2v1,v2𝒫|i,j{0,1,2}ci,j(α)Si,j(v1,v2)|+2|𝒫|x¯𝒜|V𝒫ram(x¯)|,\left|{\mathcal{S}}({\mathcal{A}})\right|\leq\frac{1}{\left|{\mathcal{P}}\right|^{2}}\displaystyle\sum_{v_{1},v_{2}\in{\mathcal{P}}}\left|\displaystyle\sum_{i,j\in\{0,1,2\}}c_{i,j}(\alpha)S_{i,j}(v_{1},v_{2})\right|+\frac{2}{\left|{\mathcal{P}}\right|}\displaystyle\sum_{\underline{x}\in{\mathcal{A}}}|V_{\mathcal{P}}^{\mathrm{ram}}(\underline{x})|, (2.4)

in which, for all v1,v2𝒫v_{1},v_{2}\in{\mathcal{P}} and i,j{0,1,2}i,j\in\{0,1,2\},

Si,j(v1,v2):=x¯𝒜x¯𝕏𝒪ram(v1)𝕏𝒪ram(v2)|ηv11(x¯(modv1))|i|ηv21(x¯(modv2))|jS_{i,j}(v_{1},v_{2}):=\displaystyle\sum_{\begin{subarray}{c}\underline{x}\in{\mathcal{A}}\\ \underline{x}\not\in\mathbb{X}_{\mathcal{O}}^{\mathrm{ram}}(v_{1})\cup\mathbb{X}_{\mathcal{O}}^{\mathrm{ram}}(v_{2})\end{subarray}}\left|\eta_{v_{1}}^{-1}(\underline{x}\,(\operatorname{mod}v_{1}))\right|^{i}\cdot\left|\eta_{v_{2}}^{-1}(\underline{x}\,(\operatorname{mod}v_{2}))\right|^{j} (2.5)

and

ci,j(α):={(αdegη)2if (i,j)=(0,0),(αdegη)(1+degη)if (i,j)=(1,0) or (0,1),(1+degη)2if (i,j)=(1,1),(αdegη)if (i,j)=(2,0) or (0,2),(1+degη)if (i,j)=(2,1) or (1,2),1if (i,j)=(2,2).c_{i,j}(\alpha):=\left\{\begin{array}[]{cc}(\alpha-\operatorname{deg}\eta)^{2}&\text{if $(i,j)=(0,0)$,}\\ (\alpha-\operatorname{deg}\eta)(1+\operatorname{deg}\eta)&\text{if $(i,j)=(1,0)$ or $(0,1)$,}\\ (1+\operatorname{deg}\eta)^{2}&\text{if $(i,j)=(1,1)$,}\\ -(\alpha-\operatorname{deg}\eta)&\text{if $(i,j)=(2,0)$ or $(0,2)$,}\\ -(1+\operatorname{deg}\eta)&\text{if $(i,j)=(2,1)$ or $(1,2)$,}\\ 1&\text{if $(i,j)=(2,2)$.}\end{array}\right.

In this formulation, we see that (2.4) is analogous to the polynomial sieve in [Bro15, Thm. 1.1], except that our formulation omits the weight that is typically present in previous sieve inequalities of this type (see §2.4).

2.3. Proof of Theorem 1.3: optimizing the choice of α\alpha

The sieve inequality in Theorem 1.3 follows from inequality (1.21) when specialized to the case of prime degree cyclic covers of projective space, once we have computed the optimal choice of α\alpha to minimize the first term on the right-hand side of (1.21).

To compute this choice, recall that, since η\eta is a cyclic map of prime degree \ell, for each vVKv\in V_{K} of good reduction for η\eta and for each x¯𝕏(𝒪)\underline{x}\in\mathbb{X}(\mathcal{O}),

|ηv1(x¯(modv))|={0if x¯(modv)Imηv,1if x¯(modv)Imηvx¯𝕏𝒪ram(v),if x¯(modv)Imηvx¯𝕏𝒪ram(v).\left|\eta_{v}^{-1}(\underline{x}\,(\operatorname{mod}v))\right|=\left\{\begin{array}[]{cc}0&\text{if $\underline{x}\,(\operatorname{mod}v)\not\in\operatorname{Im}\eta_{v}$,}\\ 1&\text{if $\underline{x}\,(\operatorname{mod}v)\in\operatorname{Im}\eta_{v}$, $\underline{x}\in\mathbb{X}_{\mathcal{O}}^{\text{ram}}(v)$,}\\ \ell&\text{if $\underline{x}\,(\operatorname{mod}v)\in\operatorname{Im}\eta_{v}$, $\underline{x}\not\in\mathbb{X}_{\mathcal{O}}^{\text{ram}}(v)$.}\end{array}\right.

Note that in the above we used the primality of degη\operatorname{deg}\eta.

We deduce that, for each vVKv\in V_{K} of good reduction for η\eta and for each x¯𝕏(𝒪)\underline{x}\in\mathbb{X}(\mathcal{O}),

(|ηv1(x¯(modv))|1)2={1if x¯(modv)Imηv,0if x¯(modv)Imηvx¯𝕏𝒪ram(v),(1)2if x¯(modv)Imηvx¯𝕏𝒪ram(v).\left(\left|\eta_{v}^{-1}(\underline{x}\,(\operatorname{mod}v))\right|-1\right)^{2}=\left\{\begin{array}[]{cc}1&\text{if $\underline{x}\,(\operatorname{mod}v)\not\in\operatorname{Im}\eta_{v}$,}\\ 0&\text{if $\underline{x}\,(\operatorname{mod}v)\in\operatorname{Im}\eta_{v}$, $\underline{x}\in\mathbb{X}_{\mathcal{O}}^{\text{ram}}(v)$,}\\ (\ell-1)^{2}&\text{if $\underline{x}\,(\operatorname{mod}v)\in\operatorname{Im}\eta_{v}$, $\underline{x}\not\in\mathbb{X}_{\mathcal{O}}^{\text{ram}}(v)$.}\end{array}\right. (2.6)

As such, for each vVKv\in V_{K} of good reduction for η\eta and for each x¯𝕏(𝒪)\𝕏𝒪ram(v),\underline{x}\in\mathbb{X}(\mathcal{O})\backslash\mathbb{X}_{\mathcal{O}}^{\text{ram}}(v),

(|ηv1(x¯(modv))|1)2=(1)+(2)(|ηv1(x¯(modv))|1).\displaystyle\left(\left|\eta_{v}^{-1}(\underline{x}\,(\operatorname{mod}v))\right|-1\right)^{2}=(\ell-1)+(\ell-2)\left(\left|\eta_{v}^{-1}(\underline{x}\,(\operatorname{mod}v))\right|-1\right). (2.7)

Noting that x¯𝒮(𝒜)\underline{x}\in{\mathcal{S}}({\mathcal{A}}) implies x¯(modv)Imηv\underline{x}\,(\operatorname{mod}v)\in\operatorname{Im}\eta_{v}, we infer that, for each vVKv\in V_{K} of good reduction for η\eta and for each x¯𝒮(𝒜)\underline{x}\in{\mathcal{S}}({\mathcal{A}}), we have

|ηv1(x¯(modv))|1.\left|\eta_{v}^{-1}(\underline{x}\,(\operatorname{mod}v))\right|\geq 1. (2.8)

Now, recall the notation of the model (1.22) for the prime degree cyclic cover 𝕐\mathbb{Y}. Motivated by our upcoming expressions in terms of character sums in §6, for each vVKv\in V_{K} of good reduction for η\eta and for each x¯𝕏(𝒪)\underline{x}\in\mathbb{X}(\mathcal{O}), we set

Ψv(F(x¯)):=|ηv1(x¯(modv))|1.\Psi_{v}(F(\underline{x})):=\left|\eta_{v}^{-1}(\underline{x}\,(\operatorname{mod}v))\right|-1. (2.9)

Then (2.7) implies that, for each vVKv\in V_{K} of good reduction for η\eta and for each x¯𝕏(𝒪)\𝕏𝒪ram(v)\underline{x}\in\mathbb{X}(\mathcal{O})\backslash\mathbb{X}_{\mathcal{O}}^{\text{ram}}(v),

Ψv(F(x¯))2=(1)+(2)Ψv(F(x¯)),\displaystyle\Psi_{v}(F(\underline{x}))^{2}=(\ell-1)+(\ell-2)\Psi_{v}(F(\underline{x})), (2.10)

while (2.8) implies that, for each vVKv\in V_{K} of good reduction for η\eta and for each x¯𝒮(𝒜)\underline{x}\in{\mathcal{S}}({\mathcal{A}}),

Ψv(F(x¯))0.\Psi_{v}(F(\underline{x}))\geq 0.

We apply these observations in inequality (1.21) of Theorem 1.2. For this, fix α1\alpha\geq 1 and x¯𝒮(𝒜)\underline{x}\in{\mathcal{S}}(\mathcal{A}), and expand

Iα(x¯)2=(v𝒫\V𝒫ram(x¯)(α+Ψv(F(x¯))(1Ψv(F(x¯)))))2I_{\alpha}(\underline{x})^{2}=\left(\displaystyle\sum_{v\in{\mathcal{P}}\backslash V_{\mathcal{P}}^{\text{ram}}(\underline{x})}\left(\alpha+\Psi_{v}(F(\underline{x}))(\ell-1-\Psi_{v}(F(\underline{x})))\right)\right)^{2}

as

v1,v2𝒫\V𝒫ram(x¯)(α2+α(1)(Ψv1(F(x¯))+Ψv2(F(x¯)))α(Ψv1(F(x¯))2+Ψv2(F(x¯))2)+(1)2Ψv1(F(x¯))Ψv2(F(x¯))(1)(Ψv1(F(x¯))2Ψv2(F(x¯))+Ψv1(F(x¯))Ψv2(F(x¯))2)+Ψv1(F(x¯))2Ψv2(F(x¯))2).\displaystyle\sum_{v_{1},v_{2}\in{\mathcal{P}}\backslash V_{\mathcal{P}}^{\text{ram}}(\underline{x})}\left(\alpha^{2}+\alpha(\ell-1)\left(\Psi_{v_{1}}(F(\underline{x}))+\Psi_{v_{2}}(F(\underline{x}))\right)\right.\\ \left.-\alpha\left(\Psi_{v_{1}}(F(\underline{x}))^{2}+\Psi_{v_{2}}(F(\underline{x}))^{2}\right)+(\ell-1)^{2}\Psi_{v_{1}}(F(\underline{x}))\Psi_{v_{2}}(F(\underline{x}))\right.\\ \left.-(\ell-1)\left(\Psi_{v_{1}}(F(\underline{x}))^{2}\Psi_{v_{2}}(F(\underline{x}))+\Psi_{v_{1}}(F(\underline{x}))\Psi_{v_{2}}(F(\underline{x}))^{2}\right)+\Psi_{v_{1}}(F(\underline{x}))^{2}\Psi_{v_{2}}(F(\underline{x}))^{2}\right).

Applying (2.10), the above sum simplifies precisely to

v1,v2𝒫\V𝒫ram(x¯)((α(1))2+(α(1))(Ψv1(F(x¯))+Ψv2(F(x¯)))+Ψv1(F(x¯))Ψv2(F(x¯))),\displaystyle\sum_{v_{1},v_{2}\in{\mathcal{P}}\backslash V_{\mathcal{P}}^{\text{ram}}(\underline{x})}\left((\alpha-(\ell-1))^{2}+(\alpha-(\ell-1))\left(\Psi_{v_{1}}(F(\underline{x}))+\Psi_{v_{2}}(F(\underline{x}))\right)+\Psi_{v_{1}}(F(\underline{x}))\Psi_{v_{2}}(F(\underline{x}))\right),

which reveals that the optimal choice of α\alpha to minimize Iα(x¯)2I_{\alpha}(\underline{x})^{2} is

α:=1.\alpha:=\ell-1.

Using the above choice of α\alpha in (1.21) of Theorem 1.2, we obtain that

|𝒮(𝒜)|\displaystyle\left|{\mathcal{S}}(\mathcal{A})\right| \displaystyle\leq 1|𝒫|2x¯𝒜I1(x¯)2+2|𝒫|x¯𝒜|V𝒫ram(x¯)|\displaystyle\frac{1}{\left|{\mathcal{P}}\right|^{2}}\displaystyle\sum_{\underline{x}\in{\mathcal{A}}}\;I_{\ell-1}(\underline{x})^{2}+\frac{2}{|{\mathcal{P}}|}\;\displaystyle\sum_{\underline{x}\in{\mathcal{A}}}|V_{\mathcal{P}}^{\text{ram}}(\underline{x})| (2.11)
=\displaystyle= 1|𝒫|2x¯𝒜v1,v2𝒫\V𝒫ram(x¯)Ψv1(F(x¯))Ψv2(F(x¯))+2|𝒫|x¯𝒜|V𝒫ram(x¯)|.\displaystyle\frac{1}{\left|{\mathcal{P}}\right|^{2}}\displaystyle\sum_{\underline{x}\in{\mathcal{A}}}\;\displaystyle\sum_{v_{1},v_{2}\in{\mathcal{P}}\backslash V_{\mathcal{P}}^{\text{ram}}(\underline{x})}\Psi_{v_{1}}(F(\underline{x}))\Psi_{v_{2}}(F(\underline{x}))+\frac{2}{|{\mathcal{P}}|}\displaystyle\sum_{\underline{x}\in{\mathcal{A}}}|V_{\mathcal{P}}^{\text{ram}}(\underline{x})|.

We leave the ramified sieve term as is, and treat the first term on the right-hand side as follows. By interchanging summations, we write this term as

1|𝒫|2v1,v2𝒫x¯𝒜\(𝕏𝒪ram(v1)𝕏𝒪ram(v2))Ψv1(F(x¯))Ψv2(F(x¯)).\frac{1}{\left|{\mathcal{P}}\right|^{2}}\displaystyle\sum_{v_{1},v_{2}\in{\mathcal{P}}}\;\displaystyle\sum_{\underline{x}\in{\mathcal{A}}\backslash\left(\mathbb{X}_{\mathcal{O}}^{\text{ram}}(v_{1})\cup\mathbb{X}_{\mathcal{O}}^{\text{ram}}(v_{2})\right)}\Psi_{v_{1}}(F(\underline{x}))\Psi_{v_{2}}(F(\underline{x})).

Distinguishing between v1=v2v_{1}=v_{2} and v1v2v_{1}\neq v_{2}, we obtain that the above expression equals

1|𝒫|2v1𝒫x¯𝒜\𝕏𝒪ram(v1)Ψv1(F(x¯))2+1|𝒫|2v1,v2𝒫v1v2x¯𝒜\(𝕏𝒪ram(v1)𝕏𝒪ram(v2))Ψv1(F(x¯))Ψv2(F(x¯)).\frac{1}{\left|{\mathcal{P}}\right|^{2}}\displaystyle\sum_{v_{1}\in{\mathcal{P}}}\;\displaystyle\sum_{\underline{x}\in{\mathcal{A}}\backslash\mathbb{X}_{\mathcal{O}}^{\text{ram}}(v_{1})}\Psi_{v_{1}}(F(\underline{x}))^{2}+\frac{1}{\left|{\mathcal{P}}\right|^{2}}\displaystyle\sum_{\begin{subarray}{c}v_{1},v_{2}\in{\mathcal{P}}\\ v_{1}\neq v_{2}\end{subarray}}\;\displaystyle\sum_{\underline{x}\in{\mathcal{A}}\backslash\left(\mathbb{X}_{\mathcal{O}}^{\text{ram}}(v_{1})\cup\mathbb{X}_{\mathcal{O}}^{\text{ram}}(v_{2})\right)}\Psi_{v_{1}}(F(\underline{x}))\Psi_{v_{2}}(F(\underline{x})). (2.12)

By (2.6), for each summand in the inner sum of the first double sum above we have

Ψv1(F(x¯))2(1)2,\Psi_{v_{1}}(F(\underline{x}))^{2}\leq(\ell-1)^{2},

so (2.12) is bounded from above by

(1)2|𝒜||𝒫|+1|𝒫|2v1,v2𝒫v1v2|x¯𝒜\(𝕏𝒪ram(v1)𝕏𝒪ram(v2))Ψv1(F(x¯))Ψv2(F(x¯))|.(\ell-1)^{2}\frac{\left|{\mathcal{A}}\right|}{\left|{\mathcal{P}}\right|}+\frac{1}{\left|{\mathcal{P}}\right|^{2}}\displaystyle\sum_{\begin{subarray}{c}v_{1},v_{2}\in{\mathcal{P}}\\ v_{1}\neq v_{2}\end{subarray}}\left|\displaystyle\sum_{\underline{x}\in{\mathcal{A}}\backslash\left(\mathbb{X}_{\mathcal{O}}^{\text{ram}}(v_{1})\cup\mathbb{X}_{\mathcal{O}}^{\text{ram}}(v_{2})\right)}\Psi_{v_{1}}(F(\underline{x}))\Psi_{v_{2}}(F(\underline{x}))\right|. (2.13)

Inserting this in (2.11) then proves the theorem.

2.4. Comparison to square and polynomial sieves

It is informative to compare Theorems 1.2 and 1.3 to their antecedents in a wide range of settings over \mathbb{Q}. These include the square sieve of Heath-Brown [HB84], which counts perfect square values of a polynomial; the power sieve in [Mun09], [Bra15], which counts perfect rr-th power values of a polynomial; and stronger versions of the square or power sieve, combined with the qq-analogue of Van der Corput’s method, in [Pie06], [HBP12]. Moreover, our work is motivated by the polynomial sieve of Browning [Bro15], also developed in a different direction involving trace functions, by Bonolis [Bon21].

All of these works develop a sieve method to tackle a problem roughly of the following form (over \mathbb{Q}): given an appropriate polynomial f(z;x¯)f(z;\underline{x}) and a set of interest 𝒜\mathcal{A}, how many x¯𝒜\underline{x}\in\mathcal{A} of bounded height have f(z;x¯)f(z;\underline{x}) solvable? This is clearly related to both our aims and our methods, now in the setting of function fields.

One of the most obvious differences is that we do not include a weight to count multiplicities of values. To understand the two main roles of the multiplicity-counting weight in the earlier settings, suppose one fixes a set 𝒜\mathcal{A} and aims to bound from above the quantity

S(𝒜):=x¯𝒜f(z;x¯)solublew(x¯)S(\mathcal{A}):=\sum_{\begin{subarray}{c}\underline{x}\in\mathcal{A}\\ f(z;\underline{x})\;\text{soluble}\end{subarray}}w(\underline{x}) (2.14)

for a fixed integer-coefficient polynomial f(z;x¯)=p0(x¯)zd++pd(x¯)f(z;\underline{x})=p_{0}(\underline{x})z^{d}+\cdots+p_{d}(\underline{x}) of interest, with x¯=(x0,,xn)\underline{x}=(x_{0},\ldots,x_{n}). In the most straightforward comparison to our setting, Browning’s polynomial sieve then provides an upper bound for S(𝒜)S(\mathcal{A}) in the form of an inequality like (2.4), but with Si,j(v1,v2)S_{i,j}(v_{1},v_{2}) containing the coefficient w(x¯)w(\underline{x}) within the sum expressed in (2.5).

In the classical setting of the square sieve, in order to bound from above the number of perfect square values of a fixed polynomial, say F(y1,,yk)[Y1,,Yk]F(y_{1},\ldots,y_{k})\in\mathbb{Z}[Y_{1},\ldots,Y_{k}] with y¯=(y1,,yk)\underline{y}=(y_{1},\ldots,y_{k}) lying in a certain region \mathcal{R}, we would set f(z;x)=z2xf(z;x)=z^{2}-x and choose w(x)=#{y¯:F(y1,,yk)=x}w(x)=\#\{\underline{y}\in\mathcal{R}:F(y_{1},\ldots,y_{k})=x\}. Then we would apply the sieve inequality, including the weight ww, to bound (2.14) from above. Similarly for the rr-power sieve and its variants in [Mun09], [HBP12], [Bra15], to count perfect rr-power values of some polynomial F(y1,,yk)F(y_{1},\ldots,y_{k}) of interest, we would set f(z;x)=zrxf(z;x)=z^{r}-x and define w(x)w(x) as above. With the more recent polynomial sieve now available in [Bro15] (and furthermore in our present geometric treatment), this type of multiplicity-counting weight is no longer as relevant, since one could instead nominally set w(x)=1w(x)=1 and instead choose p2(y¯)=F(y¯)p_{2}(\underline{y})=F(\underline{y}) for the square sieve or more generally pr(y¯)=F(y¯)p_{r}(\underline{y})=F(\underline{y}) for the rr-power sieve. Indeed, in the application of [Bro15, p. 11], the weight is defined simply to be the indicator function of a finite set. (Bonolis [Bon21, p. 27] defines a weight that counts such solutions in a “smoothed” sense.)

This brings us to the second use of the weight: it is used to eliminate what we have here called the ramified sieve term, that is,

2|𝒫|x¯𝒜|V𝒫ram(x¯)|.\frac{2}{|{\mathcal{P}}|}\;\displaystyle\sum_{\underline{x}\in{\mathcal{A}}}|V_{\mathcal{P}}^{\text{ram}}(\underline{x})|.

Previous sieve lemmas in number field settings assume that w(x¯)w(\underline{x}) vanishes for x¯\underline{x} sufficiently large, e.g., that w(x¯)=0w(\underline{x})=0 if |x¯|exp(|𝒫|)|\underline{x}|\geq\operatorname{exp}(|\mathcal{P}|). In the classical settings, in the derivation of the sieve inequality, at step (2.1) the expression |V𝒫ram(x¯)||V_{\mathcal{P}}^{\mathrm{ram}}(\underline{x})| counted the number of primes in the sieving set 𝒫\mathcal{P} that divided a certain (nonzero) polynomial expression HH in x¯\underline{x}, so that for some degree DD,

|V𝒫ram(H(x¯))|ω(H(x¯))=o(|x¯|D)=o(|𝒫|),|V_{\mathcal{P}}^{\mathrm{ram}}(H(\underline{x}))|\leq\omega(H(\underline{x}))=\operatorname{o}(|\underline{x}|^{D})=\operatorname{o}(|\mathcal{P}|),

under the relevant hypothesis assumed for the support of ww. (Here, as is standard, ω\omega counts the number of distinct prime factors, with ω(m)logm/loglog(3m)\omega(m)\ll\operatorname{log}m/\operatorname{log}\operatorname{log}(3m) for any integer m1m\geq 1.) In such a setting, we would conclude in place of (2.3) that

Σ|𝒫|2|𝒮(𝒜)|+|𝒮(𝒜)|o(|𝒫|2),\Sigma\geq|\mathcal{P}|^{2}\cdot|\mathcal{S}(\mathcal{A})|+|\mathcal{S}(\mathcal{A})|\cdot\operatorname{o}(|\mathcal{P}|^{2}),

so that the contribution of the ramified sieve term is dominated by the unramified sieve term, and hence omitted from the final sieve inequality (see e.g. [Bro15, Thm. 1.1]). In our present treatment, we prefer not to include any weight and hence we explicitly record the ramified sieve term, which must be bounded later.

Finally, in comparison to the polynomial sieve, we have emphasized not the size of the fiber |ηv1(x¯(modv))|\left|\eta_{v}^{-1}(\underline{x}\,(\operatorname{mod}v))\right| but the quantity |ηv1(x¯(modv))|1\left|\eta_{v}^{-1}(\underline{x}\,(\operatorname{mod}v))\right|-1 as in (2.9). As we have seen in the argument of §2.3, at least in the case of cyclic morphisms, this normalization is more informative of the correct choice of α\alpha, rather than the expansion in terms of the fibers, which leads to the hard-to-interpret coefficients in (2.4).

3. Preliminaries on function field arithmetic

Since Theorem 1.1 is a result about 𝔽q(T)\mathbb{F}_{q}(T), we gather in this section standard function field notation and remarks pertinent to our forthcoming arguments.

We recall the general setting of Section 1: KK is a global field of arbitrary characteristic; 𝒪{\mathcal{O}} is the ring of integers in KK; VKV_{K} is the set of places of KK; the valuations ||v|\cdot|_{v} associated to the places vVKv\in V_{K} are normalized such that the product formula (1.1) holds; htK\operatorname{ht}_{K} is the height function on Kn\mathbb{P}^{n}_{K}, defined in (1.2) and its companion height function on KK, also denoted htK\operatorname{ht}_{K}, is defined in (1.3).

We now focus on the particular global function field

K=𝔽q(T)K=\mathbb{F}_{q}(T)

and on the ring

𝒪=𝒪K=𝔽q[T],\mathcal{O}={\mathcal{O}}_{K}=\mathbb{F}_{q}[T],

where qq is the power of an odd rational prime pp. This setting will be kept throughout the remainder of the paper, unless explicitly specified otherwise.

In the setting of Theorem 1.1 and Theorem 5.1, we assume that pmp\nmid m, where mm is the degree of the polynomial FF therein. Furthermore, we assume that (q1)\ell\mid(q-1), where \ell is a rational prime defined as the degree of the cyclic cover η\eta therein; this divisibility assumption ensures that 𝔽q\mathbb{F}_{q} contains the \ell-th roots of unity. Finally, the assumption m\ell\mid m ensures that the weighted projective space 𝔽q(T)n+1(1,,1,m)\mathbb{P}_{\mathbb{F}_{q}(T)}^{n+1}\left(1,\ldots,1,\frac{m}{\ell}\right) is well-defined.

Note that K=𝔽q(T)K=\mathbb{F}_{q}(T) is the simplest instance of a global function field with field of constants 𝔽q\mathbb{F}_{q}, and that 𝒪K\mathcal{O}_{K} is the ring of elements of KK which have only 1T\frac{1}{T} as a pole. In particular, 𝒪K\mathcal{O}_{K} is a Dedekind domain (and, actually, a Euclidean domain) and plays the role of the ring \mathbb{Z} in the analogy between the arithmetic of 𝔽q(T)\mathbb{F}_{q}(T) and that of \mathbb{Q}.

As usual, we identify a place vVKv\in V_{K} with a generator of its associated unique maximal ideal. For our particular K=𝔽q(T)K=\mathbb{F}_{q}(T), we have either that v=1Tv=\frac{1}{T}, which we refer to as the place at infinity of KK, or that any vVK\{1T}v\in V_{K}\backslash\{\frac{1}{T}\}, which we refer to as a finite place or simply as a prime of KK, may be thought of as a monic irreducible polynomial in 𝔽q[T]\mathbb{F}_{q}[T]. We recall that, for a nonzero polynomial x𝒪K=𝔽q[T]x\in{\mathcal{O}}_{K}=\mathbb{F}_{q}[T], we use degT(x)\operatorname{deg}_{T}(x) to denote its degree in TT.

To simplify the exposition, we use the symbol \infty for 1T\frac{1}{T} and the symbol π\pi for an arbitrary prime of KK (that is, a monic irreducible in 𝒪K\mathcal{O}_{K}). We denote by KK_{\infty} the completion of KK with respect to the topology defined by |||\cdot|_{\infty} and recall that

K=𝔽q((1T))={jNajTj:N,aj𝔽qjN,aN0}.K_{\infty}=\mathbb{F}_{q}\left(\left(\frac{1}{T}\right)\right)=\left\{\displaystyle\sum_{j\leq N}a_{j}T^{j}:N\in\mathbb{Z},a_{j}\in\mathbb{F}_{q}\;\forall j\leq N,a_{N}\neq 0\right\}.

We denote by kπk_{\pi} the residue field 𝔽q[T]/(π)\mathbb{F}_{q}[T]/(\pi) of π\pi and by k¯π\overline{k}_{\pi} a fixed algebraic closure of kπk_{\pi}, and recall that kπk_{\pi} is a finite field with qdegT(π)q^{\operatorname{deg}_{T}(\pi)} elements. We denote by ordπ(x)\operatorname{ord}_{\pi}(x) the power of π\pi that exactly divides xx.

The absolute values |||\cdot|_{\infty}, ||π|\cdot|_{\pi} on KK are defined by

|0|:=0,|xy|:=qdegTxdegTy,|0|_{\infty}:=0,\quad\left|\frac{x}{y}\right|_{\infty}:=q^{\operatorname{deg}_{T}x-\operatorname{deg}_{T}y},

and

|0|π:=0,|xy|π:=qordπ(y)ordπ(x),|0|_{\pi}:=0,\quad\left|\frac{x}{y}\right|_{\pi}:=q^{\operatorname{ord}_{\pi}(y)-\operatorname{ord}_{\pi}(x)},

where x,y𝔽q[T]\{0}x,y\in\mathbb{F}_{q}[T]\backslash\{0\}. With these definitions, the associated valuations satisfy the product formula (1.1), the height function on Kn\mathbb{P}_{K}^{n} becomes

htK([x0:x1::xn])=max{|xi|:0in}\operatorname{ht}_{K}([x_{0}:x_{1}:\ldots:x_{n}])=\max\{|x_{i}|_{\infty}:0\leq i\leq n\}

for any [x0:x1::xn]Kn[x_{0}:x_{1}:\ldots:x_{n}]\in\mathbb{P}_{K}^{n}, and the height function on KK becomes

htK(x)=|x|\operatorname{ht}_{K}(x)=|x|_{\infty} (3.1)

for any xKx\in K^{\ast}.

Remark that, since K=𝔽q(T)K=\mathbb{F}_{q}(T) has class number 1, for any [x]=[x0:x1::xn]Kn[x]=[x_{0}:x_{1}:\ldots:x_{n}]\in\mathbb{P}_{K}^{n} we can find x0,x1,,xn𝒪Kx_{0}^{\prime},x_{1}^{\prime},\ldots,x_{n}^{\prime}\in{\mathcal{O}}_{K} with gcd(x0,x1,,xn)=1\operatorname{gcd}(x_{0}^{\prime},x_{1}^{\prime},\ldots,x_{n}^{\prime})=1, such that [x]=[x0:x1::xn][x]=[x_{0}^{\prime}:x_{1}^{\prime}:\ldots:x_{n}^{\prime}] and htK([x])=max{|xi|:0in}\operatorname{ht}_{K}([x])=\max\{|x_{i}^{\prime}|_{\infty}:0\leq i\leq n\}. Reasoning similarly, we can extend the height htK\operatorname{ht}_{K} to 𝔸Kn+1\mathbb{A}_{K}^{n+1} by setting htK((x0,x1,,xn)):=max{htK(xi):0in}\operatorname{ht}_{K}((x_{0},x_{1},\ldots,x_{n})):=\max\{\operatorname{ht}_{K}(x_{i}):0\leq i\leq n\} for any x¯:=(x0,x1,,xn)𝔸Kn+1\underline{x}:=(x_{0},x_{1},\ldots,x_{n})\in\mathbb{A}_{K}^{n+1}, which is consistent with (1.17).

In order to employ Fourier analysis on KK, we extend |||\cdot|_{\infty} to Kn+1K_{\infty}^{n+1} by

|(x0,x1,,xn)|:=max{|xi|:0in}|(x_{0},x_{1},\ldots,x_{n})|_{\infty}:=\max\left\{|x_{i}|_{\infty}:0\leq i\leq n\right\}

and consider the additive character

ψ:K,\psi_{\infty}:K_{\infty}\longrightarrow\mathbb{C}^{\ast},
ψ(jNajTj):=exp(2πiTr𝔽q/𝔽p(a1)p),\psi_{\infty}\left(\displaystyle\sum_{j\leq N}a_{j}T^{j}\right):=\operatorname{exp}\left(\frac{2\pi i\operatorname{Tr}_{\mathbb{F}_{q}/\mathbb{F}_{p}}(a_{-1})}{p}\right), (3.2)

where Tr𝔽q/𝔽p\operatorname{Tr}_{\mathbb{F}_{q}/\mathbb{F}_{p}} is the trace map.

4. Preliminaries on duals and reductions

Our proof of Theorem 1.1 will be an application of the sieve inequality in Theorem 1.3 for K=𝔽q(T)K=\mathbb{F}_{q}(T). To choose a sieving set 𝒫{\mathcal{P}} and to state the appropriate Weil-Deligne bounds needed to estimate the resulting unramified sieve term, we require certain standard facts about duals and reductions. An appendix by Joseph Rabinoff provides a self-contained reference for all the facts we require; we now summarize the consequences for our specific application.

Proposition 4.1.

Let qq be an odd rational prime power, n2n\geq 2 an integer, and H𝔽q[T][X0,,Xn]H\in\mathbb{F}_{q}[T][X_{0},\ldots,X_{n}] a homogeneous polynomial of degree m:=degX¯(H)m:=\operatorname{deg}_{\underline{X}}(H) with m2m\geq 2. Denote by WW the projective hypersurface defined in 𝔽q(T)n\mathbb{P}_{\mathbb{F}_{q}(T)}^{n} by

W:H(X0,,Xn)=0W:\quad H(X_{0},\ldots,X_{n})=0

and assume that WW is smooth over 𝔽q(T)\mathbb{F}_{q}(T). Then the following properties hold.

  1. (i)

    WW and its dual WW^{\ast} are geometrically integral, with dimW=dimW=n1\operatorname{dim}W=\operatorname{dim}W^{\ast}=n-1.

  2. (ii)

    There exists an absolutely irreducible, homogeneous polynomial H𝔽q[T][X0,,Xn],H^{\ast}\in\mathbb{F}_{q}[T][X_{0},\ldots,X_{n}], which depends only on HH, such that WW^{\ast} is defined in 𝔽q(T)n\mathbb{P}_{\mathbb{F}_{q}(T)}^{n} by

    W:H(X0,,Xn)=0.W^{\ast}:\quad H^{\ast}(X_{0},\ldots,X_{n})=0.
  3. (iii)

    There exists a finite set of finite places 𝒫𝔽q[T],\mathcal{P}\subseteq\mathbb{F}_{q}[T], which depends only on HH and HH^{*}, such that:

    1. (iii.1)

      for all finite places π𝒫\pi\not\in\mathcal{P}, degX¯(H(modπ))=degX¯(H)\operatorname{deg}_{\underline{X}}(H(\operatorname{mod}\pi))=\operatorname{deg}_{\underline{X}}(H) and the projective variety WπW_{\pi} defined by the equation

      Wπ:H(X0,,Xn)0(modπ)W_{\pi}:\quad H(X_{0},\ldots,X_{n})\equiv 0\,(\operatorname{mod}\pi)

      is a smooth and geometrically integral hypersurface;

    2. (iii.2)

      for all finite places π𝒫\pi\not\in\mathcal{P}, the dual variety (Wπ)(W_{\pi})^{*} is geometrically integral and is defined by the equation

      (Wπ):H(X0,,Xn)0(modπ);(W_{\pi})^{\ast}:\quad H^{\ast}(X_{0},\ldots,X_{n})\equiv 0\,(\operatorname{mod}\pi);

      in particular, (Wπ)=(W)π(W_{\pi})^{*}=(W^{*})_{\pi} and the notation WπW_{\pi}^{*} is well-defined.

    3. (iii.3)

      for all finite places π𝒫\pi\not\in\mathcal{P}, dimWπ=dimWπ=n1.\operatorname{dim}W_{\pi}=\operatorname{dim}W^{\ast}_{\pi}=n-1.

Proof.

Assertion (i) is the statement of Proposition 11.2(1). As observed in the appendix, for any field kk, a hypersurface in kn\mathbb{P}_{k}^{n} is the zero set of a nonzero homogeneous polynomial in k[X0,,Xn]k[X_{0},\ldots,X_{n}]. Thus implies that WW^{*} is defined by some H𝔽q(T)[X0,,Xn]H^{*}\in\mathbb{F}_{q}(T)[X_{0},\ldots,X_{n}]; after clearing denominators, we may assume H𝔽q[T][X0,,Xn]H^{*}\in\mathbb{F}_{q}[T][X_{0},\ldots,X_{n}]. Absolute irreducibility of HH^{*} is equivalent to geometric irreducibility of WW^{*}, so this proves (ii). Let 𝒫\mathcal{P} be the finite set SS in Proposition 11.5. Then (iii.1) and (iii.2) amount to statements (2) and (3) of Proposition 11.5, using Lemma 11.1 and Proposition 11.2(1) for geometric integrality. Assertion (iii.3) is a consequence of (iii.1) and (iii.2). ∎

5. Initiating the sieve to prove Theorem 1.1

We are now ready to start the proof of Theorem 1.1 on counting points on cyclic covers of Kn\mathbb{P}_{K}^{n} for K=𝔽q(T)K=\mathbb{F}_{q}(T) and an arbitrary integer n2n\geq 2.

Recall that, in Theorem 1.3, we take 𝕏:=Kn\mathbb{X}:=\mathbb{P}^{n}_{K} and 𝕐\mathbb{Y} the projective scheme in the weighted projective space Kn+1(1,,1,m)\mathbb{P}_{K}^{n+1}\left(1,\dots,1,\frac{m}{\ell}\right) defined by a model Xn+1=F(X0,,Xn)X_{n+1}^{\ell}=F(X_{0},\ldots,X_{n}) with \ell prime such that m\ell\mid m, and with F𝒪K[X0,,Xn]F\in{\mathcal{O}}_{K}[X_{0},\ldots,X_{n}] homogeneous of degree m1m\geq 1 and having the property that the hypersurface in K¯n\mathbb{P}^{n}_{\overline{K}} defined by F(X0,,Xn)=0F(X_{0},\ldots,X_{n})=0 is smooth. Moreover, recall that η:𝕐𝕏\eta:\mathbb{Y}\longrightarrow\mathbb{X} is the cyclic map of degree \ell defined over 𝔸K¯n\mathbb{A}_{\overline{K}}^{n} by η(x0,x1,,xn,xn+1):=(x0,x1,,xn)\eta(x_{0},x_{1},\ldots,x_{n},x_{n+1}):=(x_{0},x_{1},\ldots,x_{n}).

Note that each element [y][y] of 𝕐\mathbb{Y} is of the form [y]=[x0::xn:xn+1][y]=[x_{0}:\ldots:x_{n}:x_{n+1}], where [x0:xn]Kn[x_{0}:\ldots x_{n}]\in\mathbb{P}^{n}_{K} and where xn+1Kx_{n+1}\in K satisfies the equation

xn+1=F(x0,,xn).x_{n+1}^{\ell}=F(x_{0},\ldots,x_{n}).

In order to bound the number of projective solutions, we will think of the above equation as an affine model. As such, it is equally natural to restate the result of Theorem 1.1 as:

Theorem 5.1.

Let qq be an odd rational prime power, n2n\geq 2 an integer, 2\ell\geq 2 a rational prime, and F𝔽q[T][X0,,Xn]F\in\mathbb{F}_{q}[T][X_{0},\ldots,X_{n}] a homogeneous polynomial of degree m2m\geq 2 in X0,,XnX_{0},\ldots,X_{n}, with char𝔽qm.\operatorname{char}\mathbb{F}_{q}\nmid m. Assume the conditions:

  1. (i)

    gcd(m,q1)\ell\mid\operatorname{gcd}(m,q-1);

  2. (ii)

    F(X0,,Xn)=0F(X_{0},\ldots,X_{n})=0 defines a nonsingular projective hypersurface in 𝔽q(T)¯n\mathbb{P}^{n}_{\overline{\mathbb{F}_{q}(T)}}.

For every b>0b>0, let Mn(F;b)M_{n}(F;b) denote the cardinality of the set

{(x0,,xn)𝔽q[T]n+1:degT(xi)<b0in,xn+1𝔽q[T]such thatxn+1=F(x0,,xn)}.\left\{(x_{0},\ldots,x_{n})\in\mathbb{F}_{q}[T]^{n+1}:\;\operatorname{deg}_{T}(x_{i})<b\;\forall 0\leq i\leq n,\exists x_{n+1}\in\mathbb{F}_{q}[T]\;\text{such that}\;x_{n+1}^{\ell}=F(x_{0},\ldots,x_{n})\right\}.

Then for any b1b\geq 1

Mn(F;b),m,n,q,F(qb)(n+1)n+1n+2bn+1n+2,M_{n}(F;b)\ll_{\ell,m,n,q,F}\left(q^{b}\right)^{(n+1)-\frac{n+1}{n+2}}b^{\frac{n+1}{n+2}},

in which the implicit constant depends on ,m,n,q\ell,m,n,q and FF.

More precisely the implicit constant depends on FF in terms of degT(F)\operatorname{deg}_{T}(F) and degT(F)\operatorname{deg}_{T}(F^{*}) as well as a finite exceptional set of primes determined by FF; see §9. (Here, and in all that follows, given a polynomial H𝔽q[T][X0,,Xn],H\in\mathbb{F}_{q}[T][X_{0},\ldots,X_{n}], degTH\operatorname{deg}_{T}H refers to the maximum degree in TT of any coefficient of HH.)

5.1. Choosing the sieving set

We prove Theorem 5.1 (and hence Theorem 1.1) by means of the geometric sieve derived in Theorem 1.3. To phrase our goal as a sieve problem, from here onwards, unless otherwise stated, we keep the setting of Theorem 5.1 and the notation of §3, and work with the sets

𝒜\displaystyle{\mathcal{A}} :=\displaystyle:= {x¯=(x0,,xn)𝒪Kn+1:degT(xi)<b0in},\displaystyle\left\{\underline{x}=(x_{0},\ldots,x_{n})\in{\mathcal{O}}_{K}^{n+1:}\;\operatorname{deg}_{T}(x_{i})<b\;\forall 0\leq i\leq n\right\},
𝒮F(𝒜)\displaystyle{\mathcal{S}}_{F}({\mathcal{A}}) :=\displaystyle:= {x¯𝒜:y𝒪Ksuch thaty=F(x¯)}\displaystyle\left\{\underline{x}\in{\mathcal{A}}:\;\exists y\in{\mathcal{O}}_{K}\;\text{such that}\;y^{\ell}=F(\underline{x})\right\}

for a fixed arbitrary integer b>0b>0. Note that the trivial upper bound for |𝒮F(𝒜)||\mathcal{S}_{F}(\mathcal{A})| is

|𝒜|qb(n+1).|\mathcal{A}|\leq q^{b(n+1)}. (5.1)

To improve upon this bound using Theorem 1.3, we seek a suitable sieving set 𝒫{\mathcal{P}} of primes π\pi of KK, which we describe below.

We consider the smooth projective hypersurface

W(F):F(X0,,Xn)=0,W(F):\quad F(X_{0},\ldots,X_{n})=0,

whose projective dual we denote by W(F)W(F)^{\ast}. By parts (i), (ii) of Proposition 4.1, there exists a homogeneous polynomial F𝒪K[X0,,Xn]F^{\ast}\in{\mathcal{O}_{K}}[X_{0},\ldots,X_{n}] , absolutely irreducible over KK, that defines W(F)W(F)^{\ast}, that is,

W(F):F(X0,,Xn)=0.W(F)^{\ast}:\quad F^{\ast}(X_{0},\ldots,X_{n})=0.

Recall that by part (iii) of Proposition 4.1 there exists a finite set of finite primes

𝒫exc=𝒫exc(F){\mathcal{P}}_{\text{exc}}={\mathcal{P}}_{\text{exc}}(F) (5.2)

satisfying that proposition.

Upon fixing some positive integer Δ\Delta, we define the sieving set of primes as

𝒫:={π𝒪K:degT(π)=Δ,π𝒫exc}.{\mathcal{P}}:=\{\pi\in{\mathcal{O}}_{K}:\operatorname{deg}_{T}(\pi)=\Delta,\pi\not\in{\mathcal{P}}_{\text{exc}}\}. (5.3)

Later on in (9.4), we will choose Δ=Δ(n,b)\Delta=\Delta(n,b) optimally in terms of nn and bb; see also the observation below (5.4).

To understand the growth of |𝒫||{\mathcal{P}}|, let us recall the Prime Polynomial Theorem [Ros02, Thm. 2.2, p. 14]:

|#{π𝒪K:degT(π)=Δ}qΔΔ|qΔ2Δ+qΔ3.\left|\#\{\pi\in{\mathcal{O}}_{K}:\operatorname{deg}_{T}(\pi)=\Delta\}-\frac{q^{\Delta}}{\Delta}\right|\leq\frac{q^{\frac{\Delta}{2}}}{\Delta}+q^{\frac{\Delta}{3}}. (5.4)

We infer from (5.4) that, upon taking Δ(n+1)b/(n+2)\Delta\approx\lfloor(n+1)b/(n+2)\rfloor (see the precise choice in (9.4)), we have that

|𝒫|qΔΔ(qΔ2Δ+qΔ3)|𝒫exc|qΔ2Δ|\mathcal{P}|\geq\frac{q^{\Delta}}{\Delta}-\left(\frac{q^{\frac{\Delta}{2}}}{\Delta}+q^{\frac{\Delta}{3}}\right)-|\mathcal{P}_{\text{exc}}|\geq\frac{q^{\Delta}}{2\Delta} (5.5)

for all bb that are sufficiently large relative to nn and |𝒫exc||\mathcal{P}_{\text{exc}}|. We denote the least such bb by

b(n,q,F)b(n,q,F)

and refer the reader to §9.2 for more details. While we still allow Δ>0\Delta>0 to be arbitrary, from now on we assume the lower bound (5.5) for |𝒫||\mathcal{P}|.

Our task for this particular sieve problem is to bound non-trivially the right-hand side of the resulting sieve inequality of Theorem 1.3. Recall that there are three terms – the main term, the ramified term, and the unramified term. In this section, we bound the first two terms quite easily, and in the remaining sections we focus on bounding the third term.

We make the notational convention that, for x¯:=(x0,,xn)𝒪Kn+1\underline{x}:=(x_{0},\ldots,x_{n})\in\mathcal{O}_{K}^{n+1}, we write

degT(x¯)<b\operatorname{deg}_{T}(\underline{x})<b

to mean

degT(xi)<b0in.\operatorname{deg}_{T}(x_{i})<b\ \forall 0\leq i\leq n.

5.2. Upper bound for the main sieve term

Assuming Δ\Delta is chosen and bb is sufficiently large, relative to nn and |𝒫exc(F)||\mathcal{P}_{\text{exc}}(F)|, so that (5.5) holds (again, see §9.2 for the explicit requirements on bb), then by (5.1) and (5.5), the first term on the right-hand side of (1.24) is bounded above by

|𝒜||𝒫|(1)22Δqb(n+1)Δ(1)2Δqb(n+1)Δ.\frac{|\mathcal{A}|}{|\mathcal{P}|}(\ell-1)^{2}\leq 2\Delta q^{b(n+1)-\Delta}(\ell-1)^{2}\ll_{\ell}\Delta q^{b(n+1)-\Delta}. (5.6)

5.3. Upper bound for the ramified sieve term

For the second term on the right-hand side of (1.24), note that for each prime π𝒪K\pi\in{\mathcal{O}}_{K} of good reduction for η\eta, the condition that ηπ\eta_{\pi} is ramified at x¯(modπ)\underline{x}\,(\operatorname{mod}\pi) means that the extension of residue fields defined by x¯(modπ)\underline{x}\,(\operatorname{mod}\pi) is inseparable, which cannot happen in our case since the residue fields are finite, or that ηπ\eta_{\pi} has a ramification index at (x¯,y)(modπ)(\underline{x},y)\,(\operatorname{mod}\pi) larger than 1, which is equivalent to y(modπ)y\,(\operatorname{mod}\pi) being zero. Then,

1|𝒫|x¯𝒜|V𝒫ram(x¯)|=1|𝒫|x¯𝒪Kn+1degT(x¯)<b#{π𝒪K:degT(π)=Δ,π𝒫exc,F(x¯)0(modπ)}.\frac{1}{\left|{\mathcal{P}}\right|}\displaystyle\sum_{\underline{x}\in{\mathcal{A}}}|V_{\mathcal{P}}^{\mathrm{ram}}(\underline{x})|=\frac{1}{|{\mathcal{P}}|}\displaystyle\sum_{\begin{subarray}{c}\underline{x}\in\mathcal{O}_{K}^{n+1}\\ \operatorname{deg}_{T}(\underline{x})<b\end{subarray}}\#\left\{\pi\in{\mathcal{O}_{K}}:\operatorname{deg}_{T}(\pi)=\Delta,\pi\not\in{\mathcal{P}}_{\text{exc}},\;F(\underline{x})\equiv 0\,(\operatorname{mod}\pi)\right\}. (5.7)

We will bound each term according to whether F(x¯)0F(\underline{x})\neq 0 or F(x¯)=0F(\underline{x})=0. First we observe that for any nonconstant polynomial u𝒪Ku\in\mathcal{O}_{K} we have

#{π𝒪K:πu}degT(u).\#\left\{\pi\in\mathcal{O}_{K}:\pi\mid u\right\}\leq\operatorname{deg}_{T}(u).

We also restrict the degree of π\pi, so more precisely, let ωΔ(G)\omega_{\Delta}(G) denote the number of distinct primes π\pi of degree Δ\Delta that divide a fixed G(T)𝔽q[T]G(T)\in\mathbb{F}_{q}[T]. Then

(πGdegT(π)=Δπ)G.\Big{(}\prod_{\begin{subarray}{c}\pi\mid G\\ \operatorname{deg}_{T}(\pi)=\Delta\end{subarray}}\pi\Big{)}\mid G.

Taking norms, we have qΔωΔ(G)|G|q^{\Delta\omega_{\Delta}(G)}\leq|G| and therefore,

ωΔ(G)degT(G)Δ.\omega_{\Delta}(G)\leq\frac{\operatorname{deg}_{T}(G)}{\Delta}.

Thus, after we fix x¯𝒪Kn+1\underline{x}\in\mathcal{O}_{K}^{n+1} with degT(x¯)<b\operatorname{deg}_{T}(\underline{x})<b and F(x¯)0F(\underline{x})\neq 0,

#{π𝒪K:degT(π)=Δ,π𝒫exc,F(x¯)0(modπ)}\displaystyle\#\left\{\pi\in\mathcal{O}_{K}:\operatorname{deg}_{T}(\pi)=\Delta,\pi\not\in{\mathcal{P}}_{\text{exc}},\;F(\underline{x})\equiv 0\,(\operatorname{mod}\pi)\right\} \displaystyle\leq 1ΔdegT(F(x¯))\displaystyle\frac{1}{\Delta}\operatorname{deg}_{T}(F(\underline{x}))
\displaystyle\leq 1Δ(degT(F)+degX¯(F)degT(x¯))\displaystyle\frac{1}{\Delta}(\operatorname{deg}_{T}(F)+\operatorname{deg}_{\underline{X}}(F)\cdot\operatorname{deg}_{T}(\underline{x}))
=\displaystyle= 1Δ(degT(F)+mb).\displaystyle\frac{1}{\Delta}(\operatorname{deg}_{T}(F)+mb).

Here we emphasize, m=degX¯(F)m=\operatorname{deg}_{\underline{X}}(F) is the degree of F𝒪K[X0,,Xn]=𝔽q[T,X0,,Xn]F\in\mathcal{O}_{K}[X_{0},\ldots,X_{n}]=\mathbb{F}_{q}[T,X_{0},\ldots,X_{n}] in X0,,XnX_{0},\ldots,X_{n} and degT(F)\operatorname{deg}_{T}(F) is the maximum degree of (any coefficient of) FF in TT.

Applying the above observations to each summand with F(x¯)0F(\underline{x})\neq 0 on the right-hand side of (5.7), trivially counting the number of x¯\underline{x} in the sum, and applying (5.5) to bound 1/|𝒫|1/|\mathcal{P}| from above, we deduce that this contribution to the ramified sieve term is

m,n,degT(F),bqb(n+1)Δ,\ll_{m,n,\operatorname{deg}_{T}(F),}bq^{b(n+1)-\Delta},

for all bb(n,q,F)b\geq b(n,q,F).

On the other hand, the contribution to (5.7) from those x¯\underline{x} such that F(x¯)=0F(\underline{x})=0 is

|𝒫||𝒫|#{x¯𝒪Kn+1:degT(x¯)<b,F(x¯)=0}degX¯(F)qbn,\frac{|\mathcal{P}|}{|\mathcal{P}|}\#\{\underline{x}\in\mathcal{O}_{K}^{n+1}:\operatorname{deg}_{T}(\underline{x})<b,F(\underline{x})=0\}\leq\operatorname{deg}_{\underline{X}}(F)q^{bn},

by the trivial bound (see e.g. Lemma 10.1, which we include at the end of the paper for completeness). Thus in total the ramified sieve term is

m,n,degT(F)bqb(n+1)Δ+qbn,\ll_{m,n,\operatorname{deg}_{T}(F)}bq^{b(n+1)-\Delta}+q^{bn},

for all bb(n,q,F)b\geq b(n,q,F).

5.4. Remaining work

So far we have proved the upper bound

|𝒮F(𝒜)|m,n,degT(F)bqb(n+1)Δ+qbn+1|𝒫|2π1,π2𝒫π1π2|x¯𝒪Kn+1degT(x¯)<bF(x¯)0(modπ1π2)(|ηπ11(x¯(modπ1))|1)(|ηπ21(x¯(modπ2))|1)|.\left|{\mathcal{S}}_{F}({\mathcal{A}})\right|\ll_{m,n,\operatorname{deg}_{T}(F)}bq^{b(n+1)-\Delta}+q^{bn}\\ +\frac{1}{\left|{\mathcal{P}}\right|^{2}}\displaystyle\sum_{\begin{subarray}{c}\pi_{1},\pi_{2}\in{\mathcal{P}}\\ \pi_{1}\neq\pi_{2}\end{subarray}}\left|\displaystyle\sum_{\begin{subarray}{c}\underline{x}\in\mathcal{O}_{K}^{n+1}\\ \operatorname{deg}_{T}(\underline{x})<b\\ F(\underline{x})\not\equiv 0\,(\operatorname{mod}\pi_{1}\pi_{2})\end{subarray}}\left(\left|\eta_{\pi_{1}}^{-1}(\underline{x}\,(\operatorname{mod}\pi_{1}))\right|-1\right)\left(\left|\eta_{\pi_{2}}^{-1}(\underline{x}\,(\operatorname{mod}\pi_{2}))\right|-1\right)\right|. (5.8)

Recall that our goal is to prove an upper bound for 𝒮F(𝒜)\mathcal{S}_{F}(\mathcal{A}) that improves on the trivial upper bound qb(n+1)q^{b(n+1)} recorded in (5.1). The first term on the right-hand side of (5.8) will be nontrivial as long as Δ>0\Delta>0. Since the trivial upper bound for the unramified sieve term, the last term on the right-hand side of (5.8), is qb(n+1)\ll q^{b(n+1)}, which is as large as the aforementioned trivial bound (5.1), we seek any bound for the unramified sieve term that improves upon this. Ultimately, we will choose Δ\Delta to balance our upper-estimates for the terms on the right-hand side of (5.8).

To tackle the unramified sieve term, we will break our treatment into two main steps: in §6, we expand the term into a sum of complete character sums; in §8, we apply Weil-Deligne bounds (proved in §7) to each of these sums, and then use point-counting results to average over π1,π2𝒫\pi_{1},\pi_{2}\in\mathcal{P}, and sum up the resulting Weil-Deligne bounds.

6. Expansion of unramified sieve term as a mixed character sum

In this section we recall the basic notions of Fourier analysis in our function field setting and use them to expand the unramified sieve term as a mixed character sum. For ease of reference, we first record the main outcome of this section and then proceed with rigorous definitions and derivations.

For any element w¯𝒪Kn+1\underline{w}\in\mathcal{O}_{K}^{n+1}, prime π𝒪K\pi\in\mathcal{O}_{K}, and non-principal multiplicative character χπ\chi_{\pi} of 𝒪K\mathcal{O}_{K}, of modulus π\pi, we define a mixed character sum relative to a polynomial G𝒪K[X0,,Xn]G\in\mathcal{O}_{K}[X_{0},\ldots,X_{n}] by

SG(w¯,χπ):=a¯(modπ)kπn+1χπ(G(a¯))ψ(w¯a¯π),S_{G}(\underline{w},\chi_{\pi}):=\displaystyle\sum_{\underline{a}\,(\operatorname{mod}\pi)\in k_{\pi}^{n+1}}\chi_{\pi}(G(\underline{a}))\;\psi_{\infty}\left(-\frac{\underline{w}\cdot\underline{a}}{\pi}\right), (6.1)

where the additive character ψ(/π)\psi_{\infty}(\cdot/\pi) is defined in (3.2).

Proposition 6.1.

Let qq be an odd rational prime power, n2n\geq 2 an integer, 2\ell\geq 2 a rational prime, and F𝒪K[X0,,Xn]F\in\mathcal{O}_{K}[X_{0},\ldots,X_{n}] a homogeneous polynomial of degree m2m\geq 2 in X0,,XnX_{0},\ldots,X_{n}, with charKm\mathrm{char}K\nmid m, where, as before, K=𝔽q(T)K=\mathbb{F}_{q}(T). Assume gcd(m,q1)\ell\mid\operatorname{gcd}(m,q-1). Let b,Δ>0b,\Delta>0 be integers and assume that

b<2Δ.b<2\Delta.

Defining 𝒫\mathcal{P} as in (5.3), for all primes π1,π2𝒫\pi_{1},\pi_{2}\in\mathcal{P} with π1π2\pi_{1}\neq\pi_{2}, the unramified sieve term can be expanded as

x¯𝒪Kn+1degT(x¯)<bF(x¯)0(modπ1π2)(|ηπ11(x¯(modπ1))|1)(|ηπ21(x¯(modπ2))|1)=q(n+1)(2Δb)χπ1χ0χπ2χ0x¯𝒪Kn+1degT(x¯)<2ΔbSF(π¯2x¯,χπ1)SF(π¯1x¯,χπ2),\displaystyle\sum_{\begin{subarray}{c}\underline{x}\in\mathcal{O}_{K}^{n+1}\\ \operatorname{deg}_{T}(\underline{x})<b\\ F(\underline{x})\not\equiv 0\,(\operatorname{mod}\pi_{1}\pi_{2})\end{subarray}}\left(\left|\eta_{\pi_{1}}^{-1}(\underline{x}\,(\operatorname{mod}\pi_{1}))\right|-1\right)\left(\left|\eta_{\pi_{2}}^{-1}(\underline{x}\,(\operatorname{mod}\pi_{2}))\right|-1\right)\\ =q^{-(n+1)(2\Delta-b)}\displaystyle\sum_{\begin{subarray}{c}\chi_{\pi_{1}}\neq\chi_{0}\\ \chi_{\pi_{2}}\neq\chi_{0}\end{subarray}}\displaystyle\sum_{\begin{subarray}{c}\underline{x}\in\mathcal{O}_{K}^{n+1}\\ \operatorname{deg}_{T}(\underline{x})<2\Delta-b\end{subarray}}S_{F}\left(\bar{\pi}_{2}\underline{x},\chi_{\pi_{1}}\right)S_{F}\left(\bar{\pi}_{1}\underline{x},\chi_{\pi_{2}}\right), (6.2)

in which, for each πi\pi_{i}, the sum is over all non-principal characters χπi\chi_{\pi_{i}} of order \ell.

For later reference, we remark that the left-hand side of (6.2) is unchanged if we omit the condition F(x¯)0(modπ1π2)F(\underline{x})\not\equiv 0\,(\operatorname{mod}\pi_{1}\pi_{2}). Indeed, upon observing that

ηπi1(x¯(modπi))={zkπi:zm=F(x¯)},\eta_{\pi_{i}}^{-1}(\underline{x}\,(\operatorname{mod}\pi_{i}))=\{z\in k_{\pi_{i}}:z^{m}=F(\underline{x})\},

it follows that (|ηπi1(x¯(modπi))|1)=0(|\eta_{\pi_{i}}^{-1}(\underline{x}\,(\operatorname{mod}\pi_{i}))|-1)=0 whenever F(x¯)=0F(\underline{x})=0.

We note that for any polynomial F𝒪K[X0,,Xn]F\in\mathcal{O}_{K}[X_{0},\ldots,X_{n}], the trivial bound for SF(w¯,χπ)S_{F}(\underline{w},\chi_{\pi}), valid for every prime π𝒪K\pi\in\mathcal{O}_{K} and every w¯𝒪Kn+1\underline{w}\in\mathcal{O}_{K}^{n+1}, is

|SF(w¯,χπ)||π|n+1=q(n+1)degT(π).\left|S_{F}(\underline{w},\chi_{\pi})\right|\leq|\pi|_{\infty}^{n+1}=q^{(n+1)\operatorname{deg}_{T}(\pi)}. (6.3)

(Of course, this is far from sharp.) We then see that for all primes π1,π2𝒫\pi_{1},\pi_{2}\in\mathcal{P} with π1π2\pi_{1}\neq\pi_{2}, the trivial bound for the right-hand side of (6.2) is

2q(n+1)2Δ,\ll\ell^{2}q^{(n+1)2\Delta},

which is larger than the trivial bound for the left-hand side whenever b2Δb\leq 2\Delta, as will occur in our ultimate choice for Δ\Delta. The transformation is nevertheless worthwhile, since we have passed from an incomplete (multiplicative) character sum on the left-hand side to a sum of complete (mixed) character sums on the right-hand side, to which we can apply Weil-Deligne bounds.

We will not try to average nontrivially over the characters χπi\chi_{\pi_{i}} or over x¯\underline{x}; for our current scope, it will suffice to prove a nontrivial bound for the individual sums SFS_{F}, which we will return to in §7.

6.1. Expansion in terms of multiplicative characters

We will now rewrite the unramified sieve term in (5.8) in terms of characters. Let us fix primes π1,π2𝒪K\pi_{1},\pi_{2}\in\mathcal{O}_{K} with π1π2\pi_{1}\neq\pi_{2} (the condition π1,π2𝒫\pi_{1},\pi_{2}\in{\mathcal{P}} need only be specified later). For any x¯𝒪Kn+1\underline{x}\in\mathcal{O}_{K}^{n+1}, we will rewrite each of the quantities (|ηπi1(x¯(modπi))|1)(\left|\eta_{\pi_{i}}^{-1}(\underline{x}\,(\operatorname{mod}\pi_{i}))\right|-1) as a character sum by using the following proposition, whose proof we defer to Section 6.4.

Proposition 6.2.

Let qq be an odd rational prime power and, as before, take K=𝔽q(T)K=\mathbb{F}_{q}(T). Fix a rational prime \ell with (q1)\ell\mid(q-1). For any prime π𝒪K\pi\in\mathcal{O}_{K} and any a𝒪Ka\in\mathcal{O}_{K}, we have

#{y(modπ)kπ:ay(modπ)}=χπχπ(a),\#\left\{y\,(\operatorname{mod}\pi)\in k_{\pi}:a\equiv y^{\ell}\,(\operatorname{mod}\pi)\right\}=\displaystyle\sum_{\chi_{\pi}}\chi_{\pi}(a),

where χπ\chi_{\pi} runs over all multiplicative characters on 𝒪K\mathcal{O}_{K} of modulus π\pi and order dividing \ell, whose definition is extended from kπk_{\pi}^{\ast} to 𝒪K\mathcal{O}_{K} via the rule

x𝒪Kwithπx,χπ(x):={0if χπ is non-principal,1if χπ is principal.\forall x\in\mathcal{O}_{K}\;\text{with}\;\pi\mid x,\;\;\chi_{\pi}(x):=\left\{\begin{array}[]{cl}0&\text{if $\chi_{\pi}$ is non-principal,}\\ 1&\text{if $\chi_{\pi}$ is principal.}\end{array}\right. (6.4)

By Proposition 6.2, for each of π1,π2\pi_{1},\pi_{2} and any x¯𝒪Kn+1\underline{x}\in\mathcal{O}_{K}^{n+1},

|ηπi1(x¯(modπi))|\displaystyle\left|\eta_{\pi_{i}}^{-1}(\underline{x}\,(\operatorname{mod}\pi_{i}))\right| =\displaystyle= #{y(modπi)kπi:F(x0,,xn)y(modπi)}\displaystyle\#\left\{y\,(\operatorname{mod}\pi_{i})\in k_{\pi_{i}}:F(x_{0},\ldots,x_{n})\equiv y^{\ell}\,(\operatorname{mod}\pi_{i})\right\} (6.5)
=\displaystyle= χπiχπi(F(x0,,xn)),\displaystyle\displaystyle\sum_{\chi_{\pi_{i}}}\chi_{\pi_{i}}(F(x_{0},\ldots,x_{n})),

where χπi\chi_{\pi_{i}} runs over all multiplicative characters on 𝒪K\mathcal{O}_{K} of modulus πi\pi_{i} and order dividing \ell, extended to 𝒪K\mathcal{O}_{K} as in (6.4). From this, we can write

|ηπi1(x¯(modπi))|1=χπiχ0χπi(F(x0,,xn)),\left|\eta_{\pi_{i}}^{-1}(\underline{x}\,(\operatorname{mod}\pi_{i}))\right|-1=\displaystyle\sum_{\chi_{\pi_{i}}\not=\chi_{0}}\chi_{\pi_{i}}(F(x_{0},\ldots,x_{n})), (6.6)

in which the sum is now over all (non-principal) characters of order \ell and modulus πi\pi_{i}.

6.2. Fourier expansion in terms of additive characters

Next, we perform a Fourier expansion in terms of the additive character defined in Section 3. We defer the proof of the required proposition to Section 6.5.

Proposition 6.3.

Let qq be an odd rational prime power and, as before, take K=𝔽q(T)K=\mathbb{F}_{q}(T). For all primes π,π𝒪K\pi,\pi^{\prime}\in\mathcal{O}_{K} with ππ\pi\neq\pi^{\prime}, for all χπ,χπ\chi_{\pi},\chi_{\pi^{\prime}} non-principal multiplicative characters of 𝒪K\mathcal{O}_{K} of moduli π\pi, π\pi^{\prime} (respectively), for all G𝒪K[X0,,Xn]G\in\mathcal{O}_{K}[X_{0},\ldots,X_{n}], and for all integers bb such that

0<b<degT(ππ),0<b<\operatorname{deg}_{T}(\pi\pi^{\prime}),

we have

x¯𝒪Kn+1degT(x¯)<bχπ(G(x¯))χπ(G(x¯))=q(n+1)(degT(ππ)b)x¯𝒪Kn+1degT(x¯)<degT(ππ)bSG(π¯x¯,χπ)SG(π¯x¯,χπ),\displaystyle\displaystyle\sum_{\begin{subarray}{c}\underline{x}\in\mathcal{O}_{K}^{n+1}\\ \operatorname{deg}_{T}(\underline{x})<b\end{subarray}}\chi_{\pi}(G(\underline{x}))\chi_{\pi^{\prime}}(G(\underline{x}))=q^{-(n+1)(\operatorname{deg}_{T}(\pi\pi^{\prime})-b)}\displaystyle\sum_{\begin{subarray}{c}\underline{x}\in\mathcal{O}_{K}^{n+1}\\ \operatorname{deg}_{T}(\underline{x})<\operatorname{deg}_{T}(\pi\pi^{\prime})-b\end{subarray}}S_{G}\left(\bar{\pi}^{\prime}\underline{x},\chi_{\pi}\right)\;S_{G}\left(\bar{\pi}\underline{x},\chi_{\pi^{\prime}}\right),

where π¯,π¯𝒪K\bar{\pi}^{\prime},\bar{\pi}^{\prime}\in\mathcal{O}_{K} are determined by the congruences ππ¯1(modπ),\pi\bar{\pi}\equiv 1\,(\operatorname{mod}\pi^{\prime}), ππ¯1(modπ),\pi^{\prime}\bar{\pi}^{\prime}\equiv 1\,(\operatorname{mod}\pi), and the sum SG(w¯,χπ)S_{G}(\underline{w},\chi_{\pi}) is defined in (6.1).

6.3. Application to the unramified sieve term

By (6.6) and Proposition 6.3, we observe that, for any primes π1,π2𝒫\pi_{1},\pi_{2}\in\mathcal{P} with π1π2\pi_{1}\neq\pi_{2},

x¯𝒪Kn+1degT(x¯)<b(|ηπ11(x¯(modπ1))|1)(|ηπ21(x¯(modπ2))|1)=q(n+1)(2Δb)χπ1χ0χπ2χ0x¯𝒪Kn+1degT(x¯)<2ΔbSF(π¯2x¯,χπ1)SF(π¯1x¯,χπ2).\displaystyle\sum_{\begin{subarray}{c}\underline{x}\in\mathcal{O}_{K}^{n+1}\\ \operatorname{deg}_{T}(\underline{x})<b\end{subarray}}\left(\left|\eta_{\pi_{1}}^{-1}(\underline{x}\,(\operatorname{mod}\pi_{1}))\right|-1\right)\left(\left|\eta_{\pi_{2}}^{-1}(\underline{x}\,(\operatorname{mod}\pi_{2}))\right|-1\right)\\ =q^{-(n+1)(2\Delta-b)}\displaystyle\sum_{\begin{subarray}{c}\chi_{\pi_{1}}\neq\chi_{0}\\ \chi_{\pi_{2}}\neq\chi_{0}\end{subarray}}\displaystyle\sum_{\begin{subarray}{c}\underline{x}\in\mathcal{O}_{K}^{n+1}\\ \operatorname{deg}_{T}(\underline{x})<2\Delta-b\end{subarray}}S_{F}\left(\bar{\pi}_{2}\underline{x},\chi_{\pi_{1}}\right)S_{F}\left(\bar{\pi}_{1}\underline{x},\chi_{\pi_{2}}\right). (6.7)

Here we must assume that

b<2Δb<2\Delta (6.8)

in order to apply Proposition 6.3. We keep this assumption from here onwards. Finally, recall the sum over x¯\underline{x} in (5.8) actually restricts to those x¯\underline{x} for which F(x¯)0(modπ1π2)F(\underline{x})\not\equiv 0\,(\operatorname{mod}\pi_{1}\pi_{2}). But as remarked immediately below (6.2), the left-hand side of (6.2) does not change when the restriction F(x¯)0(modπ1π2)F(\underline{x})\not\equiv 0\,(\operatorname{mod}\pi_{1}\pi_{2}) is removed. Thus, (6.7) is actually equal to the unramified sieve term. As such, Proposition 6.1 is proved, as long as we verify Proposition 6.2 and Proposition 6.3.

6.4. Proof of Proposition 6.2

As in the setting of the proposition, let qq be an odd rational prime power, take K=𝔽q(T)K=\mathbb{F}_{q}(T), let \ell be a rational prime with (q1)\ell\mid(q-1), and let π𝒪K\pi\in\mathcal{O}_{K} be a prime. Since (q1)\ell\mid(q-1), then (qdegT(π)1)=|kπ|\ell\mid(q^{\operatorname{deg}_{T}(\pi)}-1)=|k_{\pi}^{*}|. We can then consider the set of multiplicative characters

χπ:kπ\chi_{\pi}:k_{\pi}^{*}\rightarrow\mathbb{C}

of order dividing \ell on kπk_{\pi}^{*}, which forms a group of order \ell. We can extend χπ\chi_{\pi} to be defined over 𝒪K\mathcal{O}_{K} by setting χπ(a)=0\chi_{\pi}(a)=0 when πa\pi\mid a. These characters can be given explicitely by \ell-power symbols, as described in [Ros02, ch. 3]. It is known that

χπ(a)=1πa and Xa(modπ)is solvable;\chi_{\pi}(a)=1\;\Leftrightarrow\;\pi\nmid a\mbox{ and }X^{\ell}\equiv a\,(\operatorname{mod}\pi)\;\text{is solvable}; (6.9)

see [Ros02, Prop. 3.1, p. 24].

The character sums are given by

χπχπ(a)={1if πa,if πa and Xa(modπ) is solvable,0if πa and Xa(modπ) is not solvable,\displaystyle\sum_{\chi_{\pi}}\chi_{\pi}(a)=\left\{\begin{array}[]{cl}1&\text{if $\pi\mid a$,}\\ \ell&\text{if $\pi\nmid a$ and $X^{\ell}\equiv a\,(\operatorname{mod}\pi)$ is solvable,}\\ 0&\text{if $\pi\nmid a$ and $X^{\ell}\equiv a\,(\operatorname{mod}\pi)$ is not solvable,}\end{array}\right. (6.10)

where χπ\chi_{\pi} runs over all the multiplicative characters of 𝒪K\mathcal{O}_{K} of modulus π\pi and order dividing \ell. See [Ros02, Prop. 4.2, p. 35]. Consequently,

χπχπ(a)=#{y(modπ)kπ:ay(modπ)},\displaystyle\sum_{\chi_{\pi}}\chi_{\pi}(a)=\#\left\{y\,(\operatorname{mod}\pi)\in k_{\pi}:a\equiv y^{\ell}\,(\operatorname{mod}\pi)\right\},

which completes the proof of Proposition 6.2.

6.5. Proof of Proposition 6.3

As before, let qq be an odd rational prime power and take K=𝔽q(T)K=\mathbb{F}_{q}(T). We first prove a lemma about detecting congruences with the additive character ψ\psi_{\infty}.

Lemma 6.4.

Let u𝒪Ku\in\mathcal{O}_{K} and a¯𝒪Kn+1\underline{a}\in\mathcal{O}_{K}^{n+1}. Let bb be an integer such that

0<b<degT(u).0<b<\operatorname{deg}_{T}(u).

Then

x¯𝒪Kn+1degT(x¯)<bx¯a¯(modu)1=q(n+1)(degT(u)b)x¯𝒪Kn+1degT(x¯)<degT(u)bψ(x¯ua¯).\displaystyle\sum_{\begin{subarray}{c}\underline{x}\in\mathcal{O}_{K}^{n+1}\\ \operatorname{deg}_{T}(\underline{x})<b\\ \underline{x}\equiv\underline{a}\,(\operatorname{mod}u)\end{subarray}}1=q^{-(n+1)(\operatorname{deg}_{T}(u)-b)}\displaystyle\sum_{\begin{subarray}{c}\underline{x}\in\mathcal{O}_{K}^{n+1}\\ \operatorname{deg}_{T}(\underline{x})<\operatorname{deg}_{T}(u)-b\end{subarray}}\psi_{\infty}\left(-\frac{\underline{x}}{u}\cdot\underline{a}\right).
Proof.

We rewrite the left-hand side by using the following indicator functions:

w,b:𝒪K{0,1},w_{\infty,b}:\mathcal{O}_{K}\longrightarrow\{0,1\},
w,b(x):={1if degT(x)<b,0otherwise,w_{\infty,b}(x):=\left\{\begin{array}[]{cl}1&\text{if $\operatorname{deg}_{T}(x)<b$,}\\ 0&\text{otherwise},\end{array}\right.

and

wb:𝒪Kn+1{0,1},w_{b}:\mathcal{O}_{K}^{n+1}\longrightarrow\{0,1\},
wb(x¯):=1inw,b(xi).w_{b}(\underline{x}):=\displaystyle\prod_{1\leq i\leq n}w_{\infty,b}(x_{i}).

With these definitions, we write

x¯𝒪Kn+1degT(x¯)<bx¯a¯(modu)1=x¯𝒪Kn+1wb(a¯+x¯u).\displaystyle\sum_{\begin{subarray}{c}\underline{x}\in\mathcal{O}_{K}^{n+1}\\ \operatorname{deg}_{T}(\underline{x})<b\\ \underline{x}\equiv\underline{a}\,(\operatorname{mod}u)\end{subarray}}1=\displaystyle\sum_{\underline{x}^{\prime}\in\mathcal{O}_{K}^{n+1}}w_{b}(\underline{a}+\underline{x}^{\prime}u).

To understand the above sum, we use Fourier analysis on Kn+1K_{\infty}^{n+1}, as detailed in [BV15, Section 2]. By [BV15, Lemma 2.1], we obtain

x¯𝒪Kn+1wb(a¯+x¯u)\displaystyle\displaystyle\sum_{\underline{x}^{\prime}\in\mathcal{O}_{K}^{n+1}}w_{b}(\underline{a}+\underline{x}^{\prime}u) =\displaystyle= x¯′′𝒪Kn+1Kn+1wb(a¯+s¯u)ψ(x¯′′s¯)𝑑s¯\displaystyle\displaystyle\sum_{\underline{x}^{\prime\prime}\in\mathcal{O}_{K}^{n+1}}\displaystyle\int_{K_{\infty}^{n+1}}w_{b}(\underline{a}+\underline{s}u)\ \psi_{\infty}(\underline{x}^{\prime\prime}\cdot\underline{s})\ d\underline{s}
=\displaystyle= x¯′′𝒪Kn+1{s¯Kn+1:|a¯+s¯u|<qb}ψ(x¯′′s¯)𝑑s¯.\displaystyle\displaystyle\sum_{\underline{x}^{\prime\prime}\in\mathcal{O}_{K}^{n+1}}\displaystyle\int_{\left\{\underline{s}\in K_{\infty}^{n+1}:\ \left|\underline{a}+\underline{s}u\right|_{\infty}<q^{b}\right\}}\psi_{\infty}(\underline{x}^{\prime\prime}\cdot\underline{s})\ d\underline{s}.

Making the change of variables

t¯:=a¯+s¯u,\underline{t}:=\underline{a}+\underline{s}\ u,

for which

dt¯=q(n+1)degT(u)ds¯,d\underline{t}=q^{(n+1)\operatorname{deg}_{T}(u)}\ d\underline{s},

by [BV15, Lemma 2.3] we infer that

{s¯Kn+1:|a¯+s¯u|<qb}\displaystyle\displaystyle\int_{\left\{\underline{s}\in K_{\infty}^{n+1}:\ \left|\underline{a}+\underline{s}u\right|_{\infty}<q^{b}\right\}} ψ(x¯′′s¯)ds¯\displaystyle\psi_{\infty}(\underline{x}^{\prime\prime}\cdot\underline{s})\ d\underline{s}
=\displaystyle= q(n+1)degT(u){t¯Kn+1:|t¯|<qb}ψ(x¯′′t¯a¯u)𝑑t¯\displaystyle q^{-(n+1)\operatorname{deg}_{T}(u)}\displaystyle\int_{\left\{\underline{t}\in K_{\infty}^{n+1}:\ |\underline{t}|_{\infty}<q^{b}\right\}}\psi_{\infty}\left(\underline{x}^{\prime\prime}\cdot\frac{\underline{t}-\underline{a}}{u}\right)\ d\underline{t}
=\displaystyle= q(n+1)degT(u)ψ(x¯′′a¯u){t¯Kn+1:|t¯|<qb}ψ(x¯′′ut¯)𝑑t¯.\displaystyle q^{-(n+1)\operatorname{deg}_{T}(u)}\psi_{\infty}\left(-\frac{\underline{x}^{\prime\prime}\cdot\underline{a}}{u}\right)\displaystyle\int_{\left\{\underline{t}\in K_{\infty}^{n+1}:\ |\underline{t}|_{\infty}<q^{b}\right\}}\psi_{\infty}\left(\frac{\underline{x}^{\prime\prime}}{u}\cdot\underline{t}\right)\ d\underline{t}.

By [BV15, Lemma 2.2],

{t¯Kn+1:|t¯|<qb}ψ(x¯′′ut¯)𝑑t¯={q(n+1)bif |x¯′′u|<qb,0otherwise.\displaystyle\int_{\left\{\underline{t}\in K_{\infty}^{n+1}:\ |\underline{t}|_{\infty}<q^{b}\right\}}\psi_{\infty}\left(\frac{\underline{x}^{\prime\prime}}{u}\cdot\underline{t}\right)\ d\underline{t}=\left\{\begin{array}[]{cl}q^{(n+1)b}&\text{if $\left|\frac{\underline{x}^{\prime\prime}}{u}\right|_{\infty}<q^{-b}$,}\\ 0&\text{otherwise}.\end{array}\right.

Hence

x¯𝒪Kn+1wb(a¯+x¯u)\displaystyle\displaystyle\sum_{\underline{x}^{\prime}\in\mathcal{O}_{K}^{n+1}}w_{b}(\underline{a}+\underline{x}^{\prime}u) =\displaystyle= q(n+1)(degT(u)b)x¯′′𝒪Kn+1degT(x¯′′)<degT(u)bψ(x¯′′ua¯).\displaystyle q^{-(n+1)(\operatorname{deg}_{T}(u)-b)}\displaystyle\sum_{\begin{subarray}{c}\underline{x}^{\prime\prime}\in\mathcal{O}_{K}^{n+1}\\ \operatorname{deg}_{T}(\underline{x}^{\prime\prime})<\operatorname{deg}_{T}(u)-b\end{subarray}}\psi_{\infty}\left(-\frac{\underline{x}^{\prime\prime}}{u}\cdot\underline{a}\right).

Now we prove Proposition 6.3. Let π,π𝒪K\pi,\pi^{\prime}\in\mathcal{O}_{K} be primes with ππ\pi\neq\pi^{\prime}, let χπ,χπ\chi_{\pi},\chi_{\pi^{\prime}} be non-principal multiplicative characters of 𝒪K\mathcal{O}_{K} of moduli π\pi, π\pi^{\prime} (respectively), and let G𝒪K[X0,,Xn]G\in\mathcal{O}_{K}[X_{0},\ldots,X_{n}]. Fix an integer bb such that

0<b<degT(π1π2).0<b<\operatorname{deg}_{T}(\pi_{1}\pi_{2}).

We partition 𝒪Kn+1\mathcal{O}_{K}^{n+1} according to the residue classes modulo ππ\pi\pi^{\prime}:

x¯𝒪Kn+1degT(x¯)<bχπ(G(x¯))χπ(G(x¯))\displaystyle\displaystyle\sum_{\begin{subarray}{c}\underline{x}\in\mathcal{O}_{K}^{n+1}\\ \operatorname{deg}_{T}(\underline{x})<b\end{subarray}}\chi_{\pi}(G(\underline{x}))\chi_{\pi^{\prime}}(G(\underline{x})) =\displaystyle= a¯(modππ)(𝒪K/(ππ))n+1χπ(G(a¯))χπ(G(a¯))x¯𝒪Kn+1degT(x¯)<bx¯a¯(modππ)1.\displaystyle\displaystyle\sum_{\underline{a}\,(\operatorname{mod}\pi\pi^{\prime})\in(\mathcal{O}_{K}/(\pi\pi^{\prime}))^{n+1}}\chi_{\pi}(G(\underline{a}))\chi_{\pi^{\prime}}(G(\underline{a}))\displaystyle\sum_{\begin{subarray}{c}\underline{x}\in\mathcal{O}_{K}^{n+1}\\ \operatorname{deg}_{T}(\underline{x})<b\\ \underline{x}\equiv\underline{a}\,(\operatorname{mod}\pi\pi^{\prime})\end{subarray}}1.

By Lemma 6.4, the above double sum equals

q(n+1)(degT(ππ))ba¯(modππ)(𝒪K/(ππ))n+1χπ(G(a¯))χπ(G(a¯))x¯𝒪Kn+1degT(x¯)<degT(ππ)bψ(x¯ππa¯).q^{-(n+1)(\operatorname{deg}_{T}(\pi\pi^{\prime}))-b}\displaystyle\sum_{\underline{a}\,(\operatorname{mod}\pi\pi^{\prime})\in(\mathcal{O}_{K}/(\pi\pi^{\prime}))^{n+1}}\chi_{\pi}(G(\underline{a}))\chi_{\pi^{\prime}}(G(\underline{a}))\displaystyle\sum_{\begin{subarray}{c}\underline{x}\in\mathcal{O}_{K}^{n+1}\\ \operatorname{deg}_{T}(\underline{x})<\operatorname{deg}_{T}(\pi\pi^{\prime})-b\end{subarray}}\psi_{\infty}\left(-\frac{\underline{x}}{\pi\pi^{\prime}}\cdot\underline{a}\right). (6.11)

Recall that ππ\pi\neq\pi^{\prime}. Then, on one hand, there exist uniquely determined π¯(modπ)kπ\overline{\pi}^{\prime}\,(\operatorname{mod}\pi)\in k_{\pi}, π¯(modπ)kπ\overline{\pi}\,(\operatorname{mod}\pi^{\prime})\in k_{\pi^{\prime}} such that

ππ¯1(modπ),\pi\overline{\pi}\equiv 1\,(\operatorname{mod}\pi^{\prime}),
ππ¯1(modπ);\pi^{\prime}\overline{\pi}^{\prime}\equiv 1\,(\operatorname{mod}\pi);

on the other hand, by the Chinese Remainder Theorem, there exist uniquely determined elements a¯1(modπ)kπn+1\underline{a}_{1}\,(\operatorname{mod}\pi)\in k_{\pi}^{n+1}, a¯2(modπ)kπn+1\underline{a}_{2}\,(\operatorname{mod}\pi^{\prime})\in k_{\pi^{\prime}}^{n+1} such that

a¯a¯1π+a¯2π(modππ).\underline{a}\equiv\underline{a}_{1}\pi^{\prime}+\underline{a}_{2}\pi\,(\operatorname{mod}\pi\pi^{\prime}).

Consequently,

a¯(modππ)(𝒪K/(ππ))n+1χπ(G(a¯))χπ(G(a¯))x¯𝒪Kn+1degT(x¯)<degT(ππ)bψ(x¯ππa¯)\displaystyle\displaystyle\sum_{\underline{a}\,(\operatorname{mod}\pi\pi^{\prime})\in(\mathcal{O}_{K}/(\pi\pi^{\prime}))^{n+1}}\chi_{\pi}(G(\underline{a}))\chi_{\pi^{\prime}}(G(\underline{a}))\displaystyle\sum_{\begin{subarray}{c}\underline{x}\in\mathcal{O}_{K}^{n+1}\\ \operatorname{deg}_{T}(\underline{x})<\operatorname{deg}_{T}(\pi\pi^{\prime})-b\end{subarray}}\psi_{\infty}\left(-\frac{\underline{x}}{\pi\pi^{\prime}}\cdot\underline{a}\right)
=\displaystyle= a¯1(modπ)kπn+1a¯2(modπ)kπn+1χπ(G(a¯1π))χπ(G(a¯2π))x¯𝒪Kn+1degT(x¯)<degT(ππ)bψ(x¯a¯1π)ψ(x¯a¯2π).\displaystyle\displaystyle\sum_{\begin{subarray}{c}\underline{a}_{1}\,(\operatorname{mod}\pi)\in k_{\pi}^{n+1}\\ \underline{a}_{2}\,(\operatorname{mod}\pi^{\prime})\in k_{\pi^{\prime}}^{n+1}\end{subarray}}\chi_{\pi}\left(G(\underline{a}_{1}\pi^{\prime})\right)\chi_{\pi^{\prime}}\left(G(\underline{a}_{2}\pi)\right)\displaystyle\sum_{\begin{subarray}{c}\underline{x}\in\mathcal{O}_{K}^{n+1}\\ \operatorname{deg}_{T}(\underline{x})<\operatorname{deg}_{T}(\pi\pi^{\prime})-b\end{subarray}}\psi_{\infty}\left(-\frac{\underline{x}\cdot\underline{a}_{1}}{\pi}\right)\ \psi_{\infty}\left(-\frac{\underline{x}\cdot\underline{a}_{2}}{\pi^{\prime}}\right).

Using that ππ\pi\neq\pi^{\prime}, we make the change of variables a1¯a1¯π¯(modπ)\underline{a_{1}}\mapsto\underline{a_{1}}\overline{\pi}^{\prime}\,(\operatorname{mod}{\pi}) and a2¯a2¯π¯(modπ)\underline{a_{2}}\mapsto\underline{a_{2}}\overline{\pi}\,(\operatorname{mod}{\pi^{\prime}}) and deduce that the above expression equals

a¯1(modπ)kπn+1a¯2(modπ)kπn+1χπ(G(a¯1))χπ(G(a¯2))x¯𝒪Kn+1degT(x¯)<degT(ππ)bψ((π¯x¯)a¯1π)ψ((π¯x¯)a¯2π).\displaystyle\sum_{\begin{subarray}{c}\underline{a}_{1}\,(\operatorname{mod}\pi)\in k_{\pi}^{n+1}\\ \underline{a}_{2}\,(\operatorname{mod}\pi^{\prime})\in k_{\pi^{\prime}}^{n+1}\end{subarray}}\chi_{\pi}\left(G(\underline{a}_{1})\right)\chi_{\pi^{\prime}}\left(G(\underline{a}_{2})\right)\displaystyle\sum_{\begin{subarray}{c}\underline{x}\in\mathcal{O}_{K}^{n+1}\\ \operatorname{deg}_{T}(\underline{x})<\operatorname{deg}_{T}(\pi\pi^{\prime})-b\end{subarray}}\psi_{\infty}\left(-\frac{(\overline{\pi}^{\prime}\underline{x})\cdot\underline{a}_{1}}{\pi}\right)\ \psi_{\infty}\left(-\frac{(\overline{\pi}\underline{x})\cdot\underline{a}_{2}}{\pi^{\prime}}\right).

This completes the proof of Proposition 6.3.

7. The Weil-Deligne bounds

We now state and prove the Weil-Deligne estimates that will be used to bound the unramified sieve term in the form of the expansion in Proposition 6.1.

Proposition 7.1.

(Weil-Deligne bounds)

Let qq be the power of an odd rational prime and take K=𝔽q(T).K=\mathbb{F}_{q}(T). Let n1n\geq 1 be an integer and H𝒪K[X0,,Xn]H\in\mathcal{O}_{K}[X_{0},\ldots,X_{n}] a homogeneous polynomial of degree m2m\geq 2 such that the projective hypersurface defined by H(X0,,Xn)=0H(X_{0},\ldots,X_{n})=0 in K¯n\mathbb{P}^{n}_{\overline{K}} is nonsingular. Assume that char𝔽qm.\operatorname{char}\mathbb{F}_{q}\nmid m. Denote by WW, 𝒲{\mathcal{W}} the nonsingular projective hypersurface, respectively the affine hypersurface, defined by

H(X0,,Xn)=0.H(X_{0},\ldots,X_{n})=0.

Denote by 𝒫exc(H)𝒪K\mathcal{P}_{\mathrm{exc}}(H)\subseteq\mathcal{O}_{K} the finite set of exceptional primes introduced in Proposition 4.1. For a prime π𝒫exc(H)\pi\notin\mathcal{P}_{\mathrm{exc}}(H), denote by WπW_{\pi}, 𝒲π{\mathcal{W}}_{\pi} the nonsingular projective hypersurface, respectively the affine hypersurface, defined by

H(X0,,Xn)0(modπ),H(X_{0},\ldots,X_{n})\equiv 0\,(\operatorname{mod}\pi),

and denote by WπW_{\pi}^{\ast}, 𝒲π{\mathcal{W}}_{\pi}^{\ast} their duals. Fix a rational prime \ell with (q1)\ell\mid(q-1). Then, for every prime π𝒫exc(H)\pi\notin\mathcal{P}_{\mathrm{exc}}(H), for every (non-principal) multiplicative character χπ\chi_{\pi} of 𝒪K\mathcal{O}_{K} of modulus π\pi and order \ell, and for every w¯𝒪Kn+1\underline{w}\in\mathcal{O}_{K}^{n+1}, the mixed character sum SH(w¯,χπ)S_{H}(\underline{w},\chi_{\pi}) defined in (6.1) satisfies the bounds:

  1. (i)

    provided w¯0¯(modπ)\underline{w}\equiv\underline{0}\,(\operatorname{mod}\pi),

    |SH(w¯,χπ)|n,degX¯(H)q(n+2)degT(π)2;\left|S_{H}(\underline{w},\chi_{\pi})\right|\ll_{n,\operatorname{deg}_{\underline{X}}(H)}q^{\frac{(n+2)\operatorname{deg}_{T}(\pi)}{2}};
  2. (ii)

    provided w¯0¯(modπ)\underline{w}\not\equiv\underline{0}\,(\operatorname{mod}\pi),

    |SH(w¯,χπ)|n,degX¯(H)q(n+2)degT(π)2;\left|S_{H}(\underline{w},\chi_{\pi})\right|\ll_{n,\operatorname{deg}_{\underline{X}}(H)}q^{\frac{(n+2)\operatorname{deg}_{T}(\pi)}{2}};
  3. (iii)

    provided w¯0¯(modπ)\underline{w}\not\equiv\underline{0}\,(\operatorname{mod}\pi) and w¯𝒲π\underline{w}\not\in{\mathcal{W}}_{\pi}^{\ast},

    |SH(w¯,χπ)|n,degX¯(H)q(n+1)degT(π)2.\left|S_{H}(\underline{w},\chi_{\pi})\right|\ll_{n,\operatorname{deg}_{\underline{X}}(H)}q^{\frac{(n+1)\operatorname{deg}_{T}(\pi)}{2}}.

The proof of Proposition 7.1 is based on estimates for character sums of polynomials in several variables, pioneered by Deligne [Del74] and further generalized by Katz [Kat02], [Kat07], Rojas-León, and others. In the proof we will call upon the following definition:

Definition 7.2.

Let kk be a finite field and let d,r1d,r\geq 1 be integers. Let fk[X1,,Xr]f\in k[X_{1},\ldots,X_{r}] be a polynomial of degree dd, which we write as

f=fd+fd1++f0f=f_{d}+f_{d-1}+\cdots+f_{0}

for uniquely determined homogeneous polynomials fik[X1,,Xr]f_{i}\in k[X_{1},\ldots,X_{r}] with degX¯(fi)=i\operatorname{deg}_{\underline{X}}(f_{i})=i. We call ff a Deligne polynomial over kk if:

  1. (i)

    charkd\operatorname{char}k\nmid d;

  2. (ii)

    the equation fd=0f_{d}=0 defines a smooth, degree dd hypersurface in kr1\mathbb{P}_{k}^{r-1}.

Relative to the trivial bound for |SH(w¯,χπ)||S_{H}(\underline{w},\chi_{\pi})| given in (6.3), we see that the non-trivial bounds given in cases (i) and (ii) provide square-root cancellation in all but one of the n+1n+1 variables of the sum, while case (iii) provides square-root cancellation in all n+1n+1 variables. We think of cases (i) and (ii) as exceptional “zero” or “bad” cases, respectively, and of case (iii) as the “good” case. In our application in §8, we use that for a fixed prime π\pi, as the parameter w¯\underline{w} varies over 𝒪Kn+1\mathcal{O}_{K}^{n+1}, case (iii) is generic, with cases (i) and (ii) being rare.

7.1. Proof of part (i) of Proposition 7.1

We consider the case w¯0¯(modπ)\underline{w}\equiv\underline{0}\,(\operatorname{mod}\pi) and seek to bound

|SH(0¯,χπ)|=|a¯(modπ)kπn+1χπ(H(a¯))|.\left|S_{H}\left(\underline{0},\chi_{\pi}\right)\right|=\left|\displaystyle\sum_{\underline{a}\,(\operatorname{mod}\pi)\in k_{\pi}^{n+1}}\chi_{\pi}(H(\underline{a}))\right|. (7.1)

Our main tool is the following estimate for multiplicative character sums, which is a special case of a much more general result in [RL05].

Theorem 7.3.

(special case of [RL05, Thm. 1.1(a)])

Let kk be a finite field, r1r\geq 1 an integer, and χ:k\chi:k^{\ast}\longrightarrow\mathbb{C}^{\ast} a non-principal multiplicative character, extended to kk by χ(0):=0\chi(0):=0. Let Y=krY=\mathbb{P}^{r}_{k}, let Hk[X0,,Xr]H\in k[X_{0},\ldots,X_{r}] be a homogeneous polynomial of degree mm, and let Zk[X0,,Xr]Z\in k[X_{0},\ldots,X_{r}] be a homogeneous polynomial of degree ee. Assume that (e,m)=1(e,m)=1, (e,chark)=1(e,\mathrm{char}k)=1, and χe\chi^{e} is nontrivial. Let HH denote the hypersurface H=0H=0 in r\mathbb{P}^{r} and similarly let ZZ denote the hypersurface Z=0Z=0 in r\mathbb{P}^{r}. Assume that YHZY\cap H\cap Z has codimension 22 in YY, and let δ\delta denote the dimension of the singular locus of YHZrY\cap H\cap Z\subset\mathbb{P}^{r}. Set V=Y(HZ)V=Y-(H\cup Z) and define f:Vkf:V\rightarrow k^{\ast} by f(X¯)=H(X¯)e/Z(X¯)m.f(\underline{X})=H(\underline{X})^{e}/Z(\underline{X})^{m}. Then

|x¯V(k)χ(f(x¯))|3(3+e+m)r+2|k|r+δ+22.\left|\displaystyle\sum_{\underline{x}\in V(k)}\chi(f(\underline{x}))\right|\leq 3(3+e+m)^{r+2}|k|^{\frac{r+\delta+2}{2}}.
Corollary 7.4.

Let kk be a finite field, n1n\geq 1 an integer, and χ:k\chi:k^{\ast}\longrightarrow\mathbb{C}^{\ast} a non-principal multiplicative character, extended to kk by χ(0):=0\chi(0):=0. Let Hk[X0,,Xn]H\in k[X_{0},\ldots,X_{n}] be a homogeneous polynomial of degree m1m\geq 1 such that the projective hypersurface defined by H(X0,,Xn)=0H(X_{0},\ldots,X_{n})=0 in k¯n\mathbb{P}^{n}_{\overline{k}} is nonsingular. Then

|x¯kn+1χ(H(x¯))|3(m+4)n+3|k|n+22.\left|\displaystyle\sum_{\underline{x}\in k^{n+1}}\chi(H(\underline{x}))\right|\leq 3(m+4)^{n+3}|k|^{\frac{n+2}{2}}.
Proof of Corollary 7.4.

Note that H(x¯)H(\underline{x}) is not a well-defined map on a projective variety, since for example x¯=cx¯\underline{x}=c\underline{x} as a projective point, for each ckc\in k^{\ast}, while H(cx¯)=cmH(x¯)H(x¯)H(c\underline{x})=c^{m}H(\underline{x})\neq H(\underline{x}) unless cm=1c^{m}=1 in kk. The setting of Theorem 7.3 corrects for this, as follows. We set Y=n+1Y=\mathbb{P}^{n+1} in the variables X0,,Xn,TX_{0},\ldots,X_{n},T, so that r=n+1.r=n+1. We define H(X0,,Xn,T)=H(X0,,Xn)H(X_{0},\ldots,X_{n},T)=H(X_{0},\ldots,X_{n}) with degree mm and Z(X0,,Xn,T)=TZ(X_{0},\ldots,X_{n},T)=T so that e=1e=1. Then f(X¯)=H(X¯)/Tm=H(cx¯)/(cT)mf(\underline{X})=H(\underline{X})/T^{m}=H(c\underline{x})/(cT)^{m} for any ck,c\in k^{\ast}, so is well-defined as a polynomial map on V=n+1((H=0)(T=0)).V=\mathbb{P}^{n+1}-((H=0)\cup(T=0)). Furthermore, since χ(0)=0\chi(0)=0, and using the homogeneity described above,

(x¯,t)V(k)χ(f(x¯,t))=(x¯,t)kn+1(t=0)χ(H(x¯)/td)=(x¯,1)kn+1χ(H(x¯))=x¯kn+1χ(H(x¯)).\sum_{(\underline{x},t)\in V(k)}\chi(f(\underline{x},t))=\sum_{(\underline{x},t)\in\mathbb{P}_{k}^{n+1}\setminus(t=0)}\chi(H(\underline{x})/t^{d})=\sum_{(\underline{x},1)\in\mathbb{P}_{k}^{n+1}}\chi(H(\underline{x}))=\sum_{\underline{x}\in k^{n+1}}\chi(H(\underline{x})).

Note that

n+1(H=0)(T=0)=n(H=0),\mathbb{P}^{n+1}\cap(H=0)\cap(T=0)=\mathbb{P}^{n}\cap(H=0),

so that in the notation of the theorem, δ=1\delta=-1 since by assumption H=0H=0 is nonsingular as a projective hypersurface in n.\mathbb{P}^{n}. Moreover, the codimension of n+1(H=0)(T=0)\mathbb{P}^{n+1}\cap(H=0)\cap(T=0) in n+1\mathbb{P}^{n+1} is 2, as required. Hence by Theorem 7.3, the corollary holds.

The corollary immediately implies (i) in Proposition 7.1, upon taking HH as in the proposition, and any π𝒫exc(H)\pi\not\in\mathcal{P}_{\mathrm{exc}}(H), with k=kπk=k_{\pi} and χπ\chi_{\pi} as in the proposition.

7.2. Proof of part (ii) of Proposition 7.1

We consider the case w¯0¯(modπ)\underline{w}\not\equiv\underline{0}\,(\operatorname{mod}\pi) and seek to bound

|SH(w¯,χπ)|=|a¯(modπ)kπn+1χπ(H(a¯))ψ(w¯a¯π)|.\left|S_{H}\left(\underline{w},\chi_{\pi}\right)\right|=\left|\displaystyle\sum_{\underline{a}\,(\operatorname{mod}\pi)\in k_{\pi}^{n+1}}\chi_{\pi}(H(\underline{a}))\psi_{\infty}\left(-\frac{\underline{w}\cdot\underline{a}}{\pi}\right)\right|.

Our main tool is the following estimate for non-singular additive character sums, due to Deligne:

Theorem 7.5.

([Del74, Thm. 8.4 p. 302])

Let kk be a finite field, r1r\geq 1 an integer, and ψ:(k,+)(,)\psi:(k,+)\longrightarrow(\mathbb{C}^{\ast},\cdot) a non-trivial additive character. Let gk[X1,,Xr]g\in k[X_{1},\ldots,X_{r}] be a polynomial of degree d2d\geq 2. Assume that gg is a Deligne polynomial over kk. Then

|a¯krψ(g(a¯))|(d1)r|k|r2.\left|\displaystyle\sum_{\underline{a}\in k^{r}}\psi(g(\underline{a}))\right|\leq(d-1)^{r}|k|^{\frac{r}{2}}.

We apply Theorem 7.5 to the finite field kπk_{\pi}, the integer r=n+1r=n+1, the character ψ\psi_{\infty}, and each of qdegT(π)1q^{\operatorname{deg}_{T}(\pi)}-1 instances of polynomials gg, derived from HH, as explained in what follows.

Recall that the Gauss sum

τ(χπ):=αkπχπ(α)ψ(απ)\tau(\chi_{\pi}):=\displaystyle\sum_{\alpha\in k_{\pi}}\chi_{\pi}(\alpha)\psi_{\infty}\left(\frac{\alpha}{\pi}\right)

satisfies the Riemann Hypothesis (see, for example, [IK04, Prop 11.5 p. 275]):

|τ(χπ)|=qdegT(π)2.\left|\tau(\chi_{\pi})\right|=q^{\frac{\operatorname{deg}_{T}(\pi)}{2}}.

Note that for each a¯(modπ)kπn+1\underline{a}\,(\operatorname{mod}\pi)\in k_{\pi}^{n+1} such that H(a¯)0(modπ)H(\underline{a})\not\equiv 0\,(\operatorname{mod}\pi), we have a bijection

kπ\displaystyle k_{\pi}^{\ast} \displaystyle\longrightarrow kπ,\displaystyle k_{\pi}^{\ast},
α\displaystyle\alpha \displaystyle\mapsto αH(a¯)1.\displaystyle\alpha H(\underline{a})^{-1}.

Then, using properties of the Gauss sum (e.g. [IR90, Prop. 8.2.2 p. 92]), we obtain

SH(w¯,χπ)\displaystyle S_{H}(\underline{w},\chi_{\pi}) =\displaystyle= χπ(1)τ(χπ)τ(χπ¯)qdegT(π)a¯(modπ)kπn+1χπ(H(a¯))ψ(w¯a¯π)\displaystyle\frac{\chi_{\pi}(-1)\tau(\chi_{\pi})\tau\left(\overline{\chi_{\pi}}\right)}{q^{\operatorname{deg}_{T}(\pi)}}\displaystyle\sum_{\underline{a}\,(\operatorname{mod}\pi)\in k_{\pi}^{n+1}}\chi_{\pi}\left(H(\underline{a})\right)\psi_{\infty}\left(-\frac{\underline{w}\cdot\underline{a}}{\pi}\right) (7.2)
=\displaystyle= χπ(1)τ(χπ)τ(χπ¯)qdegT(π)a¯(modπ)kπn+1H(a¯)0(modπ)χπ(H(a¯))ψ(w¯a¯π)\displaystyle\frac{\chi_{\pi}(-1)\tau(\chi_{\pi})\tau\left(\overline{\chi_{\pi}}\right)}{q^{\operatorname{deg}_{T}(\pi)}}\displaystyle\sum_{\begin{subarray}{c}\underline{a}\,(\operatorname{mod}\pi)\in k_{\pi}^{n+1}\\ H(\underline{a})\not\equiv 0\,(\operatorname{mod}\pi)\end{subarray}}\chi_{\pi}\left(H(\underline{a})\right)\psi_{\infty}\left(-\frac{\underline{w}\cdot\underline{a}}{\pi}\right)
=\displaystyle= χπ(1)τ(χπ)qdegT(π)αkπa¯(modπ)kπn+1H(a¯)0(modπ)χπ(α1H(a¯))ψ(αw¯a¯π)\displaystyle\frac{\chi_{\pi}(-1)\tau(\chi_{\pi})}{q^{\operatorname{deg}_{T}(\pi)}}\displaystyle\sum_{\alpha\in k_{\pi}^{\ast}}\displaystyle\sum_{\begin{subarray}{c}\underline{a}\,(\operatorname{mod}\pi)\in k_{\pi}^{n+1}\\ H(\underline{a})\not\equiv 0\,(\operatorname{mod}\pi)\end{subarray}}\chi_{\pi}\left(\alpha^{-1}H(\underline{a})\right)\psi_{\infty}\left(\frac{\alpha-\underline{w}\cdot\underline{a}}{\pi}\right)
=\displaystyle= χπ(1)τ(χπ)qdegT(π)βkπχπ(β)a¯(modπ)kπn+1ψ(βH(a¯)w¯a¯π).\displaystyle\frac{\chi_{\pi}(-1)\tau(\chi_{\pi})}{q^{\operatorname{deg}_{T}(\pi)}}\displaystyle\sum_{\beta\in k_{\pi}^{\ast}}\chi_{\pi}(\beta)\displaystyle\sum_{\underline{a}\,(\operatorname{mod}\pi)\in k_{\pi}^{n+1}}\psi_{\infty}\left(\frac{\beta H(\underline{a})-\underline{w}\cdot\underline{a}}{\pi}\right).

In the last identity, in order to sum back in the contribution from H(a¯)0(modπ),H(\underline{a})\equiv 0\,(\operatorname{mod}{\pi}), we used that

βkπχπ(β)a¯(modπ)kπn+1H(a¯)0(modπ)ψ(βH(a¯)w¯a¯π)=(βkπχπ(β))(a¯(modπ)kπn+1H(a¯)0(modπ)ψ(w¯a¯π))=0,\displaystyle\sum_{\beta\in k_{\pi}^{\ast}}\chi_{\pi}(\beta)\displaystyle\sum_{\begin{subarray}{c}\underline{a}\,(\operatorname{mod}\pi)\in k_{\pi}^{n+1}\\ H(\underline{a})\equiv 0\,(\operatorname{mod}\pi)\end{subarray}}\psi_{\infty}\left(\frac{\beta H(\underline{a})-\underline{w}\cdot\underline{a}}{\pi}\right)=\left(\displaystyle\sum_{\beta\in k_{\pi}^{\ast}}\chi_{\pi}(\beta)\right)\left(\displaystyle\sum_{\begin{subarray}{c}\underline{a}\,(\operatorname{mod}\pi)\in k_{\pi}^{n+1}\\ H(\underline{a})\equiv 0\,(\operatorname{mod}\pi)\end{subarray}}\psi_{\infty}\left(\frac{-\underline{w}\cdot\underline{a}}{\pi}\right)\right)=0,

which follows from the orthogonality of the characters χπ.\chi_{\pi}. By taking absolute values in (7.2), we deduce that

|SH(w¯,χπ)|qdegT(π)2βkπ|a¯kπn+1ψ(βH(a¯)w¯a¯π)|.\displaystyle\left|S_{H}(\underline{w},\chi_{\pi})\right|\leq q^{-\frac{\operatorname{deg}_{T}(\pi)}{2}}\displaystyle\sum_{\beta\in k_{\pi}^{\ast}}\left|\displaystyle\sum_{\underline{a}\in k_{\pi}^{n+1}}\psi_{\infty}\left(\frac{\beta H(\underline{a})-\underline{w}\cdot\underline{a}}{\pi}\right)\right|. (7.3)

We estimate the inner sum above using Theorem 7.5 for each of the polynomials over kπk_{\pi} defined by the congruence

g(X0,,Xn)βH(X0,,Xn)w¯(X0,,Xn)(modπ)g(X_{0},\ldots,X_{n})\equiv\beta H(X_{0},\ldots,X_{n})-\underline{w}\cdot(X_{0},\ldots,X_{n})\,(\operatorname{mod}\pi)

and for the additive character ψ\psi_{\infty}. Using that β0\beta\neq 0, degX¯(H)=m2\operatorname{deg}_{\underline{X}}(H)=m\geq 2, and π𝒫exc(H)\pi\not\in{\mathcal{P}}_{\text{exc}}(H), we obtain that degX¯(g)=m\operatorname{deg}_{\underline{X}}(g)=m. Let us write g=g0++gmg=g_{0}+\cdots+g_{m}, where g0g_{0}, ,\ldots, gmg_{m} are the uniquely determined homogeneous polynomials in kπ[X0,,Xn]k_{\pi}[X_{0},\ldots,X_{n}] such that degX¯(gi)=i\operatorname{deg}_{\underline{X}}(g_{i})=i. Then gmβH(modπ)g_{m}\equiv\beta H\,(\operatorname{mod}\pi). By hypothesis, H=0H=0 is nonsingular in K¯n\mathbb{P}^{n}_{\overline{K}} and π𝒫exc(H)\pi\not\in\mathcal{P}_{\mathrm{exc}}(H) so that WπW_{\pi} is nonsingular; also char𝔽qdeg(H)\mathrm{char}\;\mathbb{F}_{q}\nmid\operatorname{deg}(H). Thus H(modπ)H(\operatorname{mod}\pi) is a Deligne polynomial over kπk_{\pi}, and hence gg also is. By (7.3) and Theorem 7.5, we deduce that

|SH(w¯,χπ)|qdegT(π)2βkπ(m1)n+1q(n+1)degT(π)2<(m1)n+1q(n+2)degT(π)2,\displaystyle\left|S_{H}(\underline{w},\chi_{\pi})\right|\leq q^{-\frac{\operatorname{deg}_{T}(\pi)}{2}}\displaystyle\sum_{\beta\in k_{\pi}^{\ast}}(m-1)^{n+1}q^{\frac{(n+1)\operatorname{deg}_{T}(\pi)}{2}}<(m-1)^{n+1}q^{\frac{(n+2)\operatorname{deg}_{T}(\pi)}{2}},

which completes the proof for case (ii).

7.3. Proof of part (iii) of Proposition 7.1

We consider the case w¯0¯(modπ)\underline{w}\not\equiv\underline{0}\,(\operatorname{mod}\pi), w¯𝒲π\underline{w}\not\in{\mathcal{W}}_{\pi}^{\ast}, and seek to bound

|SH(w¯,χπ)|=|a¯(modπ)kπn+1χπ(H(a¯))ψ(w¯a¯π)|\left|S_{H}\left(\underline{w},\chi_{\pi}\right)\right|=\left|\displaystyle\sum_{\underline{a}\,(\operatorname{mod}\pi)\in k_{\pi}^{n+1}}\chi_{\pi}(H(\underline{a}))\psi_{\infty}\left(-\frac{\underline{w}\cdot\underline{a}}{\pi}\right)\right|

in such a way that we improve upon the bound in part (ii). Our main tool is the following estimate for non-singular mixed character sums, again due to Katz:

Theorem 7.6.

([Kat07, Thm. 1.1 p. 3])

Let kk be a finite field, r1r\geq 1 an integer, χ:(k,)(,)\chi:(k^{\ast},\cdot)\longrightarrow(\mathbb{C}^{\ast},\cdot) a non-principal multiplicative character, extended to kk by χ(0):=0\chi(0):=0, and ψ:(k,+)(,)\psi:(k,+)\longrightarrow(\mathbb{C}^{\ast},\cdot) a non-trivial additive character. Let f,gk[X1,,Xr]f,g\in k[X_{1},\ldots,X_{r}] be polynomials of degrees d,e1d,e\geq 1 with leading homogeneous forms fdf_{d} of degree dd, geg_{e} of degree ee (respectively). Assume that:

  1. (i)

    ff is a Deligne polynomial;

  2. (ii)

    gg is a Deligne polynomial;

  3. (iii)

    in case r2r\geq 2, the smooth hypersurfaces in kr1\mathbb{P}_{k}^{r-1} defined by fd=0f_{d}=0 and by ge=0g_{e}=0 are transverse, in the sense that their intersection is smooth and of codimension 22 in kr1\mathbb{P}_{k}^{r-1}.

Then

|a¯krχ(f(a¯))ψ(g(a¯))|r,d,e|k|r2.\left|\displaystyle\sum_{\underline{a}\in k^{r}}\chi(f(\underline{a}))\psi(g(\underline{a}))\right|\ll_{r,d,e}|k|^{\frac{r}{2}}.

We apply Theorem 7.6 to the finite field kπk_{\pi}, the integer r=n+1r=n+1, the characters ψ\psi_{\infty}, χπ\chi_{\pi}, and the polynomials over kπk_{\pi} defined by the congruences

f(X0,,Xn)\displaystyle f(X_{0},\ldots,X_{n}) \displaystyle\equiv H(X0,,Xn)(modπ),\displaystyle H(X_{0},\ldots,X_{n})\,(\operatorname{mod}\pi),
g(X0,,Xn)\displaystyle g(X_{0},\ldots,X_{n}) \displaystyle\equiv w0X0w1X1wnXn(modπ).\displaystyle-w_{0}X_{0}-w_{1}X_{1}-\cdots-w_{n}X_{n}\,(\operatorname{mod}\pi).

Note that f(modπ)f(\operatorname{mod}\pi) is homogeneous of degree degX¯(H)=m2\operatorname{deg}_{\underline{X}}(H)=m\geq 2 (because π𝒫exc(H)\pi\not\in{\mathcal{P}}_{\text{exc}}(H)) and that g(modπ)g(\operatorname{mod}\pi) is homogeneous of degree 11 (because w¯0(modπ)\underline{w}\not\equiv 0\,(\operatorname{mod}\pi)). Recalling that char𝔽qdegX¯(H)\mathrm{char}\,\mathbb{F}_{q}\nmid\operatorname{deg}_{\underline{X}}(H) by hypothesis, we deduce that f(modπ)f(\operatorname{mod}\pi) and g(modπ)g(\operatorname{mod}\pi) are indeed Deligne polynomials. Since w¯𝒲π\underline{w}\not\in{\mathcal{W}}_{\pi}^{\ast}, they are also transverse. Thus Theorem 7.6 applies, giving

|SH(w¯,χπ)|n,mq(n+1)degT(π)2.\left|S_{H}(\underline{w},\chi_{\pi})\right|\ll_{n,m}q^{\frac{(n+1)\operatorname{deg}_{T}(\pi)}{2}}.

With this, we completed the verification of all Weil-Deligne bounds of Proposition 7.1.

8. Application of the Weil-Deligne bounds to the unramified sieve term

In this section, we apply the Weil-Deligne bounds of §7 to the unramified sieve term (given by the right hand side of (6.2)).

Proposition 8.1.

Let qq be an odd rational prime power, n2n\geq 2 an integer, 2\ell\geq 2 a rational prime, and F𝒪K[X0,,Xn]F\in\mathcal{O}_{K}[X_{0},\ldots,X_{n}] a homogeneous polynomial of degree m2m\geq 2 in X0,,XnX_{0},\ldots,X_{n}, with charKm\mathrm{char}K\nmid m, where, as before, K=𝔽q(T)K=\mathbb{F}_{q}(T). Assume the conditions:

  1. (i)

    gcd(m,q1)\ell\mid\operatorname{gcd}(m,q-1);

  2. (ii)

    the projective hypersurface F(X0,,Xn)=0K¯nF(X_{0},\ldots,X_{n})=0\subset\mathbb{P}^{n}_{\overline{K}} is nonsingular.

Let b,Δ>0b,\Delta>0 be integers such that

Δ<b<2Δ.\Delta<b<2\Delta. (8.1)

Defining 𝒫\mathcal{P} as in (5.3),

q(n+1)(2Δb)|𝒫|2π1,π2𝒫π1π2χπ1χ0χπ2χ0x¯𝒪Kn+1degT(x¯)<2Δb|SF(π¯2x¯,χπ1)SF(π¯1x¯,χπ2)|,m,ndegT(F)b2qnΔ+q(n+1)Δ+q(n+1)bΔ.\frac{q^{-(n+1)(2\Delta-b)}}{|\mathcal{P}|^{2}}\sum_{\begin{subarray}{c}\pi_{1},\pi_{2}\in{\mathcal{P}}\\ \pi_{1}\neq\pi_{2}\end{subarray}}\displaystyle\sum_{\begin{subarray}{c}\chi_{\pi_{1}}\neq\chi_{0}\\ \chi_{\pi_{2}}\neq\chi_{0}\end{subarray}}\displaystyle\sum_{\begin{subarray}{c}\underline{x}\in\mathcal{O}_{K}^{n+1}\\ \operatorname{deg}_{T}(\underline{x})<2\Delta-b\end{subarray}}|S_{F}\left(\bar{\pi}_{2}\underline{x},\chi_{\pi_{1}}\right)S_{F}\left(\bar{\pi}_{1}\underline{x},\chi_{\pi_{2}}\right)|\\ \ll_{\ell,m,n\operatorname{deg}_{T}(F^{*})}b^{2}q^{n\Delta}+q^{(n+1)\Delta}+q^{(n+1)b-\Delta}. (8.2)

To prove Proposition 8.1, we will estimate the absolute value of the innermost term according to different cases of x¯\underline{x} suggested by Proposition 7.1.

8.1. Proof of Proposition 8.1: dissecting the inner sum into cases

Let π1,π2𝒫\pi_{1},\pi_{2}\in\mathcal{P} with π1π2\pi_{1}\neq\pi_{2} and let χπ1,χπ2χ0\chi_{\pi_{1}},\chi_{\pi_{2}}\neq\chi_{0} be fixed. We dissect the set {x¯𝒪Kn+1:degT(x¯)<2Δb}\{\underline{x}\in\mathcal{O}_{K}^{n+1}:\operatorname{deg}_{T}(\underline{x})<2\Delta-b\} into subsets suggested by Proposition 7.1, as follows.

For each i{1,2}i\in\{1,2\}, we consider the smooth projective hypersurface

W(F)πi:F(X0,,Xn)0(modπi).W(F)_{\pi_{i}}:\quad F(X_{0},\ldots,X_{n})\equiv 0\,(\operatorname{mod}\pi_{i}).

By part (iii) of Proposition 4.1, the reduction modulo πi\pi_{i} of FF^{\ast} remains absolutely irreducible and defines the projective dual W(F)πiW(F)_{\pi_{i}}^{\ast}, that is,

W(F)π:F(X0,,Xn)0(modπi).W(F)_{\pi}^{\ast}:\quad F^{\ast}(X_{0},\ldots,X_{n})\equiv 0\,(\operatorname{mod}\pi_{i}).

We denote by

𝒲(F)πi,𝒲(F)πi{\mathcal{W}}(F)_{\pi_{i}},\;{\mathcal{W}}(F)_{\pi_{i}}^{\ast}

the affine varieties defined by the homogenous polynomials F(modπi)F\,(\operatorname{mod}\pi_{i}), F(modπi)F^{\ast}\,(\operatorname{mod}\pi_{i}) (respectively). We partition {x¯𝒪Kn+1:degT(x¯)<2Δb}\{\underline{x}\in\mathcal{O}_{K}^{n+1}:\operatorname{deg}_{T}(\underline{x})<2\Delta-b\} according to the following cases:

  1. (C1)

    π¯1x¯𝒲(F)π2\bar{\pi}_{1}\underline{x}\not\in{\mathcal{W}}(F)_{\pi_{2}}^{\ast}, π¯2x¯𝒲(F)π1\bar{\pi}_{2}\underline{x}\not\in{\mathcal{W}}(F)_{\pi_{1}}^{\ast} (good-good);

  2. (C2)

    π¯1x¯𝒲(F)π2\bar{\pi}_{1}\underline{x}\not\in{\mathcal{W}}(F)_{\pi_{2}}^{\ast}, π¯2x¯𝒲(F)π1\bar{\pi}_{2}\underline{x}\in{\mathcal{W}}(F)_{\pi_{1}}^{\ast} (good-bad);

  3. (C2’)

    π¯1x¯𝒲(F)π2\bar{\pi}_{1}\underline{x}\in{\mathcal{W}}(F)_{\pi_{2}}^{\ast}, π¯2x¯𝒲(F)π1\bar{\pi}_{2}\underline{x}\not\in{\mathcal{W}}(F)_{\pi_{1}}^{\ast} (bad-good);

  4. (C3)

    π¯1x¯𝒲(F)π2\bar{\pi}_{1}\underline{x}\in{\mathcal{W}}(F)_{\pi_{2}}^{\ast}, π¯2x¯𝒲(F)π1\bar{\pi}_{2}\underline{x}\in{\mathcal{W}}(F)_{\pi_{1}}^{\ast} (bad-bad).

In the above partition, the case π¯1x¯𝒲(F)π2\bar{\pi}_{1}\underline{x}\in{\mathcal{W}}(F)_{\pi_{2}}^{\ast} includes the subcase in which π¯1x¯0¯(modπ2)\bar{\pi}_{1}\underline{x}\equiv\underline{0}\,(\operatorname{mod}{\pi_{2}}); similarly, the case π¯2x¯𝒲(F)π1\bar{\pi}_{2}\underline{x}\in{\mathcal{W}}(F)_{\pi_{1}}^{\ast} includes the subcase in which π¯2x¯0¯(modπ1)\bar{\pi}_{2}\underline{x}\equiv\underline{0}\,(\operatorname{mod}{\pi_{1}}). Thus the condition π¯1x¯𝒲(F)π2\bar{\pi}_{1}\underline{x}\not\in{\mathcal{W}}(F)_{\pi_{2}}^{\ast} implies that π¯1x¯0¯(modπ2)\bar{\pi}_{1}\underline{x}\not\equiv\underline{0}\,(\operatorname{mod}{\pi_{2}}); similarly, π¯2x¯𝒲(F)π1\bar{\pi}_{2}\underline{x}\not\in{\mathcal{W}}(F)_{\pi_{1}}^{\ast} implies that π¯2x¯0¯(modπ1)\bar{\pi}_{2}\underline{x}\not\equiv\underline{0}\,(\operatorname{mod}{\pi_{1}}). We will treat each case separately; we begin with the bad-bad case (C3), which is most difficult. Here we will crucially use nontrivial averaging over the primes π1,π2\pi_{1},\pi_{2} in order to produce an efficient upper bound. The good-good case sums over x¯\underline{x} and π\pi more trivially, but this is allowable because of the square-root cancellation achieved in case (iii) of Proposition 7.1. The strategy for bounding the good-bad case is a hybrid of these two methods. (The case (C2’) is analogous to the case (C2), and thus we only explicitly describe the treatment of (C2).)

8.2. Proof of Proposition 8.1: the bad-bad case (C3)

We break the left-hand side of (8.2) into two sums Σ1+Σ2,\Sigma_{1}+\Sigma_{2}, according to whether the sum of x¯𝒪Kn+1\underline{x}\in\mathcal{O}_{K}^{n+1} takes place over the x¯\underline{x} satisfying F(x¯)0F^{*}(\underline{x})\neq 0 or F(x¯)=0F^{*}(\underline{x})=0; that is to say,

Σ1\displaystyle\Sigma_{1} =q(n+1)(2Δb)|𝒫|2π1,π2𝒫π1π2χπ1χ0χπ2χ0x¯𝒪Kn+1degT(x¯)<2ΔbF(x¯)0, case (C3)|SF(π¯2x¯,χπ1)SF(π¯1x¯,χπ2)|\displaystyle=\frac{q^{-(n+1)(2\Delta-b)}}{|\mathcal{P}|^{2}}\sum_{\begin{subarray}{c}\pi_{1},\pi_{2}\in{\mathcal{P}}\\ \pi_{1}\neq\pi_{2}\end{subarray}}\displaystyle\sum_{\begin{subarray}{c}\chi_{\pi_{1}}\neq\chi_{0}\\ \chi_{\pi_{2}}\neq\chi_{0}\end{subarray}}\displaystyle\sum_{\begin{subarray}{c}\underline{x}\in\mathcal{O}_{K}^{n+1}\\ \operatorname{deg}_{T}(\underline{x})<2\Delta-b\\ \text{$F^{*}(\underline{x})\neq 0$, case (C3)}\end{subarray}}|S_{F}\left(\bar{\pi}_{2}\underline{x},\chi_{\pi_{1}}\right)S_{F}\left(\bar{\pi}_{1}\underline{x},\chi_{\pi_{2}}\right)|
Σ2\displaystyle\Sigma_{2} =q(n+1)(2Δb)|𝒫|2π1,π2𝒫π1π2χπ1χ0χπ2χ0x¯𝒪Kn+1degT(x¯)<2ΔbF(x¯)=0, case (C3)|SF(π¯2x¯,χπ1)SF(π¯1x¯,χπ2)|.\displaystyle=\frac{q^{-(n+1)(2\Delta-b)}}{|\mathcal{P}|^{2}}\sum_{\begin{subarray}{c}\pi_{1},\pi_{2}\in{\mathcal{P}}\\ \pi_{1}\neq\pi_{2}\end{subarray}}\displaystyle\sum_{\begin{subarray}{c}\chi_{\pi_{1}}\neq\chi_{0}\\ \chi_{\pi_{2}}\neq\chi_{0}\end{subarray}}\displaystyle\sum_{\begin{subarray}{c}\underline{x}\in\mathcal{O}_{K}^{n+1}\\ \operatorname{deg}_{T}(\underline{x})<2\Delta-b\\ \text{$F^{*}(\underline{x})=0,$ case (C3)}\end{subarray}}|S_{F}\left(\bar{\pi}_{2}\underline{x},\chi_{\pi_{1}}\right)S_{F}\left(\bar{\pi}_{1}\underline{x},\chi_{\pi_{2}}\right)|.

Within each term, the subscript case (C3) means that we restrict to those π1,π2,x¯\pi_{1},\pi_{2},\underline{x} such that case (C3) holds. In Σ1,\Sigma_{1}, we note that by applying cases (i) and (ii) of Proposition 7.1,

|Σ1|,m,nq(n+1)(2Δb)q(n+2)Δ|𝒫|2x¯𝒪Kn+1degT(x¯)<2ΔbF(x¯)0π1,π2𝒫π1π2case (C3)1.|\Sigma_{1}|\ll_{\ell,m,n}\frac{q^{-(n+1)(2\Delta-b)}q^{(n+2)\Delta}}{|\mathcal{P}|^{2}}\displaystyle\sum_{\begin{subarray}{c}\underline{x}\in\mathcal{O}_{K}^{n+1}\\ \operatorname{deg}_{T}(\underline{x})<2\Delta-b\\ F^{*}(\underline{x})\neq 0\end{subarray}}\sum_{\begin{subarray}{c}\pi_{1},\pi_{2}\in{\mathcal{P}}\\ \pi_{1}\neq\pi_{2}\\ \text{case (C3)}\end{subarray}}1. (8.3)

Here we used that for each πi\pi_{i}, the number of characters of order \ell modulo πi\pi_{i} is \ell. To bound this efficiently, we will use the fact that F(x¯)0F^{*}(\underline{x})\neq 0 in order to show that relatively few pairs of π1,π2\pi_{1},\pi_{2} can correspond to the bad-bad case.

In contrast, in Σ2\Sigma_{2}, since F(x¯)=0F^{*}(\underline{x})=0, then certainly x¯\underline{x} is “bad” for all πi\pi_{i}, since F(x¯)0(modπ1)F^{*}(\underline{x})\equiv 0\,(\operatorname{mod}{\pi_{1}}) for all πi𝒫\pi_{i}\in\mathcal{P}. Thus, by applying cases (i) and (ii) of Proposition 7.1, we write

|Σ2|,m,nq(n+1)(2Δb)q(n+2)Δx¯𝒪Kn+1degT(x¯)<2ΔbF(x¯)=01.|\Sigma_{2}|\ll_{\ell,m,n}q^{-(n+1)(2\Delta-b)}q^{(n+2)\Delta}\displaystyle\sum_{\begin{subarray}{c}\underline{x}\in\mathcal{O}_{K}^{n+1}\\ \operatorname{deg}_{T}(\underline{x})<2\Delta-b\\ F^{*}(\underline{x})=0\end{subarray}}1. (8.4)

The heart of the argument in the bad-bad case is thus to count efficiently those x¯\underline{x} for which F(x¯)=0F^{*}(\underline{x})=0. We return to this momentarily.

8.2.1. Bounding Σ1\Sigma_{1}

To bound Σ1\Sigma_{1}, we begin with (8.3). Case (C3) requires that π¯1x¯𝒲(F)π2\bar{\pi}_{1}\underline{x}\in\mathcal{W}(F)_{\pi_{2}}^{*}, so that by homogeneity π2|F(x¯)\pi_{2}|F^{*}(\underline{x}), and analogously π1|F(x¯)\pi_{1}|F^{*}(\underline{x}). Consequently, the innermost sum over π1π2\pi_{1}\neq\pi_{2} in (8.3) is bounded by

#{π1π2𝒫:π1π2|F(x¯)}(ω(F(x¯)))2,\#\{\pi_{1}\neq\pi_{2}\in\mathcal{P}:\pi_{1}\pi_{2}|F^{*}(\underline{x})\}\leq(\omega(F^{*}(\underline{x})))^{2},

where we let ω(y)\omega(y) denote the number of distinct prime divisors of an element y𝒪Ky\in\mathcal{O}_{K}. Now let degT(F)\operatorname{deg}_{T}(F^{*}) denote the largest degree of TT that appears in a coefficient of F.F^{*}. Then for x¯𝒪Kn+1\underline{x}\in\mathcal{O}_{K}^{n+1} with degT(x¯)<2Δb\operatorname{deg}_{T}(\underline{x})<2\Delta-b and F(x¯)0F^{*}(\underline{x})\neq 0,

ω(F(x¯))degT(F)+degX¯FdegT(x¯)degT(F)+m(m1)n1(2Δb),\omega(F^{*}(\underline{x}))\leq\operatorname{deg}_{T}(F^{*})+\operatorname{deg}_{\underline{X}}F^{*}\cdot\operatorname{deg}_{T}(\underline{x})\leq\operatorname{deg}_{T}(F^{*})+m(m-1)^{n-1}(2\Delta-b), (8.5)

where in the last inequality we applied Proposition 11.2 (3) to bound degX¯(F)\operatorname{deg}_{\underline{X}}(F^{*}). We also note that 2Δb<b2\Delta-b<b under the hypothesis (8.1). In conclusion,

|Σ1|,m,n,degT(F)b2q(n+1)(2Δb)q2Δq(n+2)Δq(n+1)(2Δb),m,n,degT(F)b2qnΔ.|\Sigma_{1}|\ll_{\ell,m,n,\operatorname{deg}_{T}(F^{*})}b^{2}q^{-(n+1)(2\Delta-b)}q^{-2\Delta}q^{(n+2)\Delta}q^{(n+1)(2\Delta-b)}\ll_{\ell,m,n,\operatorname{deg}_{T}(F^{*})}b^{2}q^{n\Delta}.

8.2.2. Bounding Σ2\Sigma_{2}

Now we turn to bounding Σ2\Sigma_{2}, starting from (8.4). The following lemma is the main tool for bounding Σ2\Sigma_{2} in the bad-bad case (C3), as well as for bounding an analogous sum in the good-bad cases (C2) and (C2’); we defer its proof to §8.5. (It is also possible to apply [BV15, Lemma 2.9]; we nevertheless include the more flexible lemma below, in case of independent interest.)

Lemma 8.2.

Let G𝒪K[X0,,Xn]G\in\mathcal{O}_{K}[X_{0},\dots,X_{n}] be an irreducible homogeneous polynomial of degree degX¯G2\operatorname{deg}_{\underline{X}}G\geq 2, and let LN1L\geq N\geq 1 be such that there is an irreducible polynomial π\pi of degree LL such that G(X¯)G(\underline{X}) remains irreducible in kπk_{\pi}. Then

x¯𝒪Kn+1degT(x¯)<NG(x¯)=01n,degX¯(G)q(n+1)NL+q(n1)L.\sum_{\begin{subarray}{c}\underline{x}\in\mathcal{O}_{K}^{n+1}\\ \operatorname{deg}_{T}(\underline{x})<N\\ G(\underline{x})=0\end{subarray}}1\ll_{n,\operatorname{deg}_{\underline{X}}(G)}q^{(n+1)N-L}+q^{(n-1)L}.

We remark that arguing as in the proof of Proposition 4.1, Part (iii.1) implies that for all but finitely many π\pi, G(X¯)G(\underline{X}) remains irreducible in kπk_{\pi}. Thus, choosing LL sufficiently large relative to a fixed polynomial GG of interest, the conditions of Lemma 8.2 will be met.

We are now ready to bound Σ2\Sigma_{2}. Apply Lemma 8.2 to (8.4) with G=FG=F^{*} ( recall degX¯(F)m,n1\operatorname{deg}_{\underline{X}}(F^{*})\ll_{m,n}1 by Proposition 11.2 (3)), and the choices N=2ΔbN=2\Delta-b and L=ΔL=\Delta (note that NLN\leq L by (8.1)). We obtain

|Σ2|,m,n\displaystyle|\Sigma_{2}|\ll_{\ell,m,n} q(n+1)(b2Δ)q(n+2)Δ(q(n+1)(2Δb)Δ+q(n1)Δ)\displaystyle q^{(n+1)(b-2\Delta)}q^{(n+2)\Delta}\left(q^{(n+1)(2\Delta-b)-\Delta}+q^{(n-1)\Delta}\right)
,m,n\displaystyle\ll_{\ell,m,n} q(n+1)Δ+q(n+1)bΔ.\displaystyle q^{(n+1)\Delta}+q^{(n+1)b-\Delta}.

This is valid as long as there is a prime π𝒪K\pi\in\mathcal{O}_{K} of degree Δ\Delta for which FF^{*} is irreducible modulo π\pi. Under the hypothesis of Proposition 8.1, this this will be true for all Δ\Delta sufficiently large, say

ΔL0(F)\Delta\geq L_{0}(F^{*}) (8.6)

for a finite parameter provided by Proposition 4.1, Part (iii.1); we will ensure this with our final choice of Δ\Delta in (9.4); see §9. Combining the bounds for Σ1\Sigma_{1} and Σ2\Sigma_{2}, we obtain that (under (8.6)) the total contribution of the bad-bad case (C3) to the left-hand side of (8.2) is

,m,n,degT(F)b2qnΔ+q(n+1)Δ+q(n+1)bΔ.\ll_{\ell,m,n,\operatorname{deg}_{T}(F^{*})}b^{2}q^{n\Delta}+q^{(n+1)\Delta}+q^{(n+1)b-\Delta}. (8.7)

Note that, in order for this to be strictly better than the trivial bound q(n+1)b\ll q^{(n+1)b}, we must have that

Δ<b.\Delta<b. (8.8)

Combined with (6.8), this motivates the hypothesis Δ<b<2Δ\Delta<b<2\Delta we currently assume.

8.3. Proof of Proposition 8.1: the good-good case (C1)

When x¯\underline{x} is in case (C1), we apply part (iii) of Proposition 7.1 to estimate each of the character sums SF(π¯2x¯,χπ1)S_{F}\left(\bar{\pi}_{2}\underline{x},\chi_{\pi_{1}}\right), SF(π¯1x¯,χπ2)S_{F}\left(\bar{\pi}_{1}\underline{x},\chi_{\pi_{2}}\right). We obtain

q(n+1)(2Δb)|𝒫|2π1,π2𝒫π1π2χπ1χ0χπ2χ0x¯𝒪Kn+1degT(x¯)<2Δbx¯in case (C1)|SF(π¯2x¯,χπ1)SF(π¯1x¯,χπ2)|,m,nq(n+1)(bΔ)N1,\frac{q^{-(n+1)(2\Delta-b)}}{|\mathcal{P}|^{2}}\sum_{\begin{subarray}{c}\pi_{1},\pi_{2}\in{\mathcal{P}}\\ \pi_{1}\neq\pi_{2}\end{subarray}}\displaystyle\sum_{\begin{subarray}{c}\chi_{\pi_{1}}\neq\chi_{0}\\ \chi_{\pi_{2}}\neq\chi_{0}\end{subarray}}\displaystyle\sum_{\begin{subarray}{c}\underline{x}\in\mathcal{O}_{K}^{n+1}\\ \operatorname{deg}_{T}(\underline{x})<2\Delta-b\\ \underline{x}\;\text{in case (C1)}\end{subarray}}\left|S_{F}\left(\bar{\pi}_{2}\underline{x},\chi_{\pi_{1}}\right)S_{F}\left(\bar{\pi}_{1}\underline{x},\chi_{\pi_{2}}\right)\right|\ll_{\ell,m,n}q^{(n+1)(b-\Delta)}N_{1},

where

N1\displaystyle N_{1} :=maxπ1π2𝒫#{x¯𝒪Kn+1:degT(x¯)<2Δb,π¯2x¯𝒲(F)π1,π¯1x¯𝒲(F)π2},\displaystyle:=\max_{\pi_{1}\neq\pi_{2}\in\mathcal{P}}\#\left\{\underline{x}\in\mathcal{O}_{K}^{n+1}:\operatorname{deg}_{T}(\underline{x})<2\Delta-b,\bar{\pi}_{2}\underline{x}\not\in{\mathcal{W}}(F)^{\ast}_{\pi_{1}},\bar{\pi}_{1}\underline{x}\not\in{\mathcal{W}}(F)^{\ast}_{\pi_{2}}\right\},

and we again used that for each πi\pi_{i}, the number of characters of order \ell modulo πi\pi_{i} is \ell. Note that

N1#{x¯𝒪Kn+1:degT(x¯)<2Δb}q(2Δb)(n+1).\displaystyle N_{1}\leq\#\left\{\underline{x}\in\mathcal{O}_{K}^{n+1}:\operatorname{deg}_{T}(\underline{x})<2\Delta-b\right\}\leq q^{\left(2\Delta-b\right)(n+1)}.

Thus the total contribution of x¯\underline{x} in case (C1) into the unramified sieve term is

,m,nq(n+1)Δ,\ll_{\ell,m,n}q^{(n+1)\Delta},

which we note is comparable to a term in the bad-bad contribution (8.7).

8.4. Proof of Proposition 8.1: cases (C2) and (C2’)

As in the case (C3), we break the sum in (8.2) into two sums Σ1+Σ2\Sigma_{1}+\Sigma_{2}, according to whether the sum of x¯𝒪Kn+1\underline{x}\in\mathcal{O}_{K}^{n+1} takes place over the x¯\underline{x} satisfying F(x¯)0F^{*}(\underline{x})\not=0 or F(x¯)=0F^{*}(\underline{x})=0. We define

Σ1=q(n+1)(2Δb)|𝒫|2π1,π2𝒫π1π2χπ1χ0χπ2χ0x¯𝒪Kn+1degT(x¯)<2ΔbF(x¯)0, case (C2)|SF(π¯2x¯,χπ1)SF(π¯1x¯,χπ2)|.\Sigma_{1}=\frac{q^{-(n+1)(2\Delta-b)}}{|\mathcal{P}|^{2}}\sum_{\begin{subarray}{c}\pi_{1},\pi_{2}\in{\mathcal{P}}\\ \pi_{1}\neq\pi_{2}\end{subarray}}\displaystyle\sum_{\begin{subarray}{c}\chi_{\pi_{1}}\neq\chi_{0}\\ \chi_{\pi_{2}}\neq\chi_{0}\end{subarray}}\displaystyle\sum_{\begin{subarray}{c}\underline{x}\in\mathcal{O}_{K}^{n+1}\\ \operatorname{deg}_{T}(\underline{x})<2\Delta-b\\ \text{$F^{*}(\underline{x})\neq 0$, case (C2)}\end{subarray}}|S_{F}\left(\bar{\pi}_{2}\underline{x},\chi_{\pi_{1}}\right)S_{F}\left(\bar{\pi}_{1}\underline{x},\chi_{\pi_{2}}\right)|.

The sum Σ2\Sigma_{2} is formally defined analogously, with the condition F(x¯)0F^{*}(\underline{x})\neq 0 replaced by the condition F(x¯)=0.F^{*}(\underline{x})=0. However, note that under the conditions of case (C2) the sum in Σ2\Sigma_{2} is empty. Indeed, if F(x¯)=0F^{*}(\underline{x})=0 then all primes are bad (that is to say, x¯𝒲(F)π\underline{x}\in\mathcal{W}(F)_{\pi}^{*} for all π\pi in the sieving set), whereas in case (C2) at least one prime is good (since by assumption π¯1x¯𝒲(F)π2\overline{\pi}_{1}\underline{x}\not\in\mathcal{W}(F)^{*}_{\pi_{2}}).

It only remains to bound Σ1\Sigma_{1}, which we accomplish by using cases (i) and (ii) of Propositions 7.1 and (8.5) as before, to conclude that

|Σ1|,m,n,degT(F)bq(2n1)Δ2.|\Sigma_{1}|\ll_{\ell,m,n,\operatorname{deg}_{T}(F^{*})}bq^{\frac{(2n-1)\Delta}{2}}.

This is clearly dominated by the bad-bad contribution (8.7). The same arguments for Σ1,Σ2\Sigma_{1},\Sigma_{2} apply for the (C2’) contribution.

To recap, the work above for cases (C1), (C2), (C2’), (C3), has shown that the unramified sieve term satisfies the bound claimed in Proposition 8.1. All that remains is to prove the counting result in Lemma 8.2, to which we now turn.

8.5. Proof of Lemma 8.2

Let π\pi be as in the statement of the proposition. We have that

x¯𝒪Kn+1degT(x¯)<NG(x¯)=01x¯𝒪Kn+1degT(x¯)<NG(x¯)=0(modπ)1.\sum_{\begin{subarray}{c}\underline{x}\in\mathcal{O}_{K}^{n+1}\\ \operatorname{deg}_{T}(\underline{x})<N\\ G(\underline{x})=0\end{subarray}}1\leq\sum_{\begin{subarray}{c}\underline{x}\in\mathcal{O}_{K}^{n+1}\\ \operatorname{deg}_{T}(\underline{x})<N\\ G(\underline{x})=0\,(\operatorname{mod}\pi)\end{subarray}}1.

We complete the sum, by counting for each αkπn+1\alpha\in k_{\pi}^{n+1} such that G(α)=0(modπ)G(\alpha)=0\,(\operatorname{mod}\pi), those x¯\underline{x} with degT(x¯)<N\operatorname{deg}_{T}(\underline{x})<N such that x¯=α¯(modπ)\underline{x}=\underline{\alpha}\,(\operatorname{mod}\pi):

x¯𝒪Kn+1degT(x¯)<NG(x¯)=0(modπ)1=\displaystyle\sum_{\begin{subarray}{c}\underline{x}\in\mathcal{O}_{K}^{n+1}\\ \operatorname{deg}_{T}(\underline{x})<N\\ G(\underline{x})=0\,(\operatorname{mod}\pi)\end{subarray}}1= α¯kπn+1G(α¯)=0(modπ)x¯𝒪Kn+1degT(x¯)<N1q(n+1)Lβ¯kπn+1ψ(β¯(α¯x¯)π)\displaystyle\sum_{\begin{subarray}{c}\underline{\alpha}\in k_{\pi}^{n+1}\\ G(\underline{\alpha})=0\,(\operatorname{mod}\pi)\end{subarray}}\sum_{\begin{subarray}{c}\underline{x}\in\mathcal{O}_{K}^{n+1}\\ \operatorname{deg}_{T}(\underline{x})<N\end{subarray}}\frac{1}{q^{(n+1)L}}\sum_{\underline{\beta}\in k_{\pi}^{n+1}}\psi_{\infty}\left(\frac{\underline{\beta}\cdot(\underline{\alpha}-\underline{x})}{\pi}\right)
=\displaystyle= 1q(n+1)Lβ¯kπn+1x¯𝒪Kn+1degT(x¯)<Nψ(β¯x¯π)α¯kπn+1G(α¯)=0(modπ)ψ(β¯α¯π),\displaystyle\frac{1}{q^{(n+1)L}}\sum_{\underline{\beta}\in k_{\pi}^{n+1}}\sum_{\begin{subarray}{c}\underline{x}\in\mathcal{O}_{K}^{n+1}\\ \operatorname{deg}_{T}(\underline{x})<N\end{subarray}}\psi_{\infty}\left(\frac{-\underline{\beta}\cdot\underline{x}}{\pi}\right)\sum_{\begin{subarray}{c}\underline{\alpha}\in k_{\pi}^{n+1}\\ G(\underline{\alpha})=0\,(\operatorname{mod}\pi)\end{subarray}}\psi_{\infty}\left(\frac{\underline{\beta}\cdot\underline{\alpha}}{\pi}\right), (8.9)

where the additive character ψ(/π)\psi_{\infty}(\cdot/\pi) is defined in (3.2) and we have applied Lemma 6.4. We remark that

x¯𝒪Kn+1degT(x¯)<Nψ(β¯x¯π)=j=0n(xj𝒪KdegT(xj)<Nψ(βjxjπ))=j=0n(xj𝒪KdegT(xj)<Nψ(Trkπ/𝔽q(βjxj))).\sum_{\begin{subarray}{c}\underline{x}\in\mathcal{O}_{K}^{n+1}\\ \operatorname{deg}_{T}(\underline{x})<N\end{subarray}}\psi_{\infty}\left(\frac{-\underline{\beta}\cdot\underline{x}}{\pi}\right)=\prod_{j=0}^{n}\left(\sum_{\begin{subarray}{c}x_{j}\in\mathcal{O}_{K}\\ \operatorname{deg}_{T}(x_{j})<N\end{subarray}}\psi_{\infty}\left(\frac{-\beta_{j}x_{j}}{\pi}\right)\right)=\prod_{j=0}^{n}\left(\sum_{\begin{subarray}{c}x_{j}\in\mathcal{O}_{K}\\ \operatorname{deg}_{T}(x_{j})<N\end{subarray}}\psi_{\infty}\left(-\operatorname{Tr}_{k_{\pi}/\mathbb{F}_{q}}(\beta_{j}x_{j})\right)\right).

Thus we can work one coordinate at a time. Fix any βjkπ.\beta_{j}\in k_{\pi}. Write kπ=𝔽q[ρ]k_{\pi}=\mathbb{F}_{q}[\rho], where ρ\rho is a root of π\pi. Then {1,ρ,ρL1}\{1,\rho\dots,\rho^{L-1}\} is a basis of kπk_{\pi} as an 𝔽q\mathbb{F}_{q}-vector space. Thus, x𝒪Kx\in\mathcal{O}_{K} with degT(x)<N\operatorname{deg}_{T}(x)<N can be expressed as x=a0+a1ρ++aN1ρN1x=a_{0}+a_{1}\rho+\cdots+a_{N-1}\rho^{N-1} with the coefficients a𝔽qa_{\ell}\in\mathbb{F}_{q} uniquely determined; this leads to

xj𝒪KdegT(xj)<Nψ(Trkπ/𝔽q(βjxj))==0N1(a𝔽qψ(Trkπ/𝔽qa(ρβj))).\sum_{\begin{subarray}{c}x_{j}\in\mathcal{O}_{K}\\ \operatorname{deg}_{T}(x_{j})<N\end{subarray}}\psi_{\infty}\left(-\operatorname{Tr}_{k_{\pi}/\mathbb{F}_{q}}(\beta_{j}x_{j})\right)=\prod_{\ell=0}^{N-1}\left(\sum_{a_{\ell}\in\mathbb{F}_{q}}\psi_{\infty}\left(-\operatorname{Tr}_{k_{\pi}/\mathbb{F}_{q}}a_{\ell}(\rho^{\ell}\beta_{j})\right)\right). (8.10)

Also notice that we have for each fixed \ell that

a𝔽qψ(Trkπ/𝔽qa(ρβj))={qifTrkπ/𝔽q(ρβj)=0,0otherwise.\sum_{a_{\ell}\in\mathbb{F}_{q}}\psi_{\infty}\left(-\operatorname{Tr}_{k_{\pi}/\mathbb{F}_{q}}a_{\ell}(\rho^{\ell}\beta_{j})\right)=\begin{cases}q&\text{if}\,\operatorname{Tr}_{k_{\pi}/\mathbb{F}_{q}}(\rho^{\ell}\beta_{j})=0,\\ 0&\text{otherwise}.\end{cases}

Combining the above with (8.10), we obtain

xj𝒪KdegT(xj)<Nψ(Trkπ/𝔽q(βjxj))={qNifTrkπ/𝔽q(ρβj)=0for=0,,N1,0otherwise.\sum_{\begin{subarray}{c}x_{j}\in\mathcal{O}_{K}\\ \operatorname{deg}_{T}(x_{j})<N\end{subarray}}\psi_{\infty}\left(-\operatorname{Tr}_{k_{\pi}/\mathbb{F}_{q}}(\beta_{j}x_{j})\right)=\begin{cases}q^{N}&\text{if}\,\operatorname{Tr}_{k_{\pi}/\mathbb{F}_{q}}(\rho^{\ell}\beta_{j})=0\,\text{for}\,\ell=0,\dots,N-1,\\ 0&\text{otherwise}.\end{cases}

Define for any 0NL0\leq N\leq L,

SN={βkπ:Trkπ/𝔽q(ρβ)=0,for=0,,N1}.S_{N}=\{\beta\in k_{\pi}\,:\,\operatorname{Tr}_{k_{\pi}/\mathbb{F}_{q}}(\rho^{\ell}\beta)=0,\,\text{for}\,\ell=0,\dots,N-1\}.

We claim that #SN=qLN\#S_{N}=q^{L-N}. To see this, consider for a given 0L10\leq\ell\leq L-1,

H:={βkπ:Trkπ/𝔽q(ρβ)=0}.H_{\ell}:=\{\beta\in k_{\pi}\,:\,\operatorname{Tr}_{k_{\pi}/\mathbb{F}_{q}}(\rho^{\ell}\beta)=0\}.

Note that HH_{\ell} is a hyperplane in the vector space kπk_{\pi} for any =0,,L1\ell=0,\dots,L-1. We will prove that SL==0L1H={0}S_{L}=\cap_{\ell=0}^{L-1}H_{\ell}=\{0\}. Indeed, if γSL\gamma\in S_{L}, then, since {1,ρ,,ρL1}\{1,\rho,\dots,\rho^{L-1}\} is a basis for kπk_{\pi}, linearity of trace implies that Trkπ/𝔽q(γy)=0\operatorname{Tr}_{k_{\pi}/\mathbb{F}_{q}}(\gamma y)=0 for all ykπy\in k_{\pi}. Now the trace pairing is non-degenerate if and only if the extension is separable (see, for example, [Jan96, Ch.1, Sec.5.2]), and since 𝔽q\mathbb{F}_{q} is perfect, we conclude that γ=0\gamma=0. Since #S0=qL\#S_{0}=q^{L} and #SL=1\#S_{L}=1, we conclude that each hyperplane H+1H_{\ell+1} lowers the dimension by exactly 11 when going from SS_{\ell} to S+1=SHS_{\ell+1}=S_{\ell}\cap H_{\ell} as long as +1L\ell+1\leq L. Once we reach L\ell\geq L, the dimension remains 0. Thus, in particular for a given NLN\leq L, we conclude that #SN=qLN\#S_{N}=q^{L-N}.

Back to the identity in (8.5), we have shown that

x¯𝒪Kn+1degT(x¯)<NG(x¯)=0(modπ)1=\displaystyle\sum_{\begin{subarray}{c}\underline{x}\in\mathcal{O}_{K}^{n+1}\\ \operatorname{deg}_{T}(\underline{x})<N\\ G(\underline{x})=0\,(\operatorname{mod}\pi)\end{subarray}}1= q(n+1)Nq(n+1)Lβ¯kπn+1𝟏β¯SNn+1α¯kπn+1G(α¯)=0(modπ)ψ(β¯α¯π)\displaystyle\frac{q^{(n+1)N}}{q^{(n+1)L}}\sum_{\underline{\beta}\in k_{\pi}^{n+1}}\mathbf{1}_{\underline{\beta}\in S_{N}^{n+1}}\sum_{\begin{subarray}{c}\underline{\alpha}\in k_{\pi}^{n+1}\\ G(\underline{\alpha})=0\,(\operatorname{mod}\pi)\end{subarray}}\psi_{\infty}\left(\frac{\underline{\beta}\cdot\underline{\alpha}}{\pi}\right)
n,degX¯(G)\displaystyle\ll_{n,\operatorname{deg}_{\underline{X}}(G)} q(n+1)Nq(n+1)L|SN|n+1maxβ¯kπn+1β¯0|α¯kπn+1G(α¯)=0(modπ)ψ(β¯α¯π)|+q(n+1)NL.\displaystyle\frac{q^{(n+1)N}}{q^{(n+1)L}}|S_{N}|^{n+1}\max_{\begin{subarray}{c}\underline{\beta}\in k_{\pi}^{n+1}\\ \underline{\beta}\not=0\end{subarray}}\left|\sum_{\begin{subarray}{c}\underline{\alpha}\in k_{\pi}^{n+1}\\ G(\underline{\alpha})=0\,(\operatorname{mod}\pi)\end{subarray}}\psi_{\infty}\left(\frac{\underline{\beta}\cdot\underline{\alpha}}{\pi}\right)\right|+q^{(n+1)N-L}.

In the last term, corresponding to β¯=0\underline{\beta}=0, we have applied the Lang-Weil bound [LW54, Thm. 1] to count

{α¯kπn+1:G(α¯)=0(modπ)}n,degX¯(G)qnL.\{\underline{\alpha}\in k_{\pi}^{n+1}:G(\underline{\alpha})=0(\operatorname{mod}\pi)\}\ll_{n,\operatorname{deg}_{\underline{X}}(G)}q^{nL}.

We now consider the additive character sum above. Given β¯0\underline{\beta}\neq 0, we start by writing

α¯kπn+1G(α¯)=0(modπ)ψ(β¯α¯π)=γkπψ(γπ)α¯kπn+1G(α¯)=0(modπ)α¯β¯=γ(modπ)1.\sum_{\begin{subarray}{c}\underline{\alpha}\in k_{\pi}^{n+1}\\ G(\underline{\alpha})=0\,(\operatorname{mod}\pi)\end{subarray}}\psi_{\infty}\left(\frac{\underline{\beta}\cdot\underline{\alpha}}{\pi}\right)=\sum_{\gamma\in k_{\pi}}\psi_{\infty}\left(\frac{\gamma}{\pi}\right)\sum_{\begin{subarray}{c}\underline{\alpha}\in k_{\pi}^{n+1}\\ G(\underline{\alpha})=0\,(\operatorname{mod}\pi)\\ \underline{\alpha}\cdot\underline{\beta}=\gamma\,(\operatorname{mod}\pi)\end{subarray}}1. (8.11)

Consider the sum over α¯\underline{\alpha}. This sum is independent of γ\gamma for γkπ\gamma\in k_{\pi}^{*}. Indeed, for γ,γ0kπ\gamma,\gamma_{0}\in k_{\pi}^{*}, write γ0=δγ\gamma_{0}=\delta\gamma. Setting d=degX¯Gd=\operatorname{deg}_{\underline{X}}G, and using the homogeneity of G(X¯)G(\underline{X}), this gives

α¯kπn+1G(α¯)=0(modπ)α¯β¯=γ(modπ)1=\displaystyle\sum_{\begin{subarray}{c}\underline{\alpha}\in k_{\pi}^{n+1}\\ G(\underline{\alpha})=0\,(\operatorname{mod}\pi)\\ \underline{\alpha}\cdot\underline{\beta}=\gamma\,(\operatorname{mod}\pi)\end{subarray}}1= α¯kπn+1δdG(α¯)=0(modπ)α¯β¯=γ(modπ)1=α¯kπn+1G(δα¯)=0(modπ)α¯β¯=γ(modπ)1\displaystyle\sum_{\begin{subarray}{c}\underline{\alpha}\in k_{\pi}^{n+1}\\ \delta^{d}G(\underline{\alpha})=0\,(\operatorname{mod}\pi)\\ \underline{\alpha}\cdot\underline{\beta}=\gamma\,(\operatorname{mod}\pi)\end{subarray}}1=\sum_{\begin{subarray}{c}\underline{\alpha}\in k_{\pi}^{n+1}\\ G(\delta\underline{\alpha})=0\,(\operatorname{mod}\pi)\\ \underline{\alpha}\cdot\underline{\beta}=\gamma\,(\operatorname{mod}\pi)\end{subarray}}1
=\displaystyle= α0¯kπn+1G(α0¯)=0(modπ)α0¯β¯=δγ(modπ)1=α0¯kπn+1G(α0¯)=0(modπ)α0¯β¯=γ0(modπ)1.\displaystyle\sum_{\begin{subarray}{c}\underline{\alpha_{0}}\in k_{\pi}^{n+1}\\ \ G(\underline{\alpha_{0}})=0\,(\operatorname{mod}\pi)\\ \underline{\alpha_{0}}\cdot\underline{\beta}=\delta\gamma\,(\operatorname{mod}\pi)\end{subarray}}1=\sum_{\begin{subarray}{c}\underline{\alpha_{0}}\in k_{\pi}^{n+1}\\ \ G(\underline{\alpha_{0}})=0\,(\operatorname{mod}\pi)\\ \underline{\alpha_{0}}\cdot\underline{\beta}=\gamma_{0}\,(\operatorname{mod}\pi)\end{subarray}}1.

Combining the above observation with (8.11), we obtain

α¯kπn+1G(α¯)=0(modπ)ψ(β¯α¯π)=\displaystyle\sum_{\begin{subarray}{c}\underline{\alpha}\in k_{\pi}^{n+1}\\ G(\underline{\alpha})=0\,(\operatorname{mod}\pi)\end{subarray}}\psi_{\infty}\left(\frac{\underline{\beta}\cdot\underline{\alpha}}{\pi}\right)= α¯kπn+1G(α¯)=0(modπ)α¯β¯=1(modπ)1+α¯kπn+1G(α¯)=0(modπ)α¯β¯=0(modπ)1.\displaystyle-\sum_{\begin{subarray}{c}\underline{\alpha}\in k_{\pi}^{n+1}\\ G(\underline{\alpha})=0\,(\operatorname{mod}\pi)\\ \underline{\alpha}\cdot\underline{\beta}=1\,(\operatorname{mod}\pi)\end{subarray}}1+\sum_{\begin{subarray}{c}\underline{\alpha}\in k_{\pi}^{n+1}\\ G(\underline{\alpha})=0\,(\operatorname{mod}\pi)\\ \underline{\alpha}\cdot\underline{\beta}=0\,(\operatorname{mod}\pi)\end{subarray}}1.

We conclude by applying the next lemma.

Lemma 8.3.

Let G𝒪K[X0,,Xn]G\in\mathcal{O}_{K}[X_{0},\dots,X_{n}] be an irreducible homogeneous polynomial of degree degX¯G2\operatorname{deg}_{\underline{X}}G\geq 2, and let π\pi be an irreducible polynomial of degree LL. Fix β¯kπn+1\underline{\beta}\in k_{\pi}^{n+1}, β¯(0,,0)\underline{\beta}\not=(0,\dots,0). Then

α¯kπn+1G(α¯)=0(modπ)α¯β¯=1(modπ)1n,degX¯(G)q(n1)L\sum_{\begin{subarray}{c}\underline{\alpha}\in k_{\pi}^{n+1}\\ G(\underline{\alpha})=0\,(\operatorname{mod}\pi)\\ \underline{\alpha}\cdot\underline{\beta}=1\,(\operatorname{mod}\pi)\end{subarray}}1\ll_{n,\operatorname{deg}_{\underline{X}}(G)}q^{(n-1)L}

and

α¯kπn+1G(α¯)=0(modπ)α¯β¯=0(modπ)1n,degX¯(G)q(n1)L.\sum_{\begin{subarray}{c}\underline{\alpha}\in k_{\pi}^{n+1}\\ G(\underline{\alpha})=0\,(\operatorname{mod}\pi)\\ \underline{\alpha}\cdot\underline{\beta}=0\,(\operatorname{mod}\pi)\end{subarray}}1\ll_{n,\operatorname{deg}_{\underline{X}}(G)}q^{(n-1)L}.
Proof of Lemma 8.3.

We consider an embedding of kπn+1k_{\pi}^{n+1} in kπn+1\mathbb{P}_{k_{\pi}}^{n+1} by adding an extra coordinate TT and interpreting kπn+1k_{\pi}^{n+1} as the subset where T=1T=1.

α¯kπn+1G(α¯)=0modπα¯β¯=1modπ1=(α¯,T)n+1G(α¯)=0modπα¯β¯=TmodπT01(α¯,T)n+1G(α¯)=0modπα¯β¯=Tmodπ1n,degX¯(G)q(n1)L,\displaystyle\sum_{\begin{subarray}{c}\underline{\alpha}\in k_{\pi}^{n+1}\\ G(\underline{\alpha})=0\operatorname{mod}\pi\\ \underline{\alpha}\cdot\underline{\beta}=1\operatorname{mod}\pi\end{subarray}}1=\sum_{\begin{subarray}{c}(\underline{\alpha},T)\in\mathbb{P}^{n+1}\\ G(\underline{\alpha})=0\operatorname{mod}\pi\\ \underline{\alpha}\cdot\underline{\beta}=T\operatorname{mod}\pi\\ T\not=0\end{subarray}}1\leq\sum_{\begin{subarray}{c}(\underline{\alpha},T)\in\mathbb{P}^{n+1}\\ G(\underline{\alpha})=0\operatorname{mod}\pi\\ \underline{\alpha}\cdot\underline{\beta}=T\operatorname{mod}\pi\end{subarray}}1\ll_{n,\operatorname{deg}_{\underline{X}}(G)}q^{(n-1)L},

where in the last identity we applied the fact that GG is irreducible and has degree at least 2, so that we can apply the result of Lang-Weil for a variety of codimension 2 [LW54, Thm 1].

For the second inequality, write for α¯kπn+1\underline{\alpha}\in k_{\pi}^{n+1}, α¯=aγ¯\underline{\alpha}=a\underline{\gamma} with akπa\in k_{\pi} and γ¯kπn\underline{\gamma}\in\mathbb{P}_{k_{\pi}}^{n}. This gives

α¯kπn+1G(α¯)=0modπα¯β¯=0modπ1=1+a0modkπγ¯nG(γ¯)=0modπγ¯β¯=0modπ1n,degX¯(G)1+(qL1)q(n2)L\displaystyle\sum_{\begin{subarray}{c}\underline{\alpha}\in k_{\pi}^{n+1}\\ G(\underline{\alpha})=0\operatorname{mod}\pi\\ \underline{\alpha}\cdot\underline{\beta}=0\operatorname{mod}\pi\end{subarray}}1=1+\sum_{\begin{subarray}{c}a\not=0\operatorname{mod}{k_{\pi}}\end{subarray}}\sum_{\begin{subarray}{c}\underline{\gamma}\in\mathbb{P}^{n}\\ G(\underline{\gamma})=0\operatorname{mod}\pi\\ \underline{\gamma}\cdot\underline{\beta}=0\operatorname{mod}\pi\end{subarray}}1\ll_{n,\operatorname{deg}_{\underline{X}}(G)}1+(q^{L}-1)q^{(n-2)L}

by a second application of Lang-Weil for a variety of codimension 2, and we conclude. ∎

This concludes the proof of Lemma 8.2, and the proof of Proposition 8.1 is complete.

9. Completing the proof of Theorem 5.1: choice of parameters

It is time to wrap up the proof of Theorem 5.1 (and Theorem 1.1). Putting together all our estimates for the main sieve term, the ramified sieve term, and the unramified sieve term, we obtain that, for fixed q,n,,Fq,n,\ell,F as in the statement of the aforementioned theorems, for any sufficiently large positive integer bb and for any positive integer Δ\Delta chosen such that (5.5) holds and

Δ<b<2Δ,\Delta<b<2\Delta,

we have

|𝒮F(𝒜)|,m,n,degT(F),degT(F)bq(n+1)bΔ+qnb+b2qnΔ+q(n+1)Δ+q(n+1)bΔbq(n+1)bΔ+q(n+1)Δ,\left|{\mathcal{S}}_{F}({\mathcal{A}})\right|\ll_{\ell,m,n,\operatorname{deg}_{T}(F),\operatorname{deg}_{T}(F^{*})}bq^{(n+1)b-\Delta}+q^{nb}+b^{2}q^{n\Delta}+q^{(n+1)\Delta}+q^{(n+1)b-\Delta}\ll bq^{(n+1)b-\Delta}+q^{(n+1)\Delta}, (9.1)

where in the last inequality we applied the fact that Δ<b\Delta<b so that nb<(n+1)bΔnb<(n+1)b-\Delta and we imposed the additional assumption

b2qΔ,b^{2}\leq q^{\Delta}, (9.2)

which we will verify momentarily. Our remaining goal is to choose Δ\Delta optimally such that the resulting upper bound for |𝒮F(𝒜)|\left|{\mathcal{S}}_{F}({\mathcal{A}})\right| improves upon the trivial bound q(n+1)bq^{(n+1)b}. In what follows, we address the choice of Δ\Delta, after which we address the existence of a minimal b(n,q,F)b(n,q,F) such that our previous working assumptions on Δ\Delta hold for all bb(n,q,F)b\geq b(n,q,F) (namely Δ<b<2Δ\Delta<b<2\Delta, and the assumptions (5.5), (8.6), (9.2)).

9.1. Choice of Δ=Δ(n,b)\Delta=\Delta(n,b)

Looking at the right-most side of (9.1), we see that an initial choice Δ0\Delta_{0} of Δ\Delta could be made such that the two terms are balanced, that is,

bq(n+1)bΔ0=q(n+1)Δ0,bq^{(n+1)b-\Delta_{0}}=q^{(n+1)\Delta_{0}},

which is equivalent to choosing

Δ0:=(n+1)b+logqbn+2.\Delta_{0}:=\frac{(n+1)b+\operatorname{log}_{q}b}{n+2}. (9.3)

Since the above Δ0\Delta_{0} is not necessarily an integer, while Δ\Delta, as the degree of a polynomial, must be an integer, we write Δ0\Delta_{0} in terms of its integral and fractional parts, and choose Δ\Delta to be the former:

Δ0=(n+1)b+logqbn+2+{(n+1)b+logqbn+2}=Δ+δ0,\Delta_{0}=\left\lfloor\frac{(n+1)b+\operatorname{log}_{q}b}{n+2}\right\rfloor+\left\{\frac{(n+1)b+\operatorname{log}_{q}b}{n+2}\right\}=\Delta+\delta_{0},

where, we highlight,

Δ=Δ(n,b):=(n+1)b+logqbn+2(including zero)\Delta=\Delta(n,b):=\left\lfloor\frac{(n+1)b+\operatorname{log}_{q}b}{n+2}\right\rfloor\in\mathbb{N}\quad\text{(including zero)} (9.4)

and

δ0:={(n+1)b+logqbn+2}[0,1).\delta_{0}:=\left\{\frac{(n+1)b+\operatorname{log}_{q}b}{n+2}\right\}\in[0,1).

With this choice of Δ\Delta, we obtain:

|𝒮F(𝒜)|\displaystyle\left|{\mathcal{S}}_{F}({\mathcal{A}})\right| ,m,n,degT(F),degT(F)\displaystyle\ll_{\ell,m,n,\operatorname{deg}_{T}(F),\operatorname{deg}_{T}(F^{*})} bq(n+1)bΔ+q(n+1)Δ\displaystyle bq^{(n+1)b-\Delta}+q^{(n+1)\Delta}
2bq(n+1)bΔ(since (n+1)Δ=(n+2)ΔΔ(n+2)Δ0Δ=(n+1)bΔ+logqb)\displaystyle\leq 2bq^{(n+1)b-\Delta}\quad\text{(since $(n+1)\Delta=(n+2)\Delta-\Delta\leq(n+2)\Delta_{0}-\Delta=(n+1)b-\Delta+\operatorname{log}_{q}b)$}
=2b(qb)(n+1)Δb\displaystyle=2b\left(q^{b}\right)^{(n+1)-\frac{\Delta}{b}}
<2b(qb)(n+1)n+1n+2logqbb(n+2)+1b,\displaystyle<2b\left(q^{b}\right)^{(n+1)-\frac{n+1}{n+2}-\frac{\operatorname{log}_{q}b}{b(n+2)}+\frac{1}{b}},

since

Δb=n+1n+2logqbb(n+2)+δ0b<n+1n+2logqbb(n+2)+1b.-\frac{\Delta}{b}=-\frac{n+1}{n+2}-\frac{\operatorname{log}_{q}b}{b(n+2)}+\frac{\delta_{0}}{b}<-\frac{n+1}{n+2}-\frac{\operatorname{log}_{q}b}{b(n+2)}+\frac{1}{b}.

We recognize the bound above as b11n+2(qb)(n+1)n+1n+2q\ll b^{1-\frac{1}{n+2}}\left(q^{b}\right)^{(n+1)-\frac{n+1}{n+2}}q. Recalling that qq is fixed, we conclude that

|𝒮F(𝒜)|,m,n,q,degT(F),degT(F)bn+1n+2(qb)(n+1)n+1n+2,\left|{\mathcal{S}}_{F}({\mathcal{A}})\right|\ll_{\ell,m,n,q,\operatorname{deg}_{T}(F),\operatorname{deg}_{T}(F^{*})}b^{\frac{n+1}{n+2}}(q^{b})^{(n+1)-\frac{n+1}{n+2}}, (9.5)

under the assumption that bb(n,q,F)b\geq b(n,q,F) is such that the inequalities Δ<b<2Δ\Delta<b<2\Delta and (5.5), (8.6), and (9.2) hold for Δ=Δ(n,b)\Delta=\Delta(n,b).

9.2. Choice of b(n,q,F)b(n,q,F)

First, let us note that, as a first constraint, bb must be chosen sufficiently large to ensure that Δ(n,b)0\Delta(n,b)\neq 0; it suffices to have bb1b\geq b_{1} for some b1=b1(n)b_{1}=b_{1}(n) chosen such that

(n+1)b1+logqb1n+2(n+1)b1n+20.\left\lfloor\frac{(n+1)b_{1}+\operatorname{log}_{q}b_{1}}{n+2}\right\rfloor\geq\left\lfloor\frac{(n+1)b_{1}}{n+2}\right\rfloor\neq 0. (9.6)

Next, recall that, in (5.5) of §5, we introduced the assumption

qΔΔ(qΔ2Δ+qΔ3)|𝒫exc(F)|qΔ2Δ,\frac{q^{\Delta}}{\Delta}-\left(\frac{q^{\frac{\Delta}{2}}}{\Delta}+q^{\frac{\Delta}{3}}\right)-|\mathcal{P}_{\text{exc}}(F)|\geq\frac{q^{\Delta}}{2\Delta}, (9.7)

which gives rise to a second set of constraints on b=b(n,q,F)b=b(n,q,F). Observe that, if we choose bb such that

|𝒫exc(F)|qΔ4Δ|\mathcal{P}_{\text{exc}}(F)|\leq\frac{q^{\Delta}}{4\Delta} (9.8)

and

qΔ2Δ+qΔ3qΔ4Δ,\frac{q^{\frac{\Delta}{2}}}{\Delta}+q^{\frac{\Delta}{3}}\leq\frac{q^{\Delta}}{4\Delta}, (9.9)

then

qΔΔ(qΔ2Δ+qΔ3)|𝒫exc(F)|qΔΔqΔ4ΔqΔ4Δ=qΔ2Δ,\frac{q^{\Delta}}{\Delta}-\left(\frac{q^{\frac{\Delta}{2}}}{\Delta}+q^{\frac{\Delta}{3}}\right)-|\mathcal{P}_{\text{exc}}(F)|\geq\frac{q^{\Delta}}{\Delta}-\frac{q^{\Delta}}{4\Delta}-\frac{q^{\Delta}}{4\Delta}=\frac{q^{\Delta}}{2\Delta},

which ensures (9.7). Since 𝒫exc(F)\mathcal{P}_{\text{exc}}(F) is a finite set and its cardinality depends on qq and FF, we can find b2,1=b2,1(q,F)b_{2,1}=b_{2,1}(q,F) such that, for any bb2,1b\geq b_{2,1}, (9.8) holds for Δ=Δ(n,b)\Delta=\Delta(n,b) as chosen in (9.4). Observe that

qΔ2+ΔqΔ3qΔ2+ΔqΔ2<qΔ2+bqΔ2=(b+1)qΔ2,q^{\frac{\Delta}{2}}+\Delta q^{\frac{\Delta}{3}}\leq q^{\frac{\Delta}{2}}+\Delta q^{\frac{\Delta}{2}}<q^{\frac{\Delta}{2}}+bq^{\frac{\Delta}{2}}=(b+1)q^{\frac{\Delta}{2}},

where we used that Δ<b\Delta<b. Thus, to ensure (9.9), it suffices to choose b2,2=b2,2(n,q)b_{2,2}=b_{2,2}(n,q) such that, for any bb2,2b\geq b_{2,2}, we have

4(b+1)q12(n+1)bn+2.4(b+1)\leq q^{\frac{1}{2}\left\lfloor\frac{(n+1)b}{n+2}\right\rfloor}.

In addition, remark that the above condition ensures (9.2).

Next we have to ensure (8.6) holds, namely ΔL0(F).\Delta\geq L_{0}(F^{*}). This similarly will hold as long as bb3b\geq b_{3} for some b3(F)b_{3}(F) sufficiently large.

Now note that the inequalities Δ<b<2Δ\Delta<b<2\Delta give rise to a final set of constraints for bb. With our final choice (9.4) of Δ\Delta, these inequalities become

(n+1)b+logqbn+2<b<2(n+1)b+logqbn+2.\left\lfloor\frac{(n+1)b+\operatorname{log}_{q}b}{n+2}\right\rfloor<b<2\left\lfloor\frac{(n+1)b+\operatorname{log}_{q}b}{n+2}\right\rfloor. (9.10)

We claim that, for any n1n\geq 1, the inequalities (9.10) hold for any bb4b\geq b_{4} for some b4=b4(n)b_{4}=b_{4}(n). In what follows, we verify this claim.

Let b3(n+2)b\geq 3(n+2). Dividing bb with quotient and remainder by n+2n+2, we find uniquely determined non-negative integers b0,r0b_{0},r_{0} such that

b=(n+2)b0+r0b=(n+2)b_{0}+r_{0} (9.11)

and

0r0n+1.0\leq r_{0}\leq n+1. (9.12)

Note that

b03.b_{0}\geq 3.

Rewriting bb using (9.11), we obtain that (9.10) is equivalent to

(n+1)b0+(n+1)r0+logqbn+2<(n+1)b0+b0+r0<2(n+1)b0+2(n+1)r0+logqbn+2.(n+1)b_{0}+\left\lfloor\frac{(n+1)r_{0}+\operatorname{log}_{q}b}{n+2}\right\rfloor<(n+1)b_{0}+b_{0}+r_{0}<2(n+1)b_{0}+2\left\lfloor\frac{(n+1)r_{0}+\operatorname{log}_{q}b}{n+2}\right\rfloor. (9.13)

The left-hand side inequality in (9.13) is equivalent to

(n+1)r0+logqbn+2<b0+r0;\left\lfloor\frac{(n+1)r_{0}+\operatorname{log}_{q}b}{n+2}\right\rfloor<b_{0}+r_{0};

for this it suffices that

(n+1)r0+logqbn+2<b0+r0=br0n+2+r0=b+(n+1)r0n+2,\frac{(n+1)r_{0}+\operatorname{log}_{q}b}{n+2}<b_{0}+r_{0}=\frac{b-r_{0}}{n+2}+r_{0}=\frac{b+(n+1)r_{0}}{n+2},

which certainly holds.

The right-hand side inequality in (9.13) is equivalent to

b0+r0<(n+1)b0+2(n+1)r0+logqbn+2;b_{0}+r_{0}<(n+1)b_{0}+2\left\lfloor\frac{(n+1)r_{0}+\operatorname{log}_{q}b}{n+2}\right\rfloor;

this will certainly hold if

r0<nb0+2(n+1)r0n+2.r_{0}<nb_{0}+2\left\lfloor\frac{(n+1)r_{0}}{n+2}\right\rfloor.

Since b03b_{0}\geq 3 and 0r0n+10\leq r_{0}\leq n+1, the above inequality will hold if

n+1<3n,n+1<3n,

which holds for any n1n\geq 1.

In conclusion, provided n1n\geq 1, there exists a positive integer b(n,q,F)b(n,q,F) such that, for any b>b(n,q,F)b>b(n,q,F), the inequalities Δ<b<2Δ\Delta<b<2\Delta, (5.5), (8.6), and (9.2) hold for Δ=Δ(n,b)\Delta=\Delta(n,b), so that the sieve process has proved (9.5). On the other hand, for each bb(n,q,F)b\leq b(n,q,F), we can apply the trivial bound

𝒮F(𝒜)(qb)n+1q(n+1)b(n,q,F)n,q,F1.\mathcal{S}_{F}(\mathcal{A})\leq(q^{b})^{n+1}\leq q^{(n+1)b(n,q,F)}\ll_{n,q,F}1.

Thus by enlarging the implicit constant in (9.5) if necessary, it holds for all bb. This completes the proof of Theorem 1.1 (and of Theorem 5.1).

10. Counting bound

For completeness, we record below a simple counting lemma, which can be considered a “trivial bound” (sometimes also called the Schwartz-Zippel bound); we applied this in §5.3.

Lemma 10.1.

Let AA be a domain, n1n\geq 1 an integer, and GA[X0,,Xn]G\in A[X_{0},\ldots,X_{n}] a homogeneous polynomial of degree e1e\geq 1 in X0,,XnX_{0},\ldots,X_{n}. Then, for any finite subset SAS\subseteq A, we have

#{(γ0,,γn)Sn+1:G(γ0,,γn)=0}e|S|n.\#\left\{(\gamma_{0},\ldots,\gamma_{n})\in S^{n+1}:G(\gamma_{0},\ldots,\gamma_{n})=0\right\}\leq e|S|^{n}. (10.1)

We recall the standard proof, which proceeds by induction on nn (e.g., see [HB02, Thm. 1] for a version of this result when A=A=\mathbb{Z}).

Proof.

In what follows, LL is the field of fractions of AA and L¯\overline{L} is a fixed algebraic closure of LL. In our argument below, it suffices that G(X0,,Xn)L¯[X0,,Xn]G(X_{0},\ldots,X_{n})\in\overline{L}[X_{0},\ldots,X_{n}], that is, we do not need to assume that the coefficients of GG are in AA.

Since GG is homogenous of degree e1e\geq 1, for each 0ie0\leq i\leq e there exists a homogenous polynomial GiL¯[X1,,Xn]G_{i}\in\overline{L}[X_{1},\ldots,X_{n}], of degree eie-i, such that

G(X0,,Xn)=0ieX0iGi(X1,,Xn).G(X_{0},\ldots,X_{n})=\displaystyle\sum_{0\leq i\leq e}X_{0}^{i}\ G_{i}(X_{1},\ldots,X_{n}).

Take i0i_{0} to be the maximal index 0ie0\leq i\leq e such that GiG_{i} is not identically zero. Thus,

G(X0,,Xn)=0ii0X0iGi(X1,,Xn).G(X_{0},\ldots,X_{n})=\displaystyle\sum_{0\leq i\leq i_{0}}X_{0}^{i}\ G_{i}(X_{1},\ldots,X_{n}).

Our goal is to show that the number of solutions in Sn+1S^{n+1} to the equation

0ii0x0iGi(x1,,xn)=0\displaystyle\sum_{0\leq i\leq i_{0}}x_{0}^{i}\ G_{i}(x_{1},\ldots,x_{n})=0 (10.2)

is at most e|S|ne|S|^{n}. We prove this statement by induction on nn.

When n=1n=1, (10.2) becomes the equation

0ii0x0iGi(x1)=0,\displaystyle\sum_{0\leq i\leq i_{0}}x_{0}^{i}\ G_{i}(x_{1})=0, (10.3)

whose degree is i0i_{0}. Choosing γ1S\gamma_{1}\in S to be a root of the polynomial Gi0(X1)G_{i_{0}}(X_{1}), provided such a root exists, we note that (10.3) may be satisfied by the pair (x0,x1)=(γ0,γ1)(x_{0},x_{1})=(\gamma_{0},\gamma_{1}) for any γ0S\gamma_{0}\in S. Since Gi0(X1)G_{i_{0}}(X_{1}) has at most degX1(G1)=ei0\operatorname{deg}_{X_{1}}(G_{1})=e-i_{0} roots in L¯\overline{L}, there are at most ei0e-i_{0} choices for γ1\gamma_{1}. There are at most |S||S| choices for γ0\gamma_{0}. As such, in this case, there are at most (ei0)|S|(e-i_{0})|S| possible solutions (γ0,γ1)S2(\gamma_{0},\gamma_{1})\in S^{2} of (10.3). Choosing γ1S\gamma_{1}\in S to not be a root of the polynomial Gi0(X1)G_{i_{0}}(X_{1}), provided such γ1\gamma_{1} exists, we see that (10.3) is a degree i0i_{0} equation with unknown x0x_{0}. Viewed over L¯\overline{L}, this equation has i0i_{0} solutions x0=γ0x_{0}=\gamma_{0}. In total, in this case, there are at most |S|i0|S|i_{0} solutions (γ0,γ1)S2(\gamma_{0},\gamma_{1})\in S^{2} of (10.3). Altogether, we obtain that (10.3) has at most (ei0)|S|+i0|S|=e|S|(e-i_{0})|S|+i_{0}|S|=e|S| solutions in S2S^{2}.

When n2n\geq 2, we make the inductive hypothesis that

#{(γ1,,γn)Sn:G(γ1,,γn)=0}degX¯(G)|S|n1\#\left\{(\gamma_{1},\ldots,\gamma_{n})\in S^{n}:G^{\prime}(\gamma_{1},\ldots,\gamma_{n})=0\right\}\leq\operatorname{deg}_{\underline{X}}(G^{\prime})|S|^{n-1} (10.4)

for any homogenous polynomial G(X1,,Xn)L¯[X1,,Xn]G^{\prime}(X_{1},\ldots,X_{n})\in\overline{L}[X_{1},\ldots,X_{n}]. In particular, we assume that (10.4) holds for Gi0(X1,,Xn)G_{i_{0}}(X_{1},\ldots,X_{n}). Choosing (γ1,,γn)Sn(\gamma_{1},\ldots,\gamma_{n})\in S^{n} to be a root of the polynomial Gi0(X1,,Xn)G_{i_{0}}(X_{1},\ldots,X_{n}), provided it exists, we note that (10.2) might be satisfied by (x0,x1,,xn)=(γ0,γ1,,γn)(x_{0},x_{1},\ldots,x_{n})=(\gamma_{0},\gamma_{1},\ldots,\gamma_{n}) for any γ0S\gamma_{0}\in S. Since degX¯(Gi0)=ei0\operatorname{deg}_{\underline{X}}(G_{i_{0}})=e-i_{0}, by the induction hypothesis we know that there are at most (ei0)|S|n1(e-i_{0})|S|^{n-1} roots (γ1,,γn)Sn(\gamma_{1},\ldots,\gamma_{n})\in S^{n}. As such, in this case, there are at most (ei0)|S|n(e-i_{0})|S|^{n} solutions (γ0,γ1,,γn)Sn+1(\gamma_{0},\gamma_{1},\ldots,\gamma_{n})\in S^{n+1} of (10.2). Choosing (γ1,,γn)Sn(\gamma_{1},\ldots,\gamma_{n})\in S^{n} to not be a root of the polynomial Gi0(X1,,Xn)G_{i_{0}}(X_{1},\ldots,X_{n}), provided it exists, we see that (10.2) gives rise to the degree i0i_{0} equation

0ii0x0iGi(γ1,,γn)=0,\displaystyle\sum_{0\leq i\leq i_{0}}x_{0}^{i}\ G_{i}(\gamma_{1},\ldots,\gamma_{n})=0,

with unknown x0x_{0}. Viewed over L¯\overline{L}, this equation has at most i0i_{0} solutions x0=γ0x_{0}=\gamma_{0}. In total, in this case, there are at most i0|S|ni_{0}|S|^{n} solutions (γ0,γ1,,γn)Sn+1(\gamma_{0},\gamma_{1},\ldots,\gamma_{n})\in S^{n+1} of (10.2). Altogether, we obtain that (10.2) has at most (ei0)|S|n+i0|S|n=e|S|n(e-i_{0})|S|^{n}+i_{0}|S|^{n}=e|S|^{n} solutions in Sn+1S^{n+1}. ∎

Funding

This work was initiated at the Women In Numbers 3rd workshop organized at the Banff International Research Station (Alberta, Canada) in Spring 2014. We thank the workshop organizers, Ling Long, Rachel Pries, and Katherine Stange, for providing us with a focused and stimulating research environment for starting this work. We thank the Department of Mathematics, Statistics, and Computer Science at the University of Illinois at Chicago (USA) for sponsoring and hosting a visit of the authors to continue this work. AB has been partially supported by NSF grant DMS-2002716, Simons Foundation collaboration grants No. 244988 and No. 524015, and by Institute for Advanced Study, which includes funding from NSF grant DMS-1638352. ACC has been partially supported by the Simons Collaboration Grants No. 318454 and No. 709008. ML has been partially supported by the Natural Sciences and Engineering Research Council of Canada under Discovery Grant 355412-2013 and the Fonds de recherche du Québec - Nature et technologies, Projet de recherche en équipe 256442 and 300951. LBP thanks the Hausdorff Center for Mathematics for a productive research environment, and has been partially supported by NSF CAREER grant DMS-1652173, a Sloan Research Fellowship, by the Charles Simonyi Endowment and NSF Grant No. 1128155 at the IAS, and by a Joan and Joseph Birman Fellowship.

11. Appendix: Projective Duality in Fibers for Smooth Hypersurfaces
by Joseph Rabinoff

In this appendix we gather some results from the literature on projective duality for smooth hypersurfaces. In the case of global fields, we can “spread out from the generic fiber” to conclude that the same results hold at all but finitely many places. We will use the language of algebraic varieties to the extent possible, and our proofs will use only elementary facts from commutative algebra, but as we will be working with non-algebraically closed fields and coefficient rings like \mathbb{Z}, some basic scheme theory will be required to make certain constructions precise.

Let kk be a field, let k¯\bar{k} be an algebraic closure, and let kn\mathbb{P}^{n}_{k} be the nn-dimensional projective space over kk, with homogeneous coordinate ring k[X0,,Xn]k[X_{0},\ldots,X_{n}]. A hypersurface in kn\mathbb{P}^{n}_{k} is the zero set of a nonzero homogeneous polynomial Hk[X0,,Xn]H\in k[X_{0},\ldots,X_{n}]. Equivalently, a hypersurface is a closed subvariety (or subscheme) of pure dimension n1n-1 (see the proof of [Har77, Proposition II.6.4]). A hypersurface WW defined by Hk[X0,,Xn]H\in k[X_{0},\ldots,X_{n}] is smooth if the homogeneous polynomials H,H/X0,,H/XnH,\partial H/\partial X_{0},\ldots,\partial H/\partial X_{n} have no common zeros in k¯n\mathbb{P}^{n}_{\bar{k}}. The hypersurface WW is integral if HH is an irreducible polynomial over kk, and it is geometrically integral if HH is irreducible over k¯\bar{k}.

Lemma 11.1.

Let n2n\geq 2 and let WknW\subset\mathbb{P}^{n}_{k} be a smooth hypersurface. Then WW is geometrically integral.

Proof.

Suppose that WW is defined by Hk[X0,,Xn]H\in k[X_{0},\ldots,X_{n}]. Then the hypersurface Wk¯k¯nW_{\bar{k}}\subset\mathbb{P}^{n}_{\bar{k}} defined by Hk¯[X0,,Xn]H\in\bar{k}[X_{0},\ldots,X_{n}] is again smooth, since H,H/X0,,H/XnH,\partial H/\partial X_{0},\ldots,\partial H/\partial X_{n} have no common zeros in k¯\bar{k}. This implies that Wk¯W_{\bar{k}} is irreducible, since any two irreducible components would intersect in a singular point [Har77, Theorem I.7.2]. This means that Wk¯W_{\bar{k}} is a nonsingular variety [Har77, Example 10.0.3], which is necessarily defined by an irreducible polynomial. ∎

In the following we assume n2n\geq 2. Let WknW\subset\mathbb{P}^{n}_{k} be the smooth hypersurface defined by Hk[X0,,Xn]H\in k[X_{0},\ldots,X_{n}]. The tangent space to WW at a rational point PW(k)P\in W(k) is the hyperplane TP(W)T_{P}(W) defined by

HX0(P)X0++HXn(P)Xn=0.\frac{\partial H}{\partial X_{0}}(P)\,X_{0}+\cdots+\frac{\partial H}{\partial X_{n}}(P)\,X_{n}=0.

This is in fact a hyperplane, as H(P)=0H(P)=0 implies H/Xi(P)0\partial H/\partial X_{i}(P)\neq 0 for some ii. This construction can be improved in the following way. Let kn\mathbb{P}^{n*}_{k} be the dual projective space parameterizing hyperplanes in affine (n+1)(n+1)-space. Concretely, we have knkn\mathbb{P}^{n*}_{k}\cong\mathbb{P}^{n}_{k}, with a point [a0::an]kn[a_{0}:\cdots:a_{n}]\in\mathbb{P}^{n}_{k} corresponding to the hyperplane a0X0++anXn=0a_{0}X_{0}+\cdots+a_{n}X_{n}=0. The map W(k)knW(k)\to\mathbb{P}^{n*}_{k} defined by PTP(W)P\mapsto T_{P}(W) can be promoted to a regular map 𝒢W:Wkn\mathcal{G}_{W}\colon W\to\mathbb{P}^{n*}_{k}; using the identification knkn\mathbb{P}^{n*}_{k}\cong\mathbb{P}^{n}_{k}, it is given by the homogeneous polynomials [H/X0::H/Xn][\partial H/\partial X_{0}:\ldots:\partial H/\partial X_{n}]. We call 𝒢W\mathcal{G}_{W} the Gauss map; its image WW^{*} is the dual variety of WW. Algebraically, the Gauss map is defined by the kk-algebra homomorphism gW:k[Y0,,Yn]k[X0,,Xn]/(H)g_{W}\colon k[Y_{0},\ldots,Y_{n}]\to k[X_{0},\ldots,X_{n}]/(H) sending YiY_{i} to H/Xi\partial H/\partial X_{i}, and WW^{*} is defined by ker(gW)\ker(g_{W}).

More generally, if WW is singular then WW^{*} is defined to be the closure of the image of the nonsingular locus under the Gauss map.

The following proposition summarizes the main facts about projective duality for smooth hypersurfaces in arbitrary characteristic. All results are extracted from [Kle86], which is an excellent reference.

Proposition 11.2.

Let n2n\geq 2, let WknW\subset\mathbb{P}^{n}_{k} be a smooth hypersurface defined by Hk[X0,,Xn]H\in k[X_{0},\ldots,X_{n}], and let WknW^{*}\subset\mathbb{P}^{n*}_{k} be the dual variety. Suppose that WW is not a hyperplane, i.e. that deg(H)>1\operatorname{deg}(H)>1.

  1. (1)

    The dual WW^{*} is a geometrically integral hypersurface.

  2. (2)

    The Gauss map 𝒢W:WW\mathcal{G}_{W}\colon W\to W^{*} is generically finite.

  3. (3)

    If WW^{*} is defined by a homogeneous polynomial Hk[X0,,Xn]H^{*}\in k[X_{0},\ldots,X_{n}], then

    [k(W):k(W)]deg(H)=deg(H)(deg(H)1)n1,[k(W):k(W^{*})]\operatorname{deg}(H^{*})=\operatorname{deg}(H)(\operatorname{deg}(H)-1)^{n-1},

    where k()k(\phantom{n}) denotes the field of rational functions.

  4. (4)

    If the field extension k(W)/k(W)k(W)/k(W^{*}) is separable (e.g. if char(k)=0\operatorname{char}(k)=0), then k(W)=k(W)k(W)=k(W^{*}).

  5. (5)

    (Reciprocity) If k(W)=k(W)k(W)=k(W^{*}) then (W)=W(W^{*})^{*}=W.

Proof.

The image of a (geometrically) integral variety under a regular map is again (geometrically) integral. This is the geometric version of the following algebraic fact: if f:ABf\colon A\to B is a homomorphism from a non-zero unitary commutative ring AA to an integral domain BB, then ker(f)\ker(f) is prime. If dim(W)=n1\operatorname{dim}(W^{*})=n-1 then 𝒢W\mathcal{G}_{W} is generically finite, as it is then a dominant morphism of varieties of the same dimension. The fact that dim(W)=n1\operatorname{dim}(W^{*})=n-1, along with the degree formula in (3), follow from [Kle86, Proposition II.2(iv) and Proposition II.9]. Assertions (4) and (5) follow from [Kle86, Proposition II.15a]. ∎

Remark 11.3.

Proposition 11.2(4) does not assert that k(W)k(W) is a purely inseparable extension of k(W)k(W^{*}). Indeed, there is no restriction on the separable degree of k(W)/k(W)k(W)/k(W^{*}) (provided that k(W)/k(W)k(W)/k(W^{*}) is not separable): see [Kle86, §II.3].

Remark 11.4.

With the notation in Proposition 11.2, suppose that (W)=W(W^{*})^{*}=W. Let d=deg(H)d=\operatorname{deg}(H) and d=deg(H)d^{*}=\operatorname{deg}(H^{*}), so that d=d(d1)n1d^{*}=d(d-1)^{n-1}. If WW^{*} is also smooth then d=d(d1)n1d=d^{*}(d^{*}-1)^{n-1}, which is true only if d=2d=2. Hence WW^{*} is not smooth if (W)=W(W^{*})^{*}=W and WW is not a quadric. See [Kle86, Corollary II.10] for more details.

Now we apply the above considerations when kk is a global field (of any characteristic). If SS is a finite set of finite places of kk then we let 𝒪S\mathscr{O}_{S} denote the ring of SS-integers. A finite place 𝔭\mathfrak{p} not contained in SS corresponds to a maximal ideal of 𝒪S\mathscr{O}_{S}; we denote the residue field by κ(𝔭)=𝒪S/𝔭\kappa(\mathfrak{p})=\mathscr{O}_{S}/\mathfrak{p}. We wish to prove a version of Proposition 11.2 that holds for all finite places outside of some SS depending only on HH. We will do so by “spreading out from the generic fiber”: we will consider varieties (or schemes) defined over Spec(𝒪S)\operatorname{Spec}(\mathscr{O}_{S}), and take the closure of WW in 𝒪Sn\mathbb{P}^{n}_{\mathscr{O}_{S}}. If char(k)=p>0\operatorname{char}(k)=p>0 then kk is the function field of a smooth, projective, geometrically integral curve CC defined over a finite field 𝔽q\mathbb{F}_{q}, and SS may be identified with a finite set of (closed) points of CC; in this case, Spec(𝒪S)\operatorname{Spec}(\mathscr{O}_{S}) is simply the variety CSC\setminus S. In characteristic zero, we are forced to use some scheme theory, as Spec(𝒪S)\operatorname{Spec}(\mathscr{O}_{S}) is not a variety over a field; in both cases, our proofs are written in the language of elementary commutative algebra.

Proposition 11.5.

Let kk be a global field, let n2n\geq 2, and let WknW\subset\mathbb{P}^{n}_{k} be a smooth hypersurface defined by Hk[X0,,Xn]H\in k[X_{0},\ldots,X_{n}] of degree at least 22. Let WknW^{*}\subset\mathbb{P}^{n*}_{k} be the dual variety, and let Hk[Y0,,Yn]H^{*}\in k[Y_{0},\ldots,Y_{n}] be a homogeneous polynomial defining WW^{*}. There exists a finite set SS of finite places of kk, depending only on HH and HH^{*}, such that the following hold.

  1. (1)

    The polynomials HH and HH^{*} have coefficients in 𝒪S\mathscr{O}_{S}.

  2. (2)

    For all finite places 𝔭\mathfrak{p} not in SS, the polynomial H(mod𝔭)κ(𝔭)[X0,,Xn]H\pmod{\mathfrak{p}}\in\kappa(\mathfrak{p})[X_{0},\ldots,X_{n}] is nonzero (thus deg(Hmod𝔭)=deg(H)\operatorname{deg}(H\operatorname{mod}\mathfrak{p})=\operatorname{deg}(H) since HH is homogeneous), and the hypersurface W𝔭κ(𝔭)nW_{\mathfrak{p}}\subset\mathbb{P}^{n}_{\kappa(\mathfrak{p})} defined by H(mod𝔭)H\pmod{\mathfrak{p}} is smooth.

  3. (3)

    For all finite places 𝔭\mathfrak{p} not in SS, the dual variety (W𝔭)κ(𝔭)n(W_{\mathfrak{p}})^{*}\subset\mathbb{P}^{n*}_{\kappa(\mathfrak{p})} is defined by H(mod𝔭)H^{*}\pmod{\mathfrak{p}}.

Proof.

The first assertion is true once SS contains all places with respect to which some coefficient of HH or HH^{*} has negative valuation. The reduction H(mod𝔭)H\pmod{\mathfrak{p}} is nonzero as long as 𝔭\mathfrak{p} is not one of the finite set of places whose valuation is strictly positive on all coefficients of HH; we include such places in SS as well. Finally, we enlarge SS to assume that 𝒪S\mathscr{O}_{S} is a unique factorization domain. Note that HH is a primitive polynomial over 𝒪S\mathscr{O}_{S} by construction: its coefficients have no common factors because we included those in SS. Similarly, by enlarging SS if necessary, we may assume that H(mod𝔭)H^{*}\pmod{\mathfrak{p}} is nonzero for all 𝔭S\mathfrak{p}\notin S, so that HH^{*} is primitive over 𝒪S\mathscr{O}_{S}.

Consider the closed subscheme W¯𝒪Sn\overline{W}\subset\mathbb{P}^{n}_{\mathscr{O}_{S}} defined by HH. (If char(k)=p>0\operatorname{char}(k)=p>0 then this is a subvariety of 𝒪Sn=𝔽qn×(CS)\mathbb{P}^{n}_{\mathscr{O}_{S}}=\mathbb{P}^{n}_{\mathbb{F}_{q}}\times(C\setminus S).) Let II be the (homogeneous) ideal of 𝒪S[X0,,Xn]\mathscr{O}_{S}[X_{0},\ldots,X_{n}] defined by HH and H/X0,,H/Xn\partial H/\partial X_{0},\ldots,\partial H/\partial X_{n}. Since WW is smooth, the extended ideal Ik[X0,,Xn]Ik[X_{0},\ldots,X_{n}] (the ideal of k[X0,,Xn]k[X_{0},\ldots,X_{n}] generated by the image of II) contains (X0m,,Xnm)(X_{0}^{m},\ldots,X_{n}^{m}) for some m>0m>0. This means that each XimX_{i}^{m} is a linear combination of HH and the H/Xj\partial H/\partial X_{j} with coefficients in kk. Enlarging SS to contain all places with negative valuation on some coefficient of one of these linear combinations, we may assume (X0m,,Xnm)I(X_{0}^{m},\ldots,X_{n}^{m})\subset I. Then for 𝔭S\mathfrak{p}\notin S we have (X0m,,Xnm)(H,H/X0,,H/Xn)(mod𝔭)(X_{0}^{m},\ldots,X_{n}^{m})\subset(H,\partial H/\partial X_{0},\ldots,\partial H/\partial X_{n})\pmod{\mathfrak{p}}, so W𝔭W_{\mathfrak{p}} is smooth.

(Geometrically, the generic fiber of W¯Spec(𝒪S)\overline{W}\to\operatorname{Spec}(\mathscr{O}_{S}) is the hypersurface WW, and the fiber over a place 𝔭Spec(𝒪S)\mathfrak{p}\in\operatorname{Spec}(\mathscr{O}_{S}) is W𝔭W_{\mathfrak{p}}. Lemma 11.6 below shows that W¯\overline{W} is the closure of WW in 𝒪Sn\mathbb{P}^{n}_{\mathscr{O}_{S}}. The singular locus of W¯\overline{W} is a closed subscheme not intersecting the generic fiber of W¯Spec(𝒪S)\overline{W}\to\operatorname{Spec}(\mathscr{O}_{S}), so its image in Spec(𝒪S)\operatorname{Spec}(\mathscr{O}_{S}) is a finite set of closed points. Deleting these points allows us to assume W¯Spec(𝒪S)\overline{W}\to\operatorname{Spec}(\mathscr{O}_{S}) is smooth.)

Now consider the morphism (regular map) 𝒢:W¯𝒪Sn\mathcal{G}\colon\overline{W}\to\mathbb{P}^{n}_{\mathscr{O}_{S}} defined by the homogeneous polynomials [H/X0::H/Xn][\partial H/\partial X_{0}:\ldots:\partial H/\partial X_{n}]. As before, this is well-defined because HH and its partial derivatives have no common zeros. Let W¯\overline{W}{}^{*} denote the image of 𝒢\mathcal{G}. Algebraically, the morphism 𝒢\mathcal{G} corresponds to the 𝒪S\mathscr{O}_{S}-algebra map g:𝒪S[Y0,,Yn]𝒪S[X0,,Xn]/(H)g\colon\mathscr{O}_{S}[Y_{0},\ldots,Y_{n}]\to\mathscr{O}_{S}[X_{0},\ldots,X_{n}]/(H) sending YiY_{i} to H/Xi\partial H/\partial X_{i}, and W¯\overline{W}{}^{*} is defined by J=ker(g)J=\ker(g). The Gauss map 𝒢W\mathcal{G}_{W} corresponds to gW=g𝒪Sk:k[Y0,,Yn]k[X0,,Xn]/(H)g_{W}=g\otimes_{\mathscr{O}_{S}}k\colon k[Y_{0},\ldots,Y_{n}]\to k[X_{0},\ldots,X_{n}]/(H), and 𝒢W𝔭\mathcal{G}_{W_{\mathfrak{p}}} is defined by g𝔭=g(mod𝔭):k(𝔭)[Y0,,Yn]k(𝔭)[X0,,Xn]/(H(mod𝔭))g_{\mathfrak{p}}=g\pmod{\mathfrak{p}}\colon k(\mathfrak{p})[Y_{0},\ldots,Y_{n}]\to k(\mathfrak{p})[X_{0},\ldots,X_{n}]/(H\pmod{\mathfrak{p}}) for 𝔭S\mathfrak{p}\notin S. Hence the dual WW^{*} is defined by the ideal ker(gW)=Jk[Y0,,Yn]\ker(g_{W})=Jk[Y_{0},\ldots,Y_{n}], and the dual (W𝔭)(W_{\mathfrak{p}})^{*} is defined by ker(g𝔭)=J(mod𝔭)\ker(g_{\mathfrak{p}})=J\pmod{\mathfrak{p}}. But ker(gW)\ker(g_{W}) is generated by HH^{*}, and we are assuming HH^{*} to be primitive, so ker(gW)𝒪S[Y0,,Yn]=(H)\ker(g_{W})\cap\mathscr{O}_{S}[Y_{0},\ldots,Y_{n}]=(H^{*}) by Lemma 11.6. On the other hand, we have ker(gW)𝒪S[Y0,,Yn]=J\ker(g_{W})\cap\mathscr{O}_{S}[Y_{0},\ldots,Y_{n}]=J by [AM16, Proposition 3.11(iv)], since JJ is prime and ker(gW)=Jk[Y0,,Yn]\ker(g_{W})=Jk[Y_{0},\ldots,Y_{n}] is not the unit ideal. Hence (W𝔭)(W_{\mathfrak{p}})^{*} is defined by J(mod𝔭)=(H(mod𝔭))J\pmod{\mathfrak{p}}=(H^{*}\pmod{\mathfrak{p}}), as desired.

(Geometrically, the restriction of 𝒢\mathcal{G} to the generic fiber of W¯Spec(𝒪S)\overline{W}\to\operatorname{Spec}(\mathscr{O}_{S}) is the Gauss map 𝒢W\mathcal{G}_{W}, and the restriction to the fiber over 𝔭\mathfrak{p} is 𝒢W𝔭\mathcal{G}_{W_{\mathfrak{p}}}. Since HH^{*} is irreducible, it defines an integral hypersurface XX in 𝒪Sn\mathbb{P}^{n}_{\mathscr{O}_{S}}, which is thus the closure of its generic fiber. But W¯\overline{W}{}^{*} is also irreducible, and W¯\overline{W}{}^{*} and XX both have generic fiber WW.) ∎

We used the following lemma in the above proof.

Lemma 11.6.

Let RR be a unique factorization domain with fraction field kk, let HR[X1,,Xn]H\in R[X_{1},\ldots,X_{n}] be a primitive polynomial of positive degree, let II be the ideal of R[X1,,Xn]R[X_{1},\ldots,X_{n}] generated by HH, and let JJ be the ideal of k[X0,,Xn]k[X_{0},\ldots,X_{n}] generated by HH. Then JR[X1,,Xn]=IJ\cap R[X_{1},\ldots,X_{n}]=I.

Proof.

Since HH is primitive, it is a prime element of R[X1,,Xn]R[X_{1},\ldots,X_{n}], so II is prime. Since HH has positive degree, the ideal JJ is not the unit ideal. Now use [AM16, Proposition 3.11(iv)]. ∎

Joseph Rabinoff: Department of Mathematics, Duke University, 120 Science Drive, Durham, North Carolina 27708, USA. Email: jdr@math.duke.edu



References

  • [AM16] M. F. Atiyah and I. G. Macdonald, Introduction to commutative algebra, economy ed., Addison-Wesley Series in Mathematics, Westview Press, Boulder, CO, 2016, For the 1969 original see [ MR0242802]. MR 3525784
  • [Bon21] D. Bonolis, A polynomial sieve and sums of Deligne type, Int. Math. Res. Not. IMRN (2021), no. 2, 1096–1137. MR 4201962
  • [Bra15] J. Brandes, Sums and differences of power-free numbers, Acta Arith. 169 (2015), no. 2, 169–180. MR 3359952
  • [Bro09] N. Broberg, Rational points on finite covers of 1{\mathbb{P}}^{1} and 2{\mathbb{P}}^{2}, J. Number Theory 101 (2009), 195–207. MR 1979659
  • [Bro15] T. D. Browning, The polynomial sieve and equal sums of like polynomials, Int. Math. Res. Not. IMRN (2015), no. 7, 1987–2019. MR 3335239
  • [BV15] T. D. Browning and P. Vishe, Rational points on cubic hypersurfaces over 𝔽q(t)\mathbb{F}_{q}(t), Geom. Funct. Anal. 25 (2015), no. 3, 671–732. MR 3361770
  • [CD08] A. C. Cojocaru and C. David, Frobenius fields for Drinfeld modules of rank 2, Compos. Math. 144 (2008), no. 4, 827–848. MR 2441247
  • [Coh81] S. D. Cohen, The distribution of Galois groups and Hilbert’s irreducibility theorem, Proc. London Math. Soc. (3) 43 (1981), no. 2, 227–250. MR 628276
  • [Del74] P. Deligne, La conjecture de Weil I, Inst. Hautes Études Sc. Publ. Math. No. 43 (1974), 273–307. MR 340258
  • [Har77] R. Hartshorne, Algebraic geometry, Springer-Verlag, New York-Heidelberg, 1977, Graduate Texts in Mathematics, No. 52. MR 0463157
  • [HB84] D. R. Heath-Brown, The square sieve and consecutive square-free numbers, Math. Ann. 266 (1984), 251–259. MR 730168
  • [HB02] by same author, The density of rational points on curves and surfaces, Ann. of Math. 155 (2002), 553–595. MR 1906595
  • [HBP12] D. R. Heath-Brown and L. B. Pierce, Counting rational points on smooth cyclic covers, J. Number Theory 132 (2012), no. 8, 1741–1757. MR 2922342
  • [Hoo78] C. Hooley, On the representations of a number as the sum of four cubes. I, Proc. London Math. Soc. (3) 36 (1978), no. 1, 117–140. MR 506025
  • [Hsu96] Chih-Nung Hsu, A large sieve inequality for rational function fields, J. Number Theory 58 (1996), no. 2, 267–287. MR 1393616
  • [IK04] H. Iwaniec and E. Kowalski, Analytic Number Theory, vol. 53, Amer. Math. Soc. Colloquium Publications, Providence RI, 2004. MR 2061214
  • [IR90] K. Ireland and M. Rosen, A classical introduction to modern number theory, second ed., Graduate Texts in Mathematics, vol. 84, Springer-Verlag, New York, 1990. MR 1070716
  • [Jan96] Gerald J. Janusz, Algebraic number fields, second ed., Graduate Studies in Mathematics, vol. 7, American Mathematical Society, Providence, RI, 1996. MR 1362545
  • [Kat02] N. M. Katz, Estimates for nonsingular multiplicative character sums, Int. Math. Res. Not. 2002 (2002), 333–349. MR 1883179
  • [Kat07] by same author, Estimates for nonsingular mixed character sums, Int. Math. Res. Not. 2007 (2007), 19 pp. MR 2359542
  • [Kle86] S. L. Kleiman, Tangency and duality, Proceedings of the 1984 Vancouver conference in algebraic geometry, CMS Conf. Proc., vol. 6, Amer. Math. Soc., Providence, RI, 1986, pp. 163–225. MR 846021
  • [LW54] Serge Lang and André Weil, Number of points of varieties in finite fields, Amer. J. Math. 76 (1954), 819–827. MR 65218
  • [Mun09] R. Munshi, Density of rational points on cyclic covers of n{\mathbb{P}}^{n}, Journal de Théorie des Nombres de Bordeaux 21 (2009), 335–341. MR 2541429
  • [Pie06] L. B. Pierce, A bound for the 3-part of class numbers of quadratic fields by means of the square sieve, Forum Math. 18 (2006), 677–698. MR 2254390
  • [RL05] A. Rojas-León, Estimates for singular multiplicative character sums, Int. Math. Res. Not. (2005), no. 20, 1221–1234. MR 2144086
  • [Ros02] M. Rosen, Number theory in function fields, Graduate Texts in Mathematics, vol. 210, Springer-Verlag, New York, 2002. MR 1876657
  • [Ser97] J.-P. Serre, Lectures on the Mordell-Weil theorem, third ed., Aspects of Mathematics, Friedr. Vieweg & Sohn, Braunschweig, 1997, Translated from the French and edited by Martin Brown from notes by Michel Waldschmidt, With a foreword by Brown and Serre. MR 1757192