Geometric generalizations of the square sieve,
with an application to cyclic covers
Abstract.
We formulate a general problem: given projective schemes and over a global field and a -morphism from to of finite degree, how many points in of height at most have a pre-image under in ? 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 and is a prime degree cyclic cover of . 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 be a global field of arbitrary characteristic and let be a ring of integers in . We denote by the set of places of . For each , we denote by the associated valuation, normalized such that the product formula holds, namely, for every ,
(1.1) |
Using the notation
for projective points, we consider the height function
(1.2) |
and note that it gives rise to the height function
(1.3) |
General Problem
Given a global field of arbitrary characteristic, a ring of integers in , projective schemes , over with fixed models over , and a -morphism , defined over and of finite degree, find an upper bound for the cardinality of the set
that holds for every .
Note that upper bounds for the above cardinality are always given by one of the two cardinalities below,
For example, when is a number field of degree over , by Schanuel’s theorem (e.g. [Ser97, §2.5 p. 17]), there exists an explicit positive constant such that, as ,
(1.4) |
As a second example, when is the function field of an absolutely irreducible projective curve over , of genus , as an immediate consequence of [Ser97, §2.5, Thm. p. 19], there exists an explicit positive constant such that, as ,
(1.5) |
As usual when navigating between the number field and the function field settings, the parameter in (1.4) was replaced by in (1.5).
In our general problem, for nontrivial choices of , we seek a nontrivial upper bound, namely a bound that grows more slowly than the trivial bound, as (respectively, as as a function of , or of , or of both and ).
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 be a number field of degree , let be an integer, and let be an irreducible (non-linear) projective variety in . Serre seeks an upper bound in for the cardinality of the set
(1.6) |
The trivial upper bound for (1.6) is , 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 such that, for all ,
(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 . Serre then poses the question of whether (1.7) can be improved to
(1.8) |
for some , without specifying whether the implied -constant might depend on any of ; 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 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:
-
(I)
to provide a nontrivial upper bound for the General Problem when , , , is a prime degree cyclic cover of , and is the natural projection;
-
(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 .
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: , , , a prime degree cyclic cover of , and the natural projection.
To be precise, let be an odd rational prime power, an integer, an integer, and a rational prime such that . We set
and take as the projective scheme in the weighted projective space defined by the weighted projective model
(1.9) |
for some polynomial of total degree in . We take
as the natural projection defined by
(1.10) |
Our interest is in estimating, from above and as a function of ( fixed, ), the counting function
(1.11) |
or, equivalently, the counting function
where denotes the degree of as a polynomial in .
Note that, in this setting, the trivial bound is
(1.12) |
In contrast, the function field analogue of Serre’s conjecture (1.8) suggests that it might be possible to prove, under appropriate conditions on , that there exists some constant for which
(1.13) |
with the implicit constant possibly depending on .
If of degree is such that defines a nonsingular projective hypersurface, Browning and Vishe’s work in [BV15] implies that
(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 which is irreducible under the condition on ; 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 (upon omitting an analysis of how the limit impacts the dependence of the -constant on ), the upper bound appearing in (1.13), as long as the defining polynomial is such that defines a nonsingular projective hypersurface.
Our first main theorem is:
Theorem 1.1 (Counting Rational Points on a Prime Degree Cyclic Cover of ).
Let be an odd rational prime power, an integer, a rational prime, and a homogeneous polynomial of degree in , with . Assume that:
-
(i)
;
-
(ii)
defines a nonsingular projective hypersurface in .
Let be the projective scheme in the weighted projective space defined by the weighted projective model (1.9). Let be the projection (1.10). Then for all the quantity defined in (1.11) satisfies the bound
where the implicit constant depends on , and .
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 -th power values of a homogeneous polynomial . For more information on the way in which the implicit constant depends on , see §9.
To put Theorem 1.1 in context, let us recall the current state of knowledge toward Serre’s conjecture (1.8) when . For , Broberg [Bro09] proved a weak form of Serre’s conjecture (with in place of a logarithmic factor) via the determinant method. For and in the case of cyclic covers of degree , the power sieve argument presented by Munshi in [Mun09] leads to the upper bound
(1.15) |
where is the defining polynomial of the cover and is its degree. Recently, Bonolis [Bon21] refined the argument given in [Mun09] and obtained the upper bound
(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 ).) Our result in Theorem 1.1 is thus an analogue over of the result (1.16) over . Note that, in the limit and aside from an analysis of how the limit impacts the dependence of the -constant on , the upper bound (1.15) or (1.16) approaches one of the form conjectured in (1.8).
Over , the best known result is due to Heath-Brown and the third author [HBP12], who proved Serre’s conjecture (1.8) for all in the case of cyclic covers, by combining a sieve method with the -analogue of Van der Corput’s method. It would be interesting to to adapt Heath-Brown and Pierce’s -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 a global field of arbitrary characteristic, a ring of integers in , the height function (1.3) constructed from valuations that satisfy the product formula (1.1), and projective schemes over with fixed models over , and a -morphism, defined over 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 in the affine sense. To clarify, using the notation
for affine points, we will work with the height function in the affine space given by
(1.17) |
focus on the set
(1.18) |
and seek an upper bound, in terms of , for the cardinality of the set
(1.19) |
The typical sieve approach is to derive information about from reductions of modulo primes. Towards this goal, for each finite place , we denote by the associated discrete valuation ring and by the associated residue field. For all but finitely many finite places (in which case we refer to as a place of good reduction for ), we denote by the reduction of modulo . We call ramified at some if is a branch point for the function . For each , we define the set
Moreover, for any nonempty finite set of finite places of good reduction for , we define the subset
(1.20) |
With this notation, we prove the following sieve inequality:
Theorem 1.2.
(Geometric Sieve)
Let be a global field of arbitrary characteristic,
be the ring of integers in ,
and
the height function (1.3) constructed from valuations that satisfy the product formula (1.1).
Let , be projective schemes over with fixed models over
and let be a -morphism, defined over and of finite degree.
For an arbitrary ,
define the sets and as in (1.18), (1.19).
Then, for
any real number
and for
any nonempty finite set
of
finite
places of good reduction for ,
(1.21) |
where, for any ,
Impetus for inequality (1.21) comes from a sequence of papers on sieve inequalities, beginning with the square sieve over of [HB84] (itself inspired by [Hoo78]), which was later developed into the power sieve over in [Mun09] and [Bra15], and into the square sieve over in [CD08]. The square and power sieves over were strengthened by being combined with the -analogue of Van der Corput’s method in [Pie06] and [HBP12]. Most recently, Browning [Bro15] expanded the square sieve over into a polynomial sieve over , while Bonolis [Bon21] developed a related polynomial sieve over 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 and over ) 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 .
Theorem 1.3.
(Geometric Sieve for Prime Degree Cyclic Covers of )
Let be a global field of arbitrary characteristic, let be the ring of integers in , and let be the height function (1.3) constructed from valuations that satisfy the product formula (1.1). Let be integers, take , and let be the projective scheme in the weighted projective space , defined by a model
(1.22) |
with prime such that , and with homogeneous of degree and having the property that the hypersurface in defined by
is nonsingular. Let be the cyclic map of degree defined over by
(1.23) |
For an arbitrary , define and as in (1.18), (1.19). Then, for any nonempty finite set of finite places of good reduction for ,
(1.24) |
where, for each finite place of good reduction for ,
(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 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 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 in Theorem 1.1 be Dwork-regular could be weakened to assume that 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 , we consider the sum
and note that each therein is added with a non-negative weight. Moreover, observe that, if , then the fiber above has at least one element and at most elements, that is,
Therefore, in this case, for each , we have
We deduce that, for every ,
(2.1) | ||||
(2.2) |
which implies that
(2.3) | |||||
Rearranging the terms and using the non-negativity of to enlarge the sum over to , we deduce that
which completes the proof of (1.21).
2.2. Equivalent formulation of Theorem 1.2
Recalling the notation in (1.25), we see that the bound for may be rewritten by expanding the square inside . This shows that
(2.4) |
in which, for all and ,
(2.5) |
and
2.3. Proof of Theorem 1.3: optimizing the choice of
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 to minimize the first term on the right-hand side of (1.21).
To compute this choice, recall that, since is a cyclic map of prime degree , for each of good reduction for and for each ,
Note that in the above we used the primality of .
We deduce that, for each of good reduction for and for each ,
(2.6) |
As such, for each of good reduction for and for each
(2.7) |
Noting that implies , we infer that, for each of good reduction for and for each , we have
(2.8) |
Now, recall the notation of the model (1.22) for the prime degree cyclic cover . Motivated by our upcoming expressions in terms of character sums in §6, for each of good reduction for and for each , we set
(2.9) |
Then (2.7) implies that, for each of good reduction for and for each ,
(2.10) |
while (2.8) implies that, for each of good reduction for and for each ,
We apply these observations in inequality (1.21) of Theorem 1.2. For this, fix and , and expand
as
Applying (2.10), the above sum simplifies precisely to
which reveals that the optimal choice of to minimize is
Using the above choice of in (1.21) of Theorem 1.2, we obtain that
(2.11) | |||||
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
Distinguishing between and , we obtain that the above expression equals
(2.12) |
By (2.6), for each summand in the inner sum of the first double sum above we have
so (2.12) is bounded from above by
(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 . 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 -th power values of a polynomial; and stronger versions of the square or power sieve, combined with the -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 ): given an appropriate polynomial and a set of interest , how many of bounded height have 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 and aims to bound from above the quantity
(2.14) |
for a fixed integer-coefficient polynomial of interest, with . In the most straightforward comparison to our setting, Browning’s polynomial sieve then provides an upper bound for in the form of an inequality like (2.4), but with containing the coefficient 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 with lying in a certain region , we would set and choose . Then we would apply the sieve inequality, including the weight , to bound (2.14) from above. Similarly for the -power sieve and its variants in [Mun09], [HBP12], [Bra15], to count perfect -power values of some polynomial of interest, we would set and define 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 and instead choose for the square sieve or more generally for the -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,
Previous sieve lemmas in number field settings assume that vanishes for sufficiently large, e.g., that if . In the classical settings, in the derivation of the sieve inequality, at step (2.1) the expression counted the number of primes in the sieving set that divided a certain (nonzero) polynomial expression in , so that for some degree ,
under the relevant hypothesis assumed for the support of . (Here, as is standard, counts the number of distinct prime factors, with for any integer .) In such a setting, we would conclude in place of (2.3) that
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 but the quantity 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 , 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 , we gather in this section standard function field notation and remarks pertinent to our forthcoming arguments.
We recall the general setting of Section 1: is a global field of arbitrary characteristic; is the ring of integers in ; is the set of places of ; the valuations associated to the places are normalized such that the product formula (1.1) holds; is the height function on , defined in (1.2) and its companion height function on , also denoted , is defined in (1.3).
We now focus on the particular global function field
and on the ring
where is the power of an odd rational prime . 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 , where is the degree of the polynomial therein. Furthermore, we assume that , where is a rational prime defined as the degree of the cyclic cover therein; this divisibility assumption ensures that contains the -th roots of unity. Finally, the assumption ensures that the weighted projective space is well-defined.
Note that is the simplest instance of a global function field with field of constants , and that is the ring of elements of which have only as a pole. In particular, is a Dedekind domain (and, actually, a Euclidean domain) and plays the role of the ring in the analogy between the arithmetic of and that of .
As usual, we identify a place with a generator of its associated unique maximal ideal. For our particular , we have either that , which we refer to as the place at infinity of , or that any , which we refer to as a finite place or simply as a prime of , may be thought of as a monic irreducible polynomial in . We recall that, for a nonzero polynomial , we use to denote its degree in .
To simplify the exposition, we use the symbol for and the symbol for an arbitrary prime of (that is, a monic irreducible in ). We denote by the completion of with respect to the topology defined by and recall that
We denote by the residue field of and by a fixed algebraic closure of , and recall that is a finite field with elements. We denote by the power of that exactly divides .
The absolute values , on are defined by
and
where . With these definitions, the associated valuations satisfy the product formula (1.1), the height function on becomes
for any , and the height function on becomes
(3.1) |
for any .
Remark that, since has class number 1, for any we can find with , such that and . Reasoning similarly, we can extend the height to by setting for any , which is consistent with (1.17).
In order to employ Fourier analysis on , we extend to by
and consider the additive character
(3.2) |
where 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 . To choose a sieving set 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 be an odd rational prime power, an integer, and a homogeneous polynomial of degree with . Denote by the projective hypersurface defined in by
and assume that is smooth over . Then the following properties hold.
-
(i)
and its dual are geometrically integral, with .
-
(ii)
There exists an absolutely irreducible, homogeneous polynomial which depends only on , such that is defined in by
-
(iii)
There exists a finite set of finite places which depends only on and , such that:
-
(iii.1)
for all finite places , and the projective variety defined by the equation
is a smooth and geometrically integral hypersurface;
-
(iii.2)
for all finite places , the dual variety is geometrically integral and is defined by the equation
in particular, and the notation is well-defined.
-
(iii.3)
for all finite places ,
-
(iii.1)
Proof.
Assertion (i) is the statement of Proposition 11.2(1). As observed in the appendix, for any field , a hypersurface in is the zero set of a nonzero homogeneous polynomial in . Thus implies that is defined by some ; after clearing denominators, we may assume . Absolute irreducibility of is equivalent to geometric irreducibility of , so this proves (ii). Let be the finite set 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 for and an arbitrary integer .
Recall that, in Theorem 1.3, we take and the projective scheme in the weighted projective space defined by a model with prime such that , and with homogeneous of degree and having the property that the hypersurface in defined by is smooth. Moreover, recall that is the cyclic map of degree defined over by .
Note that each element of is of the form , where and where satisfies the equation
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 be an odd rational prime power, an integer, a rational prime, and a homogeneous polynomial of degree in , with Assume the conditions:
-
(i)
;
-
(ii)
defines a nonsingular projective hypersurface in .
For every , let denote the cardinality of the set
Then for any
in which the implicit constant depends on and .
More precisely the implicit constant depends on in terms of and as well as a finite exceptional set of primes determined by ; see §9. (Here, and in all that follows, given a polynomial refers to the maximum degree in of any coefficient of .)
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
for a fixed arbitrary integer . Note that the trivial upper bound for is
(5.1) |
To improve upon this bound using Theorem 1.3, we seek a suitable sieving set of primes of , which we describe below.
We consider the smooth projective hypersurface
whose projective dual we denote by . By parts (i), (ii) of Proposition 4.1, there exists a homogeneous polynomial , absolutely irreducible over , that defines , that is,
Recall that by part (iii) of Proposition 4.1 there exists a finite set of finite primes
(5.2) |
satisfying that proposition.
Upon fixing some positive integer , we define the sieving set of primes as
(5.3) |
Later on in (9.4), we will choose optimally in terms of and ; see also the observation below (5.4).
To understand the growth of , let us recall the Prime Polynomial Theorem [Ros02, Thm. 2.2, p. 14]:
(5.4) |
We infer from (5.4) that, upon taking (see the precise choice in (9.4)), we have that
(5.5) |
for all that are sufficiently large relative to and . We denote the least such by
and refer the reader to §9.2 for more details. While we still allow to be arbitrary, from now on we assume the lower bound (5.5) for .
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 , we write
to mean
5.2. Upper bound for the main sieve term
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 of good reduction for , the condition that is ramified at means that the extension of residue fields defined by is inseparable, which cannot happen in our case since the residue fields are finite, or that has a ramification index at larger than 1, which is equivalent to being zero. Then,
(5.7) |
We will bound each term according to whether or . First we observe that for any nonconstant polynomial we have
We also restrict the degree of , so more precisely, let denote the number of distinct primes of degree that divide a fixed . Then
Taking norms, we have and therefore,
Thus, after we fix with and ,
Here we emphasize, is the degree of in and is the maximum degree of (any coefficient of) in .
5.4. Remaining work
So far we have proved the upper bound
(5.8) |
Recall that our goal is to prove an upper bound for that improves on the trivial upper bound recorded in (5.1). The first term on the right-hand side of (5.8) will be nontrivial as long as . Since the trivial upper bound for the unramified sieve term, the last term on the right-hand side of (5.8), is , 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 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 , 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 , prime , and non-principal multiplicative character of , of modulus , we define a mixed character sum relative to a polynomial by
(6.1) |
where the additive character is defined in (3.2).
Proposition 6.1.
Let be an odd rational prime power, an integer, a rational prime, and a homogeneous polynomial of degree in , with , where, as before, . Assume . Let be integers and assume that
Defining as in (5.3), for all primes with , the unramified sieve term can be expanded as
(6.2) |
in which, for each , the sum is over all non-principal characters of order .
For later reference, we remark that the left-hand side of (6.2) is unchanged if we omit the condition . Indeed, upon observing that
it follows that whenever .
We note that for any polynomial , the trivial bound for , valid for every prime and every , is
(6.3) |
(Of course, this is far from sharp.) We then see that for all primes with , the trivial bound for the right-hand side of (6.2) is
which is larger than the trivial bound for the left-hand side whenever , as will occur in our ultimate choice for . 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 or over ; for our current scope, it will suffice to prove a nontrivial bound for the individual sums , 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 with (the condition need only be specified later). For any , we will rewrite each of the quantities as a character sum by using the following proposition, whose proof we defer to Section 6.4.
Proposition 6.2.
Let be an odd rational prime power and, as before, take . Fix a rational prime with . For any prime and any , we have
where runs over all multiplicative characters on of modulus and order dividing , whose definition is extended from to via the rule
(6.4) |
By Proposition 6.2, for each of and any ,
(6.5) | |||||
where runs over all multiplicative characters on of modulus and order dividing , extended to as in (6.4). From this, we can write
(6.6) |
in which the sum is now over all (non-principal) characters of order and modulus .
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 be an odd rational prime power and, as before, take . For all primes with , for all non-principal multiplicative characters of of moduli , (respectively), for all , and for all integers such that
we have
where are determined by the congruences and the sum is defined in (6.1).
6.3. Application to the unramified sieve term
(6.7) |
Here we must assume that
(6.8) |
in order to apply Proposition 6.3. We keep this assumption from here onwards. Finally, recall the sum over in (5.8) actually restricts to those for which . But as remarked immediately below (6.2), the left-hand side of (6.2) does not change when the restriction 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 be an odd rational prime power, take , let be a rational prime with , and let be a prime. Since , then . We can then consider the set of multiplicative characters
of order dividing on , which forms a group of order . We can extend to be defined over by setting when . These characters can be given explicitely by -power symbols, as described in [Ros02, ch. 3]. It is known that
(6.9) |
see [Ros02, Prop. 3.1, p. 24].
6.5. Proof of Proposition 6.3
As before, let be an odd rational prime power and take . We first prove a lemma about detecting congruences with the additive character .
Lemma 6.4.
Let and . Let be an integer such that
Then
Proof.
We rewrite the left-hand side by using the following indicator functions:
and
With these definitions, we write
To understand the above sum, we use Fourier analysis on , as detailed in [BV15, Section 2]. By [BV15, Lemma 2.1], we obtain
Making the change of variables
for which
by [BV15, Lemma 2.3] we infer that
By [BV15, Lemma 2.2],
Hence
∎
Now we prove Proposition 6.3. Let be primes with , let be non-principal multiplicative characters of of moduli , (respectively), and let . Fix an integer such that
We partition according to the residue classes modulo :
Recall that . Then, on one hand, there exist uniquely determined , such that
on the other hand, by the Chinese Remainder Theorem, there exist uniquely determined elements , such that
Consequently,
Using that , we make the change of variables and and deduce that the above expression equals
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 be the power of an odd rational prime and take Let be an integer and a homogeneous polynomial of degree such that the projective hypersurface defined by in is nonsingular. Assume that Denote by , the nonsingular projective hypersurface, respectively the affine hypersurface, defined by
Denote by the finite set of exceptional primes introduced in Proposition 4.1. For a prime , denote by , the nonsingular projective hypersurface, respectively the affine hypersurface, defined by
and denote by , their duals. Fix a rational prime with . Then, for every prime , for every (non-principal) multiplicative character of of modulus and order , and for every , the mixed character sum defined in (6.1) satisfies the bounds:
-
(i)
provided ,
-
(ii)
provided ,
-
(iii)
provided and ,
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 be a finite field and let be integers. Let be a polynomial of degree , which we write as
for uniquely determined homogeneous polynomials with . We call a Deligne polynomial over if:
-
(i)
;
-
(ii)
the equation defines a smooth, degree hypersurface in .
Relative to the trivial bound for 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 variables of the sum, while case (iii) provides square-root cancellation in all 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 , as the parameter varies over , 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 and seek to bound
(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 be a finite field, an integer, and a non-principal multiplicative character, extended to by . Let , let be a homogeneous polynomial of degree , and let be a homogeneous polynomial of degree . Assume that , , and is nontrivial. Let denote the hypersurface in and similarly let denote the hypersurface in . Assume that has codimension in , and let denote the dimension of the singular locus of . Set and define by Then
Corollary 7.4.
Let be a finite field, an integer, and a non-principal multiplicative character, extended to by . Let be a homogeneous polynomial of degree such that the projective hypersurface defined by in is nonsingular. Then
Proof of Corollary 7.4.
Note that is not a well-defined map on a projective variety, since for example as a projective point, for each , while unless in . The setting of Theorem 7.3 corrects for this, as follows. We set in the variables , so that We define with degree and so that . Then for any so is well-defined as a polynomial map on Furthermore, since , and using the homogeneity described above,
Note that
so that in the notation of the theorem, since by assumption is nonsingular as a projective hypersurface in Moreover, the codimension of in is 2, as required. Hence by Theorem 7.3, the corollary holds.
∎
The corollary immediately implies (i) in Proposition 7.1, upon taking as in the proposition, and any , with and as in the proposition.
7.2. Proof of part (ii) of Proposition 7.1
We consider the case and seek to bound
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 be a finite field, an integer, and a non-trivial additive character. Let be a polynomial of degree . Assume that is a Deligne polynomial over . Then
We apply Theorem 7.5 to the finite field , the integer , the character , and each of instances of polynomials , derived from , as explained in what follows.
Recall that the Gauss sum
satisfies the Riemann Hypothesis (see, for example, [IK04, Prop 11.5 p. 275]):
Note that for each such that , we have a bijection
Then, using properties of the Gauss sum (e.g. [IR90, Prop. 8.2.2 p. 92]), we obtain
(7.2) | |||||
In the last identity, in order to sum back in the contribution from we used that
which follows from the orthogonality of the characters By taking absolute values in (7.2), we deduce that
(7.3) |
We estimate the inner sum above using Theorem 7.5 for each of the polynomials over defined by the congruence
and for the additive character . Using that , , and , we obtain that . Let us write , where , are the uniquely determined homogeneous polynomials in such that . Then . By hypothesis, is nonsingular in and so that is nonsingular; also . Thus is a Deligne polynomial over , and hence also is. By (7.3) and Theorem 7.5, we deduce that
which completes the proof for case (ii).
7.3. Proof of part (iii) of Proposition 7.1
We consider the case , , and seek to bound
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 be a finite field, an integer, a non-principal multiplicative character, extended to by , and a non-trivial additive character. Let be polynomials of degrees with leading homogeneous forms of degree , of degree (respectively). Assume that:
-
(i)
is a Deligne polynomial;
-
(ii)
is a Deligne polynomial;
-
(iii)
in case , the smooth hypersurfaces in defined by and by are transverse, in the sense that their intersection is smooth and of codimension in .
Then
We apply Theorem 7.6 to the finite field , the integer , the characters , , and the polynomials over defined by the congruences
Note that is homogeneous of degree (because ) and that is homogeneous of degree (because ). Recalling that by hypothesis, we deduce that and are indeed Deligne polynomials. Since , they are also transverse. Thus Theorem 7.6 applies, giving
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 be an odd rational prime power, an integer, a rational prime, and a homogeneous polynomial of degree in , with , where, as before, . Assume the conditions:
-
(i)
;
-
(ii)
the projective hypersurface is nonsingular.
Let be integers such that
(8.1) |
Defining as in (5.3),
(8.2) |
To prove Proposition 8.1, we will estimate the absolute value of the innermost term according to different cases of suggested by Proposition 7.1.
8.1. Proof of Proposition 8.1: dissecting the inner sum into cases
Let with and let be fixed. We dissect the set into subsets suggested by Proposition 7.1, as follows.
For each , we consider the smooth projective hypersurface
By part (iii) of Proposition 4.1, the reduction modulo of remains absolutely irreducible and defines the projective dual , that is,
We denote by
the affine varieties defined by the homogenous polynomials , (respectively). We partition according to the following cases:
-
(C1)
, (good-good);
-
(C2)
, (good-bad);
-
(C2’)
, (bad-good);
-
(C3)
, (bad-bad).
In the above partition, the case includes the subcase in which ; similarly, the case includes the subcase in which . Thus the condition implies that ; similarly, implies that . 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 in order to produce an efficient upper bound. The good-good case sums over and 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 according to whether the sum of takes place over the satisfying or ; that is to say,
Within each term, the subscript case (C3) means that we restrict to those such that case (C3) holds. In we note that by applying cases (i) and (ii) of Proposition 7.1,
(8.3) |
Here we used that for each , the number of characters of order modulo is . To bound this efficiently, we will use the fact that in order to show that relatively few pairs of can correspond to the bad-bad case.
In contrast, in , since , then certainly is “bad” for all , since for all . Thus, by applying cases (i) and (ii) of Proposition 7.1, we write
(8.4) |
The heart of the argument in the bad-bad case is thus to count efficiently those for which . We return to this momentarily.
8.2.1. Bounding
To bound , we begin with (8.3). Case (C3) requires that , so that by homogeneity , and analogously . Consequently, the innermost sum over in (8.3) is bounded by
where we let denote the number of distinct prime divisors of an element . Now let denote the largest degree of that appears in a coefficient of Then for with and ,
(8.5) |
where in the last inequality we applied Proposition 11.2 (3) to bound . We also note that under the hypothesis (8.1). In conclusion,
8.2.2. Bounding
Now we turn to bounding , starting from (8.4). The following lemma is the main tool for bounding 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 be an irreducible homogeneous polynomial of degree , and let be such that there is an irreducible polynomial of degree such that remains irreducible in . Then
We remark that arguing as in the proof of Proposition 4.1, Part (iii.1) implies that for all but finitely many , remains irreducible in . Thus, choosing sufficiently large relative to a fixed polynomial of interest, the conditions of Lemma 8.2 will be met.
We are now ready to bound . Apply Lemma 8.2 to (8.4) with ( recall by Proposition 11.2 (3)), and the choices and (note that by (8.1)). We obtain
This is valid as long as there is a prime of degree for which is irreducible modulo . Under the hypothesis of Proposition 8.1, this this will be true for all sufficiently large, say
(8.6) |
for a finite parameter provided by Proposition 4.1, Part (iii.1); we will ensure this with our final choice of in (9.4); see §9. Combining the bounds for and , we obtain that (under (8.6)) the total contribution of the bad-bad case (C3) to the left-hand side of (8.2) is
(8.7) |
Note that, in order for this to be strictly better than the trivial bound , we must have that
(8.8) |
Combined with (6.8), this motivates the hypothesis we currently assume.
8.3. Proof of Proposition 8.1: the good-good case (C1)
When is in case (C1), we apply part (iii) of Proposition 7.1 to estimate each of the character sums , . We obtain
where
and we again used that for each , the number of characters of order modulo is . Note that
Thus the total contribution of in case (C1) into the unramified sieve term is
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 , according to whether the sum of takes place over the satisfying or . We define
The sum is formally defined analogously, with the condition replaced by the condition However, note that under the conditions of case (C2) the sum in is empty. Indeed, if then all primes are bad (that is to say, for all in the sieving set), whereas in case (C2) at least one prime is good (since by assumption ).
8.5. Proof of Lemma 8.2
Let be as in the statement of the proposition. We have that
We complete the sum, by counting for each such that , those with such that :
(8.9) |
where the additive character is defined in (3.2) and we have applied Lemma 6.4. We remark that
Thus we can work one coordinate at a time. Fix any Write , where is a root of . Then is a basis of as an -vector space. Thus, with can be expressed as with the coefficients uniquely determined; this leads to
(8.10) |
Also notice that we have for each fixed that
Combining the above with (8.10), we obtain
Define for any ,
We claim that . To see this, consider for a given ,
Note that is a hyperplane in the vector space for any . We will prove that . Indeed, if , then, since is a basis for , linearity of trace implies that for all . 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 is perfect, we conclude that . Since and , we conclude that each hyperplane lowers the dimension by exactly when going from to as long as . Once we reach , the dimension remains 0. Thus, in particular for a given , we conclude that .
Back to the identity in (8.5), we have shown that
In the last term, corresponding to , we have applied the Lang-Weil bound [LW54, Thm. 1] to count
We now consider the additive character sum above. Given , we start by writing
(8.11) |
Consider the sum over . This sum is independent of for . Indeed, for , write . Setting , and using the homogeneity of , this gives
Combining the above observation with (8.11), we obtain
We conclude by applying the next lemma.
Lemma 8.3.
Let be an irreducible homogeneous polynomial of degree , and let be an irreducible polynomial of degree . Fix , . Then
and
Proof of Lemma 8.3.
We consider an embedding of in by adding an extra coordinate and interpreting as the subset where .
where in the last identity we applied the fact that 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 , with and . This gives
by a second application of Lang-Weil for a variety of codimension 2, and we conclude. ∎
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 as in the statement of the aforementioned theorems, for any sufficiently large positive integer and for any positive integer chosen such that (5.5) holds and
we have
(9.1) |
where in the last inequality we applied the fact that so that and we imposed the additional assumption
(9.2) |
which we will verify momentarily. Our remaining goal is to choose optimally such that the resulting upper bound for improves upon the trivial bound . In what follows, we address the choice of , after which we address the existence of a minimal such that our previous working assumptions on hold for all (namely , and the assumptions (5.5), (8.6), (9.2)).
9.1. Choice of
Looking at the right-most side of (9.1), we see that an initial choice of could be made such that the two terms are balanced, that is,
which is equivalent to choosing
(9.3) |
Since the above is not necessarily an integer, while , as the degree of a polynomial, must be an integer, we write in terms of its integral and fractional parts, and choose to be the former:
where, we highlight,
(9.4) |
and
With this choice of , we obtain:
since
We recognize the bound above as . Recalling that is fixed, we conclude that
(9.5) |
under the assumption that is such that the inequalities and (5.5), (8.6), and (9.2) hold for .
9.2. Choice of
First, let us note that, as a first constraint, must be chosen sufficiently large to ensure that ; it suffices to have for some chosen such that
(9.6) |
Next, recall that, in (5.5) of §5, we introduced the assumption
(9.7) |
which gives rise to a second set of constraints on . Observe that, if we choose such that
(9.8) |
and
(9.9) |
then
which ensures (9.7). Since is a finite set and its cardinality depends on and , we can find such that, for any , (9.8) holds for as chosen in (9.4). Observe that
where we used that . Thus, to ensure (9.9), it suffices to choose such that, for any , we have
In addition, remark that the above condition ensures (9.2).
Next we have to ensure (8.6) holds, namely This similarly will hold as long as for some sufficiently large.
Now note that the inequalities give rise to a final set of constraints for . With our final choice (9.4) of , these inequalities become
(9.10) |
We claim that, for any , the inequalities (9.10) hold for any for some . In what follows, we verify this claim.
Let . Dividing with quotient and remainder by , we find uniquely determined non-negative integers such that
(9.11) |
and
(9.12) |
Note that
Rewriting using (9.11), we obtain that (9.10) is equivalent to
(9.13) |
The left-hand side inequality in (9.13) is equivalent to
for this it suffices that
which certainly holds.
The right-hand side inequality in (9.13) is equivalent to
this will certainly hold if
Since and , the above inequality will hold if
which holds for any .
In conclusion, provided , there exists a positive integer such that, for any , the inequalities , (5.5), (8.6), and (9.2) hold for , so that the sieve process has proved (9.5). On the other hand, for each , we can apply the trivial bound
Thus by enlarging the implicit constant in (9.5) if necessary, it holds for all . 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 be a domain, an integer, and a homogeneous polynomial of degree in . Then, for any finite subset , we have
(10.1) |
We recall the standard proof, which proceeds by induction on (e.g., see [HB02, Thm. 1] for a version of this result when ).
Proof.
In what follows, is the field of fractions of and is a fixed algebraic closure of . In our argument below, it suffices that , that is, we do not need to assume that the coefficients of are in .
Since is homogenous of degree , for each there exists a homogenous polynomial , of degree , such that
Take to be the maximal index such that is not identically zero. Thus,
Our goal is to show that the number of solutions in to the equation
(10.2) |
is at most . We prove this statement by induction on .
When , (10.2) becomes the equation
(10.3) |
whose degree is . Choosing to be a root of the polynomial , provided such a root exists, we note that (10.3) may be satisfied by the pair for any . Since has at most roots in , there are at most choices for . There are at most choices for . As such, in this case, there are at most possible solutions of (10.3). Choosing to not be a root of the polynomial , provided such exists, we see that (10.3) is a degree equation with unknown . Viewed over , this equation has solutions . In total, in this case, there are at most solutions of (10.3). Altogether, we obtain that (10.3) has at most solutions in .
When , we make the inductive hypothesis that
(10.4) |
for any homogenous polynomial . In particular, we assume that (10.4) holds for . Choosing to be a root of the polynomial , provided it exists, we note that (10.2) might be satisfied by for any . Since , by the induction hypothesis we know that there are at most roots . As such, in this case, there are at most solutions of (10.2). Choosing to not be a root of the polynomial , provided it exists, we see that (10.2) gives rise to the degree equation
with unknown . Viewed over , this equation has at most solutions . In total, in this case, there are at most solutions of (10.2). Altogether, we obtain that (10.2) has at most solutions in . ∎
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 , some basic scheme theory will be required to make certain constructions precise.
Let be a field, let be an algebraic closure, and let be the -dimensional projective space over , with homogeneous coordinate ring . A hypersurface in is the zero set of a nonzero homogeneous polynomial . Equivalently, a hypersurface is a closed subvariety (or subscheme) of pure dimension (see the proof of [Har77, Proposition II.6.4]). A hypersurface defined by is smooth if the homogeneous polynomials have no common zeros in . The hypersurface is integral if is an irreducible polynomial over , and it is geometrically integral if is irreducible over .
Lemma 11.1.
Let and let be a smooth hypersurface. Then is geometrically integral.
Proof.
Suppose that is defined by . Then the hypersurface defined by is again smooth, since have no common zeros in . This implies that is irreducible, since any two irreducible components would intersect in a singular point [Har77, Theorem I.7.2]. This means that is a nonsingular variety [Har77, Example 10.0.3], which is necessarily defined by an irreducible polynomial. ∎
In the following we assume . Let be the smooth hypersurface defined by . The tangent space to at a rational point is the hyperplane defined by
This is in fact a hyperplane, as implies for some . This construction can be improved in the following way. Let be the dual projective space parameterizing hyperplanes in affine -space. Concretely, we have , with a point corresponding to the hyperplane . The map defined by can be promoted to a regular map ; using the identification , it is given by the homogeneous polynomials . We call the Gauss map; its image is the dual variety of . Algebraically, the Gauss map is defined by the -algebra homomorphism sending to , and is defined by .
More generally, if is singular then 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 , let be a smooth hypersurface defined by , and let be the dual variety. Suppose that is not a hyperplane, i.e. that .
-
(1)
The dual is a geometrically integral hypersurface.
-
(2)
The Gauss map is generically finite.
-
(3)
If is defined by a homogeneous polynomial , then
where denotes the field of rational functions.
-
(4)
If the field extension is separable (e.g. if ), then .
-
(5)
(Reciprocity) If then .
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 is a homomorphism from a non-zero unitary commutative ring to an integral domain , then is prime. If then is generically finite, as it is then a dominant morphism of varieties of the same dimension. The fact that , 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.
Remark 11.4.
Now we apply the above considerations when is a global field (of any characteristic). If is a finite set of finite places of then we let denote the ring of -integers. A finite place not contained in corresponds to a maximal ideal of ; we denote the residue field by . We wish to prove a version of Proposition 11.2 that holds for all finite places outside of some depending only on . We will do so by “spreading out from the generic fiber”: we will consider varieties (or schemes) defined over , and take the closure of in . If then is the function field of a smooth, projective, geometrically integral curve defined over a finite field , and may be identified with a finite set of (closed) points of ; in this case, is simply the variety . In characteristic zero, we are forced to use some scheme theory, as 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 be a global field, let , and let be a smooth hypersurface defined by of degree at least . Let be the dual variety, and let be a homogeneous polynomial defining . There exists a finite set of finite places of , depending only on and , such that the following hold.
-
(1)
The polynomials and have coefficients in .
-
(2)
For all finite places not in , the polynomial is nonzero (thus since is homogeneous), and the hypersurface defined by is smooth.
-
(3)
For all finite places not in , the dual variety is defined by .
Proof.
The first assertion is true once contains all places with respect to which some coefficient of or has negative valuation. The reduction is nonzero as long as is not one of the finite set of places whose valuation is strictly positive on all coefficients of ; we include such places in as well. Finally, we enlarge to assume that is a unique factorization domain. Note that is a primitive polynomial over by construction: its coefficients have no common factors because we included those in . Similarly, by enlarging if necessary, we may assume that is nonzero for all , so that is primitive over .
Consider the closed subscheme defined by . (If then this is a subvariety of .) Let be the (homogeneous) ideal of defined by and . Since is smooth, the extended ideal (the ideal of generated by the image of ) contains for some . This means that each is a linear combination of and the with coefficients in . Enlarging to contain all places with negative valuation on some coefficient of one of these linear combinations, we may assume . Then for we have , so is smooth.
(Geometrically, the generic fiber of is the hypersurface , and the fiber over a place is . Lemma 11.6 below shows that is the closure of in . The singular locus of is a closed subscheme not intersecting the generic fiber of , so its image in is a finite set of closed points. Deleting these points allows us to assume is smooth.)
Now consider the morphism (regular map) defined by the homogeneous polynomials . As before, this is well-defined because and its partial derivatives have no common zeros. Let denote the image of . Algebraically, the morphism corresponds to the -algebra map sending to , and is defined by . The Gauss map corresponds to , and is defined by for . Hence the dual is defined by the ideal , and the dual is defined by . But is generated by , and we are assuming to be primitive, so by Lemma 11.6. On the other hand, we have by [AM16, Proposition 3.11(iv)], since is prime and is not the unit ideal. Hence is defined by , as desired.
(Geometrically, the restriction of to the generic fiber of is the Gauss map , and the restriction to the fiber over is . Since is irreducible, it defines an integral hypersurface in , which is thus the closure of its generic fiber. But is also irreducible, and and both have generic fiber .) ∎
We used the following lemma in the above proof.
Lemma 11.6.
Let be a unique factorization domain with fraction field , let be a primitive polynomial of positive degree, let be the ideal of generated by , and let be the ideal of generated by . Then .
Proof.
Since is primitive, it is a prime element of , so is prime. Since has positive degree, the ideal 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 and , 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 , 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 , 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