Sums of cubes and the Ratios Conjectures
Abstract.
Works of Hooley and Heath-Brown imply a near-optimal bound on the number of integral solutions to in expanding regions, conditional on automorphy and GRH for certain Hasse–Weil -functions; for regions of diameter , the bound takes the form (). We attribute the to several subtly interacting proof factors; we then remove the assuming some standard number-theoretic hypotheses, mainly featuring the Ratios and Square-free Sieve Conjectures. In fact, our softest hypotheses imply conjectures of Hooley and Manin on , and show that almost all integers are sums of three cubes. Our fullest hypotheses are capable of proving power-saving asymptotics for , and producing almost all primes .
Key words and phrases:
Cubic form, circle method, rational points, Hasse–Weil -functions, correlations1991 Mathematics Subject Classification:
Primary 11D45; Secondary 11D25, 11G40, 11M50, 11P551. Introduction
Let . For each , the cubic surface has a fairly rich set of rational points [segre1943note]. On the other hand, Mordell has suggested that producing large, general integer solutions to for (or for any other fixed ) could be as hard as “finding when an assigned sequence, e.g. , occurs in the decimal expansion of ” [mordell1953integer]*p. 505. The recent work [booker2021question] of Booker and Sutherland resolves Mordell’s specific question for , but the spirit of Mordell’s suggestion certainly remains.
Heath-Brown has conjectured that should have infinitely many solutions for any fixed (see [heath1992density]*p. 623). To represent all even once, one must allow both positive and negative values of . The set has upper density in [davenport1939waring]; Deshouillers, Hennecart, and Landreau have given a model and evidence suggesting a precise density of [deshouillers2006density]. Wooley has shown, unconditionally, that contains integers for reals [wooley1995breaking, wooley2000sums, wooley2015sums]. We now recall a result of Hooley [hooley1986HasseWeil, hooley_greaves_harman_huxley_1997] and Heath-Brown [heath1998circle], and state our main result building on it; we then give further details and background.
Theorem (Hooley; Heath-Brown).
For certain Hasse–Weil -functions, assume automorphy and GRH. Then integers lie in , for any .
Theorem.
For certain Hasse–Weil -functions, assume automorphy, GRH, and the Ratios Conjectures. For a certain polynomial , assume the Square-free Sieve Conjecture. Then integers lie in , and of integers lie in .
Both results require estimating sums that roughly take the following form:
(1.1) |
The utility of GRH in this context has been highlighted by Bombieri; see [bombieri2006riemann]*p. 111. Also, in a function-field setting, [glas2022question] has made [heath1998circle] unconditional, and one could likely simplify our present hypotheses accordingly (since GRH and the Square-free Sieve Conjecture are known over function fields). We only focus on for practical reasons.
For a typical , the integer solutions to are expected to be at least exponentially sparse, if they in fact exist. Heath-Brown’s conjecture would imply that the only obstructions to solubility for , for any , are local. The naive local-to-global analog of Heath-Brown’s conjecture for is known to fail (see e.g. [ghosh2017integral]*p. 691, footnote 3), due to Brauer–Manin obstructions that do not apply to [colliot2012groupe]*p. 1304.
We restrict ourselves to a statistical analysis of over , for lying in carefully chosen regions, conducted using second moments and the variance framework of [ghosh2017integral, diaconu2019admissible, wang2022thesis, wang2023prime]. This connects naturally to difficult open questions in variables, e.g. [hooley1986some]*Conjecture 2, which lie beyond the square-root barrier in the classical Hardy–Littlewood circle method. We will attack these questions under standard number-theoretic hypotheses, primarily regarding -function statistics of Random Matrix Theory (RMT) type. Our work opens with the delta method of [duke1993bounds, heath1996new] (a clean modern form of the Kloosterman method of [kloosterman1926representation], more precise than the upper-bound variant used in [hooley1986HasseWeil, hooley_greaves_harman_huxley_1997]), whose harmonic analysis in principle allows for cancellation over the difficult classical minor arcs. We prove three levels of results, under three levels of hypotheses (the first two levels being relatively soft and qualitative; see Conjectures 1.4 and 1.8). We first recall a general weighted version of Hooley’s conjecture.
Fix a cubic form in variables with nonzero discriminant. Let be the hypersurface in . Let be the set of -dimensional vector spaces over on which vanishes. On a first reading, we suggest assuming that and is diagonal, though we will often work generally.
Given a real , and a function , let
(1.2) |
and if (our main case of interest, in which ), let
(1.3) |
where is the familiar singular series defined in [wang2023_isolating_special_solutions]*§6, and where
(1.4) |
One could attribute to Hooley [hooley1986some]*Conjecture 2, Manin (see e.g. [franke1989rational]), Vaughan–Wooley [vaughan1995certain]*Appendix, Peyre [peyre1995hauteurs], et al. the following conjecture:
(1.5) |
(The original [hooley1986some]*Conjecture 2 for would follow from (1.5) for , applied to a suitable sequence of weights .) See [ding2020variance] for another related problem.
Unconditionally, for [vaughan2020some], when and is diagonal. Under standard hypotheses on the varieties cut out by , one can prove the near-optimal Theorem 1.1 for the same . (Here .) Let be the discriminant polynomial defined in §2. Let
(1.6) |
For each , one can package local data on into a Hasse–Weil -function , defined in §3 along with the rest of the list (1.7).
Theorem 1.1 ([hooley1986HasseWeil, hooley_greaves_harman_huxley_1997]; [heath1998circle]).
Assume and is diagonal. For each , assume Conjecture 1.2 for . Then for all .
Conjecture 1.2 concerns automorphy and the Grand Riemann Hypothesis (GRH). Strictly speaking, Hooley and Heath-Brown assume (in their “Hypothesis HW”) GRH plus certain Selberg-type axioms, e.g. analyticity, but it is natural to assume automorphy in place of such axioms. A nice reference bridging these two perspectives is [farmer2019analytic].
The proof of Theorem 1.1 involves GRH on , surprisingly, coupled with subtle algebro-geometric factors of a different nature. See §2 for details. The -function ingredients extend directly to general , but some other ingredients have yet to be generalized.
GRH gives a pointwise upper bound on for , sufficient for Theorem 1.1. Going past the critical line (in mean value over ) turns out to require much new work, both with -functions and with other factors. In terms of -functions, we mainly use for , as well as , , and for . It will be convenient to assume Conjecture 1.2 in full, but average versions might also suffice.
Conjecture 1.2 (HW2).
Let . Let be one of the Hasse–Weil -functions
(1.7) |
Let be the local factors (including , the gamma factor) associated to .
-
(1)
There exists an integer , and an isobaric automorphic representation of , such that at all places .
-
(2)
has no zeros in the half-plane .
Theorem 1.3.
We use mean-value RMT-type predictions derived from the Moments and Ratios Conjectures of [conrey2005integral, conrey2007applications, conrey2008autocorrelation]. (See also [diaconu2003multiple, vcech2022ratios], and references within, for another important perspective on such conjectures.) The Hasse–Weil -functions form a geometric family, in the sense of [sarnak2016families]. For each , define the Dirichlet series
(1.9) |
Note that and are independent of .
Conjecture 1.4 (R2’).
Suppose . Let be entire, with on the strip for all . Let with . If , then
(1.10) |
The contour in (1.10) runs from to . The left-hand side of (1.10) is independent of unconditionally, or further under (HW2). Other versions of Conjecture 1.4 might also suffice for our purposes; for instance, an analog of (1.10) (with some nontrivial adjustments) might suffice, the precise norm of on the right-hand side of (1.10) is not very important, and one might not need to allow such general .
There are no log factors on the right-hand side of (1.10); the factor in (1.9), and the integral over in (1.10), play a mollifying role. The statement (R2’) can be derived from (HW2) and the Ratios Conjecture 6.3 (R2); see Proposition 6.8. However, there may well be another route to (R2’) not passing through (R2). The statement (R2’) essentially concerns cancellation in the coefficients of over moduli in a dyadic range; see Proposition 6.10. A similar log-free cancellation statement, [li2022moments]*(1.3), has recently played a crucial role in another context. Furthermore, over function fields (or under (HW2) over ), it should already be possible to obtain partial results towards (R2’), using ideas of [soundararajan2009moments, harper2013sharp, bui2021ratios, florea2021negative, bui2023negative] (after Cauchy–Schwarz over ).
For our main results, we also need the Square-free Sieve Conjecture (cf. [miller2004one]*p. 956 and [granville1998abc, poonen2003squarefree, bhargava2014geometric]) for the polynomial , restricted to a certain range . This hypothesis concerns “unlikely divisors” of the outputs of . Such hypotheses can be made unconditional over function fields; see e.g. [poonen2003squarefree]*Lemma 7.1.
Conjecture 1.5 (SFSCp,3).
There exists such that if and , then .
The Square-free Sieve Conjecture (SFSCp,3) is used to confront some novel algebro-geometric issues in our work. We use it to prove, for diagonal , a geometric relative (Conjecture 9.8) of the automorphic Sarnak–Xue Density Hypothesis ([sarnak1991bounds]*Conjecture 1, concerning the extent to which a naive generalization of the Ramanujan Conjecture can fail). This involves sieve-theoretic ideas weighted by somewhat dangerous factors. Even though we do not currently see how to eliminate the use of (SFSCp,3) over , it is fortunate to be able to reduce Conjecture 9.8 to (SFSCp,3) when is diagonal. One can also reduce (SFSCp,3) for diagonal to the case (but not to , it seems); see Proposition 9.13.
After Theorem 1.3, our next main result is the following:
Theorem 1.6.
For another conditional approach to the Hasse principle for when is diagonal, see [swinnerton2001solubility]. Over function fields of characteristic , the Hasse principle for is already known in general when [tian2017hasse]. But our approach has quantitative advantages, which become qualitative when applied to sums of three cubes.
We expect that the condition (1.11) could be removed with enough work. When , it is in fact possible to do this for free: Theorem 1.6 has the following corollary. (A similar but messier statement is possible for arbitrary diagonal when .)
Corollary 1.7.
Theorem 1.6 makes use of a first-moment estimate for the quantity (1.9) over , and over some mildly localized pieces of (“adelic perturbations” of restricted by a parameter ). For and , consider the box
(1.12) |
For , let denote the dilate . For each , let
(1.13) |
Conjecture 1.8 (RA1).
Here is a Dirichlet series with an Euler product, absolutely convergent on the half-plane . The terms are required to tend to as (when , are fixed). See Proposition 6.13 for the main use of Conjecture 1.8.
The Ratios Conjectures include (RA1), even with a power saving; see §6.3.1 for details. The choice (1.13) is permissible according to [conrey2007applications]*(2.11b). The essential feature of (1.13) for us is that is positive and independent of (but its precise constant value is not important). We could get away with a larger choice of if we assumed a correspondingly stronger error term in (1.14); but the present formulation of (RA1) is clean, and easy to compare with other literature.111Note that if , then in (1.9), we have and , where is given explicitly for diagonal by [wang2022thesis]*Lemma 8.6.7. The moments of we consider (over our orthogonal family of -functions )) are thus analogous to central moments over symplectic families. The paper [florea2021negative] seems to come close to proving (RA1) for a different family of -functions, over a function field.
We believe (RA1), like (R2’), represents a tantalizing research direction. There is another direction worth mentioning. In light of the log-free square-root cancellation in conjectured in (1.10), one may hope that “better than square-root cancellation” occurs in , by analogy with [harper2023typical]*(1.2) (an attractive conjecture based on random multiplicative functions and multiplicative chaos; see e.g. [gorodetsky2021magic, harper2023typical] for details and references). If true, this would provide additional cancellation in Proposition 7.5 (one of the key ingredients for Theorem 1.3), and thus provide an alternative approach to Theorem 1.6 (but not to Theorem 1.9).
Theorem 1.9.
Our methods would allow one to prove a version of Theorem 1.9 uniform over small archimedean and non-archimedean perturbations to . By [wang2023prime]*Theorem 1.2, one could then show under the hypotheses of Theorem 1.9 that if , then of primes lie in . One would also be able to give a power-saving analog of Corollary 1.7. But to give full details would obscure our exposition.
The power saving in Theorem 1.9 is small and complicated. (Egregious “exponent divisions” occur in Lemma 7.13 and Proposition 9.5, due to our use of (4.3); and similarly in (10.8), due to (4.4).) It would be very interesting to understand the limits of what one can hope for.
Conjecture 1.10 (RA1).
Suppose , and assume Conjecture 1.2. Let and . Let and . There exists a real , depending only on , such that if and , then
(1.15) |
For Theorem 1.9, certain degenerate residue classes play a larger role in local calculations than for Theorem 1.6. To pacify these residue classes, we need effective control on the variation of an individual local factor over . We work -adically for convenience, using Remark 3.1 to define for each with .
Conjecture 1.11 (EKL).
There exists a nonzero homogeneous polynomial , with , such that for all primes and tuples with and , we have and .
Conjecture 1.11 is an effective Krasner-type statement for . A soft version follows from [kisin1999local], and suffices for Theorem 1.6 but not for Theorem 1.9. When , it should be possible to prove Conjecture 1.11 (with ) using a minimal model for the Jacobian of . In general, one might hope to take to be a power of (or perhaps itself).
1.1. Proof overview
§2 gives background on discriminants and the delta method. The delta method (see (2.10)) connects the point count (from (1.2)) to the local behavior of the intersections over , , , and other rings, as and vary. Cf. (1.1). We highlight several distinct sources of epsilon in the Hooley–Heath-Brown Theorem 1.1, and state a result from [wang2023_isolating_special_solutions] (over ) addressing one such source.
§3 provides background on Hasse–Weil -functions and automorphic -functions.
§4 gives some local control on polynomials and -factors (based in part on [kisin1999local]); we need this for some local estimates and calculations.
§5 gives a useful “smooth framework” for dyadic decomposition and separation of variables. This lets us break certain key sums throughout the paper into more manageable pieces.
§6 derives some -function statistics over , after first doing local calculations (in the spirit of the Deligne–Katz equidistribution theorem) connected to RMT Symmetry Types (cf. [sarnak2016families]*Universality Conjecture). In particular, we state, and build on, some cases of the Ratios Conjectures for . Importantly here, the Ratios Recipe (see §6.3) can only apply once we restrict to . The recipe, naively extended to all , would give false results (failing to detect the special subvarieties on isolated in [wang2023_isolating_special_solutions]).
§7 begins to connect the “pure” -function statistics from §6 to the delta method. We approximate certain Dirichlet series “past” the critical line, in a reasonably simple and uniform way over , and control the resulting “approximation errors” on average. Handling these “errors” demands careful use of Hölder and other ideas. For example, by algorithmic tree-like means, we construct in §7.3 a small exceptional set away from which one may apply the “pure” Conjectures 1.8 and 1.10; it is also here that Conjecture 1.11 plays some key role.
The “mollified” series from (1.9) not only makes the formulas in §6 nicer, but (as we will explain in §7) also holds significance in (2.10); this double significance, though innocent at first glance, secretly reflects a randomness property (connected to Deligne–Katz) stemming from the fact that . Throughout the paper, we take much advantage of the structure of (1.9); this is essential in Conjecture 1.4 and Lemma 6.15, for instance.
§8 proves new integral bounds sensitive to some real geometry (involving discriminants). Our approach shares some important features with [huang2020density] (a beautiful recent paper on approximate integral points). In addition, we have several parameters of interest, and must obtain genuinely multivariate decay. Keeping track of uniformity is tricky.
§9, like §8, proves some new “discriminating” pointwise estimates, but on complete exponential sums instead of oscillatory integrals. We then apply these to formulate and address (under Conjecture 1.5) a geometric analog of the Sarnak–Xue Density Hypothesis. Finite-field geometry (see [wang2023dichotomous]) and the geometric sieve (see [ekedahl1991infinite, bhargava2014geometric]) both play an important role here, as does a nice result of Busé and Jouanolou on discriminants (see Theorem 2.2). It is crucial throughout §9 that (see e.g. Remark 9.2); this reflects a “randomness” not present for quadrics.
§10 ties everything together to prove our main results. We also isolate “axioms” that—if true—would allow for non-diagonal ; see Theorems 10.5, 10.7, and 10.8. Here we only consider ; there are also separate issues for (see [wang2023_isolating_special_solutions]*Remark 1.6).
For Theorem 1.3, see §10.2. Here we use an “entirely positive” Hölder argument over : we do not detect any cancellation over that would go beyond a log-free “Mertens-type heuristic on average” (cf. [ng2004distribution]*Theorem 1(iii)) over . Despite the “decoupling” convenience and power of Hölder, we must therefore be careful in §10.2 to obtain -free bounds. (The structure of §10.2 is inspired by our work with the large sieve in [wang2023_large_sieve_diagonal_cubic_forms].)
The proof of Theorem 1.3 tells us (conditionally) that even if one takes absolute values over in (2.10), the allowance of [hooley_greaves_harman_huxley_1997, heath1998circle] is unnecessary; see (10.14) in Theorem 10.5. Theorem 1.3 also highlights a nontrivial use of the log-free order of magnitude in Conjecture 1.4, a robust qualitative prediction that (if true) could perhaps be explained in other ways (not just following the rather arithmetic Ratios Recipe).
For Theorem 1.6, Corollary 1.7, and Theorem 1.9, see §10.3. In most ranges, the “entirely positive” moment estimates of §10.2 still suffice. But this time, in a few key ranges, we identify cancellation over (via §7). One critical step here is a reduction, via §8, to large moduli in (2.10) (over ), over which certain “mollified” RMT-type main terms vanish (cf. Lemma 6.15). There is also an alternative approach to cancellation (which we do not pursue): instead of Lemma 6.15, we could use the fact that (provided ), where is defined as in (2.9); cf. [wang2021_HLH_vs_RMT]*Observation 10.7.
1.2. Conventions
We write , or , to mean for some . We let denote a quantity that is . We write if . We let denote a quantity such that for every , there exists such that holds for all . When making estimates, we think of , , as fixed, but may still occasionally write (or similar) for emphasis.
We frequently use indicator notation, letting if holds, and otherwise. For any nonempty set with an obvious measure (e.g. the counting measure on a finite set, or the usual Haar measure on ), we let denote the average of over .
We let , and similarly define sets like , , . For , we let denote the -adic valuation of . For , we let denote the totient function, the number of distinct prime factors of , and the radical of .
We let (resp. ) denote the set of smooth compactly supported functions (resp. ). For any function , we let denote the closure of in the domain of ; so for instance, the left-hand side of (1.11) equals .
We let , and . In integrals, we use notation analogous to summation notation. For instance, we write to mean (in conventional notation), and we then write to mean .
We need concise notation for -norms and -norms. If is a quantity depending on a scalar or vector variable (and possibly also on other variables), we write or to denote the -norm of over , according as is a continuous or discrete variable, respectively. If the variable is clear from context, we may omit it.
We let for . When doing calculus in dimensions, a multi-index is a tuple of nonnegative integers (where will always be clear from context). Given a multi-index , we let denote the sum of the coordinates of . For a vector , we let , and write and . For example, if and , then under our conventions,
For (repeated) later use, we fix a function , supported on , with everywhere, on , and . We also fix a radial222in the Euclidean sense (so that the value of depends only on ) function , supported on the annulus (so that and ), such that for all .
2. Background on discriminants and delta
Let be the discriminant of . Let . Define the bi-homogeneous polynomial expression as in [wang2023dichotomous]*§3, so that if , then if and only if the variety in is smooth of dimension . Let
(2.1) |
then for any and prime with , the varieties and in are both smooth. Now recall , from (1.6). For each , let
(2.2) |
Lemma 2.1.
Let be integers. Then , where means . In particular, if , then .
Proof.
See e.g. [heath1998circle]*antepenultimate display of p. 683. ∎
As usual, we write . It is known that the equation defines the projective dual variety of (see e.g. [wang2023dichotomous]*Proposition 4.4). So
(2.3) |
Also, is irreducible in , and has total degree in . In particular, (since ), so by [castryck2020dimension]*Theorem 1, we have
(2.4) |
for all reals . (For diagonal , more is known, e.g. .)
One can express in terms of the discriminant of a cubic form in variables:
(2.5) |
[wang2023dichotomous]*Proposition 3.2. This lets us bring into play a nice result of [buse2014discriminant] (though the weaker classical result [hooley1988nonary]*(84) on p. 62 would also suffice for most of our needs):
Theorem 2.2 ([buse2014discriminant]*Corollary 4.30).
Let be a ring and let be a homogeneous polynomial. If , then there exists such that lies in the homogeneous ideal of generated by
Corollary 2.3.
If , then there exists such that lies in the homogeneous ideal of generated by
Fix satisfying (1.11) if is diagonal, or satisfying
(2.6) |
in general.333For convenience, we will maintain this hypothesis for the rest of the paper, except when specified otherwise. Recall from (1.2). The delta method of [duke1993bounds, heath1996new] allows one to express , up to a negligible error, as a sum over of “adelic” data. We use the precise setup from [wang2023_isolating_special_solutions]*§1 (based on [duke1993bounds, heath1996new, heath1998circle]).
Let be the smooth function given by [heath1998circle]*(2.3); we will need the full definition later, in §8. For each real , let . Let
(2.7) | ||||
(2.8) | ||||
(2.9) |
A simple rearrangement of [wang2023_isolating_special_solutions]*(1.3) gives (for all )
(2.10) |
The infinite sums in (2.10) are essentially finite, by the following standard result:
Proposition 2.4.
For some constant , we have for all . Also, holds whenever and .
Proof.
See e.g. [wang2023_isolating_special_solutions]*Proposition 5.1. ∎
We now recall some background (cf. [hooley1986HasseWeil]*§§5–6) on the sums . It is known that , and thus too, is multiplicative in . The Dirichlet series
(2.11) |
thus has an Euler product. At prime powers, is related to certain point counts in projective space. Given a prime power , let be the set of -points on the variety in , let be the set of -points on the variety in , and let
where . Then let
(2.12) |
If (e.g. if ), then by [wang2022thesis]*Proposition 3.2.4, we have
(2.13) |
If and , then by [wang2022thesis]*Proposition 3.2.6, we have
(2.14) |
Recall , from (1.6). For any set , let
(2.15) |
If , then can resemble (in some sense), leading to the following result:
Theorem 2.5 ([wang2023_isolating_special_solutions]).
Suppose and is diagonal. Then
(2.16) |
unconditionally. In particular, .
Proof.
This follows from [wang2023_isolating_special_solutions]*Corollary 1.2 and (1.7), even if we relax the condition (2.6) to . The earlier paper [heath1998circle] had proved . ∎
Since , we may thus concentrate on . The rest of §2 provides some technical context for our work (in comparison with Theorem 1.1 due to [hooley_greaves_harman_huxley_1997, heath1998circle]), but is not logically necessary for the paper.
If , then by (2.13), (2.14), and (3.4), one might expect to resemble , up to a factor absolutely convergent for ; cf. [wang2023_large_sieve_diagonal_cubic_forms]*(2.4). With this intuition in mind, let us now recall how one can prove (as is key for Theorem 1.1)
(2.17) |
under Conjecture 1.2 (HW2) for , when is even and is diagonal. We find it illuminating to work in this generality, but key here is that when .
Conditional proof sketch for (2.17).
For this proof sketch only, let . Let
(2.18) |
for ; then . Let , be the th coefficients of the Dirichlet series , , respectively. Then the following hold (see e.g. [wang2023_large_sieve_diagonal_cubic_forms]*Proposition 4.12):
-
]
-
[B1’
For , we have .
-
[B2’
For , we have .
The arguments in [hooley1986HasseWeil, hooley_greaves_harman_huxley_1997] and [heath1998circle] can then loosely be interpreted as
-
(H1)
using partial summation over to “factor out” from the sum over in (2.15) (for ), and then bounding the -contribution in ;
-
(H2)
expanding using ;
-
(H3)
using GRH to bound the -contribution in ; and
-
(H4)
using [B1’] afterwards, to bound the -contribution in .
Upon dyadic summation over , one gets .
In place of GRH, one could use an elementary statement in the spirit of a large sieve inequality. One would then use [B2’] instead of [B1’]. (See [wang2023_large_sieve_diagonal_cubic_forms].) ∎
We now diagnose (and sketch “cures” for) the key “sources of ” above:
- (1)
-
(2)
In (H4), the bound [B1’]. (Or [B2’], for the argument of [wang2023_large_sieve_diagonal_cubic_forms].) In fact, upon closer inspection, the proofs of [B1’] and [B2’] each have two sources of :
-
(a)
Good prime factors of , via the “first-order error” present in . (Cure: Replacing with a “better approximation” of .)
-
(b)
Bad prime factors of , via the sometimes large failure of square-root cancellation in individual sums of the form . (Cure: New and old pointwise bounds on , and the average-type Conjecture 1.5.)
-
(a)
-
(3)
The fact that over , each dyadic range contributes roughly equally to the final bound (2.17); cf. [wang2023_large_sieve_diagonal_cubic_forms]*Remark 5.3. If unaddressed, then the following sources of would arise (in our work):
-
(a)
A fatal factor in our proof of Theorem 1.3, via summation over .
- (b)
(Cure: New integral bounds that decay, as , fairly uniformly over .)
-
(a)
-
(4)
The lack of “-care” in bounds and “decay cutoffs” for integrals. The best recorded integral estimates (valid at least for some ) seem to be
(2.19) with , from [hooley2014octonary]*p. 252, (31). Summing over carefully, or summing over carelessly, incurs -losses. (Cure: Summing carefully over , and using the “cure to (3)” over small .)
3. Background on individual -functions
3.1. Geometric background
We need to give a precise meaning to Conjecture 1.2. We first define the necessary Hasse–Weil -functions and their local factors, following [serre1969facteurs]. (Another option, not pursued here, would be to follow [taylor2004galois].) This is technical, but allows us to capture “variation in ” in a representation-theoretic framework. At most primes, the data captured is very concrete; see e.g. (3.4).
For any perfect field , let . Let and . The case of the Riemann zeta function in (1.7) is familiar (with and ), so we focus on the other cases. Let . Let be a prime, and (viewing , as subvarieties of ) consider the -adic Galois representations
It is known that the representations , , , , of have dimensions depending only on , and are pure of weight , , , , , respectively.
Let be one of these five representations, and let , be the dimension and weight of , respectively. Define using Hodge theory, following [serre1969facteurs]*§3.2 (after passing from to a singular cohomology group independent of ); then for certain integers with , we have
Let ; then is holomorphic on the half-plane . Note that if we let , vary, the number of possible functions could be is .
For each prime , we may restrict to ; let denote the reverse characteristic polynomial of geometric Frobenius on (the inertia invariants of ), following [serre1969facteurs]*§2.2. Clearly , with equality if and only if is unramified at . Write , and let
(3.1) |
Because , are complete intersections in , it is now known444thanks to [laskar2017local]*Corollary 1.2 and its proof (cf. [saito2003weight]*Corollary 0.6), which builds on progress of [scholze2012perfectoid] on the weight-monodromy conjecture that the polynomial lies in and is independent of (so that is a multiset of algebraic numbers), and furthermore (for any embedding of into the complex numbers) we have
(3.2) |
for all , . One might also be able to directly (without automorphy) define a conductor and root number for independent of (following [serre1969facteurs] and [deligne1969constantes]), but we need not do so.
Let for any . Given a prime , let for any , and let and . Let
so that for all . Let . Make analogous definitions for and its tensor squares in (1.7), in terms of and its tensor squares. If and , then by (2.12), (3.1), smooth proper base change, and the Grothendieck–Lefschetz trace formula, we have (for instance)
(3.3) | ||||
(3.4) | ||||
(3.5) | ||||
(3.6) | ||||
(3.7) |
where for all we have (by the Weil conjectures, since )
(3.8) |
3.2. Automorphic background
We need some background on automorphic representations of for . We will only work with cuspidal ’s, or more generally, isobaric ’s. These ’s have well-defined -functions , and good formal properties (due to Rankin, Selberg, Langlands, Godement, Jacquet, Shalika, and others):
-
(1)
If is cuspidal, then is primitive in the sense of [farmer2019analytic]*Lemma 2.4, and has certain familiar analytic properties [farmer2019analytic]*Theorem 3.1.
-
(2)
For each isobaric , there is a unique multiset , consisting of cuspidals, such that . We call the ’s cuspidal constituents of .
-
(3)
Strong multiplicity one: If , are isobaric, and for all but finitely many primes , then .
Conjecture 1.2 has a host of standard consequences (which may be treated as a black box).
Proposition 3.2.
Let . Let be one of the Hasse–Weil -functions in (1.7). Assume Conjecture 1.2 for holds for some , . Then the following hold:
-
(1)
; in particular, .
-
(2)
The conductor of satisfies .
-
(3)
Each cuspidal constituent of has unitary, finite-order central character.
-
(4)
is holomorphic on , except possibly for poles at corresponding to trivial constituents of .
-
(5)
is an entire function of order .
-
(6)
has a standard functional equation (with critical line ), involving and some root number .
-
(7)
has real coefficients, is self-dual, and .
-
(8)
Let be a cuspidal or isobaric constituent of . Then for . If denotes the th coefficient of the Dirichlet series , then for all and .
Proof.
If is a prime, then has degree , with equality if and only if ; cf. [farmer2019analytic]*Axiom 3(b) and (3.3).
(1): Compare the degrees of , at a prime .
(2): For some , the local Dirichlet polynomial has rational coefficients for all primes . So by [shin2014fields]*(3.2), there exists such that for all , the representation of has field of rationality , in the sense of [shin2014fields]*Definition 2.2. So by strong multiplicity one, the field of rationality of is . Now consider any . Then has degree , so has degree , whence . By [shin2014fields]*Lemmas 3.11 and 3.13, then, . So .
(3): Let be a cuspidal constituent of , so is a cuspidal automorphic representation of for some . Let be the central character of . Then corresponds to a classical character on , where and is a Dirichlet character of conductor dividing , such that for all primes . But at each prime , if we write , then ; cf. [farmer2019analytic]*(3.3) and its proof. By (3.8) and the algebraicity of the eigenvalues , it follows that for infinitely many primes , we have and . So , i.e. is unitary; and then by the six exponentials theorem (cf. [farmer2019analytic]*proof of Lemma 4.9), so has finite order.
(7): For some , the coefficients of are all rational. Hence has real coefficients. So by (3) and strong multiplicity one, we have and thus is self-dual. The functional equation from (6) then implies , so .
(8): By (3.2), (5), (6), and Conjecture 1.2(2), we have for ; see e.g. [iwaniec2004analytic]*Theorem 5.19 and the ensuing paragraph. But , so the desired bound on follows. One can then prove using Perron’s formula ([iwaniec2004analytic]*Proposition 5.54) and contour integration; cf. [hooley1986HasseWeil]*p. 75, proof of Lemma 10. ∎
4. Local control on polynomials and -functions
We first recall some standard bounds on the local near-zero loci of a fixed polynomial; cf. [serre1981quelques]*p. 146, (57) and [ganzburg2001polynomial]. Given , let be the number of solutions to , and let be the Lebesgue measure of the set . Similarly, for , define
Proposition 4.1.
Suppose has leading term with and . Then
(4.1) |
uniformly over integers . Also, uniformly over reals , we have
(4.2) |
Proof.
Corollary 4.2.
Fix a nonconstant polynomial , where . Then
(4.3) |
uniformly over integers . Also, uniformly over reals , we have
(4.4) |
Proof.
We now turn to local -factors. (For with , we define using Remark 3.1.)
Proposition 4.3 ([kisin1999local]).
Fix a prime and a tuple with . Then there exists an integer , depending only on and , such that for all tuples with , we have and .
Proof.
Let be the open subscheme of . Let be the closed subscheme of . Let be a prime. The maps and induce local systems on . By [kisin1999local]*Theorem 5.1, case (2), and its proof, there exists a -adic neighborhood of in such that the Galois representations for all factor through in an appropriate sense. So the isomorphism class of the representation is constant over , and thus is too. ∎
Lemma 4.4.
Proof.
Suppose with and . Then and , and thus (because for every , we have and thus and ). But then , so a similar argument gives , whence
(4.6) |
Therefore, if lies in (4.5) for some , then (4.5) indeed contains the set .
Next, let . The set is closed in , and thus compact. The function on is locally constant (by (4.6)), and thus has a (finite) maximum value. Therefore, for all sufficiently large in terms of , the set (4.5) contains . But by (4.3) (or by [serre1981quelques]*p. 146, Corollaire), the measure of tends to as . ∎
Remark 4.5.
If Conjecture 1.11 holds, then (whenever ).
5. General separation technique
At several points in the paper, we need to understand quantities that vary with and . A key tool we use for this is smooth dyadic decomposition (minimizing convergence issues) followed by separation of variables (via Mellin inversion); see Lemma 5.2.
Let and for . For any and , let and . Given and , let
(5.1) |
so that Mellin inversion (see e.g. [iwaniec2004analytic]*p. 90, (4.106)) gives (for all )
(5.2) |
Proposition 5.1 (Standard Mellin bound).
Fix a compact set . Let and . Let with . Then
(5.3) |
for all . Here .
Proof.
Fix a function , supported on , with
(5.4) |
Lemma 5.2 (“Dyadic partial Mellin summation”).
Let . Let be a function. Let be a smooth function supported on for some real . Let ; let and . Then
(5.5) |
Proof.
For later use, we now do a general dyadic calculation.
Lemma 5.3.
Let with . Let . Then
are both .
Proof.
We may assume , or else the sum and integral both vanish. By subtracting , , by , we may also assume . If , then the result follows from the bound and a geometric series (or corresponding integral). If , then separately considering and leads to the result. ∎
6. Statistics of families of -functions
6.1. Computing local averages
For convenience, let and . Let be the th coefficient of the Dirichlet series . We have
(6.1) |
by (3.1). So if , then (3.4), (3.6), (3.7), and imply
(6.2) |
The local statistics of , over play a basic role in the global statistics of over . To prove that certain averages exist, we will use (3.2) and Lemma 4.4. But to estimate said averages, we will take a point-counting approach (though one could use monodromy groups instead; see e.g. [sarnak2016families]*§2.11). The result is Proposition 6.1 below.
Let be a region of the form , where are compact intervals of positive length. Let and . Let be the average of over (assuming this set is nonempty). Let .
Proposition 6.1 (LocAv).
The following two limits exist, and are independent of :
The quantity is multiplicative in : if , then
(6.3) |
The quantity is multiplicative in : if , then
(6.4) |
Now let be a prime, and let be integers. Then (uniformly over )
(6.5) |
Furthermore, if , then
(6.6) | |||
(6.7) |
Proof.
For convenience, define to be if , and if . (Here we allow , or more generally, .) By (3.2) and (6.1), we have
(6.8) |
(In contrast, for , we have and .)
Since , we have (since by (2.4)). On the other hand, if and , then by Lemma 4.4, the quantity depends only on , unless lies in one of exceptional residue classes of (or ) modulo . This, together with (6.8) and the Chinese remainder theorem, implies that (for any ) the three quantities
all exist and equal one another. Similarly, the following exist and equal one another:
This establishes the required existence, independence, and multiplicativity of limits.
6.2. The Sarnak–Shin–Templier framework
Let be an isobaric automorphic representation over corresponding to in Conjecture 1.2. Let be the set of for which is cuspidal, self-dual, and symplectic, in the sense of [sarnak2016families]*p. 533. For each , the -function has a pole at , whence there exists an isobaric automorphic representation over with
(6.12) |
Proposition 6.2.
Let . Then .
Proof.
We want to show that the family indexed by is essentially cuspidal, self-dual, and symplectic (in the sense of [sarnak2016families]*p. 538, (i)–(iii)), with a power-saving exceptional set. We follow the GRH strategy suggested in [sarnak2016families].
Let be as in §1.2. Fix and let .
By Proposition 3.2(7), each is self-dual. Let be the set of for which is cuspidal. Then has a pole at of order exactly if , and at least if ; this follows from the theory of unramified Rankin–Selberg -functions (cf. [farmer2019analytic]*proof of Lemma 2.3). A calculation with (using (3.2), GRH, and [iwaniec2004analytic]*§5.6’s Exercise 6 and §5.7’s Theorem 5.15) then yields
(6.13) |
On the other hand, has a pole at if ; this follows from [sarnak2016families]*p. 533. So a calculation with gives
(6.14) |
But for all primes and tuples with , we have , and we can use (3.4), (3.5), (3.6) to write and (since ). So the left-hand side of (6.13) equals
by Poisson summation and (6.11). Similarly, by (6.11) and (6.10), the left-hand side of (6.14) is . Thus from (6.13) we get , and from (6.14) we get . Now let . ∎
6.3. The Ratios Recipe
By Proposition 6.1 and [sarnak2016families]*pp. 534–535, Geometric Families and Remark (i), the recipe [conrey2008autocorrelation]*§5.1 makes sense for the family . We will soon derive Conjecture 1.8 accordingly, along with the following:
Conjecture 6.3 (R2).
Here is an Euler product absolutely convergent for .
Before proceeding, we make some remarks on our specific Ratios Conjectures.
Remark 6.4.
Remark 6.5.
We do not order our families by conductor (or by discriminant, for that matter). We are indexing by different level sets, as is natural for families like ours; cf. [sarnak2016families]*p. 535, Remark (i); and p. 560, second paragraph after (25).
Remark 6.6.
6.3.1. Deriving (RA1)
To derive Conjecture 1.8, first use (1.9) to write in terms of , and then replace each term on the left-hand side of (1.14) with its “naive expected value over as ” (computed using Proposition 6.1), i.e. the Dirichlet series
(6.16) |
It turns out that the series (6.16) behaves much like , as we now explain.
Define to be the product of (6.16) and , so that (6.16) factors as . Let be the th coefficient of the Dirichlet series . Then is multiplicative in by (6.3); and by (6.5), (6.6) we have
(6.17) |
(because and ). So has an Euler product, and satisfies
(6.18) |
for (for any ). On the other hand, has a pole of order at , and thus is more dominant than in (6.16).
In view of the above, the Ratios Recipe [conrey2008autocorrelation]*§5.1 produces the conjecture
(for , , as in Conjecture 1.8). This rearranges (upon division by ) to (1.14), giving Conjecture 1.8. Furthermore, the fullest Ratios Conjectures include a power-saving error term, leading naturally to Conjecture 1.10.
Here , are called polar factors (or polar terms). As we will see shortly, an additional polar factor, , arises in (R2).
Remark 6.7.
The Ratios Recipe involves the approximate functional equation for when there are ’s in the numerator, but not when there are only ’s in the denominator.
6.3.2. Deriving (R2)
For Conjecture 6.3, use (1.9) to write in terms of , and then replace each term on the left-hand side of (6.15) with its “naive expected value over as ” (computed using Proposition 6.1), i.e. the series
which by (6.4) factors as for some Euler product . If , then by (6.5) and (6.7), we have
(6.19) |
(The justification is similar to that for . Note in particular that if is the th coefficient of the double Dirichlet series , then and .)
6.4. From (R2) to (R2’) and (R2’E)
Write . Conjecture 6.3 implies, uniformly over and , that for , we have (for all )
(6.20) |
this follows for by Conjecture 6.3, and for by GRH (see Proposition 3.2(8)). Using (6.20), we proceed to derive Conjecture 1.4 (R2’).
Proof.
Let . Let with . The left-hand side of (1.10) is independent of , since each integrand is holomorphic on by GRH. Now shift contours to , and expand squares using self-duality of the -functions in (1.9) (see Proposition 3.2(7)), to equate the left-hand side of (1.10) with the quantity
After switching the order of and in (using Fubini), and plugging in the estimate (6.20) and the bound for each , we get the estimate
(6.21) |
where and
Here , so . And by Cauchy–Schwarz,
(6.22) |
Now let , and assume is large enough that . Shifting (in ) from to yields , where
comes from the residue of at , and where on we use (6.19) and the consequence of RH to be able to take
As a stepping stone from Conjecture 1.4 (R2’) to Conjecture 7.4 (R2’E’), we now state (R2’E). Let be the th coefficient of the Dirichlet series .
Conjecture 6.9 (R2’E).
Fix a function . Let . Let with . Then for some real depending only on , we have
(6.23) |
6.5. From (RA1) to (RA1’E)
We now build on Conjectures 1.8 and 1.10, introducing flexible weights over . We need some terminology on residue classes of .
Definition 6.11.
If (where ), let be the modulus of , and let and (where , are defined as in §6.3.1). Given a nonempty set of residue classes , let .
Let be a partition of into finitely many residue classes . (In §7.3, we will construct the partitions needed for our main results.) Let be a compact set. Let be a smooth function supported on . Let
(6.24) |
Conjecture 6.12 (RA1’E).
Let be reals with . Then the quantity
(6.25) |
is , for some real .
Proof.
(It would be nice to prove this only assuming (1.14) for , say, but it will be convenient to assume (1.14) for all , as in Conjecture 1.8.)
By (5.2), we have , so that
Let be a real parameter; soon below, we will let tend slowly to infinity as . Recall from (1.12). Weight is supported on , so the triangle inequality and the previous display imply that (6.25) is at most
Uniformly over and , Proposition 5.1 and (6.24) give (for all )
(6.26) |
For each pair , choose an element of , if such an element exists. Let be with in place of , and let be with in place of . Clearly .
We need to split further according to the size of , with some analytic care (keeping in mind the entireness hypothesis on in Conjecture 1.4). Let . The following hold uniformly over in any fixed finite interval:
-
(1)
For all , we have .
-
(2)
If , then .
-
(3)
If , then .
Let be with in place of , so that
(6.27) |
We first bound and . Plugging (1), (6.26) (with ), and Conjecture 1.8 (for ) into , we find that
which is in turn . Similarly, plugging (2), (6.26) (with ), GRH (see Proposition 3.2(8)), and (6.18) into reveals that
which (if , say) is (since ).
By choosing appropriately, we may ensure both (i) that as , and (ii) that the two “” terms in the previous paragraph are .
It remains to bound and . To handle both at once, we need a Fourier analog of Lemma 5.2. Let be one of the functions (given with ), (given ). Let
(6.28) |
Note that , are entire (and rapidly decaying in vertical strips), so is entire (and rapidly decaying in horizontal strips). By Parseval’s theorem and (5.4), we have
So by (6.28). By Fourier inversion applied to , we get
By analytic continuation, it follows that for any , the quantity
(6.29) |
equals , and thus has absolute value at most
(6.30) |
But for all , we have (by the definition of , if ; and then by analytic continuation, in general), and thus . Using (6.28), Proposition 5.1 (cf. (6.26)), and the definitions of (see (6.24)) to bound pointwise, we then get (by taking )
(6.31) |
for all reals and integers . (If , the factor of in (6.31) arises from the bound . If , the factor of comes from (3) over , and from the decay of in , if .)
Note that for all , we may apply (6.18) and Proposition 5.1 (after shifting contours to ) to get (provided )
(6.32) |
In view of the bound (6.30) for (6.29), we may now apply (6.31), (6.32), Conjecture 1.4 (with , , in place of , , ), and Cauchy–Schwarz to get
which is by Lemma 5.3 (provided and , and ). The same holds for . Thus Conjecture 6.12 holds with . ∎
Proposition 6.14 (RA1’E).
Proof.
Lemma 6.15.
Fix a real . Suppose and for all reals . Let be reals. Then
(6.33) |
6.6. Bounding exterior squares
For each , let denote the th coefficient of the Dirichlet series . Recall , from (2.2). Assume Conjecture 1.2.
Proposition 6.16.
Let be an even integer. Let with . Then
(6.34) |
Proof.
Let be as in §1.2. Let denote the left-hand side of (6.34). Let denote with in place of . Then by positivity,
(6.35) |
for all . But if , then (if we let and , and note that is determined by and )
by (3.2) (or (3.8)) and Poisson summation in residue classes modulo (cf. the proof of Proposition 6.2); note that , so the condition automatically implies . If , then for we conclude that
(6.36) |
We now address . Recall from Proposition 6.2. If , then (6.12) and Proposition 3.2(8) (applied to ), when combined with (3.2) at primes , yield . If , we still have the bound due to (3.2). Therefore, Proposition 6.2 yields
for all . It follows that for when , and thus (by (6.36)) for all . Now (6.34) follows from (6.35). ∎
To handle primes , we prove the following (unconditional) result:
Lemma 6.17.
If , then .
Proof.
Proposition 6.18 (2E).
Assume Conjecture 1.2. Fix and . Let with . Let . For some , we have
(6.37) |
Proof.
First suppose and . Let . By multiplicativity, , which is by (3.2). So by Hölder over , the left-hand side of (6.37) is at most times
Switching , yields (by positivity). Then Cauchy–Schwarz over , followed by Lemma 6.17 and Proposition 6.16, gives
where we use to evaluate the sum over . By Hölder, then, (6.37) holds with if , and with if . The general case follows from partial summation and Hölder, since and . ∎
7. Adapting -function statistics to delta
7.1. Factorization
We need to mold the statistics from §6 into a form friendlier for the delta method. We first split the series (from (2.11)) into more manageable pieces. Recall , from (2.2). Given , consider the factorization , where
(7.1) | ||||
(7.2) |
One can approximate using Hasse–Weil -functions. It would be nice to also relate for to -functions, even in special cases like when and . For now, we study by completely different means (see §9). In §7, we thus concentrate on .
For the rest of §7, assume . We first factor into three pieces: , , .
Definition 7.1.
The factors , in hinder any attempt to apply statistics on to . Fortunately, , turn out to behave as “error factors” on average. Proposition 6.18 lets us handle large moduli in ; note that if is not a square, and
(7.5) |
otherwise (where is as in §6.6). We now prove results to handle large moduli in .
The factor measures the quality of as an approximation to . Recall the “first-order approximation” given by , from (2.18); here . The “first-order error” is only expected to converge absolutely for . As suggested in §2, this is a “source of ” in (2.17). On the other hand, the following result establishes absolute convergence for past the critical line .
Proposition 7.2.
Uniformly over , primes , and integers , we have
(7.6) |
In particular, if , then converges absolutely over .
Proof.
Let . Let , as in §6.1. Suppose first that . Then by (7.1). So by (7.4) and (3.2), we have . So (7.6) holds.
Now suppose . Then by (2.14), and
by (2.13), (3.4), and (3.3). In particular, by (7.4) and (3.2). Furthermore,
To get further cancellation, we multiply by
to get (in view of from (3.6))
(7.7) |
By (7.4), the left-hand side of (7.7) is precisely the local factor of . Thus (7.7) completes the proof of (7.6). The convergence statement on follows from (7.6). ∎
Corollary 7.3 (3E).
Fix . Then uniformly over , we have
(7.8) |
7.2. From (R2’E) to (R2’E’)
We now build on Conjecture 6.9 (R2’E).
Conjecture 7.4 (R2’E’).
Fix a function . Let . Let with and . Then for some real , we have
(7.10) |
Proof.
Let be as in §5, so that and we have (5.4). In view of (7.1) and the factorization , we may use Lemma 5.2 (with and , and ) to write
(7.11) |
where , where , and where
(7.12) |
(Note that , are independent of , .) For all and , Proposition 5.1 gives
(7.13) |
Fix an integer for which . If , then and for all , so lies in the set
(7.14) |
Thus the equality (7.11) remains true if we restrict the integral over to the region . Now set and , for a small constant to be chosen later. Let . Letting , and using Hölder over (restricted to ), we obtain
(7.15) |
since and .
By (7.13) and (7.14), we have . Upon summing (7.15) over , we thus find that the left-hand side of (7.10) is
(7.16) |
Now let . Then (since ), so by Hölder over (writing and for brevity),
(7.17) |
We now bound the necessary -norms. First, , by (7.12) and Conjecture 6.9 (with , in place of , ). Second, by (7.12), (7.5), and (6.37) (with , with , and with , in place of , ), we have . Third, by (7.12), the bound , and Corollary 7.3, we have .
7.3. Handling variation of “error factors” for small fixed “error moduli”
We would like to build on Propositions 6.13 and 6.14, but we must first improve our understanding of certain local factors. For each , recall , from Definition 7.1, and let denote the th coefficient of the Dirichlet series
(7.19) |
so that for all , we have
(7.20) |
By Proposition 7.2, , and by (3.2) and Definition 7.1, ; so certainly
(7.21) |
Moreover, if is small (or fixed), we would like to not vary too wildly with .
Given , what data does depend on? Note that is fixed (in terms of ). So by (7.19) and (7.1), the coefficient is determined by the residue class and the local factors for . Therefore, if we define as in Lemma 4.4, then is determined by the residue class , where
(7.22) |
Proposition 7.6.
Let . Let and suppose . Let be a prime. Then if and only if .
Proof.
By symmetry, it suffices to prove the “if” direction. So, say . Then , so by (4.6). Thus . ∎
In what follows, recall the notation from Definition 6.11.
Definition 7.7.
For every , let be the set of residue classes for which there exists a tuple with .
We now construct a partition of into residue classes .
Definition 7.8.
Fix integers . We define through a recursive decomposition process. Let denote the partition of into the residue classes modulo . For each , define in terms of as follows:
-
(1)
If possible, choose a residue class with . Otherwise, let , and skip step (2).
-
(2)
Write , with . Choose with . Choose a prime with . Create by replacing the element with the lifted residue classes (with , say). Formally,
Step (2) can only occur finitely many times (because we require in step (1)). Let . Let .555The result may depend on the choices we make at each step, but all of our estimates based on will apply uniformly over all possible outcomes. Let .
In Definition 7.8, we allow the initial to branch into many different moduli. If we did not do this, then to control for might require us to work with moduli exponentially large in (for some values of ), which would be fatal to our approach to Theorem 1.9 (though perhaps OK for Theorem 1.6). We will eventually apply Propositions 6.13 and 6.14 in residue classes , for some values of . We first unravel the structure of , and provide some control on the exceptional set .
Lemma 7.9.
Let . Suppose for some . Suppose . Then . Furthermore, there exist an index , a prime , and a residue class of modulus with and , such that holds for all .
Proof.
Proposition 7.10.
Let . Let . Then . Furthermore, if and , then .
Proof.
Proposition 7.11.
Let . If , then for all .
Proof.
Suppose . By Definition 7.8, . So , whence (or else we would have by the algorithm in Definition 7.8). By Lemma 7.9 (applied repeatedly), there exists a sequence of primes , with and , such that
(7.23) |
holds for all and . If for each , we apply (7.23) with , then we get . Since , it follows that , and thus . ∎
Let be the density of a residue class in .
Lemma 7.12 (KL’).
Let . Then .
Proof.
Suppose and , where . Then by Proposition 7.11, we have . But by (7.22), we have . Therefore (since ), there exists a prime with . It follows that
for all reals . Taking (using (2.4) on the left-hand side, and Lemma 4.4 on the right-hand side; cf. the proof of Proposition 6.1), we get
(7.24) |
But by Lemma 4.4, the right-hand side of (7.24) tends to as . ∎
Lemma 7.13 (EKL’).
Assume Conjecture 1.11 (EKL). Let and . Then (uniformly over and ).
Proof.
Suppose and . As in the proof of Lemma 7.12, we have and . By Remark 4.5, we have . But by (7.22). Hence is divisible by the integer , where . Since every prime factor of is , there exists with . On the other hand, if , then (since ), so . So for every (if ), there exists , with , such that . Thus
for all reals . Taking (using (4.3) on the right-hand side), we get
But , for all with . ∎
7.4. From (RA1’E) to (RA1’E’)
We now build on Propositions 6.13 and 6.14. Let be a compact set. Let be a smooth function , supported on . Given and reals , let
(7.25) |
Conjecture 7.14 (RA1’E’).
Let , and let be a positive integer. Let be reals with . Then for some , where is defined as in (6.24).
The intermediate parameter here may seem strange, but it will ease our exposition.
Proof.
Plugging (7.20) into (7.25) reveals the equality
Fix a function with and . Let denote a real number to be chosen later. We first analyze the piece
of . For later reference, note that (since )
(7.26) | ||||
(7.27) |
We bound using the Hölder technique behind (the simplest case, , of) Proposition 7.5. Since is the th coefficient of the Dirichlet series (see §7.3), we may write in terms of , , and then apply Lemma 5.2 (with and , and ), to get
(7.28) |
(cf. (7.11)), where and
For every integer , Proposition 5.1 and (6.24) imply (uniformly over , , )
(7.29) |
(where the implied constant may depend on , as well as ).
Since , , are supported on , , , respectively, we have unless , , and for all . Fix an integer satisfying . Then identically unless lies in the set
(7.30) |
(cf. the region from (7.14)). So (7.28) holds even if we restrict to .
But if and , then (7.17) (with , , ) and the subsequent arguments up to (7.18) furnish (via Conjectures 1.2 and 6.9) the bound
(7.31) |
(where , and is as in Proposition 6.18). Upon taking absolute values in (7.28) (after restricting to ), summing over , plugging in (7.29) (with ) and then (7.31), and integrating over , we conclude that
(7.32) |
where and .
We now turn to the sums , for . We first treat . By (7.27), we have . Therefore, by the triangle inequality, is at most the sum of the quantity (6.25) (with ) and the left-hand side of (6.33) (with ). So by Conjecture 6.12 (applicable since ) and Lemma 6.15 (with and ), the sum is at most , where denotes the the expression666The replacement of with the weaker (larger) here may seem strange, but it will be convenient later.
(7.33) |
If is sufficiently small, then (since )
(7.34) |
Now suppose . Let denote an integer, with , to be chosen later. Using Definition 7.8, let
Then by (7.27) (since is a partition of ). By Proposition 7.10, we have for all , so .
By the triangle inequality,
(7.35) |
By (7.21), . However, by Lemma 5.2 (with and , and ) and Proposition 5.1, we have (cf. (7.28), (7.29), and (7.30))
for all integers . Plugging this and into (7.35), and then applying Cauchy–Schwarz over , we get (by Lemma 7.12 and Conjecture 6.9)
(7.36) |
cf. the numerics in (7.31) and (7.32). (Before applying Lemma 7.12, note that , so .)
By Definition 7.8, the quantity is constant over for each . But by (7.21), so we get
We may apply Conjecture 6.12 (with in place of ) and Lemma 6.15 (with and ) with , since and (so that and ). By the triangle inequality applied to the right-hand side of the previous display, we then obtain the bound
(7.37) |
Let denote an integer to be chosen soon. Assembling (7.26), (7.32), and our work on for (see (7.33) for , and (7.36), (7.37) for ), we get (by the triangle inequality) that the sum is at most
(7.38) |
provided and hold for all integers with . But upon replacing in (7.33) with , and summing over , we find that
(7.39) |
It remains to carefully specify parameters. Choose so that . Then let for all integers with . Let (in terms of ) be the largest integer for which ; such an integer exists, because and . Crucially, we have , since the expression is bounded on any finite set of integers .
Proposition 7.16 (RA1’E’).
Proof.
We adjust the proof of Proposition 7.15. Let be a real to be chosen later. For each integer with , let be the largest integer for which ; then (since ), and (by the maximality of ). Let
Using Proposition 6.14 in place of Conjecture 6.12, we find that and (if , then) ; cf. (7.33) and (7.37). Summing over , and writing for convenience, we obtain
(7.40) |
On the other hand, if we plug GRH (Proposition 3.2(8)) into (7.35) (after partial summation over , say, and recalling the definition of from (6.24)) and then use Lemma 7.13 (applicable since and we assume Conjecture 1.11), then (if ) we get
(7.41) |
cf. the use of Lemma 7.12 towards (7.36). On the right-hand side, (since and ), and (since and ). Thus, summing (7.41) over gives
(7.42) |
As for , the bounds (7.31) and (7.32) must be adjusted slightly, since we do not assume Conjecture 6.9. However, by Proposition 3.2(8) and partial summation, the bound (6.23) in Conjecture 6.9 still holds up to a factor of (for any ). Therefore, (7.31) and (7.32) still hold if we replace with ; so .
8. New bounds on the integral factor
Recall from (2.9). As we explained in §2, we need to go beyond the integral estimates from standard sources like [duke1993bounds, heath1996new, heath1998circle, hooley2014octonary] (and the related estimates of [hooley1986HasseWeil, hooley_greaves_harman_huxley_1997]), such as (2.19). We will prove uniform bounds free of epsilons and logs, while also bringing discriminants into the picture via the following consequence of (2.3):
(8.1) |
To give clean, general bounds, we assume (2.6). We prove the following on :
Proposition 8.1.
The fact that increasing is harmless can be interpreted as an instance of “homogeneous dimensional analysis” (and ultimately arises from the homogeneity of , via a beautiful recursive structure due to [heath1996new]; see (8.5) below). The factor measures the “degeneracy” of the real hyperplane section , and it arises in our proof for roughly the same reason that dual hypersurfaces arise in [huang2020density]*(5.4), (5.11)–(5.14).
Morally, in (2.10), Proposition 8.1 lets us “imagine that there are sharp cutoffs” and . Since by (4.4) we typically have , one might thus expect (in view of Proposition 2.4, and our -diagnosis at the end of §2) that should be the “dominant range” on average, and there we have .
To prove Proposition 8.1, we must first dig into some technical aspects of [duke1993bounds, heath1996new]’s -function. Let and ; note that is supported on . Following [heath1996new]*p. 165, let
so that is supported on . For and , let
By (2.7) and (2.9), and a change of variables from to , we have (since )
(8.2) |
The following shows that we may take in Proposition 2.4.
Lemma 8.2 (See [heath1996new]*Lemma 4).
If and , then .
Following [heath1996new], we now build a Fourier transform. Fix such that for all . Let . For any , let
be the Fourier transform of . Writing and (for and ) in (8.2), we get
(8.3) |
where ; cf. [heath1996new]*Lemma 17.
Let , so that . It turns out (see Lemma 8.3) that behaves somewhat like a “fixed” Schwartz function independent of . Thus may be compared with in the classical Dirichlet arc theory, where .
Lemma 8.3.
Let and . Then for all .
Proof.
We may assume if ; this lets us treat all simultaneously.
Since , we have
Since and , we find by induction (and the Leibniz rule) that for some constants , we have
Integrating by parts times in (repeatedly integrating and differentiating the complementary factor), and then taking absolute values, we get
We now put (8.3) in a broader framework. For any Schwartz functions and with , let
(8.4) |
Then by (8.3), we have , where .
Proposition 8.4.
Assume (2.6). Let , be as above (with , Schwartz and ). Then for all and positive reals and , we have
Before proving Proposition 8.4, we first explain why it implies the desired Proposition 8.1. In order to handle , we need a recursion, (8.5), originally observed to first order by [heath1996new]. Without such a recursion, we might suffer for small moduli , as in [hooley1986HasseWeil]*§9’s analysis of “junior arcs” (repaired in [hooley_greaves_harman_huxley_1997] for some purposes, by a clever averaging argument).
Lemma 8.5.
Let . Let and let . Let . Then
(8.5) | ||||
(8.6) |
Proof.
The formula (8.6) immediately follows upon differentiating (8.4) by . It is possible to prove (8.5) by a clever integration by parts (cf. [heath1996new]*p. 182, proof of Lemma 14). We give a slightly shorter argument. Write and in (8.4) to get
(8.7) |
Differentiating both sides of (8.7) by , using and the fact that is independent of , we get (8.5), by (8.7) applied to each of , , . ∎
Proof of Proposition 8.1, assuming Proposition 8.4.
Recall that by (8.3), we have , where . By Lemmas 8.3 and 8.5 (first using the chain rule, (8.5), and Lemma 8.3 when differentiating by , and then using (8.6) when differentiating ), we may thus write as a finite linear combination of integrals , with coefficients , running over a set of at most pairs . (For example, for and , we would use the chain rule and (8.5) to write as
where and .) Proposition 8.1 then immediately follows from Proposition 8.4 (applied to each individual ). ∎
To prove Proposition 8.4, we need the following lemma; the basic principle is familiar (see e.g. [hormander1990analysis]*Theorem 7.7.1) but the treatment of [heath1996new] is ideal for us.
Lemma 8.6 (Non-stationary phase).
Let and . Suppose is supported on , with for all . Suppose is smooth, with and for all . Then
Proof.
See [heath1996new]*Lemma 10 and its proof (repeated integration by parts); see [heath1996new]*§2 for the definition of (which allows for the required uniformity over weight functions). Note that we do not require any explicit upper bound on , or any control on the shape of the compact set . ∎
Proof of Proposition 8.4.
Certainly by (2.6). Since is smooth, we thus have for all . Since is compact, there thus exists such that for all , we have . Let be the union of all dilates of by a scale factor . The set is compact, being the continuous image of a product of compact sets. Also, since the right-hand side of (2.6) is invariant under scaling, we have
(8.8) |
Let be a constant such that for all , we have
(8.9) |
Our plan is to first consider the -integral in (see (8.4)) for a single value of at a time, and then integrate over . Given , , let and . Taking absolute values in gives the trivial bound
(8.10) |
On the other hand, . In particular, if or , then we have and for all , and thus by Lemma 8.6 (provided ). This, together with (8.10) and the bound , gives
(8.11) | ||||
(8.12) |
for all integers . Fix with
Then by (8.11), (8.12), and the triangle inequality, we have (for integers )
(8.13) |
It remains to handle . Suppose first that . Plugging (8.10) and the bound into (8.13), we get . This bound on fits in Proposition 8.4, since and .
For the rest of the proof, suppose , let and , let , let , and let
(The normalization by makes nearly independent of ; this is crucial later.) Note that . Now consider an individual with
(8.14) |
For all , we have , so and . Furthermore, the condition (8.8) implies
(8.15) |
uniformly over , by calculus (see [heath1996new]*Lemma 21 and its proof). We will split into pieces according to the size of ; cf. [huang2020density]* on p. 2061. To avoid the need for explicit stationary phase expansions (which can be messy after the leading term), we will also use a subdivision process inspired by [heath1996new]*§8.
Recall the weights , from §1.2. Let . For reals , let
(8.16) | ||||
(8.17) |
where . By the definitions of , , we have unless there exists (corresponding to ) satisfying and . Therefore, unless . Since for all , we obtain (under (8.14)) the decomposition
(8.18) |
Fix functions , all supported on , such that and we have for all . Let denote with in place of ; clearly .
We proceed based on the following rough idea: if with a large constant, integration by parts over is useful; and if with a small constant, integration by parts over is useful (if with any constant).
Let . Since , the supports of and are disjoint, so . Thus
(8.19) |
where .
Since , we have
(8.20) |
for all . For , recursively define (a sequence of smooth functions supported on ) by setting
(8.21) |
Integrating by parts times in (8.19), we get (cf. [heath1996new]*(5.4), from the proof of Lemma 10)
(8.22) |
We claim that for each , there exists a smooth function of the form , defined in terms of , , (independently of , , ), such that for all , , currently under consideration (namely, satisfying , (8.14), and ), and for all , we have
(8.23) |
This claim can be easily proven by induction on , where for we take
and for we use the product and chain rules in (8.21) (to compute in terms of and its -derivatives), and observe that (for )
(This proof uses the smoothness of , and the nonvanishing of on .)
By (8.9), we have whenever and . Inserting this and the bounds (8.20), (8.14), (8.9), and into (8.23), we find that for all . Thus (8.22) and (8.20) immediately give (for )
(8.24) |
under (8.14). Our derivation of (8.24) has essentially followed the proof of Lemma 8.6 (with a small but important twist: we use to get the factor ), but in some key ranges we need to go further. We need to decide when to integrate by parts over ; this will be informed by the next two paragraphs.
By (8.8) and the inverse function theorem, we know that for each , there exists an open neighborhood of such that the gradient map maps diffeomorphically onto . But is compact (since is compact), so there exists a constant such that for any and with , there exists with .
We apply this as follows. Let and , and suppose . Then there exists with and . Since , Taylor expansion of at gives . Furthermore, implies , so . But by (8.1). Combining the above, we get
(8.25) |
Let , and let .
Our remaining analysis breaks into two cases: and .
Suppose first that . Consider a satisfying (8.14). By (8.15), the right-hand side of (8.24) is . Choosing and inserting (8.24) into (8.18), we then get
since (via the substitution ). But if we simply take absolute values in (see (8.16)) and in (see (8.17)), and then apply (8.15), we get and for . Integrating over in the previous display, we conclude that (under (8.14)). Hence by (8.13) (after writing ) we have
But (under (8.14)). Thus , which suffices for Proposition 8.4 (since and ).
For the rest of the proof, assume , so that . We will integrate by parts over in the integrals
for . It is crucial to work in terms of rather than . Before proceeding, note that inserting (8.18) into (8.13) (and writing ) gives
(8.26) |
Let , and suppose . By the definition of , we then have and , so by (8.25), we have
(8.27) |
For convenience, let for each .
We first bound . For each , the condition (8.14) holds, so if we plug in the definition of (see (8.16)), and then switch the order of , , we may rewrite as
Here is independent of (when is fixed). Therefore, for each , the inner integral (over ) is by (8.27) and Lemma 8.6, because for and we have for all (since is Schwartz). Applying this inner integral estimate for each , and then using (8.15), we get .
One can similarly prove for .
Now suppose . Let denote with in place of . Then . By (8.22), we have (for all integers )
Since is independent of , and we have and for all , we find by (8.20), (8.23), (8.27), and Lemma 8.6 (applied to the inner integral over , for each for which there exists with ) that
for all integers , where in the final step we use (8.15).
Inserting our work from the last four paragraphs (ignoring the last one if ) into the left-hand side of (8.26), and applying (8.26) with , we get the bound
Taking , and evaluating both integrals over (the first being since , and the second being since ), we get , which suffices for Proposition 8.4. ∎
9. New bounds on bad exponential sums
We first provide some new, general, vanishing and boundedness criteria for , which when combined with classical estimates from [hooley1986HasseWeil, heath1998circle] will (under Conjecture 1.5) allow us to break a critical -barrier behind (2.17). Recall , from (2.8), (2.9).
Lemma 9.1.
Let , and let be a prime.
-
(1)
If , then .
-
(2)
If , then .
-
(3)
We have for all integers .
Proof.
(1): If , then trivially, so . Now suppose . Then by (2.1), we have
Therefore, by (2.12) and [wang2023dichotomous]*Theorem 1.1, and Proposition 2.7(2)(1) with , we have . Yet , by (2.12) and the Weil conjectures (since ). Plugging these two estimates into (2.13), we get .
(If and is diagonal, one can also improve (1) to a “codimension-three” statement, by using [wang2023dichotomous]*Theorem 1.3 in place of [wang2023dichotomous]*Theorem 1.1.)
We now turn to (2)–(3). For any vector , let . Write , where and . For integers , let
For any and , Corollary 2.3 (with in place of ) implies
(9.1) |
For any set that can be written as a finite union of residue classes , let be the density of in . Now let be an integer, and let ; then by [wang2023_isolating_special_solutions]*Proposition 7.4, we have
(9.2) |
where (the “restriction to ” of ) is defined as in [wang2023_isolating_special_solutions]*(7.5).
(2): Suppose . Then (since ), so and . In particular, by [wang2023_isolating_special_solutions]*Lemma 7.2. By (9.2) (with and ), it remains to analyze for . By (9.1), we have (since ). We now analyze .
For each , the variety in is singular at the point . Since , this variety is isomorphic to a cubic hypersurface in , and has a singular locus of dimension if (by [wang2023dichotomous]*Theorem 2.3, due to Zak). So if , then by Bézout’s theorem (applied to the system in ; note that ).
(3): Take a counterexample with minimal. Then and . So , because . Furthermore, . By (9.1), we have for all , since and . So (9.2) gives , whence . If , this contradicts [wang2023_isolating_special_solutions]*Lemma 7.2. If , then , so by [wang2023_isolating_special_solutions]*Lemma 7.2(2); again, a contradiction. Now suppose ; then and (since ), so by the minimality hypothesis. But [wang2023_isolating_special_solutions]*Lemma 7.2(2) then gives ; another contradiction. Therefore, no counterexample to (3) in fact exists.
For an alternative, more algorithmic and computational approach to (2)–(3) (at least when is diagonal), see [wang2022thesis]*§7.2 and [wang2021_HLH_vs_RMT]*Appendix D. ∎
Remark 9.2.
If were quadratic (rather than cubic), then Lemma 9.1(1) would be false whenever . See [wang2022thesis]*Remark 7.2.4 or [wang2023dichotomous]*sentence after Theorem 1.1.
Let and for each integer . For any integers , we have (see e.g. [bateman1958theorem])
(9.3) |
Proposition 9.3.
Let . Uniformly over reals with , we have
(9.4) |
Proof.
The case is trivial, so assume . Let . Let
(9.5) |
for each . For each , let be the set of tuples for which . Note that , so .
It is known (e.g. by [heath1983cubic]*Lemma 11) that for all and primes . So for all and , Lemma 9.1(1) implies
Therefore, the left-hand side of (9.4) is at most times the quantity
(9.6) |
Let , and consider an individual . Lemma 6.17, and Hölder over , give
(9.7) |
On the other hand, the scheme in has dimension (since is generically smooth, due to the absolute irreducibility of ). So by a quantitative form of [ekedahl1991infinite]’s geometric sieve (see [bhargava2014geometric]*Theorem 3.3 for the case of prime moduli, which extends to square-free moduli as in [bhargava2021galois]*§5, Case III), we have
(9.8) |
this bound follows from Lang–Weil if (since ; cf. [bhargava2014geometric]*(16)), and from elimination theory if (cf. [bhargava2014geometric]*(17) and [bhargava2021galois]*the three paragraphs after Proposition 33).
We now build on Conjecture 1.5.
Conjecture 9.4 (SFSCq,3).
There exists a real such that
(9.9) |
holds uniformly over reals with .
Proof.
Suppose are reals with .
Case 1: . Recall the notation from §4. Let and . The left-hand side of (9.9) is at most
(9.10) |
Let be a prime and an integer. By (4.3), we have . On the other hand, by Hensel’s lemma (over the smooth locus of ) and Lang–Weil (over the singular locus of ), we have , and thus . So by the Chinese remainder theorem, the right-hand side of (9.10) is
(since ). Since , it follows that the right-hand side of (9.10) is . Hence the left-hand side of (9.9) is .
Case 2: . Suppose . Let be the largest square divisor of ; then . The integer thus lies in , and hence either has a prime factor (so that for some ), or an integer factor (so that for some ). So by Conjecture 1.5 (with ) and Case 1 (with in place of ), the left-hand side of (9.9) is
Since , it follows that (9.9) holds with . ∎
For all and , let
(9.11) |
Close analogs of the moment have been considered for classically (e.g. in [heath1983cubic, hooley1986HasseWeil, heath1998circle]), and for in [wang2023_large_sieve_diagonal_cubic_forms]*Propositions 4.12 and A.1.
Conjecture 9.6 (B2).
Let with . Then .
Proposition 9.7.
Suppose is diagonal. Then Conjecture 9.6 holds. Also, for and , we have .
Proposition 9.7 is essentially due to [wang2023_large_sieve_diagonal_cubic_forms]; but we need to go one step further.
Conjecture 9.8 (B3G).
Let . There exists a nonzero polynomial , and a real , such that if , then .
One can extend Conjecture 9.8 to , almost for free (see Proposition 10.3), but it is not clear to us what the limit is. Also, for , one might hope for to be admissible, by comparing with Proposition 9.3 or [sarnak1991bounds]*Conjecture 1.
Proposition 9.9.
Remark 9.10.
We need a classical pointwise bound on . For integers , let and . Also let and .
Proposition 9.11 ([hooley1986HasseWeil, heath1998circle]; see e.g. [wang2023_large_sieve_diagonal_cubic_forms]*Proposition 4.9).
Assume is diagonal. Let . There exists such that if and , then
One could replace the in Proposition 9.11 with . This would be important if one wanted to try to prove a softer version of Proposition 9.9 (conditional on a softer version of Conjecture 1.5). We do not explore this direction in the present paper, since it would seem to involve complicated divisor-type sums with square-full parts and discriminant divisibility conditions (which might require complicated parameterizations to handle).
Proposition 9.11 is an explicit stratification result for , based on for . It would be very nice to have a usable (perhaps less explicit) replacement for Proposition 9.11 when is no longer diagonal. Work such as [denef1984rationality, pas1989uniform] could conceivably help.
Proof of Proposition 9.7.
Let . Suppose with . For each , Proposition 9.11 implies . Plugging this into (see (9.11)), and using Lemma 2.1 to bound , we get . However, by (9.3), we have
(9.12) |
Therefore, . The statement holds by a similar calculation with Proposition 9.11, Lemma 2.1, and (9.12), ending with
for , ; cf. [wang2023_large_sieve_diagonal_cubic_forms]*(4.11) in the proof of Proposition 4.12. Since , we obtain , proving Conjecture 9.6. ∎
Proof of Proposition 9.9.
Suppose . For each and , let
Let be a real parameter to be specified later. Let
Let ; then , since for . We will bound conditionally (using Conjecture 1.5), and unconditionally (using the diagonality of ). We will lose factors of at first, and then remove later.
We first handle . Given , let
Any integer can be written uniquely as , where , , are pairwise coprime integers satisfying and (for ). We then have
by the definition of . Here by Lemma 9.1(2)–(3). Also, the integer divides . Since (by Lemma 2.1) and , we conclude that
where denotes the set of tuples for which there exists with . Since (for ), the sum over on the right above is , by Propositions 9.5 and 9.3. Since , it follows that
(9.13) |
We next handle , by introducing a new Ekedahl-type idea. For each , let
Let (to be specified); suppose . Now consider an individual and . Here , so by Lemma 9.1(3). And , so
by Proposition 9.11. Thus there exists with . Clearly . Also, , so (since implies ); and . Letting , it follows (by taking with maximal, if such an exists) that
Using Lemma 2.1 to bound , we get
(9.14) |
For notational simplicity, suppose . For , let
For each , we have
by the divisor bound for . Therefore, the contribution to the right-hand side of (9.14) from with (or the analogous condition if ) is
where we use (9.12) in the final step.
On the other hand, for any , there are at most integers for which . (This is because for diagonal , the coefficient of is nonzero for all ; see e.g. [heath1998circle]*(4.2).) Therefore, the contribution to the right-hand side of (9.14) from with (or the analogous condition if ) is
and thus (by (9.12)) .
The previous two paragraphs imply that the right-hand side of (9.14) is . This, combined with (9.13), gives
Suppose , and let ; then we get
since and . It follows (upon redefining , now that we can forget about ) that for all .
It remains to replace with . This is trivial unless is very small relative to . Expanding the square in (9.11), and noting that depends only on , gives
for all . Yet for any and , there exists with , where . So
for all . Hence for all we have
since . For all (including , where ), then,
Therefore, Conjecture 9.8 holds with and (for at first, and then for by Hölder over ). ∎
Remark 9.12.
Handling takes care, because remains roughly the same size even if we restrict to (a rather sparse set). One could take in Proposition 9.9 if we were only interested in for , or for .
Proposition 9.13.
Proof.
(This result is not used in the rest of the paper, so we confine ourselves to a sketch.)
Assume . Recall from (9.5). By [bhargava2014geometric]*Theorem 3.3, we have
(9.15) |
By (9.15), we see that for any given , Conjecture 1.5 is equivalent to the statement
(9.16) |
Let be the smooth loci of the hypersurfaces , , respectively, over . The gradient defines a Gauss map ; let be the inverse image of under this map. The map is an isomorphism over (by the biduality theorem), and thus an isomorphism over (since , are flat over ). In particular, . Furthermore, it is known that lies in the open subscheme of . These two facts, plus (9.15), imply that (9.16) is equivalent to
(9.17) |
We would like to convert each “exists” into a sum. Each on the left-hand side of (9.17) satisfies , so there are at most finitely many possibilities for if . Also, by (2.4). Therefore, the statement (9.17) is equivalent to
(9.18) |
Now suppose (where ), and for each let
Then the left-hand side of (9.18) equals times
(9.19) |
since trivially (by considering the cases and separately) and (since ).
10. Delta endgame
Throughout §10, assume . Recall , from (2.15). Explicitly, we have
(10.1) |
(by Proposition 2.4 and Fubini). We are finally prepared to analyze these sums for .
10.1. Delta decomposition
Consider an individual . In view of (7.1), (7.2) and the factorization , we may use Lemma 5.2 with ,
and , to write
(10.2) |
where runs over , where , and where
(10.3) | ||||
(10.4) |
Since is supported on (by Proposition 2.4) and is supported on , we have unless and for all . Thus identically unless . So unless lies in the set
(10.5) |
(cf. the region from (7.14)). So (10.2) holds even if we restrict to .
Given , let . For integers and reals , let
Assume (2.6). The definition (10.3) and Propositions 5.1 and 8.1 imply, for and (and multi-indices ), that
(10.6) |
For each , we have and , so and . Let and ; then we deduce that
(10.7) |
for some . Also, the volume bound (4.4) implies (for all and )
(10.8) |
because for all and .
10.2. Sharp delta bounds
In [wang2023_large_sieve_diagonal_cubic_forms], we used Cauchy–Schwarz on , over to conditionally prove (2.17) (for diagonal with ) under a large-sieve hypothesis, based on the first-order approximation (see (2.18)). Now, in §10.2, we will use Hölder in a similar spirit to prove Theorem 1.3. This relies crucially on several new features, including the more precise -data captured in (compared to in [wang2023_large_sieve_diagonal_cubic_forms]).
For the next four results, let , let , let , and let . Roughly speaking, (10.11) will be useful when ; and otherwise, (10.9) will be useful for small (relative to ), and (10.13) useful for larger .
Lemma 10.1.
Assume Conjecture 7.4. Then for some , we have
(10.9) |
Proof.
Proof.
Proposition 10.3.
Assume Conjecture 9.8 with some . Then for any real , we have
(10.12) |
Proof.
Lemma 10.4.
Proof.
Case 1: . Then (see e.g. [bhargava2014geometric]*Lemma 3.1). Inserting this into (10.11), we get (10.13) upon writing (for a small ).
Case 2: . Since , we may use (10.12) (with ), Conjecture 7.4 (if ) or GRH (if , using (7.17) and Proposition 3.2(8) to prove Conjecture 7.4 up to a factor of ), and Hölder over to bound the left-hand side of (10.13) by . Since for some , we get (10.13) if is small.
Case 3: The general case. Decompose as . ∎
Theorem 10.5.
Proof.
By (10.2) and (10.5), the bound (10.15) (with and ) implies (10.14). And by (10.1), the bound (10.14) implies . It remains to prove (10.15); for this, we may assume and (since if or , we may increase or while keeping the set the same). To begin, we decompose using (10.7), and then plug in (10.6) (with ); this bounds the left-hand side of (10.15) by times
(10.16) |
where we write (for convenience) and
(10.17) |
Next, note that if , then Conjecture 7.4 holds by Propositions 6.10 and 7.5. So (10.9), (10.11), (10.13) are all at our disposal, no matter what is. Let and , say. Suppose and . Let and .
Plugging (10.8) into (10.9) (for ) and integrating (10.17) over gives
Let . Summing over (using Lemma 5.3 with , , , and , noting that and ), we get
Writing and , and then summing over using Lemma 5.3 (with and , noting that ), we get
since . But (by (10.5)), so the right-hand side is . Letting , and writing for each , we get
by integrating over for each fixed valued of . Thus
(10.18) |
On the other hand, discarding the factor in (10.17) (using ), plugging into (10.11) (for ), and integrating over gives
Writing , and then summing over , we get
But by (10.5); integrating the previous display over thus gives
(10.19) |
Now suppose ; then . Discarding in (10.17), taking in (10.13), and integrating over gives
Writing , and summing over using Lemma 5.3, gives
Here , so the right-hand side is . Let . Each satisfies , and thus , where . Thus the integral of (the left-hand side of) the previous display is
which is (by integrating first over when is fixed, and then integrating over ). This, when combined with (10.18) and (10.19), establishes (10.15) with (where , , and ), after replacing with if . ∎
Remark 10.6.
Our use of Hölder above is fairly uniform (each of (10.9), (10.11), (10.13) being based on “approximately Cauchy–Schwarz”), but the input required could maybe be slightly relaxed by applying Hölder with more varied exponents. For instance, if and , one could work with in and in (the saving from the former drowning out the loss from the latter); cf. (10.13). And if , one might hope to work with in over in a residue class to modulus (with some nontrivial archimedean restrictions on ), and then work with in afterwards.
Proof of Theorem 1.3.
Let . Let for ; then
(10.20) |
Now let . Let for ; then
(10.21) |
(by Hölder), where , say. Since , we deduce (upon inserting (10.21) into (10.20)) that if holds, then (since )
(10.22) |
Now let , and choose with . Then
(10.23) |
But satisfies (1.11) (and thus (2.6)). And we are assuming (for Theorem 1.3) Conjectures 1.2, 1.4, and 1.5; in particular, Conjectures 9.6 and 9.8 hold by Propositions 9.7 and 9.9, respectively. So Theorem 10.5 (with ) gives . But by Theorem 2.5. So by (2.10) (and the definitions (2.15), (1.6)), we have
Hence by (10.23), (10.22). Given (1.8), a standard Cauchy–Schwarz argument then leads to the desired application to (producing a “positive lower density”). ∎
10.3. Delta cancellation
To prove Theorems 1.6 and 1.9, we will complement (10.15) using Propositions 7.15 and 7.16, identifying -cancellation in some pieces of . Since is not compactly supported in , we begin with a decomposition resembling (8.18).
Recall , from §1.2. Let . For all , let (cf. (8.16), (8.17))
(10.24) | ||||
(10.25) |
We have unless there exists satisfying . Using (valid for ), we may thus write (cf. (8.18))
(10.26) |
Recall from (10.5). Let for reals . In terms of from Theorem 10.5, we have
(10.27) |
By (10.1), (10.2), (10.5), and (10.27), the bound (10.15) (when applicable) implies
(10.28) |
This leads to the following results.
Theorem 10.7.
Proof.
Before proceeding, note that Conjecture 7.14 holds by Propositions 6.10, 6.13, and 7.15, since we assume Conjectures 1.2, 1.4, and 1.8. Let be a real number to be chosen later.
Let . For each real , let (in the context of (7.25))
(noting that , so is supported on ). Then after plugging (10.4) into (10.24) and decomposing into residue classes modulo , we get
(in terms of from (7.25)). By Conjecture 7.14 (with , and with , in place of , ) and the trivial bound , we get
provided . Additionally, by (10.6) (with in place of ) and the definition (6.24) of (with , ), we have
Now, for each and (noting that ), let
Applying Conjecture 7.14 with , in place of , , we get
provided . This time, since , the bound (10.6) (with in place of ) and (6.24) (with , ) give
Inserting our bounds on , into (10.26), assuming , we find that if , then the left-hand side of (10.26) is
(10.29) |
Since and , we conclude (upon integrating over , then over with fixed, and finally over ) that
provided and hold with large enough implied constants. Therefore, there exist functions such that if and , then the left-hand side of the previous display is . It follows from (10.28) (with ) that if , then . Finally, suppose , let be the largest integer in with (so that as ), and take , to get . ∎
Proof of Theorem 1.6.
Unconditionally, by (1.3), (2.10), and (2.16), we have
(10.30) |
Now assume (1.11) (so (2.6) holds). Then Theorem 10.7 gives , since Conjectures 9.6 and 9.8 hold by Propositions 9.7 and 9.9, respectively (since we assume Conjecture 1.5). Upon plugging this bound into (10.30), we get (1.5).
The Hasse principle for follows from (1.5), upon choosing with (possible since contains a point with ). Now suppose . Then by [wang2023prime]*Theorem 1.1 (or [wang2022thesis]*Theorem 2.1.8), we find (from (1.5)) that has density in . (See [wang2023prime] for details on this last deduction, which is based on [diaconu2019admissible]. Diaconu assumes an analog of (1.5) over rather quantitatively deformed regions , whereas we work with fixed weights . It would be very interesting to see if there could be any miraculous cancellation or symmetries in the analog of over , but at the moment it seems easier to handle minimally deformed regions.) ∎
Proof of Corollary 1.7.
(Here we drop the assumption (2.6).) Let be a parameter tending to slowly as . Use (1.8) and Hölder’s inequality to upper bound the contribution to from points with . Then use Theorem 1.6 to estimate the remaining contribution to . This gives (1.5) for arbitrary . Hooley’s conjecture follows upon taking a suitable sequence of weights . ∎
Theorem 10.8.
Proof.
Proposition 7.16 applies, since we assume Conjectures 1.2, 1.10, 1.11. We now mimic the proof of Theorem 10.7, while replacing each use of Conjecture 7.14 with Proposition 7.16.
Let . Proposition 7.16 implies that for any real satisfying , we have
Furthermore, for each , Proposition 7.16 implies that
for any real satisfying .
Now let , and suppose
(10.31) |
so that and (thus) . Arguing as we did for (10.29), we find (assuming ) that the left-hand side of (10.26) is
So (upon integrating over , then over , and finally over )
Now let and , . Then (10.31) holds (since and ), and the right-hand side of the previous display is . So by (10.28) (with ) we have
where , say. ∎
Acknowledgements
Many thanks to Amit Ghosh and Peter Sarnak for suggesting the main problem addressed here. I also thank my advisor, Peter Sarnak, for his guidance and encouragement.††This work was partially supported by NSF grant DMS-1802211. I thank Calvin Deng and Yotam Hendel for sharing references for (4.1) and (4.2), and Yotam for discussions related to §9. For early comments, I thank Manjul Bhargava, Andy Booker, Tim Browning, Brian Conrey, Simona Diaconu, Bill Duke, Roger Heath-Brown, Nick Katz, Will Sawin, Bob Vaughan, and Trevor Wooley. Thanks also to everyone listed in [wang2022thesis], and to people from seminars, conferences, and other events. For more recent interactions, I thank Louis-Pierre Arguin, Emma Bailey, Paul Bourgade, Alex Gamburd, Jayce Getz, Jeff Hoffstein, Junehyuk Jung, Victor Kolyvagin, Valeriya Kovaleva, Lillian Pierce, Kannan Soundararajan, Yuri Tschinkel, Katy Woo, Max Xu, Liyang Yang, and Peter Zenz. Finally, I thank my family for their exceptional support.