Diophantine tuples and multiplicative structure of shifted multiplicative subgroups
Abstract.
In this paper, we investigate the multiplicative structure of a shifted multiplicative subgroup and its connections with additive combinatorics and the theory of Diophantine equations. Among many new results, we highlight our main contributions as follows. First, we show that if a nontrivial shift of a multiplicative subgroup contains a product set , then is essentially bounded by , refining a well-known consequence of a classical result by Vinogradov. Second, we provide a sharper upper bound of , the largest size of a set such that each pairwise product of its elements is less than a -th power, refining the recent result of Dixit, Kim, and Murty. One main ingredient in our proof is the first non-trivial upper bound on the maximum size of a generalized Diophantine tuple over a finite field. In addition, we determine the maximum size of an infinite family of generalized Diophantine tuples over finite fields with square order, which is of independent interest. We also make significant progress towards a conjecture of Sárközy on the multiplicative decompositions of shifted multiplicative subgroups. In particular, we prove that for almost all primes , the set cannot be decomposed as the product of two sets in non-trivially.
Key words and phrases:
Diophantine tuples, shifted multiplicative subgroup, multiplicative decomposition2020 Mathematics Subject Classification:
Primary 11B30, 11D72; Secondary 11D45, 11N36, 11L401. Introduction
Let be a prime power, and let be the finite field with elements. Let be a multiplicative subgroup of . While itself has a “perfect” multiplicative structure, it is natural to ask if a (non-trivial additive) shift of still possesses some multiplicative structure. Indeed, as a fundamental question in additive combinatorics, this question has drawn the attention of many researchers and it is closely related to many questions in number theory. For example, a classical result of Vinogradov [36] states that for a prime and an integer such that , if , then
(1.1) |
More generally, inequality 1.1 extends to all nontrivial multiplicative characters over all finite fields; see Proposition 3.1. Inequality 1.1 leads an estimate on the size of a product set contains in the set of shifted squares: if , , and is the subgroup of of index such that , then
(1.2) |
For more recent works related to this question and its connection with other problems, we refer to [17, 27, 31, 37] and references therein. An analogue of this question over integers is closely related to the well-studied Diophantine tuples and their generalizations; see Subsection 1.1.
In this paper, we study the multiplicative structure of a shifted multiplicative subgroup following the spirit of the aforementioned works and discuss a few new applications in additive combinatorics and Diophantine equations. More precisely, one of our contributions is the following theorem.
Theorem 1.1.
Let with . Let . Let and with . Assume further that if . If , then
Moreover, when , we have a stronger upper bound:
Clearly, Theorem 1.1 improves inequality 1.2 implied by Vinogradov’s estimate 1.1 when , and the generalization of inequality 1.2 by Gyarmati [17, Theorem 8] for general , where the upper bound is given by when is a prime. We remark that in general the condition on the binomial coefficient in the statement of Theorem 1.1 cannot be dropped when is not a prime; see Theorem 1.6.
The proof of Theorem 1.1 is based on Stepanov’s method [35], and is motivated by a recent breakthrough of Hanson and Petridis [21]. In fact, Theorem 1.1 can be viewed as a multiplicative analog of their results. Going beyond the perspective of these multiplicative analogs, we provide new insights into the application of Stepanov’s method. For example, our technique applies to all finite fields while their technique only works over prime fields. We also prove a similar result for restricted product sets (see Theorem 4.2), whereas their technique appears to only lead to a weaker bound; see Remark 4.3.
Besides Theorem 1.1, we also provide three novel applications of Theorem 1.1 and its variants. These applications significantly improve on many previous results in the literature. Unsurprisingly, to achieve these applications, we need additional tools and insights from Diophantine approximation, sieve methods, additive combinatorics, and character sums. From here, we briefly mention what applications are about. In Subsection 1.1, we improve the upper bound of generalized Diophantine tuples over integers. Interestingly, Theorem 1.1 is closely related to a bipartite version of Diophantine tuples over finite fields. This new perspective yields a substantial improvement in the result of generalized Diophantine tuples over integers. In Subsection 1.2, we obtain the following nontrivial upper bounds on generalized Diophantine tuples and strong Diophantine tuples over finite fields. Moreover, some of our new bounds are sharp. Last but not least, in Subsection 1.3, we make significant progress towards a conjecture of Sárközy [31] on multiplicative decompositions of shifted multiplicative subgroups. We elaborate on the context of these applications in the next subsections.
1.1. Diophantine tuples over integers
A set of distinct positive integers is a Diophantine -tuple if the product of any two distinct elements in the set is one less than a square. The first known example of integral Diophantine -tuples is which was studied by Fermat. The Diophantine -tuple was extended by Euler to the rational -tuple , and it had been conjectured that there is no Diophantine -tuple. The difficulty of extending Diophantine tuples can be explained by its connection to the problem of finding integral points on elliptic curves: if forms a Diophantine -tuple, in order to find a positive integer such that is a Diophantine -tuple, we need to solve the following simultaneous equation for :
This is related to the problem of finding an integral point on the following elliptic curve
From this, we can deduce that there are no infinite Diophantine -tuples by Siegel’s theorem on integral points. On the other hand, Siegel’s theorem is not sufficient to give an upper bound on the size of Diophantine tuples due to its ineffectivity. In the same vein, finding a Diophantine tuple of size greater than or equal to is related to the problem of finding integral points on hyperelliptic curves of genus . Despite the aforementioned difficulties, the conjecture on the non-existence of Diophantine -tuples was recently proved to be true in the sequel of important papers by Dujella [9], and He, Togbé, and Ziegler [22].
The definition of Diophantine -tuples has been generalized and studied in various contexts. We refer to the recent book of Dujella [10] for a thorough list of known results on the topic and their reference. In this paper, we focus on the following generalization of Diophantine tuples: for each and , we call a set of distinct positive integers a Diophantine -tuple with property if the product of any two distinct elements is less than a -th power. We write
Similar to the classical case, the problem of finding for and is related to the problem of counting the number of integral points of the superelliptic curve
The theorem of Faltings [13] guarantees that the above curve has only finitely many integral points, and this, in turn, implies that a set with property must be finite. The known upper bounds for the number of integral points depend on the coefficients of . The Caporaso-Harris-Mazur conjecture [6] implies that is uniformly bounded, independent of . For other conditional bounds, we refer the readers to Subsection 2.4.
Unconditionally, in [4], Bugeaud and Dujella showed that for , and the uniform bound for any . On the other hand, the best-known upper bound on is , due to Yip [39]. Very recently, Dixit, Kim, and Murty [7] studied the size of a generalized Diophantine -tuple with property , improving the previously best-known upper bound for given by Bérczes, Dujella, Hajdu and Luca [2] when . They showed that if is fixed and , then . Following their proof in [7], the bound can be more explicitly expressed as when is fixed, , and is the Euler phi function. Note that their upper bound on is perhaps not desirable. Indeed, it is natural to expect that would decrease if is fixed, and increases, since -th powers become sparser. Instead, our new upper bounds on support this heuristic.
In this paper, we provide a significant improvement on the upper bound of by using a novel combination of Stepanov’s method and Gallagher’s larger sieve inequality. In order to state our first result, we define the following constant
(1.3) |
for each , where the minimum is taken over all nonempty subsets of the set
and .
Theorem 1.2.
There is a constant , such that as ,
(1.4) |
holds uniformly for positive integers such that .
The constant is essentially computed via the optimal collection of “admissible” residue classes when applying Gallagher’s larger sieve (see Section 5). Note that when , we have , and hence we have . In particular, if is fixed and , inequality 1.4 implies the upper bound
(1.5) |
which already improves the best-known upper bound of [7] that we mentioned earlier substantially. In Appendix A, we illustrate the bound in inequality 1.4: for , we compute the suggested upper bound
of . From Figure A.1, one can compare the bound of in Theorem 1.2 with the bound in [7]. From inequality 1.5, we see is uniformly bounded by . Table A.2 illustrates better upper bounds on for . In particular, we use a simple greedy algorithm to determine for a fixed . We also refer to Subsection 5.3 for a simple upper bound on , which well approximates empirically.
At first glance, Theorem 1.2 improves the bound in [7] of by only a constant multiplicative factor when is fixed. Nevertheless, note that Theorem 1.2 holds uniformly for and as long as . Thus, when is assumed to be a function of which increases as increases, we can break the “ barrier” in [7], that is, , and provide a dramatic improvement.
Theorem 1.3.
There is such that , and
where is an absolute constant.
The proofs of Theorem 1.2 and Theorem 1.3 require the study of (generalized) Diophantine tuples over finite fields, which we discuss below.
1.2. Diophantine tuples over finite fields
A Diophantine -tuple with property is a set of distinct elements in such that is a -th power in or whenever . Moreover, we also define the strong Diophantine tuples in finite fields motivated by Dujella and Petričević [11]: a strong Diophantine -tuple with property is a set of distinct elements in such that is a -th power in or for any choice of and . Unlike the natural analog for the classical Diophantine tuples (of property ), it makes sense to talk about the strong Diophantine tuples in . The strong generalized Diophantine tuples with property in for general and are also meaningful to study: the problem of counting the explicit size of the strong generalized Diophantine tuples with property involves the problem of counting solutions of the equations appearing in the statement of Pillai’s conjecture. Theorem 1.2 can be improved for strong generalized Diophantine tuples with property ; see Theorem 5.2.
The generalizations of Diophantine tuples over finite fields are of independent interest. Perhaps the most interesting question to explore is the exact analog of estimating as discussed in Subsection 1.1. Indeed, estimating the size of the largest Diophantine tuple with property or with property is of particular interest for the application of Diophantine tuples (over integers) as discussed in [1, 7, 8, 16, 24]. Similarly, we denote
Note that when , it is trivial that . Thus, we always assume throughout the paper. In Section 3, we give an upper bound of and . More precisely, we prove the following proposition using a double character sum estimate. We refer to the bounds in the following proposition as the “trivial” upper bound.
Proposition 1.4 (Trivial upper bound).
Let and let be a prime power. Let and . Then and .
For the case , similar bounds of Proposition 1.4 are known previously in [1, 7, 17]. On the other hand, Proposition 1.4 gives an almost optimal bound of and when is a square (Theorem 1.6). Our next theorem improves the trivial upper bounds in Proposition 1.4 by a multiplicative constant factor or when is a prime.
Theorem 1.5.
Let . Let be a prime and let . Then
-
(1)
. Moreover, if is a -th power in , then we have a stronger upper bound:
-
(2)
.
We remark that our new bound on is sometimes sharp. For example, we get a tight bound for a prime when and . See also Theorem 4.7 and Remark 4.8 for a generalization of Theorem 1.5 over general finite fields with non-square order under some extra assumptions.
Nevertheless, in the case of finite fields of square order, we improve Proposition 1.4 by a little bit under some minor assumptions; see Theorem 4.5. Surprisingly, this tiny improvement turns out to be sharp for many infinite families of . Equivalently, we determine and exactly in those families. In the following theorem, we give a sufficient condition so that and can be determined explicitly.
Theorem 1.6.
Let be a prime power and a square, , and . Let . Suppose that there is such that and (for example, if and . Suppose further that , where is the remainder of divided by . Then . If and , then we have the stronger conclusion that .
Under the assumptions on Theorem 1.6, satisfies the required property and . Compared to Theorem 1.5, it is tempting to conjecture that such an algebraic construction (which is unique to finite fields with proper prime power order) is the optimal one with the required property. Given Proposition 1.4, to show such construction is optimal, it suffices to rule out the possibility of a Diophantine tuple with property and of size . While this seems easy, it turned out that this requires non-trivial efforts.
Next, we give concrete examples where Theorem 1.6 applies.
Example 1.7.
Note that a Diophantine tuple with property corresponds to a strong Diophantine tuple over . If is an odd square, Theorem 1.6 implies that the largest size of a strong Diophantine tuple over is given by , which is achieved by . Note that in this case we have .
We also consider the case that , , and . In this case, Theorem 1.6 also applies. Note that implies that , in which case the base- representation of only contains the digit and , so that the condition holds.
One key ingredient of the proof of Theorem 1.5 and Theorem 1.6 is Theorem 1.1. Indeed, Theorem 1.1 can also be viewed as on an upper bound of a bipartite version of Diophantine tuples over finite fields. For the applications to strong Diophantine tuples, Theorem 1.1 is sufficient. On the other hand, to obtain upper bounds on Diophantine tuples (which are not necessarily strong Diophantine tuples), we also need a version of Theorem 1.1 for restricted product sets, which can be found as Theorem 4.2. Indeed, Theorem 1.1 alone only implies a weaker bound of the form for ; see Remark 4.3.
1.3. Multiplicative decompositions of shifted multiplicative subgroups
A well-known conjecture of Sárközy [30] asserts that the set of nonzero squares cannot be written as , where and , provided that is a sufficiently large prime. This conjecture essentially predicts that the set of quadratic residues in a prime field cannot have a rich additive structure. Similarly, one expects that any non-trivial shift of cannot have a rich multiplicative structure. Indeed, this can be made precise via another interesting conjecture of Sárközy [31], which we make progress in the current paper.
Conjecture 1.8 (Sárközy).
If is a sufficiently large prime and , then the shifted subgroup cannot be written as the product , where and . In other words, has no non-trivial multiplicative decomposition.
We note that it is necessary to take out the element from the shifted subgroup, for otherwise one can always decompose as . It appears that this problem concerning multiplicative decompositions is more difficult than the one concerning additive decompositions stated previously, given that it might depend on the parameter . Inspired by 1.8, we formulate the following more general conjecture for any proper multiplicative subgroup. For simplicity, we denote to be the set of -th powers in , equivalently, the subgroup of with order .
Conjecture 1.9.
Let . If is a sufficiently large prime power, then for any , does not admit a non-trivial multiplicative decomposition, that is, there do not exist two sets with , such that .
1.9 predicts that a shifted multiplicative subgroup of a finite field admits a non-trivial multiplicative decomposition only when it has a small size. We remark that the integer version of 1.9, namely, for each , a non-trivial shift of -th powers in integers has no non-trivial multiplicative decomposition, has been proved and strengthened in a series of papers by Hajdu and Sárközy [18, 19, 20]. On the other hand, to the best knowledge of the authors, 1.9 appears to be much harder and no partial progress has been made. For the analog of 1.9 on the additive decomposition of multiplicative subgroups, we refer to recent papers [21, 31, 33, 38] for an extensive discussion on partial progress.
Our main contribution to 1.9 is the following two results. The first one is a corollary of Theorem 1.1.
Corollary 1.10.
Let and be a prime such that . Let . If can be multiplicative decomposed as the product of two sets with , then we must have , that is, all products are distinct. In other words, are multiplicatively co-Sidon.
In particular, Corollary 1.10 confirms 1.9 under the assumption that is a prime, , and is a prime. The second result provides a partial answer to 1.9 asymptotically.
Theorem 1.11.
Let be fixed and be a positive integer. As , the number of primes such that and has no non-trivial multiplicative decomposition is at least
In particular, by setting and , our result has the following significant implication to Sárközy’s conjecture [31]: for almost all odd primes , the shifted multiplicative subgroup has no non-trivial multiplicative decomposition. In other words, if , then Sárközy’s conjecture holds for almost all primes ; see Theorem 6.1 for a precise statement.
Some partial progress has been made for 1.9 when the multiplicative decomposition is assumed to have special forms [31, 33]. We also make progress in this direction in Subsection 6.2. In particular, in Theorem 6.6, we confirm the ternary version of 1.9 in a strong sense, which generalizes [31, Theorem 2].
Notations. We follow standard notations from analytic number theory. We use and to denote the standard prime-counting functions. We adopt standard asymptotic notations . We also follow the Vinogradov notation : we write if there is an absolute constant so that .
Throughout the paper, let be a prime and a power of . Let be the finite field with elements and let . We always assume that with , and denote to be the subgroup of with order . If is assumed to be fixed, for brevity, we simply write instead of .
We also need some notations for arithmetic operations among sets. Given two sets and , we write the product set , and the sumset . Given the definition of Diophantine tuples, it is also useful to define the restricted product set of , that is, .
Structure of the paper. In Section 2, we introduce more background. In Section 3, using Gauss sums and Weil’s bound, we give an upper bound on the size of the set which satisfies various multiplicative properties based on character sum estimates. In particular, we prove Proposition 1.4. In Section 4, we first prove Theorem 1.1 using Stepanov’s method. At the end of the section, we deduce applications of Theorem 1.1 to Diophantine tuples and prove Theorem 1.5 and Theorem 1.6. Via Gallagher’s larger sieve inequality and other tools from analytic number theory, in Section 5, we prove Theorem 1.2 and Theorem 1.3. In Section 6, we study multiplicative decompositions and prove Theorem 1.11.
2. Background
2.1. Stepanov’s method
We first describe Stepanov’s method [35]. If we can construct a low degree non-zero auxiliary polynomial that vanishes on each element of a set with high multiplicity, then we can give an upper bound on based on the degree of the polynomial. It turns out that the most challenging part of our proofs is to show that the auxiliary polynomial constructed is not identically zero.
To check that each root has a high multiplicity, standard derivatives might not work since we are working in a field with characteristic . To resolve this issue, we need the following notation of derivatives, known as the Hasse derivatives or hyper-derivatives; see [26, Section 6.4].
Definition 2.1.
Let . If is a non-negative integer, then the -th hyper-derivative of is
where we follow the standard convention that for , so that the -th hyper-derivative is a polynomial.
Following the definition, we have . We also need the next three lemmas.
Lemma 2.2 ([26, Lemma 6.47]).
If , then
Lemma 2.3 ([26, Corollary 6.48]).
Let be positive integers. If and , then we have
(2.1) |
Lemma 2.4 ([26, Lemma 6.51]).
Let be a non-zero polynomial in . If is a root of for , then is a root of multiplicity at least .
2.2. Gallagher’s larger sieve inequality
In this subsection, we introduce Gallagher’s larger sieve inequality and provide necessary estimations from it. Gallagher’s larger sieve inequality will be one of the main ingredients for the proof of Theorem 1.2. In 1971, Gallagher [15] discovered the following sieve inequality.
Theorem 2.5.
(Gallagher’s larger sieve inequality) Let be a natural number and . Let be a set of primes. For each prime , let . For any , we have
(2.2) |
provided that the denominator is positive.
As a preparation to apply Gallagher’s larger sieve in our proof, we need to establish a few estimates related to primes in arithmetic progressions. For , we follow the standard notation
For our purposes, could be as large as (for example, see Theorem 1.2), and we need the Siegel-Walfisz theorem to estimate .
Lemma 2.6 ([28, Corollary 11.21]).
Let be a constant. There is a constant such that
holds uniformly for and .
A standard application of partial summation with Lemma 2.6 leads to the following corollary.
Corollary 2.7.
There is a constant , such that
holds uniformly for and .
We also need the following lemma.
Lemma 2.8 ([1, page 72]).
Let be a positive integer. Then
2.3. An effective estimate for
Following [7], for each real number , we write
It is shown in [7] that as . For our application, we show a stronger result that and we will make this estimate explicit and effective. We follow the proof in [7] closely and prove the following proposition, which will be used later in the proof of Theorem 1.2.
Proposition 2.9.
If and , then , where the implicit constant is absolute.
Let . Let and be a generalized -tuple with property and . Consider the system of equations
(2.3) |
Clearly, for each , is a solution to this system, and we denote so that and . Let .
The following lemma is a generalization of [7, Lemma 3.1], showing that provides a “good” rational approximation to if is large. Note that in [7, Lemma 3.1], it was further assumed that is odd and . Nevertheless, an almost identical proof works, and we skip the proof.
Lemma 2.10.
Let
Assume that . Then for each , we have
(2.4) |
Corollary 2.11.
Assume that . Then for each and
(2.5) |
Proof.
Let . Applying the gap principle from [7, Lemma 2.4] to , we have
It follows that
In particular . Thus, if , then and inequality (2.5) follows from Lemma 2.10. ∎
Now we are ready to prove Proposition 2.9. Recall we have the inequality for , and the inequality for all positive integers . It follows that
when is odd, and
when is even. Thus, when , we always have
Therefore, when , we can apply Lemma 2.10. Note that the absolute height of is . Since , for , Lemma 2.10 implies that
moreover, . Therefore, we can apply the quantitative Roth’s theorem due to Evertse [12] (see also [7, Theorem 2.2]) to conclude that
where the implicit constant is absolute.
2.4. Implications of the Paley graph conjecture
The Paley graph conjecture on double character sums implies many results of the present paper related to the estimation of character sums. We record the statement of the conjecture (see for example [7, 16]).
Conjecture 2.12 (Paley graph conjecture).
Let be a real number. Then there is and such that for any prime , any with , and any non-trivial multiplicative character of , the following inequality holds:
The connection between the Paley graph conjecture and the problem of bounding the size of Diophantine tuples was first observed by Güloğlu and Murty in [16]. Let be fixed, , where is a prime. The Paley graph conjecture trivially implies and as . Also, the bound on in Theorem 1.2 can be improved to (see [7], [16]) when is fixed and . Furthermore, the Paley graph conjecture also immediately implies Sárközy’s conjecture (1.8) in view of Proposition 3.4. However, the Paley graph conjecture itself remains widely open, and our results are unconditional. We refer to [32] and the references therein for recent progress on the Paley graph conjecture assuming have small doubling.
3. Preliminary estimations for product sets in shifted multiplicative subgroups
3.1. Character sum estimate and the square root upper bound
The purpose of this subsection is to prove Proposition 1.4 by establishing an upper bound on the double character sum in Proposition 3.1 using basic properties of characters and Gauss sums. For any prime , and any , we follow the standard notation that , where we embed into . We refer the reader to [26, Chapter 5] for more results related to Gauss sums and character sums.
We refer to [1, Section 2] for a historical discussion of Vinogradov’s inequality 1.1. Gyarmati [17, Theorem 7] and Becker and Murty [1, Proposition 2.7] independently showed that the Legendre symbol in inequality 1.1 can be replaced with any non-trivial Dirichlet character modulo . For our purposes, we extend Vinogradov’s inequality 1.1 to all finite fields and all nontrivial multiplicative characters of , with a slightly improved upper bound.
Proposition 3.1.
Let be a non-trivial multiplicative character of and . For any , we have
Before proving Proposition 3.1, we need some preliminary estimates. Let be a multiplicative character of ; then the Gauss sum associated to is defined to be
where is the absolute trace map.
Lemma 3.2 ([26, Theorem 5.12]).
Let be a multiplicative character of . Then for any ,
Now we are ready to prove Proposition 3.1.
Proof of Proposition 3.1.
By Lemma 3.2, we can write
It is well-known that (see for example [26, Theorem 5.11]). Since for each , we can apply the triangle inequality and Cauchy-Schwarz inequality to obtain
By orthogonality relations, we have
Thus, we have
By switching the roles of and , we obtain the required character sum estimate. ∎
Let such that and denote with order . Then we prove Proposition 1.4.
Proof of Proposition 1.4.
Let be a multiplicative character of order .
Let with property , that is, . Note that for each , unless . Note that given , there is at most one such that . Therefore, by Proposition 3.1, we have
It follows that
Next we work under the weaker assumption . In this case, note that for each such that , unless . Proposition 3.1 then implies that
and it follows that . ∎
3.2. Estimates on and if
Let and . In this subsection, we provide several useful estimates on and when , which will be used in Section 6.
We need to use the following lemma, due to Karatsuba [23].
Lemma 3.3.
Let and . Then for any non-trivial multiplicative character of and any positive integer , we have
The following proposition improves and generalizes [31, Theorem 1]. It also improves [33, Lemma 17] (see Remark 3.7).
Proposition 3.4.
Let . Let such that and . If for some with , then
Proof.
Let and such that with . Without loss of generality, we assume that . We first establish a weaker lower bound that .
When , Sárközy [31] has shown that . While he only proved this estimate when is a prime [31, Theorem 1], it is clear that the same proof extends to all finite fields .
Next, assume that . Let and . Since , we have . Let be a multiplicative character of with order . Then it follows that for each , we have for each . Therefore, by a well-known consequence of Weil’s bound (see for example [26, Exercise 5.66]),
On the other hand, since , we have
Combining the above two inequalities, we obtain that
Since , it follows that and thus
Remark 3.5.
We remark that the same method could be used to refine a similar result for the additive decomposition of multiplicative subgroups, which improves a result of Shparlinski [34, Theorem 6.1] (moreover, our proof appears to be much simpler than his proof). More precisely, we can prove the following:
Let . Let such that . If for some with , then
Note that Proposition 3.4 only applies to multiplicative subgroups with . When is a prime, and is a non-trivial multiplicative subgroup of (in particular, is allowed), we have the following estimate, due to Shkredov [33].
Lemma 3.6.
If is a proper multiplicative subgroup of such that for some with and some , then
as .
Proof.
If , let ; otherwise, let . Then we have , or equivalently, . Note that and , thus, the lemma then follows immediately from [33, Lemma 17]. ∎
Remark 3.7.
Let , where with and . When is a constant, and , Proposition 3.4 is better than Lemma 3.6. Indeed, in [33, Lemma 17], Shkredov showed that , while Proposition 3.4 showed the stronger result that , removing the factor. This stronger bound would be crucial in proving results related to multiplicative decompositions (for example, Theorem 6.1). Instead, Lemma 3.6 will be useful for applications in ternary decompositions (Theorem 6.6).
Remark 3.8.
Lemma 3.6 fails to extend to , where is a proper prime power. Let be a square, , and . Let be a subset of with size . Let . Then we have while .
Remark 3.9.
Let be fixed. Let be a prime power and let , we define be the total number of pairs of sets with such that . Note that 1.9 implies when is sufficiently large, but it seems out of reach in general. Instead, one can find a non-trivial upper bound of . Using Proposition 3.4 and following the same strategy of the proof in [3, Theorem 1], we have a non-trivial upper bound of as follows:
We also note that by using Corollary 1.10, we confirmed if is a prime, , and is a prime. In addition, Theorem 1.11 gives the same result for asymptotically for a subset of primes with lower density at least .
4. Products and restricted products in shifted multiplicative subgroups
In this section, we use Stepanov’s method to study product sets and restricted product sets that are contained in shifted multiplicative subgroups. Our proofs are inspired by Stepanov’s original paper [35], and the recent breakthrough of Hanson and Pertidis [21].
Throughout the section, we assume and is a prime power. Recall that .
4.1. Product set in a shifted multiplicative subgroup
In this subsection, we prove Theorem 1.1, which can be viewed as the bipartite version of Diophantine tuples over finite fields. As a corollary of Theorem 1.1, we prove Corollary 1.10. Besides it, Theorem 1.1 will be also repeatedly used to prove several of our main results in the present paper.
Proof of Theorem 1.1.
Let . Let and such that . Since , we have
for each and . This simple observation will be used repeatedly in the following computation.
Let be the unique solution of the following system of equations:
(4.1) |
This is justified by the invertibility of the coefficient matrix of the system (a Vandermonde matrix). We claim that . Suppose otherwise that , then for all , violating the assumption in equation (4.1). Indeed, the generalized Vandermonde matrix is non-singular since it has determinant
Consider the following auxiliary polynomial
(4.2) |
Note that . Thus, if is a prime, then and thus the condition is automatically satisfied. Then is a non-zero polynomial since the coefficient of in is
by the assumption on the binomial coefficient. Also, it is clear that the degree of is at most .
Next, we compute the derivatives of on . For each , system 4.1 implies that
For each , by the assumption, , that is, for each . Thus, for each , we additionally have
Therefore, Lemma 2.4 allows us to conclude that each of is a root of with multiplicity at least , and each of is a root of with multiplicity at least . It follows that
Finally, assuming that . In this case,
And the coefficient of of is for each by the assumptions on ’s. It follows that is also a root of with multiplicity . Since , we have the stronger estimate that
∎
Remark 4.1.
More generally, one can study the same question if is instead contained in a coset of . However, note that this more general case can be always reduced to the special case studied in Theorem 1.1. Indeed, if with , then , where .
Next, we prove Corollary 1.10, an important corollary of Theorem 1.1. It would be crucial for proving results in Section 6.
Proof of Corollary 1.10.
Since , we have , thus Theorem 1.1 implies that . On the other hand, since , it follows that . Therefore, we have . ∎
4.2. Restricted product set in a shifted multiplicative subgroup
Recall that has property if and only if for each such that . In other words, . Thus, in this subsection, we are led to study restricted product sets and we establish the following restricted product analog of Theorem 1.1. In the next subsection, Theorem 1.1 and Theorem 4.2 will be applied together to prove Theorem 1.5 in the case that is a prime and Theorem 1.6 in the case that is a square.
Theorem 4.2.
Let and let be a prime power. Let and . If while , then .
The proof of Theorem 4.2 is similar to Theorem 1.1, but it is more delicate. In particular, the choice of the auxiliary polynomial 4.4 needs to be modified from that of the proof 4.2 of Theorem 1.1. In view of Theorem 1.1, we can further assume that , for otherwise we already have a good bound on ; we refer to Subsection 4.3 for details. It turns out that this additional assumption (which we get for free) is crucial in our proof since it guarantees that the auxiliary polynomial we constructed is not identically zero.
Proof of Theorem 4.2.
Since , there is such that . Let . If , then we are done. Otherwise, if is even, let ; if is odd, let , where is an arbitrary element such that . Then we have and is even. Note that , thus it suffices to show .
Let , where is even. Write . Without loss of generality, we may assume that . Let . Let be the unique solution of the following system of equations:
(4.3) |
Indeed, that coefficient matrix of the system is the generalized Vandermonde matrix which is non-singular since it has nonzero determinant Note that ; for otherwise and we must have in view of the first equations in system 4.3, which contradicts the last equation in system 4.3.
Consider the following auxiliary polynomial
(4.4) |
It is clear that the degree of is at most . Since , we have
for each . This simple observation will be used repeatedly in the following computation.
First, we claim that for each , , and , we have
(4.5) |
Indeed, by Lemma 2.3, we have
Note that in the exponent of the last summand, we always have and so that , and thus
by the assumptions in system 4.3. This proves the claim.
Similarly, using Lemma 2.2, Lemma 2.3, system 4.3, and equation 4.5, we can compute
Since for each , we have for , and thus
for all . Since , we have
Putting these altogether into the computation of , we have
where we used the fact . In particular, is not identically zero.
In conclusion, is a non-zero polynomial with degree at most , and Lemma 2.4 implies that each of is a root of with multiplicity at least . Recall that . It follows that
that is, we have Therefore, . This finishes the proof. ∎
4.3. Applications to generalized Diophantine tuples over finite fields
In this subsection, we illustrate how to apply Theorem 1.1 and Theorem 4.2 for obtaining improved upper bounds on the size of a generalized Diophantine tuple or a strong generalized Diophantine tuple over , when is a prime and is a square.
Proof of Theorem 1.5.
(1) Let with property , that is, . Theorem 1.1 implies that
It follows that . If , we have a stronger upper bound:
It follows that .
(2) Let with property , that is, . If , then (1) implies that and we are done. If , then Theorem 4.2 implies the required upper bound. ∎
Remark 4.3.
Theorem 1.1 can be used to deduce a weaker upper bound of the form for Theorem 1.5 (2). Let such that . We can write such that and differ by at most . Note that since and are disjoint, we have and thus Theorem 1.1 implies that , which further implies that . Note that such a weaker upper bound is worse than the trivial upper bound from character sums (Proposition 1.4) when , and this is one of our main motivations for establishing the bound in Theorem 1.5 (2).
Next, we consider the case is a square. First we establish a non-trivial upper bound on and under some minor assumption. While these new bounds only improve the trivial upper bound from character sums (Proposition 1.4) slightly, we will see these new bounds are sometimes sharp in the proof of Theorem 1.6. To achieve our goal, we need the following special case of Kummer’s theorem [25].
Lemma 4.4.
Let be a prime and be positive integers. If there is no carry between the addition of and in base-, then is not divisible by .
Theorem 4.5.
Let be a prime power and a square, and let .
-
(1)
Let be a divisor of . Let be the remainder of divided by . If , then .
-
(2)
Let and let be a divisor of . Let be the remainder of divided by . If , then .
Proof.
(1) Since , there is no carry between the addition of and in base-. Thus, there is no carry between the addition of and in base-. It follows from Lemma 4.4 that
Let with property such that . Note that Proposition 1.4 implies that . For the sake of contradiction, assume that . Note that and
it follows from Theorem 1.1 that
that is, , a contradiction. This completes the proof.
(2) Let with property such that . Then . If , we just apply (1). Next assume that , then Theorem 4.2 implies that
provided that . When , we have used SageMath to verify the theorem. ∎
Now we are ready to prove Theorem 1.6, which determines the maximum size of an infinitely family of generalized Diophantine tuples and strong generalized Diophantine tuples over finite fields.
Proof of Theorem 1.6.
In both cases, the upper bound follows from Theorem 4.5. To show that is a lower bound, we observe that has property (and therefore ). Indeed, since and (from the assumption ). ∎
Remark 4.6.
Our SageMath code indicates that the last statement of Theorem 1.6 does not hold when and , when and , and when and . We conjecture the same statement holds for , provided that is sufficiently large.
So far we have only considered special cases of applying Theorem 1.1. In general, to apply Theorem 1.1, the assumption on the binomial coefficient in the statement of Theorem 1.1 might be tricky to analyze. However, if the base- representation of behaves “nicely” (for example, if the order of modulo is small, then the base- representation is periodic with a small period), then it is still convenient to apply Theorem 1.1. As a further illustration, we prove the following theorem. Note that the new bound is of the same shape as that in Theorem 1.5 (2), so it can be viewed as a generalization of Theorem 1.5 (2) as changing a prime to an arbitrary power of , provided that .
Theorem 4.7.
Let , and let be a power of such that . Then for any .
Proof.
Let with property , that is, . If , we are done by Theorem 4.2. Thus, we may assume that . It suffices to show . To achieve that, we try to find an arbitrary subset of such that and is as large as possible. With such a subset , we have so that we can apply Theorem 1.1. In the rest of the proof, we aim to find such an with so that, from Theorem 1.1, we can deduce
Write in base-, that is, with for each and . Next, we construct according to the size of .
Case 1. . In this case, let be an arbitrary subset of with , that is, . It is easy to verify that using Lemma 4.4. Since , it also follows readily that .
Case 2. . In this case, let be an arbitrary subset of with
that is, . Again, it is easy to verify that using Lemma 4.4. Since , it follows that . ∎
Remark 4.8.
Under the same assumption, the proof of Theorem 4.7 can be refined to obtain improved upper bounds on . In particular, if are fixed, and is a prime, then as , we can show that uniformly among . Indeed, if with property with and , then we can assume without loss of generality that . Otherwise, we are done. Note that by Proposition 1.4. Following the notations used in the proof of Theorem 4.7, we have and thus we are always in Case 1, and the same construction of gives as . Thus, Theorem 1.1 gives .
5. Improved upper bounds on the largest size of generalized Diophantine tuples over integers
5.1. Proof of Theorem 1.2
In this subsection, we improve the upper bounds on the largest size of generalized Diophantine tuples with property . We first recall that for each and ,
For , we defined the constant in the introduction
(5.1) |
where the minimum is taken over all nonempty subset of
and
(5.2) |
Here is the proof of our main theorem, Theorem 1.2.
Proof of Theorem 1.2.
Let be a generalized Diophantine -tuple with property and . Given the assumption that , Proposition 2.9 implies that the contribution of with is is negligible. Thus, we can assume that . Let be a nonempty subset of , such that the ratio in equation 5.1 is minimized by . In other words, we have , where
To apply the Gallagher sieve inequality (Theorem 2.5), we set and define the set of primes
For each prime , denote by the image of and let .
Let . We can naturally view as a subset of . Since has property , it follows that . Note that is the multiplicative subgroup of with order . Since and , Theorem 1.5 (2) implies that
Set . Applying Gallagher’s larger sieve, we obtain that
(5.3) |
Let be the constant from Corollary 2.7. For the numerator on the right-hand side of inequality 5.3, we have
Next, we estimate the denominator on the right-hand side of inequality 5.3. Note that . Then we have , and so . Since , we deduce . Thus we have . This, together with Corollary 2.7 and Lemma 2.8, deduces that for each ,
Thus we have
Finally, recall that . It follows that
Recall that . Thus, to obtain our desired result, we need to show
and it suffices to show that
as , or equivalently,
as . We notice that . Let . Then the assumption implies
and
as required. ∎
Remark 5.1.
Note that when , that is to say, when we only consider primes such that for applying the Gallagher inequality, the condition guarantees that the -th powers are indeed -th powers modulo . We have , thus we trivially have in view of equation 5.1. In particular, if is fixed and , Theorem 1.2 implies that
(5.4) |
which already provides a substantial improvement on the best-known upper bound whenever given in [7]. Moreover, note that can be as small as when is the product of distinct primes [28, Theorem 2.9]. Thus, in view of Theorem 1.2, the inequality 5.4 already shows there is such that and
(5.5) |
Note that 5.5 already breaks the barrier. On the other hand, we can still use other primes such that for which -th powers are in fact -th powers modulo when we apply the Gallagher sieve inequality. We can take advantage of the improvement on the upper bound of . In the next two subsections, we further provide a significant improvement on inequality 5.5.
Next, we define a strong Diophantine -tuple with property to be a set of distinct positive integers such that is a -th power for any choice of and . We have a stronger upper bound for the size of a strong Diophantine tuple with property . We define
Theorem 5.2.
There is a constant , such that as ,
holds uniformly for positive integers such that . Moreover, if is even, under the same assumption (including the case ), we have the stronger bound
where is the number of divisors of .
Proof.
The proof is very similar to the proof of Theorem 1.2 and we follow all the notations and steps as in the proof of Theorem 1.2, apart from the minor modifications stated below.
We prove the first part. For each , we have the stronger upper bound that by Theorem 1.5 (2). To optimize the upper bound obtained from Gallagher’s larger sieve, we instead set .
Next, we assume that is even and prove the second part. Notice that for each , there is a positive integer , such that . Thus, is bounded by the number of positive integral solutions to the equation , which is at most . On the other hand, this also implies that all the elements in are at most . Thus, we can set instead and obtain the stronger upper bound. ∎
5.2. Proof of Theorem 1.3
In this subsection, by finding a more refined upper bound on in equation 5.1, we show that the same approach significantly improves the upper bound of in inequality 5.5 when is the product of the first few distinct primes.
We label all the primes in increasing order so that . Let be the product of first primes. Let . For , we define inductively:
(5.6) |
We note that for any . Also, it is clear that
(5.7) |
Lemma 5.3.
Following the above definitions, we have
(5.8) |
Proof.
Furthermore, we introduce the following notation which is similar to the previously introduced on equation 5.2. For each , we let
Note that . We also establish a recurrence relation on the sequence.
Lemma 5.4.
The sequence satisfies the recurrence relation
(5.9) |
Proof.
We have
It is easy to show that the inner sum consists of many and a single . It follows that
proving the lemma. ∎
We are now ready to prove Theorem 1.3.
Proof of Theorem 1.3.
For each , we choose , where is the largest integer such that . It follows that . Thus, using equations 5.7 and 5.9, we have
Note that for each prime , it is easy to verify that Recall that the inequality holds for all real , and a standard application of partial summation gives
Also, the prime number theorem implies that
and thus . Putting the above estimates altogether, we have
for some absolute constant . It follows from Theorem 1.2 that
∎
5.3. An upper bound on
In this subsection, we deduce a simple upper bound of . It turns out that this upper bound well approximates empirically.
Theorem 5.5.
For any , we have
where with
and
Proof.
We denote , where are distinct primes factors of such that
Define
Then is obviously a subset of the set consisting of residue classes that can be used in Gallagher’s larger sieve in the proof of Theorem 1.2. (In particular, when , consists of all the available residue classes with .) In view of the definition of , it suffices to show that
We first compute the size of . Equivalently, we can write
and hence, we deduce . In order to compute , we first count the number of solutions to over such that for separately, and then use the Chinese remainder theorem. Set
Note that
and . It follows that
For each , we calculate
Similarly, we have
Putting these all together, we compute
Hence,
∎
Therefore, Theorem 1.2 implies
Corollary 5.6.
There is a constant , such that as ,
holds uniformly for positive integers such that .
Remark 5.7.
Our computations indicate that when , the inequality holds for all but of them. This numerical evidence suggests that provides a good approximation for for a generic . Note that the computational complexity for computing is the same as that of the prime factorization of : a naive algorithm takes time. The best theoretical algorithm has running time using the general number field sieve [5]. On the other hand, computing requires time; we refer to Appendix A for an algorithm and some computational results.
6. Multiplicative decompositions of shifted multiplicative subgroups
In this section, we present our contributions to 1.9. In particular, we make significant progress towards Sárközy’s conjecture (1.8). We recall .
6.1. Applications to Sárközy’s conjecture
In this subsection, we show that for almost all primes , the set cannot be decomposed as the product of two sets non-trivially. This confirms the truth of Sárközy’s conjecture (1.8) as well as the truth of its generalization in the generic case (1.9) when the shift of the subgroup is given by .
Theorem 6.1.
Let be fixed. As , the number of primes such that and can be decomposed as the product of two sets non-trivially (that is, it can be written as the product of two subsets of with size at least 2) is .
Proof.
Let be the set of primes such that and admits a non-trivial multiplicative decomposition. By the prime number theorem for arithmetic progressions, it suffices to show that .
Let . Then we can write as the product of two sets such that . Then Corollary 1.10 implies that that is,
(6.1) |
On the other hand, Proposition 3.4 implies that we can find an absolute constant such that
It follows that has a divisor in the interval . To summarize, if , then we have where denotes the number of divisors of in the interval . Now, we use results by Ford [14, Theorem 6] on the distribution of shift primes with a divisor in a given interval. Denote
(6.2) | ||||
(6.3) |
Setting and , [14, Theorem 6] and [14, Theorem 1, third case of (v)] imply that
where and . It follows that as , we have . Therefore, we have
We conclude that as ,
∎
Using a similar argument, we can prove Theorem 1.11:
Proof of Theorem 1.11.
Consider the family of primes such that and is a -th power modulo . By a standard application of the Chebotarev density theorem, the density of such primes is given by . Among the family of such primes , we can repeat the same argument as in the proof of Theorem 6.1 to show that if admits a non-trivial multiplicative decomposition, then necessarily has a divisor which is “close to” . We remark that it is important to assume that is a -th power modulo , so that we can take advantage of Corollary 1.10. Similar to the proof of Theorem 6.1, we can show that among the family of primes , the property that has a divisor with the desired magnitude fails to hold for almost all . This finishes the proof of the theorem. ∎
Remark 6.2.
When is a fixed negative integer, one can obtain a similar result to Theorem 1.11 following the idea of the above proof.
Remark 6.3.
Theorem 1.11 essentially states if is fixed, is a prime, and , then it is very unlikely that we can decompose as the product of two subsets of non-trivially. On the other hand, when , the above technique does not apply. Nevertheless, when and we have two sets such that , Theorem 1.1 implies that
In particular, we get the following non-trivial fact: if is fixed, then is also uniquely fixed.
6.2. Applications to special multiplicative decompositions
In this subsection, we verify the ternary version of 1.9 in a strong sense, which generalizes [31, Theorem 2].
Shkredov [33, Theorem 3] showed if is a multiplicative subgroup of with , then there is no and such that . In fact, due to the analytic nature of the proof, he pointed out that his proof can be slightly modified to show something stronger, namely , as long as is small(see also [33, Remark 15]). The following corollary of Theorem 1.1 is of a similar flavor.
Corollary 6.4.
Let be a prime. If is a proper multiplicative subgroup of with , and , then there is no such that .
Proof.
We assume, otherwise, that for some . Then we observe that for each , it follows that
Since , it follows that . Let , and . Then we have and thus Theorem 1.1 implies that . Comparing the above two inequalities, we obtain that
which implies that , contradicting the assumption that . ∎
Lemma 6.5.
Let be nonempty subsets of and let . Then .
Proof.
It suffices to show , which is a special case of [29, Theorem 5.1] due to Ruzsa. ∎
The following two theorems generalize Sárközy [31, Theorem 2] and confirm the ternary version of 1.9 in a strong form.
Theorem 6.6.
There exists an absolute constant , such that whenever is a prime, is a proper multiplicative subgroup of with , and , there is no ternary multiplicative decomposition with and .
Proof.
Assume that there are sets with , such that for some proper multiplicative subgroup of and some .
Then we can write in three different ways: , so that we can apply the results in previous sections to each of them. Note that Lemma 3.6 implies that
On the other hand, Theorem 1.1 implies that
Therefore, from Lemma 6.5 and the fact , we have
It follows that
that is, , where the implicit constant is absolute. This completes the proof of the theorem. ∎
Theorem 6.7.
Let . There is a constant , such that for each prime power and a divisor of with , there is no ternary multiplicative decomposition with , , and .
Proof.
The proof is similar to the proof of Theorem 6.6. While Lemma 3.6 does not hold in the new setting (see Remark 3.8), we can instead use Proposition 3.4. If , then Proposition 3.4 implies that
A similar computation leads to , which implies that since we assume that . ∎
Acknowledgements
The authors thank Andrej Dujella, Greg Martin, and József Solymosi for helpful discussions. The third author was supported by the KIAS Individual Grant (CG082701) at the Korea Institute for Advanced Study and the Institute for Basic Science (IBS-R029-C1).
References
- [1] R. Becker and M. R. Murty. Diophantine -tuples with the property . Glas. Mat. Ser. III, 54(74)(1):65–75, 2019.
- [2] A. Bérczes, A. Dujella, L. Hajdu, and F. Luca. On the size of sets whose elements have perfect power -shifted products. Publ. Math. Debrecen, 79(3-4):325–339, 2011.
- [3] S. R. Blackburn, S. V. Konyagin, and I. E. Shparlinski. Counting additive decompositions of quadratic residues in finite fields. Funct. Approx. Comment. Math., 52(2):223–227, 2015.
- [4] Y. Bugeaud and A. Dujella. On a problem of Diophantus for higher powers. Math. Proc. Cambridge Philos. Soc., 135(1):1–10, 2003.
- [5] J. P. Buhler, H. W. Lenstra, Jr., and C. Pomerance. Factoring integers with the number field sieve. In The development of the number field sieve, volume 1554 of Lecture Notes in Math., pages 50–94. Springer, Berlin, 1993.
- [6] L. Caporaso, J. Harris, and B. Mazur. Uniformity of rational points. J. Amer. Math. Soc., 10(1):1–35, 1997.
- [7] A. B. Dixit, S. Kim, and M. R. Murty. Generalized Diophantine -tuples. Proc. Amer. Math. Soc., 150(4):1455–1465, 2022.
- [8] A. Dujella. On the size of Diophantine -tuples. Math. Proc. Cambridge Philos. Soc., 132(1):23–33, 2002.
- [9] A. Dujella. There are only finitely many Diophantine quintuples. J. Reine Angew. Math., 566:183–214, 2004.
- [10] A. Dujella. Diophantine -tuples and Elliptic Curves, volume 79 of Developments in Mathematics. Springer, Cham, 2024.
- [11] A. Dujella and V. Petričević. Strong Diophantine triples. Experiment. Math., 17(1):83–89, 2008.
- [12] J.-H. Evertse. On the quantitative subspace theorem. Zap. Nauchn. Sem. S.-Peterburg. Otdel. Mat. Inst. Steklov. (POMI), 377(Issledovaniya po Teorii Chisel. 10):217–240, 245, 2010.
- [13] G. Faltings. Endlichkeitssätze für abelsche Varietäten über Zahlkörpern. Invent. Math., 73(3):349–366, 1983.
- [14] K. Ford. The distribution of integers with a divisor in a given interval. Ann. of Math. (2), 168(2):367–433, 2008.
- [15] P. X. Gallagher. A larger sieve. Acta Arith., 18:77–81, 1971.
- [16] A. M. Güloğlu and M. R. Murty. The Paley graph conjecture and Diophantine -tuples. J. Combin. Theory Ser. A, 170:105155, 9, 2020.
- [17] K. Gyarmati. On a problem of Diophantus. Acta Arith., 97(1):53–65, 2001.
- [18] L. Hajdu and A. Sárközy. On multiplicative decompositions of polynomial sequences, I. Acta Arith., 184(2):139–150, 2018.
- [19] L. Hajdu and A. Sárközy. On multiplicative decompositions of polynomial sequences, II. Acta Arith., 186(2):191–200, 2018.
- [20] L. Hajdu and A. Sárközy. On multiplicative decompositions of polynomial sequences, III. Acta Arith., 193(2):193–216, 2020.
- [21] B. Hanson and G. Petridis. Refined estimates concerning sumsets contained in the roots of unity. Proc. Lond. Math. Soc. (3), 122(3):353–358, 2021.
- [22] B. He, A. Togbé, and V. Ziegler. There is no Diophantine quintuple. Trans. Amer. Math. Soc., 371(9):6665–6709, 2019.
- [23] A. A. Karatsuba. Distribution of values of Dirichlet characters on additive sequences. Dokl. Akad. Nauk SSSR, 319(3):543–545, 1991.
- [24] S. Kim, C. H. Yip, and S. Yoo. Explicit constructions of Diophantine tuples over finite fields. Ramanujan J., 65(1):163–172, 2024.
- [25] E. Kummer. Über die ergänzungssätze zu den allgemeinen reciprocitätsgesetzen. Journal für die reine und angewandte Mathematik, 44:93–146, 1852.
- [26] R. Lidl and H. Niederreiter. Finite fields, volume 20 of Encyclopedia of Mathematics and its Applications. Cambridge University Press, Cambridge, second edition, 1997.
- [27] S. Macourt, I. D. Shkredov, and I. E. Shparlinski. Multiplicative energy of shifted subgroups and bounds on exponential sums with trinomials in finite fields. Canad. J. Math., 70(6):1319–1338, 2018.
- [28] H. L. Montgomery and R. C. Vaughan. Multiplicative number theory. I. Classical theory, volume 97 of Cambridge Studies in Advanced Mathematics. Cambridge University Press, Cambridge, 2007.
- [29] I. Z. Ruzsa. Cardinality questions about sumsets. In Additive combinatorics, volume 43 of CRM Proc. Lecture Notes, pages 195–205. Amer. Math. Soc., Providence, RI, 2007.
- [30] A. Sárközy. On additive decompositions of the set of quadratic residues modulo . Acta Arith., 155(1):41–51, 2012.
- [31] A. Sárközy. On multiplicative decompositions of the set of the shifted quadratic residues modulo . In Number theory, analysis, and combinatorics, De Gruyter Proc. Math., pages 295–307. De Gruyter, Berlin, 2014.
- [32] T. Schoen and I. D. Shkredov. Character sums estimates and an application to a problem of Balog. Indiana Univ. Math. J., 71(3):953–964, 2022.
- [33] I. D. Shkredov. Any small multiplicative subgroup is not a sumset. Finite Fields Appl., 63:101645, 15, 2020.
- [34] I. E. Shparlinski. Additive decompositions of subgroups of finite fields. SIAM J. Discrete Math., 27(4):1870–1879, 2013.
- [35] S. A. Stepanov. On the number of points of a hyperelliptic curve over a finite prime field. Izv. Akad. Nauk SSSR, Ser. Mat., 33:1171–1181, 1969.
- [36] I. M. Vinogradov. Elements of number theory. Dover Publications, Inc., New York, 1954. Translated by S. Kravetz.
- [37] I. V. Vyugin and I. D. Shkredov. On additive shifts of multiplicative subgroups. Mat. Sb., 203(6):81–100, 2012.
- [38] C. H. Yip. Additive decompositions of large multiplicative subgroups in finite fields. Acta Arith., 213(2):97–116, 2024.
- [39] C. H. Yip. Improved upper bounds on Diophantine tuples with the property . 2024. arXiv:2406.00840. Bull. Aust. Math. Soc., to appear.
Appendix A Algorithm and Computations
We continue our discussion from the introduction on the following constant
It is implicit in [7] that . We also write
Our main result, Theorem 1.2, implies that . In particular, in view of Remark 5.1, it follows that for all and when . In Figure A.1, we pictorially compare our new bound with the bound when .

Recall that for , we defined the constant where the minimum is taken over all nonempty subsets of and
To compute , we use the following simple greedy algorithm with running time . The observation is as follows. If is fixed, our goal is to minimize . Thus, we should choose those residue classes with as large as possible to maximize . Then, we can sort these gcds in decreasing order, and when is fixed, we pick those residue classes corresponding to the largest gcds. The following is a precise description of the algorithm:
Algorithm A.1.
Let . We follow the notations defined in Subsection 5.2.
-
Step 1.
Let . We list the elements of by such that is decreasing by using a sorting algorithm.
-
Step 2.
Set , , and .
-
Step 3.
Return and terminate the algorithm.
Note that the running time of the above algorithm is : sorting takes time, while other steps take linear time.
Next, we also consider the minimum value of the upper bounds for each . Table A.1 shows the values of for when they are changed.
2 | 2.6071 |
---|---|
4 | 1.37258 |
6 | 0.80385 |
8 | 0.72776 |
12 | 0.44134 |
24 | 0.31910 |
36 | 0.29027 |
48 | 0.25836 |
60 | 0.21636 |
120 | 0.16570 |
180 | 0.15191 |
---|---|
240 | 0.13876 |
360 | 0.11708 |
720 | 0.09693 |
840 | 0.09266 |
1260 | 0.08465 |
1440 | 0.08445 |
1680 | 0.07624 |
2520 | 0.06465 |
5040 | 0.05317 |
7560 | 0.05171 |
---|---|
10080 | 0.04592 |
15120 | 0.04252 |
20160 | 0.04111 |
25200 | 0.03887 |
27720 | 0.03665 |
30240 | 0.03647 |
50400 | 0.03343 |
55440 | 0.02997 |
83160 | 0.02877 |
110880 | 0.02574 |
---|---|
166320 | 0.02343 |
221760 | 0.02280 |
277200 | 0.02138 |
332640 | 0.02008 |
498960 | 0.01985 |
554400 | 0.01827 |
665280 | 0.01774 |
720720 | 0.01654 |
1081080 | 0.01587 |
We also report our computations on for in the following table.
2 | 2.6071 |
---|---|
3 | 4.0000 |
4 | 1.3726 |
5 | 2.6667 |
6 | 0.8038 |
7 | 2.4000 |
8 | 0.7278 |
9 | 1.1077 |
10 | 0.7295 |
11 | 2.2222 |
12 | 0.4413 |
13 | 2.1818 |
14 | 0.7185 |
15 | 0.7222 |
16 | 0.5383 |
17 | 2.1333 |
18 | 0.4522 |
19 | 2.1176 |
20 | 0.4450 |
21 | 0.7355 |
22 | 0.7251 |
23 | 2.0952 |
24 | 0.3191 |
25 | 1.1180 |
26 | 0.7313 |
27 | 0.7508 |
28 | 0.4552 |
29 | 2.0741 |
30 | 0.3351 |
31 | 2.0690 |
32 | 0.4555 |
33 | 0.7709 |
34 | 0.7438 |
35 | 0.7311 |
36 | 0.2903 |
37 | 2.0571 |
38 | 0.7497 |
39 | 0.7873 |
40 | 0.3353 |
41 | 2.0513 |
42 | 0.3465 |
---|---|
43 | 2.0488 |
44 | 0.4739 |
45 | 0.4581 |
46 | 0.7604 |
47 | 2.0444 |
48 | 0.2584 |
49 | 1.1716 |
50 | 0.4974 |
51 | 0.8163 |
52 | 0.4818 |
53 | 2.0392 |
54 | 0.3596 |
55 | 0.7678 |
56 | 0.3486 |
57 | 0.8290 |
58 | 0.7742 |
59 | 2.0351 |
60 | 0.2164 |
61 | 2.0339 |
62 | 0.7783 |
63 | 0.4746 |
64 | 0.4124 |
65 | 0.7860 |
66 | 0.3643 |
67 | 2.0308 |
68 | 0.4950 |
69 | 0.8515 |
70 | 0.3548 |
71 | 2.0290 |
72 | 0.2171 |
73 | 2.0282 |
74 | 0.7892 |
75 | 0.5005 |
76 | 0.5006 |
77 | 0.7644 |
78 | 0.3713 |
79 | 2.0260 |
80 | 0.2730 |
81 | 0.6359 |
82 | 0.7956 |
---|---|
83 | 2.0247 |
84 | 0.2263 |
85 | 0.8195 |
86 | 0.7985 |
87 | 0.8796 |
88 | 0.3682 |
89 | 2.0230 |
90 | 0.2239 |
91 | 0.7794 |
92 | 0.5103 |
93 | 0.8877 |
94 | 0.8040 |
95 | 0.8346 |
96 | 0.2261 |
97 | 2.0211 |
98 | 0.5376 |
99 | 0.5038 |
100 | 0.3344 |
101 | 2.0202 |
102 | 0.3827 |
103 | 2.0198 |
104 | 0.3757 |
105 | 0.3626 |
106 | 0.8114 |
107 | 2.0190 |
108 | 0.2355 |
109 | 2.0187 |
110 | 0.3768 |
111 | 0.9094 |
112 | 0.2864 |
113 | 2.0180 |
114 | 0.3874 |
115 | 0.8621 |
116 | 0.5220 |
117 | 0.5160 |
118 | 0.8179 |
119 | 0.8087 |
120 | 0.1657 |
121 | 1.2615 |
122 | 0.8200 |
---|---|
123 | 0.9220 |
124 | 0.5254 |
125 | 0.8820 |
126 | 0.2335 |
127 | 2.0160 |
128 | 0.3877 |
129 | 0.9278 |
130 | 0.3869 |
131 | 2.0155 |
132 | 0.2413 |
133 | 0.8226 |
134 | 0.8256 |
135 | 0.3709 |
136 | 0.3878 |
137 | 2.0148 |
138 | 0.3955 |
139 | 2.0146 |
140 | 0.2400 |
141 | 0.9387 |
142 | 0.8291 |
143 | 0.7812 |
144 | 0.1809 |
145 | 0.8976 |
146 | 0.8307 |
147 | 0.5506 |
148 | 0.5342 |
149 | 2.0136 |
150 | 0.2468 |
151 | 2.0134 |
152 | 0.3928 |
153 | 0.5366 |
154 | 0.3772 |
155 | 0.9081 |
156 | 0.2471 |
157 | 2.0129 |
158 | 0.8353 |
159 | 0.9532 |
160 | 0.2385 |
161 | 0.8485 |
162 | 0.3140 |
---|---|
163 | 2.0124 |
164 | 0.5392 |
165 | 0.3857 |
166 | 0.8382 |
167 | 2.0121 |
168 | 0.1737 |
169 | 1.2965 |
170 | 0.4049 |
171 | 0.5453 |
172 | 0.5416 |
173 | 2.0117 |
174 | 0.4053 |
175 | 0.5104 |
176 | 0.3077 |
177 | 0.9660 |
178 | 0.8422 |
179 | 2.0113 |
180 | 0.1519 |
181 | 2.0112 |
182 | 0.3854 |
183 | 0.9700 |
184 | 0.4014 |
185 | 0.9365 |
186 | 0.4080 |
187 | 0.8001 |
188 | 0.5459 |
189 | 0.3843 |
190 | 0.4129 |
191 | 2.0106 |
192 | 0.2075 |
193 | 2.0105 |
194 | 0.8471 |
195 | 0.3948 |
196 | 0.3652 |
197 | 2.0103 |
198 | 0.2493 |
199 | 2.0102 |
200 | 0.2517 |
201 | 0.9811 |