Diophantine tuples and product sets in shifted powers
Abstract.
Let and . A Diophantine tuple with property is a set of positive integers such that is a -th power for all with . Such generalizations of classical Diophantine tuples have been studied extensively. In this paper, we prove several results related to robust versions of such Diophantine tuples and discuss their applications to product sets contained in a nontrivial shift of the set of all perfect powers or some of its special subsets. In particular, we substantially improve several results by Bérczes–Dujella–Hajdu–Luca, and Yip. We also prove several interesting conditional results. Our proofs are based on a novel combination of ideas from sieve methods, Diophantine approximation, and extremal graph theory.
Key words and phrases:
Diophantine tuples, perfect powers, sieve methods, sums of powers2020 Mathematics Subject Classification:
Primary 11B30, 11D72; Secondary 11N36, 11D41, 05C351. Introduction
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. In particular, many generalizations and variants of Diophantine tuples have been studied. We refer to the recent book of Dujella [16] for a comprehensive overview of the topic. In this paper, we provide improved unconditional and conditional upper bounds on various well-studied generalizations of Diophantine tuples and explore their connections.
Let be a nonzero integer and let , a set of distinct positive integers a Diophantine tuple with property if the product of any two distinct elements in is less than a perfect -th power. Following the standard notations, we also write
These natural generalized notions of Diophantine tuples are of particular interest; see for example [3, 8, 13, 14, 18, 22, 23, 32, 41, 42]. The best-known upper bound on is of the form ; see [32, 41, 42] for the best-known implied constant depending on . Under the Uniformity Conjecture [10] (a consequence of the Bombieri–Lang conjecture), it is well-known that for each , there is a constant such that holds for all nonzero integers [14, 32]; see also Remark 1.13. In Corollary 1.14, we show that assuming both the Uniformity Conjecture and the Lander–Parkin–Selfridge conjecture [34], there is a constant such that holds for all and .
In 2002, Gyarmati, Sárközy, and Stewart [24] initialized the study of two closely related variants of Diophantine tuples by enlarging the set of perfect -th powers (with a fixed ) to the set of perfect -th powers with bounded or the set of all perfect powers. More precisely, set
for each integer ; then the set of perfect powers is
They studied the size of a set of positive integers such that is in (where ) for all that are distinct. More generally, one can study the same question by replacing the shift with any nonzero shift; see for example [3, 4]. For brevity, we extend the familiar notions and in this more general setting. Let be a nonzero integer and let , we say a set of positive integers a Diophantine tuple with property if the product of any two distinct elements in is less than an element in , and we denote
In particular, the case is well-studied. Gyarmati, Sárközy, and Stewart [24] proved that if and is a Diophantine tuple with property , then . They then deduced that if is a Diophantine tuple with property , then . These two results have been improved by various authors [9, 12, 36, 25, 39]. Regarding their first result, the best-known improvement is due to Bugeaud and Gyarmati [9], where they showed that for . As for the second result, the best-known bound is , due to Stewart [39].
1.1. Main results
Our first result extends the result of Bugeaud and Gyarmati [9] to for a general nonzero integer .
Theorem 1.1.
Let be integers with and . Then we have
where is an absolute constant, and the implied constant is absolute. In particular, if is fixed, then .
The constant in the above theorem (and a few other results in this paper) is related to Linnik’s constant. It naturally comes from a quantitative version of Linnik’s theorem used in our proof. Also note that if is fixed, we have , which is of the same shape as the best-known upper bound on .
Our second result concerns the case .
Theorem 1.2.
Let be a nonzero integer and be a positive integer. If is a Diophantine tuple with property , then there is an absolute constant such that
where the implied constant is absolute. In particular, if , then .
Following Bérczes, Dujella, Hajdu, and Luca [3], for , let be the maximum such that there exists a set with elements and some such that is a perfect power for all with . Equivalently,
For our purpose, it is also natural to define a similar function to include those negative ’s:
In [3, Theorem 2 and Remark 2], Bérczes, Dujella, Hajdu, and Luca showed that as , and for . We provide a significant improvement on their upper bound on in the following corollary, which follows from 1.2 immediately.
Corollary 1.3.
There is an absolute constant such that
1.2. Improved bounds on bipartite Diophantine tuples
Tools from graph theory have been applied frequently to connect the property and the property . To see a quick connection, recall that the (multi-colored) Ramsey number is the smallest positive integer such that any coloring of the edges of the complete graph on vertices with colors results in the existence of an , such that there exists a complete monochromatic subgraph in the -th color with vertices. Let , and let be all the primes at most . If is a nonzero integer, then we have
(1.1) |
Indeed, if is a Diophantine tuple with property , then we can build a complete graph with vertex set . For each edge in the graph, we color it with color if is the smallest number such that is a perfect -th power. This simple observation immediately shows that is finite since we know each , except that the upper bound obtained from inequality (1.1) is potentially very large and far from the truth.
Our proofs of Theorem 1.1 and Theorem 1.2 are based on a novel combination of ideas from sieve methods, Diophantine approximation, and extremal graph theory. In particular, to prove these two theorems, it is natural to study the following question related to a “robust version” of Diophantine tuples, which is also of independent interest.
Question 1.4.
Let be a real number. Let and let be an nonzero integer. Suppose that is a subset of positive integers and among all subsets of with size , there are at least pairs of them such that is a perfect -th power. Can we give an upper bound on ?
Note that when , a set satisfying the condition in the question above is a Diophantine tuple with property . Also note that if is a sufficiently large integer and is a Diophantine tuple with property for some , then there is a prime with , such that satisfies the assumption in Question 1.4 with .
To answer Question 1.4, we take advantage of tools from extremal graph theory; see Section 2.1. In particular, in view of Kövari–Sós–Turán theorem (Lemma 2.2), it suffices to analyze the bipartite variant of Diophantine tuples; see Remark 2.3. Bipartite Diophantine tuples were recently introduced by the second author [41, 43] in connection with the work of Hajdu and Sárközy on the multiplicative decompositions of a small perturbation of the set of shifted -th powers [26, 27, 28]. However, the same objects have been studied for example by Gyarmati [23], Bugeaud and Dujella [8], and Bugeaud and Gyarmati [9], since two decades ago. More precisely, following [41], for each and each nonzero integer , we call a pair of sets a bipartite Diophantine tuple with property if are two subsets of with size at least , such that is a -th power for each and . From the viewpoint of number theory, such a bipartite variant of Diophantine tuples are also natural and useful local objects to study, since forbidden local structure often helps us to understand the global structure.
Generally speaking, bipartite Diophantine tuples are much harder to study compared to Diophantine tuples since being a Diophantine tuple imposes much more restrictions. For example, the quantities and were well-studied [15, 17] and eventually He, Togbé, and Ziegler [29] proved that , and Bonciocat, Cipu, and Mignotte [5] proved that . Nevertheless, it remains an open question to show that if is a bipartite Diophantine tuple with property (or , resp.), then is bounded by an absolute constant [2, 9, 43]; and the more general question for with appears to be even harder. The second author [41] studied the same question for . In particular, it was shown in [41, Theorem 2.2] that if is fixed, then holds for all bipartite Diophantine tuples with property . In the same paper, the following stronger result was proved. To state it, following [41], we define the following constants:
Theorem 1.5 ([41, Theorem 2.3]).
Let be fixed and be a nonzero integer. Let be a bipartite Diophantine tuple with property . Then we have:
-
(1)
If , then as .
-
(2)
If , then as .
The proof of 1.5 is based on a combination of tools from Diophantine approximation and sieve methods, and it relied on a result of Bourgain and Demeter [6] on the number of -th powers inside arithmetic progressions (see [41, Proposition 4.6]). We provide a substantial improvement on Theorem 1.5 in the next two results. In our proof, we use a more sophisticated application of sieve methods (building on the refined finite field models in Section 4) instead of considering -th powers in arithmetic progressions.
For positive integers and , we define the following constant:
(1.2) |
Theorem 1.6.
Let be fixed. Let be a nonzero integer and be a bipartite Diophantine tuple with property . Then as , the following estimate holds uniformly over an integer and :
Let be a bipartite Diophantine tuple with property , where and . Let us compare Theorem 1.5(1) and Theorem 1.6. Since for all and , we have Thus, if , by setting , Theorem 1.6 implies that
(1.3) |
always improving Theorem 1.5(1). Indeed, if is a prime, then inequality (1.3) becomes
when is composite, we get an even better exponent. When , Theorem 1.6 shows that , which improves Theorem 1.5(2). Our next result provides further refinement in this setting.
Theorem 1.7.
Let be fixed and be a nonzero integer with sufficiently large. Assume that is a bipartite Diophantine tuple with property . Then there are two constants only depending on , such that either or .
For our applications to Theorem 1.2, the following uniform version of Theorem 1.7 will be a key ingredient.
Theorem 1.8.
Let and be integers such that and . Assume that such that is a bipartite Diophantine tuple with property . Then there are two absolute constants such that either or .
1.3. Conditional results
In this section, we state our new conditional upper bounds on various notions of Diophantine tuples discussed earlier.
There are many conjectures related to sums of perfect powers, and they are often useful in Diophantine questions. The following conjecture, due to Lander, Parkin, and Selfridge [34], is widely believed.
Conjecture 1.9 (Lander–Parkin–Selfridge conjecture).
Let be positive integers. If where are positive integers such that for all and , then .
For our purpose, we need the following special case of Conjecture 1.9.
Conjecture 1.10 (Lander–Parkin–Selfridge conjecture, special case).
If is an integer, then there do not exist distinct positive integers such that
We remark that the results listed in this section still hold, at the cost of larger implied constants, if Conjecture 1.10 holds under the weaker assumption that for some fixed .
Our next result assumes Conjecture 1.10.
Theorem 1.11.
Assume Conjecture 1.10. If , , and is a bipartite Diophantine tuple with property , then . In particular, for all and .
We did not attempt to optimize the constant in the above theorem. In the next two remarks, we explain some motivations behind Theorem 1.11.
Remark 1.12.
Theorem 1.11 partially addresses a question asked in [41, Remark 4.9], where it is asked whether the ABC conjecture could be used to show that if and is a nonzero integer, then is absolutely bounded among all bipartite Diophantine tuples with property 111Note that when , because of Pell equations, for example when , there exists an infinite set such that is contained in the set of perfect squares. In particular, Theorem 1.11 does not hold for .. Theorem 1.11 answers this question in the affirmative in a much stronger form provided that assuming the Conjecture 1.10. While Conjecture 1.10 does not follow from the ABC conjecture, a slightly weaker version of the conjecture follows immediately from the -conjecture [7], which is a generalization of the ABC conjecture to more variables.
Remark 1.13.
The Uniformity Conjecture, due to Caporaso, Harris, and Mazur [10], states that if , then there is a constant , such that the number of -points of each curve defined over the rationals of genus is bounded by . Note that for , the hyperelliptic curve
has genus and has at least integral points. In particular, as .
If is a Diophantine tuple with property with and nonzero, we can take distinct elements from and consider the superelliptic curve
Since each corresponds to an integral point of , we have assuming the Uniformity Conjecture, where is the genus of . From the Riemann-Hurwitz formula, if and if . Thus, the Uniformity Conjecture only predicts that for all , where as . This is much weaker compared to the prediction from Theorem 1.11.
If is a fixed nonzero integer, then as increases, it is likely that decreases since being a perfect -th power for a larger is more restrictive. Indeed, it is proved in [41, Theorem 2.4 and Corollary 2.6] that for all provided that , and if and . Thus, it is plausible that there is an absolute constant such that for all and . In view of 1.13, by applying the Uniformity Conjecture for , and Theorem 1.11 for , this heuristic follows immediately.
Corollary 1.14.
Assume the Uniformity Conjecture and Conjecture 1.10. There is an absolute constant such that for all and .
Next, we consider conditional upper bounds on and .
Theorem 1.15.
Assume the Uniformity Conjecture and Conjecture 1.10. There is an absolute constant such that for all and . In particular, .
We also prove a weaker bound only assuming Conjecture 1.10.
Theorem 1.16.
Assume Conjecture 1.10. Then .
Finally, we turn our attention to conditional upper bounds on . So far there is no unconditional proof that , let alone for a general nonzero integer . The ABC conjecture states that for each , there is a constant , such that whenever are nonzero integers with and , we have
Under the ABC conjecture, Luca [36] showed that is finite. More generally, Bérczes, Dujella, Hajdu, and Luca [3, Theorem 4] showed that under the ABC conjecture, is finite whenever is a nonzero integer. More precisely, their proof shows that , where is an absolute constant and
is given by the Ramsey number, where , , and for . Note by a classical result of Erdös [20] on lower bounds of Ramsey numbers, thus their upper bound on is at least exponential in . Our next theorem provides a dramatic improvement on their bound.
Theorem 1.17.
Assume the ABC conjecture. If is a nonzero integer, then . In particular, there is an absolute constant , such that .
The following corollary follows immediately by combining Theorem 1.17 with Theorem 1.16 or Theorem 1.15.
Corollary 1.18.
Assume the ABC conjecture and Conjecture 1.10. Then ; and if we further assume the Uniformity Conjecture.
Notation. We follow standard notations in analytic number theory. In this paper, always denotes a prime, and and represent sums and products over all primes. We also use the Vinogradov notation ; we write if there is an absolute constant so that .
Structure of the paper. In Section 2, we provide additional background and prove some preliminary results. In Section 3, we bound the contribution of large elements in a Diophantine tuple with the desired property. In Section 4, we study various finite field analogues of Diophantine tuples as a preparation to apply sieve methods. In Section 5, we apply sieve methods to prove the improved bounds on bipartite Diophantine tuples stated in Section 1.2. Then, in Section 6, we combine results from all previous sections to prove Theorem 1.1 and Theorem 1.2. Finally, in Section 7, we prove the conditional results stated in Section 1.3.
2. Preliminaries
2.1. Tools from extremal graph theory
We first introduce a few basic terminologies from graph theory. A bipartite graph with bipartition (where ) is a graph with vertex set such that no two vertices in (resp. ) are adjacent. denotes the complete graph on vertices, namely, there is an edge between any pair of distinct vertices. denotes a complete bipartite graph with bipartition , where and , that is, there is an edge between and for all and .
Below we state two fundamental results in extremal graph theory. The first one is Turán’s theorem [40] regarding forbidden complete subgraphs, and the second one is the Kövari–Sós–Turán theorem [33] concerning forbidden complete bipartite subgraphs.
Lemma 2.1 (Turán’s theorem).
Let be a graph with vertices such that it does not contain as a subgraph. Then the number of edges of is at most .
Lemma 2.2 (Kövari–Sós–Turán theorem).
Let be a graph with vertices. If does not contain as a subgraph, where , then the number of edges of is at most .
In the next remark, we explain how a non-existence result of bipartite Diophantine tuples could be combined with the Kövari–Sós–Turán theorem to address Question 1.4.
Remark 2.3.
Let and . Let be a real number and let be a subset of positive integers, such that among all subsets of with size , there are at least pairs of them such that is a perfect -th power. Then we can build a graph with vertex set , such that two distinct vertices are adjacent if and only if is a perfect -th power. Then the given condition is equivalent to that the number of edges of is at least .
Suppose that we can show that there is no bipartite Diophantine tuple with property , where (in our actual applications, and might not be absolute constants, and could be slowly growing functions of , where ). Then in the graph theory language, this means that does not contain as a subgraph, and thus 2.2 implies that the number of edges of is at most . By comparing the lower and upper bound on the number of edges of , we have
It follows that .
Kövari–Sós–Turán theorem is essentially a consequence of the Cauchy-Schwartz inequality. For our purpose, we also need the following two helpful variants of Lemma 2.2.
Lemma 2.4 ([9, Lemma 4]).
Let be a bipartite graph with bipartition , where and the vertices of are labeled by positive integers. Suppose that for each and (and for each and ) with and such that , the induced subgraph is not complete bipartite. Then the number of edges of is at most .
Lemma 2.5 ([25, Lemma 2.4]).
Let be a graph with vertices, with the edge colored by colors. Suppose that does not contain a cycle through vertices such that the edges and have the same color, and the edges and have the same color. Then the number of edges of is at most .
2.2. Larger sieve
We will use sieve methods in our proofs. In particular, we will use a few variants of the larger sieve. We first recall Gallagher’s larger sieve [21].
Lemma 2.6 (Gallagher’s larger sieve).
Let and . Let be a set of primes. For each prime , let . For any , we have
provided that the denominator is positive.
Croot and Elsholtz [11] proved several variants of Gallagher’s larger sieve. In particular, the following variant refines Lemma 2.6 in certain ranges. Indeed, in the proof of Theorem 1.6, this variant outweighs the original version.
Lemma 2.7 ([11, Theorem 3]).
Let and . Let be a set of primes. For each prime , let . For any , we have
When one applies larger sieves, one usually needs to apply the prime number theorem for arithmetic progressions. However, in our applications, the modulus is not always fixed, so we need the following quantitative version of Linnik’s theorem; see for example [30, Corollary 18.8].
Lemma 2.8 (Quantitative Linnik’s theorem).
There exist positive constants and such that whenever is an integer, , and is an integer with , then
where the implied constant is absolute.
Combining Lemma 2.8 with the prime number theorem for arithmetic progressions with modulus bounded, we obtain the following corollary.
Corollary 2.9.
There exists positive constants and such that whenever is a positive integer, , and is an integer with , then
The following auxiliary estimate will be useful.
Lemma 2.10.
Let be a finite set of primes. Then we have .
Proof.
Let , and let be the first odd primes. Since the function is decreasing when , it follows that
3. Contribution of large elements
In this section, we deduce upper bounds on the number of “large” elements in Diophantine tuples with property . Our proofs are inspired by various ideas in [9, 24, 25, 42, 43].
We begin by collecting some known results regarding upper bounds on the number of “large” elements in Diophantine tuples with property .
The following lemma is due to Dujella [14].
Lemma 3.1 (Dujella).
Let be a nonzero integer. If is a Diophantine tuple with property , then .
The following lemma is a special case of [41, Proposition 4.1] due to the second author.
Lemma 3.2 ([41]).
Let and be a nonzero integer. Let if , and if . There do not exist integers and , such that , , and is a perfect -th power for all and .
Lemma 3.3 ([41, Proposition 4.3]).
Let and let be a nonzero integer. be a bipartite Diophantine tuple with property . If , then in both sets and , at most elements are at least .
Proposition 3.4.
Let and let be a nonzero integer. Let be two finite sets of positive integers in with . Then the number of pairs such that , and is is a perfect -th power is at most when , and at most when .
Proof.
Note that and for all . We first consider the case . In this case, the corollary follows immediately from Lemma 2.4 and 3.2 by building a bipartite graph with bipartition such that and are adjacent if and only if is a perfect -th power and the inequality . The proof of the case is similar; Lemma 2.4 implies that the number of edges is at most . ∎
Corollary 3.5.
Let and let be a nonzero integer. If a finite set of positive integers in , then the number of pairs such that with and is is a perfect -th power is at most when , at most when , and at most when .
Proof.
When , the corollary follows from Proposition 3.4 by setting .
The next proposition bounds the number of large elements in a Diophantine tuple with property , with finite.
Proposition 3.6.
Let be integers with and . If is a Diophantine tuple with property , then , where the implied constant is absolute.
Proof.
Since satisfies property , for each with , there is a prime such that is a perfect -th power. It follows from Corollary 3.5 that
It follows that , as required. ∎
We also need to bound the number of large elements in a Diophantine tuple with property . To achieve that, we apply tools from Diophantine approximation. We recall the following fundamental result in linear forms of logarithms of algebraic numbers; see for example [1]. For a nonzero rational number , where are coprime integers, its height is defined to .
Lemma 3.7 ([1]).
Let and be non-zero integers and let and be positive rational numbers. Put for , , and . Then there exists an effectively computable positive constant such that if , then
Proposition 3.8.
Let be a nonzero integer and let . There is an absolute constant such that if are distinct integers in , then there do not exist two primes such that there exist positive integers with
(3.1) |
Proof.
Let be the constant from Lemma 3.7. We show that the proposition is true with .
For the sake of contradiction, suppose that there exist primes and positive integers satisfying equation (3.1). Let . Without loss of generality, assume that . Note that . Also, we have
Observe that
It follows that
Thus,
where we used the assumption that .
Set
then and . It is easy to verify that if is a real number such that , then ; and if , then . It follows that and thus
On the other hand, we can apply Lemma 3.7 with , , , to get a lower bound on . Note that we have and . Then Lemma 3.7 implies that
It follows that
which contradicts the assumption that since . ∎
Now we are ready to prove a key result of the section.
Proposition 3.9.
Let be a nonzero integer and let . Let be a Diophantine tuple with property , then , where the implied constant is absolute.
Proof.
Let , and inductively we define for each integer . Let for each positive integer ; then we have , where . Thus, it suffices to show that for each .
Fix an integer with . Since is a Diophantine tuple with property , for each pair with , is a perfect -th power for some prime ; however, since , it is necessary that . Let be the constant from 3.8. By Corollary 3.5, the number of such pairs corresponding to primes is at most
Next, we give an upper bound on the number of pairs corresponding to primes using Lemma 2.5. Build a graph with vertex set . If with such that is a -th power for some prime ; then there is an edge between and and we color it using the smallest such . Then the edges of are colored by at most colors. By Proposition 3.8, does not contain a cycle through vertices such that the edges and have the same color, and the edges and have the same color. Thus, Lemma 2.5 implies that the number of edges in is at most
It follows that
It follows that , as required. ∎
4. Finite field models
In this section, we use character sum estimates to study bounds on various finite field analogues of Diophantine tuples as a preparation to apply sieve methods in Section 5. Throughout, let be the finite field with elements, and .
4.1. Diophantine tuples with property
The following Vinogradov-type double character sum estimate is well-known; see for example [23, Theorem 7] and [32, Proposition 3.1].
Lemma 4.1 (Vinogradov).
Let be a non-trivial multiplicative character of and . For any , we have
Next, we use Lemma 4.1 to deduce an upper bound on the finite field analogue of a Diophantine tuple with property .
Proposition 4.2.
Let be distinct primes and let be a prime such that for each . Let . For each , let . If a subset of such that for all with , then
Proof.
For a set , we use to denote , and to denote the indicator function of the set . For each , let be a multiplicative character of of order ; then by the orthogonality relation, we have
(4.1) |
Let . Since for all with , and for each , there is at most one such that , it follows that
(4.2) |
4.2. Bipartite Diophantine tuples
Next, we consider finite field analogues of bipartite Diophantine tuples. We begin by applying Weil’s bound on complete character sums.
Lemma 4.3.
Let and be a prime such that . Let , , and , such that for all and . If , then .
Proof.
Let be a multiplicative character of with order . Let . Note that for each and each , we have and thus . In particular, each is a solution to the system of equations
Now a classical application of Weil’s bound (see for example [35, Exercise 5.66]) implies that
as required. ∎
Next, we consider an explicit version of a variant of a double character sum estimate due to Karatsuba [31] (see also [38, Lemma 2.2]). For our applications, we establish an explicit dependence on the parameter for the upper bound below. Karatsuba’s estimate is better than Vinogradov’s estimate when and are asymmetric in the sense that the sizes of and are not comparable.
Lemma 4.4.
Let , and . Then for any non-trivial multiplicative character of and any positive integer , we have
Proof.
By Hölder’s inequality, we have
Note that
(4.7) |
where the sum on the right-hand side runs over all -triples . If there exists with for each , then a standard application of Weil’s bound implies that the second summand on the right-hand side of equation (4.7) is (see for example [30, Corollary 11.24]); otherwise the summand is trivially bounded by . Note that in the latter case, each appears at least twice, so the number of such tuples is at most . Therefore,
and the required estimate follows. ∎
Corollary 4.5.
Let be an integer and be a prime such that . Let and , such that for all and . If is a positive integer such that , then .
5. Bounds on Bipartite Diophantine tuples
In this section, we use sieve methods to prove the results stated in Section 1.2.
5.1. Proof of Theorem 1.6
Let . Let such that with . Then . Let and . It follows from Lemma 3.3 that and . We replace with a subset of of size if .
Consider the set of primes , where and
(5.1) |
has been defined in equation (1.2). Note that . Since , as , we have , and thus
(5.2) |
For each , let be the image of modulo and view as a subset of the finite field ; similarly, define . Let
We claim that . Indeed, if and , then there are two distinct elements such that . However, there are less than pairs of distinct elements , and for each such pair , the number of distinct prime factors of is at most since . A similar argument shows that and .
Let . Next we apply Lemma 2.7 to bound using the set of primes . By [37, Theorem 2.7], we have
and thus
(5.3) |
Let ; we claim that
(5.4) |
Note that we have , and . If , then we simply apply the trivial bound ; if , then we have
with and , and thus Lemma 4.3 implies that inequality (5.4).
It follows from inequality (5.4) that
(5.5) |
Note that from equation (5.2), we have
(5.6) |
as . On the other hand, by the prime number theorem for arithmetic progressions, we have
(5.7) |
By Lemma 2.10, for each , we have
(5.8) |
Thus, combining equation (5.7) and inequality (5.8), we have
(5.9) |
Combining inequalities (5.5), (5.6), and (5.9),
(5.10) |
Combining inequalities (5.3) and (5.10) with Lemma 2.7,
as required.
5.2. Proof of Theorems 1.7 and 1.8
The proof of Theorems 1.7 and 1.8 are inspired by several arguments used in [23, 41] for bipartite Diophantine tuples, as well as the “inverse sieve argument” developed by Elsholtz [19] in the inverse Goldbach problem.
Proof of Theorem 1.7.
Let . In view of 3.3, we may assume that . Let . Set
For each prime , let be the image of modulo , and define similarly. Let and let .
For each prime , we can view and as subsets of . For each and , is a perfect -th power. It follows that . Thus, by Corollary 4.5, if , that is, , then . Let
By the prime number for arithmetic progressions, we have
(5.11) |
We consider the following two cases.
Next, we use a similar idea to prove Theorem 1.8.
Proof of Theorem 1.8.
Let . Let and be the positive constants from Corollary 2.9. Without loss of generality, we may assume that and . Let
We define in the same way as in the proof of Theorem 1.7.
Since , Corollary 2.9 implies that
(5.14) |
We consider the following two cases.
Case 1: . As in the proof of Theorem 1.7, we have . It follows from inequality (5.14) that
(5.15) |
Applying Gallagher’s sieve (Lemma 2.6), inequalities (5.14) and (5.15), and the prime number theorem, we conclude that
where we used the assumption that in the last step, and the implied constants are all absolute.
Case 2: . As in the proof of Theorem 1.7, we have Now Gallagher’s sieve implies that
where we used the assumption that in the last step, and the implied constants are all absolute.
Thus, we conclude that either or , as desired. ∎
6. Bounds on Diophantine tuples with property
6.1. Proof of Theorem 1.1
Let and let be a Diophantine tuple with property . In view of Proposition 3.6, to prove the theorem, we can additionally assume that is sufficiently large and . In particular, since [32, Lemma 2.8], we can choose large enough so that
(6.1) |
List the primes up to by and set to be the product of these primes. By the prime number theorem, . Let and be the positive constants from Corollary 2.9. Without loss of generality, we may assume that and . Let
For each prime , let be the image of modulo and view as a subset of . Let with ; then , that is, there are and , such that . Since for each , it follows that , viewed as an element in , lies in the set . Thus, Proposition 4.2 implies that , where
Thus, . By the prime number theorem,
(6.2) |
Since , Corollary 2.9 implies that
(6.3) |
Thus, inequalities (6.1) and (6.3) imply that
(6.4) |
Applying Gallagher’s sieve (Lemma 2.6), inequalities (6.2) and (6.4), and the fact and , we conclude that
for some absolute constant , where the implied constant is absolute, as required.
6.2. Proof of Theorem 1.2
By Proposition 3.9, we may assume that , where . We may assume that is sufficiently large. Let be the complete graph with vertex set . For each with , by definition, is a perfect power; we can thus write for some positive integer and a prime and we color the edge by the smallest such . Note that each prime we used to color some edge satisfies that . Let be the two absolute constants from 1.8. Then it follows from 1.8 that does not contain a monochromatic as a subgraph. Thus, for each prime , Lemma 2.2 implies that the number of edges in with color is at most
Since the total number of edges in is , it follows that
and thus
where the implied constant is absolute, as required.
7. Proofs of conditional results
7.1. Conditional estimates on : Proof of Theorem 1.11
Assume that are two elements in . For each , there exist positive integers and such that and ; it follows that . In particular, for two distinct , they induce a nontrivial integer solution to the Diophantine equation in the sense that .
Suppose , then we can pick elements of to generate
More precisely, take distinct elements from , and for each , define via the following rule:
(7.1) |
It follows that the matrix
has determinant since . Thus, we have by the Leibniz formula for the determinant of that
(7.2) |
where runs over , the set of all permutations of . Since , if we can show that the 24 numbers are pairwise distinct for , then equation (7.2) will violate Conjecture 1.10.
Thus, it suffices to show that if , then we can always choose distinct elements from and for each , define via equation (7.1), such that the 24 numbers are pairwise distinct for , then we are done. Equivalently, we study the matrix
For convenience, we introduce the following terminology. For each , we call a positive integer an factor of if it appears as a term (up to the sign) in the Leibniz formula for the determinant of some submatrix of . Thus, our goal is to guarantee that all the factors of are distinct and we describe an algorithm to achieve this purpose.
The key of the algorithm will be based on the following simple observation: the function is monotone whenever . Since
it follows that if is fixed, then is also uniquely fixed. Similarly, if is fixed, then is also uniquely fixed. In particular, as long as are distinct, we automatically have for each . Similarly, as long as are distinct, we automatically have for each .
We choose from step by step such that they are in strictly increasing order, such that at each step, for each , all the factors in the matrix (with entries already determined) are distinct. We use the following algorithm to choose the elements :
-
(1)
choose to be the smallest element from .
-
(2)
choose the smallest element such that and .
-
(3)
choose the smallest element such that and .
-
(4)
choose the smallest element such that and .
Since , obviously the above algorithm went through. Note that for , all the factors in the matrix with entries already determined are distinct.
Next we pick step by step. For each , we pick using the following algorithm:
-
(1)
Let be the collection of all possible factors in the matrix (with entries already determined).
-
(2)
Let and let be the quotient set of .
-
(3)
Pick such that .
We first note that is upper bounded by the total number of factors, so
It follows from that . Note that for all , we have . It follows that the above algorithm works since .
Next, we claim that the choice of from the above algorithm guarantees that for each , all the factors in the matrix (with entries already determined) are distinct. Let be fixed, consider two factors and coming from different entries. Let be the set of labels of the entries building and . We consider the following four cases:
-
(1)
Both and are in . Without the loss of generality, we may assume that contains as an entry, and contains as an entry. Then if , we have and , and thus , violating the assumption that .
-
(2)
is in , and is not. Without the loss of generality, we may assume that contains as an entry. If also contains as an entry, then before we pick , we can already guarantee ; if does not contain as an entry, then would imply that and thus , violating the assumption.
-
(3)
is in , and is not. Similar to the analysis in the previous case, we have .
-
(4)
None of and are in . Then before we pick , we can already guarantee .
We conclude that ; this proves the claim.
Since , we can run the above algorithm to construct the desired step by step, as required.
7.2. Condition estimates on and : Proof of Theorem 1.15 and Theorem 1.16
Proof of Theorem 1.15.
Let and be a nonzero integer. Let be a Diophantine tuple with property . Then for all with . Let be the complete graph with the vertex set ; for each edge in , we color it with the smallest prime such that is a -th power. By definition, all the primes we used to color the edges of are at most . Let be the constant from 1.14; then does not contain a monochromatic as a subgraph. Let be the Ramsey number , where the number of ’s is . Since there are primes less than , it follows that contains no complete subgraph such that all edges have color in . Thus, Lemma 2.1 implies that the number of edges in with color in is at most . For each prime such that , 1.11 implies that does not contain a monochromatic in color as a subgraph and thus the number of edges in with color is by Lemma 2.2. We conclude that
and thus . This finishes the proof that , where the implied constant is absolute.
Next, we give an upper bound on . Let such that there is an integer with such that is a perfect power for all with ; since and , it is necessary for to be a -th power for some . This shows that is a Diophantine tuple with property and thus
as required. ∎
Proof of Theorem 1.16.
Assume that is sufficiently large. Pick a set with , such that there is some such that is a perfect power for all with . Let be the graph with the vertex set , such that there is an edge between two distinct vertices and if and only if . By definition, is a complete graph. For each such that , we color the edge with if , otherwise we color the edge with the smallest prime such that is a -th power. Note that each prime we used to color some edge satisfies that . By Theorem 1.1, there is an absolute constant , such that does not contain a monochromatic with color as a subgraph. Lemma 2.1 then implies that the number of edges in with color is at most . This implies that the number of edges in with color being a prime is at least
For each prime with , by Theorem 1.11, does not contain a monochromatic with color as a subgraph; thus, Lemma 2.2 implies that the number of edges in with color is . It follows that and thus . ∎
7.3. Conditional estimates on : Proof of Theorem 1.17
Proof of Theorem 1.17.
Let be a Diophantine tuple with property . Under the ABC conjecture, by [3, Lemma 1], there is an absolute constant , such that there do not exist distinct positive integers , each at least , with the property that for each , for some integers and with . Let . Define and . By definition, . It suffices to show is absolutely bounded, so that we have
for some absolute constant from Corollary 1.3.
Next, we show that is absolutely bounded. Label the primes less than by . We build a complete graph with vertex set , and we color the edge connecting two distinct elements with the color by the following rule:
-
•
If is a perfect -th power for some , then ;
-
•
Otherwise, set to be the smallest number such that is a perfect -th power.
In view of the above discussion, contains no monochromatic complete subgraph in color . By Lemma 3.1, contains no monochromatic complete subgraph in color . Let be the Ramsey number ; then by definition, contains no complete subgraph with all edges in color or . Lemma 2.1 then implies that the number of edges in with color or is at most . By Corollary 3.5, the number of edges in with color is at most , and for each , the number of edges in with color is at most . It follows that
and thus
We conclude that is absolutely bounded, as required. ∎
Acknowledgments
The second author thanks Andrej Dujella, Greg Martin, József Solymosi, and Zixiang Xu for helpful discussions at the early stages of the project. The research of the second author was supported in part by an NSERC fellowship.
References
- [1] A. Baker and G. Wüstholz. Logarithmic forms and group varieties. J. Reine Angew. Math., 442:19–62, 1993.
- [2] G. Batta, L. Hajdu, and A. Pongrácz. On Diophantine graphs, 2024. arXiv:2410.20120.
- [3] 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.
- [4] A. Bérczes, A. Dujella, L. Hajdu, and S. Tengely. Finiteness results for -Diophantine sets. Monatsh. Math., 180(3):469–484, 2016.
- [5] N. C. Bonciocat, M. Cipu, and M. Mignotte. There is no Diophantine -quadruple. J. Lond. Math. Soc. (2), 105(1):63–99, 2022.
- [6] J. Bourgain and C. Demeter. On the number of th powers inside arithmetic progressions. arXiv e-prints, 2018. arXiv:1811.11919.
- [7] J. Browkin and J. Brzeziński. Some remarks on the -conjecture. Math. Comp., 62(206):931–939, 1994.
- [8] Y. Bugeaud and A. Dujella. On a problem of Diophantus for higher powers. Math. Proc. Cambridge Philos. Soc., 135(1):1–10, 2003.
- [9] Y. Bugeaud and K. Gyarmati. On generalizations of a problem of Diophantus. Illinois J. Math., 48(4):1105–1115, 2004.
- [10] L. Caporaso, J. Harris, and B. Mazur. Uniformity of rational points. J. Amer. Math. Soc., 10(1):1–35, 1997.
- [11] E. S. Croot, III and C. Elsholtz. On variants of the larger sieve. Acta Math. Hungar., 103(3):243–254, 2004.
- [12] R. Dietmann, C. Elsholtz, K. Gyarmati, and M. Simonovits. Shifted products that are coprime pure powers. J. Combin. Theory Ser. A, 111(1):24–36, 2005.
- [13] A. B. Dixit, S. Kim, and M. R. Murty. Generalized Diophantine -tuples. Proc. Amer. Math. Soc., 150(4):1455–1465, 2022.
- [14] A. Dujella. On the size of Diophantine -tuples. Math. Proc. Cambridge Philos. Soc., 132(1):23–33, 2002.
- [15] A. Dujella. Bounds for the size of sets with the property . Glas. Mat. Ser. III, 39(59)(2):199–205, 2004.
- [16] A. Dujella. Diophantine -tuples and Elliptic Curves, volume 79 of Developments in Mathematics. Springer, Cham, 2024.
- [17] A. Dujella and C. Fuchs. Complete solution of a problem of Diophantus and Euler. J. London Math. Soc. (2), 71(1):33–52, 2005.
- [18] A. Dujella and F. Luca. Diophantine -tuples for primes. Int. Math. Res. Not., (47):2913–2940, 2005.
- [19] C. Elsholtz. The inverse Goldbach problem. Mathematika, 48(1-2):151–158, 2001.
- [20] P. Erdös. Some remarks on the theory of graphs. Bull. Amer. Math. Soc., 53:292–294, 1947.
- [21] P. X. Gallagher. A larger sieve. Acta Arith., 18:77–81, 1971.
- [22] A. M. Güloğlu and M. R. Murty. The Paley graph conjecture and Diophantine -tuples. J. Combin. Theory Ser. A, 170:105155, 9, 2020.
- [23] K. Gyarmati. On a problem of Diophantus. Acta Arith., 97(1):53–65, 2001.
- [24] K. Gyarmati, A. Sárközy, and C. L. Stewart. On shifted products which are powers. Mathematika, 49(1-2):227–230, 2002.
- [25] K. Gyarmati and C. L. Stewart. On powers in shifted products. Glas. Mat. Ser. III, 42(62)(2):273–279, 2007.
- [26] L. Hajdu and A. Sárközy. On multiplicative decompositions of polynomial sequences, I. Acta Arith., 184(2):139–150, 2018.
- [27] L. Hajdu and A. Sárközy. On multiplicative decompositions of polynomial sequences, II. Acta Arith., 186(2):191–200, 2018.
- [28] L. Hajdu and A. Sárközy. On multiplicative decompositions of polynomial sequences, III. Acta Arith., 193(2):193–216, 2020.
- [29] B. He, A. Togbé, and V. Ziegler. There is no Diophantine quintuple. Trans. Amer. Math. Soc., 371(9):6665–6709, 2019.
- [30] H. Iwaniec and E. Kowalski. Analytic number theory, volume 53 of American Mathematical Society Colloquium Publications. American Mathematical Society, Providence, RI, 2004.
- [31] A. A. Karatsuba. Distribution of values of Dirichlet characters on additive sequences. Dokl. Akad. Nauk SSSR, 319(3):543–545, 1991.
- [32] S. Kim, C. H. Yip, and S. Yoo. Multiplicative structure of shifted multiplicative subgroups and its applications to Diophantine tuples, 2023. Canad. J. Math., to appear. arXiv:2309.09124.
- [33] T. Kövari, V. T. Sós, and P. Turán. On a problem of K. Zarankiewicz. Colloq. Math., 3:50–57, 1954.
- [34] L. J. Lander, T. R. Parkin, and J. L. Selfridge. A survey of equal sums of like powers. Math. Comp., 21:446–459, 1967.
- [35] R. Lidl and H. Niederreiter. Finite fields, volume 20 of Encyclopedia of Mathematics and its Applications. Cambridge University Press, Cambridge, second edition, 1997.
- [36] F. Luca. On shifted products which are powers. Glas. Mat. Ser. III, 40(60)(1):13–20, 2005.
- [37] 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.
- [38] I. E. Shparlinski. Additive decompositions of subgroups of finite fields. SIAM J. Discrete Math., 27(4):1870–1879, 2013.
- [39] C. L. Stewart. On sets of integers whose shifted products are powers. J. Combin. Theory Ser. A, 115(4):662–673, 2008.
- [40] P. Turán. Eine Extremalaufgabe aus der Graphentheorie. Mat. Fiz. Lapok, 48:436–452, 1941.
- [41] C. H. Yip. Multiplicatively reducible subsets of shifted perfect -th powers and bipartite Diophantine tuples, 2023. Acta Arith., to appear. arXiv:2312.14450.
- [42] C. H. Yip. Improved upper bounds on Diophantine tuples with the property , 2024. Bull. Aust. Math. Soc., to appear. arXiv:2406.00840.
- [43] C. H. Yip. Multiplicatively irreducibility of small perturbations of shifted -th powers, 2025. arXiv:2501.16620.