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

\stackMath

Diophantine tuples and product sets in shifted powers

Ernie Croot School of Mathematics
Georgia Institute of Technology
Atlanta, GA 30332
United States
ernest.croot@math.gatech.edu
 and  Chi Hoi Yip School of Mathematics
Georgia Institute of Technology
Atlanta, GA 30332
United States
cyip30@gatech.edu
Abstract.

Let k2k\geq 2 and n0n\neq 0. A Diophantine tuple with property Dk(n)D_{k}(n) is a set of positive integers AA such that ab+nab+n is a kk-th power for all a,bAa,b\in A with aba\neq b. 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 powers
2020 Mathematics Subject Classification:
Primary 11B30, 11D72; Secondary 11N36, 11D41, 05C35

1. Introduction

A set {a1,a2,,am}\{a_{1},a_{2},\ldots,a_{m}\} of distinct positive integers is a Diophantine mm-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 nn be a nonzero integer and let k2k\geq 2, a set AA of distinct positive integers a Diophantine tuple with property Dk(n)D_{k}(n) if the product of any two distinct elements in AA is nn less than a perfect kk-th power. Following the standard notations, we also write

Mk(n)=sup{|A|:A satisfies the property Dk(n)}.M_{k}(n)=\sup\{|A|\colon A\subseteq{\mathbb{N}}\text{ satisfies the property }D_{k}(n)\}.

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 Mk(n)M_{k}(n) is of the form Mk(n)klog|n|M_{k}(n)\ll_{k}\log|n|; see [32, 41, 42] for the best-known implied constant depending on kk. Under the Uniformity Conjecture [10] (a consequence of the Bombieri–Lang conjecture), it is well-known that for each k2k\geq 2, there is a constant CkC_{k} such that Mk(n)CkM_{k}(n)\leq C_{k} holds for all nonzero integers nn [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 CC such that Mk(n)CM_{k}(n)\leq C holds for all k2k\geq 2 and n0n\neq 0.

In 2002, Gyarmati, Sárközy, and Stewart [24] initialized the study of two closely related variants of Diophantine tuples by enlarging the set of perfect kk-th powers (with a fixed kk) to the set of perfect kk-th powers with bounded kk or the set of all perfect powers. More precisely, set

Vd={xk:x,2kd}V_{d}=\{x^{k}:x\in\mathbb{N},2\leq k\leq d\}

for each integer d2d\geq 2; then the set of perfect powers is

V:=d2Vd={xk:x,k,k2}.V_{\infty}:=\bigcup_{d\geq 2}V_{d}=\{x^{k}:x,k\in\mathbb{N},k\geq 2\}.

They studied the size of a set of positive integers AA such that aa+1aa^{\prime}+1 is in VdV_{d} (where 2d2\leq d\leq\infty) for all a,aAa,a^{\prime}\in A that are distinct. More generally, one can study the same question by replacing the shift 11 with any nonzero shift; see for example [3, 4]. For brevity, we extend the familiar notions Dk(n)D_{k}(n) and Mk(n)M_{k}(n) in this more general setting. Let nn be a nonzero integer and let 2d2\leq d\leq\infty, we say a set AA of positive integers a Diophantine tuple with property Dd(n)D_{\leq d}(n) if the product of any two distinct elements in AA is nn less than an element in VdV_{d}, and we denote

Md(n)=sup{|A|:A satisfies the property Dd(n)}.M_{\leq d}(n)=\sup\{|A|\colon A\subseteq{\mathbb{N}}\text{ satisfies the property }D_{\leq d}(n)\}.

In particular, the case n=1n=1 is well-studied. Gyarmati, Sárközy, and Stewart [24] proved that if 2d<2\leq d<\infty and A{1,2,,N}A\subseteq\{1,2,\ldots,N\} is a Diophantine tuple with property Dd(1)D_{\leq d}(1), then |A|d2(logd)2loglogN|A|\ll\frac{d^{2}}{(\log d)^{2}}\log\log N. They then deduced that if A{1,2,,N}A\subseteq\{1,2,\ldots,N\} is a Diophantine tuple with property D(1)D_{\leq\infty}(1), then |A|(logN)2/loglogN|A|\ll(\log N)^{2}/\log\log N. 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 Md(1)(d/logd)2M_{\leq d}(1)\ll(d/\log d)^{2} for 2d<2\leq d<\infty. As for the second result, the best-known bound is |A|(logN)2/3(loglogN)1/3|A|\ll(\log N)^{2/3}(\log\log N)^{1/3}, due to Stewart [39].

1.1. Main results

Our first result extends the result of Bugeaud and Gyarmati [9] to Md(n)M_{\leq d}(n) for a general nonzero integer nn.

Theorem 1.1.

Let d,nd,n be integers with 2d<2\leq d<\infty and n0n\neq 0. Then we have

Md(n)d2(logd)2+edLlog|n|,M_{\leq d}(n)\ll\frac{d^{2}}{(\log d)^{2}}+e^{dL}\log|n|,

where LL is an absolute constant, and the implied constant is absolute. In particular, if 2d<2\leq d<\infty is fixed, then Md(n)log|n|M_{\leq d}(n)\ll\log|n|.

The constant LL 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 d<d<\infty is fixed, we have Md(n)dlog|n|M_{\leq d}(n)\ll_{d}\log|n|, which is of the same shape as the best-known upper bound on Md(n)M_{d}(n).

Our second result concerns the case d=d=\infty.

Theorem 1.2.

Let nn be a nonzero integer and NN be a positive integer. If A[1,N]A\subseteq[1,N] is a Diophantine tuple with property D(n)D_{\leq\infty}(n), then there is an absolute constant LL such that

|A|exp(L(loglog|n|)2)+logN,|A|\ll\exp\big{(}L(\log\log|n|)^{2}\big{)}+\log N,

where the implied constant is absolute. In particular, if L(loglog|n|)2loglogNL(\log\log|n|)^{2}\leq\log\log N, then |A|logN|A|\ll\log N.

Following Bérczes, Dujella, Hajdu, and Luca [3], for x1x\geq 1, let f(x)f(x) be the maximum KK such that there exists a set A[1,x]A\subseteq[1,x]\cap\mathbb{N} with KK elements and some 1nx1\leq n\leq x such that ab+nab+n is a perfect power for all a,bAa,b\in A with aba\neq b. Equivalently,

f(x)=max{|A|:A[1,x] has property D(n) for some 1nx}.f(x)=\max\{|A|:A\subseteq[1,x]\cap\mathbb{N}\text{ has property }D_{\leq\infty}(n)\text{ for some }1\leq n\leq x\}.

For our purpose, it is also natural to define a similar function to include those negative nn’s:

f~(x)=max{|A|:A[1,x] has property D(n) for some 1|n|x}.\widetilde{f}(x)=\max\{|A|:A\subseteq[1,x]\cap\mathbb{N}\text{ has property }D_{\leq\infty}(n)\text{ for some }1\leq|n|\leq x\}.

In [3, Theorem 2 and Remark 2], Bérczes, Dujella, Hajdu, and Luca showed that f(x)x2/3+o(1)f(x)\leq x^{2/3+o(1)} as xx\to\infty, and f(x)(loglogx/2logloglogx)1/3f(x)\geq\lfloor(\log\log x/2\log\log\log x)^{1/3}\rfloor for x>eeex>e^{e^{e}}. We provide a significant improvement on their upper bound on f(x)f(x) in the following corollary, which follows from 1.2 immediately.

Corollary 1.3.

There is an absolute constant LL such that

f(x)f~(x)exp(L(loglogx)2).f(x)\leq\widetilde{f}(x)\ll\exp\big{(}L(\log\log x)^{2}\big{)}.

In the next subsection, we provide an overview of our proof of Theorem 1.1 and Theorem 1.2, provide some additional backgrounds, and state our additional contributions.

1.2. Improved bounds on bipartite Diophantine tuples

Tools from graph theory have been applied frequently to connect the property Dk(n)D_{k}(n) and the property Dd(n)D_{\leq d}(n). To see a quick connection, recall that the (multi-colored) Ramsey number R(m1,m2,,ms)R(m_{1},m_{2},\ldots,m_{s}) is the smallest positive integer RR such that any coloring of the edges of the complete graph on RR vertices with ss colors results in the existence of an 1is1\leq i\leq s, such that there exists a complete monochromatic subgraph in the ii-th color with mim_{i} vertices. Let 2d<2\leq d<\infty, and let p1<p2<<pp_{1}<p_{2}<\ldots<p_{\ell} be all the primes at most dd. If nn is a nonzero integer, then we have

Md(n)R(Mp1(n),Mp2(n),,Mp(n)).M_{\leq d}(n)\leq R(M_{p_{1}}(n),M_{p_{2}}(n),\ldots,M_{p_{\ell}}(n)). (1.1)

Indeed, if AA is a Diophantine tuple with property Md(n)M_{\leq d}(n), then we can build a complete graph with vertex set AA. For each edge abab in the graph, we color it with color ii if ii is the smallest number such that ab+nab+n is a perfect pip_{i}-th power. This simple observation immediately shows that Md(n)M_{\leq d}(n) is finite since we know each Mpi(n)pilog|n|M_{p_{i}}(n)\ll_{p_{i}}\log|n|, 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 δ(0,1]\delta\in(0,1] be a real number. Let k2k\geq 2 and let nn be an nonzero integer. Suppose that XX is a subset of positive integers and among all subsets {a,b}\{a,b\} of XX with size 22, there are at least δ(|X|2)\delta\binom{|X|}{2} pairs of them such that ab+nab+n is a perfect kk-th power. Can we give an upper bound on |X||X|?

Note that when δ=1\delta=1, a set AA satisfying the condition in the question above is a Diophantine tuple with property Dk(n)D_{k}(n). Also note that if NN is a sufficiently large integer and A[1,N]A\subseteq[1,N] is a Diophantine tuple with property D(n)D_{\leq\infty}(n) for some 1|n|N1\leq|n|\leq N, then there is a prime kk with 2klogN2\leq k\leq\log N, such that AA satisfies the assumption in Question 1.4 with δ=1logN\delta=\frac{1}{\log N}.

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 kk-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 k2k\geq 2 and each nonzero integer nn, we call a pair of sets (A,B)(A,B) a bipartite Diophantine tuple with property BDk(n)BD_{k}(n) if A,BA,B are two subsets of \mathbb{N} with size at least 22, such that ab+nab+n is a kk-th power for each aAa\in A and bBb\in B. 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 M2(1)M_{2}(1) and M2(1)M_{2}(-1) were well-studied [15, 17] and eventually He, Togbé, and Ziegler [29] proved that M2(1)=4M_{2}(1)=4, and Bonciocat, Cipu, and Mignotte [5] proved that M2(1)=3M_{2}(-1)=3. Nevertheless, it remains an open question to show that if (A,B)(A,B) is a bipartite Diophantine tuple with property BD2(1)BD_{2}(1) (or BD2(1)BD_{2}(-1), resp.), then min{|A|,|B|}\min\{|A|,|B|\} is bounded by an absolute constant [2, 9, 43]; and the more general question for BD2(n)BD_{2}(n) with |n|2|n|\geq 2 appears to be even harder. The second author [41] studied the same question for k3k\geq 3. In particular, it was shown in [41, Theorem 2.2] that if k3k\geq 3 is fixed, then min{|A|,|B|}log|n|\min\{|A|,|B|\}\ll\log|n| holds for all bipartite Diophantine tuples (A,B)(A,B) with property BDk(n)BD_{k}(n). In the same paper, the following stronger result was proved. To state it, following [41], we define the following constants:

r3=9,r4=6,r5=5, and rk=4 for k6;\displaystyle r_{3}=9,\quad r_{4}=6,\quad r_{5}=5,\quad\text{ and }\quad r_{k}=4\quad\text{ for }k\geq 6;
s3=6,s4=4,s5=3, and sk=2 for k6;\displaystyle s_{3}=6,\quad s_{4}=4,\quad s_{5}=3,\quad\text{ and }\quad s_{k}=2\quad\text{ for }k\geq 6;
t3=15399938,t4=343,t5=9723,t6=294, and tk=k2+k4k26k+6 for k7.t_{3}=\frac{15399}{938},\quad t_{4}=\frac{34}{3},\quad t_{5}=\frac{97}{23},\quad t_{6}=\frac{29}{4},\text{ and }\quad t_{k}=\frac{k^{2}+k-4}{k^{2}-6k+6}\quad\text{ for }k\geq 7.
Theorem 1.5 ([41, Theorem 2.3]).

Let k3k\geq 3 be fixed and nn be a nonzero integer. Let (A,B)(A,B) be a bipartite Diophantine tuple with property BDk(n)BD_{k}(n). Then we have:

  1. (1)

    If |A|rk|A|\geq r_{k}, then |B||n|tkk+o(1)|B|\leq|n|^{\frac{t_{k}}{k}+o(1)} as |n||n|\to\infty.

  2. (2)

    If |A|loglog|n|+3.3log(k1)+8|A|\geq\frac{\log\log|n|+3.3}{\log(k-1)}+8, then |B||n|1k2+o(1)|B|\leq|n|^{\frac{1}{k-2}+o(1)} as |n||n|\to\infty.

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 kk-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 kk-th powers in arithmetic progressions.

For positive integers kk and mm, we define the following constant:

θk,m:=1ikgcd(i,k)=1gcd(i1,k)m.\theta_{k,m}:=\sum_{\begin{subarray}{c}1\leq i\leq k\\ \gcd(i,k)=1\end{subarray}}\gcd(i-1,k)^{m}. (1.2)
Theorem 1.6.

Let k3k\geq 3 be fixed. Let nn be a nonzero integer and (A,B)(A,B) be a bipartite Diophantine tuple with property BDk(n)BD_{k}(n). Then as |n||n|\to\infty, the following estimate holds uniformly over an integer mrkskm\geq r_{k}-s_{k} and m=o(loglog|n|)m=o(\log\log|n|):

|A|m+sk|B||n|tkϕ(k)θk,m+o(1).|A|\geq m+s_{k}\implies|B|\leq|n|^{\frac{t_{k}\phi(k)}{\theta_{k,m}}+o(1)}.

Let (A,B)(A,B) be a bipartite Diophantine tuple with property BDk(n)BD_{k}(n), where k3k\geq 3 and n0n\neq 0. Let us compare Theorem 1.5(1) and Theorem 1.6. Since gcd(i1,k)=11\gcd(i-1,k)=1\geq 1 for all ii and gcd(0,k)=k\gcd(0,k)=k, we have θk,mkm+ϕ(k)1.\theta_{k,m}\geq k^{m}+\phi(k)-1. Thus, if |A|rk|A|\geq r_{k}, by setting m=rkskm=r_{k}-s_{k}, Theorem 1.6 implies that

|B||n|tkϕ(k)k2+ϕ(k)1+o(1),|B|\leq|n|^{\frac{t_{k}\phi(k)}{k^{2}+\phi(k)-1}+o(1)}, (1.3)

always improving Theorem 1.5(1). Indeed, if kk is a prime, then inequality (1.3) becomes

|B||n|tk(k1)k2+k2+o(1)=|n|tkk+2+o(1);|B|\leq|n|^{\frac{t_{k}(k-1)}{k^{2}+k-2}+o(1)}=|n|^{\frac{t_{k}}{k+2}+o(1)};

when kk is composite, we get an even better exponent. When |A|loglog|n||A|\gg\log\log|n|, Theorem 1.6 shows that |B|=|n|o(1)|B|=|n|^{o(1)}, which improves Theorem 1.5(2). Our next result provides further refinement in this setting.

Theorem 1.7.

Let k3k\geq 3 be fixed and nn be a nonzero integer with |n||n| sufficiently large. Assume that (A,B)(A,B) is a bipartite Diophantine tuple with property BDk(n)BD_{k}(n). Then there are two constants C1,C2C_{1},C_{2} only depending on kk, such that either |A|C1loglog|n||A|\leq C_{1}\log\log|n| or |B|C2log|n|(loglog|n|)2|B|\leq C_{2}\log|n|\cdot(\log\log|n|)^{2}.

For our applications to Theorem 1.2, the following uniform version of Theorem 1.7 will be a key ingredient.

Theorem 1.8.

Let k2k\geq 2 and n,Nn,N be integers such that 1|n|N1\leq|n|\leq N and k5logNk\leq 5\log N. Assume that A,B{1,2,,N}A,B\subseteq\{1,2,\ldots,N\} such that (A,B)(A,B) is a bipartite Diophantine tuple with property BDk(n)BD_{k}(n). Then there are two absolute constants L1,L2L_{1},L_{2} such that either |A|L1loglogN|A|\leq L_{1}\log\log N or |B|(logN)L2|B|\leq(\log N)^{L_{2}}.

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 m,n,km,n,k be positive integers. If i=1naik=j=1mbjk,\sum_{i=1}^{n}a_{i}^{k}=\sum_{j=1}^{m}b_{j}^{k}, where a1,a2,,an,b1,b2,,bma_{1},a_{2},\ldots,a_{n},b_{1},b_{2},\ldots,b_{m} are positive integers such that aibja_{i}\neq b_{j} for all 1in1\leq i\leq n and 1jm1\leq j\leq m, then m+nkm+n\geq k.

For our purpose, we need the following special case of Conjecture 1.9.

Conjecture 1.10 (Lander–Parkin–Selfridge conjecture, special case).

If k25k\geq 25 is an integer, then there do not exist distinct positive integers a1,b1,,a12,b12a_{1},b_{1},\ldots,a_{12},b_{12} such that i=112aik=i=112bik.\sum_{i=1}^{12}a_{i}^{k}=\sum_{i=1}^{12}b_{i}^{k}.

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 kk0k\geq k_{0} for some fixed k0k_{0}.

Our next result assumes Conjecture 1.10.

Theorem 1.11.

Assume Conjecture 1.10. If k25k\geq 25, n0n\neq 0, and (A,B)(A,B) is a bipartite Diophantine tuple with property BDk(n)BD_{k}(n), then max{|A|,|B|}21736\max\{|A|,|B|\}\leq 21736. In particular, Mk(n)21738M_{k}(n)\leq 21738 for all k25k\geq 25 and n0n\neq 0.

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 k3k\geq 3 and nn is a nonzero integer, then |A||B||A||B| is absolutely bounded among all bipartite Diophantine tuples (A,B)(A,B) with property BDk(n)BD_{k}(n) 111Note that when k=2k=2, because of Pell equations, for example when A={1,2}A=\{1,2\}, there exists an infinite set BB such that AB+1AB+1 is contained in the set of perfect squares. In particular, Theorem 1.11 does not hold for k=2k=2.. Theorem 1.11 answers this question in the affirmative in a much stronger form provided that k25k\geq 25 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 nn-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 g2g\geq 2, then there is a constant CgC_{g}, such that the number of \mathbb{Q}-points of each curve defined over the rationals \mathbb{Q} of genus gg is bounded by CgC_{g}. Note that for i{1,2}i\in\{1,2\}, the hyperelliptic curve

y2=(x1)(x2)(x(2g+i))y^{2}=(x-1)(x-2)\cdots(x-(2g+i))

has genus gg and has at least 2g+i2g+i integral points. In particular, CgC_{g}\to\infty as gg\to\infty.

If AA is a Diophantine tuple with property Dk(n)D_{k}(n) with k4k\geq 4 and nn nonzero, we can take 33 distinct elements a1,a2,a3a_{1},a_{2},a_{3} from AA and consider the superelliptic curve

E:yk=(a1x+n)(a2x+n)(a3x+n).E:y^{k}=(a_{1}x+n)(a_{2}x+n)(a_{3}x+n).

Since each aA{a1,a2,a3}a\in A\setminus\{a_{1},a_{2},a_{3}\} corresponds to an integral point of EE, we have |A||E()|+3|E()|+3Cg+3|A|\leq|E(\mathbb{Z})|+3\leq|E(\mathbb{Q})|+3\leq C_{g}+3 assuming the Uniformity Conjecture, where gg is the genus of EE. From the Riemann-Hurwitz formula, g=k2g=k-2 if 3k3\mid k and g=k1g=k-1 if 3k3\nmid k. Thus, the Uniformity Conjecture only predicts that Mk(n)Ck~M_{k}(n)\leq\widetilde{C_{k}} for all n0n\neq 0, where Ck~\widetilde{C_{k}}\to\infty as kk\to\infty. This is much weaker compared to the prediction from Theorem 1.11.

If nn is a fixed nonzero integer, then as kk increases, it is likely that Mk(n)M_{k}(n) decreases since being a perfect kk-th power for a larger kk is more restrictive. Indeed, it is proved in [41, Theorem 2.4 and Corollary 2.6] that Mk(n)9M_{k}(n)\leq 9 for all k2k\geq 2 provided that |n|=1|n|=1, and Mk(n)19M_{k}(n)\leq 19 if |n|2|n|\geq 2 and k2log|n|+2k\geq 2\log|n|+2. Thus, it is plausible that there is an absolute constant CC such that Mk(n)CM_{k}(n)\leq C for all k2k\geq 2 and n0n\neq 0. In view of 1.13, by applying the Uniformity Conjecture for k24k\leq 24, and Theorem 1.11 for k25k\geq 25, this heuristic follows immediately.

Corollary 1.14.

Assume the Uniformity Conjecture and Conjecture 1.10. There is an absolute constant CC such that Mk(n)CM_{k}(n)\leq C for all k2k\geq 2 and n0n\neq 0.

Next, we consider conditional upper bounds on Md(n)M_{\leq d}(n) and f~(x)\widetilde{f}(x).

Theorem 1.15.

Assume the Uniformity Conjecture and Conjecture 1.10. There is an absolute constant CC^{\prime} such that Md(n)C(d/logd)2M_{\leq d}(n)\leq C^{\prime}(d/\log d)^{2} for all 2d<2\leq d<\infty and n0n\neq 0. In particular, f~(x)(logx/loglogx)2\widetilde{f}(x)\ll(\log x/\log\log x)^{2}.

We also prove a weaker bound only assuming Conjecture 1.10.

Theorem 1.16.

Assume Conjecture 1.10. Then f~(x)(logx)4\widetilde{f}(x)\ll(\log x)^{4}.

Finally, we turn our attention to conditional upper bounds on M(n)M_{\leq\infty}(n). So far there is no unconditional proof that M(1)<M_{\leq\infty}(1)<\infty, let alone M(n)M_{\leq\infty}(n) for a general nonzero integer nn. The ABC conjecture states that for each ϵ>0\epsilon>0, there is a constant CϵC_{\epsilon}, such that whenever a,b,ca,b,c are nonzero integers with gcd(a,b,c)=1\gcd(a,b,c)=1 and a+b=ca+b=c, we have

max{|a|,|b|,|c|}Cϵrad(abc)1+ϵ.\max\{|a|,|b|,|c|\}\leq C_{\epsilon}\operatorname{rad}(abc)^{1+\epsilon}.

Under the ABC conjecture, Luca [36] showed that M(1)M_{\leq\infty}(1) is finite. More generally, Bérczes, Dujella, Hajdu, and Luca [3, Theorem 4] showed that under the ABC conjecture, M(n)M_{\leq\infty}(n) is finite whenever nn is a nonzero integer. More precisely, their proof shows that M(n)c0|n|3+R(n)M_{\leq\infty}(n)\leq c_{0}|n|^{3}+R(n), where c0c_{0} is an absolute constant and

R(n)=R(C1(2,n),C1(3,n),C1(5,n),,C1(3203,n),5),R(n)=R(C_{1}(2,n),C_{1}(3,n),C_{1}(5,n),\ldots,C_{1}(3203,n),5),

is given by the Ramsey number, where C1(2,n)=31+15.476log|n|C_{1}(2,n)=31+15.476\log|n|, C1(3,n)=2|n|17+6C_{1}(3,n)=2|n|^{17}+6, and C1(k,n)=2|n|5+3C_{1}(k,n)=2|n|^{5}+3 for k5k\geq 5. Note R(n)(2)|n|5R(n)\gg(\sqrt{2})^{|n|^{5}} by a classical result of Erdös [20] on lower bounds of Ramsey numbers, thus their upper bound on M(n)M_{\leq\infty}(n) is at least exponential in |n||n|. Our next theorem provides a dramatic improvement on their bound.

Theorem 1.17.

Assume the ABC conjecture. If nn is a nonzero integer, then M(n)f~(2|n|17)M_{\leq\infty}(n)\ll\widetilde{f}(2|n|^{17}). In particular, there is an absolute constant LL, such that M(n)exp(L(loglog|n|)2)M_{\leq\infty}(n)\ll\exp\big{(}L(\log\log|n|)^{2}\big{)}.

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 M(n)(log|n|)4M_{\leq\infty}(n)\ll(\log|n|)^{4}; and M(n)(log|n|/loglog|n|)2M_{\leq\infty}(n)\ll(\log|n|/\log\log|n|)^{2} if we further assume the Uniformity Conjecture.

Notation. We follow standard notations in analytic number theory. In this paper, pp always denotes a prime, and p\sum_{p} and p\prod_{p} represent sums and products over all primes. We also use the Vinogradov notation \ll; we write XYX\ll Y if there is an absolute constant C>0C>0 so that |X|CY|X|\leq CY.

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 GG with bipartition (A,B)(A,B) (where AB=A\cap B=\emptyset) is a graph with vertex set ABA\cup B such that no two vertices in AA (resp. BB) are adjacent. KrK_{r} denotes the complete graph on rr vertices, namely, there is an edge between any pair of distinct vertices. Ks,tK_{s,t} denotes a complete bipartite graph with bipartition (A,B)(A,B), where |A|=s|A|=s and |B|=t|B|=t, that is, there is an edge between aa and bb for all aAa\in A and bBb\in B.

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 GG be a graph with nn vertices such that it does not contain KrK_{r} as a subgraph. Then the number of edges of GG is at most 12(11r1)n2\frac{1}{2}(1-\frac{1}{r-1})n^{2}.

Lemma 2.2 (Kövari–Sós–Turán theorem).

Let GG be a graph with nn vertices. If GG does not contain Ks,tK_{s,t} as a subgraph, where sts\leq t, then the number of edges of GG is at most (t1)1/sn21/s+(s1)n(t-1)^{1/s}n^{2-1/s}+(s-1)n.

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 k2k\geq 2 and n0n\neq 0. Let δ(0,1]\delta\in(0,1] be a real number and let XX be a subset of positive integers, such that among all subsets {a,b}\{a,b\} of XX with size 22, there are at least δ(|X|2)\delta\binom{|X|}{2} pairs of them such that ab+nab+n is a perfect kk-th power. Then we can build a graph GG with vertex set XX, such that two distinct vertices a,bXa,b\in X are adjacent if and only if ab+nab+n is a perfect kk-th power. Then the given condition is equivalent to that the number of edges of GG is at least δ(|X|2)\delta\binom{|X|}{2}.

Suppose that we can show that there is no bipartite Diophantine tuple (A,B)(A,B) with property BDk(n)BD_{k}(n), where 2s=|A||B|=t2\leq s=|A|\leq|B|=t (in our actual applications, ss and tt might not be absolute constants, and could be slowly growing functions of NN, where A,B[N]A,B\subseteq[N]). Then in the graph theory language, this means that GG does not contain Ks,tK_{s,t} as a subgraph, and thus 2.2 implies that the number of edges of GG is at most (t1)1/s|X|21/s+(s1)|X|(t-1)^{1/s}|X|^{2-1/s}+(s-1)|X|. By comparing the lower and upper bound on the number of edges of GG, we have

δ(|X|2)(t1)1/s|X|21/s+(s1)|X|.\delta\binom{|X|}{2}\leq(t-1)^{1/s}|X|^{2-1/s}+(s-1)|X|.

It follows that |X|(2/δ)s(t1)+2s/δ|X|\leq(2/\delta)^{s}(t-1)+2s/\delta.

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 GG be a bipartite graph with bipartition (A,B)(A,B), where |A|=n|B|=m|A|=n\leq|B|=m and the vertices of GG are labeled by positive integers. Suppose that for each XAX\subseteq A and YBY\subseteq B (and for each XBX\subseteq B and YAY\subseteq A) with |X|=r|X|=r and |Y|=t|Y|=t such that maxxXx<minyYy\max_{x\in X}x<\min_{y\in Y}y, the induced subgraph G[XY]G[X\cup Y] is not complete bipartite. Then the number of edges of GG is at most 2(t1)1/rmn11/r+2(r1)m2(t-1)^{1/r}mn^{1-1/r}+2(r-1)m.

Lemma 2.5 ([25, Lemma 2.4]).

Let GG be a graph with nn vertices, with the edge colored by kk colors. Suppose that GG does not contain a cycle through vertices v1,v2,v3,v4v_{1},v_{2},v_{3},v_{4} such that the edges v1v2v_{1}v_{2} and v4v1v_{4}v_{1} have the same color, and the edges v2v3v_{2}v_{3} and v3v4v_{3}v_{4} have the same color. Then the number of edges of GG is at most k1/2n3/2+knk^{1/2}n^{3/2}+kn.

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 NN\in\mathbb{N} and A{1,2,,N}A\subseteq\{1,2,\ldots,N\}. Let 𝒫{\mathcal{P}} be a set of primes. For each prime p𝒫p\in{\mathcal{P}}, let Ap=A(modp)A_{p}=A\pmod{p}. For any 1<QN1<Q\leq N, we have

|A|pQ,p𝒫logplogNpQ,p𝒫logp|Ap|logN,|A|\leq\frac{\underset{p\leq Q,p\in\mathcal{P}}{\sum}\log p-\log N}{\underset{p\leq Q,p\in\mathcal{P}}{\sum}\frac{\log p}{|A_{p}|}-\log N},

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 NN\in\mathbb{N} and A{1,2,,N}A\subseteq\{1,2,\ldots,N\}. Let 𝒫[2,Q]{\mathcal{P}}\subseteq[2,Q] be a set of primes. For each prime p𝒫p\in{\mathcal{P}}, let Ap=A(modp)A_{p}=A\pmod{p}. For any 1<QN1<Q\leq N, we have

|A|max{Q,23Nexp(p𝒫logpp)exp(p𝒫logp|Ap|)}.|A|\leq\max\biggl{\{}Q,\frac{23N\exp\big{(}\sum_{p\in\mathcal{P}}\frac{\log p}{p}\big{)}}{\exp\big{(}\sum_{p\in\mathcal{P}}\frac{\log p}{|A_{p}|}\big{)}}\biggr{\}}.

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 q0q_{0} and LL such that whenever qq0q\geq q_{0} is an integer, xqLx\geq q^{L}, and aa is an integer with gcd(a,q)=1\gcd(a,q)=1, then

pxpa(modq)logpxϕ(q)q,\sum_{\begin{subarray}{c}p\leq x\\ p\equiv a\pmod{q}\end{subarray}}\log p\gg\frac{x}{\phi(q)\sqrt{q}},

where the implied constant is absolute.

Combining Lemma 2.8 with the prime number theorem for arithmetic progressions with modulus qq0q\leq q_{0} bounded, we obtain the following corollary.

Corollary 2.9.

There exists positive constants cc and LL such that whenever qq is a positive integer, xqLx\geq q^{L}, and aa is an integer with gcd(a,q)=1\gcd(a,q)=1, then

pxpa(modq)logpcxϕ(q)q.\sum_{\begin{subarray}{c}p\leq x\\ p\equiv a\pmod{q}\end{subarray}}\log p\geq\frac{cx}{\phi(q)\sqrt{q}}.

The following auxiliary estimate will be useful.

Lemma 2.10.

Let 𝒫\mathcal{P} be a finite set of primes. Then we have p𝒫logpplog|𝒫|\sum_{p\in\mathcal{P}}\frac{\log p}{p}\ll\log|\mathcal{P}|.

Proof.

Let 𝒫={q1,q2,,qm}\mathcal{P}=\{q_{1},q_{2},\ldots,q_{m}\}, and let p1<p2<<pmp_{1}<p_{2}<\ldots<p_{m} be the first mm odd primes. Since the function logxx\frac{\log x}{x} is decreasing when x3x\geq 3, it follows that

p𝒫logpp=j=1mlogqjqjlog22+ppmlogpplogpmlog(mlogm)log|𝒫|.\sum_{p\in\mathcal{P}}\frac{\log p}{p}=\sum_{j=1}^{m}\frac{\log q_{j}}{q_{j}}\leq\frac{\log 2}{2}+\sum_{p\leq p_{m}}\frac{\log p}{p}\ll\log p_{m}\ll\log(m\log m)\ll\log|\mathcal{P}|.\qed

3. Contribution of large elements

In this section, we deduce upper bounds on the number of “large” elements in Diophantine tuples with property Dd(n)D_{\leq d}(n). 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 Dk(n)D_{k}(n).

The following lemma is due to Dujella [14].

Lemma 3.1 (Dujella).

Let nn be a nonzero integer. If A[|n|3,+)A\subseteq[|n|^{3},+\infty) is a Diophantine tuple with property D2(n)D_{2}(n), then |A|21|A|\leq 21.

The following lemma is a special case of [41, Proposition 4.1] due to the second author.

Lemma 3.2 ([41]).

Let k3k\geq 3 and nn be a nonzero integer. Let =3\ell=3 if k=3k=3, and =2\ell=2 if k4k\geq 4. There do not exist integers a1,a2,,aa_{1},a_{2},\ldots,a_{\ell} and b1,b2,,bsk+1b_{1},b_{2},\ldots,b_{s_{k}+1}, such that a1<a2<<ab1<b2<<bsk+1a_{1}<a_{2}<\ldots<a_{\ell}\leq b_{1}<b_{2}<\ldots<b_{s_{k}+1}, 2|n|tkb12|n|^{t_{k}}\leq b_{1}, and aibj+na_{i}b_{j}+n is a perfect kk-th power for all 1i1\leq i\leq\ell and 1jsk+11\leq j\leq s_{k}+1.

Lemma 3.3 ([41, Proposition 4.3]).

Let k3k\geq 3 and let nn be a nonzero integer. (A,B)(A,B) be a bipartite Diophantine tuple with property BDk(n)BD_{k}(n). If min{|A|,|B|}rk\min\{|A|,|B|\}\geq r_{k}, then in both sets AA and BB, at most sks_{k} elements are at least 2|n|tk2|n|^{t_{k}}.

Next, we use 3.2 and Lemma 2.4 to deduce the following proposition.

Proposition 3.4.

Let k3k\geq 3 and let nn be a nonzero integer. Let A,BA,B be two finite sets of positive integers in [2|n|17,)[2|n|^{17},\infty) with |A||B||A|\leq|B|. Then the number of pairs (a,b)(a,b) such that aAa\in A, bBb\in B and ab+nab+n is is a perfect kk-th power is at most 7|B||A|1/27|B||A|^{1/2} when k4k\geq 4, and at most 8|B||A|2/38|B||A|^{2/3} when k=3k=3.

Proof.

Note that tk<17t_{k}<17 and sk6s_{k}\leq 6 for all k3k\geq 3. We first consider the case k4k\geq 4. In this case, the corollary follows immediately from Lemma 2.4 and 3.2 by building a bipartite graph with bipartition (A,B)(A,B) such that aAa\in A and bBb\in B are adjacent if and only if ab+nab+n is a perfect kk-th power and the inequality 26|B||A|1/2+2|B|7|B||A|1/22\sqrt{6}|B||A|^{1/2}+2|B|\leq 7|B||A|^{1/2}. The proof of the case k=3k=3 is similar; Lemma 2.4 implies that the number of edges is at most 261/3|B||A|2/3+4|B|<8|B||A|2/32\cdot 6^{1/3}|B||A|^{2/3}+4|B|<8|B||A|^{2/3}. ∎

Corollary 3.5.

Let k2k\geq 2 and let nn be a nonzero integer. If AA a finite set of positive integers in [2|n|17,)[2|n|^{17},\infty), then the number of pairs {a,b}\{a,b\} such that a,bAa,b\in A with aba\neq b and ab+nab+n is is a perfect kk-th power is at most 7|A|3/27|A|^{3/2} when k4k\geq 4, at most 8|A|5/38|A|^{5/3} when k=3k=3, and at most 10|A|2/2110|A|^{2}/21 when k=2k=2.

Proof.

When k3k\geq 3, the corollary follows from Proposition 3.4 by setting B=AB=A.

It remains to prove the case k=2k=2. Let GG be the graph with vertex set AA such that two distinct vertices a,ba,b are adjacent if and only if ab+nab+n is a square. By Lemma 3.1, GG does not contain K22K_{22} as a subgraph. It follows from Lemma 2.1 that the number of edges of GG is at most 10|A|2/2110|A|^{2}/21. ∎

The next proposition bounds the number of large elements in a Diophantine tuple with property Dd(n)D_{\leq d}(n), with dd finite.

Proposition 3.6.

Let n,dn,d be integers with n0n\neq 0 and d2d\geq 2. If A[2|n|17,)A\subseteq[2|n|^{17},\infty) is a Diophantine tuple with property Dd(n)D_{\leq d}(n), then |A|(d/logd)2|A|\ll(d/\log d)^{2}, where the implied constant is absolute.

Proof.

Since AA satisfies property Dd(n)D_{\leq d}(n), for each a,bAa,b\in A with aba\neq b, there is a prime pdp\leq d such that ab+nab+n is a perfect pp-th power. It follows from Corollary 3.5 that

(|A|2)10|A|221+8|A|5/3+7|A|3/2(π(d)2).\binom{|A|}{2}\leq\frac{10|A|^{2}}{21}+8|A|^{5/3}+7|A|^{3/2}\cdot(\pi(d)-2).

It follows that |A|(π(d))2|A|\ll(\pi(d))^{2}, as required. ∎

We also need to bound the number of large elements in a Diophantine tuple with property D(n)D_{\leq\infty}(n). 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 α=a/b\alpha=a/b, where a,ba,b are coprime integers, its height H(α)H(\alpha) is defined to max{|a|,|b|}\max\{|a|,|b|\}.

Lemma 3.7 ([1]).

Let b1b_{1} and b2b_{2} be non-zero integers and let α1\alpha_{1} and α2\alpha_{2} be positive rational numbers. Put Ai=max{2,H(αi)}A_{i}=\max\{2,H(\alpha_{i})\} for i{1,2}i\in\{1,2\}, B=max{|b1|,|b2|,2}B=\max\{|b_{1}|,|b_{2}|,2\}, and Λ=b1logα1+b2logα2\Lambda=b_{1}\log\alpha_{1}+b_{2}\log\alpha_{2}. Then there exists an effectively computable positive constant CC such that if Λ0\Lambda\neq 0, then

log|Λ|>ClogA1logA2logB.\log|\Lambda|>-C\log A_{1}\cdot\log A_{2}\cdot\log B.
Proposition 3.8.

Let nn be a nonzero integer and let M(4|n|)17M\geq(4|n|)^{17}. There is an absolute constant CC^{\prime} such that if a1,a2,a3,a4a_{1},a_{2},a_{3},a_{4} are distinct integers in [M,M][\sqrt{M},M], then there do not exist two primes p1,p2(ClogMloglogM)1/2p_{1},p_{2}\geq(C^{\prime}\log M\log\log M)^{1/2} such that there exist positive integers x1,x2,x3,x4x_{1},x_{2},x_{3},x_{4} with

a1a2+n=x1p1,a2a3+n=x2p2,a3a4+n=x3p2,a4a1+n=x4p1.a_{1}a_{2}+n=x_{1}^{p_{1}},\quad a_{2}a_{3}+n=x_{2}^{p_{2}},\quad a_{3}a_{4}+n=x_{3}^{p_{2}},\quad a_{4}a_{1}+n=x_{4}^{p_{1}}. (3.1)
Proof.

Let CC be the constant from Lemma 3.7. We show that the proposition is true with C=max{1,21C}C^{\prime}=\max\{1,21C\}.

For the sake of contradiction, suppose that there exist primes p1,p2(ClogMloglogM)1/2p_{1},p_{2}\geq(C^{\prime}\log M\log\log M)^{1/2} and positive integers x1,x2,x3,x4x_{1},x_{2},x_{3},x_{4} satisfying equation (3.1). Let t=min{p1,p2}t=\min\{p_{1},p_{2}\}. Without loss of generality, assume that x1p1=max{x1p1,x2p2,x3p2,x4p1}x_{1}^{p_{1}}=\max\{x_{1}^{p_{1}},x_{2}^{p_{2}},x_{3}^{p_{2}},x_{4}^{p_{1}}\}. Note that x3p2=a3a4+nM|n|>M/2x_{3}^{p_{2}}=a_{3}a_{4}+n\geq M-|n|>M/2. Also, we have max{x1,x2,x3,x4)(M2+|n|)1/tM3/t,andmax{p1,p2}log2(M2+|n|)3log2M<4.5logM.\max\{x_{1},x_{2},x_{3},x_{4})\leq(M^{2}+|n|)^{1/t}\leq M^{3/t},\text{and}\max\{p_{1},p_{2}\}\leq\log_{2}(M^{2}+|n|)\leq 3\log_{2}M<4.5\log M.

Observe that

(x1p1n)(x3p2n)=a1a2a3a4=(x2p2n)(x4p1n).(x_{1}^{p_{1}}-n)(x_{3}^{p_{2}}-n)=a_{1}a_{2}a_{3}a_{4}=(x_{2}^{p_{2}}-n)(x_{4}^{p_{1}}-n).

It follows that

x1p1x3p2x2p2x4p1\displaystyle x_{1}^{p_{1}}x_{3}^{p_{2}}-x_{2}^{p_{2}}x_{4}^{p_{1}} =n(x1p1+x3p2x2p2x4p1)\displaystyle=n(x_{1}^{p_{1}}+x_{3}^{p_{2}}-x_{2}^{p_{2}}-x_{4}^{p_{1}})
=n(a1a2+a3a4a2a3a4a1)=n(a1a3)(a2a4)0.\displaystyle=n(a_{1}a_{2}+a_{3}a_{4}-a_{2}a_{3}-a_{4}a_{1})=n(a_{1}-a_{3})(a_{2}-a_{4})\neq 0.

Thus,

0<|x2p2x4p1x1p1x3p21|2|n|x1p1x1p1x3p22|n|x3p24|n|MM1/17M=M1617,0<\bigg{|}\frac{x_{2}^{p_{2}}x_{4}^{p_{1}}}{x_{1}^{p_{1}}x_{3}^{p_{2}}}-1\bigg{|}\leq\frac{2|n|x_{1}^{p_{1}}}{x_{1}^{p_{1}}x_{3}^{p_{2}}}\leq\frac{2|n|}{x_{3}^{p_{2}}}\leq\frac{4|n|}{M}\leq\frac{M^{1/17}}{M}=M^{-\frac{16}{17}},

where we used the assumption that M(4|n|)17M\geq(4|n|)^{17}.

Set

Λ=logx2p2x4p1x1p1x3p2=p1log(x4x1)+p2log(x2x3);\Lambda=\log\frac{x_{2}^{p_{2}}x_{4}^{p_{1}}}{x_{1}^{p_{1}}x_{3}^{p_{2}}}=p_{1}\log\bigg{(}\frac{x_{4}}{x_{1}}\bigg{)}+p_{2}\log\bigg{(}\frac{x_{2}}{x_{3}}\bigg{)};

then Λ0\Lambda\neq 0 and |eΛ1|M16/17416|e^{\Lambda}-1|\leq M^{-16/17}\leq 4^{-16}. It is easy to verify that if zz is a real number such that |ez1|<1/8|e^{z}-1|<1/8, then |z|<1/2|z|<1/2; and if |z|<1/2|z|<1/2, then |ez1||z|/2|e^{z}-1|\geq|z|/2. It follows that |Λ|2M16/17|\Lambda|\leq 2M^{-16/17} and thus

log|Λ|log21617logM<1517logM.\log|\Lambda|\leq\log 2-\frac{16}{17}\log M<-\frac{15}{17}\log M.

On the other hand, we can apply Lemma 3.7 with α1=x4/x1\alpha_{1}=x_{4}/x_{1}, α2=x2/x3\alpha_{2}=x_{2}/x_{3}, b1=p1b_{1}=p_{1}, b2=p2b_{2}=p_{2} to get a lower bound on log|Λ|\log|\Lambda|. Note that we have logA=max{logH(α1),logH(α2)}(3logM)/t\log A=\max\{\log H(\alpha_{1}),\log H(\alpha_{2})\}\leq(3\log M)/t and logB=logmax{p1,p2}<log(4.5logM)<2loglogM\log B=\log\max\{p_{1},p_{2}\}<\log(4.5\log M)<2\log\log M. Then Lemma 3.7 implies that

1517logM>log|Λ|>C9log2Mt2(2loglogM).-\frac{15}{17}\log M>\log|\Lambda|>-C\frac{9\log^{2}M}{t^{2}}(2\log\log M).

It follows that

t<21ClogMloglogM,t<\sqrt{21C\log M\log\log M},

which contradicts the assumption that tClogMloglogMt\geq\sqrt{C^{\prime}\log M\log\log M} since C21CC^{\prime}\geq 21C. ∎

Now we are ready to prove a key result of the section.

Proposition 3.9.

Let nn be a nonzero integer and let N4|n|17N\geq 4|n|^{17}. Let A[(4|n|)17,N]A\subseteq[(4|n|)^{17},N] be a Diophantine tuple with property D(n)D_{\leq\infty}(n), then |A|logN|A|\ll\log N, where the implied constant is absolute.

Proof.

Let M1=4|n|17M_{1}=4|n|^{17}, and inductively we define Mi=Mi12M_{i}=M_{i-1}^{2} for each integer i2i\geq 2. Let Ai=A[Mi,Mi+1]A_{i}=A\cap[M_{i},M_{i+1}] for each positive integer ii; then we have A=i=1AiA=\bigcup_{i=1}^{\ell}A_{i}, where loglogN\ell\ll\log\log N. Thus, it suffices to show that |Ai|logNloglogN|A_{i}|\ll\frac{\log N}{\log\log N} for each ii.

Fix an integer ii with 1i1\leq i\leq\ell. Since AiA_{i} is a Diophantine tuple with property D(n)D_{\leq\infty}(n), for each pair {a,b}Ai\{a,b\}\subseteq A_{i} with aba\neq b, ab+nab+n is a perfect pp-th power for some prime pp; however, since ab+nN2+n<4N2ab+n\leq N^{2}+n<4N^{2}, it is necessary that plog2(4N2)4logNp\leq\log_{2}(4N^{2})\leq 4\log N. Let CC^{\prime} be the constant from 3.8. By Corollary 3.5, the number of such pairs corresponding to primes pt:=(ClogNloglogN)1/2p\leq t:=(C^{\prime}\log N\log\log N)^{1/2} is at most

10|Ai|221+8|Ai|5/3+7|A|3/2π(t).\frac{10|A_{i}|^{2}}{21}+8|A_{i}|^{5/3}+7|A|^{3/2}\pi(t).

Next, we give an upper bound on the number of pairs corresponding to primes t<p4logNt<p\leq 4\log N using Lemma 2.5. Build a graph GiG_{i} with vertex set AiA_{i}. If a,bAia,b\in A_{i} with aba\neq b such that ab+nab+n is a pp-th power for some prime p(t,4logN]p\in(t,4\log N]; then there is an edge between aa and bb and we color it using the smallest such pp. Then the edges of GiG_{i} are colored by at most π(4logN)\pi(4\log N) colors. By Proposition 3.8, GiG_{i} does not contain a cycle through vertices v1,v2,v3,v4v_{1},v_{2},v_{3},v_{4} such that the edges v1v2v_{1}v_{2} and v4v1v_{4}v_{1} have the same color, and the edges v2v3v_{2}v_{3} and v3v4v_{3}v_{4} have the same color. Thus, Lemma 2.5 implies that the number of edges in GiG_{i} is at most

π(4logN)|Ai|3/2+π(4logN)|Ai|.\sqrt{\pi(4\log N)}|A_{i}|^{3/2}+\pi(4\log N)|A_{i}|.

It follows that

(|Ai|2)10|Ai|221+8|Ai|5/3+7|A|3/2π(t)+π(4logN)|Ai|3/2+π(4logN)|Ai|.\binom{|A_{i}|}{2}\leq\frac{10|A_{i}|^{2}}{21}+8|A_{i}|^{5/3}+7|A|^{3/2}\pi(t)+\sqrt{\pi(4\log N)}|A_{i}|^{3/2}+\pi(4\log N)|A_{i}|.

It follows that |Ai|π(t)2+π(4logN)logNloglogN|A_{i}|\ll\pi(t)^{2}+\pi(4\log N)\ll\frac{\log N}{\log\log N}, 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 𝔽p\mathbb{F}_{p} be the finite field with pp elements, and 𝔽p=𝔽p{0}\mathbb{F}_{p}^{*}=\mathbb{F}_{p}\setminus\{0\}.

4.1. Diophantine tuples with property Dd(n)D_{\leq d}(n)

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 χ\chi be a non-trivial multiplicative character of 𝔽p\mathbb{F}_{p} and λ𝔽p\lambda\in\mathbb{F}_{p}^{*}. For any A,B𝔽pA,B\subseteq\mathbb{F}_{p}^{*}, we have

|aA,bBχ(ab+λ)|p|A||B|.\bigg{|}\sum_{a\in A,\,b\in B}\chi(ab+\lambda)\bigg{|}\leq\sqrt{p|A||B|}.

Next, we use Lemma 4.1 to deduce an upper bound on the finite field analogue of a Diophantine tuple with property Dd(n)D_{\leq d}(n).

Proposition 4.2.

Let p1,p2,,pmp_{1},p_{2},\ldots,p_{m} be distinct primes and let pp be a prime such that p1(modpi)p\equiv 1\pmod{p_{i}} for each 1im1\leq i\leq m. Let λ𝔽p\lambda\in\mathbb{F}_{p}^{*}. For each 1im1\leq i\leq m, let Hi={xpi:x𝔽p}H_{i}=\{x^{p_{i}}:x\in\mathbb{F}_{p}\}. If AA a subset of 𝔽p\mathbb{F}_{p}^{*} such that ab+λi=1mHiab+\lambda\in\bigcup_{i=1}^{m}H_{i} for all a,bAa,b\in A with aba\neq b, then

|A|(2mp+2)i=1m(11pi)1.|A|\leq(2^{m}\sqrt{p}+2)\cdot\prod_{i=1}^{m}\bigg{(}1-\frac{1}{p_{i}}\bigg{)}^{-1}.
Proof.

For a set S𝔽pS\subseteq\mathbb{F}_{p}, we use SS^{*} to denote S{0}S\setminus\{0\}, and 𝟏S\mathbf{1}_{S} to denote the indicator function of the set SS. For each 1im1\leq i\leq m, let χpi\chi_{p_{i}} be a multiplicative character of 𝔽p\mathbb{F}_{p} of order (p1)/pi(p-1)/p_{i}; then by the orthogonality relation, we have

𝟏Hi=1piji=0pi1χpiji.\mathbf{1}_{H_{i}^{*}}=\frac{1}{p_{i}}\sum_{j_{i}=0}^{p_{i}-1}\chi_{p_{i}}^{j_{i}}. (4.1)

Let H=i=1mHiH=\bigcup_{i=1}^{m}H_{i}. Since ab+λHab+\lambda\in H for all a,bAa,b\in A with aba\neq b, and for each aAa\in A, there is at most one bAb\in A such that ab+λ=0ab+\lambda=0, it follows that

a,bA𝟏H(ab+λ)a,bA𝟏H(ab+λ)|A||A|22|A|.\sum_{a,b\in A}\mathbf{1}_{H^{*}}(ab+\lambda)\geq\sum_{a,b\in A}\mathbf{1}_{H}(ab+\lambda)-|A|\geq|A|^{2}-2|A|. (4.2)

On the other hand, by equation (4.1), we have

𝟏H=1i=1m(1𝟏Hi)=1i=1m(11pij=0pi1χpij)=I[m]|I|iI(1piji=0pi1χpiji).\mathbf{1}_{H^{*}}=1-\prod_{i=1}^{m}(1-\mathbf{1}_{H_{i}^{*}})=1-\prod_{i=1}^{m}\bigg{(}1-\frac{1}{p_{i}}\sum_{j=0}^{p_{i}-1}\chi_{p_{i}}^{j}\bigg{)}=-\sum_{\begin{subarray}{c}I\subseteq[m]\\ |I|\neq\emptyset\end{subarray}}\prod_{i\in I}\bigg{(}-\frac{1}{p_{i}}\sum_{j_{i}=0}^{p_{i}-1}\chi_{p_{i}}^{j_{i}}\bigg{)}. (4.3)

Fix a nonempty subset II of [m][m]. Let |I|=k|I|=k and write {pi:iI}={q1,q2,,qk}\{p_{i}:i\in I\}=\{q_{1},q_{2},\ldots,q_{k}\}. Then we have

iI(ji=0pi1χpiji)=j1=0q11jk=0qk1i=1kχqiji;\prod_{i\in I}\bigg{(}\sum_{j_{i}=0}^{p_{i}-1}\chi_{p_{i}}^{j_{i}}\bigg{)}=\sum_{j_{1}=0}^{q_{1}-1}\cdots\sum_{j_{k}=0}^{q_{k}-1}\prod_{i=1}^{k}\chi_{q_{i}}^{j_{i}}; (4.4)

note that the character i=1kχqiji\prod_{i=1}^{k}\chi_{q_{i}}^{j_{i}} is trivial if and only if j1=j2==0j_{1}=j_{2}=\cdots=0. It then follows from Lemma 4.1 that

a,bAiI(ji=0pi1χpiji(ab+λ))=j1=0q11jk=0qk1a,bA(i=1kχqiji)(ab+λ)=|A|2+E(I),\sum_{a,b\in A}\prod_{i\in I}\bigg{(}\sum_{j_{i}=0}^{p_{i}-1}\chi_{p_{i}}^{j_{i}}(ab+\lambda)\bigg{)}=\sum_{j_{1}=0}^{q_{1}-1}\cdots\sum_{j_{k}=0}^{q_{k}-1}\sum_{a,b\in A}\bigg{(}\prod_{i=1}^{k}\chi_{q_{i}}^{j_{i}}\bigg{)}(ab+\lambda)=|A|^{2}+E(I), (4.5)

where |E(I)|iIpip|A|.|E(I)|\leq\prod_{i\in I}p_{i}\cdot\sqrt{p}|A|.

Combining equations (4.3), (4.4), and (4.5),

a,bA𝟏H(ab+λ)\displaystyle\sum_{a,b\in A}\mathbf{1}_{H^{*}}(ab+\lambda) =I[m]Ia,bAiI(1piji=0pi1χpiji(ab+λ))\displaystyle=-\sum_{\begin{subarray}{c}I\subseteq[m]\\ I\neq\emptyset\end{subarray}}\sum_{a,b\in A}\prod_{i\in I}\bigg{(}-\frac{1}{p_{i}}\sum_{j_{i}=0}^{p_{i}-1}\chi_{p_{i}}^{j_{i}}(ab+\lambda)\bigg{)}
=I[m]I(|A|2iI(pi)+E(I)iI(pi))\displaystyle=-\sum_{\begin{subarray}{c}I\subseteq[m]\\ I\neq\emptyset\end{subarray}}\bigg{(}\frac{|A|^{2}}{\prod_{i\in I}(-p_{i})}+\frac{E(I)}{\prod_{i\in I}(-p_{i})}\bigg{)}
(1i=1m(11pi))|A|2+2mp|A|.\displaystyle\leq\bigg{(}1-\prod_{i=1}^{m}\bigg{(}1-\frac{1}{p_{i}}\bigg{)}\bigg{)}|A|^{2}+2^{m}\sqrt{p}|A|. (4.6)

Comparing inequalities (4.2) and (4.6), we get

(2mp+2)|A|i=1m(11pi)|A|2,(2^{m}\sqrt{p}+2)|A|\geq\prod_{i=1}^{m}\bigg{(}1-\frac{1}{p_{i}}\bigg{)}\cdot|A|^{2},

as required. ∎

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 d2d\geq 2 and pp be a prime such that p1(modd)p\equiv 1\pmod{d}. Let A𝔽pA\subseteq\mathbb{F}_{p}^{*}, B𝔽pB\subseteq\mathbb{F}_{p}, and λ𝔽p\lambda\in\mathbb{F}_{p}^{*}, such that ab+λ{xd:x𝔽p}ab+\lambda\in\{x^{d}:x\in\mathbb{F}_{p}\} for all aAa\in A and bBb\in B. If |A|=m|A|=m, then |B|pdm+mp|B|\leq\frac{p}{d^{m}}+m\sqrt{p}.

Proof.

Let χ\chi be a multiplicative character of 𝔽p\mathbb{F}_{p} with order dd. Let A={a1,a2,,am}A=\{a_{1},a_{2},\ldots,a_{m}\}. Note that for each bBb\in B and each 1im1\leq i\leq m, we have χ(aib+λ)=1\chi(a_{i}b+\lambda)=1 and thus χ(b+ai1λ)=χ(ai)¯\chi(b+a_{i}^{-1}\lambda)=\overline{\chi(a_{i})}. In particular, each bBb\in B is a solution to the system of equations

χ(x+ai1λ)=χ(ai)¯,1im.\chi(x+a_{i}^{-1}\lambda)=\overline{\chi(a_{i})},\quad\forall 1\leq i\leq m.

Now a classical application of Weil’s bound (see for example [35, Exercise 5.66]) implies that

|B|pdm+(m1md+1dm)p+md<pdm+mp,|B|\leq\frac{p}{d^{m}}+\bigg{(}m-1-\frac{m}{d}+\frac{1}{d^{m}}\bigg{)}\sqrt{p}+\frac{m}{d}<\frac{p}{d^{m}}+m\sqrt{p},

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 ν\nu for the upper bound below. Karatsuba’s estimate is better than Vinogradov’s estimate when AA and BB are asymmetric in the sense that the sizes of AA and BB are not comparable.

Lemma 4.4.

Let A𝔽p,B𝔽pA\subseteq\mathbb{F}_{p},B\subseteq\mathbb{F}_{p}^{*}, and λ𝔽p\lambda\in\mathbb{F}_{p}^{*}. Then for any non-trivial multiplicative character χ\chi of 𝔽p\mathbb{F}_{p} and any positive integer ν\nu, we have

|aAbBχ(ab+λ)||B|(2ν1)/2ν(2ν|A|2νp+(2ν)ν|A|νp)1/2ν.\bigg{|}\sum_{\begin{subarray}{c}a\in A\\ b\in B\end{subarray}}\chi(ab+\lambda)\bigg{|}\leq|B|^{(2\nu-1)/2\nu}\big{(}2\nu|A|^{2\nu}\sqrt{p}+(2\nu)^{\nu}|A|^{\nu}p\big{)}^{1/2\nu}.
Proof.

By Hölder’s inequality, we have

|aAbBχ(ab+λ)|\displaystyle\bigg{|}\sum_{\begin{subarray}{c}a\in A\\ b\in B\end{subarray}}\chi(ab+\lambda)\bigg{|} =|bBχ(b)aAχ(a+b1λ)|\displaystyle=\bigg{|}\sum_{b\in B}\chi(b)\sum_{a\in A}\chi(a+b^{-1}\lambda)\bigg{|}
(bB|χ(b)|2ν/(2ν1))(2ν1)/2ν(bB|aAχ(a+b1λ)|2ν)1/2ν\displaystyle\leq\bigg{(}\sum_{b\in B}|\chi(b)|^{2\nu/(2\nu-1)}\bigg{)}^{(2\nu-1)/2\nu}\bigg{(}\sum_{b\in B}\bigg{|}\sum_{a\in A}\chi(a+b^{-1}\lambda)\bigg{|}^{2\nu}\bigg{)}^{1/2\nu}
|B|(2ν1)/2ν(x𝔽p|aAχ(a+x)|2ν)1/2ν.\displaystyle\leq|B|^{(2\nu-1)/2\nu}\bigg{(}\sum_{x\in\mathbb{F}_{p}}\bigg{|}\sum_{a\in A}\chi(a+x)\bigg{|}^{2\nu}\bigg{)}^{1/2\nu}.

Note that

x𝔽p|aAχ(a+x)|2ν=a1,a2,,a2νA(x𝔽pχ(j=1ν(x+aj))χ¯(k=ν+12ν(x+ak))),\displaystyle\sum_{x\in\mathbb{F}_{p}}\bigg{|}\sum_{a\in A}\chi(a+x)\bigg{|}^{2\nu}=\sum_{a_{1},a_{2},\ldots,a_{2\nu}\in A}\bigg{(}\sum_{x\in\mathbb{F}_{p}}\chi\bigg{(}\prod_{j=1}^{\nu}(x+a_{j})\bigg{)}\cdot\overline{\chi}\bigg{(}\prod_{k=\nu+1}^{2\nu}(x+a_{k})\bigg{)}\bigg{)}, (4.7)

where the sum on the right-hand side runs over all 2ν2\nu-triples (a1,a2,,a2ν)A2ν(a_{1},a_{2},\ldots,a_{2\nu})\in A^{2\nu}. If there exists ii with aiaja_{i}\neq a_{j} for each jij\neq i, then a standard application of Weil’s bound implies that the second summand on the right-hand side of equation (4.7) is 2νp2\nu\sqrt{p} (see for example [30, Corollary 11.24]); otherwise the summand is trivially bounded by pp. Note that in the latter case, each aia_{i} appears at least twice, so the number of such tuples (a1,a2,,a2ν)A2ν(a_{1},a_{2},\ldots,a_{2\nu})\in A^{2\nu} is at most (2νν)ν!|A|ν=(2ν)!ν!|A|ν\binom{2\nu}{\nu}\nu!|A|^{\nu}=\frac{(2\nu)!}{\nu!}|A|^{\nu}. Therefore,

x𝔽p|aAχ(a+x)|2ν2ν|A|2νp+(2ν)!ν!|A|νp<2ν|A|2νp+(2ν)ν|A|νp,\sum_{x\in\mathbb{F}_{p}}\bigg{|}\sum_{a\in A}\chi(a+x)\bigg{|}^{2\nu}\leq 2\nu|A|^{2\nu}\sqrt{p}+\frac{(2\nu)!}{\nu!}|A|^{\nu}p<2\nu|A|^{2\nu}\sqrt{p}+(2\nu)^{\nu}|A|^{\nu}p,

and the required estimate follows. ∎

Corollary 4.5.

Let k2k\geq 2 be an integer and pp be a prime such that gcd(k,p1)>1\gcd(k,p-1)>1. Let A,B𝔽pA,B\subseteq\mathbb{F}_{p} and λ𝔽p\lambda\in\mathbb{F}_{p}^{*}, such that ab+λ{xk:x𝔽p}ab+\lambda\in\{x^{k}:x\in\mathbb{F}_{p}\} for all aAa\in A and bBb\in B. If ν\nu is a positive integer such that |A|2νp1/2ν|A|\geq 2\nu p^{1/2\nu}, then |B|12νp|B|\leq 12\nu\sqrt{p}.

Proof.

Note that {xk:x𝔽p}={xgcd(k,p1):x𝔽p}\{x^{k}:x\in\mathbb{F}_{p}\}=\{x^{\gcd(k,p-1)}:x\in\mathbb{F}_{p}\}. Let χ\chi be a multiplicative character of 𝔽p\mathbb{F}_{p} with order gcd(k,p1)\gcd(k,p-1) and let B=B{0}𝔽pB^{\prime}=B\setminus\{0\}\subseteq\mathbb{F}_{p}^{*}. Note that for each bBb\in B^{\prime}, there is at most one aAa\in A such that ab+λ=0ab+\lambda=0. Thus, Lemma 4.4 implies that

|B|(|A|1)aAbBχ(ab+λ)|B|(2ν1)/2ν(2ν|A|2νp+(2ν)ν|A|νp)1/2ν.|B^{\prime}|(|A|-1)\leq\sum_{\begin{subarray}{c}a\in A\\ b\in B^{\prime}\end{subarray}}\chi(ab+\lambda)\leq|B^{\prime}|^{(2\nu-1)/2\nu}(2\nu|A|^{2\nu}\sqrt{p}+(2\nu)^{\nu}|A|^{\nu}p)^{1/2\nu}. (4.8)

Since |A|2νp1/2ν|A|\geq 2\nu p^{1/2\nu}, it follows that 2ν|A|2νp>(2ν)ν|A|νp2\nu|A|^{2\nu}\sqrt{p}>(2\nu)^{\nu}|A|^{\nu}p. Therefore, inequality (4.8) implies

|B|(|A|1)2ν4ν|A|2νp.|B^{\prime}|(|A|-1)^{2\nu}\leq 4\nu|A|^{2\nu}\sqrt{p}.

Since |A|2ν+1|A|\geq 2\nu+1, it follows that

|B|4νp(|A||A|1)2ν4νp(1+12ν)2ν<4eνp<12νp1.|B^{\prime}|\leq 4\nu\sqrt{p}\cdot\bigg{(}\frac{|A|}{|A|-1}\bigg{)}^{2\nu}\leq 4\nu\sqrt{p}\cdot\bigg{(}1+\frac{1}{2\nu}\bigg{)}^{2\nu}<4e\nu\sqrt{p}<12\nu\sqrt{p}-1.

It follows that |B||B|+112νp|B|\leq|B^{\prime}|+1\leq 12\nu\sqrt{p}, as required. ∎

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 N=2|n|tkN=2|n|^{t_{k}}. Let A,BA,B\subseteq\mathbb{N} such that AB+n{xk:x}AB+n\subseteq\{x^{k}:x\in\mathbb{N}\} with |B||A|m+sk|B|\geq|A|\geq m+s_{k}. Then |B||A|m+sk(rksk)+sk=rk|B|\geq|A|\geq m+s_{k}\geq(r_{k}-s_{k})+s_{k}=r_{k}. Let A=A[1,N]A^{\prime}=A\cap[1,N] and B=B[1,N]B^{\prime}=B\cap[1,N]. It follows from Lemma 3.3 that |A||A|skm|A^{\prime}|\geq|A|-s_{k}\geq m and |B||B|sk|B|3|B^{\prime}|\geq|B|-s_{k}\geq|B|-3. We replace AA^{\prime} with a subset of AA^{\prime} of size mm if |A|>m|A^{\prime}|>m.

Consider the set of primes 𝒫={p:pQ}\mathcal{P}=\{p:p\leq Q\}, where Q=Nϕ(k)/θk,m,Q=N^{\phi(k)/\theta_{k,m}}, and

θk,m=1ikgcd(i,k)=1gcd(i1,k)m\theta_{k,m}=\sum_{\begin{subarray}{c}1\leq i\leq k\\ \gcd(i,k)=1\end{subarray}}\gcd(i-1,k)^{m} (5.1)

has been defined in equation (1.2). Note that ϕ(k)<kmθk,mkm+1\phi(k)<k^{m}\leq\theta_{k,m}\leq k^{m+1}. Since m=o(loglog|n|)m=o(\log\log|n|), as |n||n|\to\infty, we have m2k3m+1=o(log|n|)m^{2}k^{3m+1}=o(\log|n|), and thus

mk2m=o(logNkm+1)=o(logQ).mk^{2m}=o\bigg{(}\frac{\log N}{k^{m+1}}\bigg{)}=o(\log Q). (5.2)

For each p𝒫p\in\mathcal{P}, let ApA_{p} be the image of AA^{\prime} modulo pp and view ApA_{p} as a subset of the finite field 𝔽p\mathbb{F}_{p}; similarly, define BpB_{p}. Let

𝒫1={p𝒫:|Ap|<m},𝒫2={p𝒫:0Ap},𝒫3={p𝒫:pn},\mathcal{P}_{1}=\{p\in\mathcal{P}:|A_{p}|<m\},\quad\mathcal{P}_{2}=\{p\in\mathcal{P}:0\in A_{p}\},\quad\mathcal{P}_{3}=\{p\in\mathcal{P}:p\mid n\},\quad

We claim that |𝒫1|m2log2N|\mathcal{P}_{1}|\leq m^{2}\log_{2}N. Indeed, if p𝒫p\in\mathcal{P} and |Ap|<|A|=m|A_{p}|<|A^{\prime}|=m, then there are two distinct elements a1,a2Aa_{1},a_{2}\in A^{\prime} such that p(a1a2)p\mid(a_{1}-a_{2}). However, there are less than m2m^{2} pairs of distinct elements a1,a2Aa_{1},a_{2}\in A^{\prime}, and for each such pair a1,a2a_{1},a_{2}, the number of distinct prime factors of |a1a2||a_{1}-a_{2}| is at most log2N\log_{2}N since 0<|a1a2|N0<|a_{1}-a_{2}|\leq N. A similar argument shows that |𝒫2|mlog2N|\mathcal{P}_{2}|\leq m\log_{2}N and |𝒫3|log2N|\mathcal{P}_{3}|\leq\log_{2}N.

Let 𝒫~=𝒫𝒫1𝒫2𝒫3\widetilde{\mathcal{P}}=\mathcal{P}\setminus\mathcal{P}_{1}\setminus\mathcal{P}_{2}\setminus\mathcal{P}_{3}. Next we apply Lemma 2.7 to bound |B||B| using the set of primes 𝒫~\widetilde{\mathcal{P}}. By [37, Theorem 2.7], we have

p𝒫~logpppQlogpp=logQ+O(1)=ϕ(k)θk,mlogN+O(1),\sum_{p\in\widetilde{\mathcal{P}}}\frac{\log p}{p}\leq\sum_{p\leq Q}\frac{\log p}{p}=\log Q+O(1)=\frac{\phi(k)}{\theta_{k,m}}\log N+O(1),

and thus

exp(p𝒫~logpp)Q.\exp\bigg{(}\sum_{p\in\widetilde{\mathcal{P}}}\frac{\log p}{p}\bigg{)}\ll Q. (5.3)

Let p𝒫~p\in\widetilde{\mathcal{P}}; we claim that

|Bp|pgcd(k,p1)m+mp.|B_{p}|\leq\frac{p}{\gcd(k,p-1)^{m}}+m\sqrt{p}. (5.4)

Note that we have pnp\nmid n, |Ap|=m|A_{p}|=m and 0Ap0\notin A_{p}. If gcd(k,p1)=1\gcd(k,p-1)=1, then we simply apply the trivial bound |Bp|p|B_{p}|\leq p; if gcd(k,p1)>1\gcd(k,p-1)>1, then we have

ApBp+n{xk:x𝔽p}={xgcd(k,p1):x𝔽p}A_{p}B_{p}+n\subseteq\{x^{k}:x\in\mathbb{F}_{p}\}=\{x^{\gcd(k,p-1)}:x\in\mathbb{F}_{p}\}

with pnp\nmid n and Ap𝔽pA_{p}\subseteq\mathbb{F}_{p}^{*}, and thus Lemma 4.3 implies that inequality (5.4).

It follows from inequality (5.4) that

p𝒫~logp|Bp|p𝒫~logppgcd(k,p1)m+mp=p𝒫~logppgcd(k,p1)mp𝒫~logpmppgcd(k,p1)m(pgcd(k,p1)m+mp).\sum_{p\in\widetilde{\mathcal{P}}}\frac{\log p}{|B_{p}|}\geq\sum_{p\in\widetilde{\mathcal{P}}}\frac{\log p}{\frac{p}{\gcd(k,p-1)^{m}}+m\sqrt{p}}=\sum_{p\in\widetilde{\mathcal{P}}}\frac{\log p}{\frac{p}{\gcd(k,p-1)^{m}}}-\sum_{p\in\widetilde{\mathcal{P}}}\frac{\log p\cdot m\sqrt{p}}{\frac{p}{\gcd(k,p-1)^{m}}(\frac{p}{\gcd(k,p-1)^{m}}+m\sqrt{p})}. (5.5)

Note that from equation (5.2), we have

p𝒫~logpmppgcd(k,p1)m(pgcd(k,p1)m+mp)\displaystyle\sum_{p\in\widetilde{\mathcal{P}}}\frac{\log p\cdot m\sqrt{p}}{\frac{p}{\gcd(k,p-1)^{m}}(\frac{p}{\gcd(k,p-1)^{m}}+m\sqrt{p})} p𝒫~k2mlogpmpp2\displaystyle\ll\sum_{p\in\widetilde{\mathcal{P}}}\frac{k^{2m}\log p\cdot m\sqrt{p}}{p^{2}}
mk2mplogpp3/2mk2m=o(logQ)\displaystyle\ll mk^{2m}\sum_{p}\frac{\log p}{p^{3/2}}\ll mk^{2m}=o(\log Q) (5.6)

as |n||n|\to\infty. On the other hand, by the prime number theorem for arithmetic progressions, we have

pQlogppgcd(k,p1)m\displaystyle\sum_{p\leq Q}\frac{\log p}{\frac{p}{\gcd(k,p-1)^{m}}} =1ikgcd(i,k)=1pQpi(modk)logppgcd(k,p1)m\displaystyle=\sum_{\begin{subarray}{c}1\leq i\leq k\\ \gcd(i,k)=1\end{subarray}}\sum_{\begin{subarray}{c}p\leq Q\\ p\equiv i\pmod{k}\end{subarray}}\frac{\log p}{\frac{p}{\gcd(k,p-1)^{m}}}
=1ikgcd(i,k)=1(1+o(1))gcd(k,p1)mϕ(k)logQ=(1+o(1))θk,mϕ(k)logQ.\displaystyle=\sum_{\begin{subarray}{c}1\leq i\leq k\\ \gcd(i,k)=1\end{subarray}}\frac{(1+o(1))\gcd(k,p-1)^{m}}{\phi(k)}\log Q=\frac{(1+o(1))\theta_{k,m}}{\phi(k)}\log Q. (5.7)

By Lemma 2.10, for each j{1,2,3}j\in\{1,2,3\}, we have

p𝒫jlogppgcd(k,p1)mkmp𝒫jlogppkmlog|𝒫j|kmloglogN=o(logQ).\sum_{p\in\mathcal{P}_{j}}\frac{\log p}{\frac{p}{\gcd(k,p-1)^{m}}}\leq k^{m}\sum_{p\in\mathcal{P}_{j}}\frac{\log p}{p}\ll k^{m}\log|\mathcal{P}_{j}|\ll k^{m}\log\log N=o(\log Q). (5.8)

Thus, combining equation (5.7) and inequality (5.8), we have

p𝒫~logppgcd(k,p1)mpQlogppgcd(k,p1)mj=13p𝒫jlogppgcd(k,p1)m=(1+o(1))θk,mϕ(k)logQ.\sum_{p\in\widetilde{\mathcal{P}}}\frac{\log p}{\frac{p}{\gcd(k,p-1)^{m}}}\geq\sum_{p\leq Q}\frac{\log p}{\frac{p}{\gcd(k,p-1)^{m}}}-\sum_{j=1}^{3}\sum_{p\in\mathcal{P}_{j}}\frac{\log p}{\frac{p}{\gcd(k,p-1)^{m}}}=\frac{(1+o(1))\theta_{k,m}}{\phi(k)}\log Q. (5.9)

Combining inequalities (5.5), (5.6), and (5.9),

p𝒫~logp|Bp|(1+o(1))θk,mϕ(k)logQ.\sum_{p\in\widetilde{\mathcal{P}}}\frac{\log p}{|B_{p}|}\geq\frac{(1+o(1))\theta_{k,m}}{\phi(k)}\log Q. (5.10)

Combining inequalities (5.3) and (5.10) with Lemma 2.7,

|B|QNo(1)=Nϕ(k)θk,m+o(1),|B|\ll QN^{o(1)}=N^{\frac{\phi(k)}{\theta_{k,m}}+o(1)},

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 N=2|n|17N=2|n|^{17}. In view of 3.3, we may assume that A,B[1,N]A,B\subseteq[1,N]. Let ν=loglogN\nu=\lceil\log\log N\rceil. Set

Q=(48νϕ(k)logN)2,𝒫={pQ:p1(modk),pn}.Q=(48\nu\phi(k)\log N)^{2},\quad\mathcal{P}=\{p\leq Q:p\equiv 1\pmod{k},\,p\nmid n\}.

For each prime pp, let ApA_{p} be the image of AA modulo pp, and define BpB_{p} similarly. Let 𝒫1={p𝒫:|Ap|2νp1/2ν}\mathcal{P}_{1}=\{p\in\mathcal{P}:|A_{p}|\geq 2\nu p^{1/2\nu}\} and let 𝒫2=𝒫𝒫1\mathcal{P}_{2}=\mathcal{P}\setminus\mathcal{P}_{1}.

For each prime p𝒫p\in\mathcal{P}, we can view ApA_{p} and BpB_{p} as subsets of 𝔽p\mathbb{F}_{p}. For each aAa\in A and bBb\in B, ab+nab+n is a perfect kk-th power. It follows that ApBp+n{xk:x𝔽p}A_{p}B_{p}+n\subseteq\{x^{k}:x\in\mathbb{F}_{p}\}. Thus, by Corollary 4.5, if p𝒫1p\in\mathcal{P}_{1}, that is, |Ap|2νp1/2ν|A_{p}|\geq 2\nu p^{1/2\nu}, then |Bp|12νp|B_{p}|\leq 12\nu\sqrt{p}. Let

S:=p𝒫logp,S1:=p𝒫1logp,S2:=p𝒫2logp.S:=\sum_{\begin{subarray}{c}p\in\mathcal{P}\end{subarray}}\log p,\quad S_{1}:=\sum_{\begin{subarray}{c}p\in\mathcal{P}_{1}\end{subarray}}\log p,\quad S_{2}:=\sum_{\begin{subarray}{c}p\in\mathcal{P}_{2}\end{subarray}}\log p.

By the prime number for arithmetic progressions, we have

Sp1(modk)pQlogppnlogp(1o(1))Qϕ(k)logN=(1o(1))Qϕ(k).S\geq\sum_{\begin{subarray}{c}p\equiv 1\pmod{k}\\ p\leq Q\end{subarray}}\log p-\sum_{p\mid n}\log p\geq(1-o(1))\frac{Q}{\phi(k)}-\log N=(1-o(1))\frac{Q}{\phi(k)}. (5.11)

We consider the following two cases.

Case 1: S1S2S_{1}\geq S_{2}. Then we have S1S/2S_{1}\geq S/2 and thus

p𝒫1logp|Bp|p𝒫1logp12νpS112νQS24νQ.\sum_{\begin{subarray}{c}p\in\mathcal{P}_{1}\end{subarray}}\frac{\log p}{|B_{p}|}\geq\sum_{\begin{subarray}{c}p\in\mathcal{P}_{1}\end{subarray}}\frac{\log p}{12\nu\sqrt{p}}\geq\frac{S_{1}}{12\nu\sqrt{Q}}\geq\frac{S}{24\nu\sqrt{Q}}.

It follows from inequality (5.11) that

p𝒫1logp|Bp|logNS24νQlogN(1o(1))Q24νϕ(k)logN=(1o(1))logN.\sum_{\begin{subarray}{c}p\in\mathcal{P}_{1}\end{subarray}}\frac{\log p}{|B_{p}|}-\log N\geq\frac{S}{24\nu\sqrt{Q}}-\log N\geq(1-o(1))\frac{\sqrt{Q}}{24\nu\phi(k)}-\log N=(1-o(1))\log N. (5.12)

Applying Gallagher’s sieve (Lemma 2.6), and inequalities (5.11) and (5.12), we conclude that

|B|p𝒫1logplogNp𝒫1logp|Bp|logN(1+o(1))Qϕ(k)logNklogN(loglogN)2log|n|(loglog|n|)2.|B|\leq\frac{\underset{p\in\mathcal{P}_{1}}{\sum}\log p-\log N}{\underset{p\in\mathcal{P}_{1}}{\sum}\frac{\log p}{|B_{p}|}-\log N}\leq(1+o(1))\frac{Q}{\phi(k)\log N}\ll_{k}\log N(\log\log N)^{2}\ll\log|n|(\log\log|n|)^{2}.

Case 2: S1<S2S_{1}<S_{2}. Then we have S2S/2S_{2}\geq S/2 and thus

p𝒫2logp|Ap|p𝒫2logp2νp1/2νS22νQ1/2νS4νQ1/2ν.\sum_{\begin{subarray}{c}p\in\mathcal{P}_{2}\end{subarray}}\frac{\log p}{|A_{p}|}\geq\sum_{\begin{subarray}{c}p\in\mathcal{P}_{2}\end{subarray}}\frac{\log p}{2\nu p^{1/2\nu}}\geq\frac{S_{2}}{2\nu Q^{1/2\nu}}\geq\frac{S}{4\nu Q^{1/2\nu}}. (5.13)

Applying Gallagher’s sieve (Lemma 2.6), and inequalities (5.11) and (5.13), we conclude that

|A|p𝒫2logplogNp𝒫2logp|Ap|logNSS4νQ1/2νlogNνQ1/2νkloglogNloglog|n|.|A|\leq\frac{\underset{p\in\mathcal{P}_{2}}{\sum}\log p-\log N}{\underset{p\in\mathcal{P}_{2}}{\sum}\frac{\log p}{|A_{p}|}-\log N}\leq\frac{S}{\frac{S}{4\nu Q^{1/2\nu}}-\log N}\ll\nu Q^{1/2\nu}\ll_{k}\log\log N\ll\log\log|n|.

Thus, we conclude that either |A|kloglog|n||A|\ll_{k}\log\log|n| or |B|klog|n|(loglog|n|)2|B|\ll_{k}\log|n|(\log\log|n|)^{2}, as desired. ∎

Next, we use a similar idea to prove Theorem 1.8.

Proof of Theorem 1.8.

Let ν=loglogN\nu=\lceil\log\log N\rceil. Let cc and LL be the positive constants from Corollary 2.9. Without loss of generality, we may assume that L10L\geq 10 and c1c\leq 1. Let

Q=kL(24c1νϕ(k)klogN)2,𝒫={pQ:p1(modk),pn}.Q=k^{L}(24c^{-1}\nu\phi(k)\sqrt{k}\log N)^{2},\quad\mathcal{P}=\{p\leq Q:p\equiv 1\pmod{k},\,p\nmid n\}.

We define 𝒫1,𝒫2,S,S1,S2\mathcal{P}_{1},\mathcal{P}_{2},S,S_{1},S_{2} in the same way as in the proof of Theorem 1.7.

Since QkLQ\geq k^{L}, Corollary 2.9 implies that

Sp1(modk)pQlogppnlogpcQϕ(k)klogN.S\geq\sum_{\begin{subarray}{c}p\equiv 1\pmod{k}\\ p\leq Q\end{subarray}}\log p-\sum_{p\mid n}\log p\geq\frac{cQ}{\phi(k)\sqrt{k}}-\log N. (5.14)

We consider the following two cases.

Case 1: S1S2S_{1}\geq S_{2}. As in the proof of Theorem 1.7, we have p𝒫1logp|Bp|S24νQ\sum_{\begin{subarray}{c}p\in\mathcal{P}_{1}\end{subarray}}\frac{\log p}{|B_{p}|}\geq\frac{S}{24\nu\sqrt{Q}}. It follows from inequality (5.14) that

p𝒫1logp|Bp|logNcQ24νϕ(k)k2logN(kL/22)logNlogN.\sum_{\begin{subarray}{c}p\in\mathcal{P}_{1}\end{subarray}}\frac{\log p}{|B_{p}|}-\log N\geq\frac{c\sqrt{Q}}{24\nu\phi(k)\sqrt{k}}-2\log N\geq(k^{L/2}-2)\log N\gg\log N. (5.15)

Applying Gallagher’s sieve (Lemma 2.6), inequalities (5.14) and (5.15), and the prime number theorem, we conclude that

|B|p𝒫1logplogNp𝒫1logp|Bp|logNQlogNkL+3ν2logN(logN)L+5,|B|\leq\frac{\underset{p\in\mathcal{P}_{1}}{\sum}\log p-\log N}{\underset{p\in\mathcal{P}_{1}}{\sum}\frac{\log p}{|B_{p}|}-\log N}\ll\frac{Q}{\log N}\ll k^{L+3}\nu^{2}\log N\ll(\log N)^{L+5},

where we used the assumption that k5logNk\leq 5\log N in the last step, and the implied constants are all absolute.

Case 2: S1<S2S_{1}<S_{2}. As in the proof of Theorem 1.7, we have p𝒫2logp|Ap|S4νQ1/2ν.\sum_{\begin{subarray}{c}p\in\mathcal{P}_{2}\end{subarray}}\frac{\log p}{|A_{p}|}\geq\frac{S}{4\nu Q^{1/2\nu}}. Now Gallagher’s sieve implies that

|A|p𝒫2logplogNp𝒫2logp|Ap|logNSS4νQ1/2νlogNνQ1/2νloglogN,|A|\leq\frac{\underset{p\in\mathcal{P}_{2}}{\sum}\log p-\log N}{\underset{p\in\mathcal{P}_{2}}{\sum}\frac{\log p}{|A_{p}|}-\log N}\leq\frac{S}{\frac{S}{4\nu Q^{1/2\nu}}-\log N}\ll\nu Q^{1/2\nu}\ll\log\log N,

where we used the assumption that k5logNk\leq 5\log N in the last step, and the implied constants are all absolute.

Thus, we conclude that either |A|loglogN|A|\ll\log\log N or |B|(logN)L+5|B|\ll(\log N)^{L+5}, as desired. ∎

6. Bounds on Diophantine tuples with property Dd(n)D_{\leq d}(n)

In this section, we combine the results from all previous sections to prove Theorems 1.1 and 1.2.

6.1. Proof of Theorem 1.1

Let N=2|n|17N=2|n|^{17} and let AA be a Diophantine tuple with property Dd(n)D_{\leq d}(n). In view of Proposition 3.6, to prove the theorem, we can additionally assume that NN is sufficiently large and A{1,2,,N}A\subseteq\{1,2,\ldots,N\}. In particular, since pnlogp/p(log|n|)1/2\sum_{p\mid n}\log p/\sqrt{p}\ll(\log|n|)^{1/2} [32, Lemma 2.8], we can choose NN large enough so that

pnlogpplogN.\sum_{p\mid n}\frac{\log p}{\sqrt{p}}\leq\log N. (6.1)

List the primes up to dd by p1,p2,,pmp_{1},p_{2},\ldots,p_{m} and set rr to be the product of these primes. By the prime number theorem, logrd\log r\ll d. Let cc and LL be the positive constants from Corollary 2.9. Without loss of generality, we may assume that L10L\geq 10 and c1c\leq 1. Let

Q=rL(c1Dϕ(r)rlogN)2,𝒫={pQ:p1(modr),pn}.Q=r^{L}(c^{-1}D\phi(r)\sqrt{r}\log N)^{2},\quad\mathcal{P}=\{p\leq Q:p\equiv 1\pmod{r},p\nmid n\}.

For each prime p𝒫p\in\mathcal{P}, let ApA_{p} be the image of AA modulo pp and view ApA_{p} as a subset of 𝔽p\mathbb{F}_{p}. Let a,bAa,b\in A with aba\neq b; then ab+nVdab+n\in V_{d}, that is, there are xx\in\mathbb{N} and 1jm1\leq j\leq m, such that ab+n=xpjab+n=x^{p_{j}}. Since p1(modpi)p\equiv 1\pmod{p_{i}} for each 1im1\leq i\leq m, it follows that ab+nab+n, viewed as an element in 𝔽p\mathbb{F}_{p}, lies in the set i=1m{ypi:y𝔽p}\bigcup_{i=1}^{m}\{y^{p_{i}}:y\in\mathbb{F}_{p}\}. Thus, Proposition 4.2 implies that |Ap|Dp|A_{p}|\leq D\sqrt{p}, where

D=(2m+2)i=1m(11/pi)14m.D=(2^{m}+2)\prod_{i=1}^{m}(1-1/p_{i})^{-1}\ll 4^{m}.

Thus, logDmd\log D\ll m\ll d. By the prime number theorem,

p𝒫logppQlogpQrL+3D2(logN)2.\sum_{p\in\mathcal{P}}\log p\leq\sum_{p\leq Q}\log p\ll Q\ll r^{L+3}D^{2}(\log N)^{2}. (6.2)

Since QrLQ\geq r^{L}, Corollary 2.9 implies that

p1(modr)pQlogpp1Qp1(modr)pQlogpcQϕ(r)rrL/2DlogN.\sum_{\begin{subarray}{c}p\equiv 1\pmod{r}\\ p\leq Q\end{subarray}}\frac{\log p}{\sqrt{p}}\geq\frac{1}{\sqrt{Q}}\sum_{\begin{subarray}{c}p\equiv 1\pmod{r}\\ p\leq Q\end{subarray}}\log p\geq\frac{c\sqrt{Q}}{\phi(r)\sqrt{r}}\geq r^{L/2}D\log N. (6.3)

Thus, inequalities (6.1) and (6.3) imply that

p𝒫logp|Ap|1Dp1(modr)pQlogpppnlogppLrL/2logNlogNrL/2logN.\sum_{p\in\mathcal{P}}\frac{\log p}{|A_{p}|}\geq\frac{1}{D}\sum_{\begin{subarray}{c}p\equiv 1\pmod{r}\\ p\leq Q\end{subarray}}\frac{\log p}{\sqrt{p}}-\sum_{p\mid n}\frac{\log p}{\sqrt{p}}\geq Lr^{L/2}\log N-\log N\gg r^{L/2}\log N. (6.4)

Applying Gallagher’s sieve (Lemma 2.6), inequalities (6.2) and (6.4), and the fact logrd\log r\ll d and logDd\log D\ll d, we conclude that

|A|p𝒫logplogNp𝒫logp|Ap|logNr3+L/2D2logNedLlogN|A|\leq\frac{\underset{p\in\mathcal{P}}{\sum}\log p-\log N}{\underset{p\in\mathcal{P}}{\sum}\frac{\log p}{|A_{p}|}-\log N}\ll r^{3+L/2}D^{2}\log N\ll e^{dL^{\prime}}\log N

for some absolute constant LL^{\prime}, where the implied constant is absolute, as required.

6.2. Proof of Theorem 1.2

By Proposition 3.9, we may assume that A[1,M]A\subseteq[1,M], where M=(4|n|)17M=(4|n|)^{17}. We may assume that MM is sufficiently large. Let GG be the complete graph with vertex set AA. For each abAab\in A with aba\neq b, by definition, ab+nab+n is a perfect power; we can thus write ab+n=xpab+n=x^{p} for some positive integer xx and a prime pp and we color the edge abab by the smallest such pp. Note that each prime pp we used to color some edge satisfies that plog2(ab+n)log2(M2+n)5logMp\leq\log_{2}(ab+n)\leq\log_{2}(M^{2}+n)\leq 5\log M. Let L1,L2L_{1},L_{2} be the two absolute constants from 1.8. Then it follows from 1.8 that GG does not contain a monochromatic KL1loglogM,(logM)L2+1K_{\lceil L_{1}\log\log M\rceil,\lfloor(\log M)^{L_{2}}\rfloor+1} as a subgraph. Thus, for each prime p5logMp\leq 5\log M, Lemma 2.2 implies that the number of edges in GG with color pp is at most

(logM)L2/(L1loglogM)|A|21/L1loglogM+(L1loglogM)|A|.(\log M)^{L_{2}/(L_{1}\log\log M)}|A|^{2-1/\lceil L_{1}\log\log M\rceil}+(L_{1}\log\log M)|A|.

Since the total number of edges in GG is (|A|2)\binom{|A|}{2}, it follows that

(|A|2)5logM((logM)L2/(L1loglogM)|A|21/L1loglogM+(L1loglogM)|A|),\binom{|A|}{2}\leq 5\log M\bigg{(}(\log M)^{L_{2}/(L_{1}\log\log M)}|A|^{2-1/\lceil L_{1}\log\log M\rceil}+(L_{1}\log\log M)|A|\bigg{)},

and thus

|A|(10eL2/L1logM)L1loglogM+2L1loglogMexp(2L1(loglog|n|)2),|A|\leq(10e^{L_{2}/L_{1}}\log M)^{\lceil L_{1}\log\log M\rceil}+2L_{1}\log\log M\ll\exp(2L_{1}(\log\log|n|)^{2}),

where the implied constant is absolute, as required.

7. Proofs of conditional results

7.1. Conditional estimates on Mk(n)M_{k}(n): Proof of Theorem 1.11

Assume that u<vu<v are two elements in AA. For each bBb\in B, there exist positive integers xx and yy such that ub+n=xkub+n=x^{k} and vb+n=ykvb+n=y^{k}; it follows that uykvxk=(uv)nuy^{k}-vx^{k}=(u-v)n. In particular, for two distinct b,bbb,b^{\prime}\in b, they induce a nontrivial integer solution (x,y,z,w)(x,y,z,w) to the Diophantine equation vxkuykvzk+uwk=0vx^{k}-uy^{k}-vz^{k}+uw^{k}=0 in the sense that (x,y)(z,w)(x,y)\neq(z,w).

Suppose |B|8|B|\geq 8, then we can pick 88 elements of BB to generate

vx1kuy1kvz1k+uw1k=0\displaystyle vx_{1}^{k}-uy_{1}^{k}-vz_{1}^{k}+uw_{1}^{k}=0
vx2kuy2kvz2k+uw2k=0\displaystyle vx_{2}^{k}-uy_{2}^{k}-vz_{2}^{k}+uw_{2}^{k}=0
vx3kuy3kvz3k+uw3k=0\displaystyle vx_{3}^{k}-uy_{3}^{k}-vz_{3}^{k}+uw_{3}^{k}=0
vx4kuy4kvz4k+uw4k=0\displaystyle vx_{4}^{k}-uy_{4}^{k}-vz_{4}^{k}+uw_{4}^{k}=0

More precisely, take 88 distinct elements b1,b2,b3,b4,c1,c2,c3,c4b_{1},b_{2},b_{3},b_{4},c_{1},c_{2},c_{3},c_{4} from BB, and for each 1i41\leq i\leq 4, define xi,yi,zi,wix_{i},y_{i},z_{i},w_{i} via the following rule:

ubi+n=xik,vbi+n=yik,uci+n=zik,vci+n=wik.ub_{i}+n=x_{i}^{k},\quad vb_{i}+n=y_{i}^{k},\quad uc_{i}+n=z_{i}^{k},\quad vc_{i}+n=w_{i}^{k}. (7.1)

It follows that the matrix

M:=M(b1,b2,b3,b4,c1,c2,c3,c4)=(x1ky1kz1kw1kx2ky2kz2kw2kx3ky3kz3kw3kx4ky4kz4kw4k)M:=M(b_{1},b_{2},b_{3},b_{4},c_{1},c_{2},c_{3},c_{4})=\begin{pmatrix}x_{1}^{k}&y_{1}^{k}&z_{1}^{k}&w_{1}^{k}\\ x_{2}^{k}&y_{2}^{k}&z_{2}^{k}&w_{2}^{k}\\ x_{3}^{k}&y_{3}^{k}&z_{3}^{k}&w_{3}^{k}\\ x_{4}^{k}&y_{4}^{k}&z_{4}^{k}&w_{4}^{k}\end{pmatrix}

has determinant 0 since M(v,u,v,u)T=0M(v,-u,-v,u)^{T}=\textbf{0}. Thus, we have by the Leibniz formula for the determinant of MM that

0=σ(1)sgn(σ)(xσ(1)yσ(2)zσ(3)wσ(4))k,0=\sum_{\sigma}(-1)^{\operatorname{sgn}(\sigma)}(x_{\sigma(1)}y_{\sigma(2)}z_{\sigma(3)}w_{\sigma(4)})^{k}, (7.2)

where σ\sigma runs over S4S_{4}, the set of all 2424 permutations of 1,2,3,41,2,3,4. Since k25k\geq 25, if we can show that the 24 numbers xσ(1)yσ(2)zσ(3)wσ(4)x_{\sigma(1)}y_{\sigma(2)}z_{\sigma(3)}w_{\sigma(4)} are pairwise distinct for σS4\sigma\in S_{4}, then equation (7.2) will violate Conjecture 1.10.

Thus, it suffices to show that if |B|21737|B|\geq 21737, then we can always choose 88 distinct elements b1,b2,b3,b4,c1,c2,c3,c4b_{1},b_{2},b_{3},b_{4},c_{1},c_{2},c_{3},c_{4} from BB and for each 1i41\leq i\leq 4, define xi,yi,zi,wix_{i},y_{i},z_{i},w_{i} via equation (7.1), such that the 24 numbers xσ(1)yσ(2)zσ(3)wσ(4)x_{\sigma(1)}y_{\sigma(2)}z_{\sigma(3)}w_{\sigma(4)} are pairwise distinct for σS4\sigma\in S_{4}, then we are done. Equivalently, we study the matrix

N:=N(b1,b2,b3,b4,c1,c2,c3,c4)=(x1y1z1w1x2y2z2w2x3y3z3w3x4y4z4w4).N:=N(b_{1},b_{2},b_{3},b_{4},c_{1},c_{2},c_{3},c_{4})=\begin{pmatrix}x_{1}&y_{1}&z_{1}&w_{1}\\ x_{2}&y_{2}&z_{2}&w_{2}\\ x_{3}&y_{3}&z_{3}&w_{3}\\ x_{4}&y_{4}&z_{4}&w_{4}\end{pmatrix}.

For convenience, we introduce the following terminology. For each 1i41\leq i\leq 4, we call a positive integer an i×ii\times i factor of NN if it appears as a term (up to the sign) in the Leibniz formula for the determinant of some i×ii\times i submatrix of NN. Thus, our goal is to guarantee that all the 4×44\times 4 factors of NN 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 g(t)=ut+nvt+ng(t)=\frac{ut+n}{vt+n} is monotone whenever vt+n>0vt+n>0. Since

ubi+nvbi+n=xikyik,\frac{ub_{i}+n}{vb_{i}+n}=\frac{x_{i}^{k}}{y_{i}^{k}},

it follows that if xiyi\frac{x_{i}}{y_{i}} is fixed, then bib_{i} is also uniquely fixed. Similarly, if ziwi\frac{z_{i}}{w_{i}} is fixed, then cic_{i} is also uniquely fixed. In particular, as long as b1,b2,b3,b4b_{1},b_{2},b_{3},b_{4} are distinct, we automatically have xiyjxjyix_{i}y_{j}\neq x_{j}y_{i} for each 1i<j41\leq i<j\leq 4. Similarly, as long as c1,c2,c3,c4c_{1},c_{2},c_{3},c_{4} are distinct, we automatically have ziwjzjwiz_{i}w_{j}\neq z_{j}w_{i} for each 1i<j41\leq i<j\leq 4.

We choose b1,b2,b3,b4,c1,c2,c3,c4b_{1},b_{2},b_{3},b_{4},c_{1},c_{2},c_{3},c_{4} from BB step by step such that they are in strictly increasing order, such that at each step, for each i{1,2,3,4}i\in\{1,2,3,4\}, all the i×ii\times i factors in the matrix NN (with entries already determined) are distinct. We use the following algorithm to choose the elements b1,b2,b3,b4b_{1},b_{2},b_{3},b_{4}:

  1. (1)

    choose b1b_{1} to be the smallest element from BB.

  2. (2)

    choose the smallest element b2Bb_{2}\in B such that b2>b1b_{2}>b_{1} and x2y1x_{2}\neq y_{1}.

  3. (3)

    choose the smallest element b3Bb_{3}\in B such that b3>b2b_{3}>b_{2} and x3{y1,y2}x_{3}\notin\{y_{1},y_{2}\}.

  4. (4)

    choose the smallest element b4Bb_{4}\in B such that b4>b3b_{4}>b_{3} and x4{y1,y2,y3}x_{4}\notin\{y_{1},y_{2},y_{3}\}.

Since |B|21737|B|\geq 21737, obviously the above algorithm went through. Note that for i{1,2}i\in\{1,2\}, all the i×ii\times i factors in the matrix NN with entries already determined are distinct.

Next we pick c1,c2,c3,c4c_{1},c_{2},c_{3},c_{4} step by step. For each 1j41\leq j\leq 4, we pick cjc_{j} using the following algorithm:

  1. (1)

    Let SjS_{j} be the collection of all possible factors in the matrix NN (with entries already determined).

  2. (2)

    Let Sj=Sj{1}S_{j}^{\prime}=S_{j}\cup\{1\} and let Qj=Sj/SjQ_{j}=S_{j}^{\prime}/S_{j}^{\prime} be the quotient set of SS^{\prime}.

  3. (3)

    Pick cjBc_{j}\in B such that zj,wj,wj/zjQjz_{j},w_{j},w_{j}/z_{j}\notin Q_{j}.

We first note that |Sj||S_{j}| is upper bounded by the total number of factors, so

|Sj|16+1692!+16943!+24=208.|S_{j}|\leq 16+\frac{16\cdot 9}{2!}+\frac{16\cdot 9\cdot 4}{3!}+24=208.

It follows from that |Qj(1,)|(2092)=21736|Q_{j}\cap(1,\infty)|\leq\binom{209}{2}=21736. Note that for all cjBc_{j}\in B, we have zj,wj,wjzj>1z_{j},w_{j},\frac{w_{j}}{z_{j}}>1. It follows that the above algorithm works since |B|21737|B|\geq 21737.

Next, we claim that the choice of cjc_{j} from the above algorithm guarantees that for each i{1,2,3,4}i\in\{1,2,3,4\}, all the i×ii\times i factors in the matrix NN (with entries already determined) are distinct. Let ii be fixed, consider two i×ii\times i factors m1m_{1} and m2m_{2} coming from different entries. Let EE be the set of labels of the entries building m1m_{1} and m2m_{2}. We consider the following four cases:

  1. (1)

    Both zjz_{j} and wjw_{j} are in EE. Without the loss of generality, we may assume that m1m_{1} contains zjz_{j} as an entry, and m2m_{2} contains wjw_{j} as an entry. Then if m1=m2m_{1}=m_{2}, we have m1/zjSjm_{1}/z_{j}\in S_{j}^{\prime} and m2/wj=m1/wjSjm_{2}/w_{j}=m_{1}/w_{j}\in S_{j}^{\prime}, and thus wj/zjQjw_{j}/z_{j}\in Q_{j}, violating the assumption that wj/zjQjw_{j}/z_{j}\notin Q_{j}.

  2. (2)

    zjz_{j} is in EE, and wjw_{j} is not. Without the loss of generality, we may assume that m1m_{1} contains zjz_{j} as an entry. If m2m_{2} also contains zjz_{j} as an entry, then before we pick cjc_{j}, we can already guarantee m1/zjm2/zjm_{1}/z_{j}\neq m_{2}/z_{j}; if m2m_{2} does not contain zjz_{j} as an entry, then m1=m2m_{1}=m_{2} would imply that m1/zj,m1Sjm_{1}/z_{j},m_{1}\in S_{j}^{\prime} and thus zjQjz_{j}\in Q_{j}, violating the assumption.

  3. (3)

    wjw_{j} is in EE, and zjz_{j} is not. Similar to the analysis in the previous case, we have m1m2m_{1}\neq m_{2}.

  4. (4)

    None of wjw_{j} and zjz_{j} are in EE. Then before we pick cjc_{j}, we can already guarantee m1m2m_{1}\neq m_{2}.

We conclude that m1m2m_{1}\neq m_{2}; this proves the claim.

Since |B|21737|B|\geq 21737, we can run the above algorithm to construct the desired c1,c2,c3,c4c_{1},c_{2},c_{3},c_{4} step by step, as required.

7.2. Condition estimates on Md(n)M_{\leq d}(n) and f~(x)\widetilde{f}(x): Proof of Theorem 1.15 and Theorem 1.16

Proof of Theorem 1.15.

Let 2d<2\leq d<\infty and nn be a nonzero integer. Let AA be a Diophantine tuple with property Dd(n)D_{\leq d}(n). Then ab+nVdab+n\in V_{d} for all a,bAa,b\in A with aba\neq b. Let GG be the complete graph with the vertex set AA; for each edge abab in GG, we color it with the smallest prime pp such that ab+nab+n is a pp-th power. By definition, all the primes we used to color the edges of GG are at most dd. Let CC be the constant from 1.14; then GG does not contain a monochromatic KC+1K_{C+1} as a subgraph. Let RR be the Ramsey number R(C+1,C+1,,C+1)R(C+1,C+1,\cdots,C+1), where the number of (C+1)(C+1)’s is 99. Since there are 99 primes less than 2525, it follows that GG contains no complete subgraph KR+1K_{R+1} such that all edges have color in {2,3,5,7,11,13,17,19,23}\{2,3,5,7,11,13,17,19,23\}. Thus, Lemma 2.1 implies that the number of edges in GG with color in {2,3,5,7,11,13,17,19,23}\{2,3,5,7,11,13,17,19,23\} is at most 12(11R1)|A|2\frac{1}{2}(1-\frac{1}{R-1})|A|^{2}. For each prime pp such that 25<pd25<p\leq d, 1.11 implies that GG does not contain a monochromatic K2,21737K_{2,21737} in color pp as a subgraph and thus the number of edges in GG with color pp is |A|3/2\ll|A|^{3/2} by Lemma 2.2. We conclude that

(|A|2)12(11R1)|A|2π(d)|A|3/2\binom{|A|}{2}-\frac{1}{2}(1-\frac{1}{R-1})|A|^{2}\ll\pi(d)|A|^{3/2}

and thus |A|π(d)2(d/logd)2|A|\ll\pi(d)^{2}\ll(d/\log d)^{2}. This finishes the proof that Md(n)(d/logd)2M_{\leq d}(n)\ll(d/\log d)^{2}, where the implied constant is absolute.

Next, we give an upper bound on f~(x)\widetilde{f}(x). Let A[1,x]A\subset[1,x]\cap\mathbb{N} such that there is an integer nn with 1|n|x1\leq|n|\leq x such that ab+nab+n is a perfect power for all a,bAa,b\in A with aba\neq b; since ab+nx2+xab+n\leq x^{2}+x and log2(x2+x)5logx\log_{2}(x^{2}+x)\leq 5\log x, it is necessary for ab+nab+n to be a dd-th power for some d5logxd\leq 5\log x. This shows that AA is a Diophantine tuple with property D5logx(n)D_{\leq\lfloor 5\log x\rfloor}(n) and thus

|A|M5logx(n)(logx/loglogx)2,|A|\leq M_{\leq\lfloor 5\log x\rfloor}(n)\ll(\log x/\log\log x)^{2},

as required. ∎

Proof of Theorem 1.16.

Assume that xx is sufficiently large. Pick a set A[1,x]A\subseteq[1,x]\cap\mathbb{N} with |A|=f~(x)|A|=\widetilde{f}(x), such that there is some 1|n|x1\leq|n|\leq x such that ab+nab+n is a perfect power for all a,bAa,b\in A with aba\neq b. Let GG be the graph with the vertex set AA, such that there is an edge between two distinct vertices aa and bb if and only if ab+nVab+n\in V_{\infty}. By definition, GG is a complete graph. For each a,bAa,b\in A such that aba\neq b, we color the edge abab with 11 if ab+nV25ab+n\in V_{25}, otherwise we color the edge abab with the smallest prime pp such that ab+nab+n is a pp-th power. Note that each prime pp we used to color some edge satisfies that 25<plog2(x2+n)5logx25<p\leq\log_{2}(x^{2}+n)\leq 5\log x. By Theorem 1.1, there is an absolute constant CC, such that GG does not contain a monochromatic KClogx+1K_{\lfloor C\log x\rfloor+1} with color 11 as a subgraph. Lemma 2.1 then implies that the number of edges in GG with color 11 is at most 12(11Clogx)|A|2\frac{1}{2}(1-\frac{1}{\lfloor C\log x\rfloor})|A|^{2}. This implies that the number of edges in GG with color being a prime is at least

(|A|2)12(11Clogx)|A|2|A|2Clogx|A||A|2logx.\binom{|A|}{2}-\frac{1}{2}(1-\frac{1}{\lfloor C\log x\rfloor})|A|^{2}\geq\frac{|A|^{2}}{\lfloor C\log x\rfloor}-|A|\gg\frac{|A|^{2}}{\log x}.

For each prime pp with 25<p<5logx25<p<5\log x, by Theorem 1.11, GG does not contain a monochromatic K2,21737K_{2,21737} with color pp as a subgraph; thus, Lemma 2.2 implies that the number of edges in GG with color pp is |A|3/2\ll|A|^{3/2}. It follows that |A|2logxlogx|A|3/2\frac{|A|^{2}}{\log x}\ll\log x\cdot|A|^{3/2} and thus f~(x)=|A|(logx)4\widetilde{f}(x)=|A|\ll(\log x)^{4}. ∎

7.3. Conditional estimates on M(n)M_{\leq\infty}(n): Proof of Theorem 1.17

Proof of Theorem 1.17.

Let AA be a Diophantine tuple with property M(n)M_{\leq\infty}(n). Under the ABC conjecture, by [3, Lemma 1], there is an absolute constant c0c_{0}, such that there do not exist 55 distinct positive integers a1,a2,a3,a4,a5a_{1},a_{2},a_{3},a_{4},a_{5}, each at least c0|n|3c_{0}|n|^{3}, with the property that for each 1i<j32051\leq i<j\leq 3205, aiaj+n=xijkija_{i}a_{j}+n={x_{ij}}^{k_{ij}} for some integers xijx_{ij} and kijk_{ij} with kij3205k_{ij}\geq 3205. Let N=max(c0|n|3,2|n|17)N=\max(c_{0}|n|^{3},2|n|^{17}). Define A1=A[1,N]A_{1}=A\cap[1,N] and A2=A(N,]A_{2}=A\cap(N,\infty]. By definition, |A1|f~(N)f~(2|n|17)|A_{1}|\ll\widetilde{f}(N)\ll\widetilde{f}(2|n|^{17}). It suffices to show |A2||A_{2}| is absolutely bounded, so that we have

|A|=|A1|+|A2|f~(2|n|17)exp(L(loglog|n|)2)|A|=|A_{1}|+|A_{2}|\ll\widetilde{f}(2|n|^{17})\ll\exp(L(\log\log|n|)^{2})

for some absolute constant LL from Corollary 1.3.

Next, we show that |A2||A_{2}| is absolutely bounded. Label the 452452 primes less than 32053205 by 2=p1<p2<<p452=32032=p_{1}<p_{2}<\ldots<p_{452}=3203. We build a complete graph GG with vertex set A2A_{2}, and we color the edge connecting two distinct elements a,aA2a,a^{\prime}\in A_{2} with the color ii by the following rule:

  • If aa+naa^{\prime}+n is a perfect kk-th power for some k3205k\geq 3205, then i=0i=0;

  • Otherwise, set ii to be the smallest number such that aa+naa^{\prime}+n is a perfect pip_{i}-th power.

In view of the above discussion, GG contains no monochromatic complete subgraph K5K_{5} in color 0. By Lemma 3.1, GG contains no monochromatic complete subgraph K22K_{22} in color 11. Let CC be the Ramsey number R(5,22)R(5,22); then by definition, GG contains no complete subgraph KCK_{C} with all edges in color 0 or 11. Lemma 2.1 then implies that the number of edges in GG with color 0 or 11 is at most 12(11C1)|A2|2\frac{1}{2}(1-\frac{1}{C-1})|A_{2}|^{2}. By Corollary 3.5, the number of edges in GG with color 22 is at most 8|A2|5/38|A_{2}|^{5/3}, and for each 3i4523\leq i\leq 452, the number of edges in GG with color ii is at most 7|A2|3/27|A_{2}|^{3/2}. It follows that

|A2|(|A2|1)212(11C1)|A2|2+8|A2|5/3+4507|A2|3/2,\frac{|A_{2}|(|A_{2}|-1)}{2}\leq\frac{1}{2}\bigg{(}1-\frac{1}{C-1}\bigg{)}|A_{2}|^{2}+8|A_{2}|^{5/3}+450\cdot 7|A_{2}|^{3/2},

and thus

|A2|(C1)+16(C1)|A2|2/3+6300|A2|1/2.|A_{2}|\leq(C-1)+16(C-1)|A_{2}|^{2/3}+6300|A_{2}|^{1/2}.

We conclude that |A2||A_{2}| 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 nn-shifted products. Publ. Math. Debrecen, 79(3-4):325–339, 2011.
  • [4] A. Bérczes, A. Dujella, L. Hajdu, and S. Tengely. Finiteness results for FF-Diophantine sets. Monatsh. Math., 180(3):469–484, 2016.
  • [5] N. C. Bonciocat, M. Cipu, and M. Mignotte. There is no Diophantine D(1)D(-1)-quadruple. J. Lond. Math. Soc. (2), 105(1):63–99, 2022.
  • [6] J. Bourgain and C. Demeter. On the number of kkth powers inside arithmetic progressions. arXiv e-prints, 2018. arXiv:1811.11919.
  • [7] J. Browkin and J. Brzeziński. Some remarks on the abcabc-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 mm-tuples. Proc. Amer. Math. Soc., 150(4):1455–1465, 2022.
  • [14] A. Dujella. On the size of Diophantine mm-tuples. Math. Proc. Cambridge Philos. Soc., 132(1):23–33, 2002.
  • [15] A. Dujella. Bounds for the size of sets with the property D(n)D(n). Glas. Mat. Ser. III, 39(59)(2):199–205, 2004.
  • [16] A. Dujella. Diophantine mm-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 mm-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 mm-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 kk-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 D(n)D(n), 2024. Bull. Aust. Math. Soc., to appear. arXiv:2406.00840.
  • [43] C. H. Yip. Multiplicatively irreducibility of small perturbations of shifted kk-th powers, 2025. arXiv:2501.16620.