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

Rational configuration problems and a family of curves

Jonathan Love McGill University [email protected]
(Date: November 2021)
Abstract.

Given η=(abcd)GL2()\eta=\begin{pmatrix}a&b\\ c&d\end{pmatrix}\in\operatorname{GL}_{2}(\mathbb{Q}), we consider the number of rational points on the genus one curve

Hη:y2=(a(1x2)+b(2x))2+(c(1x2)+d(2x))2.H_{\eta}:y^{2}=(a(1-x^{2})+b(2x))^{2}+(c(1-x^{2})+d(2x))^{2}.

We prove that the set of η\eta for which Hη()H_{\eta}(\mathbb{Q})\neq\emptyset has density zero, and that if a rational point (x0,y0)Hη()(x_{0},y_{0})\in H_{\eta}(\mathbb{Q}) exists, then Hη()H_{\eta}(\mathbb{Q}) is infinite unless a certain explicit polynomial in a,b,c,d,x0,y0a,b,c,d,x_{0},y_{0} vanishes.

Curves of the form HηH_{\eta} naturally occur in the study of configurations of points in n\mathbb{R}^{n} with rational distances between them. As one example demonstrating this framework, we prove that if a line through the origin in 2\mathbb{R}^{2} passes through a rational point on the unit circle, then it contains a dense set of points PP such that the distances from PP to each of the three points (0,0)(0,0), (0,1)(0,1), and (1,1)(1,1) are all rational. We also prove some results regarding whether a rational number can be expressed as a sum or product of slopes of rational right triangles.

Supported by CRM-ISM postdoctoral fellowship

1. Introduction

1.1. A family of curves

Fix η:=(abcd)GL2()\eta:=\begin{pmatrix}a&b\\ c&d\end{pmatrix}\in\operatorname{GL}_{2}(\mathbb{Q}), and let HηH_{\eta} be the curve defined by

(1) Hη:y2=(a(z2x2)+b(2xz))2+(c(z2x2)+d(2xz))2\displaystyle H_{\eta}:y^{2}=(a(z^{2}-x^{2})+b(2xz))^{2}+(c(z^{2}-x^{2})+d(2xz))^{2}

in the weighted projective plane where x,y,zx,y,z have degree 1,2,11,2,1, respectively. Rational points on this curve correspond to vectors (uv)2{0}\begin{pmatrix}u\\ v\end{pmatrix}\in\mathbb{Q}^{2}\setminus\{0\} such that both (uv)\begin{pmatrix}u\\ v\end{pmatrix} and η(uv)\eta\begin{pmatrix}u\\ v\end{pmatrix} have rational length, and as a result, curves of this form can be used to describe solutions to a collection of rational configuration problems; see Section 1.2 for more details. In this paper we study the loci of points η\eta for which HηH_{\eta} has zero, finitely many, or infinitely many rational points.

First, we show that for most values of η\eta, the curve HηH_{\eta} has no rational points.

Theorem 1.1.

Let (X)\mathcal{L}(X) be the set of ηGL2()\eta\in\operatorname{GL}_{2}(\mathbb{Q}) with a,b,c,d[X,X]a,b,c,d\in\mathbb{Z}\cap[-X,X] such that Hη(v)H_{\eta}(\mathbb{Q}_{v}) is nonempty for all v{,2,3,5,7,}v\in\{\infty,2,3,5,7,\ldots\}. Then for some constant C>0C>0,

|(X)|(2X)4<C(logX)1/4.\frac{|\mathcal{L}(X)|}{(2X)^{4}}<C(\log X)^{-1/4}.

The proof is given in Section 4. Note that HηHmηH_{\eta}\simeq H_{m\eta} for any positive integer mm, so by clearing denominators, every HηH_{\eta} is isomorphic to one of the curves counted in Theorem 1.1. For the sake of comparison, consider the following result by Bhargava, Cremona, and Fisher.

Theorem 1.2 ([4, Theorem 3]).

Let (X)\mathcal{L}^{\prime}(X) denote the set of (a,b,c,d,e)([X,X])5(a,b,c,d,e)\in(\mathbb{Z}\cap[-X,X])^{5} such that

y2=ax4+bx3+cx2+dx+ey^{2}=ax^{4}+bx^{3}+cx^{2}+dx+e

has a v\mathbb{Q}_{v} point for all v{,2,3,5,7,}v\in\{\infty,2,3,5,7,\ldots\}. Then

limX|(X)|(2X)50.7596.\lim_{X\to\infty}\frac{|\mathcal{L}^{\prime}(X)|}{(2X)^{5}}\approx 0.7596.

We see that the subfamily HηH_{\eta} differs from the larger family, in that far fewer specializations are everywhere locally soluble.

Now suppose we restrict our attention to the collection of points η\eta for which HηH_{\eta} does contain a rational point. In this case we have a stronger classification. Let ηt\eta^{t} denote the transpose of η\eta.

Proposition 1.3.

Suppose HηH_{\eta} (Eq. 1) has a rational point. If ηηt\eta\eta^{t} is a scalar matrix, then detη=λ2\det\eta=\lambda^{2} for some λ×\lambda\in\mathbb{Q}^{\times} and HηH_{\eta} is a union of two rational conics,

y=±λ(x2+1).y=\pm\lambda(x^{2}+1).

Otherwise HηH_{\eta} is isomorphic to

Er,s:y2=x3+(1+r2+s2)x2+s2xE_{r,s}:y^{2}=x^{3}+(1+r^{2}+s^{2})x^{2}+s^{2}x

for some r,sr,s\in\mathbb{Q} with s0s\neq 0 and (r,s)(0,±1)(r,s)\neq(0,\pm 1).

A proof is given in Section 3.4 using the fact that the isomorphism type of HηH_{\eta} is invariant under acting on the left and right of η\eta by elements of the orthogonal group O2()\operatorname{O}_{2}(\mathbb{Q}). An explicit change of variables expressing r,sr,s in terms of a,b,c,da,b,c,d and the rational point (x0:y0:z0)Hη()(x_{0}:y_{0}:z_{0})\in H_{\eta}(\mathbb{Q}) is given by Lemma 3.2.

Theorem 1.4.

Let r,sr,s\in\mathbb{Q} with s0s\neq 0 and (r,s)(0,±1)(r,s)\neq(0,\pm 1). The point (1,r)Er,s()(-1,r)\in E_{r,s}(\mathbb{Q}) is non-torsion if and only if r0r\neq 0, s±1s\neq\pm 1, and 4r2s±(1s2)24r^{2}s\neq\pm(1-s^{2})^{2}.

In particular, for most of the values η\eta such that Hη()H_{\eta}(\mathbb{Q}) is nonempty, Hη()H_{\eta}(\mathbb{Q}) is actually infinite. The proof of this result is given in Section 3.5. We discuss several applications of this result to rational distance problems in Section 1.2, but mention one here as a representative example.

Corollary 1.5.

On any line of the form x=0x=0 or y=2t1t2xy=\frac{2t}{1-t^{2}}x for t{1,0,1}t\in\mathbb{Q}\setminus\{-1,0,1\}, there exists a dense set of points with rational distance from each of (0,0)(0,0), (0,1)(0,1), and (1,1)(1,1).

In fact we prove a stronger result: there is an infinite collection of curves CnC_{n} in the plane such that the intersection points of the curves CnC_{n} with any fixed line y=2t1t2xy=\frac{2t}{1-t^{2}}x (for t{1,0,1}t\in\mathbb{Q}\setminus\{-1,0,1\}) gives a dense set of solutions to the three-distance problem within the given line (Corollary 5.4).

Even in the cases where (1,r)Er,s()(-1,r)\in E_{r,s}(\mathbb{Q}) is torsion, there are still several cases in which we can prove Er,s()E_{r,s}(\mathbb{Q}) has positive rank. We discuss these in more depth in Section 5.4, but note one special case here. Let

𝒮\displaystyle\mathcal{S}^{\prime} ={α:α2+1}\displaystyle=\{\alpha\in\mathbb{Q}:\sqrt{\alpha^{2}+1}\in\mathbb{Q}\}

denote the set of slopes of rational right triangles (including negatives and zero).

Proposition 1.6.

For all tt\in\mathbb{Q}, the equations x1+x2+x3=tx_{1}+x_{2}+x_{3}=t and x1x2x3=tx_{1}x_{2}x_{3}=t each have an infinite set of solutions with x1,x2,x3𝒮x_{1},x_{2},x_{3}\in\mathcal{S}^{\prime}.

See Section 5.4 for a proof.

1.2. Rational configuration problems

Given a finite simple graph G=(V,E)G=(V,E), an embedding ϕ:Vn\phi:V\hookrightarrow\mathbb{R}^{n} is a rational configuration if the distance d(ϕ(v),ϕ(w))d(\phi(v),\phi(w)) is rational for all (v,w)E(v,w)\in E. We may add some additional constraints to the set of allowable embeddings (for instance, we may require some pairs of edges to be the same length, or to meet at right angles), and in doing so we obtain a corresponding rational configuration problem: to determine whether there exists a rational configuration satisfying the desired constraints, and if so, to classify or count the number of rational configurations. We describe a list of sample rational configurations below; the corresponding graphs can be found in Table 1.

  • “Adjacent rectangles:” Find two rectangles sharing an edge such that the distance between any two vertices is rational.

  • “Detour:” Fix parameters r,s,t×r,s,t\in\mathbb{Q}^{\times}. Find a point xx such that (x,0)(x,0) has rational distance to (0,0)(0,0), (r,0)(r,0), (0,s)(0,s), and (r,t)(r,t). (A traveller is going from (0,s)(0,s) to (r,t)(r,t), but has to take a detour to stop at the xx-axis along the way; can they do so using only two straight paths of rational length?)

  • “Perfect cuboid:” Find a rectangular prism such that the distance between any two vertices is rational.

  • “Body cuboid:” Find a rectangular prism such that the distance between any two vertices that share a face are rational.

  • “Square four-distance:” Find a point (x,y)2(x,y)\in\mathbb{R}^{2} such that the distance to each of (0,0)(0,0), (1,0)(1,0), (0,1)(0,1), and (1,1)(1,1) is rational.

  • “Square three-distance:” Find a point (x,y)2(x,y)\in\mathbb{R}^{2} such that the distance to each of (0,0)(0,0), (0,1)(0,1), and (1,1)(1,1) is rational.

  • “Rectangle four-distance:” Find r×r\in\mathbb{Q}^{\times} and a point (x,y)2(x,y)\in\mathbb{R}^{2} such that the distance to each of (0,0)(0,0), (0,1)(0,1), (r,0)(r,0), and (r,1)(r,1) is rational.

  • “Rational distances under Möbius transformation:” Fix a,b,c,da,b,c,d\in\mathbb{Q} with adbc0ad-bc\neq 0. Find zz\in\mathbb{C} such that zz and az+bcz+d\frac{az+b}{cz+d} both have rational distance from 0.

Configuration Graph Equation Solutions given by
Adjacent rectangles α1\alpha_{1}α2\alpha_{2}11 α1+α2=α3\alpha_{1}+\alpha_{2}=\alpha_{3} Eα3,1()E_{\alpha_{3},1}(\mathbb{Q})
(\infty for all α3𝒮\alpha_{3}\in\mathcal{S}:
Proposition 5.5)
Detour (r,s×r,s\in\mathbb{Q}^{\times}) sα1s\alpha_{1}α2\alpha_{2}ss11rr sα1+α2=rs\alpha_{1}+\alpha_{2}=r Er,s()E_{r,s}(\mathbb{Q})
(\infty if |s|1|s|\neq 1 and 4r2s±(1s2)24r^{2}s\neq\pm(1-s^{2})^{2}:
Theorem 1.4)
Perfect cuboid α1\alpha_{1}α2\alpha_{2}11 α12+α22=α32\alpha_{1}^{2}+\alpha_{2}^{2}=\alpha_{3}^{2} Unknown
Body cuboid α1\alpha_{1}α2\alpha_{2}11 α1α3=α2\alpha_{1}\alpha_{3}=\alpha_{2} E0,α3()E_{0,\alpha_{3}}(\mathbb{Q})
(Square) Four-distance 11αα2\alpha\alpha_{2}α3\alpha_{3}α1\alpha_{1} α1α2=α3α4=α1+α31\alpha_{1}\alpha_{2}=\alpha_{3}\alpha_{4}=\alpha_{1}+\alpha_{3}-1 Unknown
(Square) three-distance 11αα2\alpha\alpha_{2}α3\alpha_{3}α1\alpha_{1} α1α2=α1+α31\alpha_{1}\alpha_{2}=\alpha_{1}+\alpha_{3}-1 E1,1α3()E_{-1,1-\alpha_{3}}(\mathbb{Q})
(\infty for all α3𝒮\alpha_{3}\in\mathcal{S}:
Corollary 1.5)
(Rectangle) four-distance 11α1α2\alpha_{1}\alpha_{2}α1\alpha_{1}α3\alpha_{3} α1α2=α3α4\alpha_{1}\alpha_{2}=\alpha_{3}\alpha_{4} E0,α3α4E_{0,\alpha_{3}\alpha_{4}}
Rational distances under Möbius transformation, η=(abcd)GL2()\eta=\begin{pmatrix}a&b\\ c&d\end{pmatrix}\in\operatorname{GL}_{2}(\mathbb{Q}) cα1+dc\alpha_{1}+daα1+ba\alpha_{1}+bα1\alpha_{1}11 (aα1+b)α2=(cα1+d)(a\alpha_{1}+b)\alpha_{2}=(c\alpha_{1}+d) Hη()H_{\eta}(\mathbb{Q})
Table 1. Diagrams of rational distance problems. Rational configurations are given by solutions to the given equation with αi𝒮\alpha_{i}\in\mathcal{S} for all ii. By Proposition 1.8, these configurations are parametrized by rational points on a curve HηH_{\eta}. If a distinguished rational point on HηH_{\eta} is known, then the isomorphic elliptic curve Er,sE_{r,s} is given (Proposition 1.3), and labeled with \infty in the cases that Er,s()E_{r,s}(\mathbb{Q}) is known to be infinite.

The perfect cuboid problem and square four-distance problem are classic unsolved problems (see Section 2); this paper does not present a solution to either of them. However, we can put all the remaining problems in this list into a common framework. Define

𝒮\displaystyle\mathcal{S} ={(u:v)1()u2+v2},\displaystyle=\{(u:v)\in\mathbb{P}^{1}(\mathbb{Q})\mid\sqrt{u^{2}+v^{2}}\in\mathbb{Q}\},

so that whenever u,vu,v, not both zero, are the legs of a (possibly degenerate) rational right triangle, the slope of the triangle is in 𝒮\mathcal{S}. Then for distinct P1,P22P_{1},P_{2}\in\mathbb{Q}^{2}, the distance between P1P_{1} and P2P_{2} is rational if and only if the line between P1P_{1} and P2P_{2} has slope in 𝒮\mathcal{S}. Using this observation, we can parametrize solutions to many rational configuration problems by finding elements of 𝒮\mathcal{S} satisfying simple polynomial relations.

Example 1.7.

Given a hypothetical solution to the perfect cuboid problem, we can scale the solution so that one edge length has length 11; this implies there exist α1,α2{0}\alpha_{1},\alpha_{2}\in\mathbb{Q}\setminus\{0\} such that

1+α12,1+α22,α12+α22,and1+α12+α221+\alpha_{1}^{2},\quad 1+\alpha_{2}^{2},\quad\alpha_{1}^{2}+\alpha_{2}^{2},\quad\text{and}\quad 1+\alpha_{1}^{2}+\alpha_{2}^{2}

are all perfect squares. If we set α3=α12+α22\alpha_{3}=\sqrt{\alpha_{1}^{2}+\alpha_{2}^{2}}, then the polynomial constraints above are equivalent to requiring

α12+α22=α32for some α1,α2,α3𝒮{(1:0),(0:1)}.\alpha_{1}^{2}+\alpha_{2}^{2}=\alpha_{3}^{2}\qquad\text{for some }\alpha_{1},\alpha_{2},\alpha_{3}\in\mathcal{S}\setminus\{(1:0),(0:1)\}.

Similar polynomial constraints for each of the problems above are listed in Table 1. Note that for every problem in Table 1 besides the perfect cuboid problem and the square four-distance problem, rational configurations correspond to solutions in 𝒮\mathcal{S} to a single polynomial in multiple variables that is linear in each variable.

Proposition 1.8.

Let η=(abcd)GL2()\eta=\begin{pmatrix}a&b\\ c&d\end{pmatrix}\in\operatorname{GL}_{2}(\mathbb{Q}), and let FηF_{\eta} be the curve in 1×1\mathbb{P}^{1}\times\mathbb{P}^{1} defined by

Fη:ax1x2+bx1z2+cz1x2+dz1z2=0.F_{\eta}:ax_{1}x_{2}+bx_{1}z_{2}+cz_{1}x_{2}+dz_{1}z_{2}=0.

There is a degree 44 morphism Φ:HηFη\Phi:H_{\eta}\to F_{\eta} inducing a surjection

Hη()Fη()(𝒮×𝒮).H_{\eta}(\mathbb{Q})\to F_{\eta}(\mathbb{Q})\cap(\mathcal{S}\times\mathcal{S}).

This follows from Proposition 5.1. Proposition 1.8 shows that for a wide collection of problems, rational configurations can be classified using rational points on curves of the form HηH_{\eta}. We can use this observation to show that some rational configuration problems have infinitely many rational configurations. In some cases, such as the detour problem and the square three-distance problem, the infinitude of solutions will be a consequence of Theorem 1.4. For others, including the adjacent rectangles, body cuboid, and rectangle four-distance problems, the corresponding curve Er,sE_{r,s} lands in one of the exceptional cases of Theorem 1.4, and so we cannot immediately conclude that there are infinitely many solutions.

1.3. Outline

We begin with a discussion of some related problems and their histories in Section 2. In Section 3 we analyze the algebraic structure of the family HηH_{\eta}, in particular showing that the isomorphism type of HηH_{\eta} is invariant under a left- and right-action of the orthogonal group (Section 3.3). We then analyze the singular fibers (Section 3.4), followed by the non-singular fibers that contain a rational point (Section 3.5), proving that most fibers of this type have infinitely many rational points (Theorem 1.4). Completing our study of rational points on the fibers, Section 4 contains a proof the set of fibers containing a rational point has low density (Theorem 1.1). Note that Section 4 only requires Section 3.1 and Section 3.2 from Section 3.

We conclude with some applications of these results in Section 5, focusing primarily on the square three-distance problem.

1.4. Acknowledgments

The author was supported by a CRM-ISM Postdoctoral Fellowship during the writing of this article, and would like to thank Andrew Granville, Sun-Kai Leung, Michael Lipnowski, Henri Darmon, Eyal Goren, Allysa Lumley, Olivier Mila, Wanlin Li, and Valeriya Kovaleva for helpful discussions.

2. Prior work on related problems

There are a number of open problems regarding rational configurations; in this section we will focus on two of them, namely the perfect cuboid problem in Section 2.1 and the square four-distance problem in Section 2.2 (both of these are discussed at greater length in [9]). In each case, we show that the problem is equivalent to the existence of a Pythagorean solution of a certain polynomial or system of polynomials. Finally, in Section 2.3, we compare to the congruent number problem.

2.1. Perfect cuboid problem

While the perfect cuboid problem is open, significant progress has been made towards studying the “body cuboid” problem, which is to give a cuboid in which all edges and all face diagonals (but not necessarily the body diagonal) have rational lengths. If 1,α1,α21,\alpha_{1},\alpha_{2} are the edge lengths of a body cuboid, then α12+1\alpha_{1}^{2}+1, α22+1\alpha_{2}^{2}+1, and α12+α22\alpha_{1}^{2}+\alpha_{2}^{2} are all perfect squares; the first two conditions say that α1,α2𝒮\alpha_{1},\alpha_{2}\in\mathcal{S} and the third is equivalent to requiring α2α1𝒮\frac{\alpha_{2}}{\alpha_{1}}\in\mathcal{S}.

For each fixed α3𝒮\alpha_{3}\in\mathcal{S}, the values α1,α2𝒮\alpha_{1},\alpha_{2}\in\mathcal{S} satisfying α1α3=α2\alpha_{1}\alpha_{3}=\alpha_{2} are parametrized by an elliptic curve (Proposition 1.8 and Proposition 1.3). This association between body cuboids and a family of elliptic curves is well-studied; Luijk has an in-depth survey [13] that mentions this association as well as many other known results about perfect cuboids. Halbeisen and Hungerbüler [10] investigate this problem as well. Given a fixed α3=ba\alpha_{3}=\frac{b}{a}, they associate solutions α1,α2𝒮{0}\alpha_{1},\alpha_{2}\in\mathcal{S}\setminus\{0\} satisfying α1α3=α2\alpha_{1}\alpha_{3}=\alpha_{2} to rational points on the elliptic curve

(2) E:y2=x3+(a2+b2)x2+a2b2x.\displaystyle E:y^{2}=x^{3}+(a^{2}+b^{2})x^{2}+a^{2}b^{2}x.

Proposition 1.8 and Proposition 1.3 recovers this classification. They show that there is a subgroup of E()E(\mathbb{Q}) isomorphic to /2×/4\mathbb{Z}/2\mathbb{Z}\times\mathbb{Z}/4\mathbb{Z} which give degenerate solutions to the corresponding rational distance problem. Ruling out other possible torsion points, they conclude [10, Theorem 8] that nondegnerate solutions exist if and only if E()E(\mathbb{Q}) has positive rank. In this case they call (a,b)(a,b) a double-pythapotent pair.

2.2. Four-distance problem

As with the perfect cuboid problem, the four-distance problem is currently out of reach, but a slightly weaker variant has many known solutions. The three-distance problem is to find points P=(x,y)2P=(x,y)\in\mathbb{R}^{2} with rational distance to (0,0)(0,0), (0,1)(0,1), and (1,1)(1,1). The coordinates x,yx,y are not a priori assumed to be rational, but since x2+y2x^{2}+y^{2}, x2+(1y)2x^{2}+(1-y)^{2}, and (1x)2+(1y)2(1-x)^{2}+(1-y)^{2} must all be rational, the differences 2y12y-1 and 2x12x-1 must also be rational, so in fact P2P\in\mathbb{Q}^{2}. We can then scale by an element of ×\mathbb{Q}^{\times} so that x=1x=1, and a solution to the square three-distance problem is equivalent to the existence of α1,α2,α3𝒮\alpha_{1},\alpha_{2},\alpha_{3}\in\mathcal{S} satsifying 1+α1α2=α1+α31+\alpha_{1}\alpha_{2}=\alpha_{1}+\alpha_{3}.

For many years it was believed that there were no solutions to the three-distance problem aside from points on the coordinate axes. The first one-parameter family of nontrivial solutions was found in 1967 by J.H. Hunter, and then many more infinite families were found in rapid succession; a historical overview is given by Berry, who also presents an “extraordinary abundance” of solutions lying in infinitely many one-parameter families [3]. We observe that the families of solutions obtained in Corollary 1.5 are distinct from those that appear in [3, Table 4], though it is unclear whether any (or all) of the one-parameter families we consider are eventually accounted for by Berry’s construction.

2.3. Congruent number problem

A rational number nn\in\mathbb{Q} is a congruent number if it is the area of a right triangle with rational edge lengths; that is, if there is a solution to

(3) a2+b2=c2and12ab=n,a,b,c×.\displaystyle a^{2}+b^{2}=c^{2}\qquad\text{and}\qquad\frac{1}{2}ab=n,\qquad a,b,c\in\mathbb{Q}^{\times}.

The “congruent number problem” is to determine whether a given nn\in\mathbb{Q} is a congruent number. This problem is not a rational configuration problem, but the underlying methods used to study these two problems are similar enough that a comparison is worthwhile.

There is a well-known approach to studying the congruent number problem; see for example the expositions [6] and [5]. For fixed nn, any solution to Eq. 3 corresponds to a rational point on an elliptic curve over \mathbb{Q} defined by

(4) E(n):y2=x3n2x.\displaystyle E^{(n)}:y^{2}=x^{3}-n^{2}x.

There are “degenerate points” in E(n)()E^{(n)}(\mathbb{Q}) that do not correspond to solutions; it can be shown that the set of degenerate points equals the torsion subgroup of E(n)()E^{(n)}(\mathbb{Q}). Thus nn is a congruent number if and only if E(n)()E^{(n)}(\mathbb{Q}) has positive rank. A formula due to Tunnell can be used to determine whether the analytic rank of E(n)E^{(n)} is zero or positive [16], so by assuming the Birch and Swinnerton-Dyer conjecture, this gives a criterion that determines whether a given number is congruent.

Many aspects of this paper are modeled off of the approach described for studying the congruent number problem. To put the two problems on a common footing, note that nn is a congruent number if and only if x1=a2x_{1}=a^{2} and x2=bax_{2}=\frac{b}{a} give a solution to

(5) x1x22n=0,x1(×)2,x2𝒮.\displaystyle x_{1}x_{2}-2n=0,\qquad x_{1}\in(\mathbb{Q}^{\times})^{2},\;x_{2}\in\mathcal{S}.

Both 𝒮\mathcal{S} and (×)2(\mathbb{Q}^{\times})^{2} can be represented as the image of 𝔸1()\mathbb{A}^{1}(\mathbb{Q}) under the image of a degree 22 rational map 𝔸1𝔸1\mathbb{A}^{1}\to\mathbb{A}^{1}. The curve E(n)E^{(n)} comes equipped with a degree 44 rational map to the variety defined by x1x22n=0x_{1}x_{2}-2n=0, and non-degenerate points in E(n)()E^{(n)}(\mathbb{Q}) map to solutions to Eq. 5. This is directly analogous to the relation between Hη()H_{\eta}(\mathbb{Q}) and solutions to rational distance problems (Proposition 1.8).

However, it is worth highlighting a few key differences between the congruent number problem and the family of rational distance problems we consider.

  • Size of parameter space. The isomorphism class of E(n)E^{(n)} is determined by the class of nn in ×/(×)2\mathbb{Q}^{\times}/(\mathbb{Q}^{\times})^{2}, while the isomorphism class of HηH_{\eta} is determined by the class of a corresponding matrix in a double quotient of GL2()\operatorname{GL}_{2}(\mathbb{Q}).

  • Existence of rational points. Every nn determines an elliptic curve E(n)E^{(n)}, which has a rational point. By contrast, the genus one curves HηH_{\eta} typically have no rational points (Theorem 1.1).

  • Closure under addition of degenerate points. In both problems, the corresponding genus one curve has a set of “degenerate” rational points, which do not yield valid solutions to the original problem. For the congruent number problem, the set of degenerate points equals the torsion subgroup of E(n)()E^{(n)}(\mathbb{Q}). For rational configuration problems, however, even if η\mathcal{H}_{\eta} is isomorphic to an elliptic curve (Proposition 1.3), the degenerate points in Hη(){H}_{\eta}(\mathbb{Q}) may not form a subgroup. This is to our advantage: we can often add together degenerate points to produce non-degenerate points, something that is not possible in the congruent number problem. This is the key idea behind Theorem 1.4.

  • Geometric variation in the family. The curves E(n)E^{(n)} are quadratic twists of the curve y2=x3xy^{2}=x^{3}-x, and are therefore all isomorphic over ¯\overline{\mathbb{Q}}. This fact is used in a key way in the proof of Tunnell’s theorem, as he applies a result due to Waldspurger [17] relating the central value of the LL-function of an elliptic curve with that of each of its quadratic twists. By contrast, the curves HηH_{\eta} do not have constant jj-invariant. This means that Tunnell’s approach to computing the analytic rank does not apply to this family.

3. The structure of the family

3.1. Assumptions and notation

Let KK be a field of characteristic not equal to 22, in which 1-1 is not a square; later we will restrict to K=K=\mathbb{Q}, but many of our results hold in more generality. Throughout this paper, all schemes will be defined over KK unless otherwise indicated, and if XX and YY are schemes then X×Y:=X×KYX\times Y:=X\times_{K}Y.

Throughout, 1\mathbb{P}^{1} will denote the projective line over KK, while 2\mathbb{P}^{2} will denote a weighted projective space over KK, where the variables x,y,zx,y,z have weights 1,2,11,2,1, respectively. We use the notation (x:z)(x:z) and (x:y:z)(x:y:z) to denote elements of 1(K)\mathbb{P}^{1}(K) and 2(K)\mathbb{P}^{2}(K), respectively. That is, for (x,z)K2{(0,0)}(x,z)\in K^{2}\setminus\{(0,0)\} we have

(x:z)={(λx:λz)λK×},(x:z)=\{(\lambda x:\lambda z)\mid\lambda\in K^{\times}\},

and for (x,y,z)K3{(0,0,0)}(x,y,z)\in K^{3}\setminus\{(0,0,0)\} we have

(x:y:z)={(λx,λ2y:λz)λK×}.(x:y:z)=\{(\lambda x,\lambda^{2}y:\lambda z)\mid\lambda\in K^{\times}\}.

Let GL2=SpecK[a,b,c,d,(adbc)1]\operatorname{GL}_{2}=\operatorname{Spec}K[a,b,c,d,(ad-bc)^{-1}] denote the algebraic group of 2×22\times 2 invertible matrices, with identity element II. Given a matrix ηGL2(K)\eta\in\operatorname{GL}_{2}(K), its transpose will be denoted ηt\eta^{t}. Let O2\operatorname{O}_{2} denote the orthogonal group of 2×22\times 2 matrices, that is, the algebraic subgroup of GL2{\operatorname{GL}_{2}} defined by the condition that MGL2(K¯)M\in{\operatorname{GL}_{2}}(\overline{K}) is in O2(K¯)\operatorname{O}_{2}(\overline{K}) if and only if MMt=MtM=IMM^{t}=M^{t}M=I.

3.2. Definition of \mathcal{H} and basic properties

Using the coordinates ((x:y:z),(abcd))\left((x:y:z),\begin{pmatrix}a&b\\ c&d\end{pmatrix}\right) on 2×GL2\mathbb{P}^{2}\times\operatorname{GL}_{2}, define the variety \mathcal{H} by the equation

(6) :y2\displaystyle\mathcal{H}:y^{2} =(a(z2x2)+b(2xz))2+(c(z2x2)+d(2xz))2.\displaystyle=(a(z^{2}-x^{2})+b(2xz))^{2}+(c(z^{2}-x^{2})+d(2xz))^{2}.

If we let N:K2KN:K^{2}\to K be defined by N(u,v)=u2+v2N(u,v)=u^{2}+v^{2}, then this can equivalently be written

(7) :y2\displaystyle\mathcal{H}:y^{2} =N((abcd)(z2x22xz)).\displaystyle=N\left(\begin{pmatrix}a&b\\ c&d\end{pmatrix}\begin{pmatrix}z^{2}-x^{2}\\ 2xz\end{pmatrix}\right).

This variety comes equipped with a morphism π:GL2\pi:\mathcal{H}\to\operatorname{GL}_{2}, which equips \mathcal{H} with the structure of a flat family of curves. Given ηGL2(K)\eta\in\operatorname{GL}_{2}(K), Hη{H}_{\eta} is the fiber of π\pi over η\eta.

The generic fiber of π\pi is a genus one hyperelliptic curve over the function field K(a,b,c,d)K(a,b,c,d), with discriminant

(8) Δ()=216(adbc)4((a+d)2+(bc)2)((ad)2+(b+c)2).\displaystyle\Delta(\mathcal{H})=2^{16}(ad-bc)^{4}((a+d)^{2}+(b-c)^{2})((a-d)^{2}+(b+c)^{2}).

The Jacobian variety of this curve is an elliptic curve over K(a,b,c,d)K(a,b,c,d), which by classical invariant theory (see for example [18, 2]) has a model

(9) E:y2=x3+(a2+b2+c2+d2)x2+(adbc)2x.\displaystyle E:y^{2}=x^{3}+(a^{2}+b^{2}+c^{2}+d^{2})x^{2}+(ad-bc)^{2}x.

We have two commuting involutions on \mathcal{H} as a scheme over GL2\operatorname{GL}_{2}, given by

(10) σ1:(x:y:z)(x:y:z)andσ2:(x:y:z)(z:y:x),\displaystyle\sigma_{1}:(x:y:z)\mapsto(x:-y:z)\qquad\text{and}\qquad\sigma_{2}:(x:y:z)\mapsto(-z:y:x),

generating a Klein four-group

(11) Γ:=σ1,σ2\displaystyle\Gamma:=\langle\sigma_{1},\sigma_{2}\rangle

acting on \mathcal{H}. (Note that σ2\sigma_{2} is an involution because yy has weight 22, and so (x:y:z)=(x:y:z)(-x:y:-z)=(x:y:z).)

3.3. Double cosets and reduction

We show that the isomorphism class of ηGL2(K)\eta\in\operatorname{GL}_{2}(K) is invariant on double cosets in

O2(K)\GL2(K)/(K×O2(K)),\operatorname{O}_{2}(K)\backslash\operatorname{GL}_{2}(K)/(K^{\times}\cdot\operatorname{O}_{2}(K)),

and use this to show that η\mathcal{H}_{\eta} has a KK-point if and only if η\eta is in the same double coset as (1r0s)\begin{pmatrix}1&r\\ 0&s\end{pmatrix} for some r,sKr,s\in K.

Lemma 3.1.

Let η,ηGL2(K)\eta,\eta^{\prime}\in{\operatorname{GL}_{2}}(K). If ηK×O2(K)ηO2(K)\eta^{\prime}\in K^{\times}\operatorname{O}_{2}(K)\eta\operatorname{O}_{2}(K), then there is an isomorphism τ:ηη\tau:\mathcal{H}_{\eta}\to\mathcal{H}_{\eta^{\prime}} over KK that commutes with the action of Γ\Gamma.

Proof.

Let η=λr1ηr21\eta^{\prime}=\lambda r_{1}\eta r_{2}^{-1}, where λK×\lambda\in K^{\times} and r1,r2O2(K)r_{1},r_{2}\in\operatorname{O}_{2}(K). Write r2=(uvϵvϵu)r_{2}=\begin{pmatrix}u&-v\\ \epsilon v&\epsilon u\end{pmatrix}, where u,vKu,v\in K, ϵ=±1\epsilon=\pm 1, and u2+v2=1u^{2}+v^{2}=1. There exists s,tKs,t\in K so that u=t2s2t2+s2u=\frac{t^{2}-s^{2}}{t^{2}+s^{2}} and v=2stt2+s2v=\frac{2st}{t^{2}+s^{2}}. Then for any (x:y:z)η(K¯)(x:y:z)\in\mathcal{H}_{\eta}(\overline{K}),

(λ(s2+t2)y)2\displaystyle(\lambda(s^{2}+t^{2})y)^{2} =N(λ(s2+t2)(abcd)(z2x22xz))\displaystyle=N\left(\lambda(s^{2}+t^{2})\begin{pmatrix}a&b\\ c&d\end{pmatrix}\begin{pmatrix}z^{2}-x^{2}\\ 2xz\end{pmatrix}\right)
=N(r1λ(abcd)r21(t2s22st2ϵstϵ(t2s2))(z2x22xz))\displaystyle=N\left(r_{1}\lambda\begin{pmatrix}a&b\\ c&d\end{pmatrix}r_{2}^{-1}\begin{pmatrix}t^{2}-s^{2}&-2st\\ 2\epsilon st&\epsilon(t^{2}-s^{2})\end{pmatrix}\begin{pmatrix}z^{2}-x^{2}\\ 2xz\end{pmatrix}\right)
=N(λr1(abcd)r21((tzsx)2(tx+sz)22ϵ(tzsx)(tx+sz))).\displaystyle=N\left(\lambda r_{1}\begin{pmatrix}a&b\\ c&d\end{pmatrix}r_{2}^{-1}\begin{pmatrix}(tz-sx)^{2}-(tx+sz)^{2}\\ 2\epsilon(tz-sx)(tx+sz)\end{pmatrix}\right).

Thus the map

τ:(x:y:z)(ϵ(tx+sz):λ(s2+t2)y:tzsx)\tau:(x:y:z)\mapsto(\epsilon(tx+sz):\lambda(s^{2}+t^{2})y:tz-sx)

defines an isomorphism ηη\mathcal{H}_{\eta}\to\mathcal{H}_{\eta^{\prime}}, and the involutions yyy\mapsto-y and (x:z)(z:x)(x:z)\mapsto(-z:x) are preserved. ∎

Given ηGL2(K)\eta\in{\operatorname{GL}_{2}}(K), suppose η\eta is in the same double coset as an element of the form η=(1r0s)GL2(K)\eta^{\prime}=\begin{pmatrix}1&r\\ 0&s\end{pmatrix}\in\operatorname{GL}_{2}(K). We have (0:1:1)η(K)(0:1:1)\in\mathcal{H}_{\eta^{\prime}}(K), so by Lemma 3.1, we can conclude that η(K)\mathcal{H}_{\eta}(K) is nonempty. The following lemma gives us the converse result: if η(K)\mathcal{H}_{\eta}(K) is nonempty then η\eta is in the same double-coset as a matrix of the form η=(1r0s)\eta^{\prime}=\begin{pmatrix}1&r\\ 0&s\end{pmatrix}.

Lemma 3.2.

Let η=(abcd)GL2(K)\eta=\begin{pmatrix}a&b\\ c&d\end{pmatrix}\in\operatorname{GL}_{2}(K). Suppose there is a point P=(x0:y0:z0)η(K)P=(x_{0}:y_{0}:z_{0})\in\mathcal{H}_{\eta}(K). Define

(12) r:=(ab+cd)((z02x02)2(2x0z0)2)(a2b2+c2d2)(z02x02)(2x0z0)y02,s:=(adbc)(z02+x02)2y02.\displaystyle\begin{aligned} r&:=\frac{(ab+cd)((z_{0}^{2}-x_{0}^{2})^{2}-(2x_{0}z_{0})^{2})-(a^{2}-b^{2}+c^{2}-d^{2})(z_{0}^{2}-x_{0}^{2})(2x_{0}z_{0})}{y_{0}^{2}},\\ s&:=\frac{(ad-bc)(z_{0}^{2}+x_{0}^{2})^{2}}{y_{0}^{2}}.\end{aligned}

Then ηK×O2(K)(1r0s)O2(K)\eta\in K^{\times}\operatorname{O}_{2}(K)\begin{pmatrix}1&r\\ 0&s\end{pmatrix}\operatorname{O}_{2}(K).

Proof.

Suppose x02+z02=0x_{0}^{2}+z_{0}^{2}=0. If z00z_{0}\neq 0, then (x0z0)2=1\left(\frac{x_{0}}{z_{0}}\right)^{2}=-1, contradicting the assumption that 1-1 is not a square in KK. Hence z0=0z_{0}=0, and likewise x0=0x_{0}=0. But this implies y0=0y_{0}=0, which contradicts the fact that (x0:y0:z0)2~(K)(x_{0}:y_{0}:z_{0})\in\widetilde{\mathbb{P}^{2}}(K).

If y0=0y_{0}=0, then a similar argument shows that we must have

a(z02x02)+b(2x0z0)=c(z02x02)+d(2x0z0)=0.a(z_{0}^{2}-x_{0}^{2})+b(2x_{0}z_{0})=c(z_{0}^{2}-x_{0}^{2})+d(2x_{0}z_{0})=0.

But this implies that the nonzero vector (z02x02,2x0z0)(z_{0}^{2}-x_{0}^{2},2x_{0}z_{0}) is in the kernel of (abcd)\begin{pmatrix}a&b\\ c&d\end{pmatrix}, contradicting the assumption that ηGL2(K)\eta\in\operatorname{GL}_{2}(K). Hence y00y_{0}\neq 0.

Since x02+z020x_{0}^{2}+z_{0}^{2}\neq 0 and y00y_{0}\neq 0, the matrices

r1\displaystyle r_{1} =1y0(a(z02x02)+b(2x0z0)c(z02x02)+d(2x0z0)c(z02x02)d(2x0z0)a(z02x02)+b(2x0z0))\displaystyle=\frac{1}{y_{0}}\begin{pmatrix}a(z_{0}^{2}-x_{0}^{2})+b(2x_{0}z_{0})&c(z_{0}^{2}-x_{0}^{2})+d(2x_{0}z_{0})\\ -c(z_{0}^{2}-x_{0}^{2})-d(2x_{0}z_{0})&a(z_{0}^{2}-x_{0}^{2})+b(2x_{0}z_{0})\end{pmatrix}
r2\displaystyle r_{2} =1z02+x02(z02x022x0z02x0z0z02x02)\displaystyle=\frac{1}{z_{0}^{2}+x_{0}^{2}}\begin{pmatrix}z_{0}^{2}-x_{0}^{2}&-2x_{0}z_{0}\\ 2x_{0}z_{0}&z_{0}^{2}-x_{0}^{2}\end{pmatrix}

are both well-defined elements of SO2(K)\operatorname{SO}_{2}(K). We can check by direct computation that z02+x02y0r1ηr2=(1r0s)\frac{z_{0}^{2}+x_{0}^{2}}{y_{0}}r_{1}\eta r_{2}=\begin{pmatrix}1&r\\ 0&s\end{pmatrix}. ∎

3.4. Isomorphism classes of fibers

The curve HηH_{\eta} is singular when the discriminant (Eq. 8) vanishes. Since adbc0ad-bc\neq 0 for all ηGL2(K)\eta\in\operatorname{GL}_{2}(K) and KK does not contain a square root of 1-1, this can only occur if a=da=-d and b=cb=c, or if a=da=d and b=cb=-c. One of these two conditions holds if and only if a2+b2=c2+d2a^{2}+b^{2}=c^{2}+d^{2} and ac+bd=0ac+bd=0; thus the singular fibers HηH_{\eta} are exactly those with

ηηt=(a2+b2)I.\eta\eta^{t}=(a^{2}+b^{2})I.

In this case HηH_{\eta} reduces to the form

y2=(a2+b2)(z2+x2)2.y^{2}=(a^{2}+b^{2})(z^{2}+x^{2})^{2}.

If a2+b2=λ2a^{2}+b^{2}=\lambda^{2} then HηH_{\eta} splits into two conics, y=±λ(z2+x2)y=\pm\lambda(z^{2}+x^{2}). If a2+b2a^{2}+b^{2} is not a square in KK, then there are no solutions in 2(K)\mathbb{P}^{2}(K).

If HηH_{\eta} has a rational point and the discriminant (Eq. 8) does not vanish at η\eta, then HηH_{\eta} is isomorphic to its Jacobian. Using Lemma 3.2 and Eq. 9, we can conclude that HηH_{\eta} is isomorphic to

Er,s:y2=x3+(1+r2+s2)x2+s2xE_{r,s}:y^{2}=x^{3}+(1+r^{2}+s^{2})x^{2}+s^{2}x

for some r,sKr,s\in K; the non-vanishing of the discriminant says that s0s\neq 0 and (r,s)(0,±1)(r,s)\neq(0,\pm 1). This completes the proof of Proposition 1.3.

3.5. Nonsingular fibers with a rational point

We now restrict our attention to K=K=\mathbb{Q} in order to prove Theorem 1.4, which we recall for convenience.

See 1.4

Proof.

Assume R:=(1,r)R:=(-1,r) is torsion in Er,s()E_{r,s}(\mathbb{Q}). Since x2+(1+r2+s2)x+s2x^{2}+(1+r^{2}+s^{2})x+s^{2} is positive on an open interval around x=0x=0, there exists 1<x<0-1<x<0 for which Er,s()E_{r,s}(\mathbb{R}) does not contain any point of the form (x,y)(x,y). Thus Er,s()E_{r,s}(\mathbb{R}) has two components, with R:=(1,r)R:=(-1,r) on the non-identity component and T:=(0,0)T:=(0,0) on the identity component. This shows RR is not a multiple of 22 in Er,s()E_{r,s}(\mathbb{R}), and hence RR cannot have odd order. By Mazur’s classification of torsion subgroups, we can conclude that if RR is torsion then its order must be an even number at most 1212. If RR has order 1010 then the only possibility for the torsion subgroup of Er,s()E_{r,s}(\mathbb{Q}) is /10\mathbb{Z}/{10}\mathbb{Z}, so that TT is the unique element of order 22. This implies T=5RT=5R, which again leads to a contradiction when we consider the component group of Er,s()E_{r,s}(\mathbb{R}).

We can conclude that if RR is torsion, it must have order {2,4,6,8,12}\ell\in\{2,4,6,8,12\}. For each such \ell, let ψ(r,s,x)[r,s,x]\psi_{\ell}(r,s,x)\in\mathbb{Z}[r,s,x] denote the \ell-th division polynomial on Er,sE_{r,s}; this is a polynomial with the property that ψ(r,s,x)=0\psi_{\ell}(r,s,x)=0 for x¯x\in\overline{\mathbb{Q}} if and only if (x,y)Er,s(¯)[](x,y)\in E_{r,s}(\overline{\mathbb{Q}})[\ell] for some y¯y\in\overline{\mathbb{Q}} (see for instance [15, Exercise 3.7]). We compute the division polynomial ψ(r,s,x)\psi_{\ell}(r,s,x), and determine all possible (r,s)2(r,s)\in\mathbb{Q}^{2} such that ψ(r,s,1)=0\psi_{\ell}(r,s,-1)=0.

  • We have ψ2(r,s,1)=r2\psi_{2}(r,s,-1)=-r^{2}, so RR has order 22 if and only if r=0r=0.

  • We have

    ψ4(r,s,1)ψ2(r,s,1)=2(s1)(s+1)(2r2s2+2r2+(s21)2).\frac{\psi_{4}(r,s,-1)}{\psi_{2}(r,s,-1)}=-2(s-1)(s+1)(2r^{2}s^{2}+2r^{2}+(s^{2}-1)^{2}).

    The last factor is a sum of non-negative terms, including at least one positive term because (r,s)(0,±1)(r,s)\neq(0,\pm 1). Hence RR has order 44 if and only if s=±1s=\pm 1.

  • The quotient of ψ6(r,s,1)\psi_{6}(r,s,-1) by ψ2(r,s,1)ψ3(r,s,1)\psi_{2}(r,s,-1)\psi_{3}(r,s,-1) factors into two irreducible polynomials in [r,s]\mathbb{Q}[r,s]. The first factor is 4r2s2+(s21)24r^{2}s^{2}+(s^{2}-1)^{2}, which is positive for all (r,s)(0,±1)(r,s)\neq(0,\pm 1). The second factor is

    16s2r44(s21)2(s2+1)r23(s21)4.16s^{2}r^{4}-4(s^{2}-1)^{2}(s^{2}+1)r^{2}-3(s^{2}-1)^{4}.

    Considering this as a quadratic polynomial in r2r^{2}, the discriminant is equal to

    16(s21)4(s4+14s2+1).16(s^{2}-1)^{4}(s^{4}+14s^{2}+1).

    In order for r2r^{2} to be rational (let alone rr), this discriminant must equal a rational square. Thus we consider rational points on the curve CC defined by y2=s4+14s2+1y^{2}=s^{4}+14s^{2}+1. There are eight rational points (s,y)C()(s,y)\in C(\mathbb{Q}): two at infinity, as well as (1,±4)(-1,\pm 4), (0,±1)(0,\pm 1), and (1,±4)(1,\pm 4). Using the Weierstrass form y2=x37x2+12xy^{2}=x^{3}-7x^{2}+12x for CC we can confirm that CC has no other rational points, so the only possibilities for ss are 1,0,1-1,0,1. If s=0s=0 then we have r2=34r^{2}=-\frac{3}{4}, yielding no rational solutions. If s=±1s=\pm 1 then we have r=0r=0, contradicting (r,s)(0,±1)(r,s)\neq(0,\pm 1).

  • The quotient of ψ8(r,s,1)\psi_{8}(r,s,-1) by ψ4(r,s,1)\psi_{4}(r,s,-1) factors into three irreducible polynomials in [r,s]\mathbb{Q}[r,s]. The first two factors are 4r2s(s21)24r^{2}s-(s^{2}-1)^{2} and 4r2s+(s21)24r^{2}s+(s^{2}-1)^{2}; these each have infinitely many rational solutions. The third factor is positive for all (r,s)(0,±1)(r,s)\neq(0,\pm 1).

  • If we eliminate common factors with ψ6(r,s,1)\psi_{6}(r,s,-1) and ψ4(r,s,1)\psi_{4}(r,s,-1) from ψ12(r,s,x)\psi_{12}(r,s,x), we are left with three irreducible polynomials in [r,s]\mathbb{Q}[r,s]. The first factor is

    16s(s2s+1)r4+8s(s21)2r2+(s21)4.16s(s^{2}-s+1)r^{4}+8s(s^{2}-1)^{2}r^{2}+(s^{2}-1)^{4}.

    Considered as a quadratic in r2r^{2}, the discriminant is 64s(s1)6(s+1)4-64s(s-1)^{6}(s+1)^{4}, which is a square if and only if s=k2s=-k^{2} for some kk\in\mathbb{Q}. Plugging this in and solving for r2r^{2}, we find that either

    r2=(k41)24k(k2+k+1)orr2=(k41)24k(k2k+1).r^{2}=\frac{(k^{4}-1)^{2}}{4k(k^{2}+k+1)}\qquad\text{or}\qquad r^{2}=-\frac{(k^{4}-1)^{2}}{4k(k^{2}-k+1)}.

    For the first option, we obtain rr\in\mathbb{Q} if and only if k3+k2+kk^{3}+k^{2}+k is a nonzero square. The only rational points on the elliptic curve y2=k3+k2+ky^{2}=k^{3}+k^{2}+k are the point at infinity and (k,y)=(0,0)(k,y)=(0,0), so there is no kk\in\mathbb{Q} for which rr is rational. For the second option, we obtain rr\in\mathbb{Q} if and only if (k)3+(k)2+(k)(-k)^{3}+(-k)^{2}+(-k) is a nonzero square, and by the same reasoning there is no such kk. Hence this factor is nonzero for all (r,s)2(r,s)\in\mathbb{Q}^{2}.

    The second factor is obtained from the first by sss\mapsto-s, so it has no rational solutions either. The third factor is positive for all (r,s)(0,±1)(r,s)\neq(0,\pm 1).

To summarize, we obtain the following possibilities:

  • (1,r)(-1,r) has order 22 if and only if r=0r=0;

  • (1,r)(-1,r) has order 44 if and only if s=±1s=\pm 1;

  • (1,r)(-1,r) has order 88 if and only if 4r2s=±(1s2)24r^{2}s=\pm(1-s^{2})^{2}.

  • There are no values of (r,s)2(r,s)\in\mathbb{Q}^{2} for which (1,r)(-1,r) has any other finite order.∎

4. Upper bound on locally soluble curves

In this section we prove Theorem 1.1. The main idea is to prove a local obstruction to the existence of p\mathbb{Q}_{p}-points (Lemma 4.1). This obstruction is “large,” in the sense that the proportion of curves satisfying the obstruction is approximately a constant multiple of 1p\frac{1}{p}. By contrast, each local obstruction in Theorem 1.2 only affects O(1p2)O(\frac{1}{p^{2}}) of all curves. At a high level, the difference in behavior between the two families stems from the fact that 1p\sum\frac{1}{p} diverges but 1p2\sum\frac{1}{p^{2}} converges.

4.1. Local Obstructions

Given a ring AA we use M2(A)M_{2}(A) to denote the ring of 2×22\times 2 matrices over AA. For primes pp we define

Rp:={ηM2(p)GL2(p):Hη(p)=};R_{p}:=\left\{\eta\in M_{2}(\mathbb{Z}_{p})\cap\operatorname{GL}_{2}(\mathbb{Q}_{p}):H_{\eta}(\mathbb{Q}_{p})=\emptyset\right\};

Thus (X)\mathcal{L}(X) counts the set of ηM2()GL2()\eta\in M_{2}(\mathbb{Z})\cap\operatorname{GL}_{2}(\mathbb{Q}) with entries of absolute value at most XX such that ηRp\eta\notin R_{p} for all primes pp (note that there is never a real obstruction: Hη()H_{\eta}(\mathbb{R})\neq\emptyset for all ηGL2()\eta\in\operatorname{GL}_{2}(\mathbb{Q}).)

The main contribution to RpR_{p} will come from the following constraint.

Lemma 4.1.

Let pp be an odd prime and η=(abcd)M2(p)GL2(p)\eta=\begin{pmatrix}a&b\\ c&d\end{pmatrix}\in M_{2}(\mathbb{Z}_{p})\cap\operatorname{GL}_{2}(\mathbb{Q}_{p}). Suppose padbcp\mid ad-bc, and aa, a2+b2a^{2}+b^{2}, and a2+c2a^{2}+c^{2} are all nonzero mod pp. Then Hη(p)H_{\eta}(\mathbb{Q}_{p}) is nonempty if and only if at least one of a2+b2a^{2}+b^{2} or a2+c2a^{2}+c^{2} is a square modulo pp.

Proof.

Assume Hη(p)H_{\eta}(\mathbb{Q}_{p}) has a point (x:y:z)(x:y:z); without loss of generality we can assume that x,y,zx,y,z are in p\mathbb{Z}_{p} and at least one of x,zx,z is in p×\mathbb{Z}_{p}^{\times}. Reducing modulo pp we have

(13) a2y2=(a2+c2)(a(z2x2)+b(2xz))2.\displaystyle{a}^{2}{y}^{2}=({a}^{2}+{c}^{2})({a}({z}^{2}-{x}^{2})+{b}(2{x}{z}))^{2}.

If a(z2x2)+b(2xz)0{a}({z}^{2}-{x}^{2})+{b}(2{x}{z})\neq 0, then

a2+c2=(aya(z2x2)+b(2xz))2{a}^{2}+{c}^{2}=\left(\frac{a}{y}{{a}({z}^{2}-{x}^{2})+{b}(2{x}{z})}\right)^{2}

is a square. On the other hand, suppose a(z2x2)+b(2xz)=0{a}({z}^{2}-{x}^{2})+{b}(2{x}{z})=0. Since a0a\neq 0 and at least one of x,zx,z is nonzero we must have xz0xz\neq 0, so

a2+b2\displaystyle{a}^{2}+{b}^{2} =a2+(a(z2x2)2xz)2=(a(z2+x2)2xz)2\displaystyle={a}^{2}+\left(\frac{-{a}({z}^{2}-{x}^{2})}{2{x}{z}}\right)^{2}=\left(\frac{{a}({z}^{2}+{x}^{2})}{2{x}{z}}\right)^{2}

is a square.

Conversely, if a2+c2a^{2}+c^{2} is a nonzero square mod pp then Eq. 13 clearly has solutions with y0{y}\neq 0; these are smooth points on Hη(𝔽p)H_{\eta}(\mathbb{F}_{p}) so they lift to points on Hη(p)H_{\eta}(\mathbb{Q}_{p}). If a2+b2a^{2}+b^{2} is a nonzero square mod pp, then it is a square in p\mathbb{Q}_{p}, so there exist x,zp×x,z\in\mathbb{Q}_{p}^{\times} with ba=x2z22xz\frac{b}{a}=\frac{x^{2}-z^{2}}{2xz}. This implies a(z2x2)+b(2xz)=0a(z^{2}-x^{2})+b(2xz)=0, so (x:c(z2x2)+d(2xz):z)(x:c(z^{2}-x^{2})+d(2xz):z) is a point in Hη(p)H_{\eta}(\mathbb{Q}_{p}). ∎

In light of the above, set

Ωp:={μ=(abcd)M2(𝔽p):adbc=0,(a2+b2p)=1,(a2+c2p)=1},\displaystyle\Omega_{p}:=\left\{\mu=\begin{pmatrix}{a}&{b}\\ {c}&{d}\end{pmatrix}\in M_{2}(\mathbb{F}_{p}):{a}{d}-{b}{c}=0,\;\bigg{(}\frac{{a}^{2}+{b}^{2}}{p}\bigg{)}=-1,\;\bigg{(}\frac{{a}^{2}+{c}^{2}}{p}\bigg{)}=-1\right\},

where (p)\left(\frac{\cdot}{p}\right) denotes the Legendre symbol. If by abuse of notation we associate Ωp\Omega_{p} with its preimage in M2(p)M_{2}(\mathbb{Z}_{p}) under reduction mod pp, we have

(14) (ΩpGL2(p))Rp.\displaystyle(\Omega_{p}\cap\operatorname{GL}_{2}(\mathbb{Q}_{p}))\subseteq R_{p}.

This follows from Lemma 4.1: note that the Legendre symbol conditions in the definition of Ωp\Omega_{p} force a{a} to be nonzero mod pp.

Lemma 4.2.

We have |Ωp|=14p3+O(p2)|\Omega_{p}|=\frac{1}{4}p^{3}+O(p^{2}).

Proof.

Let (abcd)Ωp\begin{pmatrix}{a}&{b}\\ {c}&{d}\end{pmatrix}\in\Omega_{p}. Since a2+b2{a}^{2}+{b}^{2} is not a square, we have a0{a}\neq 0. Note that a2+r2{a}^{2}+{r}^{2} is a square in 𝔽p\mathbb{F}_{p} if and only if r{r} is in the image of the map ϕ:𝔽p×𝔽p\phi:\mathbb{F}_{p}^{\times}\to\mathbb{F}_{p} given by ϕ(t)=a1t22t\phi({t})={a}\frac{1-{t}^{2}}{2{t}}. We have ϕ(t)=ϕ(s)\phi({t})=\phi({s}) if and only if s=1t{s}=-\frac{1}{{t}}, and so there are 12(p±1)\frac{1}{2}(p\pm 1) values in the range of ϕ\phi (with the ±\pm sign depending on whether or not 1-1 is a square mod pp). Hence b{b} and c{c} must each be one of the 12(p1)\frac{1}{2}(p\mp 1) values not in the range of ϕ\phi. Since there are p1p-1 choices for a{a}, there are 12(p1)\frac{1}{2}(p\mp 1) choices for each of b{b} and c{c}, and the value of d=bca{d}=\frac{{b}{c}}{{a}} is fixed, we obtain the desired count. ∎

The following result is not used in the sequel, but justifies the claim that Ωp\Omega_{p} is the main contribution to RpR_{p}.

Lemma 4.3.

Let Rp¯\overline{R_{p}} denote the image of RpR_{p} under reduction mod pp. We have

|Rp¯Ωp|=O(p2).|\overline{R_{p}}\setminus\Omega_{p}|=O(p^{2}).
Proof.

We produce a collection of pairs of linear equations over 𝔽p\mathbb{F}_{p} with the property that every element of Rp¯Ωp\overline{R_{p}}\setminus\Omega_{p} satisfies one of these pairs of equations. Let η¯=(abcd)M2(𝔽p)\bar{\eta}=\begin{pmatrix}a&b\\ c&d\end{pmatrix}\in M_{2}(\mathbb{F}_{p}). If

(adbc)((ad)2+(b+c)2)((a+d)2+(bc)2)0,(ad-bc)((a-d)^{2}+(b+c)^{2})((a+d)^{2}+(b-c)^{2})\neq 0,

then for any lift ηM2(p)GL2(p)\eta\in M_{2}(\mathbb{Z}_{p})\cap\operatorname{GL}_{2}(\mathbb{Q}_{p}), the discriminant of HηH_{\eta} (Eq. 8) is not divisible by pp, and so Hη(𝔽p)H_{\eta}(\mathbb{F}_{p}) is nonempty by the Hasse-Weil bound. As these are all smooth points, they lift to points in Hη(p)H_{\eta}(\mathbb{Q}_{p}). Thus η¯\bar{\eta} is not in Rp¯\overline{R_{p}}.

We can therefore assume that exactly one of the following constraints holds:

  1. (a)

    adbc=0{a}{d}-{b}{c}=0 and a=0{a}=0,

  2. (b)

    adbc=0{a}{d}-{b}{c}=0 and a0{a}\neq 0,

  3. (c)

    adbc0{a}{d}-{b}{c}\neq 0, (ad)2+(b±c)2=0({a}\mp{d})^{2}+({b}\pm{c})^{2}=0, and ad=0{a}\mp{d}=0,

  4. (d)

    adbc0{a}{d}-{b}{c}\neq 0, (ad)2+(b±c)2=0({a}\mp{d})^{2}+({b}\pm{c})^{2}=0 and ad0{a}\mp{d}\neq 0.

In case (a) or (c), η¯\bar{\eta} must satisfy one of the following pairs of linear equations over 𝔽p\mathbb{F}_{p}:

a=b=0,a=c=0,orad=b±c=0.{a}={b}=0,\qquad{a}={c}=0,\qquad\text{or}\qquad{a}\mp{d}={b}\pm{c}=0.

Now consider case (b). If a2+b2=0{a}^{2}+{b}^{2}=0 or a2+c2=0{a}^{2}+{c}^{2}=0, then using a0{a}\neq 0 and ad=bc{a}{d}={b}{c} we can conclude that

a+ib=c+id=0ora+ic=b+id=0{a}+{i}{b}={c}+{i}{d}=0\qquad\text{or}\qquad{a}+{i}{c}={b}+{i}{d}=0

for some i𝔽p{i}\in\mathbb{F}_{p} satisfying i2=1{i}^{2}=-1. If on the other hand a2+b2{a}^{2}+{b}^{2} and a2+c2{a}^{2}+{c}^{2} are both nonzero, then η¯\bar{\eta} is not in Rp¯\overline{R_{p}} by Lemma 4.1.

Finally we consider case (d). We assume (ad)2+(b+c)2=0(a-d)^{2}+(b+c)^{2}=0 and ad0a-d\neq 0 as the other case is similar. Then b+c=i(ad)b+c=i(a-d) for some i𝔽pi\in\mathbb{F}_{p} with i2=1i^{2}=-1. The reduction of HηH_{\eta} modulo pp is given by

y2=(z+ix)2h(x,z)y^{2}=(z+ix)^{2}h(x,z)

where

h(x,z):=(a2+c2)(z2x2)+(2a(b+c)+i(c2a2+2bc))(2xz).h(x,z):=(a^{2}+c^{2})(z^{2}-x^{2})+(2a(b+c)+i(c^{2}-a^{2}+2bc))(2xz).

If the discriminant of h(x,z)h(x,z) is nonzero, then the equation r2=h(x,z)r^{2}=h(x,z) defines a smooth projective conic over 𝔽p\mathbb{F}_{p}, and there must exist at least one point (x,z,r)(x,z,r) on this conic with z+ix0z+ix\neq 0. Then (x:(z+ix)r:z)(x:(z+ix)r:z) defines a smooth point in Hη(𝔽p)H_{\eta}(\mathbb{F}_{p}), which lifts to a point in Hη(p)H_{\eta}(\mathbb{Q}_{p}). Hence η¯\bar{\eta} is not in Rp¯\overline{R_{p}}. Computing the discriminant of h(x,z)h(x,z), we find that elements in Rp¯Ωp\overline{R_{p}}\setminus\Omega_{p} in this case must satisfy

b+ci(ad)=(bia)(a+ic)(b+c)=0.b+c-i(a-d)=(b-ia)(a+ic)(b+c)=0.

4.2. Proof of Theorem 1.1

The author would like to thank Sun-Kai Leung for suggesting the following proof.

Recall that (X)\mathcal{L}(X) is the set of all ηM2()\eta\in M_{2}(\mathbb{Z}) with nonzero determinant, entries having absolute value at most XX, and with ηRp\eta\notin R_{p} for all pp. Set

(X):={ηM2()GL2():ηΩp for all p}.\mathcal{M}(X):=\left\{\eta\in M_{2}(\mathbb{Z})\cap\operatorname{GL}_{2}(\mathbb{Q}):\eta\notin\Omega_{p}\text{ for all }p\right\}.

By Eq. 14 we can see that (X)(X)\mathcal{L}(X)\subseteq\mathcal{M}(X), and so it suffices to find an upper bound for (X)\mathcal{M}(X).

Set

𝒮(Y)\displaystyle\mathscr{S}(Y) :=mYμ2(m)prime pm|Ωp|p4|Ωp|,\displaystyle:=\sum_{m\leq Y}\mu^{2}(m)\prod_{\text{prime }p\mid m}\frac{|\Omega_{p}|}{p^{4}-|\Omega_{p}|},

where μ(m)\mu(m) is the Möbius function (so μ2(m)=1\mu^{2}(m)=1 if mm is squarefree and μ2(m)=0\mu^{2}(m)=0 otherwise). Applying the nn-dimensional large sieve [8, Lemma B], we obtain

|(X)||(X)|X4𝒮(X),\displaystyle|\mathcal{L}(X)|\leq|\mathcal{M}(X)|\ll\frac{X^{4}}{\mathscr{S}(\sqrt{X})},

where f(X)g(X)f(X)\ll g(X) means that for some positive constant CC we have f(X)<Cg(X)f(X)<Cg(X) for sufficiently large XX. Theorem 1.1 follows immediately from this bound after applying the following lemma.

Lemma 4.4.

We have

𝒮(Y)(logY)1/4.\mathscr{S}(Y)\gg(\log Y)^{1/4}.
Proof.

We have |Ωp|0|\Omega_{p}|\geq 0, and by Lemma 4.2 we have |Ωp|14(p3Cp2)|\Omega_{p}|\geq\frac{1}{4}(p^{3}-Cp^{2}) for some positive constant CC, so

𝒮(Y)\displaystyle\mathscr{S}(Y) mYμ2(m)prime pmp3Cp24p4\displaystyle\geq\sum_{m\leq Y}\mu^{2}(m)\prod_{\text{prime }p\mid m}\frac{p^{3}-Cp^{2}}{4p^{4}}
=mYμ2(m)mpm14(1Cp).\displaystyle=\sum_{m\leq Y}\frac{\mu^{2}(m)}{m}\prod_{p\mid m}\frac{1}{4}\left(1-\frac{C}{p}\right).

Set

f(m):=μ2(m)prime pm14(1Cp).f(m):=\mu^{2}(m)\prod_{\text{prime }p\mid m}\frac{1}{4}\left(1-\frac{C}{p}\right).

Note that ff is a multiplicative function, it is positive for sufficiently large pp, and

prime pYf(p)logpp\displaystyle\sum_{\text{prime }p\leq Y}\frac{f(p)\log p}{p} =14pYlogppCpYlogpp2\displaystyle=\frac{1}{4}\sum_{p\leq Y}\frac{\log p}{p}-C\sum_{p\leq Y}\frac{\log p}{p^{2}}
=14log(Y)+O(1)\displaystyle=\frac{1}{4}\log(Y)+O(1)

(see for instance [7, p. 57]). Hence, by Wirsing’s Theorem ([11, Theorem 14.3] with κ=14\kappa=\frac{1}{4}, c=0c=0, k=1k=1) we have

𝒮(Y)\displaystyle\mathscr{S}(Y) mYf(m)m(logY)1/4.\displaystyle\geq\sum_{m\leq Y}\frac{f(m)}{m}\gg(\log Y)^{1/4}.

5. Applications to Rational Distance Problems

5.1. From Hη(K)H_{\eta}(K) to rational configurations

We return temporarily to the more general setting of a field KK in which 1-1 is not a square. Let η=(abcd)GL2(K)\eta=\begin{pmatrix}a&b\\ c&d\end{pmatrix}\in\operatorname{GL}_{2}(K). In Eq. 6, we defined the variety

:y2\displaystyle\mathcal{H}:y^{2} =N((abcd)(z2x22xz)),\displaystyle=N\left(\begin{pmatrix}a&b\\ c&d\end{pmatrix}\begin{pmatrix}z^{2}-x^{2}\\ 2xz\end{pmatrix}\right),

where N(u,v):=u2+v2N(u,v):=u^{2}+v^{2}. We also define the subvariety \mathcal{F} of 1×1×GL2\mathbb{P}^{1}\times\mathbb{P}^{1}\times\operatorname{GL}_{2} (with coordinates ((u1:v1),(u2:v2),(abcd))((u_{1}:v_{1}),(u_{2}:v_{2}),\begin{pmatrix}a&b\\ c&d\end{pmatrix}) by

:au1u2+bu1v2+cv1u2+dv1v2=0.\mathcal{F}:au_{1}u_{2}+bu_{1}v_{2}+cv_{1}u_{2}+dv_{1}v_{2}=0.

As with \mathcal{H}, there is a projection map to GL2\operatorname{GL}_{2} giving \mathcal{F} the structure of a flat family of curves, with FηF_{\eta} denoting the fiber over ηGL2(K)\eta\in\operatorname{GL}_{2}(K). We define

𝒮K\displaystyle\mathcal{S}_{K} :={(u:v)1(K)u2+v2 is a square in K}\displaystyle:=\{(u:v)\in\mathbb{P}^{1}(K)\mid u^{2}+v^{2}\text{ is a square in }K\}
={(z2x2:2xz)(x:z)1(K)}.\displaystyle=\{(z^{2}-x^{2}:2xz)\mid(x:z)\in\mathbb{P}^{1}(K)\}.
Proposition 5.1.

There is a morphism Φ:\Phi:\mathcal{H}\to\mathcal{F} over GL2\operatorname{GL}_{2} defined by sending (x:y:z)(x:y:z) to

(15) ((c(z2x2)d(2xz):a(z2x2)+b(2xz)),(z2x2:2xz)).\displaystyle\left((-c(z^{2}-x^{2})-d(2xz):a(z^{2}-x^{2})+b(2xz)),\,(z^{2}-x^{2}:2xz)\right).

Further, Φ\Phi induces a bijection between the set of Γ\Gamma-orbits in Hη(K){H}_{\eta}(K) and the set Fη(K)(𝒮K×𝒮K)F_{\eta}(K)\cap(\mathcal{S}_{K}\times\mathcal{S}_{K}).

Proof.

Notice that the equation defining \mathcal{F} can be written

(16) 0\displaystyle 0 =(u1v1)(abcd)(u2v2).\displaystyle=\begin{pmatrix}u_{1}&v_{1}\end{pmatrix}\begin{pmatrix}a&b\\ c&d\end{pmatrix}\begin{pmatrix}u_{2}\\ v_{2}\end{pmatrix}.

The morphism Φ\Phi is well-defined because

((cdab)(z2x22xz))t(abcd)(z2x22xz)\displaystyle\left(\begin{pmatrix}-c&-d\\ a&b\end{pmatrix}\begin{pmatrix}z^{2}-x^{2}\\ 2xz\end{pmatrix}\right)^{t}\begin{pmatrix}a&b\\ c&d\end{pmatrix}\begin{pmatrix}z^{2}-x^{2}\\ 2xz\end{pmatrix}
=(z2x22xz)(0adbcad+bc0)(z2x22xz)\displaystyle\qquad=\begin{pmatrix}z^{2}-x^{2}&2xz\end{pmatrix}\begin{pmatrix}0&ad-bc\\ -ad+bc&0\end{pmatrix}\begin{pmatrix}z^{2}-x^{2}\\ 2xz\end{pmatrix}
=0.\displaystyle\qquad=0.

Given (x:y:z)η(K)(x:y:z)\in\mathcal{H}_{\eta}(K), we have

(z2x2)2+(2xz)2=(z2+x2)2(z^{2}-x^{2})^{2}+(2xz)^{2}=(z^{2}+x^{2})^{2}

and

(c(z2x2)d(2xz))2+(a(z2x2)+b(2xz))2=y2,(-c(z^{2}-x^{2})-d(2xz))^{2}+(a(z^{2}-x^{2})+b(2xz))^{2}=y^{2},

so that Φ(x:y:z)𝒮K×𝒮K\Phi(x:y:z)\in\mathcal{S}_{K}\times\mathcal{S}_{K}. Conversely, given any (α1,α2)Fη(K)(𝒮K×𝒮K)(\alpha_{1},\alpha_{2})\in F_{\eta}(K)\cap(\mathcal{S}_{K}\times\mathcal{S}_{K}), we can write α2=(z2x2:2xz)\alpha_{2}=(z^{2}-x^{2}:2xz) and α1=(z2x2:2xz)\alpha_{1}=(z^{\prime 2}-x^{\prime 2}:2x^{\prime}z^{\prime}) for some x,z,x,zKx,z,x^{\prime},z^{\prime}\in K. The fact that (α1,α2)Fη(K)(\alpha_{1},\alpha_{2})\in F_{\eta}(K) is then equivalent to

(z2x22xz)(abcd)(z2x22xz)=0.\begin{pmatrix}z^{\prime 2}-x^{\prime 2}&2x^{\prime}z^{\prime}\end{pmatrix}\begin{pmatrix}a&b\\ c&d\end{pmatrix}\begin{pmatrix}z^{2}-x^{2}\\ 2xz\end{pmatrix}=0.

This implies that (abcd)(z2x22xz)\begin{pmatrix}a&b\\ c&d\end{pmatrix}\begin{pmatrix}z^{2}-x^{2}\\ 2xz\end{pmatrix} must equal λ(2xzz2x2)\lambda\begin{pmatrix}-2x^{\prime}z^{\prime}\\ z^{\prime 2}-x^{\prime 2}\end{pmatrix} for some λK×\lambda\in K^{\times}. In particular,

N((abcd)(z2x22xz))=λ2(z2+x2)2,N\left(\begin{pmatrix}a&b\\ c&d\end{pmatrix}\begin{pmatrix}z^{2}-x^{2}\\ 2xz\end{pmatrix}\right)=\lambda^{2}(z^{\prime 2}+x^{\prime 2})^{2},

so that (x:y:z)η(K)(x:y:z)\in\mathcal{H}_{\eta}(K) for y=λ(z2+x2)y=\lambda(z^{\prime 2}+x^{\prime 2}). One can then confirm that Φ\Phi maps (x:y:z)(x:y:z) to (α1,α2)(\alpha_{1},\alpha_{2}), using the computation

(cdab)(z2x22xz)\displaystyle\begin{pmatrix}-c&-d\\ a&b\end{pmatrix}\begin{pmatrix}z^{2}-x^{2}\\ 2xz\end{pmatrix} =(0110)(abcd)(z2x22xz)\displaystyle=\begin{pmatrix}0&-1\\ 1&0\end{pmatrix}\begin{pmatrix}a&b\\ c&d\end{pmatrix}\begin{pmatrix}z^{2}-x^{2}\\ 2xz\end{pmatrix}
=λ(0110)(2xzz2x2)\displaystyle=\lambda\begin{pmatrix}0&-1\\ 1&0\end{pmatrix}\begin{pmatrix}-2x^{\prime}z^{\prime}\\ z^{\prime 2}-x^{\prime 2}\end{pmatrix}
=λ(z2x22xz).\displaystyle=-\lambda\begin{pmatrix}z^{\prime 2}-x^{\prime 2}\\ 2x^{\prime}z^{\prime}\end{pmatrix}.

Hence Φ\Phi maps Hη(K){H}_{\eta}(K) surjectively onto Fη(K)(𝒮K×𝒮K)F_{\eta}(K)\cap(\mathcal{S}_{K}\times\mathcal{S}_{K}).

Finally, observe that for each α2𝒮K\alpha_{2}\in\mathcal{S}_{K}, there are two choices for (x:z)1(K)(x:z)\in\mathbb{P}^{1}(K) with (z2x2:2xz)=α2(z^{2}-x^{2}:2xz)=\alpha_{2}, interchanged by the involution (x:z)(z:x)(x:z)\mapsto(-z:x). Once xx and zz are fixed, there are two choices for yy, interchanged by yyy\mapsto-y. Hence Γ\Gamma acts transitively on the fibers of Φ\Phi. ∎

Remark 5.2.

For many rational distance problems, solutions ((u1:v1),(u2:v2))Fη(K)(𝒮K×𝒮K)((u_{1}:v_{1}),(u_{2}:v_{2}))\in F_{\eta}(K)\cap(\mathcal{S}_{K}\times\mathcal{S}_{K}) with u1v1u2v2=0u_{1}v_{1}u_{2}v_{2}=0 will be considered degenerate (as they correspond to rational right triangles with no width). The degenerate locus u1v1u2v2=0u_{1}v_{1}u_{2}v_{2}=0 pulls back to the subvariety 𝒟\mathcal{D}\subseteq\mathcal{H} defined by

(17) 𝒟:xyz(z4x4)(a(z2x2)+b(2xz))(c(z2x2)+d(2xz))=0.\displaystyle\mathcal{D}:xyz(z^{4}-x^{4})(a(z^{2}-x^{2})+b(2xz))(c(z^{2}-x^{2})+d(2xz))=0.

5.2. Density of rational configuration solutions

For any embedding KK\hookrightarrow\mathbb{R}, if η(K)\mathcal{H}_{\eta}(K) is infinite, we can show that Fη(K)(𝒮K×𝒮K)F_{\eta}(K)\cap(\mathcal{S}_{K}\times\mathcal{S}_{K}) is dense in Fη()F_{\eta}(\mathbb{R}). This is a special case of the following result. Let EE denote the Jacobian of HηH_{\eta}.

Lemma 5.3.

Let ηGL2()\eta\in\operatorname{GL}_{2}(\mathbb{R}) and suppose Δ(η)0\Delta(\mathcal{H}_{\eta})\neq 0. Let AHη()A\subseteq{H}_{\eta}(\mathbb{R}) be the image of an infinite subgroup of E()E(\mathbb{R}) under some isomorphism E()Hη()E(\mathbb{R})\cong{H}_{\eta}(\mathbb{R}). Then the image of AA under Φ:HηFη\Phi:{H}_{\eta}\to F_{\eta} (Proposition 5.1) is dense in Fη()F_{\eta}(\mathbb{R}).

Proof.

The (topological) curve Hη(){H}_{\eta}(\mathbb{R}) has two connected components, given by points (x:y:z)(x:y:z) with y>0y>0 and those with y<0y<0 respectively: there is no equivalence between any points with y>0y>0 and points with y<0y<0 because of the weighting on 2{\mathbb{P}^{2}} (Section 3.2), and there are no points with y=0y=0 because Δ(η)0\Delta(\mathcal{H}_{\eta})\neq 0 and 1-1 is not a square in KK. Thus E()E(\mathbb{R}) has structure of a Lie group S1()×/2S^{1}(\mathbb{R})\times\mathbb{Z}/2\mathbb{Z}. Any infinite subgroup of E()E(\mathbb{R}) has dense intersection with the identity component, so AA has dense intersection with one of the components of Hη(){H}_{\eta}(\mathbb{R}).

We can express the map Φ:HηFη{\Phi}:{H}_{\eta}\to{F}_{\eta} as a composition

Hη1Fη(x:y:z)(x:z)((c(z2x2)d(2xz):a(z2x2)+b(2xz)),(z2x2:2xz)).\begin{array}[]{ccccc}{H}_{\eta}&\to&\mathbb{P}^{1}&\to&{F}_{\eta}\\ (x:y:z)&\mapsto&(x:z)&\mapsto&\left(\begin{gathered}(-c(z^{2}-x^{2})-d(2xz):a(z^{2}-x^{2})+b(2xz)),\\ (z^{2}-x^{2}:2xz)\end{gathered}\right).\end{array}

The first map induces a continuous surjection from each component of η()\mathcal{H}_{\eta}(\mathbb{R}) onto 1()\mathbb{P}^{1}(\mathbb{R}), so the image of AA is dense in 1()\mathbb{P}^{1}(\mathbb{R}). The second map induces a continuous surjection 1()Fη()\mathbb{P}^{1}(\mathbb{R})\to{F}_{\eta}(\mathbb{R}), so the image of AA is dense in Fη(){F}_{\eta}(\mathbb{R}). ∎

5.3. Application to three-distance problem

We will use Theorem 1.4 to prove the following statement.

Corollary 5.4.

There exists an infinite collection of rational functions, ρn:𝔸1𝔸2\rho_{n}:\mathbb{A}^{1}_{\mathbb{Q}}\dashrightarrow\mathbb{A}^{2}_{\mathbb{Q}} for nn\in\mathbb{Z}, with the following properties. For all t{0,±1}t\in\mathbb{Q}-\{0,\pm 1\} and all nn\in\mathbb{Z}, if ρn\rho_{n} is defined at tt, then ρn(t)\rho_{n}(t) has rational distance from each of (0,0)(0,0), (0,1)(0,1), and (1,1)(1,1). Further, for each t{0,±1}t\in\mathbb{Q}-\{0,\pm 1\}, there are only finitely many nn\in\mathbb{Z} for which ρn\rho_{n} is not defined at tt, and the set

{ρn(t):n,ρn defined at t}\{\rho_{n}(t):n\in\mathbb{Z},\,\rho_{n}\text{ defined at }t\}

is a dense subset of the line y=2t1t2xy=\frac{2t}{1-t^{2}}x in 2\mathbb{R}^{2}.

Proof.

For the sake of clarity, we begin by proving the weaker result mentioned in the introduction: for each t{0,±1}t\in\mathbb{Q}-\{0,\pm 1\}, the line y=2t1t2xy=\frac{2t}{1-t^{2}}x has a dense set of points that have rational distance from each of (0,0)(0,0), (0,1)(0,1), and (1,1)(1,1). Once this is done, we will explain how the proof can be modified to allow for families of solutions parametrized by tt.

Let t{0,±1}t\in\mathbb{Q}-\{0,\pm 1\}, and set s(t):=12t1t2s(t):=1-\frac{2t}{1-t^{2}}. There is no rational solution to 1=2t1t21=\frac{2t}{1-t^{2}}, so η(t):=(110s(t))\eta(t):=\begin{pmatrix}1&-1\\ 0&s(t)\end{pmatrix} is an element of GL2()\operatorname{GL}_{2}(\mathbb{Q}). We have |s(t)|1|s(t)|\neq 1 because we excluded the case t=0t=0 and there is no rational solution to 2=2t1t22=\frac{2t}{1-t^{2}}. Further, there is no rational solution to |1u42u|=1\left|\frac{1-u^{4}}{2u}\right|=1. Hence, by Theorem 1.4, η(t)()\mathcal{H}_{\eta(t)}(\mathbb{Q}) is infinite. By Lemma 5.3, this implies that the set of ((u1:v1),(u2:v2))𝒮×𝒮((u_{1}:v_{1}),(u_{2}:v_{2}))\in\mathcal{S}\times\mathcal{S} satisfying u1u2+v1v2=u1v2+2t1t2v1v2u_{1}u_{2}+v_{1}v_{2}=u_{1}v_{2}+\frac{2t}{1-t^{2}}v_{1}v_{2} (the defining equation of Fη(t)F_{\eta(t)}) is dense in Fη(t)()F_{\eta(t)}(\mathbb{R}).

Now define the rational function z:Fη(t)\dashedrightarrow𝔸2z:F_{\eta(t)}\dashedrightarrow\mathbb{A}^{2} by

(18) z((u1:v1),(u2:v2)):=((1t2)v1(1t2)u1+2tv1,2tv1(1t2)u1+2tv1).\displaystyle z((u_{1}:v_{1}),(u_{2}:v_{2})):=\left(\frac{(1-t^{2})v_{1}}{(1-t^{2})u_{1}+2tv_{1}},\,\frac{2tv_{1}}{(1-t^{2})u_{1}+2tv_{1}}\right).

The map zz restricts to a homeomorphism

Fη(t)(){((2t:1t2),(1t2:2t))}{(x,y)2:y=2t1t2x},F_{\eta(t)}(\mathbb{R})\setminus\left\{\left((-2t:1-t^{2}),(1-t^{2}:2t)\right)\right\}\to\left\{(x,y)\in\mathbb{R}^{2}:y=\tfrac{2t}{1-t^{2}}x\right\},

So after removing a single point from Fη(t)(K)(𝒮×𝒮)F_{\eta(t)}(K)\cap(\mathcal{S}\times\mathcal{S}), the remainder maps to a dense subset of the line y=2t1t2xy=\frac{2t}{1-t^{2}}x. For each (α1,α2)Fη(t)()(𝒮×𝒮)(\alpha_{1},\alpha_{2})\in F_{\eta(t)}(\mathbb{Q})\cap(\mathcal{S}\times\mathcal{S}) other than ((2t:1t2),(1t2:2t))\left((-2t:1-t^{2}),(1-t^{2}:2t)\right), the point (x,y):=z(α1,α2)(x,y):=z(\alpha_{1},\alpha_{2}) satisfies

x2+y2\displaystyle x^{2}+y^{2} =((1+t2)v1u1(1t2)+2tv1)2,\displaystyle=\left(\frac{(1+t^{2})v_{1}}{u_{1}(1-t^{2})+2tv_{1}}\right)^{2},
x2+(1y)2\displaystyle x^{2}+(1-y)^{2} =((1t2)u1(1t2)+2tv1)2(u12+v12),\displaystyle=\left(\frac{(1-t^{2})}{u_{1}(1-t^{2})+2tv_{1}}\right)^{2}(u_{1}^{2}+v_{1}^{2}),
(1x)2+(1y)2\displaystyle(1-x)^{2}+(1-y)^{2} =((1t2)u1+2t1t2v1v1(1t2)u1+2tv1)2+((1t2)u1(1t2)u1+2tv1)2\displaystyle=\left((1-t^{2})\frac{u_{1}+\frac{2t}{1-t^{2}}v_{1}-v_{1}}{(1-t^{2})u_{1}+2tv_{1}}\right)^{2}+\left(\frac{(1-t^{2})u_{1}}{(1-t^{2})u_{1}+2tv_{1}}\right)^{2}
=((1t2)u1u2((1t2)u1+2tv1)v2)2+((1t2)u1(1t2)u1+2tv1)2\displaystyle=\left(\frac{(1-t^{2})u_{1}u_{2}}{((1-t^{2})u_{1}+2tv_{1})v_{2}}\right)^{2}+\left(\frac{(1-t^{2})u_{1}}{(1-t^{2})u_{1}+2tv_{1}}\right)^{2}
=((1t2)u1((1t2)u1+2tv1)v2)2(u22+v22).\displaystyle=\left(\frac{(1-t^{2})u_{1}}{((1-t^{2})u_{1}+2tv_{1})v_{2}}\right)^{2}(u_{2}^{2}+v_{2}^{2}).

Since α1,α2𝒮\alpha_{1},\alpha_{2}\in\mathcal{S}, these are all squares in ×\mathbb{Q}^{\times}, so this gives a solution to the three-distance problem.

We now return to the problem of producing explicit parametrizations of solutions in terms of tt. For this, note that tη(t)t\mapsto\eta(t) defines a morphism V:=𝔸1{0,±1}GL2V:=\mathbb{A}^{1}-\{0,\pm 1\}\to\operatorname{GL}_{2}. We will define a rational map ρn:V\dashedrightarrow𝔸2\rho_{n}:V\dashedrightarrow\mathbb{A}^{2} by a composition

ρn:VτnE𝜀Φ1×1×Vz𝔸2,\rho_{n}:V\xrightarrow{\tau_{n}}E^{\prime}\xrightarrow{\varepsilon}\mathcal{H}^{\prime}\xrightarrow{\Phi^{\prime}}\mathbb{P}^{1}\times\mathbb{P}^{1}\times V\xrightarrow{z^{\prime}}\mathbb{A}^{2},

where each variety besides 𝔸2\mathbb{A}^{2} is a scheme over VV and each map besides zz^{\prime} is a morphism over VV. We consider each of these maps in turn.

  • Let EE be the subvariety of 2×GL2\mathbb{P}^{2}\times\operatorname{GL}_{2} parametrizing the Jacobian varieties of \mathcal{H} (defined by Eq. 9). Let EE^{\prime} be the fiber product of VV with EE, so that Et=Eη(t)E^{\prime}_{t}=E_{\eta(t)} for all tV()t\in V(\mathbb{Q}). We have a section VEV\to E^{\prime} given by t(1,1)t\mapsto(-1,-1). Using the group law on the generic fiber of EE^{\prime}, define the rational map τn:V\dashedrightarrowE\tau_{n}:V\dashedrightarrow E^{\prime} by the property that τn(t)=n(1,1)Et()\tau_{n}(t)=n(-1,-1)\in E^{\prime}_{t}(\mathbb{Q}) for all tV()t\in V(\mathbb{Q}). The proof of Theorem 1.4 shows that (1,1)(-1,-1) is non-torsion in Et()E^{\prime}_{t}(\mathbb{Q}) for all tV()t\in V(\mathbb{Q}), so for each such tt, the set {τn(t):n}\{\tau_{n}(t):n\in\mathbb{Z}\} is an infinite subgroup of Et()E^{\prime}_{t}(\mathbb{Q}).

  • The fiber product of VV with \mathcal{H} is a one-parameter family \mathcal{H}^{\prime} of curves over VV, with the property that t=η(t)\mathcal{H}^{\prime}_{t}=\mathcal{H}_{\eta(t)}. We have a section VV\to\mathcal{H}^{\prime} given by t(0:1:1)t\mapsto(0:1:1), allowing us to define a birational map ε:E\dashedrightarrow\varepsilon:E^{\prime}\dashedrightarrow\mathcal{H}^{\prime} over \mathbb{Q} sending the zero section of EE^{\prime} to the given section of \mathcal{H}^{\prime}. This map restricts to an isomorphism on all fibers over points in V()V(\mathbb{Q}).

  • The rational map Φ:\dashedrightarrow1×1×V\Phi^{\prime}:\mathcal{H}^{\prime}\dashedrightarrow\mathbb{P}^{1}\times\mathbb{P}^{1}\times V is defined by

    ((x:y:z),t)((d(t)(2xz):(z2x2)(2xz)),(z2x2:2xz),t).((x:y:z),t)\mapsto\left((-d(t)(2xz):(z^{2}-x^{2})-(2xz)),(z^{2}-x^{2}:2xz),\,t\right).

    Note that after restricting to a fiber t\mathcal{H}^{\prime}_{t}, the first two components of Φ\Phi^{\prime} agree with the map Φ:Hη(t)Fη(t)\Phi:{H}_{\eta(t)}\to F_{\eta(t)} (Proposition 5.1). So for any tV()t\in V(\mathbb{Q}), the set

    St:={(Φετn)(t):n}S_{t}:=\{(\Phi^{\prime}\circ\varepsilon\circ\tau_{n})(t):n\in\mathbb{Z}\}

    is contained in Fη(t)()(𝒮×𝒮)×{t}F_{\eta(t)}(\mathbb{Q})\cap(\mathcal{S}\times\mathcal{S})\times\{t\}, and is a dense subset of Fη(t)()×{t}F_{\eta(t)}(\mathbb{R})\times\{t\} by Lemma 5.3.

  • The rational map z:1×1×V\dashedrightarrow𝔸2z^{\prime}:\mathbb{P}^{1}\times\mathbb{P}^{1}\times V\dashedrightarrow\mathbb{A}^{2} is defined on an appropriate dense open subset by

    z((u1:v1),(u2:v2),t)=((1t2)v1(1t2)u1+2tv1,2tv1(1t2)u1+2tv1).z^{\prime}((u_{1}:v_{1}),(u_{2}:v_{2}),t)=\left(\frac{(1-t^{2})v_{1}}{(1-t^{2})u_{1}+2tv_{1}},\,\frac{2tv_{1}}{(1-t^{2})u_{1}+2tv_{1}}\right).

    Note that when restricted to Fη(t)×{t}F_{\eta(t)}\times\{t\}, the map agrees with z:Fη(t)\dashedrightarrow𝔸2z:F_{\eta(t)}\dashedrightarrow\mathbb{A}^{2} defined in Eq. 18. So the same proof as above shows that for each tt, zz^{\prime} maps StS_{t} minus a point to a dense subset of the line y=2t1t2xy=\frac{2t}{1-t^{2}}x consisting of solutions to the three-distance problem.∎

5.4. Special cases

Some rational configuration problems fall under the exceptional cases of Theorem 1.4. We consider a few of these here. Let

𝒮:={uv(u:v)𝒮}\displaystyle\mathcal{S}^{\prime}:=\left\{\frac{u}{v}\in\mathbb{Q}\mid(u:v)\in\mathcal{S}\right\}

be the affine elements of 𝒮\mathcal{S}.

Proposition 5.5.

Let α3𝒮\alpha_{3}\in\mathcal{S}^{\prime}. There exist infinitely many pairs α1,α2𝒮\alpha_{1},\alpha_{2}\in\mathcal{S}^{\prime} such that α1+α2=α3\alpha_{1}+\alpha_{2}=\alpha_{3}.

That is, for any rectangle with rational distances between every two vertices, there are infinitely many ways to split it into two rectangles with rational distances between every two vertices.

Proof.

Let α3=1t22t\alpha_{3}=\frac{1-t^{2}}{2t} for some t{0,±1}t\in\mathbb{Q}-\{0,\pm 1\}. The equation x1+x2α3=0x_{1}+x_{2}-\alpha_{3}=0 defines FηF_{\eta} for η=(011α3)\eta=\begin{pmatrix}0&1\\ 1&-\alpha_{3}\end{pmatrix}, and η\mathcal{H}_{\eta} (which contains the rational point (0:1:1)(0:1:1)) is isomorphic to its Jacobian (as in Eq. 9),

E:y2=x3+((1t22t)2+2)x2+x.E:y^{2}=x^{3}+\left(\left(\frac{1-t^{2}}{2t}\right)^{2}+2\right)x^{2}+x.

We consider EηE_{\eta} as an elliptic curve over the function field (t)\mathbb{Q}(t), and note that E((t))E(\mathbb{Q}(t)) has a rational point P=(t,12(t+1)2)P=\left(t,\frac{1}{2}(t+1)^{2}\right). By computing nPnP for n=1,,12n=1,\ldots,12 and checking that the denominators of the coordinates have no rational roots, we can confirm that PP is non-torsion for all t{0,±1}t\in\mathbb{Q}-\{0,\pm 1\}. Hence η\mathcal{H}_{\eta} has infinitely many rational points, so Fη()(𝒮×𝒮)F_{\eta}(\mathbb{Q})\cap(\mathcal{S}\times\mathcal{S}) is infinite by Proposition 5.1. ∎

This allows us to prove that every rational number can be written as a sum of three elements of 𝒮\mathcal{S}^{\prime} in infinitely many ways; in other words, for any rational t>0t>0, there are infinitely many ways to cut a 1×t1\times t rectangle into three rectangles, each of which has rational distances between every pair of vertices. We also prove that every rational number can be written as a product of three elements of 𝒮\mathcal{S}^{\prime} in infinitely many ways.

Proof of Proposition 1.6.

The equation x1+2x2t=0x_{1}+2x_{2}-t=0 defines FηF_{\eta} for η=(012t)\eta=\begin{pmatrix}0&1\\ 2&-t\end{pmatrix}. Now η\mathcal{H}_{\eta} is isomorphic to η\mathcal{H}_{\eta^{\prime}} for η=(1t/201/2)\eta^{\prime}=\begin{pmatrix}1&-t/2\\ 0&-1/2\end{pmatrix}, and by Theorem 1.4, η()\mathcal{H}_{\eta}(\mathbb{Q}) is infinite for all t0t\neq 0. Hence every nonzero tt\in\mathbb{Q} can be written as α1+α2+α2\alpha_{1}+\alpha_{2}+\alpha_{2} for infinitely many pairs (α1,α2)𝒮2(\alpha_{1},\alpha_{2})\in\mathcal{S}_{\mathbb{Q}}^{2}. The case t=0t=0 follows from Proposition 5.5 because 𝒮\mathcal{S} is closed under negation.

Next we will show that for any t×t\in\mathbb{Q}^{\times}, there exists u{0,±1}u\in\mathbb{Q}-\{0,\pm 1\} such that when s=t(2u1u2)s=-t\left(\frac{2u}{1-u^{2}}\right), the polynomial x1x2+sx_{1}x_{2}+s has infinitely many solutions with x1,x2𝒮x_{1},x_{2}\in\mathcal{S}^{\prime}. Each of these solutions can then be multiplied by 1u22u𝒮\frac{1-u^{2}}{2u}\in\mathcal{S}^{\prime} to exhibit tt as a product of three elements of 𝒮\mathcal{S}^{\prime}.

Let η=(100s)\eta=\begin{pmatrix}1&0\\ 0&s\end{pmatrix}. We consider the elliptic curve

Eη:y2=x(x+1)(x+s2)=x(x+1)(x+t2(2u1u2)2).E_{\eta}:y^{2}=x(x+1)(x+s^{2})=x(x+1)\left(x+t^{2}\left(\frac{2u}{1-u^{2}}\right)^{2}\right).

If we set u=t2+2u=t^{2}+2, then the elliptic curve

y2=x(x+1)(x+t2(2(t2+2)1(t2+2)2)2)y^{2}=x(x+1)\left(x+t^{2}\left(\frac{2(t^{2}+2)}{1-(t^{2}+2)^{2}}\right)^{2}\right)

over \mathbb{Q} has a point

(t2(t2+1)2(t2+2)(t2+3)2,t2(t2+2)(t8+4t6+6t4+8t2+9)(t2+3)3),\left(\frac{t^{2}(t^{2}+1)^{2}(t^{2}+2)}{(t^{2}+3)^{2}},\,\frac{t^{2}(t^{2}+2)(t^{8}+4t^{6}+6t^{4}+8t^{2}+9)}{(t^{2}+3)^{3}}\right),

which has infinite order when t0,±1t\neq 0,\pm 1. So for all t{0,±1}t\in\mathbb{Q}-\{0,\pm 1\}, x1x2t(2u1u2)=0x_{1}x_{2}-t\left(\frac{2u}{1-u^{2}}\right)=0 has infinitely many solutions x1,x2𝒮x_{1},x_{2}\in\mathcal{S}^{\prime}, so tt can be written as a product of three elements of 𝒮\mathcal{S}^{\prime} in infinitely many different ways.

We finally must handle t=±1t=\pm 1. In this case, we can set u=56u=\frac{5}{6}. For η=(100±6011)\eta=\begin{pmatrix}1&0\\ 0&\pm\frac{60}{11}\end{pmatrix} we have the elliptic curve

Eη:y2=x3+(1+(6011)2)x2+(6011)2x,E_{\eta}:y^{2}=x^{3}+\left(1+\left(\frac{60}{11}\right)^{2}\right)x^{2}+\left(\frac{60}{11}\right)^{2}x,

which has a non-torsion point (1211,204121)(-\frac{12}{11},\frac{204}{121}) (in fact Eη()E_{\eta}(\mathbb{Q}) has rank 22). Thus there are infinitely many Pythagorean solutions of x1x26011=0x_{1}x_{2}\mp\frac{60}{11}=0, allowing us to write ±1\pm 1 as a product α1α21160\alpha_{1}\alpha_{2}\frac{11}{60} of three elements of 𝒮\mathcal{S}^{\prime} in infinitely many ways. ∎

Remark 5.6.

The substitution s=t2+2s=t^{2}+2 was found essentially by trial and error, guided by inspiration from a MathOverflow answer by Siksek [1] describing how to find a positive rank subfamily of the family y2=x(x+1)(x+(1ss)2)y^{2}=x(x+1)(x+(\frac{1-s}{s})^{2}), and from Naskrkecki [14] who used a similar method to find a positive rank subfamily of the curve y2=x(x1)(x(2s1s2)2)y^{2}=x(x-1)\left(x-\left(\frac{2s}{1-s^{2}}\right)^{2}\right).

For the t=1t=1 case, the existence of a solution to α1α2α3=1\alpha_{1}\alpha_{2}\alpha_{3}=1 is equivalent to the existence of a body cuboid Section 2.1. The existence of a body cuboid with edge lengths (240,117,44)(240,117,44) leads to the choice of uu.

By Proposition 5.5, every element of 𝒮\mathcal{S}^{\prime} can be written as a sum of two elements of 𝒮\mathcal{S}^{\prime} in infinitely many ways, but we have no comparable result for products. A natural question then is to determine which rational numbers tt can be written as a product of two elements of 𝒮\mathcal{S}^{\prime} in infinitely many ways. This line of inquiry is explored in more depth in [12].

References

  • [1] Siksek (https://mathoverflow.net/users/4140/siksek) “The rank of a class elliptic curves” URL:https://mathoverflow.net/q/63856 (version: 2011-05-03), MathOverflow URL: https://mathoverflow.net/q/63856
  • [2] Sang Yook An et al. “Jacobians of Genus One Curves” In Journal of Number Theory 90.2, 2001, pp. 304–315 DOI: https://doi.org/10.1006/jnth.2000.2632
  • [3] T.. Berry “Points at rational distance from the corners of a unit square” In Annali della Scuola Normale Superiore di Pisa - Classe di Scienze Ser. 4, 17.4 Scuola normale superiore, 1990, pp. 505–529 URL: http://www.numdam.org/item/ASNSP_1990_4_17_4_505_0/
  • [4] Manjul Bhargava, John Cremona and Tom Fisher “The proportion of genus one curves over \mathbb{Q} defined by a binary quartic that everywhere locally have a point” In Int. J. Number Theory 17.4, 2021, pp. 903–923 DOI: 10.1142/S1793042121500147
  • [5] V. Chandrasekar “The congruent number problem” In Resonance 3, 1998, pp. 33–45
  • [6] Keith Conrad “The congruent number problem” Accessed: 20210-11-01, https://kconrad.math.uconn.edu/blurbs/ugradnumthy/congnumber.pdf
  • [7] Harold Davenport “Multiplicative number theory” Revised and with a preface by Hugh L. Montgomery 74, Graduate Texts in Mathematics Springer-Verlag, New York, 2000, pp. xiv+177
  • [8] P.. Gallagher “The large sieve and probabilistic Galois theory” In Analytic number theory (Proc. Sympos. Pure Math., Vol. XXIV, St. Louis Univ., St. Louis, Mo., 1972) Vol. XXIV, Proc. Sympos. Pure Math. Amer. Math. Soc., Providence, RI, 1973, pp. 91–101
  • [9] Richard K. Guy “Unsolved problems in number theory”, Problem Books in Mathematics Springer-Verlag, New York, 2004, pp. xviii+437 DOI: 10.1007/978-0-387-26677-0
  • [10] Lorenz Halbeisen and Norbert Hungerbühler “Pairing pythagorean pairs” In Journal of Number Theory, 2021 DOI: https://doi.org/10.1016/j.jnt.2021.07.002
  • [11] Dimitris Koukoulopoulos “The distribution of prime numbers” 203, Graduate Studies in Mathematics American Mathematical Society, Providence, RI, [2019] ©2019, pp. xii + 356 DOI: 10.1090/gsm/203
  • [12] Jonathan Love “Root numbers of a family of elliptic curves and two applications”, 2023 arXiv:2201.04708 [math.NT]
  • [13] R Luijk “On Perfect Cuboids”, 2000
  • [14] Bartosz Naskręcki “Mordell-Weil ranks of families of elliptic curves associated to Pythagorean triples” In Acta Arith. 160.2, 2013, pp. 159–183 DOI: 10.4064/aa160-2-5
  • [15] Joseph H. Silverman “The arithmetic of elliptic curves” 106, Graduate Texts in Mathematics Springer, Dordrecht, 2009, pp. xx+513 DOI: 10.1007/978-0-387-09494-6
  • [16] J.. Tunnell “A classical Diophantine problem and modular forms of weight 3/23/2 In Invent. Math. 72.2, 1983, pp. 323–334 DOI: 10.1007/BF01389327
  • [17] J.-L. Waldspurger “Sur les coefficients de Fourier des formes modulaires de poids demi-entier” In J. Math. Pures Appl. (9) 60.4, 1981, pp. 375–484
  • [18] André Weil “Remarques sur un mémoire d’Hermite” In Arch. Math. (Basel) 5, 1954, pp. 197–202 DOI: 10.1007/BF01899338