Rational configuration problems and a family of curves
Abstract.
Given , we consider the number of rational points on the genus one curve
We prove that the set of for which has density zero, and that if a rational point exists, then is infinite unless a certain explicit polynomial in vanishes.
Curves of the form naturally occur in the study of configurations of points in with rational distances between them. As one example demonstrating this framework, we prove that if a line through the origin in passes through a rational point on the unit circle, then it contains a dense set of points such that the distances from to each of the three points , , and 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.
1. Introduction
1.1. A family of curves
Fix , and let be the curve defined by
(1) |
in the weighted projective plane where have degree , respectively. Rational points on this curve correspond to vectors such that both and 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 for which has zero, finitely many, or infinitely many rational points.
First, we show that for most values of , the curve has no rational points.
Theorem 1.1.
Let be the set of with such that is nonempty for all . Then for some constant ,
The proof is given in Section 4. Note that for any positive integer , so by clearing denominators, every 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 denote the set of such that
has a point for all . Then
We see that the subfamily 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 for which does contain a rational point. In this case we have a stronger classification. Let denote the transpose of .
Proposition 1.3.
Suppose (Eq. 1) has a rational point. If is a scalar matrix, then for some and is a union of two rational conics,
Otherwise is isomorphic to
for some with and .
A proof is given in Section 3.4 using the fact that the isomorphism type of is invariant under acting on the left and right of by elements of the orthogonal group . An explicit change of variables expressing in terms of and the rational point is given by Lemma 3.2.
Theorem 1.4.
Let with and . The point is non-torsion if and only if , , and .
In particular, for most of the values such that is nonempty, 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 or for , there exists a dense set of points with rational distance from each of , , and .
In fact we prove a stronger result: there is an infinite collection of curves in the plane such that the intersection points of the curves with any fixed line (for ) gives a dense set of solutions to the three-distance problem within the given line (Corollary 5.4).
Even in the cases where is torsion, there are still several cases in which we can prove has positive rank. We discuss these in more depth in Section 5.4, but note one special case here. Let
denote the set of slopes of rational right triangles (including negatives and zero).
Proposition 1.6.
For all , the equations and each have an infinite set of solutions with .
See Section 5.4 for a proof.
1.2. Rational configuration problems
Given a finite simple graph , an embedding is a rational configuration if the distance is rational for all . 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 . Find a point such that has rational distance to , , , and . (A traveller is going from to , but has to take a detour to stop at the -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 such that the distance to each of , , , and is rational.
-
•
“Square three-distance:” Find a point such that the distance to each of , , and is rational.
-
•
“Rectangle four-distance:” Find and a point such that the distance to each of , , , and is rational.
-
•
“Rational distances under Möbius transformation:” Fix with . Find such that and both have rational distance from .
Configuration | Graph | Equation | Solutions given by |
---|---|---|---|
Adjacent rectangles |
( for all : Proposition 5.5) |
||
Detour () |
( if and : Theorem 1.4) |
||
Perfect cuboid | Unknown | ||
Body cuboid | |||
(Square) Four-distance | Unknown | ||
(Square) three-distance |
( for all : Corollary 1.5) |
||
(Rectangle) four-distance | |||
Rational distances under Möbius transformation, |
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
so that whenever , not both zero, are the legs of a (possibly degenerate) rational right triangle, the slope of the triangle is in . Then for distinct , the distance between and is rational if and only if the line between and has slope in . Using this observation, we can parametrize solutions to many rational configuration problems by finding elements of 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 ; this implies there exist such that
are all perfect squares. If we set , then the polynomial constraints above are equivalent to requiring
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 to a single polynomial in multiple variables that is linear in each variable.
Proposition 1.8.
Let , and let be the curve in defined by
There is a degree morphism inducing a surjection
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 . 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 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 , in particular showing that the isomorphism type of 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 are the edge lengths of a body cuboid, then , , and are all perfect squares; the first two conditions say that and the third is equivalent to requiring .
For each fixed , the values satisfying 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 , they associate solutions satisfying to rational points on the elliptic curve
(2) |
Proposition 1.8 and Proposition 1.3 recovers this classification. They show that there is a subgroup of isomorphic to 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 has positive rank. In this case they call 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 with rational distance to , , and . The coordinates are not a priori assumed to be rational, but since , , and must all be rational, the differences and must also be rational, so in fact . We can then scale by an element of so that , and a solution to the square three-distance problem is equivalent to the existence of satsifying .
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 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) |
The “congruent number problem” is to determine whether a given 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 , any solution to Eq. 3 corresponds to a rational point on an elliptic curve over defined by
(4) |
There are “degenerate points” in that do not correspond to solutions; it can be shown that the set of degenerate points equals the torsion subgroup of . Thus is a congruent number if and only if has positive rank. A formula due to Tunnell can be used to determine whether the analytic rank of 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 is a congruent number if and only if and give a solution to
(5) |
Both and can be represented as the image of under the image of a degree rational map . The curve comes equipped with a degree rational map to the variety defined by , and non-degenerate points in map to solutions to Eq. 5. This is directly analogous to the relation between 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 is determined by the class of in , while the isomorphism class of is determined by the class of a corresponding matrix in a double quotient of .
-
•
Existence of rational points. Every determines an elliptic curve , which has a rational point. By contrast, the genus one curves 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 . For rational configuration problems, however, even if is isomorphic to an elliptic curve (Proposition 1.3), the degenerate points in 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 are quadratic twists of the curve , and are therefore all isomorphic over . 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 -function of an elliptic curve with that of each of its quadratic twists. By contrast, the curves do not have constant -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 be a field of characteristic not equal to , in which is not a square; later we will restrict to , but many of our results hold in more generality. Throughout this paper, all schemes will be defined over unless otherwise indicated, and if and are schemes then .
Throughout, will denote the projective line over , while will denote a weighted projective space over , where the variables have weights , respectively. We use the notation and to denote elements of and , respectively. That is, for we have
and for we have
Let denote the algebraic group of invertible matrices, with identity element . Given a matrix , its transpose will be denoted . Let denote the orthogonal group of matrices, that is, the algebraic subgroup of defined by the condition that is in if and only if .
3.2. Definition of and basic properties
Using the coordinates on , define the variety by the equation
(6) |
If we let be defined by , then this can equivalently be written
(7) |
This variety comes equipped with a morphism , which equips with the structure of a flat family of curves. Given , is the fiber of over .
The generic fiber of is a genus one hyperelliptic curve over the function field , with discriminant
(8) |
The Jacobian variety of this curve is an elliptic curve over , which by classical invariant theory (see for example [18, 2]) has a model
(9) |
We have two commuting involutions on as a scheme over , given by
(10) |
generating a Klein four-group
(11) |
acting on . (Note that is an involution because has weight , and so .)
3.3. Double cosets and reduction
We show that the isomorphism class of is invariant on double cosets in
and use this to show that has a -point if and only if is in the same double coset as for some .
Lemma 3.1.
Let . If , then there is an isomorphism over that commutes with the action of .
Proof.
Let , where and . Write , where , , and . There exists so that and . Then for any ,
Thus the map
defines an isomorphism , and the involutions and are preserved. ∎
Given , suppose is in the same double coset as an element of the form . We have , so by Lemma 3.1, we can conclude that is nonempty. The following lemma gives us the converse result: if is nonempty then is in the same double-coset as a matrix of the form .
Lemma 3.2.
Let . Suppose there is a point . Define
(12) |
Then .
Proof.
Suppose . If , then , contradicting the assumption that is not a square in . Hence , and likewise . But this implies , which contradicts the fact that .
If , then a similar argument shows that we must have
But this implies that the nonzero vector is in the kernel of , contradicting the assumption that . Hence .
Since and , the matrices
are both well-defined elements of . We can check by direct computation that . ∎
3.4. Isomorphism classes of fibers
The curve is singular when the discriminant (Eq. 8) vanishes. Since for all and does not contain a square root of , this can only occur if and , or if and . One of these two conditions holds if and only if and ; thus the singular fibers are exactly those with
In this case reduces to the form
If then splits into two conics, . If is not a square in , then there are no solutions in .
If has a rational point and the discriminant (Eq. 8) does not vanish at , then is isomorphic to its Jacobian. Using Lemma 3.2 and Eq. 9, we can conclude that is isomorphic to
for some ; the non-vanishing of the discriminant says that and . This completes the proof of Proposition 1.3.
3.5. Nonsingular fibers with a rational point
We now restrict our attention to in order to prove Theorem 1.4, which we recall for convenience.
See 1.4
Proof.
Assume is torsion in . Since is positive on an open interval around , there exists for which does not contain any point of the form . Thus has two components, with on the non-identity component and on the identity component. This shows is not a multiple of in , and hence cannot have odd order. By Mazur’s classification of torsion subgroups, we can conclude that if is torsion then its order must be an even number at most . If has order then the only possibility for the torsion subgroup of is , so that is the unique element of order . This implies , which again leads to a contradiction when we consider the component group of .
We can conclude that if is torsion, it must have order . For each such , let denote the -th division polynomial on ; this is a polynomial with the property that for if and only if for some (see for instance [15, Exercise 3.7]). We compute the division polynomial , and determine all possible such that .
-
•
We have , so has order if and only if .
-
•
We have
The last factor is a sum of non-negative terms, including at least one positive term because . Hence has order if and only if .
-
•
The quotient of by factors into two irreducible polynomials in . The first factor is , which is positive for all . The second factor is
Considering this as a quadratic polynomial in , the discriminant is equal to
In order for to be rational (let alone ), this discriminant must equal a rational square. Thus we consider rational points on the curve defined by . There are eight rational points : two at infinity, as well as , , and . Using the Weierstrass form for we can confirm that has no other rational points, so the only possibilities for are . If then we have , yielding no rational solutions. If then we have , contradicting .
-
•
The quotient of by factors into three irreducible polynomials in . The first two factors are and ; these each have infinitely many rational solutions. The third factor is positive for all .
-
•
If we eliminate common factors with and from , we are left with three irreducible polynomials in . The first factor is
Considered as a quadratic in , the discriminant is , which is a square if and only if for some . Plugging this in and solving for , we find that either
For the first option, we obtain if and only if is a nonzero square. The only rational points on the elliptic curve are the point at infinity and , so there is no for which is rational. For the second option, we obtain if and only if is a nonzero square, and by the same reasoning there is no such . Hence this factor is nonzero for all .
The second factor is obtained from the first by , so it has no rational solutions either. The third factor is positive for all .
To summarize, we obtain the following possibilities:
-
•
has order if and only if ;
-
•
has order if and only if ;
-
•
has order if and only if .
-
•
There are no values of for which 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 -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 . By contrast, each local obstruction in Theorem 1.2 only affects of all curves. At a high level, the difference in behavior between the two families stems from the fact that diverges but converges.
4.1. Local Obstructions
Given a ring we use to denote the ring of matrices over . For primes we define
Thus counts the set of with entries of absolute value at most such that for all primes (note that there is never a real obstruction: for all .)
The main contribution to will come from the following constraint.
Lemma 4.1.
Let be an odd prime and . Suppose , and , , and are all nonzero mod . Then is nonempty if and only if at least one of or is a square modulo .
Proof.
Assume has a point ; without loss of generality we can assume that are in and at least one of is in . Reducing modulo we have
(13) |
If , then
is a square. On the other hand, suppose . Since and at least one of is nonzero we must have , so
is a square.
Conversely, if is a nonzero square mod then Eq. 13 clearly has solutions with ; these are smooth points on so they lift to points on . If is a nonzero square mod , then it is a square in , so there exist with . This implies , so is a point in . ∎
In light of the above, set
where denotes the Legendre symbol. If by abuse of notation we associate with its preimage in under reduction mod , we have
(14) |
This follows from Lemma 4.1: note that the Legendre symbol conditions in the definition of force to be nonzero mod .
Lemma 4.2.
We have .
Proof.
Let . Since is not a square, we have . Note that is a square in if and only if is in the image of the map given by . We have if and only if , and so there are values in the range of (with the sign depending on whether or not is a square mod ). Hence and must each be one of the values not in the range of . Since there are choices for , there are choices for each of and , and the value of is fixed, we obtain the desired count. ∎
The following result is not used in the sequel, but justifies the claim that is the main contribution to .
Lemma 4.3.
Let denote the image of under reduction mod . We have
Proof.
We produce a collection of pairs of linear equations over with the property that every element of satisfies one of these pairs of equations. Let . If
then for any lift , the discriminant of (Eq. 8) is not divisible by , and so is nonempty by the Hasse-Weil bound. As these are all smooth points, they lift to points in . Thus is not in .
We can therefore assume that exactly one of the following constraints holds:
-
(a)
and ,
-
(b)
and ,
-
(c)
, , and ,
-
(d)
, and .
In case (a) or (c), must satisfy one of the following pairs of linear equations over :
Now consider case (b). If or , then using and we can conclude that
for some satisfying . If on the other hand and are both nonzero, then is not in by Lemma 4.1.
Finally we consider case (d). We assume and as the other case is similar. Then for some with . The reduction of modulo is given by
where
If the discriminant of is nonzero, then the equation defines a smooth projective conic over , and there must exist at least one point on this conic with . Then defines a smooth point in , which lifts to a point in . Hence is not in . Computing the discriminant of , we find that elements in in this case must satisfy
∎
4.2. Proof of Theorem 1.1
The author would like to thank Sun-Kai Leung for suggesting the following proof.
Recall that is the set of all with nonzero determinant, entries having absolute value at most , and with for all . Set
By Eq. 14 we can see that , and so it suffices to find an upper bound for .
Set
where is the Möbius function (so if is squarefree and otherwise). Applying the -dimensional large sieve [8, Lemma B], we obtain
where means that for some positive constant we have for sufficiently large . Theorem 1.1 follows immediately from this bound after applying the following lemma.
Lemma 4.4.
We have
5. Applications to Rational Distance Problems
5.1. From to rational configurations
We return temporarily to the more general setting of a field in which is not a square. Let . In Eq. 6, we defined the variety
where . We also define the subvariety of (with coordinates by
As with , there is a projection map to giving the structure of a flat family of curves, with denoting the fiber over . We define
Proposition 5.1.
There is a morphism over defined by sending to
(15) |
Further, induces a bijection between the set of -orbits in and the set .
Proof.
Notice that the equation defining can be written
(16) |
The morphism is well-defined because
Given , we have
and
so that . Conversely, given any , we can write and for some . The fact that is then equivalent to
This implies that must equal for some . In particular,
so that for . One can then confirm that maps to , using the computation
Hence maps surjectively onto .
Finally, observe that for each , there are two choices for with , interchanged by the involution . Once and are fixed, there are two choices for , interchanged by . Hence acts transitively on the fibers of . ∎
Remark 5.2.
For many rational distance problems, solutions with will be considered degenerate (as they correspond to rational right triangles with no width). The degenerate locus pulls back to the subvariety defined by
(17) |
5.2. Density of rational configuration solutions
For any embedding , if is infinite, we can show that is dense in . This is a special case of the following result. Let denote the Jacobian of .
Lemma 5.3.
Let and suppose . Let be the image of an infinite subgroup of under some isomorphism . Then the image of under (Proposition 5.1) is dense in .
Proof.
The (topological) curve has two connected components, given by points with and those with respectively: there is no equivalence between any points with and points with because of the weighting on (Section 3.2), and there are no points with because and is not a square in . Thus has structure of a Lie group . Any infinite subgroup of has dense intersection with the identity component, so has dense intersection with one of the components of .
We can express the map as a composition
The first map induces a continuous surjection from each component of onto , so the image of is dense in . The second map induces a continuous surjection , so the image of is dense in . ∎
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, for , with the following properties. For all and all , if is defined at , then has rational distance from each of , , and . Further, for each , there are only finitely many for which is not defined at , and the set
is a dense subset of the line in .
Proof.
For the sake of clarity, we begin by proving the weaker result mentioned in the introduction: for each , the line has a dense set of points that have rational distance from each of , , and . Once this is done, we will explain how the proof can be modified to allow for families of solutions parametrized by .
Let , and set . There is no rational solution to , so is an element of . We have because we excluded the case and there is no rational solution to . Further, there is no rational solution to . Hence, by Theorem 1.4, is infinite. By Lemma 5.3, this implies that the set of satisfying (the defining equation of ) is dense in .
Now define the rational function by
(18) |
The map restricts to a homeomorphism
So after removing a single point from , the remainder maps to a dense subset of the line . For each other than , the point satisfies
Since , these are all squares in , 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 . For this, note that defines a morphism . We will define a rational map by a composition
where each variety besides is a scheme over and each map besides is a morphism over . We consider each of these maps in turn.
-
•
Let be the subvariety of parametrizing the Jacobian varieties of (defined by Eq. 9). Let be the fiber product of with , so that for all . We have a section given by . Using the group law on the generic fiber of , define the rational map by the property that for all . The proof of Theorem 1.4 shows that is non-torsion in for all , so for each such , the set is an infinite subgroup of .
-
•
The fiber product of with is a one-parameter family of curves over , with the property that . We have a section given by , allowing us to define a birational map over sending the zero section of to the given section of . This map restricts to an isomorphism on all fibers over points in .
-
•
The rational map is defined by
Note that after restricting to a fiber , the first two components of agree with the map (Proposition 5.1). So for any , the set
is contained in , and is a dense subset of by Lemma 5.3.
-
•
The rational map is defined on an appropriate dense open subset by
Note that when restricted to , the map agrees with defined in Eq. 18. So the same proof as above shows that for each , maps minus a point to a dense subset of the line 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
be the affine elements of .
Proposition 5.5.
Let . There exist infinitely many pairs such that .
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 for some . The equation defines for , and (which contains the rational point ) is isomorphic to its Jacobian (as in Eq. 9),
We consider as an elliptic curve over the function field , and note that has a rational point . By computing for and checking that the denominators of the coordinates have no rational roots, we can confirm that is non-torsion for all . Hence has infinitely many rational points, so 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 in infinitely many ways; in other words, for any rational , there are infinitely many ways to cut a 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 in infinitely many ways.
Proof of Proposition 1.6.
The equation defines for . Now is isomorphic to for , and by Theorem 1.4, is infinite for all . Hence every nonzero can be written as for infinitely many pairs . The case follows from Proposition 5.5 because is closed under negation.
Next we will show that for any , there exists such that when , the polynomial has infinitely many solutions with . Each of these solutions can then be multiplied by to exhibit as a product of three elements of .
Let . We consider the elliptic curve
If we set , then the elliptic curve
over has a point
which has infinite order when . So for all , has infinitely many solutions , so can be written as a product of three elements of in infinitely many different ways.
We finally must handle . In this case, we can set . For we have the elliptic curve
which has a non-torsion point (in fact has rank ). Thus there are infinitely many Pythagorean solutions of , allowing us to write as a product of three elements of in infinitely many ways. ∎
Remark 5.6.
The substitution 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 , and from Naskrkecki [14] who used a similar method to find a positive rank subfamily of the curve .
For the case, the existence of a solution to is equivalent to the existence of a body cuboid Section 2.1. The existence of a body cuboid with edge lengths leads to the choice of .
By Proposition 5.5, every element of can be written as a sum of two elements of in infinitely many ways, but we have no comparable result for products. A natural question then is to determine which rational numbers can be written as a product of two elements of 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 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 ” 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