Classification of rational angles in plane lattices

By Roberto Dvornicich, Francesco Veneziano, and Umberto Zannier

Abstract

This paper is concerned with configurations of points in a plane lattice which determine angles that are rational multiples of . We shall study how many such angles may appear in a given lattice and in which positions, allowing the lattice to vary arbitrarily.

This classification turns out to be much less simple than could be expected, leading even to parametrizations involving rational points on certain algebraic curves of positive genus.

1. Introduction

The present paper will be concerned with lattices in and in fact with the angles determined by an ordered triple of distinct points , , varying through the lattice. Our leading issue will involve angles which are rational multiples of , which we will call rational angles for brevity. These angles of course appear in regular polygons, in tessellations of the plane, and in other similar issues, and it seems to us interesting to study in which lattices these angles appear and how.

Suppose , are points in the lattice , and let be the angle formed between the rays and (where denotes the origin); one might wonder when this angle is a rational multiple of . It turns out that if is a rational multiple of , then is one of , , , , as shown by J. S. Calcut in Reference Cal09 (see the Appendix to this Introduction for a proof simpler than Calcut’s). Analogous properties hold for the Eisenstein lattice generated by the vertices of an equilateral triangle in the plane: the rational angles which occur are precisely the integer multiples of .

Let us recall a few historical precedents of similar problems.

Considering the simplest lattice , E. Lucas Reference Luc78 in 1878 answered in the negative the rather natural question of whether points , , can determine an equilateral triangle. In fact, it is not very difficult to show that no angle , with , , , can be equal to . (This will follow as an extremely special case of our analysis, but we anticipate it in the Appendix to the Introduction (§ 1.1) with a short and simple argument.)

In 1946, in a related direction, W. Scherrer Reference Sch46 considered all regular polygons with all vertices in a given arbitrary lattice and proved that the polygons which may occur are precisely those with , , or sides. (See the Appendix to the Introduction (§ 1.2) for an account of Scherrer’s nice proof with some comments.)

These results prompt the question: what happens when we consider other plane⁠Footnote1 lattices?

1

Of course one could consider analogous problems in higher dimensions, however we expect the complexity will increase greatly. There is some work in progress by Poonen et al. which appears to be not unrelated.

Namely, for an arbitrary lattice , how can we describe all rational angles determined by three points in ? The most ambitious goal would be to obtain a complete classification, in some sense. In particular, we can list the following issues in this direction:

1.

Which rational angles can occur in a given plane lattice ? Intuitively, we should not expect many such angles when the lattice is fixed.

2.

In which positions can they occur? Namely, which triples of points , , can determine such a given angle? Note that it is clearly sufficient to consider the case when is the origin and that we may replace , by any two points in the lattice on the lines , . So we shall consider an angle determined by the origin and two lines passing through the origin and another lattice point.

3.

Adopting a reciprocal point of view, how can we classify plane lattices according to the structure of all rational angles determined by their points?

We shall give below a more precise meaning to these questions; for instance, we shall study plane lattices according to how many rational angles (with vertex in the origin) may occur and in which geometric configurations; for instance it is relevant which pairs of angles have a side in common.

It is clear that for any arbitrarily prescribed angle we are able to find a plane lattice in which this angle appears as the angle determined by three points of the lattice. It is easily seen that even a second (rational) angle can be prescribed arbitrarily. On adding further conditions on the rationality of other angles and their relative positions, the arithmetical information deduced from these conditions increases and imposes severe restrictions on the lattice, which can lead to a classification. Let us observe that the variables in our problem are

(i)

the lattices,

(ii)

the (rational) angles,

(iii)

the lines (through lattice points) determining these angles.

We shall see that for three or more rational angles in a lattice we have roughly the following possibilities:

(A)

either the angles belong to a certain finite set which is described in Section 5,

(B)

or the lattice belongs to one of finitely many families of lattices (and corresponding angles) which are described in Section 6.4.

More precisely, with a notation that will be properly introduced later, we will prove in Sections 5 and 6 the following theorem.

Theorem 1.1.

Let be a space of one of the three types , , or . Then either is homotetic to one of the spaces described in Section 6.4, or any rational angle of , with , coprime, satisfies

Furthermore, when the angles lie in the mentioned finite set and are fixed, then either

(A1)

the lattice belongs to a certain well-described family (called CM in this paper); moreover, for a fixed lattice in this family the points determining the angles are parameterized by the rational points in finitely many rational curves (see Section 3.1 and equation Equation 3.2); or

(A2)

both the lattice and the vertices are parametrized by the rational points on finitely many suitable algebraic curves; or

(A3)

we have a finite set of lattices-vertices.

We shall analyze all of these situations, also looking at the maximal number of different configurations of rational angles which can exist in a lattice, in the sense of item 2 above. We shall prove that indeed this number is finite apart from well-described families.

As it is natural to expect, this study involves algebraic relations among roots of unity, a topic which falls into a well-established theory. However, even with these tools at our disposal, our problem turns out to be more delicate than can be expected at first sight, in that some surprising phenomena will appear on the way to a complete picture. For instance we shall see that some of the curves mentioned in (A2) have positive genus, in fact we have examples of genus up to .

In this paper we shall give an “almost” complete classification, in the sense alluded to above. In particular, this shall be complete regarding the sets of angles which may appear. Concerning the classification of specific configurations of angles, we shall confine them to a certain explicit finite list. In the present paper we shall treat in full detail only a part of them, in particular when the rational angles considered share a side.

We also give a full discussion of a case in which configurations arise that correspond to rational points on a certain elliptic curve (of positive rank). (See § 8)

We postpone to a second paper the discussion of the remaining few cases, which need not be treated differently, but which are computationally rather complicated due to the combinatorics of the configurations. We note that in view of Faltings’ theorem, there are only finitely many rational points on the curves of genus which appear; however no known method is available to calculate these points (only to estimate their number), so we shall not be able a priori to be fully explicit in these cases.

A concrete example

As an illustration of the kind of problems one is faced with, we show here an example of a curve of high genus which arises from the general treatment of lattices with three nonadjacent rational angles.

Fixing the amplitudes of the angles to be , , , we are led to study the rational solutions to the system , where

These equations define a variety in which consists of the four lines and of an irreducible curve of genus 5.

It is worth noting that the curve contains some trivial rational points such as, for instance, , but also nontrivial rational points such as , which corresponds to the lattice generated by 1 and , with

In addition to the angle spanned by 1 and , there are two more rational angles to be found between lines through the origin passing through elements of the lattice. The angle spanned by and has an amplitude of , and the one spanned by and has an amplitude of . Figure 1 illustrates these angles.

As will be clear soon, these problems are better analyzed not in plane lattices but by viewing the real plane as the complex field and considering, in place of the lattices, the -vector subspaces of generated by the lattices.

Organization of the paper

The paper will be organized roughly as follows.

In § 2 we shall introduce in detail our issues, giving also some notation and terminology.

Section 3 will be subdivided in several parts.

In the first two parts we shall find general equations corresponding to the configurations that we want to study (as described in the previous section).

The third part will be devoted to obtaining a linear relation in roots of unity, with rational coefficients, after elimination from the equations obtained formerly (depending on the configuration).

In the fourth part we will study more in depth the elimination carried out in the previous part, in order to prove some geometric results.

The fifth part will recall known results from the theory of linear relations in roots of unity, which shall be used to treat the mentioned equations.

In § 4 we shall study spaces with special symmetries (for instance those which correspond to imaginary quadratic fields).

In § 5 we shall prove the bound appearing in Theorem 1.1. For this we shall use among other things results of the geometry of numbers (not applied to the original lattice however, but to a certain region in dimension ).

In § 6 we classify the finite number of continuous families of lattices which escape the previous theorem.

In § 7 we study the configurations where there exist four nonproportional points , , , of the lattice such that every angle is rational. In this analysis we shall meet two rather surprising geometric shapes (which we shall call dodecagonal).

Section 8 will contain the complete study of an elliptic curve such that its rational points correspond to lattices with three nonadjacent rational angles. (The group of rational points will be found to be isomorphic to .)

Appendix to the Introduction

In this short appendix we give a couple of simple proofs related to the known results cited in the Introduction. These will be largely superseded by the rest of the paper, but due to their simplicity we have decided to offer independent short arguments for them.

1.1. Rational angles in the Gaussian lattice

By Gaussian lattice we mean as usual the lattice . For our issue of angles, it is equivalent to consider the -vector space generated by the lattice, i.e., . This space has the special feature of being a field, i.e., the Gaussian field , which makes our problem quite a bit simpler. In fact, let be a rational angle occurring in or , where , are coprime nonzero integers, and let . That occurs as an angle in means that there are nonzero points , such that where . Conjugating and dividing, we obtain , where . So the root of unity lies in , and it is well known that contains only the fourth roots of unity, so has order dividing . A direct argument is to observe that is an algebraic integer so must be of the shape with integers , , and being a root of unity this forces , so is a power of and , as required.

In the converse direction, observe that indeed is determined by the three points , , of , (i.e., the complex numbers , , ).

A similar argument holds for the Eisenstein lattice, which again generates over a field, namely the field generated by the roots of unity of order ; hence can be a twelfth root of unity.

These special lattices will be treated in greater generality in § 4.2.

1.2. On Sherrer’s proof

As mentioned above, Sherrer proved that the only regular polygons with vertices in some lattice are the -gons for , , . The idea of his proof is by descent: if , …, are the vertices of an -gon in a given lattice , then are again the vertices of a regular -gon in ; however, if , the new sides are smaller whence iterating the procedure, we obtain a contradiction (and similarly with a little variation for ). In fact, this argument amounts to the following: we assume as before that the lattice is inside and (after an affine map) that contains all the th roots of unity where . By appealing to the fact, already known to Gauss, that the degree is equal to , we may already conclude that , i.e., , , , , . But we may also avoid using such result, on observing that for , the ring contains nonzero elements of arbitrarily small complex absolute value, so cannot be contained in a lattice, which is discrete. To justify the assertion, consider the elements . If , they have absolute value which is decreasing to since . For , we may instead consider in place of .

2. Terminology and notation

We identify the euclidean plane with . We call a rational angle in an ordered couple of distinct lines through the origin such that the measure of the euclidean angle between them is a rational multiple of . We say that two points , such that determine (or form) a rational angle if the lines do (i.e., if the argument is a rational multiple of ). With a slight abuse of notation we write the angle as .

Let be a lattice. Given a rational angle determined by elements of , many more pairs in can be trivially found (by multiplication by integers) that determine the same angle. Therefore we prefer to tensor the whole lattice by and study angles in the tensored space. With this point of view, we can say that when we draw rational angles in we extend the sides indefinitely and we are not concerned with which points of actually meet the sides.

In this setting, the objects that we will study are two-dimensional -vector spaces that contain two -linearly independent vectors. These are precisely the sets obtained after tensoring a plane lattice by . From now on, unless otherwise stated, we will refer to these sets simply as spaces.

Any angle-preserving transformation of that sends the origin to itself will clearly establish a bijection between the rational angles of a space and the rational angles of its image.

These transformations are generated by complex homotheties of the form for a fixed and by complex conjugation. For this reason we will say that two spaces , are homothetic (and we write ) if they are sent one to the other by a homothety, and we will say that they are equivalent (and we write ) if is homothetic to or to its complex conjugate .

Remark 2.1.

Clearly every space is homothetic to a space containing 1. Given two spaces and , it is easy to see that if and only if , where , in which case the homothetic coefficient is given by .

The same condition can also be expressed by saying that if and only if there is a -linear dependence between , , , . In fact, from a linear relation is immediately obtained, and vice versa. From such a linear relation we obtain a matrix which must be invertible because .

In conclusion, we can always assume up to homothety that with . Up to equivalence, we can additionally assume that .

For ease of notation, we will often write .

In every space , given a rational angle , we can obtain other rational angles by swapping them. We call the angles thus obtained equivalent, and we are not concerned with them.

Given two adjacent rational angles and , we see immediately that is also a rational angle. For this reason it is more convenient to consider sets of adjacent angles as a single geometrical configuration, rather than as independent angles. This point of view is also supported by the shape of the equations that describe these cases. Therefore we call a rational -tuple a set of vectors such that is a rational angle for all .

According to this definition, a rational -tuple can be identified with an -element subset of . This point of view however is not particularly useful when writing up the equations that describe the configuration.

Remark 2.2.

If a rational -tuple and a rational -tuple are not disjoint, then their union is still a rational -tuple for some , .

From the shape of equations Equation 3.5 and Equation 3.11 below, which describe rational -tuples, it is clear that angles which belong to a rational -tuple containing 1 are qualitatively different from angles which do not belong to such an -tuple. In fact, an angle in a rational -tuple containing 1 leads to an equation of degree 1 in , while angles in rational -tuples not containing 1 lead to equations of degree 2 in .

In light of these considerations, we will say that a space is of type if it contains a rational -tuple; we extend this notation additively, by saying that is of type if it contains a rational -tuple and a disjoint rational -tuple, and so on. There is an obvious partial order on the possible types, and we say that has an exact type , if this type is maximal for . We will characterize in § 4.2 the spaces for which such a maximal type exists.

We remark that spaces of type correspond to triangles in which all angles are rational multiples of .

We will denote by the set of all roots of unity.

3. Equations

3.1. The equation of a rational angle

Let , and let . Let , , , be such that is a rational angle, which is to say, there exists a root of unity such that the ratio is real. Setting this ratio equal to its conjugate leads to the equation

This shows that . Equation Equation 3.1 is bihomogeneous of degree 1 in both , and , , so it defines a curve . Setting

we can rearrange equation Equation 3.1 as

The curve has genus 0 and it is irreducible. In fact, it has bidegree and, if it were not irreducible, it would have two components of bidegrees and . This happens if and only if , and a small computation shows that this happens if and only if .

There is a bijection between and the set of rational angles in with . However, is in general not defined over , but only over the field . This fact plays a role in the characterization of spaces with infinitely many rational angles.

3.2. The equations of a rational -tuple containing

Let be a space with a rational -tuple (). Up to homothety of the space and equivalence of rational angles, we can assume that , and that the -tuple is given by , where the are distinct rational numbers different from 0. We write , with and . Similarly, let for , …, .

In particular we have that for , …, . Equating these numbers and their conjugates, we can write

which we can solve for , , or obtaining

where we have set for , …, .

3.3. The equations of a rational -tuple not containing

Let and , with and . Let be a rational -tuple () which does not contain vectors proportional to 1 or . By rescaling we can assume for , …, and the distinct nonzero rational numbers. Let such that for , …, . Then we have

where we have set . Writing these equations as quadratic equations in , we get

for , …, .

3.4. Equations for the three main cases

We begin by observing that, if is a space of type , it is homothetic to with .

If a space has two nonequivalent rational angles, we have seen in Sections 3.2 and 3.3 that we can derive equations for with coefficients in cyclotomic fields; therefore .

If we have two independent such equations, we can eliminate and obtain one equation in roots of unity with rational coefficients. We shall then apply the results of Section 3.5 with the aim of bounding the degree of the roots of unity intervening in the equation, outside of certain families which admit a parametrization.

When eliminating , we could have two equations of shape Equation 3.5 (which amounts to having a rational 4-tuple), or one of shape Equation 3.5 and one of shape Equation 3.11 (which amounts to one rational triple and one more angle not adjacent to it) or two equations of shape Equation 3.11, for which we need three rational angles pairwise nonadjacent. These are the cases that we denote as type , type and type and they will constitute the main equations, whose solutions we shall seek by means of Theorem 3.2. For ease of notation, we change variables here with respect to those considered in Sections 3.2 and 3.3.

Case . In this case we have two equations of the shape Equation 3.5. Eliminating gives

where , are distinct rational numbers different from 0 and , , are distinct roots of unity different from .

Now we set , , and , and obtain

Case . In this case we have one equation of the shape Equation 3.5 and one of the shape Equation 3.11. Eliminating gives

where , , are rational, with and . Furthermore if one of , is equal to or the configuration reduces to that of case , so we may also assume that , , are all distinct and nonzero. The roots of unity , , are different from , and .

By setting , , and , , , we write

Case . have two equations of the shape Equation 3.11. Eliminating gives an expression in the four rational parameters , , , and the three roots of unity , , , where , , , , . Furthermore, if the set contains fewer than five distinct elements, we are in a case treated previously, so we may also assume that , , , are all distinct and nonzero.

By setting , , and , , , , we obtain an unwieldy polynomial equation

of degree in each of the three variables. For ease of reference, we list in Table 1 the coefficients of each term appearing in .

Remark 3.1.

Notice that, while the shape of equation Equation 3.15 is not perfectly symmetric in the three angles, by applying an homothety we can always easily permute the three angles. This amounts to saying that the equation is stable by the substitution which sends the triples , , with angles (squared) to , , , with angles (squared) .

We observe that the nondegeneracy conditions coming from the geometry of the problem ensure that equations Equation 3.13, Equation 3.14, Equation 3.15 do not identically vanish.

In cases and , given a solution of equations Equation 3.13 and Equation 3.14, one can easily obtain the corresponding value of and geometric configuration from equation Equation 3.5: we have

It is easy to check directly that, under the required conditions, this value of is never real.

In case we have two equations of degree for , and it is not always possible to go back from a solution of Equation 3.15 to a geometrical configuration.

Given a solution of Equation 3.15, if the quantity is different from 0, then the two quadratic equations for are independent, and it is possible to solve for obtaining

If instead the equality holds, either is the only common solution or the two equations are proportional, and in this case two values of are found. We can study fully the cases in which this happens.

Setting and eliminating it, we see that the two equations are proportional if the following unit equation is satisfied

Writing , , and dividing by gives

This is a rational combination of four cosines of rational multiples of , and all such combinations have been classified in Reference CJ76, Theorem 7.

3.5. Equations in roots of unity

The study of linear relations among roots of unity goes back to long ago. For instance in 1877 Gordan Reference Gor77 studied the equation

with , , rational angles, with the purpose of classifying the finite subgroups of . The matter was considered by several other authors, also studying polygons with rational angles and rational side-lengths. Among these authors we point out Mann Reference Man65 and Conway and Jones Reference CJ76. These last authors described these issues, important for the present paper, as “trigonometric diophantine equations”. We do not pause further on other references, but we remark that this problem is linked to the conjectures of Lang on torsion points on subvarieties of tori.

For simplicity we state a theorem of Reference CJ76, which we will apply to the equations in roots of unity that we have obtained before.

Theorem 3.2 (Conway and Jones).

Let

be a linear relation with rational coefficients between roots of unity , normalized with . Then either there is a vanishing subsum or the common order of the is a squarefree number satisfying

A generalization of this theorem, with a different proof, was given in Reference DZ00, and a version which takes into account reductions modulo prime numbers in Reference DZ02.

4. Spaces with special symmetries

We study now some special classes of spaces which are stable under some angle-preserving transformation. They are relevant to our program because the presence of such symmetries can lead to a richer set of rational angles.

4.1. Spaces with

The following lemma characterizes, up to homothety, the spaces which are fixed by complex conjugation.

Lemma 4.1.

Let be a space. The following conditions are equivalent:

(i)

is homothetic to with .

(ii)

is homothetic to with , a purely imaginary number.

(iii)

is homothetic to a space with .

Proof.

We have that

and if and only if is purely imaginary and nonzero; see Figure 2.

Clear.

Let . This means that and for some matrix , and .

The eigenvalues of can only be . If they were equal, then would either be of infinite order or be equal to , but this is impossible because otherwise , would be both real or both purely imaginary. Therefore has distinct eigenvalues and , and it can be diagonalized over , that is to say that with and . Now we have .

It seems natural to compare the three properties (i), (ii), (iii) with the condition . It is obvious that (iii) implies this condition. However the converse implication does not hold; see Section 4.3, especially Lemma 4.4, for this.

4.2. CM spaces

We say that a space has complex multiplication (CM) if there is a such that the multiplication by sends to itself. It is easy to see that this happens if and only if with imaginary quadratic, and in this case the homothetic coefficient can be taken as any element in .

This shows that if the space is stabilized by a nontrivial homothety there are infinitely many other nontrivial homotheties which stabilize it, and the image of any rational angle under any such homothety is again a rational angle.

The following theorem summarizes the situation and fully describes the set of rational angles in a CM space, up to equivalence of angles and the action of the space on itself by multiplication.

Theorem 4.2.

Let be a space.

(i)

is CM if and only if for a squarefree .

(ii)

For a squarefree , the rational angles in are, up to equivalence, precisely those of the form with .

(iii)

The rational angles in are, up to equivalence, precisely those of the form with and .

(iv)

The rational angles in are, up to equivalence, precisely those of the form with and , , , , , where .

Proof.

In proving part (i) we can assume up to homothety (which preserves both conditions) that with . As seen is section 1,

for a if and only if the homothety with coefficient sends to itself. If is quadratic irrational, then the coefficients of its minimal polynomial provide the suitable , , , . Vice versa, if such a matrix exists, we get immediately a quadratic polynomial satisfied by .

Let us now prove part (ii). Let be a rational angle. By considering the angle , we can reduce to the case . Then either is a rational multiple of , which is what we need to prove, or form a rational triple. Now we use the notation of Section 3.2. By equation Equation 3.3 we see that is a root of unity in , and therefore it must be . This means that lies on the imaginary axis, so it must indeed be a rational multiple of .

The proof of parts (iii) and (iv) is analogous. As before, must be a fourth (resp., sixth) root of unity, therefore lies on one of the lines whose angle with the real axis is an integral multiple of (resp., ). The points , , (resp., , , , , ) lie on these lines, and any other such point must be a rational multiple of one of them.

We remark that if is obtained as for a lattice , the condition that is a CM space in our sense is equivalent to saying that the elliptic curve has complex multiplication.

We have seen that CM spaces contain infinitely many rational angles. In fact they are the only spaces with this property.

Theorem 4.3.

Let be a non-CM space. Then has only finitely many rational angles.

Proof.

Up to homothety, we can assume that for some . We can also assume that is an algebraic number, otherwise, as remarked at the beginning of Section 3.4, the space contains, up to equivalence, only one rational angle. The number field contains finitely many roots of unity; therefore it is enough to show that for every fixed root of unity such that , the curve defined by equation Equation 3.2 has only finitely many rational points.

With the notation of Section 3.1, if the three coefficients , , are not all rationals, then there exists a Galois automorphism such that . In this case, the set is contained in the intersection of two distinct irreducible curves of bidegree in and therefore it contains at most two elements.

In the case that , , , the space is CM. In fact and , and this implies that ; by Theorem 4.2(i), this is equivalent to being CM.

We remark that the arguments of this proof allow one to bound the number of rational angles in in terms of . We will see in Section 5 how this bound can be made independent of .

We notice also that equation Equation 3.15, in the case of a CM space not homothetic to or reduces to , which defines a surface in .

4.3. Spaces with

More in general, any reflection preserves angles so we can look at spaces which are stable by any reflection (not just complex conjugation).

Lemma 4.4.

Let be a space. The following conditions are equivalent:

(i)

;

(ii)

is homothetic to with .

Proof.

We have that for some and that , so that . Now let such that ; such a exists because contains two -linearly independent vectors. Then we have that and .

Both properties are invariant by homothety, so it is enough to show that when . Indeed

We will see later in Section 7.4 some relevant nontrivial examples of spaces with this property.

We remark that if is obtained as for a lattice , the condition that is equivalent to saying that the elliptic curve has a real -invariant (this property is stable for complex homotheties, and in particular depends only on the space ).

Example 4.5.

Consider a space , with a transcendental such that . We will show that , which obviously satisfies the conditions of Lemma 4.4, does not satisfy the conditions of Lemma 4.1.

Suppose that is homothetic to a space with (as in condition (i) of Lemma 4.1). Then it follows that

with , , , rationals and transcendental. The conditions and easily imply, writing , that

which implies and .

But this yields , and since 2 is not a square, we find that , which is impossible.

5. Finiteness of angles in lattices outside special families

Let us consider the equations obtained in Section 3.4. They are of the form

where and the are homogeneous polynomials of degree 4 (in the most general case ) in the four variables , , , , and we seek solutions , , and , , , distinct nonzero integers (cases and are of the same form, only involving fewer terms, fewer variables, and with coefficients of smaller degree).

Let us fix a solution , , , , , , , and let be the minimal common order of the roots of unity , , . We will argue now about the factorization of , showing that either is bounded by an absolute constant or the solution belongs to a parametric family of solutions corresponding to a translate of an algebraic subgroup of contained inside the variety defined by equation Equation 5.1.

5.1. Odd primes appearing with exponent

Let be an odd prime dividing exactly. Let be a primitive th root of unity, and let us write , , , with , , roots of unity of order coprime with . If we denote by the scalar product of and as vectors of , we can write the monomial as with a root of unity of order prime with , and the equation can be rewritten as

It follows that the quantities

for , …, must all be equal because the fields generated by and by the ’s are linearly disjoint.

Equation Equation 5.1 has only 27 terms. This implies that, for all primes bigger than 27, the coefficients are all equal to . We consider only this case.

Not all entries of are multiples of , otherwise would not be the exact order of , , . Therefore the lattice is a lattice of volume .

By classical results in the geometry of numbers, namely the exact value of Hermite’s constant⁠Footnote2 in dimension , we obtain that there exists a of norm bounded as .

2

This result goes back to Gauss through the theory of arithmetical reduction of ternary quadratic forms; see Reference Cas71 for a thorough treatment.

We thank Davide Lombardo for suggesting the use of the exact value of Hermite’s constant in place of Minkowski’s theorem: this leads to a considerable numerical improvement.

Let us write

Let , be two vectors intervening in the same . Then we have that

So either or , and then

Notice that either or , in which case we can divide by 2 and replace with . In both cases we obtain

Then if we have that for every , intervening in the same . Let us now consider a new triple , where is an unknown. If we substitute in Equation 5.1 and collect the powers of as in Equation 5.2, we obtain

and we can collect the powers of for some exponents . But the ’s are all , and so is a solution identically in .

5.2. Primes appearing with exponent at least

Let us fix a solution , , , , , , , let be the minimal common order of the roots of unity , , , and let be a prime power dividing exactly. Let be a primitive th root of unity, and let us write , , , with roots of unity of order coprime with . As before, we can write the monomial as with a root of unity of order prime with , and the equation can be rewritten as

The exponents in the sums on the right are all multiples of , so those powers of are th roots of unity, but the degree of over the field generated by the and the th roots of unity is exactly , which implies that

for all , …, . From now on, we can argue as we did previously in the case of .

Not all entries of are multiples of , otherwise would not be the exact order of , , , therefore the lattice is a lattice of volume . As before, there exists a of norm bounded as .

Let us write

If in any of the sums Equation 5.4 two different , appear, we have that

So either or

If is odd, we can argue again that the only case in which has norm greater than 3 is when it is equal to , in which case it can be replaced by its half, and we obtain

If , we can only use that , which gives

Therefore if (or for ), we have that for every , intervening in the sums Equation 5.4. Let us now consider a new triple , , , where is an unknown. If we substitute in Equation 5.3, we obtain

and we can collect the powers of for some exponents . But we argued before that the ’s are all , and so is a solution identically in .

We have thus shown that every solution of Equation 5.1 either belongs to a parametric family given by , which represents the units in a translate of an algebraic subgroup of , or the common order is a divisor of

Combined with the proof of Theorem 4.3 this implies that, outside the parametric families just mentioned (which will be described in the next section), the number of rational angles in non-CM spaces is at most .

6. Parametric families of solutions

In this section we study which translates of algebraic subgroups of are contained in the variety defined by Equation 5.1. These are the parametric families of solutions which escape the analysis carried out in the previous section.

Suppose then that we have such a solution. This amounts to setting

where , , , is a parameter, , , are not all constant in , so we may assume , , not all and coprime.

6.1. Case

Let us assume for now that we are in the most general case , that is to say that , , , are all distinct and nonzero and , , ; in this case we may assume , , by replacing with and possibly exchanging the role of , or , .

6.1.1. If , , are all positive

In this case the term in is the one of highest degree (as polynomial in ) among those appearing in Table 1, so if Equation 5.1 is satisfied identically in its coefficient must be equal to 0. However this coefficient is , which is not .

6.1.2. If and , are positive and distinct

In this case the terms , , are those of highest degree. The leading term in is equal to , which tells us that, for the equation to be satisfied identically in , we must have and either or . If this holds, the term of highest degree is either the one in or the one in , but setting either coefficient equal to implies that both and hold. Under this further assumption, the whole polynomial reduces to , which can’t be identically 0 in if and are distinct.

6.1.3. If and

Arguing as before we see that, for the term of degree 4 to be , we must have and one of or . The part of degree 3 is now given by the two terms in and together; the coefficient to be set equal to 0 is given by

If , then the term in degree three vanishes, and looking at the terms of lower degree we reach what is indeed a parametric solution of Equation 5.1:

If , then we have

so .

Setting implies , which was discussed above, while implies , which in turn reduces the full equation to

which is not the case.

6.1.4. If and

In this case, looking at the coefficient of the term in gives

where the second factor is equivalent to the first after exchanging every one of , , , with the reciprocals of , , , , respectively. The unit equation

doesn’t have any one-term subsum equal to . Two-term subsums equal to are easily excluded, except possibly if

hold. The second equation implies that and then in turn the first one gives (the two signs being chosen independently). Of these four alternatives, three lead immediately to a contradiction, and we are left with the conditions , , which imply and . But under these conditions the full equation reduces to , which cannot be equal to .

We are left then with solutions of a four-term equation in units without subsums equal to . Therefore , must be sixth roots of unity, and by direct inspection one finds three solutions (up to Galois action and exchanging the role of , ). Each of these solutions, when plugged back into Equation 5.1, gives a nonvanishing term in degree 1.

Since we are assuming we are in the most general case , we can freely permute the angles , , by changing the base of our lattice and obtaining a new equation of the shape Equation 5.1, which is also identically satisfied. Then it is easy to see that we can always reduce up to homothety to one of the four cases discussed above.

6.2. Case

The equation is preserved by the transformations

This allows us to assume , , and .

6.2.1. If , , and

In this case the term of highest degree is , whose coefficient is , which is different from .

6.2.2. If , , and

In this case the terms of highest degree are and , and setting the coefficient equal to leads to

This implies , but so we must have and . With this substitution, equation Equation 5.1 becomes (after cancelling a factor )

Let us assume first that . In this case we see that the term of highest degree is either (if ) or (if ). Their coefficients are and , respectively, so in either case they are not .

If instead the coefficient to be set equal to is . This implies , which is impossible because in one case we get and in the other .

6.2.3. If and

In this case the terms of highest degree are and , so we get , therefore and . Substituting back into Equation 5.1 gives

Now if , the dominant term is either (if ) or (if ), and their coefficients are and , respectively, which are nonvanishing. If instead , the coefficient to be set equal to is . This implies , which gives (because , ), so , which is a contradiction.

6.2.4. If and

The terms of highest degree in are , , , . Setting the coefficient equal to gives

This is a four-term unit equation. After normalizing the equation dividing by , we need to study the vanishing subsums. The only nontrivial case is given by and , which gives a vanishing subsum. Substituting back into Equation 5.1 gives a nonvanishing coefficient either in degree 2 or in degree 1.

A simple computer check finds nine nontrivial solutions among the sixth roots of unity, but none of them leads to solutions identically in .

6.2.5. If , and

The terms of highest degree are and , so we get , so and . Substituting back into Equation 5.1 gives

Let us assume first that . In this case we see that the term of highest degree is either (if ) or (if ). Their coefficients are and , respectively, so in either case they are not , because . If instead the coefficient to be set equal to is . This implies , which is impossible because in one case we get and in the other .

6.2.6. If and

In this case the coefficient of the term of degree in factors as

It is enough to set the first factor equal to , as the second one is completely analogous (as seen by the transformation , , ). This leads to a three-term unit equation. It is readily seen that there are no vanishing subsums, and a simple computer check finds no nondegenerate solution.

6.2.7. If and

In this case the terms of highest degree are , , , , so we get the unit equation

We need to study the vanishing subsums. The only nontrivial case is given by and , which indeed gives a vanishing subsum. Substituting back into Equation 5.1 gives a nonvanishing coefficient in degree 1. A simple computer check finds nine nontrivial solutions among the sixth roots of unity, but none of them leads to solutions identically in .

6.3. Case

If we are in case we notice that the symmetries of the equation allow us to permute freely the angles , , . Furthermore, up to homotheties, we can assume that the three exponents , , are nonnegative, thanks to the substitution , , , , .

6.3.1. ,

In this case, setting the coefficients of the terms of degree and one equal to we obtain the system of equations

If we add them and divide by , we get

which is a contradiction.

6.3.2. If

In this case, the term with the highest degree is , whose coefficient is , which does not vanish.

6.3.3. If

In this case the terms of highest degree are and , and setting the coefficient equal to 0 leads to

which implies and . Substituting back into Equation 3.13 gives

and indeed setting

gives a parametric solution of Equation 5.1.

6.3.4.

Setting the coefficients of the terms in and equal to leads to

multiplying the first equation by and adding them leads to

which would imply either or ; this is a contradiction.

6.4. Infinite families

We have found the infinite families Equation 6.1 and Equation 6.2. We can now understand them better.

The family Equation 6.1 is of type whenever the parameter is a root of unity. Its feature is the presence of a right angle; in fact the spaces of this family all belong to the self-conjugated spaces studied in Section 4.1. Up to homothety they form a family parametrized by a rational number , and a root of unity , with a purely imaginary root of

as illustrated in Figure 3.

If or , the type of the resulting space becomes , and we find the second parametric family Equation 6.2, which will be studied in full detail in the next section, as part of the complete description of all spaces of type . Up to homothety, for a root of unity , is given by

as illustrated in Figure 4.

Theorem 6.1.

Every non-CM space has at most different rational angles.

Proof.

Let us consider a space and assume that it has three nonequivalent rational angles with arguments three roots of unity , , of order greater than . Therefore, by the content of Section 5, this space is homothetic to one of the families described above. However this is impossible, because it would imply that one of , , is equal to . This shows that, in total, in a fixed space only different roots of unity can occur as arguments of a rational angle. It was already remarked in the proof of Theorem 4.3 that, for a fixed root of unity and a fixed non-CM space, there are at most two nonequivalent rational angles of that argument, and this completes the proof.

7. Spaces with a rational 4-tuple

7.1. Rectangular and superrectangular spaces

We saw before that the spaces which are invariant under complex conjugation can be characterized up to homothety as those generated by a of norm 1. Among those spaces, the subfamily of those for which is a root of unity are especially relevant with respect to the rational angles.

Lemma 7.1.

Let be a space. The following conditions are equivalent:

(i)

is homothetic to with ,

(ii)

contains a rational triple with two perpendicular vectors,

(iii)

contains a rational -tuple with two perpendicular vectors.

Proof.

The equivalence between (i) and (ii) is proved as in Lemma 4.1, and (iii)(ii) is trivial, so only (ii)(iii) needs to be shown:

After applying an homothety and multiplying the vectors in the triple by suitable rational numbers, we can assume that the rational triple is , , for some purely imaginary number . We see immediately that the stability of under complex conjugation allows us to add the vector to the triple and obtain a rational quadruple.

We define a space as rectangular if it contains a rational angle of , and as superrectangular if it satisfies the conditions of Lemma 7.1.

By definition every superrectangular space satisfies the hypotheses of Lemma 4.1, whose points (i) and (ii) provide two different parametrizations of superrectangular spaces.

Looking for a purely imaginary , with the notation in Section 3.2, we have , , , and equation Equation 3.5 gives

and the 4-tuple is given by .

Choosing a root of unity as a generator instead, we get , , , , and the 4-tuple is .

7.1.1. Rational -tuples in superrectangular spaces

Let us determine when the rational -tuple of a superrectangular space can be extended to a rational -tuple. Let be a root of unity, let and be a rational -tuple, with , , always with the notation of Section 3.2. Then equation Equation 3.5 gives

so we get the unit equation

No subsum can vanish, because , and ; therefore by Theorem 3.2 the only solutions are to be found among sixth roots of unity, and this case has been already discussed in Section 4.2. This proves that the only superrectangular spaces whose -tuple can be extended to a -tuple are those homothetic to .

7.1.2. Additional angles in superrectangular spaces

Let us now check which superrectangular spaces have additional rational angles that are not part of a rational -tuple containing 1.

Using the same notation for the space , we seek a rational angle . Then equation Equation 3.11 gives us

with , distinct rationals, different from , , .

We are again in the position of using Theorem 3.2. To cut the number of cases to check, we see that the equation does not change if we swap and and invert . We also see that if , then has degree , and we get one of the two CM superrectangular spaces and , which were already discussed.

A computer search does not find any solution with common order a divisor of 30 but not of 6 (if the common order is a divisor of 6, we are again in the case of ). We are left with examining all possible subsums.

The only coefficient that might vanish is , but its vanishing implies .

Of the two-term subsums, seven directly imply that or are equal to ; two imply that ; the remaining six lead to .

Of the ten pairs of vanishing three-term subsums, nine directly imply that and the last one that .

Therefore we can sum up these computations in the following statement:

Theorem 7.2.

A superrectangular space not homothetic to or has exact type , i.e., it has only one rational -tuple up to equivalence and no other rational angle.

7.2. The general case

We have seen that every superrectangular space is of type . Let us now show that, with only finitely many exceptions, every space of type is a superrectangular space.

Let be a space with a rational 4-tuple given by . By the computations in Section 3.2 we have

Eliminating , we have

with , , distinct roots of unity and , distinct rational numbers. In order to show that the space is superrectangular, it is enough to show that one of the or a ratio is equal to .

We can rewrite this equation as

We can now apply Theorem 3.2 and conclude that either there is a vanishing subsum, or the common order of , , is a divisor of 30.

A computer search shows that there are no solutions with common order a divisor of 30 which do not belong to the family of superrectangular spaces. Therefore we are left with searching solutions with vanishing subsums.

7.3. Vanishing subsums in Equation 7.1

If there is a vanishing subsum in Equation 7.1, then there is a vanishing subsum of minimal length at most .

7.3.1. One-term subsums

If a vanishing subsum involves only one term, then the coefficient of the term must be 0, which is forbidden because , are nonzero and distinct.

7.3.2. Three-term subsums

There are 20 three-term subsums, which get paired in ten systems of two three-term linear equations. By direct examination, applying again Theorem 3.2, one sees that any solution in roots of unity leads to a variable or ratio of two variables being equal to , or to solutions where , , have common order a divisor of 6, which have been already discarded.

7.3.3. Two-term subsums

There are 15 two-term subsums. By direct inspection one checks that, after setting them equal to , 12 of them immediately imply that one variable or a ratio of two variables is equal to . The remaining three, which are those obtained by pairing terms with the same coefficient, correspond the relations

Each of these relations reduces Equation 7.1 to a four-term equation. Precisely,

We can apply Theorem 3.2 again, to find that solutions with vanishing subsums lead again to variables or ratios of variables being equal to , while the solutions without vanishing subsums are found with , , of common order a divisor of , which implies that the common order of , , is a divisor of .

7.4. Dodecagonal spaces

There are indeed solutions of common order 12. Up to homotheties, exchanging the roles of the vectors in the rational -tuple and acting with Galois automorphisms we find two spaces.

Let be a primitive twelfth root of unity.

The first space is given by

as illustrated in Figure 5. The second is given by

as illustrated in Figure 6.

It is easy to check that , , and are not -linearly dependent, and thus that is not homothetic to .

It is maybe more surprising that for , and every Galois automorphism .

Figure 5.
\renewcommand{\arraystretch}{1} \setlength{\unitlength}{1.0pt} \begin{tikzpicture}[scale=4] \coordinate(O) at (0,0); \node(pol) [draw, thick, blue!90!black,rotate=90,minimum size=8cm,regular polygon, regular polygon sides=12, rotate=195] at (0,0) {}; \coordinate(C) at (pol.corner 7); \foreach\n[count=\nu from 0, remember=\n as \lastn, evaluate={\nu+\lastn}] in {1,2,...,12} \node[anchor=\n*(360/12)+180]at(pol.corner \n){$\zeta^{\nu}$}; \draw[thin, name path=d15] (pol.corner 1)--(pol.corner 5); \draw[thin, name path=d27] (pol.corner 2)--(pol.corner 7); \draw[thin, name path=d310] (pol.corner 3)--(pol.corner 10); \draw[thin, name path=d412] (pol.corner 4)--(pol.corner 12); \draw[thin, name path=d47] (pol.corner 4)--(pol.corner 7); \draw[thin, name path=d612] (pol.corner 6)--(pol.corner 12); \path[name intersections={of=d15 and d27,by=A}]; \path[name intersections={of=d47 and d612,by=B}]; \node[circle, fill=red, inner sep=0pt, minimum size=4pt,label=45:$\tau_1$] at (A) {}; \node[circle, fill=red, inner sep=0pt, minimum size=4pt, label=-90:$O$] at (O) {}; \draw[red, very thick] (O)--(A); \draw[red, very thick] (A)--(B); \draw[red, very thick] (B)--(C); \draw[red, very thick] (C)--(O); \end{tikzpicture}

In order to check this, let us fix as a basis for ; in this basis, any Galois automorphism acts by exchanging the sign of some coordinates. Given two elements , and expressing in this basis the determinant of the matrix with column vectors , , , , one can check explicitly that it depends only on the squares of the coordinates of and , and thus that it is invariant by the action of the Galois group.

In particular this shows that the dodecagonal spaces satisfy the symmetry considered in Section 4.3.

7.4.1. Additional angles in dodecagonal spaces

It is now quite easy to check that in these dodecagonal spaces the rational -tuple that defines them cannot be extended to a rational -tuple. With a little more (computational) work, we also see that dodecagonal spaces do not contain any additional rational angle. Indeed it is enough to check which rational angles are contained in the spaces with , and the ones defined in the previous section. By the usual Equation 3.7, this amounts to finding all solutions of

with , and distinct, and with different from 1. But we see immediately that lies in , the twelfth cyclotomic field, so for , …, . By writing both sides of the equation in terms of a basis, it is easy to see that the only solutions are found when , for , or , for .

The computations of this section can be summarized in the following theorem:

Theorem 7.3.

Let be a space with a rational quadruple. Then is homothetic to either a superrectangular space or one of the two dodecagonal spaces.

Table 2 summarizes the classification obtained so far.

Remark 7.4 (An application to euclidean geometry).

It is worth noting that, as a consequence of the classification of all spaces of type obtained in this section, it can be shown that the red parallelogram in Figure 5 is the only parallelogram with the properties that:

(1)

all angles determined by sides and diagonals are rational multiples of ,

(2)

it is neither a rectangle nor a rhombus.

Quadrilaterals with property (1) have already been considered in the literature (see for example Reference Rig78) and are related to intersecting triples of diagonals in regular polygons, a topic fully analyzed in Reference PR98.

A related result that arises from the argument concerns rational products of tangents of rational angles, in a way similar but distinct from the study carried out in Reference Mye93.

8. Example of spaces corresponding to rational points on a curve genus

As an example of the possible phenomena that can appear in the more general case , we show here an infinite family of spaces parametrized by the rational points on an elliptic curve of rank over .

Setting , , in equation Equation 3.15, we obtain

Let us take one of the two factors multiplying and set it to . Any rational solution of the system

with , , , distinct and nonzero gives a solution to equation Equation 3.15, and thus a space of type . The system Equation 8.2 defines a variety in , which is the union of the two lines and and an irreducible curve of genus 1.

After eliminating and applying the transformation

this curve has equation

It is possible to put this plane curve in Weierstrass form, sending at infinity the point . Under the transformation

we obtain the Weierstrass form

This elliptic curve has -invariant 128. It has a -torsion point . The rank of the Mordell–Weil group is , with generator .

This curve has infinitely many rational points, and each quadruple provides a value of such that the space is of type . Let us see now that the set of spaces so obtained is infinite also considering them up to equivalence.

The value of , expressed in the original coordinates , , , , is given by

If , , , , we have that , and we can consider its Galois conjugates over . Let us call these four values , , , (corresponding, in the same order, to the identity and the Galois automorphisms fixing , , ).

Let us recall that the cross-ratio of four complex numbers , , , is the rational function

For every Möbius transformation we have

If and are homothetic spaces, we have that for some , and therefore ; similarly, if , then , and the cross-ratio is invariant under such a permutation of the variables.

Any commutes with Galois automorphisms, so we have that , , can be expressed as rational functions of , , , , and so can the cross-ratio . Computing its expression, we obtain

This defines a rational function on the projective curve , and by what we argued above we see that, if , are two rational points that give two equivalent spaces, then . Clearly the generic fibre of the function is finite, and this immediately proves our claim, that the rational points on give rise to infinitely many pairwise nonequivalent spaces.

Notice that this happens because of the special shape of the function ; for a “general” rational function , the equation would define in a reducible curve whose components other than the diagonal would normally have genus greater than . However for the function in question, the equation cuts in six irreducible curves of genus 1.

An even clearer picture of the geometry of this problem is obtained by putting in the form Equation 8.3. In the variables, the function is given by

Appendix A. The irreducibility of the surface defined by equation Equation 3.15

For fixed , , , equation Equation 3.15 defined a surface , whose rational points correspond to lattices of type in which the three rational angles have fixed arguments. In this appendix we study more in depth its geometric properties.

A motivation for this study, other than its interest on its own, comes from the fact that we can reduce this problem of rational points on a surface to a problem of rational points on curves, arguing in a way similar to Section 3.1, with the important distinction that here the space is not fixed.

In fact, assume that is not defined over . Then there is a Galois automorphism such that . If is irreducible, then all rational points on lie in , which is a curve.

The aim of this appendix is to study the surface , proving that is irreducible unless , and that is not defined over unless , , are either three fourth roots of unity or three sixth roots of unity.

A.1. Elimination of two quadratic equations

In order to prove some geometrical properties of the variety defined by equation Equation 3.15, we study here in general the elimination of one variable from two quadratic equations.

Let

be two equations, where for the moment , , , , are elements of a field. In our application, these two equations will be of the form Equation 3.11. On subtracting we obtain , whence, on multiplying any of the equations by and substituting for , we obtain

This is of course the resultant of the quadratic polynomials above. After some calculations we also find

Let us now consider , , , as independent variables over an algebraically closed field of characteristic , giving the weight to , . Note that the expression Equation A.2 for has decreasing weights in the variables , whereas the whole expression is homogeneous of degree with respect to these weights. The total ordinary degree in all the four variables is whereas the separate degrees in , and , are both equal to .

Proposition A.1.

The polynomial is irreducible over . More generally, it is irreducible as a polynomial in over any extension of not containing a square root of .

Of course there is a similar statement on replacing with .

Proof.

Let . Then from Equation A.1 we find that the roots of as a polynomial in are

This clearly yields what we asserted.

Let now , , be fixed roots of unity, different from .

We view , , , as variables on the affine -space . Letting , , , be new variables, we define a regular map by

This also corresponds to a ring homomorphism

In the following we shall view through this homomorphism as a polynomial in , , , , with coefficients in ; it is homogeneous of degree . Similarly, we think of , , , as polynomials in , , , as given by Equation A.3. The equation defines the surface , while the equation defines the surface .

Warning: Here a word of warning is needed since the ring homomorphism is not always injective. It is injective (i.e., the map is dominant, which in turn amounts to the algebraic independence of , , , as given by Equation A.3) except when and either or is . In these cases we have as a polynomial in , , , . We have precisely if .

For simplicity of notation we omit explicit reference to this fact in what follows, which should not create confusion.

Note that up to a factor in , equals the polynomial defined by Table 1. More precisely .

Theorem A.2.

For fixed , , the polynomial is irreducible in , unless .

We note that for the polynomial factors as .

Proof.

Set

Suppose, to start with, that we are in

Case 1. If is different from all among , , which amounts to , then we have

where for this proof we set

Under the present assumption, the above ring homomorphism is injective and yields actually a field extension which is is finite Galois with group isomorphic to the -group , acting trivially on constants and capital variables and acting on the lowercase variables by transpositions on , and , .

Remark.

This action gives a certain easy action on the original variables , , , , but we do not need to make this explicit. This may be relevant when studying the rational points , since the present action is not defined over (thinking of , , , as defined over ).

Suppose that is a nontrivial irreducible factor of ; we may assume it is homogeneous of degree or .

For , we have that is also a factor of , since the latter is invariant by . Let act by fixing , and switching , . The fixed field of is . Note that is fixed by (and equals ), whereas . Hence .

Suppose first that for a constant ; this should be necessarily since . If , then vanishes on putting , whereas does not (the term dominates). Hence . But then is invariant by , thus lies in , and then necessarily in (e.g., since , are integral over which is integrally closed). Then divides in and by Proposition A.1 we deduce that is a square in , and hence in . But , hence this latter possibility leads to , . On the other hand is excluded by the present hypothesis, whereas easily leads to , again excluded.

Therefore, since , are both irreducible, nonproportional, and divisors of , is a divisor of ; since it lies in (being a norm) and since is irreducible in this ring (because as above is not a square in ), must be a constant multiple of . Setting we have that is a square in (because it is a norm from a quadratic extension generated by , or else because , become equal after such substitution). Since it lies in it is either a square in this ring, or vanishes on setting ; this last fact is excluded by direct computation. Indeed, since , , we have

On the other hand, from Equation A.1 we see that the discriminant with respect to is, up to a nonzero square, . As above we have , hence this discriminant cannot vanish, hence the polynomial cannot be a square.

This concludes the discussion of Case 1, and therefore from now on we may deal with

Case 2. If , by symmetry we may assume, say, that , i.e., and .

Let us also assume first that .

If (i.e., ), then equation Equation A.1 proves what we need unless , since is not a square. If also , we have whence .

Hence let us assume .

By Proposition A.1 applied with in place of , if is reducible over as a polynomial in , then (by Gauss’s lemma) either is divisible by or is a square in . This last possibility is excluded as above since we are presently assuming that (so are still defined).

On the other hand the expression Equation A.2 shows that is not a multiple of .

We are left with the case and or , which are symmetric, so assume . Then we have , , and we directly can check the irreducibility of , e.g., checking it is irreducible in , since it is quadratic in this variable and with discriminant a constant multiple of , not a square.

We remark that the irreducibility of the surface could also be used to argue for the finiteness of the rational angles in a fixed non-CM lattice, but the proof of Theorem 6.1 gives an explicit bound.

Remark A.3.

We remark that the varieties and are rational. Indeed one can set and , and equation Equation A.1 becomes , which is linear in , ; this shows that is rational. As for the variety , one can see that , through the substitutions Equation A.3, gives rise to a ternary quadratic form . After dehomogenizing with respect to, say, , we can then parametrize rationally , in terms of a parameter . Through Equation A.1 we can also express in terms of , , which leads to an equation which is linear in both , , thus allowing, for example, to express as a rational function of and .

In any case the fact that these varieties are rational is not relevant to our arguments.

A.2. The field of definition of

We observed that, when is not defined over , its rational points lie in the intersection of the conjugates of , and therefore on a union of finitely many varieties of dimension at most 1. With Proposition A.4 we show that the abundance of rational angles on CM spaces is instead explained by the fact that they correspond to angles for which the surface is defined over .

Proposition A.4.

The surface is defined over if and only if either or .

Proof.

In order to understand when the surface is defined over , one can consider the following four monomials in the variables , , , from equation Equation 3.15, with their coefficients:

If the variety is defined over , then the ratio of any two of those four coefficients, when they are not zero, must be a rational number. Considering the first two, we see that either there is a rational relation between , , , , or ; considering also the second two, we obtain that either there is a rational relation between , , , , or . Similarly, pairing the first with the third and the second with the fourth, we obtain that either there is a rational relation between , , , , or . If , the equation defining the surface becomes

and for it to be defined over , must have order 4 or 6 (and analogously if ). If both a -linear relation with nonzero coefficients between , , , and between , , , exist, then by Theorem 3.2 , , are roots of unity of common order or .

Acknowledgments

We thank Yves André, Julian Demeio and Davide Lombardo for useful comments and discussions. We thank the anonymous referee for pointing out to us the references in Remark 7.4.

About the authors

Roberto Dvornicich is full professor at the University of Pisa since 1990. His interests regard mainly algebra and number theory.

Francesco Veneziano is assistant professor at the University of Genova. His research interests lie in number theory and arithmetic geometry.

Umberto Zannier is professor of geometry at Scuola Normale Superiore in Pisa. His interests mainly focus on various aspects of number theory.

Table of Contents

  1. Abstract
  2. 1. Introduction
    1. Theorem 1.1.
    2. A concrete example
    3. Organization of the paper
  3. Appendix to the Introduction
    1. 1.1. Rational angles in the Gaussian lattice
    2. 1.2. On Sherrer’s proof
  4. 2. Terminology and notation
  5. 3. Equations
    1. 3.1. The equation of a rational angle
    2. 3.2. The equations of a rational -tuple containing
    3. 3.3. The equations of a rational -tuple not containing
    4. 3.4. Equations for the three main cases
    5. 3.5. Equations in roots of unity
    6. Theorem 3.2 (Conway and Jones).
  6. 4. Spaces with special symmetries
    1. 4.1. Spaces with
    2. Lemma 4.1.
    3. 4.2. CM spaces
    4. Theorem 4.2.
    5. Theorem 4.3.
    6. 4.3. Spaces with
    7. Lemma 4.4.
    8. Example 4.5.
  7. 5. Finiteness of angles in lattices outside special families
    1. 5.1. Odd primes appearing with exponent
    2. 5.2. Primes appearing with exponent at least
  8. 6. Parametric families of solutions
    1. 6.1. Case
    2. 6.2. Case
    3. 6.3. Case
    4. 6.4. Infinite families
    5. Theorem 6.1.
  9. 7. Spaces with a rational 4-tuple
    1. 7.1. Rectangular and superrectangular spaces
    2. Lemma 7.1.
    3. Theorem 7.2.
    4. 7.2. The general case
    5. 7.3. Vanishing subsums in 7.1
    6. 7.4. Dodecagonal spaces
    7. Theorem 7.3.
  10. 8. Example of spaces corresponding to rational points on a curve genus
  11. Appendix A. The irreducibility of the surface defined by equation 3.15
    1. A.1. Elimination of two quadratic equations
    2. Proposition A.1.
    3. Theorem A.2.
    4. A.2. The field of definition of
    5. Proposition A.4.
  12. Acknowledgments
  13. About the authors

Figures

Figure 1.
\renewcommand{\arraystretch}{1} \setlength{\unitlength}{1.0pt} \begin{tikzpicture}[scale=0.5] \coordinate(O) at (0,0); \node[anchor=90] at (O) {$0$}; \node[anchor=90] at (1,0) {$1$}; \node[anchor=270] at ({2.86807*cos(108)},{2.86807*sin(108)}) {$\tau$}; \node[anchor=270] at ({2.86807*cos(108)+2},{2.86807*sin(108)}) {$\tau+2$}; \node[anchor=270] at ({2.86807*cos(108)+12},{2.86807*sin(108)}) {$\tau+12$}; \node[anchor=270] at ({2.86807*cos(108)-8},{2.86807*sin(108)}) {$\tau-8$}; \node[anchor=270] at ({2.86807*cos(108)-3},{2.86807*sin(108)}) {$\tau-3$}; \foreach\y in {-10,-9,...,12} \node[circle, fill=black, inner sep=1pt, minimum size=1pt] at ({\y},{0}) {}; \foreach\y in {-9,-8,...,13} \node[circle, fill=black, inner sep=1pt, minimum size=1pt] at ({2.86807*cos(108)+\y},{2.86807*sin(108)}) {}; \foreach\y in {-8,-7,...,13} \node[circle, fill=black, inner sep=1pt, minimum size=1pt] at ({2*2.86807*cos(108)+\y},{2*2.86807*sin(108)}) {}; \foreach\y in {-7,-6,...,14} \node[circle, fill=black, inner sep=1pt, minimum size=1pt] at ({3*2.86807*cos(108)+\y},{3*2.86807*sin(108)}) {}; \draw[->,thick,red] (O)--(1,0); \draw[->,thick,red] (O)--({2.86807*cos(108)},{2.86807*sin(108)}); \draw[->,thick,blue] (O)--({2.86807*cos(108)+2},{2.86807*sin(108)}); \draw[->,thick,blue] (O)--({2.86807*cos(108)+12},{2.86807*sin(108)}); \draw[->,thick,green] (O)--({2.86807*cos(108)-3},{2.86807*sin(108)}); \draw[->,thick,green] (O)--({2.86807*cos(108)-8},{2.86807*sin(108)}); \node[circle, fill=black, inner sep=1pt, minimum size=3pt] at ({0},{0}) {}; \end{tikzpicture}
Table 1.
Figure 2.
\renewcommand{\arraystretch}{1} \setlength{\unitlength}{1.0pt} \begin{tikzpicture}[scale=1.5] \coordinate(A) at (-1,0); \coordinate(B) at (2.5,0); \coordinate(C) at (0,-0.5); \coordinate(D) at (0,1.5); \coordinate(uno) at (1,0); \node[anchor=90]at(uno){$1$}; \coordinate(tau) at ({cos(40)},{sin(40)}); \node[anchor=0]at(tau){$\tau$}; \coordinate(tau1) at ($(tau)+(1,0)$); \node[anchor=180]at(tau1){$\tau+1$}; \coordinate(tau-1) at ($(tau)-(1,0)$); \node[anchor=0]at(tau-1){$\tau-1$}; \coordinate(2tau) at ($2*(tau)$); \node[anchor=180]at(2tau){$2\tau$}; \coordinate(O) at (0,0); \node[anchor=45]at(O){$O$}; \draw[->,thin, name path=ab] (A)--(B); \draw[->,thin, name path=cd] (C)--(D); \draw[->,thick,red, name path=ac] (O)--(tau); \draw[->,thick,red, name path=ac] (O)--(uno); \draw[thin, name path=cd] (uno)--(tau1); \draw[thin, name path=ac] (tau)--(tau1); \draw[thin,dotted, name path=cd] (uno)--(tau); \draw[->,thick,blue, name path=ac] (O)--(tau1); \draw[->,thick,blue, name path=ac] (O)--(tau-1); \draw[thin, name path=ac] (tau-1)--(2tau); \draw[thin, name path=ac] (tau1)--(2tau); \end{tikzpicture}
Figure 3.
\renewcommand{\arraystretch}{1} \setlength{\unitlength}{1.0pt} \begin{tikzpicture}[scale=1.5] \coordinate(A) at (-2.5,0); \coordinate(B) at (2.5,0); \coordinate(C) at (0,-0.5); \coordinate(D) at (0,2.5); \coordinate(uno) at (1,0); \node[anchor=90]at(uno){$1$}; \coordinate(tau) at (0,1.5); \node[anchor=0]at(tau){$\tau$}; \coordinate(tau1) at ($(tau)+(1,0)$); \node[anchor=270]at(tau1){$\tau+1$}; \coordinate(taua) at ($(tau)+(2.3,0)$); \node[anchor=270]at(taua){$\tau+a$}; \coordinate(tau-1) at ($(tau)-(1,0)$); \node[anchor=270]at(tau-1){$\tau-1$}; \coordinate(tau-a) at ($(tau)-(2.3,0)$); \node[anchor=270]at(tau-a){$\tau-a$}; \coordinate(O) at (0,0); \node[anchor=45]at(O){$O$}; \draw[->,thin, name path=ab] (A)--(B); \draw[->,thin, name path=cd] (C)--(D); \draw[->,thick,red, name path=ac] (O)--(tau); \draw[->,thick,red, name path=ac] (O)--(uno); \draw[->,thick,blue, name path=ac] (O)--(tau1); \draw[->,thick,blue, name path=ac] (O)--(taua); \draw[->,thick,green, name path=ac] (O)--(tau-1); \draw[->,thick,green, name path=ac] (O)--(tau-a); \end{tikzpicture}
Figure 4.
\renewcommand{\arraystretch}{1} \setlength{\unitlength}{1.0pt} \begin{tikzpicture}[scale=1.5] \coordinate(A) at (-2,0); \coordinate(B) at (2,0); \coordinate(C) at (0,-0.5); \coordinate(D) at (0,2.5); \coordinate(uno) at (1,0); \node[anchor=90]at(uno){$1$}; \coordinate(tau) at (0,1.5); \node[anchor=0]at(tau){$\tau$}; \coordinate(tau1) at ($(tau)+(1,0)$); \node[anchor=270]at(tau1){$\tau+1$}; \coordinate(tau-1) at ($(tau)-(1,0)$); \node[anchor=270]at(tau-1){$\tau-1$}; \coordinate(O) at (0,0); \node[anchor=45]at(O){$O$}; \draw[->,thin, name path=ab] (A)--(B); \draw[->,thin, name path=cd] (C)--(D); \draw[->,thick,blue, name path=ac] (O)--(tau); \draw[->,thick,blue, name path=ac] (O)--(uno); \draw[thin, name path=cd] (uno)--(tau1); \draw[thin, name path=ac] (tau)--(tau1); \draw[->,thick,blue, name path=ac] (O)--(tau1); \draw[->,thick,blue, name path=ac] (O)--(tau-1); \draw[thin,dotted, name path=cd] (uno)--(tau); \end{tikzpicture}
Figure 5.
\renewcommand{\arraystretch}{1} \setlength{\unitlength}{1.0pt} \begin{tikzpicture}[scale=4] \coordinate(O) at (0,0); \node(pol) [draw, thick, blue!90!black,rotate=90,minimum size=8cm,regular polygon, regular polygon sides=12, rotate=195] at (0,0) {}; \coordinate(C) at (pol.corner 7); \foreach\n[count=\nu from 0, remember=\n as \lastn, evaluate={\nu+\lastn}] in {1,2,...,12} \node[anchor=\n*(360/12)+180]at(pol.corner \n){$\zeta^{\nu}$}; \draw[thin, name path=d15] (pol.corner 1)--(pol.corner 5); \draw[thin, name path=d27] (pol.corner 2)--(pol.corner 7); \draw[thin, name path=d310] (pol.corner 3)--(pol.corner 10); \draw[thin, name path=d412] (pol.corner 4)--(pol.corner 12); \draw[thin, name path=d47] (pol.corner 4)--(pol.corner 7); \draw[thin, name path=d612] (pol.corner 6)--(pol.corner 12); \path[name intersections={of=d15 and d27,by=A}]; \path[name intersections={of=d47 and d612,by=B}]; \node[circle, fill=red, inner sep=0pt, minimum size=4pt,label=45:$\tau_1$] at (A) {}; \node[circle, fill=red, inner sep=0pt, minimum size=4pt, label=-90:$O$] at (O) {}; \draw[red, very thick] (O)--(A); \draw[red, very thick] (A)--(B); \draw[red, very thick] (B)--(C); \draw[red, very thick] (C)--(O); \end{tikzpicture}
Figure 6.
\renewcommand{\arraystretch}{1} \setlength{\unitlength}{1.0pt} \begin{tikzpicture}[scale=4] \coordinate(O) at (0,0); \node(pol) [draw, thick, blue!90!black,rotate=90,minimum size=8cm,regular polygon, regular polygon sides=12, rotate=195] at (0,0) {}; \coordinate(C) at (pol.corner 7); \foreach\n[count=\nu from 0, remember=\n as \lastn, evaluate={\nu+\lastn}] in {1,2,...,12} \node[anchor=\n*(360/12)+180]at(pol.corner \n){$\zeta^{\nu}$}; \draw[thin, name path=d14] (pol.corner 1)--(pol.corner 4); \draw[thin, name path=d17] (pol.corner 1)--(pol.corner 7); \draw[thin, name path=d27] (pol.corner 7)--($(pol.corner 2)!-1cm!(pol.corner 7)$); \draw[thin, name path=d111] (pol.corner 11)--($(pol.corner 1)!-1cm!(pol.corner 11)$); \path[name intersections={of=d27 and d111,by=B}]; \node[circle, fill=red, inner sep=0pt,minimum size=4pt, label=-75:$\tau_2+2$] at (B) {}; \draw[thin, name path=d39] (pol.corner 3)--(pol.corner 9); \draw[thin, name path=d46] (pol.corner 4)--(pol.corner 6); \draw[thin, name path=d57] (pol.corner 5)--(pol.corner 7); \path[name intersections={of=d14 and d39,by=A}]; \path[name intersections={of=d46 and d57,by=tau2}]; \draw[thin] (tau2)--(B); \draw[thin] (O)--(tau2); \node[circle, fill=red, inner sep=0pt, minimum size=4pt] at (A) {}; \node[circle, fill=red, inner sep=0pt, minimum size=4pt, label=-90:$O$] at (O) {}; \node[circle, fill=red, inner sep=0pt, minimum size=4pt,label=-15:$\tau_2$] at (tau2) {}; \draw[red, very thick] (tau2)--(A); \draw[red, very thick] (A)--(O); \draw[red, very thick] (O)--(pol.corner 7); \draw[red, very thick] (pol.corner 7)--(tau2); \end{tikzpicture}
Table 2.
Rational anglesDescriptionType
homothetic to an imaginary quadratic field different from or .CM and rectangular
homothetic to .CM and superrectangular
homothetic to .
a non-CM superrectangular space.superrectangular
homothetic to one of the dodecagonal spaces.
Expected rational and elliptic families and a finite list.

Mathematical Fragments

Theorem 1.1.

Let be a space of one of the three types , , or . Then either is homotetic to one of the spaces described in Section 6.4, or any rational angle of , with , coprime, satisfies

Equation (3.1)
Equation (3.2)
Equations (3.3), (3.4), (3.5)
Equations (3.6), (3.7), (3.8), (3.9)
Equations (3.10), (3.11)
Equation (3.13)
Equation (3.14)
Equation (3.15)
Theorem 3.2 (Conway and Jones).

Let

be a linear relation with rational coefficients between roots of unity , normalized with . Then either there is a vanishing subsum or the common order of the is a squarefree number satisfying

Lemma 4.1.

Let be a space. The following conditions are equivalent:

(i)

is homothetic to with .

(ii)

is homothetic to with , a purely imaginary number.

(iii)

is homothetic to a space with .

Theorem 4.2.

Let be a space.

(i)

is CM if and only if for a squarefree .

(ii)

For a squarefree , the rational angles in are, up to equivalence, precisely those of the form with .

(iii)

The rational angles in are, up to equivalence, precisely those of the form with and .

(iv)

The rational angles in are, up to equivalence, precisely those of the form with and , , , , , where .

Theorem 4.3.

Let be a non-CM space. Then has only finitely many rational angles.

Lemma 4.4.

Let be a space. The following conditions are equivalent:

(i)

;

(ii)

is homothetic to with .

Equation (5.1)
Equation (5.2)
Equation (5.3)
Equation (5.4)
Equation (6.1)
Equation (6.2)
Theorem 6.1.

Every non-CM space has at most different rational angles.

Lemma 7.1.

Let be a space. The following conditions are equivalent:

(i)

is homothetic to with ,

(ii)

contains a rational triple with two perpendicular vectors,

(iii)

contains a rational -tuple with two perpendicular vectors.

Equation (7.1)
Remark 7.4 (An application to euclidean geometry).

It is worth noting that, as a consequence of the classification of all spaces of type obtained in this section, it can be shown that the red parallelogram in Figure 5 is the only parallelogram with the properties that:

(1)

all angles determined by sides and diagonals are rational multiples of ,

(2)

it is neither a rectangle nor a rhombus.

Quadrilaterals with property (1) have already been considered in the literature (see for example Reference Rig78) and are related to intersecting triples of diagonals in regular polygons, a topic fully analyzed in Reference PR98.

A related result that arises from the argument concerns rational products of tangents of rational angles, in a way similar but distinct from the study carried out in Reference Mye93.

Equation (8.2)
Equation (8.3)
Equation (A.1)
Equation (A.2)
Proposition A.1.

The polynomial is irreducible over . More generally, it is irreducible as a polynomial in over any extension of not containing a square root of .

Equation (A.3)

References

[Cal09]
J. S. Calcut, Gaussian integers and arctangent identities for , Amer. Math. Monthly 116 (2009), no. 6, 515–530, DOI 10.4169/193009709X470416. MR2519490, Show rawAMSref\bib{Calcut09}{article}{ author={Calcut, Jack S.}, title={Gaussian integers and arctangent identities for $\pi $}, journal={Amer. Math. Monthly}, volume={116}, date={2009}, number={6}, pages={515--530}, issn={0002-9890}, review={\MR {2519490}}, doi={10.4169/193009709X470416}, } Close amsref.
[Cas71]
J. W. S. Cassels, An introduction to the geometry of numbers, Springer-Verlag, Berlin-New York, 1971. Second printing, corrected; Die Grundlehren der mathematischen Wissenschaften, Band 99. MR0306130, Show rawAMSref\bib{casselsIntroGeoNumbers}{book}{ author={Cassels, J. W. S.}, title={An introduction to the geometry of numbers}, note={Second printing, corrected; Die Grundlehren der mathematischen Wissenschaften, Band 99}, publisher={Springer-Verlag, Berlin-New York}, date={1971}, pages={viii+344}, review={\MR {0306130}}, } Close amsref.
[CJ76]
J. H. Conway and A. J. Jones, Trigonometric Diophantine equations (On vanishing sums of roots of unity), Acta Arith. 30 (1976), no. 3, 229–240, DOI 10.4064/aa-30-3-229-240. MR422149, Show rawAMSref\bib{ConwayJones}{article}{ author={Conway, J. H.}, author={Jones, A. J.}, title={Trigonometric Diophantine equations (On vanishing sums of roots of unity)}, journal={Acta Arith.}, volume={30}, date={1976}, number={3}, pages={229--240}, issn={0065-1036}, review={\MR {422149}}, doi={10.4064/aa-30-3-229-240}, } Close amsref.
[DZ00]
R. Dvornicich and U. Zannier, On sums of roots of unity, Monatsh. Math. 129 (2000), no. 2, 97–108, DOI 10.1007/s006050050009. MR1742911, Show rawAMSref\bib{DvoZan00}{article}{ author={Dvornicich, Roberto}, author={Zannier, Umberto}, title={On sums of roots of unity}, journal={Monatsh. Math.}, volume={129}, date={2000}, number={2}, pages={97--108}, issn={0026-9255}, review={\MR {1742911}}, doi={10.1007/s006050050009}, } Close amsref.
[DZ02]
R. Dvornicich and U. Zannier, Sums of roots of unity vanishing modulo a prime, Arch. Math. (Basel) 79 (2002), no. 2, 104–108, DOI 10.1007/s00013-002-8291-4. MR1925376, Show rawAMSref\bib{DvoZan02}{article}{ author={Dvornicich, Roberto}, author={Zannier, Umberto}, title={Sums of roots of unity vanishing modulo a prime}, journal={Arch. Math. (Basel)}, volume={79}, date={2002}, number={2}, pages={104--108}, issn={0003-889X}, review={\MR {1925376}}, doi={10.1007/s00013-002-8291-4}, } Close amsref.
[Gor77]
P. Gordan, Ueber endliche Gruppen linearer Transformationen einer Veränderlichen (German), Math. Ann. 12 (1877), no. 1, 23–46, DOI 10.1007/BF01442466. MR1509926, Show rawAMSref\bib{Gordan}{article}{ author={Gordan, P.}, title={Ueber endliche Gruppen linearer Transformationen einer Ver\"{a}nderlichen}, language={German}, journal={Math. Ann.}, volume={12}, date={1877}, number={1}, pages={23--46}, issn={0025-5831}, review={\MR {1509926}}, doi={10.1007/BF01442466}, } Close amsref.
[Luc78]
E. Lucas, Théorème sur la géométrie des quinconces (French), Bull. Soc. Math. France 6 (1878), 9–10. MR1503766, Show rawAMSref\bib{Lucas-Quinconces}{article}{ author={Lucas, E.}, title={Th\'{e}or\`eme sur la g\'{e}om\'{e}trie des quinconces}, language={French}, journal={Bull. Soc. Math. France}, volume={6}, date={1878}, pages={9--10}, issn={0037-9484}, review={\MR {1503766}}, } Close amsref.
[Man65]
H. B. Mann, On linear relations between roots of unity, Mathematika 12 (1965), 107–117, DOI 10.1112/S0025579300005210. MR191892, Show rawAMSref\bib{Mann}{article}{ author={Mann, Henry B.}, title={On linear relations between roots of unity}, journal={Mathematika}, volume={12}, date={1965}, pages={107--117}, issn={0025-5793}, review={\MR {191892}}, doi={10.1112/S0025579300005210}, } Close amsref.
[Mye93]
G. Myerson, Rational products of sines of rational angles, Aequationes Math. 45 (1993), no. 1, 70–82, DOI 10.1007/BF01844426. MR1201398, Show rawAMSref\bib{MyersonProductsSines}{article}{ author={Myerson, Gerald}, title={Rational products of sines of rational angles}, journal={Aequationes Math.}, volume={45}, date={1993}, number={1}, pages={70--82}, issn={0001-9054}, review={\MR {1201398}}, doi={10.1007/BF01844426}, } Close amsref.
[PR98]
B. Poonen and M. Rubinstein, The number of intersection points made by the diagonals of a regular polygon, SIAM J. Discrete Math. 11 (1998), no. 1, 135–156, DOI 10.1137/S0895480195281246. MR1612877, Show rawAMSref\bib{PoonenRubinstein}{article}{ author={Poonen, Bjorn}, author={Rubinstein, Michael}, title={The number of intersection points made by the diagonals of a regular polygon}, journal={SIAM J. Discrete Math.}, volume={11}, date={1998}, number={1}, pages={135--156}, issn={0895-4801}, review={\MR {1612877}}, doi={10.1137/S0895480195281246}, } Close amsref.
[Rig78]
J. F. Rigby, Adventitious quadrangles: a geometrical approach, Math. Gaz. 62 (1978), no. 421, 183–191, DOI 10.2307/3616687. MR513855, Show rawAMSref\bib{Rigby}{article}{ author={Rigby, J. F.}, title={Adventitious quadrangles: a geometrical approach}, journal={Math. Gaz.}, volume={62}, date={1978}, number={421}, pages={183--191}, issn={0025-5572}, review={\MR {513855}}, doi={10.2307/3616687}, } Close amsref.
[Sch46]
W Scherrer, Die Einlagerung eines regulären Vielecks in ein Gitter, Elemente der Mathematik 1 (1946), 97–98., Show rawAMSref\bib{Scherrer-Gitter}{article}{ author={Scherrer, W}, title={{Die Einlagerung eines regul{\"a}ren Vielecks in ein Gitter}}, date={1946}, journal={Elemente der Mathematik}, volume={1}, pages={97\ndash 98}, } Close amsref.

Article Information

MSC 2020
Primary: 11H06 (Lattices and convex bodies (number-theoretic aspects)), 14G05 (Rational points), 11D61 (Exponential Diophantine equations), 51M05 (Euclidean geometries (general) and generalizations)
Keywords
  • Plane lattices
  • trigonometric diophantine equations
  • rational points on curves
Author Information
Roberto Dvornicich
Department of mathematics, University of Pisa, Largo Bruno Pontecorvo 5, 56127 Pisa, Italy
roberto.dvornicich@unipi.it
MathSciNet
Francesco Veneziano
Department of mathematics, University of Genova, Via Dodecaneso 35, 16146 Genova, Italy
veneziano@dima.unige.it
ORCID
MathSciNet
Umberto Zannier
Scuola Normale Superiore, Piazza dei Cavalieri 7, 56126 Pisa, Italy
umberto.zannier@sns.it
MathSciNet
Journal Information
Bulletin of the American Mathematical Society, Volume 59, Issue 2, ISSN 1088-9485, published by the American Mathematical Society, Providence, Rhode Island.
Publication History
This article was received on and published on .
Copyright Information
Copyright 2021 American Mathematical Society
Article References

Settings

Change font size
Resize article panel
Enable equation enrichment

Note. To explore an equation, focus it (e.g., by clicking on it) and use the arrow keys to navigate its structure. Screenreader users should be advised that enabling speech synthesis will lead to duplicate aural rendering.

For more information please visit the AMS MathViewer documentation.