Computing Euclidean Belyi maps
Abstract.
We exhibit an explicit algorithm to compute three-point branched covers of the complex projective line when the uniformizing triangle group is Euclidean.
2010 Mathematics Subject Classification:
11G32, 11Y401. Introduction
1.1. Motivation
Grothendieck in his Esquisse d’un Programme [5] described an action of the absolute Galois group of the rational numbers on the sets of Belyi maps and dessins d’enfants, linking combinatorics, topology, geometry, and arithmetic in a deep and surprising way. Computational aspects of this program [12] remain of significant interest, and there has been recent, fundamental progress using complex analytic techniques [6, 9, 10, 1]. A common thread underlying these approaches is to realize a Belyi map via uniformization as where is one of the three classical geometries (the sphere, the Euclidean plane, or the hyperbolic plane), and is a finite-index subgroup of a triangle group. The case where is spherical is truly classical, corresponding to certain triangulations of the Platonic solids. In the hyperbolic case, analytic methods can be employed to convert this geometric description into an algebraic one, using modular forms, finite element techniques, or conformal maps. What remains is the case of Euclidean triangle groups, those arising from the familiar regular triangular tessellations of the Euclidean plane. In this paper, we fill this gap: we compute Euclidean Belyi maps explicitly from maps of complex tori, forming a bridge between the classical and the general.
1.2. Main result
A Belyi map over is a morphism of nice (projective, nonsingular, integral) curves over that is unramified away from . By the Riemann existence theorem, we may equivalently work with such a map of compact Riemann surfaces. Famously, Belyi [2, 3] proved that a curve over can be defined over the algebraic numbers if and only if admits a Belyi map.
Belyi maps admit a tidy combinatorial description, something we take as the input to our algorithm. A permutation triple of degree is a triple of permutations on elements such that . A permutation triple is transitive if it generates a transitive subgroup of . The monodromy around of a Belyi map of degree gives a permutation triple of degree , giving a bijection between isomorphism classes of Belyi maps of degree and transitive permutation triples up to simultaneous conjugation. Lifting paths, one can compute (by numerical approximation) the permutation triple attached to a Belyi map; in this paper, we consider the harder, converse computational task.
Let be a transitive permutation triple of degree and let be the orders of , respectively. By the theory of covering spaces, the permutation triple defines a homomorphism and thereby a subgroup of index (see section 2). The quotient can be given the natural structure of a Riemann surface , and the further quotient to defines a Belyi map . By the theorem of Belyi, the map can be defined over the field of algebraic numbers .
We say that (and its corresponding map ) is Euclidean if , in which case the attached triangle group is a group of symmetries of the Euclidean plane, whence . Our main result provides an algorithmic way to compute algebraic equations for given .
Theorem 1.2.1.
There exists an explicit algorithm that, given as input a transitive, Euclidean permutation triple , produces as output a model for the Belyi map associated to over .
The algorithm in Theorem 1.2.1 is specified in Algorithm 3.5.1. We implemented the algorithm in the computer algebra system Magma [4]: the running time is quite favorable. We computed a database of Euclidean Belyi maps with this implementation (see section 4) which we will upload to the LMFDB [7]. Our code is available as part of a Belyi maps package available online (https://github.com/michaelmusty/Belyi).
Remark 1.2.2.
It would be interesting to estimate the running time of our algorithm by estimating the heights of intermediate computations and the precision required in Step 4 of Algorithm 3.2.5.
1.3. Proof sketch
We now briefly indicate the idea behind the proof of Theorem 1.2.1. We first convert the permutation triple into an explicit description of the group . Next, we write as a semi-direct product, where consists of the subgroup of translations in and is generated by rotation around a particular point, which we can find explicitly. The quotients and define elliptic curves. We then have the following commutative diagram, which we call the master diagram:
(1.3.1) |
To find the Belyi map , our strategy is to compute the other three maps in our diagram, filling in by commutativity (“descending along ”). The bottom map depends only on and the choice of origin, giving six possibilities. The map is an isogeny of elliptic curves, which we compute from the inclusion of lattices implied by by applying formulas of Vélu. The top map is computed by looking at the fixed field of under the finite subgroup of automorphisms corresponding to the rotations (taking care to ensure these rotations act by automorphisms at the origin). The final step, to fill in to make the diagram commute, is obtained via explicit substitution.
1.4. Contents
1.5. Acknowledgements
The authors would like to thank Sam Schiavone and Jeroen Sijsling for discussions. Voight was supported by a Simons Collaboration grant (550029).
2. Group theory and geometry
In this section, we begin by developing some preliminary input coming from group theory and geometry.
2.1. Transitive permutation representations
First, a few basic facts and conventions. In this article, the symmetric group acts on the right on , written in exponentiated form: e.g., if and then .
Recall that if is a group, a (finite) permutation representation of is a group homomorphism for some , and we say that is transitive if its image is a transitive subgroup of . A transitive permutation triple defines a transitive permutation representation by for , and conversely.
Let be a transitive permutation representation. Let
(2.1.1) |
be the preimage of the stabilizer of under . (The stabilizer of is conjugate to in .) Then . Conversely, given of index , the action of on the cosets of gives a transitive permutation representation , and this correspondence is bijective.
2.2. Euclidean triangle groups
We refer to Magnus [8, §II.4] for classical background on Euclidean triangle groups; we briefly summarize some classical facts. Let be a triangle in the Euclidean plane with angles , , and at the vertices , , and labeled clockwise, with . Then in fact there are only three possibilities, namely
corresponding to the solutions to ; the corresponding tessellations of the Euclidean plane by triangles are sketched in Figure 2.2, with alternating triangles colored white and black.
The group generated by the reflections in the sides of generates a discrete group of isometries acting properly on , with fundamental domain . The further subgroup of orientation-preserving isometries has index , described as follows. For , let be the counterclockwise rotation about by an angle of .
Proposition 2.2.2.
The following statements hold.
-
(a)
There is a presentation
-
(b)
There is a unique group homomorphism
(2.2.3) such that
-
(c)
We have where is the subgroup of translations, giving a split exact sequence
(2.2.4) in particular,
Proof.
See Magnus [8, Theorem 2.5] for a proof of (a) using the Reidemeister–Schreier method.
For part (b), we check that the relations in are satisfied: indeed, we have , and . Alternatively, we define a group homomorphism first by taking the quotient by the commutator subgroup to surject onto , then map to via .
Since it will be of some importance to us, we prove part (c) two ways. First, we compute algebraically. We treat the case , the other two being similar. Without loss of generality, we may suppose that and . Then , where . The translations in are precisely those that translate by the orbit of , so is generated by and . We then compute directly that
is the composition of the rotation in followed by the translation in . Since , we conclude that . In particular, every transformation is of the form for and with in ; and written this way, , so indeed . (We may also verify independently that is normal in : if then
(2.2.5) |
is again translation by a point in the orbit of , so .) Finally, since is generated freely by two translations, it follows that as claimed.
We may also argue geometrically, as follows. Intuitively, each transformation rotates the plane by the corresponding interior angle , composition accumulates this rotation in an abelian way, and the resulting transformation is a translation if and only if the total amount of rotation sums to a multiple of . In other words, every element of is obtained by first rotation by a power of to put into one of positions, then translation of : see Figure 2.2.
More precisely, around there is a central hexagon or square consisting of pairs of white and black triangles. Let . Then is another hexagon or square in the tessellation, with center . It is geometrically evident (and can be verified in a straightforward manner) that there is a unique translation mapping to , so the composition fixes and maps to itself. But again visibly, the stabilizer of in is precisely . This association thereby defines a surjective group homomorphism with kernel , as claimed. Figure 2.2 gives the transformation taking (blue) to (red) by first rotating about by (applying ) then translating by , an element of . ∎
Corollary 2.2.7.
The group is generated by
(2.2.8) |
Proof.
Visibly from Figure 2.2 we have with index (halving a fundamental triangle), and . Attached to each translation subgroup is the orbit of
(2.2.9) |
which defines a lattice . We write
(2.2.10) | ||||
Remark 2.2.11.
More precisely, we work with these lattices up to homothety, rescaling by an element of ; to obtain elliptic curves defined over (see section 3.1), we must rescale by a real number (which can be given explicitly as a real period).
2.3. Fundamental domains
In this section, we describe fundamental domains for the groups under consideration. A fundamental domain for the action of is obtained from any pair of one shaded triangle and one unshaded triangle which we may take to share an edge. This gives a region where all the interior points are distinct under the identification . Furthermore, we can divide the four sides of the quadrilateral into two pairs of consecutive sides identified under the quotient by as in Figure 2.3, so that has genus 0.
Since is generated by two noncollinear translations, we can take as its fundamental domain the parallelogram determined from two sides sharing a vertex at the origin. Opposite edges are identified while consecutive edges are distinct, so the fundamental region is equivalent to a torus (genus 1). Similar statements hold for .
Finally, a fundamental domain for is constructed in the usual manner: we choose coset representatives , and then for the fundamental domain for we have the fundamental domain .
We now consider the genus of the surface . Given a permutation , let be the number of disjoint cycles in and define its excess as . Then by the Riemann–Hurwitz formula, the genus of is equal to [12, (1.5)]
(2.3.3) |
Lemma 2.3.4.
We have , with equality if and only if for all , every cycle in has length .
Proof.
For since the cycle decomposition of can contain no cycle of length greater than , we have , so
with equality if and only if all cycles in are length . Substituting this into (2.3.3), the result follows. ∎
Remark 2.3.5.
We will see later, in Corollary 2.5.7, that if and only if .
2.4. Translation subgroups
Let
(2.4.1) |
be the subgroup of translations in ; then , as by Proposition 2.2.2. Writing and similarly for , the containments of these four groups give quotient maps which fit into the diagram (1.3.1).
We again have a lattice
(2.4.2) |
with a subgroup of finite index. When no confusion can arise, we will identify translation maps by the corresponding lattice element. We define
(2.4.3) |
In the following algorithm, we compute a convenient basis for .
Algorithm 2.4.4.
This algorithm takes as input and outputs a basis for and .
-
1.
Let be the transitive permutation representation attached to , and for , let be as in Corollary 2.2.7 (a basis for ).
-
2.
Let be the cycle containing in and let be the cycle containing in . For , let be the length of .
-
3.
Compute
-
4.
Let be the matrix whose rows are the elements of . Reduce to Hermite normal form (HNF) and take its first two row vectors and .
-
5.
Return and and .
Proof of correctness.
Since and commute, any is of the form for some . By definition, such if and only if , or equivalently when . Since has order , we only need to consider for . The -span of therefore gives all pairs such that is in . Since only row operations are performed in computing the Hermite normal form, the -span does not change, hence computed in step 5 generate . Finally, we have
2.5. Rotation index
In this section, we study rotations in . Restricting the exact sequence (2.2.4) we obtain
where . Evidently, is a cyclic group with order dividing .
Definition 2.5.1.
The rotation index of is .
Lemma 2.5.2.
We have
where .
Proof.
From
we conclude . ∎
In Proposition 2.2.2(c) we split the exact sequence using . Indeed, the analogous sequence for above is again split, but not necessarily by a power of : instead, is generated by a rotation about some vertex (an element in the orbit of , , or ), as follows.
Lemma 2.5.3.
There exists a vertex whose stabilizer has a generator of , giving a split exact sequence
so in particular .
Proof.
Every element of is either a translation (and fixes no point) or fixes a unique point ( fixes if ), necessarily a vertex as every nonidentity element of finite order in is conjugate to one of the generators . So let be any element which maps to a generator of under , well-defined up to a translation in . If is a translation, which is to say , then is trivial: hence , and we may take to be any vertex (each having trivial stabilizer under ).
Otherwise, fixes a vertex with the claimed properties; the splitting follows immediately, just as we saw in the geometric proof of Proposition 2.2.2(c). ∎
Definition 2.5.4.
A vertex whose stabilizer generates is called a vertex of maximum rotation.
With Lemma 2.5.3, we can be more precise about the possible vertices of maximal rotation.
Corollary 2.5.5.
The vertices of maximal rotation, up to translation by , are in bijection with the union of the sets of cycles in with length for .
Proof.
Under the quotient map , for , the preimages of the vertex are in bijection with the cycles in and the stabilizer of a vertex with cycle has order where is the length of . Such a vertex has maximal rotation if and only if . ∎
Because a permutation triple which is simultaneously conjugate to gives an isomorphic Belyi map (with differently labelled sheets), we may suppose without loss of generality that one of is a vertex of maximal rotation: after simultaneous conjugation, we just insist that belongs to a cycle as in Corollary 2.5.5. This “preprocessing” step is given as follows.
Algorithm 2.5.6.
This algorithm takes as input a Euclidean permutation triple and gives as output the rotation index and a simultaneously conjugate triple and such that one of is a vertex of maximal rotation
-
1.
Compute using Algorithm 2.4.4 and .
-
2.
By trying all possibilities, find a cycle in with with length .
-
3.
For any , return and the simultaneous conjugation of by .
Proof.
From here forward, we may suppose without loss of generality that this “preprocessing” step has been applied.
We now see the exact circumstances when .
Corollary 2.5.7.
We have if and only if if and only if .
3. Equations
From the subgroup of index , in the previous section we defined the translation subgroups whose quotients fit into the commutative diagram (1.3.1). We now calculate equations for these curves and the maps between them. As a basic reference, we refer to Silverman [13, 14].
3.1. Fixed maps
We begin with the bottom map , which depends only on (with the choice of the origin at ). From Proposition 2.2.2(c), the map is the quotient by a cyclic group of rotations of order at a vertex , which we may take as the origin of the elliptic curve . Accordingly, these rotations act by automorphisms of the elliptic curve , and so their equations are well-known [14, §II.2] (see also Lemma 3.3.2 below). Define the elliptic curves
(3.1.1) |
over , the automorphisms
(3.1.2) | ||||||||
and the quotient maps
(3.1.3) | ||||||||
We recall the lattices defined in (2.2.10). After homothety, the Weierstrass map gives an analytic isomorphism from the complex elliptic curve to . Moreover, the rotation acts by (gently abusing notation), and the quotient map is given by in these coordinates. Similar statements hold for the two cases.
Lemma 3.1.4.
The maps for are Euclidean Belyi maps of degree .
Proof.
For , the set of preimages under has cardinality four unless or , in which case , giving ramification type . Similarly for , we have six preimages under unless or , in which case , giving ramification .
For , the map is ramified above with ramification , so to get ramification at we simply postcompose with the Möbius transformation . ∎
3.2. Isogeny
We now turn to the left map in (1.3.1), an isogeny .
We first show how to work explicitly with torsion on using exact arithmetic. To handle the three cases uniformly, let or , so that , let , and let or .
Lemma 3.2.1.
For all , there exists an effectively computable rational function such that for all .
Proof.
For an integer , the torsion group
is a cyclic -module; we use the symbol to denote a generator of as a -module.
Algorithm 3.2.2.
This algorithm takes as input and returns as output a number field and the set
for a generator .
-
1.
Compute the -division polynomial for .
-
2.
For each proper divisor , compute the -division polynomial for and divide by recursively.
-
3.
Let be an irreducible factor of over of highest degree and let with a root of .
-
4.
Return the values
Remark 3.2.3.
As an alternative to Step 4 (in place of computing the rational functions), at the cost of enlarging to include the -coordinate (if , we just need ), we can just compute directly using the group law on .
Proof of correctness.
In Step 1, we form the polynomial whose roots are the -coordinates of the -torsion points, by definition of the division polynomial. In Step 2, we remove all roots whose order is a proper divisor of ; so any remaining root will be the -coordinate of a point with exact order . Some such point generates as a module. By lemma 3.2.1, taking as a root of an irreducible factor of highest degree in step 3 guarantees that . The output of Step 4 is correct by Lemma 3.2.1. ∎
Next, we recall section 2.4, where we defined and computed in Algorithm 2.4.4 a basis for . Since , we have an isogeny
(3.2.4) | ||||
dual to our desired isogeny . From this setup, we compute an equation for using Vélu’s formulas, as in the following algorithm.
Algorithm 3.2.5.
This algorithm takes as input a basis
(3.2.6) | ||||
for and gives as output a model for the isogeny .
-
1.
Let . If is odd, let . If is even, let ).
-
2.
Let
where here indicates the dictionary order.
-
3.
Let
-
4.
Compute
to enough precision to distinguish their values.
-
5.
Call Algorithm 3.2.2 with output . Embed , and let be the set of -coordinates in whose embedding into matches a value in .
-
6.
Let
Let be the subfield of generated (over ) by the coefficients of .
-
7.
Using Vélu’s formulas [15], compute the isogeny with kernel and defined over .
-
8.
Return the dual isogeny .
Proof of correctness.
Algorithm 2.4.4 gives and , so . Note also that and , so .
Let be the -division polynomial. We determine the -coordinates of the points in from among the roots of . Since if and only if , it follows that and
To list representatives for , we proceed as follows: if we identify ordered pairs with coordinates relative to the basis for (i.e., indicates the point ), then and are equivalent modulo if and only if and for some . So the set
(3.2.7) |
with elements gives a complete set of coset representatives for . It follows then that the set
(3.2.8) | ||||
gives a complete set of coset representatives for .
We use the Weierstrass -function to map the points in the set to points on the algebraic model . On this model, since for all we only need one representative in up to inverses. Points in corresponding to the pairs and give inverses on if and only if and . Forming the set in step then avoids redundancies so that no points in the set are inverse to each other.
The algebraic recognition in Steps 4 and 5 follow since the values are the distinct -coordinates of -torsion points. With an equation for and the polynomial representing the kernel of the isogeny , we can use Vélu’s formula to calculate explicitly. Taking the dual to gives the desired isogeny . ∎
Remark 3.2.9.
If and above are coprime, then with is cyclic. So, in algorithm 3.2.2, we need only compute the set of values for a generating point of . Then, we may take those values as the roots of the kernel polynomial in step of algorithm 3.2.5. As the computation of the rational maps can be costly, this is a useful simplification. If and are not coprime, let . Then, we may factor where is the multiplication by map and is the isogeny with cyclic kernel obtained as described above replacing with and with .
3.3. Descent using automorphisms
Returning to our master diagram (1.3.1), we now consider the top map having computed in the previous section an equation for and over a number field . To do so, we apply a bit of Galois theory. Associated to our master diagram is the following diagram of inclusions of function fields (see e.g. Silverman [13, §II.2]):
(3.3.1) |
We recall our explicit equations from section 3.1 and the automorphisms (3.1.2). The inclusion realizes as the fixed field under . For example, for we have
with , and so with we have
because .
By Lemma 2.5.3, there exists a vertex of maximal rotation (Definition 2.5.4) for . At the end of section 2, we argued that up to isomorphism (without loss of generality) we may suppose that this vertex is one of . We have equal to the rotation index.
If , then and is the identity. So we may suppose that .
First suppose that is a vertex of maximal rotation under a subgroup of rotations generated by a power of . Then the quotient map is again by a subgroup of automorphisms of over as an elliptic curve, so is given in the same well-known manner as in section 3.1.
Lemma 3.3.2.
Suppose is a vertex of maximal rotation with . Then , and the following statements hold.
-
(a)
If , then has an equation of the form for some nonzero , and can be taken to be , respectively.
-
(b)
If , then for some nonzero , and .
-
(c)
If , then .
Proof.
We may suppose that has a Weierstrass equation . Any automorphism of is of the form for some with and . Considering the cases gives or as in (a) and (b). We compute the maps in (a)–(c) by considering the fixed subfields under these automorphisms, as above. ∎
Suppose now that our vertex of maximum rotation is either or , with rotations generated by an element (generating the coset representatives of ). In this case, need not induce an automorphism of , because as a rotation of the plane need not take the lattice corresponding to back to itself. However, we may simply translate, as in the following lemma.
Lemma 3.3.3.
Let be the image of of . Let denote the elliptic curve whose underlying curve is but with origin . Then we have an isomorphism
(3.3.4) | ||||
of elliptic curves, and induces an automorphism of the elliptic curve under .
Proof.
The translation isomorphism moves to the origin on ; thus the action induced by is bijective and fixes the origin on , so gives an automorphism of as an elliptic curve. ∎
Thus to compute the map , by the lemma we first compose with the isomorphism to reduce to the previous case. But rather than compute the point and the translation map, we find it computationally more convenient to translate by the point on the base curve.
Writing for the elliptic curve having origin , we have the following diagram:
(3.3.5) |
The diagram is commutative because , both points corresponding to under the complex uniformization. Note that the map has the same defining equation as the map , and still defines a finite map of curves—it just loses the property of being a homomorphism.
In this way, we have “aligned” with , and we can more simply repeat the steps above with in place of at the cost of computing translation maps with the image of either or , giving a few more fixed maps , which can be computed by composing with translation (computed using the group law).
Lemma 3.3.6.
The following statements hold, with as in (3.1.1).
-
(a)
If and , then we have
-
(b)
If and , then we have
-
(c)
If and , then we have
In case (a), we may need to extend the field of definition to include . In the remaining cases, we have taken without loss of generality, so the maps (3.1.3) may be used. After having made this reduction, we drop the superscripts (the underlying curves have the same equations) and proceed to the final step.
3.4. The Belyi map
With three of the four maps in our master diagram determined, we complete the computation of by filling in the map in the master diagram from the other three sides, using commutativity. To do this, we again apply Galois theory, referring to the field diagram (3.3.1).
Let , a map represented by a rational function where is the defining equation of . By commutativity, we have . If , then is the identity map so . So we may suppose that .
The monomial map is described by Lemma 3.3.2, corresponding to the cyclic field extension , given explicitly by . In particular, lies in this fixed field, and we need to solve
given and explicitly for . Accordingly, we can write as a rational function in the monomial , using the relation if necessary, replacing every instance of in with a new variable . Then defines the map .
Remark 3.4.1.
We have seen that Euclidean Belyi maps can be understood as descending an isogeny along a fixed quotient map; this is encoded in our master diagram. Our effort has been to take as input a permutation triple and then to compute the master diagram (associated isogeny and then its descent). One can also cut this in the middle, working directly with the master diagram by specifying a pair where is a subgroup containing an element of order and is a subgroup with and . This data defines an isogeny to dual to the one provided by the torsion subgroup, and the descent is along the subgroup of automorphisms, with stabilizing this kernel.
3.5. Proof of main result
To finish, we put all of the pieces together.
Algorithm 3.5.1.
This algorithm takes as input a Euclidean, transitive permutation triple corresponding to a homomorphism with for ; it gives as output a model for the corresponding Belyi map from to .
Theorem 3.5.2.
Algorithm 3.5.1 terminates with correct output.
Proof.
Remark 3.5.3.
In the above, we assumed throughout a Euclidean triangle group with three generators and with orders and respectively and satisfying . These three generators corresponded to rotations around the three vertices of a designated triangle in the corresponding tessellation of the plane. We took as input to our algorithm the set of all permutation triples such that taking to described a group homomorphism with transitive image. In some contexts, we might prefer to work with the relation . The change amounts to a relabeling of vertices so that and follow each other counterclockwise around a chosen triangle.
Accordingly, given a permutation triple with , we just take inverses to obtain , and we call our algorithm above with this inverted input.
4. Examples and data
We conclude with some examples computed using an implementation of Algorithm 3.5.1.
4.1. Description of implementation
We implemented Algorithm 3.5.1 using the Magma computer algebra system [4]. In particular, we used the existing implementation of Vélu’s formula in calculating our isogeny and the implementation of division polynomials. The construction of these isogenies is the most time intensive step in our calculation, as in general it involves working in a number field of possibly large degree. Even with this step, most of our example computations take no more than a few seconds to finish. An example in degree 100 took only 30 seconds.
Remark 4.1.1.
Returning to Remark 2.2.11, we see that Magma provides two periods for that span its associated lattice, so we are careful to generate our basis vectors for and to deal with lattice coordinate points relative to the lattice Magma uses in its computations. As we only need worry about this for our two canonical elliptic curves, we can see which lattice Magma uses, compare it to our own lattices described above, and convert coordinates between the two by a simple change of basis operation.
4.2. Belyi maps obtained from triples
We give here some examples to illustrate Algorithm 3.5.1. We list the final Belyi maps from and provide factorizations of the numerator, denominator, and their difference in the case of genus zero maps, confirming the correspondence between ramification at , , and respectively and the cycle structure of . (We provide monic factorizations, ignoring leading coefficients.)
Example 4.2.1.
Given the permutation triple , we will illustrate the steps in our algorithm and determine the corresponding Belyi map. First, we call Algorithm 2.5.6 and conjugate by the transposition to obtain where is then the vertex of maximal rotation. Since this conjugate triple gives an isomorphic Belyi map, we will redefine . Since and span the translations in by Corollary 2.2.7, we take and and call Algorithm 2.4.4. We find our basis vectors for are and so and .
We obtain the rotation index
and take , so the points in in the kernel of the multiplication by map from to are
with coordinates relative to and , while the points whose images on have distinct -coordinates are as in Step 3 of Algorithm 3.2.5 . Letting and be the -coordinates of the images of these three points on , we obtain the kernel polynomial
which we input to Vélu’s formula and take the dual to obtain the isogeny given by
and see that is given by the equation .
Since we are in the case, our map is given by , so the composition is given by
Finally, since , the map has . So, we wish to rewrite in terms of only . Since points on satisfy , we may replace each instance of in with ; we obtain a rational function in , which gives our final Belyi map
Let and be the numerator and denominator of respectively. Note that the preimages under of , and respectively are the roots of , and . To confirm the ramification of , we note that up to a constant multiple we have the factorizations
where the repeated factors confirm the ramification, and we note the direct correspondence between the powers of the factors and the cycle structure of .
Example 4.2.2.
Given , we determine that has genus and the corresponding Belyi map is given by
with numerator, denominator, and difference given by
Example 4.2.3.
Given the triple we obtain
with numerator, denominator, and difference respectively given by
where .
Remark 4.2.4.
Unfortunately, our algorithms do not automatically descend the Belyi map to a minimal field of definition (if such a field exists). For example, for the permutation triple we find the map
defined over ; however, it can be shown that the Belyi map descends to , given more simply by . We refer to Sijsling–Voight [12, §6] and Musty–Schiavone–Sijsling–Voight [11, §4] for further discussion.
References
- [1] Dominik Barth and Andreas Wenz, Computation of Belyi maps with prescribed ramification and applications in Galois theory, J. Algebra 569 (2021), 616–642.
- [2] G.V. Belyi, Galois extensions of a maximal cyclotomic field, Math. USSR-Izv. 14 (1980), no. 2, 247–256.
- [3] G.V. Belyi, A new proof of the three-point theorem, translation in Sb. Math. 193 (2002), no. 3–4, 329–332.
- [4] W. Bosma, J. Cannon, and C. Playoust, The Magma algebra system. I. The user language, J. Symbolic Comput. 24 (3–4), 1997, 235–265.
- [5] Alexandre Grothendieck, Sketch of a programme (translation into English), Geometric Galois Actions. 1. Around Grothendieck’s Esquisse d’un Programme, eds. Leila Schneps and Pierre Lochak, London Math. Soc. Lect. Note Series, vol. 242, Cambridge University Press, Cambridge, 1997, 243–283.
- [6] Michael Klug, Michael Musty, Sam Schiavone, and John Voight, Numerical calculation of three-point covers of the projective line, LMS J. Comput. Math. 17 (2014), no. 1, 379–430.
- [7] The LMFDB Collaboration, The L-functions and Modular Forms Database, http://www.lmfdb.org, accessed 27 May 2020.
- [8] Wilhelm Magnus, Noneuclidean tesselations and their groups, Pure and Applied Mathematics, vol. 61, Academic Press, New York, 1974.
- [9] Hartmut Monien, The sporadic group , Hauptmodul and Belyi map, 2017, arXiv:1703.05200.
- [10] Hartmut Monien, The sporadic group , Hauptmodul and Belyi map, 2018, arXiv:1802.06923.
- [11] Michael Musty, Sam Schiavone, Jeroen Sijsling, and John Voight, A database of Belyi maps, Proceedings of the Thirteenth Algorithmic Number Theory Symposium (ANTS-XIII), eds. Renate Scheidler and Jonathan Sorenson, Open Book Series 2, Mathematical Sciences Publishers, Berkeley, 2019, 375–392.
- [12] Jeroen Sijsling and John Voight, On computing Belyi maps, Publ. Math. Besançon: Algèbre Théorie Nr. 2014/1, Presses Univ. Franche-Comté, Besançon, 73–131.
- [13] Joseph Silverman, The arithmetic of elliptic curves, 2nd ed., Grad. Texts in Math., vol. 106, Springer, Dordrecht, 2009.
- [14] Joseph Silverman, Advanced topics in the arithmetic of elliptic curves, Grad. Texts in Math., vol. 151, Springer-Verlag, New York, 1994.
- [15] Jacques Vélu, Isogénies entre courbes elliptiques, C. R. Acad. Sci. Paris Sér. A-B 273 (1971), A238–A241.