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

Computing Euclidean Belyi maps

Matthew Radosevich Department of Mathematics, Dartmouth College, 6188 Kemeny Hall, Hanover, NH 03755, USA matt.j.radosevich@gmail.com  and  John Voight Department of Mathematics, Dartmouth College, 6188 Kemeny Hall, Hanover, NH 03755, USA jvoight@gmail.com
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, 11Y40

1. Introduction

1.1. Motivation

Grothendieck in his Esquisse d’un Programme [5] described an action of the absolute Galois group Gal(al|)\operatorname{Gal}(\mathbb{Q}^{\textup{al}}\,|\,\mathbb{Q}) 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 φ:Γ\Δ\\varphi\colon\Gamma\backslash{\mathcal{H}}\to\Delta\backslash{\mathcal{H}} where {\mathcal{H}} is one of the three classical geometries (the sphere, the Euclidean plane, or the hyperbolic plane), and ΓΔ\Gamma\leq\Delta is a finite-index subgroup of a triangle group. The case where {\mathcal{H}} 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 \mathbb{C} is a morphism φ:X1\varphi\colon X\to\mathbb{P}^{1}_{\mathbb{C}} of nice (projective, nonsingular, integral) curves over \mathbb{C} that is unramified away from {0,1,}\{0,1,\infty\}. 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 XX over \mathbb{C} can be defined over the algebraic numbers al\mathbb{Q}^{\textup{al}} if and only if XX 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 dd is a triple (σ0,σ1,σ)Sd3(\sigma_{0},\sigma_{1},\sigma_{\infty})\in S_{d}^{3} of permutations on dd elements such that σσ1σ0=1\sigma_{\infty}\sigma_{1}\sigma_{0}=1. A permutation triple is transitive if it generates a transitive subgroup of SdS_{d}. The monodromy around 0,1,0,1,\infty of a Belyi map of degree dd gives a permutation triple of degree dd, giving a bijection between isomorphism classes of Belyi maps of degree dd 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 σ\sigma be a transitive permutation triple of degree dd and let a,b,ca,b,c be the orders of σ0,σ1,σ\sigma_{0},\sigma_{1},\sigma_{\infty}, respectively. By the theory of covering spaces, the permutation triple σ\sigma defines a homomorphism π:Δ(a,b,c)Sd\pi\colon\Delta(a,b,c)\to S_{d} and thereby a subgroup ΓΔ(a,b,c)\Gamma\leq\Delta(a,b,c) of index dd (see section 2). The quotient Γ\\Gamma\backslash{\mathcal{H}} can be given the natural structure of a Riemann surface X(Γ)X(\Gamma), and the further quotient to Δ\\Delta\backslash{\mathcal{H}} defines a Belyi map φ:X(Γ)X(Δ)1\varphi\colon X(\Gamma)\to X(\Delta)\simeq\mathbb{P}^{1}_{\mathbb{C}}. By the theorem of Belyi, the map φ\varphi can be defined over the field of algebraic numbers al\mathbb{Q}^{\textup{al}}.

We say that σ\sigma (and its corresponding map φ\varphi) is Euclidean if 1/a+1/b+1/c=11/a+1/b+1/c=1, in which case the attached triangle group Δ(a,b,c)\Delta(a,b,c) is a group of symmetries of the Euclidean plane, whence (a,b,c)=(3,3,3),(2,3,6),(2,4,4)(a,b,c)=(3,3,3),(2,3,6),(2,4,4). Our main result provides an algorithmic way to compute algebraic equations for φ\varphi given σ\sigma.

Theorem 1.2.1.

There exists an explicit algorithm that, given as input a transitive, Euclidean permutation triple σ\sigma, produces as output a model for the Belyi map φ\varphi associated to σ\sigma over al\mathbb{Q}^{\textup{al}}.

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 σ\sigma into an explicit description of the group ΓΔ\Gamma\leq\Delta. Next, we write ΓT(Γ)R(Γ)\Gamma\simeq T(\Gamma)\rtimes R(\Gamma) as a semi-direct product, where T(Γ)T(\Gamma) consists of the subgroup of translations in Γ\Gamma and R(Γ)R(\Gamma) is generated by rotation around a particular point, which we can find explicitly. The quotients E(Γ)\colonequalsT(Γ)\E(\Gamma)\colonequals T(\Gamma)\backslash\mathbb{C} and E(Δ)\colonequalsT(Δ)\E(\Delta)\colonequals T(\Delta)\backslash\mathbb{C} define elliptic curves. We then have the following commutative diagram, which we call the master diagram:

E(Γ)βψX(Γ)φE(Δ)αX(Δ)1
\begin{gathered}\lx@xy@svg{\hbox{\raise 0.0pt\hbox{\kern 15.03473pt\hbox{\ignorespaces\ignorespaces\ignorespaces\hbox{\vtop{\kern 0.0pt\offinterlineskip\halign{\entry@#!@&&\entry@@#!@\cr&\\&\\\\&\crcr}}}\ignorespaces{\hbox{\kern-13.99306pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{E(\Gamma)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces{}\ignorespaces\ignorespaces{\hbox{\lx@xy@drawline@}}\ignorespaces\ignorespaces\ignorespaces{\hbox{\kern 29.80698pt\raise-4.04167pt\hbox{{}\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\hbox{\hbox{\kern 0.0pt\raise-1.75pt\hbox{$\scriptstyle{\beta}$}}}\kern 3.0pt}}}}}}\ignorespaces{\hbox{\kern 51.9902pt\raise-15.88698pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}\ignorespaces\ignorespaces{\hbox{\lx@xy@drawline@}}\ignorespaces{\hbox{\lx@xy@drawline@}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{\hbox{\lx@xy@droprule}}\ignorespaces\ignorespaces\ignorespaces{\hbox{\kern-10.55974pt\raise-20.30554pt\hbox{{}\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\hbox{\hbox{\kern 0.0pt\raise-1.75pt\hbox{$\scriptstyle{\psi}$}}}\kern 3.0pt}}}}}}\ignorespaces{\hbox{\kern 0.0pt\raise-30.11108pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}{\hbox{\lx@xy@droprule}}{\hbox{\lx@xy@droprule}}{\hbox{\kern 63.53882pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{}$}}}}}}}{\hbox{\kern-3.0pt\raise-20.30554pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{}$}}}}}}}{\hbox{\kern 51.9902pt\raise-20.30554pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{X(\Gamma)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{\hbox{\lx@xy@droprule}}\ignorespaces\ignorespaces\ignorespaces{\hbox{\kern 66.53882pt\raise-40.89552pt\hbox{{}\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\hbox{\hbox{\kern 0.0pt\raise-0.8264pt\hbox{$\scriptstyle{\varphi}$}}}\kern 3.0pt}}}}}}\ignorespaces{\hbox{\kern 66.53882pt\raise-49.84775pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}{\hbox{\lx@xy@droprule}}{\hbox{\lx@xy@droprule}}{\hbox{\kern-15.03473pt\raise-40.61108pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{E(\Delta)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces{}\ignorespaces\ignorespaces{\hbox{\lx@xy@drawline@}}\ignorespaces\ignorespaces\ignorespaces{\hbox{\kern 19.72876pt\raise-53.46767pt\hbox{{}\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\hbox{\hbox{\kern 0.0pt\raise-1.50694pt\hbox{$\scriptstyle{\alpha}$}}}\kern 3.0pt}}}}}}\ignorespaces{\hbox{\kern 40.86798pt\raise-53.46338pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}\ignorespaces\ignorespaces{\hbox{\lx@xy@drawline@}}\ignorespaces{\hbox{\lx@xy@drawline@}}{\hbox{\kern-3.0pt\raise-61.4855pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{}$}}}}}}}{\hbox{\kern 40.86798pt\raise-61.4855pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{X(\Delta)\simeq\mathbb{P}^{1}}$}}}}}}}\ignorespaces}}}}\ignorespaces\end{gathered}
(1.3.1)

To find the Belyi map φ\varphi, our strategy is to compute the other three maps in our diagram, filling in φ\varphi by commutativity (“descending ψ\psi along α\alpha”). The bottom map α\alpha depends only on a,b,ca,b,c and the choice of origin, giving six possibilities. The map ψ\psi is an isogeny of elliptic curves, which we compute from the inclusion of lattices implied by T(Γ)T(Δ)T(\Gamma)\leq T(\Delta) by applying formulas of Vélu. The top map β\beta is computed by looking at the fixed field of (E(Γ))\mathbb{C}(E(\Gamma)) under the finite subgroup of automorphisms corresponding to the rotations R(Γ)R(\Gamma) (taking care to ensure these rotations act by automorphisms at the origin). The final step, to fill in φ\varphi to make the diagram commute, is obtained via explicit substitution.

1.4. Contents

After reviewing background in section 2, we exhibit in section 3 the main algorithm (Algorithm 3.5.1) in pseudocode and then prove the main result (Theorem 1.2.1). In section 4 we describe an implementation in the Magma computer algebra system and then present some computed examples.

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 SdS_{d} acts on the right on {1,,d}\{1,\dots,d\}, written in exponentiated form: e.g., if τ=(1 2 3)\tau=(1\,2\,3) and μ=(2 3)\mu=(2\,3) then 1τμ=(1τ)μ=2μ=31^{\tau\mu}=(1^{\tau})^{\mu}=2^{\mu}=3.

Recall that if GG is a group, a (finite) permutation representation of GG is a group homomorphism π:GSd\pi\colon G\to S_{d} for some d1d\geq 1, and we say that π\pi is transitive if its image is a transitive subgroup of SdS_{d}. A transitive permutation triple σ\sigma defines a transitive permutation representation by π(δs)=σs\pi(\delta_{s})=\sigma_{s} for s=a,b,cs=a,b,c, and conversely.

Let π:ΔSd\pi\colon\Delta\to S_{d} be a transitive permutation representation. Let

Γ\colonequals{δΔ:1π(δ)=1}.\Gamma\colonequals\{\delta\in\Delta:1^{\pi(\delta)}=1\}. (2.1.1)

be the preimage of the stabilizer of 11 under π\pi. (The stabilizer of k{1,,d}k\in\{1,\dots,d\} is conjugate to Γ\Gamma in Δ\Delta.) Then [Δ:Γ]=d[\Delta:\Gamma]=d. Conversely, given ΓΔ\Gamma\leq\Delta of index dd, the action of Δ\Delta on the cosets of Γ\Gamma gives a transitive permutation representation π:ΔSd\pi\colon\Delta\to S_{d}, 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 TT^{*} be a triangle in the Euclidean plane \mathbb{C} with angles π/a\pi/a, π/b\pi/b, and π/c\pi/c at the vertices vav_{a}, vbv_{b}, and vcv_{c} labeled clockwise, with a,b,c2a,b,c\in\mathbb{Z}_{\geq 2}. Then in fact there are only three possibilities, namely

(a,b,c)=(3,3,3),(2,3,6),(2,4,4)(a,b,c)=(3,3,3),(2,3,6),(2,4,4)

corresponding to the solutions to 1/a+1/b+1/c=11/a+1/b+1/c=1; the corresponding tessellations of the Euclidean plane by triangles are sketched in Figure 2.2, with alternating triangles colored white and black.

[Uncaptioned image][Uncaptioned image][Uncaptioned image]Figure 2.2: Tessellations for Δ(3,3,3)Δ(2,3,6), and Δ(2,4,4)\begin{gathered}\includegraphics[scale={.25}]{pictures/333Tess.png}\includegraphics[scale={.25}]{pictures/236Tess.png}\includegraphics[scale={.25}]{pictures/244Tess.png}\\[4.0pt] \text{Figure \ref{fig:tesseuc}: Tessellations for $\Delta(3,3,3)$, $\Delta(2,3,6)$, and $\Delta(2,4,4)$}\end{gathered}

The group generated by the reflections in the sides of TT^{*} generates a discrete group of isometries acting properly on \mathbb{C}, with fundamental domain TT^{*}. The further subgroup of orientation-preserving isometries Δ(a,b,c)\Delta(a,b,c) has index 22, described as follows. For s{a,b,c}s\in\{a,b,c\}, let δs\delta_{s} be the counterclockwise rotation about vsv_{s} by an angle of 2π/s2\pi/s.

Proposition 2.2.2.

The following statements hold.

  1. (a)

    There is a presentation

    Δ=Δ(a,b,c)δa,δb,δc|δaa=δbb=δcc=δcδbδa=1.\Delta=\Delta(a,b,c)\simeq\langle\delta_{a},\delta_{b},\delta_{c}\,|\,\delta_{a}^{a}=\delta_{b}^{b}=\delta_{c}^{c}=\delta_{c}\delta_{b}\delta_{a}=1\rangle.
  2. (b)

    There is a unique group homomorphism

    ρ:Δ1c//c\rho\colon\Delta\to\tfrac{1}{c}\mathbb{Z}/\mathbb{Z}\xrightarrow{\sim}\mathbb{Z}/c\mathbb{Z} (2.2.3)

    such that

    δa,δb,δc1/a,1/b,1/cc/a,c/b,1.\delta_{a},\delta_{b},\delta_{c}\mapsto 1/a,1/b,1/c\mapsto c/a,c/b,1.
  3. (c)

    We have kerρ=T(Δ)\ker\rho=T(\Delta) where T(Δ)ΔT(\Delta)\trianglelefteq\Delta is the subgroup of translations, giving a split exact sequence

    1T(Δ)Δ𝜌/c1;1\to T(\Delta)\to\Delta\xrightarrow{\rho}\langle\mathbb{Z}/c\mathbb{Z}\rangle\to 1; (2.2.4)

    in particular,

    Δ=T(Δ)δc2/c.\Delta=T(\Delta)\langle\delta_{c}\rangle\simeq\mathbb{Z}^{2}\rtimes\mathbb{Z}/c\mathbb{Z}.
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 Δ\Delta are satisfied: indeed, we have ρ(δss)=s(c/s)0(modc)\rho(\delta_{s}^{s})=s(c/s)\equiv 0\pmod{c}, and ρ(δcδbδa)=c(1/a+1/b+1/c)0(modc)\rho(\delta_{c}\delta_{b}\delta_{a})=c(1/a+1/b+1/c)\equiv 0\pmod{c}. Alternatively, we define a group homomorphism first by taking the quotient by the commutator subgroup to surject onto (/a/b/c)/(1,1,1)(\mathbb{Z}/a\mathbb{Z}\oplus\mathbb{Z}/b\mathbb{Z}\oplus\mathbb{Z}/c\mathbb{Z})/\langle(1,1,1)\rangle, then map to /c\mathbb{Z}/c\mathbb{Z} via (x,y,z)x(c/a)+y(c/b)+z(x,y,z)\mapsto x(c/a)+y(c/b)+z.

Since it will be of some importance to us, we prove part (c) two ways. First, we compute algebraically. We treat the case Δ=Δ(2,3,6)\Delta=\Delta(2,3,6), the other two being similar. Without loss of generality, we may suppose that vc=0v_{c}=0 and vb=1v_{b}=1. Then va=(ζ6+1)/2v_{a}=(\zeta_{6}+1)/2, where ζ6=exp(2πi/6)\zeta_{6}=\exp(2\pi i/6). The translations in Δ\Delta are precisely those that translate by the Δ\Delta orbit of vc=0v_{c}=0, so T(Δ)T(\Delta) is generated by zz+(ζ6+1)=z+2vaz\mapsto z+(\zeta_{6}+1)=z+2v_{a} and zz+3i=z+(2ζ61)z\mapsto z+\sqrt{3}i=z+(2\zeta_{6}-1). We then compute directly that

δa(z)=z+(ζ6+1)\delta_{a}(z)=-z+(\zeta_{6}+1)

is the composition of the rotation zz=ζ63zz\mapsto-z=\zeta_{6}^{3}z in δc\langle\delta_{c}\rangle followed by the translation zz+(ζ6+1)z\mapsto z+(\zeta_{6}+1) in T(Δ)T(\Delta). Since δb=δc1δa1\delta_{b}=\delta_{c}^{-1}\delta_{a}^{-1}, we conclude that Δ=T(Δ)δc\Delta=T(\Delta)\langle\delta_{c}\rangle. In particular, every transformation δΔ\delta\in\Delta is of the form δ(z)=ζ6iz+β\delta(z)=\zeta_{6}^{i}z+\beta for i/ci\in\mathbb{Z}/c\mathbb{Z} and with zz+βz\mapsto z+\beta in T(Δ)T(\Delta); and written this way, ρ(δ)=i(modc)\rho(\delta)=i\pmod{c}, so indeed kerρ=T(Δ)\ker\rho=T(\Delta). (We may also verify independently that T(Δ)T(\Delta) is normal in Δ\Delta: if τ(z)=z+βT(Δ)\tau(z)=z+\beta\in T(\Delta) then

(δc1τδc)(z)=z+ζ61β=z+δc1(β)(\delta_{c}^{-1}\tau\delta_{c})(z)=z+\zeta_{6}^{-1}\beta=z+\delta_{c}^{-1}(\beta) (2.2.5)

is again translation by a point in the Δ\Delta orbit of vcv_{c}, so δc1τδcT(Δ)\delta_{c}^{-1}\tau\delta_{c}\in T(\Delta).) Finally, since T(Δ)2T(\Delta)\simeq\mathbb{Z}^{2} is generated freely by two translations, it follows that Δ2/c\Delta\simeq\mathbb{Z}^{2}\rtimes\mathbb{Z}/c\mathbb{Z} as claimed.

We may also argue geometrically, as follows. Intuitively, each transformation δs\delta_{s} rotates the plane by the corresponding interior angle 2π/s=(c/s)(2π/c)2\pi/s=(c/s)(2\pi/c), 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 2π2\pi. In other words, every element of Δ\Delta is obtained by first rotation by a power of δc\delta_{c} to put EE into one of cc positions, then translation of EE: see Figure 2.2.

[Uncaptioned image]Figure 2.2: Geometric proof of Proposition 2.2.2(b)–(c)\begin{gathered}\includegraphics[scale={.3}]{pictures/TDNormal_7-25.png}\\[4.0pt] \text{Figure \ref{fig:deltalemmapng}: Geometric proof of Proposition \ref{prop:TDNormal}(b)--(c)}\end{gathered}

More precisely, around vcv_{c} there is a central hexagon or square EE consisting of cc pairs of white and black triangles. Let δΔ\delta\in\Delta. Then δ(E)\delta(E) is another hexagon or square in the tessellation, with center δ(vc)\delta(v_{c}). It is geometrically evident (and can be verified in a straightforward manner) that there is a unique translation τδT(Δ)\tau_{\delta}\in T(\Delta) mapping δ(vc)\delta(v_{c}) to vcv_{c}, so the composition fixes vcv_{c} and maps EE to itself. But again visibly, the stabilizer of EE in Δ\Delta is precisely δc\langle\delta_{c}\rangle. This association thereby defines a surjective group homomorphism Δδc\Delta\to\langle\delta_{c}\rangle with kernel T(Δ)T(\Delta), as claimed. Figure 2.2 gives the transformation taking TT^{*} (blue) to TT^{\prime} (red) by first rotating about vcv_{c} by 5π/35\pi/3 (applying δc5\delta_{c}^{5}) then translating by zz+β1z\mapsto z+\beta_{1}, an element of T(Δ)T(\Delta). ∎

Corollary 2.2.7.

The group T(Δ)T(\Delta) is generated by

(ω1,ω2)\colonequals{(δaδc2,δbδc2), if (a,b,c)=(3,3,3);(δaδc3,δbδc4), if (a,b,c)=(2,3,6);(δaδc2,δbδc3), if (a,b,c)=(2,4,4).(\omega_{1},\omega_{2})\colonequals\begin{cases}(\delta_{a}\delta_{c}^{2},\delta_{b}\delta_{c}^{2}),&\text{ if $(a,b,c)=(3,3,3)$;}\\ (\delta_{a}\delta_{c}^{3},\delta_{b}\delta_{c}^{4}),&\text{ if $(a,b,c)=(2,3,6)$;}\\ (\delta_{a}\delta_{c}^{2},\delta_{b}\delta_{c}^{3}),&\text{ if $(a,b,c)=(2,4,4)$.}\end{cases} (2.2.8)
Proof.

In each case, ω1\omega_{1} and ω2\omega_{2} are in the kernel of the homomorphism ρ\rho described in proposition 2.2.2, and thus ω1,ω2T(Δ)\omega_{1},\omega_{2}\in T(\Delta). From Figure 2.2, it is straightforward to verify that the ω1,ω2\langle\omega_{1},\omega_{2}\rangle orbit of vcv_{c} is the same as the T(Δ)T(\Delta) orbit of vcv_{c}, so T(Δ)=ω1,ω2T(\Delta)=\langle\omega_{1},\omega_{2}\rangle. ∎

Visibly from Figure 2.2 we have Δ(3,3,3)Δ(2,3,6)\Delta(3,3,3)\trianglelefteq\Delta(2,3,6) with index 22 (halving a fundamental triangle), and T(Δ(3,3,3))=T(Δ(2,3,6))T(\Delta(3,3,3))=T(\Delta(2,3,6)). Attached to each translation subgroup is the orbit of 0

ΛΔ\colonequalsT(Δ)0\Lambda_{\Delta}\colonequals T(\Delta)\cdot 0 (2.2.9)

which defines a lattice ΛΔ2\Lambda_{\Delta}\simeq\mathbb{Z}^{2}. We write

Λ\displaystyle\Lambda_{\square} \colonequalsΛΔ(2,4,4)=[i]\displaystyle\colonequals\Lambda_{\Delta(2,4,4)}=\mathbb{Z}[i] (2.2.10)
Λ\displaystyle\Lambda_{\hexagon} \colonequalsΛΔ(3,3,3)=ΛΔ(2,3,6)=[ζ6]\displaystyle\colonequals\Lambda_{\Delta(3,3,3)}=\Lambda_{\Delta(2,3,6)}=\mathbb{Z}[\zeta_{6}]
Remark 2.2.11.

More precisely, we work with these lattices up to homothety, rescaling by an element of ×\mathbb{C}^{\times}; to obtain elliptic curves defined over \mathbb{Q} (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 Δ\Delta 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 Δ\Delta\circlearrowright\mathbb{C}. Furthermore, we can divide the four sides of the quadrilateral into two pairs of consecutive sides identified under the quotient by Δ\Delta as in Figure 2.3, so that X(Δ)X(\Delta) has genus 0.

[Uncaptioned image][Uncaptioned image][Uncaptioned image]Figure 2.3: Fundamental domains for Δ, like colored edges and vertices are identified\begin{gathered}\includegraphics[scale={.2}]{./pictures/333Dregion.png}\includegraphics[scale={.2}]{./pictures/244Dregion.png}\includegraphics[scale={.2}]{./pictures/236Dregion.png}\\[4.0pt] \text{Figure \ref{fig:DeltaFundoms}: Fundamental domains for $\Delta$, like colored edges and vertices are identified}\end{gathered}

Since T(Δ)T(\Delta) 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 T(Γ)T(\Gamma).

[Uncaptioned image][Uncaptioned image][Uncaptioned image]Figure 2.3: Fundamental domains for T(Δ), like colors identified\begin{gathered}\includegraphics[scale={.32}]{./pictures/333TDPar.png}\includegraphics[scale={.30}]{./pictures/244TDPar.png}\includegraphics[scale={.24}]{./pictures/236TDPar.png}\\[4.0pt] \text{Figure \ref{fig:FundomsTD}: Fundamental domains for $T(\Delta)$, like colors identified}\end{gathered}

Finally, a fundamental domain for Γ\Gamma is constructed in the usual manner: we choose coset representatives Δ=i=1dγiΓ\Delta=\bigsqcup_{i=1}^{d}\gamma_{i}\Gamma, and then for D(Δ)D(\Delta) the fundamental domain for Δ\Delta we have the fundamental domain D(Γ)=i=1dγiD(Δ)D(\Gamma)=\bigcup_{i=1}^{d}\gamma_{i}D(\Delta).

We now consider the genus of the surface X(Γ)\colonequalsΓ\X(\Gamma)\colonequals\Gamma\backslash\mathbb{C}. Given a permutation τSd\tau\in S_{d}, let k(τ)k(\tau) be the number of disjoint cycles in τ\tau and define its excess as e(τ)\colonequalsdk(τ)e(\tau)\colonequals d-k(\tau). Then by the Riemann–Hurwitz formula, the genus of X(Γ)X(\Gamma) is equal to [12, (1.5)]

g(X(Γ))=1d+e(σ0)+e(σ1)+e(σ)2.g(X(\Gamma))=1-d+\frac{e(\sigma_{0})+e(\sigma_{1})+e(\sigma_{\infty})}{2}. (2.3.3)
Lemma 2.3.4.

We have g(X(Γ))1g(X(\Gamma))\leq 1, with equality if and only if for all s{a,b,c}s\in\{a,b,c\}, every cycle in σs\sigma_{s} has length ss.

Proof.

For s{a,b,c}s\in\{a,b,c\} since the cycle decomposition of σs\sigma_{s} can contain no cycle of length greater than ss, we have k(σs)d/sk(\sigma_{s})\geq d/s, so

e(σa)+e(σb)+e(σc)3d(da+db+dc)=3dd=2de(\sigma_{a})+e(\sigma_{b})+e(\sigma_{c})\leq 3d-\left(\frac{d}{a}+\frac{d}{b}+\frac{d}{c}\right)=3d-d=2d

with equality if and only if all cycles in σs\sigma_{s} are length ss. Substituting this into (2.3.3), the result follows. ∎

Remark 2.3.5.

We will see later, in Corollary 2.5.7, that g(X(Γ))=1g(X(\Gamma))=1 if and only if Γ=T(Γ)\Gamma=T(\Gamma).

2.4. Translation subgroups

Let

T(Γ)\colonequalsΓT(Δ)=kerρ|ΓT(\Gamma)\colonequals\Gamma\cap T(\Delta)=\ker\rho|_{\Gamma} (2.4.1)

be the subgroup of translations in Γ\Gamma; then T(Γ)ΓT(\Gamma)\trianglelefteq\Gamma, as T(Δ)ΔT(\Delta)\trianglelefteq\Delta by Proposition 2.2.2. Writing E(Γ)\colonequalsT(Γ)\E(\Gamma)\colonequals T(\Gamma)\backslash\mathbb{C} and similarly for Δ\Delta, the containments of these four groups give quotient maps which fit into the diagram (1.3.1).

We again have a lattice

ΛΓ\colonequalsT(Γ)0\Lambda_{\Gamma}\colonequals T(\Gamma)\cdot 0 (2.4.2)

with ΛΓΛΔ\Lambda_{\Gamma}\leq\Lambda_{\Delta} a subgroup of finite index. When no confusion can arise, we will identify translation maps by the corresponding lattice element. We define

N\colonequals[T(Δ):T(Γ)].N\colonequals[T(\Delta):T(\Gamma)]. (2.4.3)

In the following algorithm, we compute a convenient basis for T(Γ)T(\Gamma).

Algorithm 2.4.4.

This algorithm takes as input σ\sigma and outputs a basis η1,η2\eta_{1},\eta_{2} for T(Γ)T(\Gamma) and N=[T(Δ):T(Γ)]N=[T(\Delta):T(\Gamma)].

  1. 1.

    Let π\pi be the transitive permutation representation attached to σ\sigma, and for i=1,2i=1,2, let ωi\omega_{i} be as in Corollary 2.2.7 (a basis for T(Δ)T(\Delta)).

  2. 2.

    Let τ1\tau_{1} be the cycle containing 11 in π(ω1)\pi(\omega_{1}) and let τ2\tau_{2} be the cycle containing 11 in π(ω21)\pi(\omega_{2}^{-1}). For i=1,2i=1,2, let i\ell_{i} be the length of τi\tau_{i}.

  3. 3.

    Compute

    V\colonequals{(b1,b2):0bii for i=1,2 and 1τ1b1=1τ2b2}.V\colonequals\bigl{\{}(b_{1},b_{2}):0\leq b_{i}\leq\ell_{i}\text{ for $i=1,2$ and }1^{\tau_{1}^{b_{1}}}=1^{\tau_{2}^{b_{2}}}\bigr{\}}.
  4. 4.

    Let AA be the matrix whose rows are the elements of VV. Reduce AA to Hermite normal form (HNF) and take its first two row vectors (n1,n2)(n_{1},n_{2}) and (0,m2)(0,m_{2}).

  5. 5.

    Return η1=ω1n1ω2n2\eta_{1}=\omega_{1}^{n_{1}}\omega_{2}^{n_{2}} and η2=ω2m2\eta_{2}=\omega_{2}^{m_{2}} and N=n1m2N=n_{1}m_{2}.

Proof of correctness.

Since ω1\omega_{1} and ω2\omega_{2} commute, any ηT(Δ)\eta\in T(\Delta) is of the form η=ω1a1ω2a2\eta=\omega_{1}^{a_{1}}\omega_{2}^{a_{2}} for some (a1,a2)2(a_{1},a_{2})\in\mathbb{Z}^{2}. By definition, such ηT(Γ)\eta\in T(\Gamma) if and only if 1(π1(ω1)a1π2(ω2)a2)=11^{(\pi_{1}(\omega_{1})^{a_{1}}\pi_{2}(\omega_{2})^{a_{2}})}=1, or equivalently when 1τ1a1=1τ2a21^{\tau_{1}^{a_{1}}}=1^{{\tau_{2}^{a_{2}}}}. Since τi\tau_{i} has order i\ell_{i}, we only need to consider 0aii0\leq a_{i}\leq\ell_{i} for i=1,2i=1,2. The \mathbb{Z}-span of VV therefore gives all pairs (a1,a2)(a_{1},a_{2}) such that η=ω1a1ω2a2\eta=\omega_{1}^{a_{1}}\omega_{2}^{a_{2}} is in T(Γ)T(\Gamma). Since only row operations are performed in computing the Hermite normal form, the \mathbb{Z}-span does not change, hence η1,η2\eta_{1},\eta_{2} computed in step 5 generate T(Γ)T(\Gamma). Finally, we have

N=[T(Δ):T(Γ)]=det(n1n20m2)=n1m2.N=[T(\Delta):T(\Gamma)]=\det\begin{pmatrix}n_{1}&n_{2}\\ 0&m_{2}\end{pmatrix}=n_{1}m_{2}.\qed

2.5. Rotation index

In this section, we study rotations in Γ\Gamma. Restricting the exact sequence (2.2.4) we obtain

1T(Γ)ΓR(Γ)11\to T(\Gamma)\to\Gamma\to R(\Gamma)\to 1

where R(Γ)\colonequalsρ(Γ)/cR(\Gamma)\colonequals\rho(\Gamma)\leq\mathbb{Z}/c\mathbb{Z}. Evidently, R(Γ)R(\Gamma) is a cyclic group with order dividing cc.

Definition 2.5.1.

The rotation index of Γ\Gamma is r(Γ)\colonequals[Γ:T(Γ)]=#R(Γ)r(\Gamma)\colonequals[\Gamma:T(\Gamma)]=\#R(\Gamma).

Lemma 2.5.2.

We have

r(Γ)=cNdr(\Gamma)=\frac{cN}{d}

where N=[T(Δ):T(Γ)]N=[T(\Delta):T(\Gamma)].

Proof.

From

[Δ:T(Γ)]=[Δ:Γ][Γ:T(Γ)]=[Δ:T(Δ)][T(Δ):T(Γ)][\Delta:T(\Gamma)]=[\Delta:\Gamma][\Gamma:T(\Gamma)]=[\Delta:T(\Delta)][T(\Delta):T(\Gamma)]

we conclude dr(Γ)=cNdr(\Gamma)=cN. ∎

In Proposition 2.2.2(c) we split the exact sequence using δc\delta_{c}. Indeed, the analogous sequence for Γ\Gamma above is again split, but not necessarily by a power of δc\delta_{c}: instead, R(Γ)R(\Gamma) is generated by a rotation about some vertex (an element in the Δ\Delta orbit of vav_{a}, vbv_{b}, or vcv_{c}), as follows.

Lemma 2.5.3.

There exists a vertex vOv_{O} whose stabilizer γOΓ\gamma_{O}\in\Gamma has ρ(γO)\rho(\gamma_{O}) a generator of R(Γ)R(\Gamma), giving a split exact sequence

1T(Γ)ΓγO11\to T(\Gamma)\to\Gamma\to\langle\gamma_{O}\rangle\to 1

so in particular Γ=T(Γ)γO2/r(Γ)\Gamma=T(\Gamma)\langle\gamma_{O}\rangle\simeq\mathbb{Z}^{2}\rtimes\mathbb{Z}/r(\Gamma)\mathbb{Z}.

Proof.

Every element of Δ\Delta is either a translation (and fixes no point) or fixes a unique point (zuz+vz\mapsto uz+v fixes z=v/(1u)z=v/(1-u) if u1u\neq 1), necessarily a vertex as every nonidentity element of finite order in Δ\Delta is conjugate to one of the generators δa,δb,δc\delta_{a},\delta_{b},\delta_{c}. So let γOΓ\gamma_{O}\in\Gamma be any element which maps to a generator of R(Γ)R(\Gamma) under ρ\rho, well-defined up to a translation in T(Γ)T(\Gamma). If γO\gamma_{O} is a translation, which is to say γOT(Γ)\gamma_{O}\in T(\Gamma), then R(Γ)R(\Gamma) is trivial: hence Γ=T(Γ)\Gamma=T(\Gamma), and we may take vOv_{O} to be any vertex (each having trivial stabilizer under Γ\Gamma).

Otherwise, γO\gamma_{O} fixes a vertex vOv_{O} 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 vOv_{O} whose stabilizer generates R(Γ)R(\Gamma) 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 T(Γ)T(\Gamma), are in bijection with the union of the sets of cycles τ\tau in σs\sigma_{s} with length s/r(Γ)s/r(\Gamma) for s{a,b,c}s\in\{a,b,c\}.

Proof.

Under the quotient map Γ\Δ\\Gamma\backslash\mathbb{C}\to\Delta\backslash\mathbb{C}, for s{a,b,c}s\in\{a,b,c\}, the preimages of the vertex vsv_{s} are in bijection with the cycles in σs\sigma_{s} and the stabilizer of a vertex with cycle τ\tau has order s/(τ)s/\ell(\tau) where (τ)\ell(\tau) is the length of τ\tau. Such a vertex has maximal rotation if and only if s/(τ)=r(Γ)s/\ell(\tau)=r(\Gamma). ∎

Because a permutation triple which is simultaneously conjugate to σ\sigma gives an isomorphic Belyi map (with differently labelled sheets), we may suppose without loss of generality that one of va,vb,vcv_{a},v_{b},v_{c} is a vertex of maximal rotation: after simultaneous conjugation, we just insist that 11 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 σ\sigma and gives as output the rotation index r(Γ)r(\Gamma) and a simultaneously conjugate triple σ\sigma^{\prime} and s{a,b,c}s\in\{a,b,c\} such that one of va,vb,vcv_{a},v_{b},v_{c} is a vertex of maximal rotation

  1. 1.

    Compute NN using Algorithm 2.4.4 and r(Γ)=cN/dr(\Gamma)=cN/d.

  2. 2.

    By trying all possibilities, find a cycle τ\tau in σs\sigma_{s} with s{a,b,c}s\in\{a,b,c\} with length (τ)=s/r(Γ)\ell(\tau)=s/r(\Gamma).

  3. 3.

    For any iτi\in\tau, return r(Γ)r(\Gamma) and the simultaneous conjugation of σ\sigma by (1i)(1\,i).

Proof.

In Step 1, the rotation index is computed correctly by Lemma 2.5.2. Step 2 will succeed by 2.5.5. By choice of Γ\Gamma as the stabilizer of 11, we conclude that vsv_{s} is a vertex of maximal rotation. ∎

From here forward, we may suppose without loss of generality that this “preprocessing” step has been applied.

We now see the exact circumstances when g(X(Γ))=1g(X(\Gamma))=1.

Corollary 2.5.7.

We have g(X(Γ))=1g(X(\Gamma))=1 if and only if r(Γ)=1r(\Gamma)=1 if and only if Γ=T(Γ)\Gamma=T(\Gamma).

Proof.

By Corollary 2.5.5, we have r(Γ)=1r(\Gamma)=1 if and only if for all s{a,b,c}s\in\{a,b,c\}, every cycle in σs\sigma_{s} has length ss; the result then follows from Lemma 2.3.4. ∎

3. Equations

From the subgroup ΓΔ\Gamma\leq\Delta of index dd, in the previous section we defined the translation subgroups T(Γ)T(Δ)T(\Gamma)\leq T(\Delta) 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 α:E(Δ)X(Δ)1\alpha\colon E(\Delta)\to X(\Delta)\simeq\mathbb{P}^{1}, which depends only on Δ\Delta (with the choice of the origin at vcv_{c}). From Proposition 2.2.2(c), the map α\alpha is the quotient by a cyclic group of rotations of order cc at a vertex vcv_{c}, which we may take as the origin of the elliptic curve E(Δ)E(\Delta). Accordingly, these rotations act by automorphisms of the elliptic curve E(Δ)E(\Delta), and so their equations are well-known [14, §II.2] (see also Lemma 3.3.2 below). Define the elliptic curves

E:y2=x3+1E:y2=x3xE_{\hexagon}\colon y^{2}=x^{3}+1\hskip 21.52771ptE_{\square}\colon y^{2}=x^{3}-x (3.1.1)

over \mathbb{Q}, the automorphisms

δ3:E\displaystyle\delta_{3}\colon E_{\hexagon} E\displaystyle\to E_{\hexagon}\hskip 21.52771pt δ4:E\displaystyle\delta_{4}\colon E_{\square} E\displaystyle\to E_{\square}\hskip 21.52771pt δ6:E\displaystyle\delta_{6}\colon E_{\hexagon} E\displaystyle\to E_{\hexagon} (3.1.2)
(x,y)\displaystyle(x,y) (ζ3x,y)\displaystyle\mapsto(\zeta_{3}x,y) (x,y)\displaystyle(x,y) (x,iy)\displaystyle\mapsto(-x,iy) (x,y)\displaystyle(x,y) (ζ31x,y).\displaystyle\mapsto(\zeta_{3}^{-1}x,-y).

and the quotient maps

α3:E\displaystyle\alpha_{3}\colon E_{\hexagon} 1\displaystyle\to\mathbb{P}^{1}\hskip 21.52771pt α4:E\displaystyle\alpha_{4}\colon E_{\square} 1\displaystyle\to\mathbb{P}^{1}\hskip 21.52771pt α6:E\displaystyle\alpha_{6}\colon E_{\hexagon} 1\displaystyle\to\mathbb{P}^{1} (3.1.3)
(x,y)\displaystyle(x,y) y+12\displaystyle\mapsto\frac{y+1}{2} (x,y)\displaystyle(x,y) x2\displaystyle\mapsto x^{2} (x,y)\displaystyle(x,y) y2.\displaystyle\mapsto y^{2}.

We recall the lattices defined in (2.2.10). After homothety, the Weierstrass map z((z),(z)/2)z\mapsto(\wp(z),\wp^{\prime}(z)/2) gives an analytic isomorphism from the complex elliptic curve /Λ\mathbb{C}/\Lambda_{\square} to E()E_{\square}(\mathbb{C}). Moreover, the rotation δ4\delta_{4} acts by (x,y)(x,iy)(x,y)\mapsto(-x,iy) (gently abusing notation), and the quotient map E(Δ)X(Δ)E(\Delta)\to X(\Delta) is given by α4\alpha_{4} in these coordinates. Similar statements hold for the two \hexagon cases.

Lemma 3.1.4.

The maps αc\alpha_{c} for c=3,4,6c=3,4,6 are Euclidean Belyi maps of degree cc.

Proof.

For α4\alpha_{4}, the set of preimages under (x,y)x2=t(x,y)\mapsto x^{2}=t has cardinality four unless t=0,t=0,\infty or y=0y=0, in which case t=x2=0,1t=x^{2}=0,1, giving ramification type (2,4,4)(2,4,4). Similarly for α6\alpha_{6}, we have six preimages under (x,y)y2=t(x,y)\mapsto y^{2}=t unless t=0,t=0,\infty or y21=x3=0y^{2}-1=x^{3}=0, in which case t=y2=1t=y^{2}=1, giving ramification (2,3,6)(2,3,6).

For α3\alpha_{3}, the map (x,y)y(x,y)\mapsto y is ramified above {±1,}\{\pm 1,\infty\} with ramification (3,3,3)(3,3,3), so to get ramification at {0,1,}\{0,1,\infty\} we simply postcompose with the Möbius transformation y(y+1)/2y\mapsto(y+1)/2. ∎

3.2. Isogeny

We now turn to the left map ψ\psi in (1.3.1), an isogeny E(Γ)E(Δ)E(\Gamma)\to E(\Delta).

We first show how to work explicitly with torsion on E(Δ)E(\Delta) using exact arithmetic. To handle the three cases uniformly, let j=ij=i or j=ζ6j=\zeta_{6}, so that Λ=ΛΔ=[j]\Lambda=\Lambda_{\Delta}=\mathbb{Z}[j], let K\colonequals(j)K\colonequals\mathbb{Q}(j)\subseteq\mathbb{C}, and let E=EE=E_{\square} or E=EE=E_{\hexagon}.

Lemma 3.2.1.

For all a+bj[j]a+bj\in\mathbb{Z}[j], there exists an effectively computable rational function ma+bj(x)K(x)m_{a+bj}(x)\in K(x) such that x([a+bj]P)=ma+bj(x(P))x([a+bj]P)=m_{a+bj}(x(P)) for all PE(al)P\in E(\mathbb{Q}^{\textup{al}}).

Proof.

When b=0b=0, the lemma is established as part of the theory of division polynomials: see Silverman [13, Exercise 3.7(d)]. When b0b\neq 0, we may similarly calculate using the explicit description of the action of jj given in (3.1.2): we still know that x([a+bj]P)x([a+bj]P) is a rational function in x(P)x(P), since x([a+bj](P))=x([a+bj]P)=x([a+bj]P)x([a+bj](-P))=x(-[a+bj]P)=x([a+bj]P). ∎

For an integer N1N\geq 1, the torsion group

E[N]1NΛ/Λ[j]/N[j]E[N]\simeq\tfrac{1}{N}\Lambda/\Lambda\simeq\mathbb{Z}[j]/N\mathbb{Z}[j]

is a cyclic [j]\mathbb{Z}[j]-module; we use the symbol PE[N]P\in E[N] to denote a generator of E[N]E[N] as a [j]\mathbb{Z}[j]-module.

Algorithm 3.2.2.

This algorithm takes as input N1N\in\mathbb{Z}_{\geq 1} and returns as output a number field LL and the set

{(a+bj,x([a+bj]P)):a,b/N}[j]/N[j]×L\{(a+bj,x([a+bj]P)):a,b\in\mathbb{Z}/N\mathbb{Z}\}\subseteq\mathbb{Z}[j]/N\mathbb{Z}[j]\times L

for a generator PP.

  1. 1.

    Compute the NN-division polynomial fN(x)[x]f_{N}(x)\in\mathbb{Q}[x] for EE.

  2. 2.

    For each proper divisor DND\mid N, compute the DD-division polynomial fD(x)[x]f_{D}(x)\in\mathbb{Q}[x] for E(Δ)E(\Delta) and divide fN(x)f_{N}(x) by gcd(fD(x),fN(x))\gcd(f_{D}(x),f_{N}(x)) recursively.

  3. 3.

    Let gN(x)g_{N}(x) be an irreducible factor of fN(x)f_{N}(x) over K[x]K[x] of highest degree and let L\colonequalsK(θ)L\colonequals K(\theta) with θ\theta a root of gN(x)g_{N}(x).

  4. 4.

    Return the values

    {(a+bj,ma+bj(θ)):a,b/N}.\{(a+bj,m_{a+bj}(\theta)):a,b\in\mathbb{Z}/N\mathbb{Z}\}.
Remark 3.2.3.

As an alternative to Step 4 (in place of computing the rational functions), at the cost of enlarging LL to include the yy-coordinate (if E:y2=f(x)E\colon y^{2}=f(x), we just need f(θ)\sqrt{f(\theta)}), we can just compute directly using the group law on EE.

Proof of correctness.

In Step 1, we form the polynomial whose roots are the xx-coordinates of the NN-torsion points, by definition of the division polynomial. In Step 2, we remove all roots whose order is a proper divisor of NN; so any remaining root will be the xx-coordinate of a point with exact order NN. Some such point PP generates E[N]E[N] as a [j]\mathbb{Z}[j] module. By lemma 3.2.1, taking θ\theta as a root of an irreducible factor of highest degree in step 3 guarantees that x(E[N])K(θ)x(E[N])\subseteq K(\theta). The output of Step 4 is correct by Lemma 3.2.1. ∎

Next, we recall section 2.4, where we defined N\colonequals[T(Δ):T(Γ)]N\colonequals[T(\Delta):T(\Gamma)] and computed in Algorithm 2.4.4 a basis for ΛΓ\Lambda_{\Gamma}. Since NΛΔΛΓN\Lambda_{\Delta}\subseteq\Lambda_{\Gamma}, we have an isogeny

ψ^:E(Δ)\displaystyle\widehat{\psi}\colon E(\Delta) E(Γ)\displaystyle\to E(\Gamma) (3.2.4)
z\displaystyle z Nz\displaystyle\mapsto Nz

dual to our desired isogeny ψ\psi. From this setup, we compute an equation for ψ\psi using Vélu’s formulas, as in the following algorithm.

Algorithm 3.2.5.

This algorithm takes as input a basis

η1\displaystyle\eta_{1} =n1ω1+n2ω2\displaystyle=n_{1}\omega_{1}+n_{2}\omega_{2} (3.2.6)
η2\displaystyle\eta_{2} =m2ω2\displaystyle=m_{2}\omega_{2}

for ΛΓΛΔ=ω1+ω2\Lambda_{\Gamma}\leq\Lambda_{\Delta}=\mathbb{Z}\omega_{1}+\mathbb{Z}\omega_{2} and gives as output a model for the isogeny ψ:E(Γ)E(Δ)\psi\colon E(\Gamma)\to E(\Delta).

  1. 1.

    Let p1\colonequals(0,n1/2)p_{1}\colonequals(0,\lceil n_{1}/2\rceil). If m2m_{2} is odd, let p2:=(m2/2,n1)p_{2}:=(\lfloor m_{2}/2\rfloor,n_{1}). If m2m_{2} is even, let p2\colonequals(m2/2,n1/2p_{2}\colonequals(m_{2}/2,\lfloor n_{1}/2\rfloor).

  2. 2.

    Let

    C\colonequals{(t1,t2)/m2×/n1:p1(t1,t2)p2}C\colonequals\{(t_{1},t_{2})\in\mathbb{Z}/m_{2}\mathbb{Z}\times\mathbb{Z}/n_{1}\mathbb{Z}:p_{1}\leq(t_{1},t_{2})\leq p_{2}\}

    where \leq here indicates the dictionary order.

  3. 3.

    Let

    K\colonequals{1N(t1n1,t1n2+t2m2):(t1,t2)C}.K\colonequals\left\{\tfrac{1}{N}(t_{1}n_{1},t_{1}n_{2}+t_{2}m_{2}):(t_{1},t_{2})\in C\right\}.
  4. 4.

    Compute

    X\colonequals{Λ(aiω1+biω2):(ai,bi)K}X\colonequals\{\wp_{\Lambda}(a_{i}\omega_{1}+b_{i}\omega_{2}):(a_{i},b_{i})\in K\}\subseteq\mathbb{C}

    to enough precision to distinguish their values.

  5. 5.

    Call Algorithm 3.2.2 with output WLW_{L}. Embed LL\hookrightarrow\mathbb{C}, and let XLLX_{L}\subseteq L be the set of xx-coordinates in WLW_{L} whose embedding into \mathbb{C} matches a value in XX.

  6. 6.

    Let

    p(x)\colonequalskXL(xk)L[x].p(x)\colonequals\prod_{k\in X_{L}}(x-k)\in L[x].

    Let KK^{\prime} be the subfield of LL generated (over \mathbb{Q}) by the coefficients of p(x)p(x).

  7. 7.

    Using Vélu’s formulas [15], compute the isogeny ψ^:EE\widehat{\psi}\colon E\to E^{\prime} with kernel p(x)K[x]p(x)\in K^{\prime}[x] and EE^{\prime} defined over KK^{\prime}.

  8. 8.

    Return the dual isogeny ψ:EE\psi\colon E^{\prime}\to E.

Proof of correctness.

Algorithm 2.4.4 gives η1=n1ω1+n2ω2\eta_{1}=n_{1}\omega_{1}+n_{2}\omega_{2} and η2=m2ω2\eta_{2}=m_{2}\omega_{2}, so ΛΓΛΔ\Lambda_{\Gamma}\subseteq\Lambda_{\Delta}. Note also that Nω1=m2η1n2η2N\omega_{1}=m_{2}\eta_{1}-n_{2}\eta_{2} and Nω2=n1η2N\omega_{2}=n_{1}\eta_{2}, so NΛΔΛΓN\Lambda_{\Delta}\subseteq\Lambda_{\Gamma}.

Let fN(x)f_{N}(x) be the NN-division polynomial. We determine the xx-coordinates of the points in kerψ^\ker\widehat{\psi} from among the roots of fNf_{N}. Since z(1/N)ΛΓz\in(1/N)\Lambda_{\Gamma} if and only if NzΛΓNz\in\Lambda_{\Gamma}, it follows that ker(ψ^)=(1/N)ΛΓ/ΛΔΛΓ/NΛΔ\ker(\widehat{\psi})=(1/N)\Lambda_{\Gamma}/\Lambda_{\Delta}\simeq\Lambda_{\Gamma}/N\Lambda_{\Delta} and

#ker(ψ^)=#(ΛΓ/NΛΔ)=det(m2n20n1)=n1m2=N.\#\ker(\widehat{\psi})=\#(\Lambda_{\Gamma}/N\Lambda_{\Delta})=\det\begin{pmatrix}m_{2}&-n_{2}\\ 0&n_{1}\end{pmatrix}=n_{1}m_{2}=N.

To list representatives for ΛΓ/NΛΔ\Lambda_{\Gamma}/N\Lambda_{\Delta}, we proceed as follows: if we identify ordered pairs (a,b)(a,b) with coordinates relative to the basis {η1,η2}\{\eta_{1},\eta_{2}\} for ΛΓ\Lambda_{\Gamma} (i.e., (a,b)(a,b) indicates the point aη1+bη2a\eta_{1}+b\eta_{2}), then (a1,b1)(a_{1},b_{1}) and (a2,b2)(a_{2},b_{2}) are equivalent modulo NΛΔN\Lambda_{\Delta} if and only if a1a2=im2a_{1}-a_{2}=im_{2} and b1b2=jn1in2b_{1}-b_{2}=jn_{1}-in_{2} for some i,ji,j\in\mathbb{Z}. So the set

{t1η1+t2η2:0t1<m2,0t2<n1}\{t_{1}\eta_{1}+t_{2}\eta_{2}:0\leq t_{1}<m_{2},0\leq t_{2}<n_{1}\} (3.2.7)

with NN elements gives a complete set of coset representatives for ΛΓ/NΛΔ\Lambda_{\Gamma}/N\Lambda_{\Delta}. It follows then that the set

A\displaystyle A \colonequals{1N(t1η1+t2η2):0t1<m2,0t2<n1}\displaystyle\colonequals\{\tfrac{1}{N}(t_{1}\eta_{1}+t_{2}\eta_{2}):0\leq t_{1}<m_{2},0\leq t_{2}<n_{1}\} (3.2.8)
={1Nt1(n1ω1+n2ω2)+1Nt2(m2ω2):0t1<m2,0t2<n1}\displaystyle\qquad=\{\tfrac{1}{N}t_{1}(n_{1}\omega_{1}+n_{2}\omega_{2})+\tfrac{1}{N}t_{2}(m_{2}\omega_{2}):0\leq t_{1}<m_{2},0\leq t_{2}<n_{1}\}
={1Nxn1ω1+1N(t1n2+t2m2)ω2:0t1<m2,0t2<n1}\displaystyle\qquad=\{\tfrac{1}{N}xn_{1}\omega_{1}+\tfrac{1}{N}(t_{1}n_{2}+t_{2}m_{2})\omega_{2}:0\leq t_{1}<m_{2},0\leq t_{2}<n_{1}\}

gives a complete set of coset representatives for (1/N)ΛΓ/ΛΔ(1/N)\Lambda_{\Gamma}/\Lambda_{\Delta}.

We use the Weierstrass \wp-function to map the points in the set zAz\in A to points P=((z),(z)/2)E()P=(\wp(z),\wp^{\prime}(z)/2)\in E(\mathbb{C}) on the algebraic model EE. On this model, since x(Q)=x(Q)x(-Q)=x(Q) for all QQ we only need one representative in AA up to inverses. Points in AA corresponding to the pairs (t1,t2)(t_{1},t_{2}) and (t1,t2)(t_{1}^{\prime},t_{2}^{\prime}) give inverses on E()E(\mathbb{C}) if and only if m2(t1+t1)m_{2}\mid(t_{1}+t_{1}^{\prime}) and n1(t2+t2)n_{1}\mid(t_{2}+t_{2}^{\prime}). Forming the set CC in step 22 then avoids redundancies so that no points in the set KK are inverse to each other.

The algebraic recognition in Steps 4 and 5 follow since the values are the distinct xx-coordinates of NN-torsion points. With an equation for EE and the polynomial representing the kernel of the isogeny ψ^:EE\widehat{\psi}\colon E\to E^{\prime}, we can use Vélu’s formula to calculate ψ^\widehat{\psi} explicitly. Taking the dual to ψ^\widehat{\psi} gives the desired isogeny ψ\psi. ∎

Remark 3.2.9.

If n1n_{1} and m2m_{2} above are coprime, then ker(ψ^)/m2×/n1/N\ker(\hat{\psi})\cong\mathbb{Z}/m_{2}\mathbb{Z}\times\mathbb{Z}/n_{1}\mathbb{Z}\cong\mathbb{Z}/N\mathbb{Z} with N=n1m2N=n_{1}m_{2} is cyclic. So, in algorithm 3.2.2, we need only compute the set of values {ma(x(P)):a/N}\{m_{a}(x(P)):a\in\mathbb{Z}/N\mathbb{Z}\} for a generating point PP of ker(ψ^)\ker(\hat{\psi}). Then, we may take those values as the roots of the kernel polynomial p(x)p(x) in step 66 of algorithm 3.2.5. As the computation of the rational maps ma+bj(x)m_{a+bj}(x) can be costly, this is a useful simplification. If n1n_{1} and m2m_{2} are not coprime, let k:=gcd(n1,m2)k:=\gcd(n_{1},m_{2}). Then, we may factor ψ=[k]ψ\psi=[k]\circ\psi^{\prime} where [k]:E(Δ)E(Δ)[k]\colon E(\Delta)\to E(\Delta) is the multiplication by kk map and ψ:E(Γ)E(Δ)\psi^{\prime}:E(\Gamma)\to E(\Delta) is the isogeny with cyclic kernel obtained as described above replacing n1n_{1} with n1/kn_{1}/k and m2m_{2} with m2/km_{2}/k.

3.3. Descent using automorphisms

Returning to our master diagram (1.3.1), we now consider the top map β:E(Γ)X(Γ)\beta\colon E(\Gamma)\to X(\Gamma) having computed in the previous section an equation for ψ\psi and E(Γ)E(\Gamma) over a number field KK^{\prime}. 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]):

(E(Γ))βψ(X(Γ))φ(E(Δ))α(X(Δ))
\begin{gathered}\lx@xy@svg{\hbox{\raise 0.0pt\hbox{\kern 22.53474pt\hbox{\ignorespaces\ignorespaces\ignorespaces\hbox{\vtop{\kern 0.0pt\offinterlineskip\halign{\entry@#!@&&\entry@@#!@\cr&\\&\\\\&\crcr}}}\ignorespaces{\hbox{\kern-21.49307pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{\mathbb{C}(E(\Gamma))\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces{}\ignorespaces\ignorespaces{\hbox{\lx@xy@drawline@}}\ignorespaces\ignorespaces\ignorespaces{\hbox{\kern 31.44293pt\raise-3.57639pt\hbox{{}\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\hbox{\hbox{\kern 0.0pt\raise-2.21529pt\hbox{$\scriptstyle{\beta^{*}}$}}}\kern 3.0pt}}}}}}\ignorespaces{}\ignorespaces\ignorespaces{\hbox{\lx@xy@drawline@}}\ignorespaces{\hbox{\lx@xy@drawline@}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{\hbox{\lx@xy@droprule}}\ignorespaces\ignorespaces\ignorespaces{\hbox{\kern-12.55974pt\raise-20.30554pt\hbox{{}\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\hbox{\hbox{\kern 0.0pt\raise-2.21529pt\hbox{$\scriptstyle{\psi^{*}}$}}}\kern 3.0pt}}}}}}\ignorespaces{}{\hbox{\lx@xy@droprule}}{\hbox{\lx@xy@droprule}}{\hbox{\kern 68.4583pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{}$}}}}}}}{\hbox{\kern-3.0pt\raise-20.30554pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{}$}}}}}}}{\hbox{\kern 49.40967pt\raise-20.30554pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{\mathbb{C}(X(\Gamma))\ignorespaces\ignorespaces\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{\hbox{\lx@xy@droprule}}\ignorespaces\ignorespaces\ignorespaces{\hbox{\kern 71.4583pt\raise-40.61108pt\hbox{{}\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\hbox{\hbox{\kern 0.0pt\raise-1.29167pt\hbox{$\scriptstyle{\varphi^{*}}$}}}\kern 3.0pt}}}}}}\ignorespaces{}{\hbox{\lx@xy@droprule}}{\hbox{\lx@xy@droprule}}{\hbox{\kern-22.53474pt\raise-40.61108pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{\mathbb{C}(E(\Delta))\ignorespaces\ignorespaces\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces{}\ignorespaces\ignorespaces{\hbox{\lx@xy@drawline@}}\ignorespaces\ignorespaces\ignorespaces{\hbox{\kern 27.71112pt\raise-55.73607pt\hbox{{}\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\hbox{\hbox{\kern 0.0pt\raise-1.97223pt\hbox{$\scriptstyle{\alpha^{*}}$}}}\kern 3.0pt}}}}}}\ignorespaces{}\ignorespaces\ignorespaces{\hbox{\lx@xy@drawline@}}\ignorespaces{\hbox{\lx@xy@drawline@}}{\hbox{\kern-3.0pt\raise-60.91663pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{}$}}}}}}}{\hbox{\kern 48.368pt\raise-60.91663pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{\mathbb{C}(X(\Delta))}$}}}}}}}\ignorespaces}}}}\ignorespaces\end{gathered}
(3.3.1)

We recall our explicit equations from section 3.1 and the automorphisms (3.1.2). The inclusion α\alpha^{*} realizes (X(Δ))\mathbb{C}(X(\Delta)) as the fixed field under δc\langle\delta_{c}^{*}\rangle. For example, for c=4c=4 we have

(E)=(x,y)\mathbb{C}(E_{\square})=\mathbb{C}(x,y)

with y2=x3xy^{2}=x^{3}-x, and so with δ4(x,y)=(x,iy)\delta_{4}(x,y)=(-x,iy) we have

(E)δ4=(x2,y4)=(x2)(E)\mathbb{C}(E_{\square})^{\langle\delta_{4}^{*}\rangle}=\mathbb{C}(x^{2},y^{4})=\mathbb{C}(x^{2})\subseteq\mathbb{C}(E_{\square})

because y4=x62x4+x2(x2)y^{4}=x^{6}-2x^{4}+x^{2}\in\mathbb{C}(x^{2}).

By Lemma 2.5.3, there exists a vertex of maximal rotation (Definition 2.5.4) for Γ\Gamma. 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 va,vb,vcv_{a},v_{b},v_{c}. We have degβ=r(Γ){1,2,3,4,6}\deg\beta=r(\Gamma)\in\{1,2,3,4,6\} equal to the rotation index.

If r(Γ)=1r(\Gamma)=1, then E(Γ)=X(Γ)E(\Gamma)=X(\Gamma) and β\beta is the identity. So we may suppose that r(Γ)>1r(\Gamma)>1.

First suppose that vc=0v_{c}=0 is a vertex of maximal rotation under a subgroup of rotations generated by a power of δc\delta_{c}. Then the quotient map β\beta is again by a subgroup of automorphisms of E(Γ)E(\Gamma) over KK^{\prime} as an elliptic curve, so is given in the same well-known manner as in section 3.1.

Lemma 3.3.2.

Suppose vcv_{c} is a vertex of maximal rotation with r(Γ)>1r(\Gamma)>1. Then X(Γ)1X(\Gamma)\simeq\mathbb{P}^{1}, and the following statements hold.

  1. (a)

    If r(Γ)=3,6r(\Gamma)=3,6, then E(Γ)E(\Gamma) has an equation of the form y2=x3+By^{2}=x^{3}+B for some nonzero BKB\in K^{\prime}, and β:E(Γ)X(Γ)\beta\colon E(\Gamma)\to X(\Gamma) can be taken to be (x,y)y,y2(x,y)\mapsto y,y^{2}, respectively.

  2. (b)

    If r(Γ)=4r(\Gamma)=4, then E(Γ):y2=x3+AxE(\Gamma)\colon y^{2}=x^{3}+Ax for some nonzero AKA\in K^{\prime}, and β(x,y)=x2\beta(x,y)=x^{2}.

  3. (c)

    If r(Γ)=2r(\Gamma)=2, then β(x,y)=x\beta(x,y)=x.

Proof.

We may suppose that E(Γ)E(\Gamma) has a Weierstrass equation y2=x3+Ax+By^{2}=x^{3}+Ax+B. Any automorphism of EE is of the form (x,y)(u2x,u3y)(x,y)\mapsto(u^{-2}x,u^{-3}y) for some u×u\in\mathbb{C}^{\times} with u4A=Au^{-4}A=A and u6B=Bu^{-6}B=B. Considering the cases r(Γ)=3,4,6r(\Gamma)=3,4,6 gives A=0A=0 or B=0B=0 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 vOv_{O} of maximum rotation is either vav_{a} or vbv_{b}, with rotations generated by an element δO\delta_{O} (generating the coset representatives of Γ/T(Γ)\Gamma/T(\Gamma)). In this case, δO\delta_{O} need not induce an automorphism of E(Γ)E(\Gamma), because as a rotation of the plane δO\delta_{O} need not take the lattice corresponding to T(Γ)T(\Gamma) back to itself. However, we may simply translate, as in the following lemma.

Lemma 3.3.3.

Let QO\colonequals((vO),(vO)/2)E(Γ)Q_{O}\colonequals(\wp(v_{O}),\wp^{\prime}(v_{O})/2)\in E(\Gamma) be the image of of vOv_{O}. Let E(Γ)E(\Gamma)^{\prime} denote the elliptic curve whose underlying curve is E(Γ)E(\Gamma) but with origin QOQ_{O}. Then we have an isomorphism

τQO:E(Γ)\displaystyle\tau_{-Q_{O}}\colon E(\Gamma) E(Γ)\displaystyle\to E(\Gamma)^{\prime} (3.3.4)
P\displaystyle P PQO\displaystyle\mapsto P-Q_{O}

of elliptic curves, and δO\delta_{O} induces an automorphism of the elliptic curve E(Γ)E(\Gamma)^{\prime} under τQO\tau_{-Q_{O}}.

Proof.

The translation isomorphism moves QOQ_{O} to the origin on E(Γ)E(\Gamma)^{\prime}; thus the action induced by δO\delta_{O} is bijective and fixes the origin on E(Γ)E(\Gamma)^{\prime}, so gives an automorphism of E(Γ)E(\Gamma)^{\prime} as an elliptic curve. ∎

Thus to compute the map β:E(Γ)X(Γ)\beta\colon E(\Gamma)\to X(\Gamma), by the lemma we first compose with the isomorphism τQO:E(Γ)E(Γ)\tau_{-Q_{O}}\colon E(\Gamma)\to E(\Gamma)^{\prime} to reduce to the previous case. But rather than compute the point QOE(Γ)Q_{O}\in E(\Gamma) and the translation map, we find it computationally more convenient to translate by the point PO\colonequals((vO),(vO)/2)E(Δ)P_{O}\colonequals(\wp(v_{O}),\wp^{\prime}(v_{O})/2)\in E(\Delta) on the base curve.

Writing E(Δ)E(\Delta)^{\prime} for the elliptic curve E(Δ)E(\Delta) having origin POP_{O}, we have the following diagram:

E(Γ)βψE(Γ)βψτQOX(Γ)φE(Δ)αE(Δ)ατPOX(Δ)
\begin{gathered}\lx@xy@svg{\hbox{\raise 0.0pt\hbox{\kern 15.03473pt\hbox{\ignorespaces\ignorespaces\ignorespaces\hbox{\vtop{\kern 0.0pt\offinterlineskip\halign{\entry@#!@&&\entry@@#!@\cr&\\&\\&&&\\&\\\\&&&\crcr}}}\ignorespaces{\hbox{\kern-3.0pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{}$}}}}}}}{\hbox{\kern 40.0764pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{E(\Gamma)^{\prime}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces{}\ignorespaces\ignorespaces{\hbox{\lx@xy@drawline@}}\ignorespaces\ignorespaces\ignorespaces{\hbox{\kern 94.68573pt\raise-14.02777pt\hbox{{}\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\hbox{\hbox{\kern 0.0pt\raise-2.30556pt\hbox{$\scriptstyle{\beta^{\prime}}$}}}\kern 3.0pt}}}}}}\ignorespaces{\hbox{\kern 125.68587pt\raise-34.35608pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}\ignorespaces\ignorespaces{\hbox{\lx@xy@drawline@}}\ignorespaces{\hbox{\lx@xy@drawline@}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{\hbox{\lx@xy@droprule}}\ignorespaces\ignorespaces\ignorespaces{\hbox{\kern 44.27972pt\raise-37.48367pt\hbox{{}\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\hbox{\hbox{\kern 0.0pt\raise-1.75pt\hbox{$\scriptstyle{\psi}$}}}\kern 3.0pt}}}}}}\ignorespaces{\hbox{\kern 54.83946pt\raise-50.41661pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}{\hbox{\lx@xy@droprule}}{\hbox{\lx@xy@droprule}}{\hbox{\kern-13.99306pt\raise-21.08331pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{E(\Gamma)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces{}\ignorespaces\ignorespaces{\hbox{\lx@xy@drawline@}}\ignorespaces\ignorespaces\ignorespaces{\hbox{\kern 72.86942pt\raise-26.14032pt\hbox{{}\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\hbox{\hbox{\kern 0.0pt\raise-1.75pt\hbox{$\scriptstyle{\beta}$}}}\kern 3.0pt}}}}}}\ignorespaces{\hbox{\kern 125.68587pt\raise-39.30034pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}\ignorespaces\ignorespaces{\hbox{\lx@xy@drawline@}}\ignorespaces{\hbox{\lx@xy@drawline@}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{\hbox{\lx@xy@droprule}}\ignorespaces\ignorespaces\ignorespaces{\hbox{\kern-10.55974pt\raise-52.3194pt\hbox{{}\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\hbox{\hbox{\kern 0.0pt\raise-1.75pt\hbox{$\scriptstyle{\psi}$}}}\kern 3.0pt}}}}}}\ignorespaces{\hbox{\kern 0.0pt\raise-73.05548pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}{\hbox{\lx@xy@droprule}}{\hbox{\lx@xy@droprule}}\ignorespaces\ignorespaces\ignorespaces{}\ignorespaces\ignorespaces{\hbox{\lx@xy@drawline@}}\ignorespaces\ignorespaces\ignorespaces{\hbox{\kern 16.2825pt\raise-4.80472pt\hbox{{}\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\hbox{\hbox{\kern 0.0pt\raise-0.27695pt\hbox{$\scriptstyle{\tau_{-Q_{O}}}$}}}\kern 3.0pt}}}}}}\ignorespaces{\hbox{\kern 40.53624pt\raise-5.5pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}\ignorespaces\ignorespaces{\hbox{\lx@xy@drawline@}}\ignorespaces{\hbox{\lx@xy@drawline@}}{\hbox{\kern 51.83946pt\raise-21.08331pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{}$}}}}}}}{\hbox{\kern-3.0pt\raise-41.38885pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{}$}}}}}}}{\hbox{\kern 51.83946pt\raise-41.38885pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{}$}}}}}}}{\hbox{\kern 94.6442pt\raise-41.38885pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{}$}}}}}}}{\hbox{\kern 125.68587pt\raise-41.38885pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{X(\Gamma)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{\hbox{\lx@xy@droprule}}\ignorespaces\ignorespaces\ignorespaces{\hbox{\kern 140.23448pt\raise-72.62494pt\hbox{{}\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\hbox{\hbox{\kern 0.0pt\raise-0.8264pt\hbox{$\scriptstyle{\varphi}$}}}\kern 3.0pt}}}}}}\ignorespaces{\hbox{\kern 140.23448pt\raise-93.36102pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}{\hbox{\lx@xy@droprule}}{\hbox{\lx@xy@droprule}}{\hbox{\kern-3.0pt\raise-62.47217pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{}$}}}}}}}{\hbox{\kern 39.03473pt\raise-62.47217pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{E(\Delta)^{\prime}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces{}\ignorespaces\ignorespaces{\hbox{\lx@xy@drawline@}}\ignorespaces\ignorespaces\ignorespaces{\hbox{\kern 94.55205pt\raise-78.1041pt\hbox{{}\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\hbox{\hbox{\kern 0.0pt\raise-2.0625pt\hbox{$\scriptstyle{\alpha^{\prime}}$}}}\kern 3.0pt}}}}}}\ignorespaces{\hbox{\kern 124.6442pt\raise-96.3247pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}\ignorespaces\ignorespaces{\hbox{\lx@xy@drawline@}}\ignorespaces{\hbox{\lx@xy@drawline@}}{\hbox{\kern-15.03473pt\raise-83.55548pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{E(\Delta)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces{}\ignorespaces\ignorespaces{\hbox{\lx@xy@drawline@}}\ignorespaces\ignorespaces\ignorespaces{\hbox{\kern 64.1211pt\raise-98.2152pt\hbox{{}\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\hbox{\hbox{\kern 0.0pt\raise-1.50694pt\hbox{$\scriptstyle{\alpha}$}}}\kern 3.0pt}}}}}}\ignorespaces{\hbox{\kern 124.6442pt\raise-101.62297pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}\ignorespaces\ignorespaces{\hbox{\lx@xy@drawline@}}\ignorespaces{\hbox{\lx@xy@drawline@}}\ignorespaces\ignorespaces\ignorespaces{}\ignorespaces\ignorespaces{\hbox{\lx@xy@drawline@}}\ignorespaces\ignorespaces\ignorespaces{\hbox{\kern 16.30922pt\raise-67.27689pt\hbox{{}\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\hbox{\hbox{\kern 0.0pt\raise-0.27695pt\hbox{$\scriptstyle{\tau_{-P_{O}}}$}}}\kern 3.0pt}}}}}}\ignorespaces{\hbox{\kern 40.53624pt\raise-67.97217pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}\ignorespaces\ignorespaces{\hbox{\lx@xy@drawline@}}\ignorespaces{\hbox{\lx@xy@drawline@}}{\hbox{\kern-3.0pt\raise-103.86102pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{}$}}}}}}}{\hbox{\kern 51.83946pt\raise-103.86102pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{}$}}}}}}}{\hbox{\kern 94.6442pt\raise-103.86102pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{}$}}}}}}}{\hbox{\kern 124.6442pt\raise-103.86102pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{X(\Delta)}$}}}}}}}\ignorespaces}}}}\ignorespaces\end{gathered}
(3.3.5)

The diagram is commutative because ψ(QO)=PO\psi(Q_{O})=P_{O}, both points corresponding to vOv_{O} under the complex uniformization. Note that the map ψ:E(Γ)E(Δ)\psi\colon E(\Gamma)^{\prime}\to E(\Delta)^{\prime} has the same defining equation as the map E(Γ)E(Δ)E(\Gamma)\to E(\Delta), and still defines a finite map of curves—it just loses the property of being a homomorphism.

In this way, we have “aligned” E(Δ)E(\Delta)^{\prime} with E(Γ)E(\Gamma)^{\prime}, and we can more simply repeat the steps above with E(Δ)E(\Delta)^{\prime} in place of E(Δ)E(\Delta) at the cost of computing translation maps τPO:E(Δ)E(Δ)\tau_{P_{O}}\colon E(\Delta)\to E(\Delta)^{\prime} with POP_{O} the image of either vav_{a} or vbv_{b}, giving a few more fixed maps α\alpha^{\prime}, which can be computed by composing α\alpha with translation (computed using the group law).

Lemma 3.3.6.

The following statements hold, with E(Δ)=E(Δ)E(\Delta)^{\prime}=E(\Delta) as in (3.1.1).

  1. (a)

    If c=6c=6 and vO=va=v2v_{O}=v_{a}=v_{2}, then we have

    α:E(Δ)\displaystyle\alpha^{\prime}\colon E(\Delta)^{\prime} 1\displaystyle\to\mathbb{P}^{1}
    (x,y)\displaystyle(x,y) (9ζ69)x2+9ζ6x+9x3+(3ζ63)x23ζ6x+1\displaystyle\mapsto\frac{(9\zeta_{6}-9)x^{2}+9\zeta_{6}x+9}{x^{3}+(3\zeta_{6}-3)x^{2}-3\zeta_{6}x+1}
  2. (b)

    If c=6c=6 and vO=vb=v3v_{O}=v_{b}=v_{3}, then we have

    α:E(Δ)\displaystyle\alpha^{\prime}\colon E(\Delta)^{\prime} 1\displaystyle\to\mathbb{P}^{1}
    (x,y)\displaystyle(x,y) x6+8x3y+8x3+16y2+32y+16x6\displaystyle\mapsto\frac{x^{6}+8x^{3}y+8x^{3}+16y^{2}+32y+16}{x^{6}}
  3. (c)

    If c=4c=4 and vO=va=v2v_{O}=v_{a}=v_{2}, then we have

    α:E(Δ)\displaystyle\alpha^{\prime}\colon E(\Delta)^{\prime} 1\displaystyle\to\mathbb{P}^{1}
    (x,y)\displaystyle(x,y) (x+1)2(x1)2\displaystyle\mapsto\frac{(x+1)^{2}}{(x-1)^{2}}

In case (a), we may need to extend the field of definition KK^{\prime} to include ζ6\zeta_{6}. In the remaining cases, we have taken vO=vcv_{O}=v_{c} 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 φ:X(Γ)X(Δ)\varphi\colon X(\Gamma)\to X(\Delta) 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 ξ\colonequalsαψ:E(Γ)X(Δ)\xi\colonequals\alpha\circ\psi\colon E(\Gamma)\to X(\Delta), a map represented by a rational function ξ(x,y)K(x,y)\xi(x,y)\in K^{\prime}(x,y) where E=E(Γ):y2=f(x)E^{\prime}=E(\Gamma)\colon y^{2}=f^{\prime}(x) is the defining equation of EE^{\prime}. By commutativity, we have ξ=φβ\xi=\varphi\circ\beta. If r(Γ)=1r(\Gamma)=1, then β\beta is the identity map so φ=ξ\varphi=\xi. So we may suppose that r(Γ)>1r(\Gamma)>1.

The monomial map β:E(Γ)X(Γ)\beta\colon E(\Gamma)\to X(\Gamma) is described by Lemma 3.3.2, corresponding to the cyclic field extension (E(Γ))(X(Γ))\mathbb{C}(E(\Gamma))\supseteq\mathbb{C}(X(\Gamma)), given explicitly by β(x,y)=y2,x2,y,x\beta(x,y)=y^{2},x^{2},y,x. In particular, φ(X(Γ))\varphi\in\mathbb{C}(X(\Gamma)) lies in this fixed field, and we need to solve

ξ(x,y)=φ(β(x,y))\xi(x,y)=\varphi(\beta(x,y))

given ξ\xi and β\beta explicitly for φ\varphi. Accordingly, we can write ξ(x,y)\xi(x,y) as a rational function in the monomial β(x,y)\beta(x,y), using the relation y2=f(x)y^{2}=f^{\prime}(x) if necessary, replacing every instance of β(x,y)\beta(x,y) in ξ(x,y)\xi(x,y) with a new variable uu. Then φ(u)K(u)\varphi(u)\in K^{\prime}(u) defines the map φ:X(Γ)11\varphi\colon X(\Gamma)\simeq\mathbb{P}^{1}\to\mathbb{P}^{1}.

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 (K,H)(K,H) where K[j]/N[j](/N)2K\leq\mathbb{Z}[j]/N\mathbb{Z}[j]\simeq(\mathbb{Z}/N\mathbb{Z})^{2} is a subgroup containing an element of order NN and HjH\leq\langle j\rangle is a subgroup with H{±1}H\neq\{\pm 1\} and HK=KHK=K. This data defines an isogeny to E(Δ)E(\Delta) dual to the one provided by the torsion subgroup, and the descent is along the subgroup of automorphisms, with HH 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 σ=(σa,σb,σc)Sd3\sigma=(\sigma_{a},\sigma_{b},\sigma_{c})\in S_{d}^{3} corresponding to a homomorphism π:ΔSd\pi\colon\Delta\to S_{d} with π(δs)=σs\pi(\delta_{s})=\sigma_{s} for s=a,b,cs=a,b,c; it gives as output a model for the corresponding Belyi map from X(Γ)X(\Gamma) to 1\mathbb{P}^{1}.

  1. 1.

    Apply the preprocessing step by calling Algorithm 2.5.6, with vertex of maximal rotation vOv_{O}.

  2. 2.

    Depending on the case of (a,b,c)(a,b,c) and vOv_{O}, look up β\beta using Lemma 3.3.2 and the map α:E(Δ)1\alpha\colon E(\Delta)\to\mathbb{P}^{1} using Lemma 3.3.6 (referring back to 3.1.3).

  3. 3.

    Call Algorithm 2.4.4 to compute a basis η1,η2\eta_{1},\eta_{2} for T(Γ)T(\Gamma) and N=[T(Δ):T(Γ)]N=[T(\Delta):T(\Gamma)].

  4. 4.

    Call Algorithm 3.2.2 to compute ψ:E(Γ)E(Δ)\psi\colon E(\Gamma)\to E(\Delta).

  5. 5.

    Compute the composition ξ\colonequalsαψ\xi\colonequals\alpha\circ\psi.

  6. 6.

    From ξ=φβ\xi=\varphi\circ\beta, compute φ:X(Γ)1\varphi\colon X(\Gamma)\to\mathbb{P}^{1} by substitution. Return φ\varphi.

Theorem 3.5.2.

Algorithm 3.5.1 terminates with correct output.

Proof.

Correctness follows from our master diagram (1.3.1) and the correctness of each step, provided by the proof of correctness of the algorithm used except for Step 6, which is justified in section 3.4. ∎

Remark 3.5.3.

In the above, we assumed throughout a Euclidean triangle group Δ\Delta with three generators δa,δb,\delta_{a},\delta_{b}, and δc\delta_{c} with orders a,b,a,b, and cc respectively and satisfying δcδbδa=1\delta_{c}\delta_{b}\delta_{a}=1. 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 σ=(σa,σb,σc)\sigma=(\sigma_{a},\sigma_{b},\sigma_{c}) such that π:ΔSn\pi\colon\Delta\to S_{n} taking δi\delta_{i} to σi\sigma_{i} described a group homomorphism with transitive image. In some contexts, we might prefer to work with the relation δaδbδc=1\delta_{a}\delta_{b}\delta_{c}=1. The change amounts to a relabeling of vertices so that va,vb,v_{a},v_{b}, and vcv_{c} follow each other counterclockwise around a chosen triangle.

Accordingly, given a permutation triple σ\sigma^{\prime} with σaσbσc=1\sigma_{a}^{\prime}\sigma_{b}^{\prime}\sigma_{c}^{\prime}=1, we just take inverses σs\colonequals(σs)1\sigma_{s}\colonequals(\sigma_{s}^{\prime})^{-1} to obtain σcσbσa=1\sigma_{c}\sigma_{b}\sigma_{a}=1, 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 ψ\psi 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 EE that span its associated lattice, so we are careful to generate our basis vectors for T(Γ)T(\Gamma) 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 φ:X(Γ)1\varphi:X(\Gamma)\to\mathbb{P}^{1} and provide factorizations of the numerator, denominator, and their difference in the case of genus zero maps, confirming the correspondence between ramification at 0, \infty, and 11 respectively and the cycle structure of σ\sigma. (We provide monic factorizations, ignoring leading coefficients.)

Example 4.2.1.

Given the permutation triple σ\colonequals((2 4 3),(1 3 4),(1 2 3))\sigma\colonequals((2\,4\,3),(1\,3\,4),(1\,2\,3)), we will illustrate the steps in our algorithm and determine the corresponding Belyi map. First, we call Algorithm 2.5.6 and conjugate σ\sigma by the transposition (1 4)(1\,4) to obtain ((2 1 3),(4 3 1),(4 2 3))((2\,1\,3),(4\,3\,1),(4\,2\,3)) where vcv_{c} is then the vertex of maximal rotation. Since this conjugate triple gives an isomorphic Belyi map, we will redefine σ\colonequals((2 1 3),(4 3 1),(4 2 3))\sigma\colonequals((2\,1\,3),(4\,3\,1),(4\,2\,3)). Since ω1\colonequalsδbδc2\omega_{1}\colonequals\delta_{b}\delta_{c}^{2} and ω2\colonequalsδb2δc\omega_{2}\colonequals\delta_{b}^{2}\delta_{c} span the translations in T(Δ)T(\Delta) by Corollary 2.2.7, we take σ1=π(ω1)=(1 3)(2 4)\sigma_{1}=\pi(\omega_{1})=(1\,3)(2\,4) and σ2=π(ω2)=(1 4)(2 3)\sigma_{2}=\pi(\omega_{2})=(1\,4)(2\,3) and call Algorithm 2.4.4. We find our basis vectors for T(Γ)T(\Gamma) are η1=ω12\eta_{1}=\omega_{1}^{2} and η2=ω22\eta_{2}=\omega_{2}^{2} so n1=2,n2=0,m1=0,n_{1}=2,n_{2}=0,m_{1}=0, and m2=2m_{2}=2.

We obtain the rotation index

r=cn1m2d=3(2)(2)4=3r=\frac{cn_{1}m_{2}}{d}=\frac{3(2)(2)}{4}=3

and take N=[T(Δ):T(Γ)]=n1m2=4N=[T(\Delta):T(\Gamma)]=n_{1}m_{2}=4, so the points in T(Δ)\T(\Delta)\backslash\mathbb{C} in the kernel of the multiplication by NN map from T(Δ)\T(\Delta)\backslash\mathbb{C} to T(Γ)\T(\Gamma)\backslash\mathbb{C} are

A={(0,0),(1/2,0),(0,1/2),(1/2,1/2)}A=\{(0,0),(1/2,0),(0,1/2),(1/2,1/2)\}

with coordinates relative to ω1\omega_{1} and ω2\omega_{2}, while the points whose images on E(Δ)E(\Delta) have distinct xx-coordinates are K:={(1/2,0),(0,1/2),(1/2,1/2)}K:=\{(1/2,0),(0,1/2),(1/2,1/2)\} as in Step 3 of Algorithm 3.2.5 . Letting k1,k2,k_{1},k_{2}, and k3k_{3} be the xx-coordinates of the images of these three points on E(Δ)E(\Delta), we obtain the kernel polynomial

p(x)=(xk1)(xk2)(xk3)=x3+1p(x)=(x-k_{1})(x-k_{2})(x-k_{3})=x^{3}+1

which we input to Vélu’s formula and take the dual to obtain the isogeny ψ:E(Γ)E(Δ)\psi\colon E(\Gamma)\to E(\Delta) given by

ψ(x,y)=((1/16)x432xx3+64,(1/64)x6y+20x3y512yx6+128x3+4096)\psi(x,y)=\left(\frac{(1/16)x^{4}-32x}{x^{3}+64},\frac{(1/64)x^{6}y+20x^{3}y-512y}{x^{6}+128x^{3}+4096}\right)

and see that E(Γ)E(\Gamma) is given by the equation y2=x3+64y^{2}=x^{3}+64.

Since we are in the Δ(3,3,3)\Delta(3,3,3) case, our map α:E(Δ)1\alpha\colon E(\Delta)\to\mathbb{P}^{1}_{\mathbb{C}} is given by α(x,y)=(y+1)/2\alpha(x,y)=(y+1)/2, so the composition ξ=αψ:E(Γ)1\xi=\alpha\circ\psi\colon E(\Gamma)\to\mathbb{P}^{1}_{\mathbb{C}} is given by

ξ(x,y)\displaystyle\xi(x,y) =α((1/16)x432xx3+64,(1/64)x6y+20x3y512yx6+128x3+4096)\displaystyle=\alpha\left(\frac{(1/16)x^{4}-32x}{x^{3}+64},\frac{(1/64)x^{6}y+20x^{3}y-512y}{x^{6}+128x^{3}+4096}\right)
=(1/128)x6y+(1/2)x6+10x3y+64x3256y+2048x6+128x3+4096\displaystyle=\frac{(1/128)x^{6}y+(1/2)x^{6}+10x^{3}y+64x^{3}-256y+2048}{x^{6}+128x^{3}+4096}

Finally, since r=6r=6, the map β:E(Γ)X(Γ)\beta\colon E(\Gamma)\to X(\Gamma) has β(x,y)=y\beta(x,y)=y. So, we wish to rewrite ξ\xi in terms of only yy. Since points on E(Γ)E(\Gamma) satisfy x3=y264x^{3}=y^{2}-64, we may replace each instance of x3x^{3} in ξ\xi with y264y^{2}-64; we obtain a rational function in yy, which gives our final Belyi map

φ(x)=(1/128)x4+(1/2)x3+9x2864x3\varphi(x)=\frac{(1/128)x^{4}+(1/2)x^{3}+9x^{2}-864}{x^{3}}

Let N(x)N(x) and D(x)D(x) be the numerator and denominator of φ\varphi respectively. Note that the preimages under φ\varphi of 0,0,\infty, and 11 respectively are the roots of N,DN,D, and NDN-D. To confirm the ramification of φ\varphi, we note that up to a constant multiple we have the factorizations

N(x)\displaystyle N(x) =(x8)(x+24)3\displaystyle=(x-8)(x+24)^{3}
D(x)\displaystyle D(x) =x3\displaystyle=x^{3}
N(x)D(x)\displaystyle N(x)-D(x) =(x+8)(x24)3\displaystyle=(x+8)(x-24)^{3}

where the repeated factors confirm the ramification, and we note the direct correspondence between the powers of the factors and the cycle structure of σ\sigma.

Example 4.2.2.

Given σ\colonequals((1 4)(2 5)(3 6),(1 3 5),(1 4 5 2 3 6))\sigma\colonequals((1\,4)(2\,5)(3\,6),(1\,3\,5),(1\,4\,5\,2\,3\,6)), we determine that X(Γ)X(\Gamma) has genus 0 and the corresponding Belyi map φ:X(Γ)1\varphi\colon X(\Gamma)\to\mathbb{P}^{1}_{\mathbb{C}} is given by

φ(x)=x6+162x5+7047x4+43740x3+413343x2+1062882x+4782969x654x5+1215x414580x3+98415x2354294x+531441\varphi(x)=\frac{x^{6}+162x^{5}+7047x^{4}+43740x^{3}+413343x^{2}+1062882x+4782969}{x^{6}-54x^{5}+1215x^{4}-14580x^{3}+98415x^{2}-354294x+531441}

with numerator, denominator, and difference given by

N(x)\displaystyle N(x) =(x3+81x2+243x+2187)2\displaystyle=(x^{3}+81x^{2}+243x+2187)^{2}
D(x)\displaystyle D(x) =(x9)6\displaystyle=(x-9)^{6}
N(x)D(x)\displaystyle N(x)-D(x) =(x2+27)(x+9)3\displaystyle=(x^{2}+27)(x+9)^{3}
Example 4.2.3.

Given the triple σ\colonequals((1 9)(2 8)(3 7)(4 6),(1 6)(2 9 10 3)(4 5 8 7),(1 2 5 4)(3 8)(6 7 10 9))\sigma\colonequals((1\,9)(2\,8)(3\,7)(4\,6),(1\,6)(2\,9\,10\,3)(4\,5\,8\,7),\\ (1\,2\,5\,4)(3\,8)(6\,7\,10\,9)) we obtain

φ(x)=1/625x10+1/125(8i+44)x8+1/25(264i+702)x6+1/5(2872i+4796)x4+(10296i+11753)x2x8+1/5(152i164)x6+1/25(18696i+1422)x4+1/125(547048i+434764)x2+1/625(1476984i9653287)\varphi(x)=\frac{\begin{multlined}1/625x^{10}+1/125(8i+44)x^{8}+1/25(264i+702)x^{6}\\ +1/5(2872i+4796)x^{4}+(10296i+11753)x^{2}\end{multlined}1/625x^{10}+1/125(8i+44)x^{8}+1/25(264i+702)x^{6}\\ +1/5(2872i+4796)x^{4}+(10296i+11753)x^{2}}{\begin{multlined}x^{8}+1/5(152i-164)x^{6}+1/25(-18696i+1422)x^{4}\\ +1/125(547048i+434764)x^{2}+1/625(-1476984i-9653287)\end{multlined}x^{8}+1/5(152i-164)x^{6}+1/25(-18696i+1422)x^{4}\\ +1/125(547048i+434764)x^{2}+1/625(-1476984i-9653287)}

with numerator, denominator, and difference respectively given by

x2(x2+10i+55)4\displaystyle x^{2}(x^{2}+10i+55)^{4}
(x2+1/5(38i41))4\displaystyle(x^{2}+1/5(38i-41))^{4}
(x4i+3)(x+4i3)(x2+(2i+14)x24i7)2(x2+(2i14)x24i7)2\displaystyle(x-4i+3)(x+4i-3)(x^{2}+(-2i+14)x-24i-7)^{2}(x^{2}+(2i-14)x-24i-7)^{2}

where i2=1i^{2}=-1.

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 σ\colonequals((1 4),(1 2 6)(3 4 5),(1 6 2 4 3 5)\sigma\colonequals((1\,4),(1\,2\,6)(3\,4\,5),(1\,6\,2\,4\,3\,5) we find the map

φ(x)=36(ζ61)(x2)(x2ζ1)2(x2+2x11)(x+2z3)6\varphi(x)=36(\zeta_{6}-1)\frac{(x-2)(x-2\zeta-1)^{2}(x^{2}+2x-11)}{(x+2z-3)^{6}}

defined over (ζ6)\mathbb{Q}(\zeta_{6}); however, it can be shown that the Belyi map descends to \mathbb{Q}, given more simply by φ(x)=9(3x63x4+x2)\varphi(x)=9(3x^{6}-3x^{4}+x^{2}). 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 J2J2, Hauptmodul and Belyi map, 2017, arXiv:1703.05200.
  • [10] Hartmut Monien, The sporadic group Co3\mathrm{Co}_{3}, 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.