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

Sporadic cubic torsion

Maarten Derickx Department of Mathematics, Massachusetts Institute of Technology, Cambridge, MA 02139, USA maarten@mderickx.nl Anastassia Etropolski Department of Mathematics, Rice University, Houston, TX 77005, USA aetropolski@rice.edu Mark van Hoeij Department of Mathematics, Florida State University, Tallahassee, FL 32306 USA hoeij@math.fsu.edu Jackson S. Morrow Department of Mathematics , Emory University, Atlanta, GA 30322 USA jmorrow4692@gmail.com  and  David Zureick-Brown Department of Mathematics, Emory University, Atlanta, GA 30322 USA dzb@mathcs.emory.edu
Abstract.

Let KK be a number field, and let E/KE/K be an elliptic curve over KK. The Mordell–Weil theorem asserts that the KK-rational points E(K)E(K) of EE form a finitely generated abelian group. In this work, we complete the classification of the finite groups which appear as the torsion subgroup of E(K)E(K) for KK a cubic number field.

To do so, we determine the cubic points on the modular curves X1(N)X_{1}(N) for

N=21,22,24,25,26,28,30,32,33,35,36,39,45,65,121.N=21,22,24,25,26,28,30,32,33,35,36,39,45,65,121.

As part of our analysis, we determine the complete list of NN for which J0(N)J_{0}(N) (resp., J1(N)J_{1}(N), resp., J1(2,2N)J_{1}(2,2N)) has rank 0. We also provide evidence to a generalized version of a conjecture of Conrad, Edixhoven, and Stein by proving that the torsion on J1(N)()J_{1}(N)(\mathbb{Q}) is generated by Gal(¯/)\operatorname{Gal}(\overline{\mathbb{Q}}/\mathbb{Q})-orbits of cusps of X1(N)¯X_{1}(N)_{\overline{\mathbb{Q}}} for N55N\leq 55, N54N\neq 54.

Key words and phrases:
modular curves; elliptic curves; finitely many cubic points
1991 Mathematics Subject Classification:
11G18, (11G05, 14H45, 11Y50)

1. Introduction

Let E/E/\mathbb{Q} be an elliptic curve defined over the rationals \mathbb{Q}. In 1901, Poincaré [poincare1901] conjectured that the set E()E(\mathbb{Q}) of \mathbb{Q}-rational points on EE is a finitely generated abelian group. Mordell [mordell1922rational] proved this conjecture in 1922, and Weil [weil1929arithmetique] generalized this in 1929 to an arbitrary abelian variety defined over a number field.

Theorem 1.1 (Mordell–Weil).

For an elliptic curve defined over a number field KK,

E(K)rE(K)tors.E(K)\cong\mathbb{Z}^{r}\oplus E(K)_{{\operatorname{tors}}}.

The free rank rr of E(K)E(K) is the rank of EE over KK and the finite group E(K)torsE(K)_{{\operatorname{tors}}} is the torsion subgroup of E(K)E(K). Since E(K)torsE(K)_{{\operatorname{tors}}} is isomorphic to a finite subgroup of (/)2\left(\mathbb{Q}/\mathbb{Z}\right)^{2}, we know that this group must be isomorphic to a group of the form

/N/NM,\mathbb{Z}/N\mathbb{Z}\oplus\mathbb{Z}/NM\mathbb{Z},

for positive integers N,MN,M. The celebrated result of Mazur classified which N,MN,\,M appear for K=K=\mathbb{Q}.

Theorem 1.2 (Mazur [Mazur:eisenstein]).

Let E/E/\mathbb{Q} be an elliptic curve. Then E()torsE(\mathbb{Q})_{{\operatorname{tors}}} is isomorphic to one of the following 15 groups:

/N1\displaystyle\mathbb{Z}/N_{1}\mathbb{Z} with 1N112,N111,\displaystyle 1\leq N_{1}\leq 12,N_{1}\neq 11,
/2/2N2\displaystyle\mathbb{Z}/2\mathbb{Z}\oplus\mathbb{Z}/2N_{2}\mathbb{Z} with 1N24.\displaystyle 1\leq N_{2}\leq 4.

Furthermore, there exist infinitely many ¯\overline{\mathbb{Q}}-isomorphism classes for each such torsion subgroup.

The modular curves X1(M,MN)X_{1}(M,MN) classify elliptic curves (and degenerations) together with independent points P,QP,Q of order MM and MNMN. In this language, Mazur’s theorem asserts that X1(M,MN)()X_{1}(M,MN)(\mathbb{Q}) has no non-cuspidal rational points for (M,MN)(M,MN) outside of the above set.

In 1996, Merel [Merel:uniformity] proved the existence of a uniform bound on the size of E(K)torsE(K)_{{\operatorname{tors}}} that depends only on the degree of the number field K.K. Merel’s result leads to the natural question of classifying (up to isomorphism) the torsion subgroups of elliptic curves defined over number fields of degree dd, for a fixed integer d1d\geq 1.

For d=2d=2, this classification was started by Kenku–Momose and completed by Kamienny.

Theorem 1.3 (Kenku–Momose [kenku1988torsion]; Kamienny [kamienny1992torsion]).

Let K/K/\mathbb{Q} be a quadratic extension and E/KE/K be an elliptic curve. Then E(K)torsE(K)_{{\operatorname{tors}}} is isomorphic to one of the following 26 groups:

/N1\displaystyle\mathbb{Z}/N_{1}\mathbb{Z} with 1N118,N117,\displaystyle 1\leq N_{1}\leq 18,N_{1}\neq 17,
/2/2N2\displaystyle\mathbb{Z}/2\mathbb{Z}\oplus\mathbb{Z}/2N_{2}\mathbb{Z} with 1N26,\displaystyle 1\leq N_{2}\leq 6,
/3/3N3\displaystyle\mathbb{Z}/3\mathbb{Z}\oplus\mathbb{Z}/3N_{3}\mathbb{Z} with N3=1,2,\displaystyle N_{3}=1,2,
/4/4.\displaystyle\mathbb{Z}/4\mathbb{Z}\oplus\mathbb{Z}/4\mathbb{Z}.

Furthermore, there exist infinitely many ¯\overline{\mathbb{Q}}-isomorphism classes for each such torsion subgroup.

The modular curves X1(M,MN)X_{1}(M,MN) for the above list all have genus 2\leq 2; all have infinitely many quadratic points.

1.4. Statement of main results

In this paper, we shall be concerned with the case of d=3d=3. Jeon, Kim, and Schweizer [Jeon2004Cubic] determined the torsion structures that appear infinitely often as one varies over all elliptic curves over all cubic number fields (i.e., they classified the pairs (M,MN)(M,MN) for which X1(M,MN)X_{1}(M,MN) is trigonal, and show there are no X1(M,MN)X_{1}(M,MN) which admit a degree 3 map to a positive rank elliptic curve over \mathbb{Q}). In [jeon2011familiescubic], Jeon, Kim, and Lee constructed infinite families of elliptic curves realizing each of these torsion structures by finding models of the relevant trigonal modular curves. The first author and Najman [derickxNajman:TorsionCyclicCubic] classified the torsion groups of elliptic curves over cubic fields with Galois group /3\mathbb{Z}/3\mathbb{Z}, complex cubic fields, and totally real cubic fields with Galois group S3S_{3}.

In 2010, Najman [najman2012torsion] discovered the first example of sporadic torsion: the elliptic curve E/E/\mathbb{Q} with Cremona label 162b1 satisfies E((ζ9)+)tors/21E(\mathbb{Q}(\zeta_{9})^{+})_{{\operatorname{tors}}}\cong\mathbb{Z}/21\mathbb{Z}, and is the only elliptic curve defined over \mathbb{Q} which admits a KK-rational 21-torsion point. He then classified the possible torsion subgroups of elliptic curves defined over \mathbb{Q} when considered over some cubic number field KK. The consideration of the base change of elliptic curves defined over \mathbb{Q} to other number fields leads to sharper results concerning the torsion subgroups which appear over number fields KK (see [alvaro2013fieldofdefn]). A computer generated table of sporadic points is given in [vanHoeij:LowDegreePlaces].

Parent [Parent:cubic-torsion-french, Parent:no-17-torsion] proved that cubic torsion points of prime order pp do not exist for p>13p>13 (reliant on Kato’s [Kato:p-adic-hodge-zeta] subsequent generalization of Kolyvagyn’s theorem to quotients of J1(N)J_{1}(N)). Momose [Momose:quadratic-torsion, Theorem B] ruled out the cyclic torsion for the cases N1=27,64N_{1}=27,64. It was formally conjectured in [wang2015thesis, Conjecture 1.1.2] that the only possible torsion structures for elliptic curves over KK are the ones identified by Jeon–Kim–Schweizer and Najman [Jeon2004Cubic, najman2012torsion]. Wang made progress on this conjecture in his thesis (see [wang:cyclictorsion, wang:cyclictorsion2, wang:cyclictorsion3] for updated versions), and ruled out the existence of cyclic torsion for N1=77,91,143,169.N_{1}=77,91,143,169. Bruin and Najman [bruinN:criterion-to-rule-out-torsion, Theorem 7] ruled out the cyclic cases N1=40,49,55N_{1}=40,49,55 and the non-cyclic case N2=10N_{2}=10.

Our main theorem completes the classification of torsion over cubic number fields.

Theorem A.

Let K/K/\mathbb{Q} be a cubic extension and E/KE/K be an elliptic curve. Then E(K)torsE(K)_{{\operatorname{tors}}} is isomorphic to one of the following 26 groups:

/N1\displaystyle\mathbb{Z}/N_{1}\mathbb{Z} with N1=1,,16,18,20,21,\displaystyle N_{1}=1,\dots,16,18,20,21,
/2/2N2\displaystyle\mathbb{Z}/2\mathbb{Z}\oplus\mathbb{Z}/2N_{2}\mathbb{Z} with N2=1,,7.\displaystyle N_{2}=1,\dots,7.

There exist infinitely many ¯\overline{\mathbb{Q}}-isomorphism classes for each such torsion subgroup except for /21\mathbb{Z}/21\mathbb{Z}. In this case, the elliptic curve with Cremona label 162b1 and minimal Weierstrass equation y2+xy+y=x3x25x+5y^{2}+xy+y=x^{3}-x^{2}-5x+5 over (ζ9)+[x]/(x33x+1)\mathbb{Q}(\zeta_{9})^{+}\cong\mathbb{Q}[x]/(x^{3}-3x+1) is the unique elliptic curve over a cubic field with /21\mathbb{Z}/21\mathbb{Z}-torsion, in particular the point (2α3,2α2)(2\alpha-3,2\alpha-2) has order 2121 where α\alpha is a root of x33x+1x^{3}-3x+1.

Remark 1.5 (Enumeration of remaining cases).

Combining the above work of Parent, Momose, Wang, and Bruin–Najman, the remaining task111Wang’s proofs for N1=22,25,39,40,49,55,65N_{1}=22,25,39,40,49,55,65 contain an error, which we address in Remark 7.5. (The remaining cases from Wang’s papers are correct.) is to determine the cubic points on the modular curves X1(N1)X_{1}(N_{1}) for

N1=21,22,24,25,26,28,30,32,33,35,36,39,45,65,121N_{1}=21,22,24,25,26,28,30,32,33,35,36,39,45,65,121

and on X1(2,2N2)X_{1}(2,2N_{2}) for N2=8,9N_{2}=8,9. (Note that there is a natural map X1(4N)X1(2,2N)X_{1}(4N)\to X_{1}(2,2N) by [JeonK:bielliptic-modular-curves, Section 1], and thus the determination of the cubic points on X1(2,16)X_{1}(2,16) and X1(2,18)X_{1}(2,18) determines the cubic points on X1(32)X_{1}(32) and X1(36)X_{1}(36).)

1.6. Strategy

We first determine the complete list of NN for which J0(N)J_{0}(N) (resp., J1(N)J_{1}(N), resp., J1(2,2N)J_{1}(2,2N)) has rank 0 (Section 3). For the NiN_{i} from Remark 1.5, J1(N1)()J_{1}(N_{1})(\mathbb{Q}) has rank 0, unless N1=65,121N_{1}=65,121.

For the rank 0 cases, we use a variety of techniques:

  • local arguments (§5.1),

  • direct computation of preimages of an Abel–Jacobi map X1(N)(3)()J1(N)()X_{1}(N)^{(3)}(\mathbb{Q})\to J_{1}(N)(\mathbb{Q})5.2),

  • passage to modular curve quotients (§5.3),

  • results on the cuspidal subgroup of J1(N)()J_{1}(N)(\mathbb{Q})4), and

  • explicit description of cusps on X1(N)X_{1}(N) via modular units (§2.9).

For the rank 1 cases, we use a modified formal immersion criterion (§7.1 and §7.3). These methods are expounded on in Sections 4 and 5.

1.7. Outline of paper

In Section 2, we recall background on the arithmetic of curves, modular curves, cuspidal subschemes of modular curves, and modular units. In Section 3, we determine the complete list of NN for which J0(N)J_{0}(N) (resp., J1(N)J_{1}(N), resp., J1(2,2N)J_{1}(2,2N)) has rank 0, and in Section 4, we investigate the torsion subgroup of J1(N)()J_{1}(N)(\mathbb{Q}) via modular symbols. We describe the various techniques used to determine the cubic points on X1(N)X_{1}(N) in Section 5, and we conclude with the determination of cubic points on modular curves in Sections 6 and 7.

1.8. Conventions

Let KK be a field and let X/KX/K be a nice curve i.e., a smooth, proper, geometrically integral scheme of dimension one. For such an XX, let K(X)K(X) denote its function field and let JXJ_{X} denote its Jacobian. For a field LKL\supset K, let XLX_{L} denote the base change of XX to LL. In some cases, we will use this notation when the curve XX is defined over a ring RR. Let DivX\operatorname{Div}X be the group of all divisors of XX, and let Div0X\operatorname{Div}^{0}X be the subgroup of divisors of degree 0. Let DivLX\operatorname{Div}_{L}X (resp. DivL0X\operatorname{Div}^{0}_{L}X) be the group of all divisors (resp. the subgroup of divisors of degree 0) of the curve XLX_{L}. We note that base change gives inclusions DivXDivLX\operatorname{Div}X\hookrightarrow\operatorname{Div}_{L}X and Div0XDivL0X\operatorname{Div}^{0}X\hookrightarrow\operatorname{Div}^{0}_{L}X. For abelian varieties A1,A2A_{1},A_{2} over KK, we will use the notation A1A2A_{1}\sim_{\mathbb{Q}}A_{2} to denote that A1A_{1} is \mathbb{Q}-isogenous to A2A_{2}. We will typically refer to a finite abelian group by its invariants [n1,,nm][n_{1},\dots,n_{m}] (ordered by divisibility).

1.9. Comments on code

This paper has a large computational component. We use the computer algebra programs MapleTM [Maple10], Magma [Magma], and Sage [sagemath] to perform these computations. The code verifying our claims is available at the Github repository

https://github.com/jmorrow4692/SporadicCubicTorsion,

on the final author’s website

https://math.emory.edu/~dzb/DEvHMZB-sporadicTorsion/,

and attached as an ancillary file on the arXiv page for this paper.

1.10. Summary of cases and techniques

In Table 1, we summarize the modular curves from Remark 1.5, their genera, and the proof technique we use to determine the cubic points on these modular curves.

Level Genus Method of proof Genus ofquotient\begin{array}[]{c}\textbf{{Genus of}}\\ \textbf{{quotient}}\end{array}
32 17 Maps to another curve in this table g(X1(2,16))=5g(X_{1}(2,16))=5
36 17 Maps to another curve in this table g(X1(2,18))=7g(X_{1}(2,18))=7
22 6 Local methods at p=3p=36.1) N/A
25 12 Local methods at p=3p=3 N/A
21 5 Direct analysis over \mathbb{Q}6.2) N/A
26 10 Direct analysis over 𝔽3\mathbb{F}_{3} N/A
30 9 Direct analysis over \mathbb{Q} on X0(30)X_{0}(30)6.4) g(X0(30))=3g(X_{0}(30))=3
33 21 Direct analysis over \mathbb{Q} on X0(33)X_{0}(33) g(X0(33))=3g(X_{0}(33))=3
35 25 Direct analysis over \mathbb{Q} on X0(35)X_{0}(35) g(X0(35))=3g(X_{0}(35))=3
39 33 Direct analysis over \mathbb{Q} on X0(39)X_{0}(39) g(X0(39))=3g(X_{0}(39))=3
(2,16) 5 Hecke bound + direct analysis over 𝔽3\mathbb{F}_{3}6.5) N/A
(2,18) 7 Hecke bound + direct analysis over 𝔽5\mathbb{F}_{5} N/A
28 10 Hecke bound + direct analysis over 𝔽3\mathbb{F}_{3}6.6) N/A
24 5 Hecke bound + additional argument (§4.13) + direct analysis over 𝔽5\mathbb{F}_{5} N/A
45 41 Hecke bound + direct analysis over \mathbb{Q} on XH(45)X_{H}(45) (§6.7\S\ref{sec:45}) g(XH(45))=5g(X_{H}(45))=5
65 121 Formal immersion criteria (§\S7.3) g(X0(65))=5g(X_{0}(65))=5
121 526 Formal immersion criteria (§\S7.1) g(X0(121))=6g(X_{0}(121))=6
Table 1. Summary of genera of modular curves and proof techniques

2. Background

In this section, we recall some basic definitions concerning curves and the definition of certain modular curves.

2.1. Geometry of curves

First, we review a few definitions from the geometry of curves. Let XX be a smooth proper geometrically connected curve defined over a field KK.

Definition 2.2.

The gonality γ(X)\gamma(X) of XX is the minimal degree among all finite morphisms XK1X\to\mathbb{P}^{1}_{K}. We say that a closed point PXP\in X has degree dd if [K(P):K]=d[K(P):K]=d.

Remark 2.3.

If XX admits a map f:XK1f\colon X\to\mathbb{P}^{1}_{K} of degree dd, or a map f:XEf\colon X\to E of degree dd, where EE is an elliptic curve with positive rank over KK, then XX also admits infinitely many points of degree dd (arising from fibers of ff). Conversely, if d<γ(X)/2d<\gamma(X)/2, then XX admits only finitely many points of degree dd [frey:curves-with-infinitely-many, Prop. 1]; if the Jacobian of XX has rank 0 over KK (and in particular XX does not admit a non-constant map to an elliptic curve with positive rank over KK), then in fact for d<γ(X)d<\gamma(X), XX admits only finitely many points of degree dd [derickxS:quintic-sextic-torsion, Proposition 2.3].

Definition 2.4.

For a positive integer dd, we define the dthd^{\text{th}}-symmetric power of XX to be X(d):=Xd/SdX^{(d)}:=X^{d}/S_{d} where SdS_{d} is the symmetric group on dd letters. The KK-points of X(d)X^{(d)} correspond to effective KK-rational divisors on XX of degree dd. In particular, a point of X/KX/K of degree dd gives rise to a divisor of degree dd, and thus a point of X(d)(K)X^{(d)}(K), and we will often identify a degree dd point of XX with a divisor of degree dd without distinguishing notation.

If X(d)(K)X^{(d)}(K) is non-empty, then a fixed KK-rational divisor EE of degree dd gives rise to a corresponding Abel–Jacobi map

fd,E:X(d)JX,DDE.f_{d,E}\colon X^{(d)}\to J_{X},\quad D\mapsto D-E.

2.5. Modular curves

We now list the various modular curves we will study.

The modular curve X1(M,MN)X_{1}(M,MN) is the moduli space whose non-cuspidal KK-rational points classify elliptic curves over KK together with independent points P,QP,Q of order MM and MNMN. By setting M=1M=1, we encounter the modular curves X1(N):=X1(1,N)X_{1}(N):=X_{1}(1,N) whose non-cuspidal KK-rational points parametrize elliptic curves over KK which have a torsion point of exact order NN defined over KK.

The modular curve X0(N)X_{0}(N) is the moduli space whose non-cuspidal KK-rational points classify elliptic curves with a cyclic subgroup of order NN (or equivalently, a cyclic isogeny of degree NN).

For a subgroup Γ1(N)HΓ0(N)\Gamma_{1}(N)\subseteq H\subseteq\Gamma_{0}(N), we can form the “intermediate” modular curve XH(N)X_{H}(N). This curve is a quotient of X1(N)X_{1}(N) by a subgroup of Aut(X1(N))\operatorname{Aut}(X_{1}(N)), and (roughly) parameterizes elliptic curves whose mod NN Galois representation has image contained in HH (see [RouseZB:2Adic, Lemma 2.1]).

Remark 2.6 (Models of XHX_{H}).

For computational purposes, we need explicit equations for many of these modular curves. For small values of NN, the Magma intrinsic SmallModularCurve(NN) produces smooth models for the modular curves X0(N)X_{0}(N). For larger values of NN (e.g., N=65,121N=65,121), we use work of Ozman–Siksek [ozman:quadraticpoints] to find canonical models for X0(N)X_{0}(N). We use the algorithm from [DvHZ, Section 2] to compute an equation for a quotient of X1(45)X_{1}(45) via relations between modular units; see Subsection 6.7.

There are several ways to compute models for X1(M,MN)X_{1}(M,MN). We use [sutherland2012optimizedX1] for the equations for X1(N)X_{1}(N) and for the jj-map j:X1(N)X(1)j\colon X_{1}(N)\rightarrow X(1). These models are generally singular. We can compute models for X1(M,MN)X_{1}(M,MN) by taking the normalization of (a particular component of) the fiber product X1(M)×X(1)X1(MN)X_{1}(M)\times_{X(1)}X_{1}(MN). A priori, the fiber product produces a singular model XX. We can desingularize using the canonical map and the Magma intrinsic CanonicalImage(X,CanonicalMap(X)). We can also compute a model for X1(2,2N)X_{1}(2,2N) by computing a quotient of X1(4N)X_{1}(4N) with [DvHZ, Section 2], or we can use [derickxS:quintic-sextic-torsion, Section 3] (we checked that our models are birational to theirs).

2.7. Cuspidal subschemes of modular curves

Our analysis of cubic points will heavily rely on understanding the cuspidal subscheme of X1(N)X_{1}(N) and X0(N)X_{0}(N).

Lemma 2.8.

Let N5N\geq 5 be a positive integer, and let R=[1/2N]R=\mathbb{Z}[1/2N].

  1. (1)

    The cuspidal subscheme of X1(N)RX_{1}(N)_{R} is isomorphic to

    dN(μN/d×/d)/[1],\bigsqcup_{d\mid N}(\mu_{N/d}\times\mathbb{Z}/d\mathbb{Z})^{\prime}/[-1],

    where the prime notation refers to points of maximal order.

  2. (2)

    The cuspidal subscheme of X0(N)RX_{0}(N)_{R} is isomorphic to

    dN(μgcd(d,N/d))\bigsqcup_{d\mid N}(\mu_{\gcd(d,N/d)})^{\prime}

    where the prime notation refers to points of maximal order.

Proof.

This can be directly computed from [derickx:thesis, Chapter 1, Sections 1.3, 1.4, and 2.2] and [derickx2014gonality, Footnote 5]. ∎

2.9. Modular units

In practice, we will work with some model of X1(N)X_{1}(N) (canonical or singular), and will need to explicitly determine the cusps on our model (e.g., to find an explicit basis for J1(N)()torsJ_{1}(N)(\mathbb{Q})_{{\operatorname{tors}}}). The naive approach is to compute the poles of the jj-map but it has high degree when NN is large. Modular units are a useful alternative.

Definition 2.10.

A non-zero element of (X1(N))\mathbb{Q}(X_{1}(N)) is called a modular unit if all of its poles and roots are cusps. Let 1(N)(X1(N))/×\mathcal{F}_{1}(N)\subset\mathbb{Q}(X_{1}(N))/\mathbb{Q}^{\times} be the group of modular units modulo ×\mathbb{Q}^{\times}.

A basis of 1(N)\mathcal{F}_{1}(N) mod ×\mathbb{Q}^{\times} is given in [derickx2014gonality, Conjecture 1] which was proved by Streng [streng2015generators, Theorem 1]. The relevance for our purposes is that this basis is expressed in terms of the same coordinates used in the defining equations for X1(N)X_{1}(N) from [sutherland:X1NTables], and we have their divisors as well [vanHoeijHanson]. For further discussion of modular units, we refer the reader to [derickx2014gonality, Section 2] and [kubertLangModularUnits].

3. Modular Jacobians of rank 0

In this section, we compute, with proof, the complete list of NN for which J0(N)()J_{0}(N)(\mathbb{Q}) (resp., J1(N)()J_{1}(N)(\mathbb{Q}), resp., J1(2,2N)()J_{1}(2,2N)(\mathbb{Q})) has rank 0. This extends the computation of [derickx2014gonality, Lemma 1 (3)] and [derickxS:quintic-sextic-torsion, Theorem 4.1].

Theorem 3.1.

Let S0S_{0} be the set

{1,,36,38,,42,44,,52,54,55,56,59,60,62,63,64,66,68,69,70,71,72,75,76,78,80,81,84,87,90,94,95,96,98,100,104,105,108,110,119,120,126,132,140,144,150,168,180},\{1,\ldots,36,38,\ldots,42,44,\ldots,52,54,55,56,59,60,62,63,64,66,68,69,70,71,72,75,76,78,\\ 80,81,84,87,90,94,95,96,98,100,104,105,108,110,119,120,126,132,140,144,150,168,180\},

and let S1S_{1} be the set

{1,,21,24,25,26,27,30,33,35,42,45}.\{1,\ldots,21,24,25,26,27,30,33,35,42,45\}.

Then the following are true.

  1. (1)

    The rank of J0(N)()J_{0}(N)(\mathbb{Q}) is zero if NS0N\in S_{0}.

  2. (2)

    The rank of J1(N)()J_{1}(N)(\mathbb{Q}) is zero if NS0{63,80,95,104,105,126,144}N\in S_{0}-\{63,80,95,104,105,126,144\}.

  3. (3)

    The rank of J1(2,2N)()J_{1}(2,2N)(\mathbb{Q}) is zero if NS1N\in S_{1}.

Under the assumption of the BSD conjecture the converses of parts (1), (2) and (3) are also true.

In preparation of the proof, we make a series of remarks about computing ranks of modular abelian varieties.

Remark 3.2 (Analytic ranks).

Let AA be a simple factor of J1(N)J_{1}(N). By Kolyvagyn’s theorem (improved to J1(N)J_{1}(N) by Kato [Kato:p-adic-hodge-zeta, Corollary 14.3]) we know that if L(A,1)L(A,1), the LL-series of AA evaluated at 11, is non-zero, then the rank of A()A(\mathbb{Q}) is zero. The BSD conjecture implies that the converse of this statement is also true. A provably correct computation of whether L(A,1)0L(A,1)\neq 0 is implemented in Magma: Decomposition(JOne(N)) computes the simple factors of J1(N)J_{1}(N), and for a simple factor AA, IsZeroAt(LSeries(A),1) provably computes whether L(A,1)0L(A,1)\neq 0. (This is similar to the approach of [derickxS:quintic-sextic-torsion, Theorem 4.1].)

For simple factors of J0(N)J_{0}(N) this computation is quite fast, even for relatively large NN. For factors of J1(N)J_{1}(N) this computation is much slower, even for NN in the 100,,180100,\ldots,180 range.

Remark 3.3 (Winding quotients).

A faster alternative to working directly with the LL-series of the simple factors of J(N)J_{*}(N), where * is either 0 or 1, is the following. The winding quotient Je(N)J_{e}(N) of J(N)J_{*}(N) is the largest quotient with analytic rank zero, and may be described as Je(N)=J(N)/IeJ(N)J_{e}(N)=J_{*}(N)/I_{e}J_{*}(N), where e={0,}H1(X(N)(),)e=\{0,\infty\}\in H_{1}(X_{*}(N)(\mathbb{C}),\mathbb{Q}) is the “winding element” and IeI_{e} is the annihilator of ee in the Hecke algebra 𝐓\mathbf{T} (see [Merel:uniformity, Prop. 1] and [derickx2017small, Definition 4.1 and Theorem 4.2]). In particular, J(N)()J_{*}(N)(\mathbb{Q}) has rank zero if and only if the projection of the image of ee under the Hecke algebra 𝐓\mathbf{T} onto the cuspidal subspace spans it.

By [LarioS:hecke, Theorem 5.1], the Hecke algebra 𝐓\mathbf{T} on Γ(N)\Gamma_{*}(N) is generated by TnT_{n} with

nk12[SL2():Γ(N)]n\leq\frac{k}{12}\cdot\left[\operatorname{SL}_{2}(\mathbb{Z})\colon\Gamma_{*}(N)\right]

(where we note that their proof, as written, also works for Γ1(N)\Gamma_{1}(N)), so this computation is finite and easily implemented in Magma; see master-ranks.m for more detail.

Remark 3.4 (Moonshine and Ogg’s Jack Daniels Challenge).

The subspace of weight 2 newforms with sign of functional equation equal to 1-1 is the subspace fixed by the Fricke involution, and thus for pp prime, J0(p)J_{0}(p) is isogenous to a product A×J0+(p)A\times J_{0}^{+}(p), where J0+(p)J_{0}^{+}(p) is the Jacobian of the quotient X0+(p)X^{+}_{0}(p) of X0(p)X_{0}(p) by the Fricke involution, and where, necessarily, each factor of AA has even rank and each factor of J0(p)J_{0}(p) has odd rank. In particular, if J0(p)()J_{0}(p)(\mathbb{Q}) has rank 0, then X0+(p)X^{+}_{0}(p) has genus 0.

As is well celebrated, Ogg [Ogg:Automorphismes-de-courbes-modulaires, Remarque 1] proved that for pp prime, X0+(p)X^{+}_{0}(p) has genus zero if and only if pp divides the order of the Monster group. Hence, if J0(N)()J_{0}(N)(\mathbb{Q}) has rank 0 and pp is a prime divisor of NN, then pp divides

24632059761121331719232931414759712^{46}\cdot 3^{20}\cdot 5^{9}\cdot 7^{6}\cdot 11^{2}\cdot 13^{3}\cdot 17\cdot 19\cdot 23\cdot 29\cdot 31\cdot 41\cdot 47\cdot 59\cdot 71

(and offered a bottle of Jack Daniels for an explanation of the coincidence).

We note that this approach does not generalize to composite level NN. For example, J0(28)()J_{0}(28)(\mathbb{Q}) has rank 0, but X0+(28)X_{0}^{+}(28) is the elliptic curve X0(14)X_{0}(14). It is still true that one can deduce the signs of the functional equations of newforms via Atkin–Lehner, but the presence of oldforms contributes additional genus to X0+(N)X_{0}^{+}(N).

Proof of Theorem 3.1.

For NS0N\in S_{0}, we compute that J0(N)()J_{0}(N)(\mathbb{Q}) has rank zero via Remark 3.2 (using Remark 3.3 to check our computation), and for any integer of the form NpN\cdot p with NS0N\in S_{0} and pp a prime divisor of the order of the Monster group, we again similarly compute that the rank is non-zero via Magma. This proves part (1).

For part (2), if J1(N)()J_{1}(N)(\mathbb{Q}) has rank 0, then J0(N)()J_{0}(N)(\mathbb{Q}) also has rank 0 (but not conversely), so we again check using Remark 3.2 (and Remark 3.3 for some of the larger NN) for which NS0N\in S_{0} J1(N)()J_{1}(N)(\mathbb{Q}) has rank 0.

For part (3), by [JeonK:bielliptic-modular-curves, Section 1], there is an isomorphism XΔ(4N)X1(2,2N)X_{\Delta}(4N)\to X_{1}(2,2N), with Δ:={±1,±(2N+1)}\Delta:=\{\pm 1,\pm(2N+1)\}, and thus a surjection X1(4N)X1(2,2N)X_{1}(4N)\to X_{1}(2,2N). Thus, if J1(4N)()J_{1}(4N)(\mathbb{Q}) has rank 0, then J1(2,2N)()J_{1}(2,2N)(\mathbb{Q}) also has rank 0. On the other hand, there are surjective maps X1(2,2N)XΔ(4N)X0(4N)X_{1}(2,2N)\cong X_{\Delta}(4N)\to X_{0}(4N) and X1(2,2N)X1(2N)X_{1}(2,2N)\to X_{1}(2N), so if either J0(4N)()J_{0}(4N)(\mathbb{Q}) or J1(2N)()J_{1}(2N)(\mathbb{Q}) have positive rank, then J1(2,2N)()J_{1}(2,2N)(\mathbb{Q}) also has positive rank. Cases (1) and (2) thus determine the rank of J1(2,2N)()J_{1}(2,2N)(\mathbb{Q}) unless N=20,26,36N=20,26,36. The remaining three cases we do by hand using the isomorphism XΔ(4N)X1(2,2N)X_{\Delta}(4N)\to X_{1}(2,2N) and Magma’s intrinsic JH to compute the rank of JΔ(4N)J_{\Delta}(4N) via Remark 3.2. ∎

See the file master-ranks.m for code verifying these computations.

4. Computing rational torsion on modular Jacobians

This section explains how to compute the rational torsion of a modular Jacobian. For an additional technique see [ozman:quadraticpoints, Section 4].

4.1. Local bounds on torsion via reduction

Let KK be a number field, 𝔭\mathfrak{p} a prime of KK lying over a rational prime p>2p>2, and A/KA/K an abelian variety with good reduction at 𝔭\mathfrak{p}. Then by [katz:galois-properties-of-torsion, Appendix], if the ramification index e𝔭e_{\mathfrak{p}} is less than p1p-1, the reduction map A(K)torsA(𝔽𝔭)A(K)_{{\operatorname{tors}}}\rightarrow A(\mathbb{F}_{\mathfrak{p}}) is injective. The GCD of #A(𝔽𝔭)\#A(\mathbb{F}_{\mathfrak{p}}), as 𝔭\mathfrak{p} ranges over such primes, gives a naive upper bound on A(K)torsA(K)_{{\operatorname{tors}}}. For rational torsion on modular Jacobians, one can quickly compute #A(𝔽p)\#A(\mathbb{F}_{{p}}) via the coefficients of the corresponding modular forms, which Magma has packaged into the intrinsic TorsionMultiple.

Comparing orders is usually insufficient (e.g., if A()torsA(\mathbb{Q})_{{\operatorname{tors}}} is cyclic and A(𝔽p)A(\mathbb{F}_{{p}}) is non cyclic, the output of TorsionMultiple is usually larger than #A()tors\#A(\mathbb{Q})_{{\operatorname{tors}}}; see Example 4.2). To improve these bounds, we compute the “GCD” of the groups A(𝔽p)A(\mathbb{F}_{{p}}) for various p{p}; more precisely, given abelian groups A1,,AnA_{1},\ldots,A_{n}, we define the GCD(A1,,An)\operatorname{GCD}(A_{1},\ldots,A_{n}) to be the largest abelian group AA such that each AiA_{i} contains a subgroup isomorphic to AA.

Example 4.2 (Torsion on J1(21)J_{1}(21)).

To demonstrate this, we consider the rational torsion on A=J1(21)A=J_{1}(21), which is a 55-dimensional modular abelian variety. The output of the intrinsic TorsionMultiple(JOne(21)) yields a bound of 728728, but A(𝔽5)A(\mathbb{F}_{5}) has invariants [2184][2184] and A(𝔽11)A(\mathbb{F}_{11}) has invariants [14,6916][14,6916], and these groups have GCD [364][364]. We computed A(𝔽p)A(\mathbb{F}_{p}) by reducing a model for X1(21)X_{1}(21) modulo pp and using Magma ’s ClassGroup intrinsic; see the Magma file master-21.m.

4.3. Better local bounds on torsion via Eichler–Shimura and modular symbols

The naive approach above does not incorporate the action of complex conjugation, and slows quickly as the genus grows. We describe here a substantial improvement, derived from the Eichler–Shimura relation. For an intermediate modular curve XH(N)X_{H}(N), let JH(N)J_{H}(N) denote its Jacobian.

For a prime q2Nq\nmid 2N, let TqT_{q} be the qq-th Hecke operator. By the Eichler–Shimura relation, the kernel of the operator

Tqqq1:JH(N)(¯)torsJH(N)(¯)torsT_{q}-q\langle q\rangle-1\colon J_{H}(N)({\overline{\mathbb{Q}}})_{{\operatorname{tors}}}\to J_{H}(N)({\overline{\mathbb{Q}}})_{{\operatorname{tors}}}

contains the prime to qq torsion in JH(N)()J_{H}(N)(\mathbb{Q}) (cf. [derickx2017small, Proposition 5.2] and [DiamondI:modularCurves, page 87]). Additionally, let

τ:JH(N)(¯)JH(N)(¯)\tau\colon J_{H}(N)(\overline{\mathbb{Q}})\to J_{H}(N)(\overline{\mathbb{Q}})

be complex conjugation. Then it is clear that τ1\tau-1 vanishes on JH(N)()J_{H}(N)(\mathbb{Q}), and thus also vanishes on JH(N)()torsJ_{H}(N)(\mathbb{Q})_{{\operatorname{tors}}}.

Thus, for a finite set of primes q1,,qnq_{1},\ldots,q_{n} (still coprime to 2N2N) with n2n\geq 2,

MH:=JH(N)(¯)tors[Tq1q1q11,,Tqnqnqn1,τ1]M_{H}:=J_{H}(N)(\overline{\mathbb{Q}})_{{\operatorname{tors}}}[T_{q_{1}}-q_{1}\langle q_{1}\rangle-1,\ldots,T_{q_{n}}-q_{n}\langle q_{n}\rangle-1,\tau-1] (4.3.1)

contains JH(N)()torsJ_{H}(N)(\mathbb{Q})_{{\operatorname{tors}}}. We can efficiently compute MHM_{H} as follows. Under the uniformization

JH(N)()H1(XH(N)(),)/H1(XH(N)(),)J_{H}(N)(\mathbb{C})\cong H_{1}(X_{H}(N)(\mathbb{C}),\mathbb{C})/H_{1}(X_{H}(N)(\mathbb{C}),\mathbb{Z})

we can identify the geometric torsion as

JH(N)(¯)torsH1(XH(N)(),)/H1(XH(N)(),).J_{H}(N)(\overline{\mathbb{Q}})_{{\operatorname{tors}}}\cong H_{1}(X_{H}(N)(\mathbb{C}),\mathbb{Q})/H_{1}(X_{H}(N)(\mathbb{C}),\mathbb{Z}).

While it does not make sense to ask this identification to be Galois equivariant, it does commute with the Hecke and diamond operators, and with complex conjugation. The right hand side lends itself very well to explicit computations with Sage using modular symbols (and is very fast, since the computations are now essentially linear algebra). This is implemented in the module CuspidalClassgroup from [derickx:X1Ncode], and is also available in our accompanying code (see 1.9).

We will call this upper bound on rational torsion we get from computing MHM_{H} the Hecke bound.

To demonstrate how the Hecke bound produces better bounds than the local bounds from Subsection 4.1, we will consider the following two examples.

Example 4.4 (Torsion on J1(28)J_{1}(28)).

Consider the rational torsion on A=J1(28)A=J_{1}(28), which is a 1010-dimensional modular abelian variety. The intrinsic TorsionMultiple(JOne(28)) produces a bound of 359424359424. We compute that A(𝔽3)A(\mathbb{F}_{3}) has invariants [4,4,24,936][4,4,24,936] and A(𝔽5)A(\mathbb{F}_{5}) has invariants [2,2,8,8,312,936][2,2,8,8,312,936], which have GCD [4,4,24,936][4,4,24,936]. The Hecke bound give an upper bound of [2,4,12,936][2,4,12,936], which improves upon the GCD bound; see the Sage file torsionComputations.py.

Example 4.5 (Torsion on J1(2,18)J_{1}(2,18)).

Consider the rational torsion on A=J1(2,18)A=J_{1}(2,18), which is a 77-dimensional modular abelian variety. We compute that A(𝔽5)A(\mathbb{F}_{5}) has invariants [6,84,252][6,84,252] and A(𝔽7)A(\mathbb{F}_{7}) has invariants [3,6,6,126,126][3,6,6,126,126], which have GCD [6,42,126][6,42,126]. The Hecke bound give an upper bound of [2,42,126][2,42,126], which improves upon the GCD bound; see the Sage file torsionComputations.py.

4.6. Cuspidal torsion and a generalized Conrad–Edixhoven–Stein conjecture

By the Manin–Drinfeld theorem [maninParabolic, drinfeldTwotheorems], cuspidal divisors are torsion; conversely, it is expected that J()torsJ(\mathbb{Q})_{{\operatorname{tors}}} is cuspidal for a modular Jacobian JJ. More precisely, we conjecture that the torsion on J1(N)()J_{1}(N)(\mathbb{Q}) is generated by the Gal(¯/)\operatorname{Gal}(\overline{\mathbb{Q}}/\mathbb{Q})-orbits of cusps (see Conjecture 4.15).

This is a conjecture of Conrad–Edixhoven–Stein [ConradES:J-connected-fibers, Conjecture 6.2.2] for the modular Jacobian J1(p)J_{1}(p) where pp is a prime. The authors prove this conjecture for all primes p157p\leq 157 except for p=29,97,101,109,p=29,97,101,109, and 113113, and the case of p=29p=29 was proved in [derickx2017small, Theorem 6.4]. Moreover, their conjecture is true for all primes pp such that J1(p)()J_{1}(p)(\mathbb{Q}) has rank 0. For composite NN, the torsion subgroup of J1(N)J_{1}(N) generated by Gal(¯/)\operatorname{Gal}(\overline{\mathbb{Q}}/\mathbb{Q})-orbits of cusps has been studied, and in some special cases, the Conrad–Edixhoven–Stein conjecture has been proved; see the following works [yu:cuspidalclassnumber, takagi:classnumberp, takai:cuspidalclassnumber3power, hazama:cuspidalclassnumber, csirik:KernelEisensteinIdeal, takagi:classnumber2power, yang:modularunits, yang:CuspidalTorsion, sun:cuspidalclassnumber, chen:CuspidalTorsion, takagi:CupsidalClassNumber2p, ohta:CuspidalTorsion, takagi:CuspidalTorsion2p].

In this subsection, we will use the construction of MHM_{H} from Equation (4.3.1) and the results from Subsection 2.9 to determine when the rational torsion on certain modular Jacobians is cuspidal. First, we need to establish some definitions.

Definition 4.7.

We define the following subgroups of DivXH(N)\operatorname{Div}X_{H}(N).

  • Let DivcXH(N)\operatorname{Div}^{\operatorname{c}}X_{H}(N) denote the free abelian group generated by the cusps of XH¯X_{H_{\overline{\mathbb{Q}}}}.

  • Let DivcXH(N):=DivcXH(N)DivXH(N)\operatorname{Div}^{\operatorname{c}}_{\mathbb{Q}}X_{H}(N):=\operatorname{Div}^{\operatorname{c}}X_{H}(N)\cap\,\operatorname{Div}_{\mathbb{Q}}X_{H}(N) denote the subgroup generated by the Gal(¯/)\operatorname{Gal}(\overline{\mathbb{Q}}/\mathbb{Q})-orbits of cusps.

Definition 4.8.

We define the following quotients of the divisor groups from Definition 4.7.

  • Let princXH(N)\,\operatorname{prin}^{\operatorname{c}}X_{H}(N) denote the principal divisors in Div0,cXH(N)\operatorname{Div}^{0,\operatorname{c}}X_{H}(N), and let ClcXH(N):=Div0,cXH(N)/princXH(N)\operatorname{Cl}^{\operatorname{c}}X_{H}(N):=\operatorname{Div}^{0,\operatorname{c}}X_{H}(N)/\operatorname{prin}^{\operatorname{c}}X_{H}(N) denote the quotient.

  • Let princXH(N)\operatorname{prin}^{\operatorname{c}}_{\mathbb{Q}}X_{H}(N) denote the principal divisors in Div0,cXH(N)\operatorname{Div}^{0,\operatorname{c}}_{\mathbb{Q}}X_{H}(N), and let ClcXH(N):=Div0,cXH(N)/princXH(N)\operatorname{Cl}^{\operatorname{c}}_{\mathbb{Q}}X_{H}(N):=\operatorname{Div}^{0,\operatorname{c}}_{\mathbb{Q}}X_{H}(N)/\operatorname{prin}^{\operatorname{c}}X_{H}(N) denote the quotient.

With these definitions it is clear that determining if the rational torsion on the modular Jacobian JHJ_{H} is cuspidal boils down to whether ClcXH(N)=JH()tors\operatorname{Cl}^{\operatorname{c}}_{\mathbb{Q}}X_{H}(N)=J_{H}(\mathbb{Q})_{{\operatorname{tors}}}.

Subsection 4.3 gave an inclusion JH()torsMHJ_{H}(\mathbb{Q})_{{\operatorname{tors}}}\subset M_{H} (the “Hecke bound”). In order to get lower bounds, we need to compute ClcXH(N)\operatorname{Cl}^{\operatorname{c}}_{\mathbb{Q}}X_{H}(N), which we now describe. Let G:=Gal((ζN)/)(/N)×G:=\operatorname{Gal}(\mathbb{Q}(\zeta_{N})/\mathbb{Q})\cong(\mathbb{Z}/N\mathbb{Z})^{\times} be the Galois group of the cyclotomic field obtained by adjoining an NN-th root of unity to \mathbb{Q}. To determine the group ClcXH(N)\operatorname{Cl}^{\operatorname{c}}_{\mathbb{Q}}X_{H}(N), we first use results and code from [derickx2014gonality] and actions of diamond operators to compute ClcXH(N)\operatorname{Cl}^{\operatorname{c}}X_{H}(N), and from here we take GG-invariants as Div0,cXH(N)=(Div0,cXH(N))G\operatorname{Div}^{0,\operatorname{c}}_{\mathbb{Q}}X_{H}(N)=(\operatorname{Div}^{0,\operatorname{c}}X_{H}(N))^{G}. An implementation to compute ClcX1(N)\operatorname{Cl}^{\operatorname{c}}_{\mathbb{Q}}X_{1}(N) is available at [vanHoeij:X1Ncode], explained in [vanHoeijHanson], and a Sage version extending this to ClcXH(N)\operatorname{Cl}^{\operatorname{c}}_{\mathbb{Q}}X_{H}(N) can be found in the module MiscellaneousFunctions from [derickx:X1Ncode]. We illustrate the utility of this code in the following examples.

Example 4.9 (Torsion on J1(28)J_{1}(28)).

The code finds ClcX1(28)[2,4,12,936]\operatorname{Cl}^{\operatorname{c}}_{\mathbb{Q}}X_{1}(28)\cong[2,4,12,936]. Combining this with the Hecke bound from Example 4.4 gives ClcX1(28)=J1(28)()tors\operatorname{Cl}^{\operatorname{c}}_{\mathbb{Q}}X_{1}(28)=J_{1}(28)(\mathbb{Q})_{{\operatorname{tors}}}, and in particular, all of the rational torsion on J1(28)J_{1}(28) is cuspidal; see the Sage file torsionComputations.py.

Example 4.10 (Torsion on J1(2,18)J_{1}(2,18)).

The code finds ClcX1(2,18)[2,42,126]\operatorname{Cl}^{\operatorname{c}}_{\mathbb{Q}}X_{1}(2,18)\cong[2,42,126]. Combining this with the Hecke bound from Example 4.5 gives ClcX1(2,18)=J1(2,18)()tors\operatorname{Cl}^{\operatorname{c}}_{\mathbb{Q}}X_{1}(2,18)=J_{1}(2,18)(\mathbb{Q})_{{\operatorname{tors}}}, and in particular, all of the rational torsion on J1(2,18)J_{1}(2,18) is cuspidal; see the Sage file torsionComputations.py.

To study the question of whether the rational torsion on the modular Jacobian is cuspidal in more generality, we proceed as follows. Since (ClcXH(N))GJH()(\operatorname{Cl}^{\operatorname{c}}X_{H}(N))^{G}\subset J_{H}(\mathbb{Q}), it is necessary that the map

Div0,cXH(N)=(Div0,cXH(N))G(ClcXH(N))G\operatorname{Div}^{0,\operatorname{c}}_{\mathbb{Q}}X_{H}(N)=(\operatorname{Div}^{0,\operatorname{c}}X_{H}(N))^{G}\to(\operatorname{Cl}^{\operatorname{c}}X_{H}(N))^{G}

is surjective in order for the rational torsion to be cuspidal.

For the remainder of this section, we will focus our attention on determining the rational torsion on modular curves X1(N)X_{1}(N) and X1(2,2N)X_{1}(2,2N) with modular Jacobians J1(N)J_{1}(N) and J1(2,2N)J_{1}(2,2N), respectively. As a first step, we show that for N55N\leq 55, the above map on divisors is surjective.

Proposition 4.11.

Let N55N\leq 55 be an integer. Then ClcX1(N)=(ClcX1(N))G\operatorname{Cl}^{\operatorname{c}}_{\mathbb{Q}}X_{1}(N)=(\operatorname{Cl}^{\operatorname{c}}X_{1}(N))^{G}.

Proof.

We verify the result by computing the maps

(Div0,cX1(N))G(ClcX1(N))G(\operatorname{Div}^{0,\operatorname{c}}X_{1}(N))^{G}\to(\operatorname{Cl}^{\operatorname{c}}X_{1}(N))^{G}

in terms of modular symbols and the computation showed it was surjective in all cases. ∎

For code verifying these claims as well as those in Theorem 4.13, Corollary 4.14, and Proposition 4.16 see the Sage file torsionComputations.py.

For the main theorem in this section, we will also need the following lemma.

Lemma 4.12.

Let RR be a commutative ring MM be an RR-module whose cardinality is finite and C,M,TC,M^{\prime},T be RR-submodules of MM such that CMC\subseteq M^{\prime}, CTC\subseteq T and MT=CM^{\prime}\cap T=C. Assume that for all maximal ideals mm of RR one has (M/C)[m]=(M/C)[m](M/C)[m]=(M^{\prime}/C)[m], then T=CT=C.

Proof.

By taking the quotient by CC, one may assume that C=0C=0. Since TT is of finite cardinality, it is isomorphic to i=1kT[mi]\otimes_{i=1}^{k}T[m_{i}^{\infty}] for some finite sequence of maximal ideals m1,,mkm_{1},\dots,m_{k}. Let m=mim=m_{i} be one of these maximal ideals. Due to the equalities M[m]=M[m]M[m]=M^{\prime}[m] and MT=0M^{\prime}\cap T=0, one gets T[m]=M[m]T=M[m]T=0T[m]=M[m]\cap T=M^{\prime}[m]\cap T=0, and thence T[m]=0T[m^{\infty}]=0. Since this holds for all mim_{i}, T=0T=0. ∎

Theorem 4.13.

Let N55N\leq 55, N54N\neq 54 be an integer. Then

ClcX1(N)=J1(N)()tors\operatorname{Cl}^{\operatorname{c}}_{\mathbb{Q}}X_{1}(N)=J_{1}(N)(\mathbb{Q})_{{\operatorname{tors}}}

If N=54N=54, then the index of ClcX1(N)\operatorname{Cl}^{\operatorname{c}}_{\mathbb{Q}}X_{1}(N) in J1(N)()torsJ_{1}(N)(\mathbb{Q})_{{\operatorname{tors}}} is a divisor of 33.

Proof.

This was verified using a Sage computation, which we now describe. We compute the MM from Equation (4.3.1) for the modular Jacobians J1(N)J_{1}(N) for N55N\leq 55. For N24,32,33,40,48,54N\neq 24,32,33,40,48,54, our computation shows that MClcX1(N)M\subseteq\operatorname{Cl}^{\operatorname{c}}X_{1}(N) holds and hence

(ClcX1(N))GJ1(N)()torsMG(ClcX1(N))G.(\operatorname{Cl}^{\operatorname{c}}X_{1}(N))^{G}\subseteq J_{1}(N)(\mathbb{Q})_{{\operatorname{tors}}}\subseteq M^{G}\subseteq(\operatorname{Cl}^{\operatorname{c}}X_{1}(N))^{G}.

Therefore, the result follows from Proposition 4.11 except for the cases of N=24,24,32,33,40,48,54N=24,24,32,33,40,48,54. For these cases, define M=MClcX1(N)M^{\prime}=M\cap\operatorname{Cl}^{\operatorname{c}}X_{1}(N). The Sage computation then shows that for all primes pp that divide #M\#M, we have (M/ClcX1(N))[p]=(M/ClcX1(N))[p](M/\operatorname{Cl}^{\operatorname{c}}_{\mathbb{Q}}X_{1}(N))[p]=(M^{\prime}/\operatorname{Cl}^{\operatorname{c}}_{\mathbb{Q}}X_{1}(N))[p] holds for all the above NN except for 5454. The theorem follows by applying Lemma 4.12 with C=ClcX1(N)C=\operatorname{Cl}^{\operatorname{c}}_{\mathbb{Q}}X_{1}(N) and T=J1(N)()torsT=J_{1}(N)(\mathbb{Q})_{{\operatorname{tors}}}.

For the N=54N=54, we computed the index i:=[M+ClcX1(N):ClcX1(N)]i:=[M+\operatorname{Cl}^{\operatorname{c}}X_{1}(N):\operatorname{Cl}^{\operatorname{c}}X_{1}(N)]; taking Gal(¯/)\operatorname{Gal}(\overline{\mathbb{Q}}/\mathbb{Q})-invariants of the sequence

0ClcX1(N)M+ClcX1(N)(M+ClcX1(N))/ClcX1(N)00\to\operatorname{Cl}^{\operatorname{c}}X_{1}(N)\to M+\operatorname{Cl}^{\operatorname{c}}X_{1}(N)\to(M+\operatorname{Cl}^{\operatorname{c}}X_{1}(N))/\operatorname{Cl}^{\operatorname{c}}X_{1}(N)\to 0

shows that [(M+ClcX1(N))Gal(¯/):(ClcX1(N))Gal(¯/)][(M+\operatorname{Cl}^{\operatorname{c}}X_{1}(N))^{\operatorname{Gal}(\overline{\mathbb{Q}}/\mathbb{Q})}:(\operatorname{Cl}^{\operatorname{c}}X_{1}(N))^{\operatorname{Gal}(\overline{\mathbb{Q}}/\mathbb{Q})}] divides 33. Since J1()tors=(M+ClcX1(N))Gal(¯/)J_{1}(\mathbb{Q})_{{\operatorname{tors}}}=(M+\operatorname{Cl}^{\operatorname{c}}X_{1}(N))^{\operatorname{Gal}(\overline{\mathbb{Q}}/\mathbb{Q})} and ClcX1(N)Gal(¯/)=ClcX1(N))\operatorname{Cl}^{\operatorname{c}}X_{1}(N)^{\operatorname{Gal}(\overline{\mathbb{Q}}/\mathbb{Q})}=\operatorname{Cl}^{\operatorname{c}}_{\mathbb{Q}}X_{1}(N)), we use this index to get the multiplicative upper bound of 33 mentioned in the second statement. ∎

The only N55N\leq 55 such that J1(N)J_{1}(N) has positive rank are 37,4337,43 and 5353 so as a corollary we immediately get that:

Corollary 4.14.

Let N55N\leq 55 be an integer. If N37,43,53,N\neq 37,43,53, or 5454, then

ClcX1(N)=J1(N)().\operatorname{Cl}^{\operatorname{c}}_{\mathbb{Q}}X_{1}(N)=J_{1}(N)(\mathbb{Q}).

Based on the results of Theorem 4.13, we make the following conjecture, which is a generalization of the conjecture of Conrad–Edixhoven–Stein [ConradES:J-connected-fibers, Conjecture 6.2.2].

Conjecture 4.15.

For any integer NN, we have that

ClcX1(N)=J1(N)()tors.\operatorname{Cl}^{\operatorname{c}}_{\mathbb{Q}}X_{1}(N)=J_{1}(N)(\mathbb{Q})_{{\operatorname{tors}}}.

Using the same proof as in Theorem 4.13, we can prove the following.

Proposition 4.16.

Let N16N\leq 16 be an integer. Then

ClcX1(2,2N)=J1(2,2N)()tors.\operatorname{Cl}^{\operatorname{c}}_{\mathbb{Q}}X_{1}(2,2N)=J_{1}(2,2N)(\mathbb{Q})_{{\operatorname{tors}}}.

Again see the Sage file torsionComputations.py.

4.17. Local to global failures

For J0(N)J_{0}(N), the local methods described above often give only an upper bound instead of the true torsion. With enough work one can sometimes fix this; see Example 4.18 below. For another compelling example see [ozman:quadraticpoints, Subsection 5.5] — the curve X0(45)X_{0}(45) is a plane quartic, and they use the explicit description of J0(45)(¯)[2]J_{0}(45)({\overline{\mathbb{Q}}})[2] via bitangents to bridge the discrepancy between their local bounds and the true torsion.

Example 4.18 (Torsion on J0(30)J_{0}(30)).

Consider the rational torsion on A=J0(30)A=J_{0}(30), which is a 33-dimensional modular abelian variety. The rational cuspidal divisors generate a subgroup with invariants [2,4,24][2,4,24]. Locally, A(𝔽7)A(\mathbb{F}_{7}) has invariants [2,2,4,48][2,2,4,48] and A(𝔽23)A(\mathbb{F}_{23}) has invariants [2,12,24,24][2,12,24,24]; these groups have GCD [2,2,4,24][2,2,4,24]. The Hecke bound and the additional argument from Theorem 4.13 do not improve this.

The curve X0(30)X_{0}(30) is hyperelliptic and admits the model y2=f(x)y^{2}=f(x), where

f(x)=(x2+3x+1)(x2+6x+4)(x4+5x3+11x2+10x+4),f(x)=(x^{2}+3x+1)\cdot(x^{2}+6x+4)\cdot(x^{4}+5x^{3}+11x^{2}+10x+4),

and we can exploit the explicit description of the 2-torsion of the Jacobian of a hyperelliptic curve. The geometric 2-torsion J0(30)(¯)[2]J_{0}(30)({\overline{\mathbb{Q}}})[2] is Galois-equivariantly in bijection with even order subsets of the set of Weierstrass points modulo the subset of all Weierstrass points (see e.g., [gross:Hanoi-lectures-on-the-arithmetic-of-hyperelliptic-curves, Section 6]), and one can compute the rational torsion by taking Galois invariants. See the file functions.m for a short routine twoTorsionRank which computes J()[2]J(\mathbb{Q})[2] for the Jacobian of a hyperelliptic curve. Using this routine, we find that rank𝔽2J0(30)()[2]=3\operatorname{rank}_{\mathbb{F}_{2}}J_{0}(30)(\mathbb{Q})[2]=3, and hence the rational torsion on J0(30)J_{0}(30) is cuspidal and isomorphic to [2,4,24][2,4,24].

Note that if JJ is the Jacobian of a hyperelliptic curve defined by an odd degree polynomial f(x)f(x), rank𝔽2J()[2]=i\operatorname{rank}_{{\mathbb{F}}_{2}}J(\mathbb{Q})[2]=i, where i+1i+1 is the number of factors f(x)f(x); this is no longer true in general, as J0(30)J_{0}(30) demonstrates.

See the Magma file master-30.m for more details.

4.19. Tables of cuspidal torsion subgroups for J1(N)J_{1}(N) and J1(2,2N)J_{1}(2,2N)

We conclude this section by giving the structure of ClcX1(N)\operatorname{Cl}^{\operatorname{c}}_{\mathbb{Q}}X_{1}(N) for 10N5510\leq N\leq 55 and of ClcX1(2,2N)\operatorname{Cl}^{\operatorname{c}}_{\mathbb{Q}}X_{1}(2,2N) for 5N165\leq N\leq 16 in terms of its invariant factor decomposition in Tables 2 and 3. For code, see the Sage file torsionComputations.py.

Remark 4.20.

In [yang:modularunits], Yang determines the cuspidal torsion on J1(N)J_{1}(N) generated by \infty-cusps (cf. [derickx2017small, Definition 5.4]). By comparing his results with Table 2, we can find NN for which one needs non-\infty cusps and/or non-rational cusps to generate the torsion on J1(N)()J_{1}(N)(\mathbb{Q}). For example, by [yang:modularunits, Table 1], the \infty-cusps on X1(20)X_{1}(20) generate a subgroup isomorphic to [20][20], but by Theorem 4.13 and Table 2, we see that J1(20)()J_{1}(20)(\mathbb{Q}) is cuspidal and isomorphic to [60][60]; in particular, the \infty-cusps do not generate all of the rational torsion on J1(20)J_{1}(20). A similar situation occurs for X1(21)X_{1}(21), which we discuss in Subsection 6.2.

NN ClcX1(N)\operatorname{Cl}^{\operatorname{c}}_{\mathbb{Q}}X_{1}(N) NN ClcX1(N)\operatorname{Cl}^{\operatorname{c}}_{\mathbb{Q}}X_{1}(N) NN ClcX1(N)\operatorname{Cl}^{\operatorname{c}}_{\mathbb{Q}}X_{1}(N)
11 [5] 27 [3,3,52497] 42 [182,1092,131040]
13 [19] 28 [2,4,12,936] 43 [2,1563552532984879906]
14 [6] 29 [4,4,64427244] 44 [4,620,3100,6575100]
15 [4] 30 [4,8160] 45 [3,9,36,16592750496]
16 [2,10] 31 [10,1772833370] 46 [408991,546949390174]
17 [584] 32 [2,2,2,4,120,11640] 47 [3279937688802933030787]
18 [21] 33 [5,42373650] 48 [2,2,2,2,4,40,40,240,1436640]
19 [4383] 34 [8760,595680] 49 [7,52367710906884085342]
20 [60] 35 [13,109148520] 50 [5,1137775,47721696825]
21 [364] 36 [12,252,7812] 51 [8,1168,7211322610146240]
22 [5,775] 37 [160516686697605] 52 [4,28,532,7980,17470957140]
23 [408991] 38 [9,4383,33595695] 53 [182427302879183759829891277]
24 [2,2,120] 39 [7,31122,3236688] 5454^{\prime} [3,3,3,9,9,1102437,1529080119]
25 [227555] 40 [2,2,2,8,120,895440] 55 [5,550,8972396739917886000]
26 [133,1995] 41 [107768799408099440]
Table 2. Cuspidal torsion in J1(N)()J_{1}(N)(\mathbb{Q}) for N=11N=11 and 13N5513\leq N\leq 55 (all torsion if N54N\neq 54).
(2,2N)(2,2N) ClcX1(2,2N)\operatorname{Cl}^{\operatorname{c}}_{\mathbb{Q}}X_{1}(2,2N) (2,2N)(2,2N) ClcX1(2,2N)\operatorname{Cl}^{\operatorname{c}}_{\mathbb{Q}}X_{1}(2,2N) (2,2N)(2,2N) ClcX1(2,2N)\operatorname{Cl}^{\operatorname{c}}_{\mathbb{Q}}X_{1}(2,2N)
(2,10)(2,10) [6] (2,18)(2,18) [2,42,126] (2,26)(2,26) [2,14,266,3990,11970]
(2,12)(2,12) [4] (2,20)(2,20) [4,60,120] (2,28)(2,28) [2,4,4,4,8,24,936,936]
(2,14)(2,14) [2,2,6,18] (2,22)(2,22) [2,10,1550,4650] (2,30)(2,30) [2,2,2,2,2,24,8160,8160]
(2,16)(2,16) [2,20,20] (2,24)(2,24) [2,4,4,120,240] (2,32)(2,32) [2,2,2,4,4,24,120,23280,23280]
Table 3. (Cuspidal) torsion on J1(2,2N)()J_{1}(2,2N)(\mathbb{Q}) for 5N165\leq N\leq 16.

5. Methods for determining cubic points on curves

In this section, we describe the variety of methods we utilize to determine the cubic points on the modular curves X1(N)X_{1}(N).

5.1. Local methods

In some cases, we can deduce that there are no non-cuspidal cubic points on X1(N)X_{1}(N) simply by determining points on X1(N)(𝔽pi)X_{1}(N)(\mathbb{F}_{p^{i}}) for i=1,2,i=1,2, and 33 for some prime p2Np\nmid 2N.

Suppose that XX is a curve of gonality at least 4 and at least one rational point, and that its Jacobian JXJ_{X} satisfies rank J()J(\mathbb{Q}) = 0. Fix a point X()\infty\in X(\mathbb{Q}). Since XX has gonality at least 4, the Abel–Jacobi maps

fi,i:X(i)JX,DDif_{i,i\infty}\colon X^{(i)}\hookrightarrow J_{X},\,D\mapsto D-i\infty

are injective for i=1,2,3i=1,2,3. Moreover, since the rank of JX()J_{X}(\mathbb{Q}) is zero, the reduction map JX()JX(𝔽p)J_{X}(\mathbb{Q})\hookrightarrow J_{X}(\mathbb{F}_{p}) is injective for p>2p>2 (see Remark 4.1). We thus get a commutative diagram

X(i)()\textstyle{X^{(i)}(\mathbb{Q})\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}JX()\textstyle{J_{X}(\mathbb{Q})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}X(i)(𝔽p)\textstyle{X^{(i)}({\mathbb{F}}_{p})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}JX(𝔽p)\textstyle{J_{X}({\mathbb{F}}_{p})}

of injections. In particular, the reduction maps X(i)()X(i)(𝔽p)X^{(i)}(\mathbb{Q})\hookrightarrow X^{(i)}({\mathbb{F}}_{p}) are injective for i=1,2,3i=1,2,3; if there is a prime pp such that these maps are also surjective, then we have determined X(i)()X^{(i)}(\mathbb{Q}). One can verify surjectivity by checking cardinalities, i.e., checking if

#X(i)()#X(i)(𝔽p).\#X^{(i)}(\mathbb{Q})\geq\#X^{(i)}({\mathbb{F}}_{p}).

In practice, #X(i)(𝔽p)\#X^{(i)}({\mathbb{F}}_{p}) can be determined very quickly in Magma (even using a singular model, by working with places instead of points), and we often a priori have a lower bound on #X(i)()\#X^{(i)}(\mathbb{Q}) coming from an explicit description of cusps (e.g., Lemma 2.8).

See Subsection 6.1 for examples.

5.2. Direct analysis of preimages of an Abel–Jacobi map

When X1(N)X_{1}(N) has gonality at least 4 and J1(N)()J_{1}(N)(\mathbb{Q}) has rank 0, one can compute the finitely many preimages of an Abel–Jacobi map ι:X1(N)(3)()J1(N)()\iota\colon X_{1}(N)^{(3)}(\mathbb{Q})\to J_{1}(N)(\mathbb{Q}). For values of NN such that the genus of X1(N)X_{1}(N) and the size of J1(N)()J_{1}(N)(\mathbb{Q}) is not too large, it is possible to do this directly over \mathbb{Q}: fixing a base point X1(N)()\infty\in X_{1}(N)(\mathbb{Q}), a divisor DJ1(N)()D\in J_{1}(N)(\mathbb{Q}) is in the image of the Abel–Jacobi map EE3E\mapsto E-3\infty if and only if the linear system |D+3||D+3\infty|\neq\emptyset. One can compute |D+3||D+3\infty| via Magma’s RiemannRoch intrinsic. If |D+3|=|D+3\infty|=\emptyset then we disregard it; otherwise it will contain a single effective divisor EE of degree 3. Thus as DD ranges over J1(N)()J_{1}(N)(\mathbb{Q}), we eventually compute all of the effective degree 3 divisors (and hence the image of Abel–Jacobi). We will refer to this as direct analysis over \mathbb{Q}.

Direct analysis over \mathbb{Q} can be very slow (cf. Remark 6.3). A much faster method from [vanHoeij:LowDegreePlaces, Footnote 1] works as follows. The diagram from Subsection 5.1

X(3)()\textstyle{X^{(3)}(\mathbb{Q})\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ι\scriptstyle{\iota}redX\scriptstyle{{\operatorname{red}}_{X}}JX()\textstyle{\ignorespaces\ignorespaces\ignorespaces\ignorespaces J_{X}(\mathbb{Q})}redJ\scriptstyle{{\operatorname{red}}_{J}}X(3)(𝔽p)\textstyle{X^{(3)}({\mathbb{F}}_{p})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ιp\scriptstyle{\iota_{p}}JX(𝔽p)\textstyle{J_{X}({\mathbb{F}}_{p})}

commutes, so the image of ιp\iota_{p} contains the reduction of the image of ι\iota. It thus suffices to:

  1. (1)

    compute the image of ιp\iota_{p} (which is generally very fast),

  2. (2)

    compute the preimage of imιp\operatorname{im}\iota_{p} under redJ{\operatorname{red}}_{J} (also fast), and

  3. (3)

    compute which elements of redJ1(imιp){\operatorname{red}}_{J}^{-1}\left(\operatorname{im}\iota_{p}\right) are in the image of ι\iota.

We will refer to this approach as direct analysis over 𝔽p{\mathbb{F}}_{p}. The set redJ1(imιp){\operatorname{red}}_{J}^{-1}\left(\operatorname{im}\iota_{p}\right) of divisors which are “locally in the image of Abel–Jacobi” is generally much smaller than JX()J_{X}(\mathbb{Q}), so step (3) is much faster. (Equivalently, since each map is injective, one can compute the intersection (imredJ)(imιp).\left(\operatorname{im}{\operatorname{red}}_{J}\right)\cap\left(\operatorname{im}\iota_{p}\right).) One could further speed up the computation by repeating this procedure at several primes; see the Mordell–Weil sieve [ozman:quadraticpoints, Section 6]. Implementing this was unnecessary for our results.

See Subsections 6.2 and 6.5 for examples.

5.3. Direct analysis on non-trigonal curve quotients

While direct analysis works in principle, we encounter many values of NN where the genus of X1(N)X_{1}(N) and the size of J1(N)J_{1}(N) are large (for example, g(X1(45))=41g(X_{1}(45))=41 and 16128153482112#J1(45)()16128153482112\mid\#J_{1}(45)(\mathbb{Q})), and hence working directly with X1(N)(3)()X_{1}(N)^{(3)}(\mathbb{Q}) and J1(N)()J_{1}(N)(\mathbb{Q}) is hard. Instead, we consider a morphism X1(N)XX_{1}(N)\to X where XX is non-trigonal and perform the direct analysis on XX. For details on how to find a model for XX, see Remark 2.6.

Remark 5.4 (Cubic points on hyperelliptic curves).

A curve XX of genus g2g\leq 2 is trigonal as it admits a base-point free g31g^{1}_{3} by Riemann–Roch, and a non-hyperelliptic curve of genus g=3g=3 or 44 is trigonal. However, a hyperelliptic curve of g3g\geq 3 is not trigonal, and thus has finitely many cubic points. In this case, X(3)()X^{(3)}(\mathbb{Q}) is still infinite; it contains the image of X()×X(2)()X(\mathbb{Q})\times X^{(2)}(\mathbb{Q}) (i.e., divisors of the form P+Q+QιP+Q+Q^{\iota}, where PX()P\in X(\mathbb{Q}) and where ι\iota is the hyperelliptic involution), and the complement of this subset contains the finitely many cubic points. Therefore, when we search for non-trigonal curve quotients, we either need a genus g3g\geq 3 hyperelliptic curve or certain non-hyperelliptic curves of g5g\geq 5.

In the hyperelliptic case, we need to know that the image of a cubic point does not reduce to a non-cuspidal rational point plus another divisor. We encounter four genus 3 hyperelliptic cases, namely X0(N)X_{0}(N) for N=30,33,35,N=30,33,35, and 3939. By [kenkuIsomClasses, Theorem 1], we know that the rational points on the curves X0(N)X_{0}(N) for these NN are all cuspidal.

See Subsection 6.4 for examples.

5.5. Formal immersion criteria

When the rank of J1(N)()J_{1}(N)(\mathbb{Q}) is positive, we apply formal immersion criteria of [derickx2017small] to determine all of the points on X1(N)(3)()X_{1}(N)^{(3)}(\mathbb{Q}). The underlying ideas of using formal immersion in the study of rational and higher degree points on modular curves comes from the foundational works of Mazur [mazur1978primedegreeisogeny] and Kamienny [kamienny:torsionpointshigherdegree].

To begin, we define formal immersions. Throughout this subsection, let RR be a discrete valuation ring with perfect residue field κ\kappa and fraction field KK.

Definition 5.6.

Let ϕ:XY\phi\colon X\to Y be a morphism of Noetherian schemes and xXx\in X a point and y=f(x)Yy=f(x)\in Y. Then ϕ\phi is a formal immersion at xx if the induced morphism of complete local rings ϕ^:𝒪Y,y^𝒪X,x^\widehat{\phi}^{*}\colon\widehat{\mathcal{O}_{Y,y}}\to\widehat{\mathcal{O}_{X,x}} is surjective.

The main reason we consider formal immersions is the following lemma.

Lemma 5.7 ([derickx2017small, Lemma 2.2]).

Let X,YX,Y be Noetherian schemes. Let RR be a Noetherian local ring with maximal ideal 𝔪\mathfrak{m} and residue field κ=R/𝔪\kappa=R/\mathfrak{m}. Suppose that f:XYf\colon X\to Y is a formal immersion at a point xX(κ)x\in X(\kappa) and suppose that P,QX(R)P,Q\in X(R) are two points such that x=Pκ=Qκx=P_{\kappa}=Q_{\kappa} and f(P)=f(Q)f(P)=f(Q). Then P=QP=Q.

To verify that a morphism is a formal immersion at a point, we will use the following criterion.

Proposition 5.8 ([derickx2017small, Proposition 3.7]).

Let CC be a smooth projective curve over RR with geometrically connected generic fiber and Jacobian JJ. Let yC(d)(κ)y\in C^{(d)}(\kappa) be a point and write y=j=1mnjyjy=\sum_{j=1}^{m}n_{j}y_{j} with yjC(dj)(κ¯)y_{j}\in C^{(d_{j})}(\overline{\kappa}) distinct and m,n1,,nmm,n_{1},\dots,n_{m}\in\mathbb{N}. Let t:JAt\colon J\to A be a map of abelian schemes over RR such that t(J1(R))={0}t(J^{1}(R))=\{0\}, where J1(R)J^{1}(R) denotes the kernel of the reduction map J(R)J(κ)J(R)\to J(\kappa). Let qjq_{j} be a uniformizer at yjy_{j}, ee be a positive integer and ω1,,ωet(Cot0Aκ¯)Cot0(Jκ¯)\omega_{1},\dots,\omega_{e}\in t^{*}(\operatorname{Cot}_{0}A_{\overline{\kappa}})\subset\operatorname{Cot}_{0}(J_{\overline{\kappa}}). For 1ie1\leq i\leq e and 1jm1\leq j\leq m, let a(ωi,qj,nj):=(a1(ωi),,anj(ωi))a(\omega_{i},q_{j},n_{j}):=(a_{1}(\omega_{i}),\dots,a_{n_{j}}(\omega_{i})) be the row vector of the first njn_{j} coefficients of ωi\omega_{i}’s qjq_{j}-expansion.

Then tfd,y:Cκ(d)Aκt\circ f_{d,y}\colon C_{\kappa}^{(d)}\to A_{\kappa} is a formal immersion at yy if the matrix

A:=(a(ω1,q1,n1)a(ω1,q2,n2)a(ω1,q1,nm)a(ω2,q1,n1)a(ω2,q2,n2)a(ω2,q1,nm)a(ωe,q1,n1)a(ωe,q2,n2)a(ωe,q1,nm))A:=\begin{pmatrix}a(\omega_{1},q_{1},n_{1})&a(\omega_{1},q_{2},n_{2})&\cdots&a(\omega_{1},q_{1},n_{m})\\ a(\omega_{2},q_{1},n_{1})&a(\omega_{2},q_{2},n_{2})&\cdots&a(\omega_{2},q_{1},n_{m})\\ \vdots&\vdots&\ddots&\vdots\\ a(\omega_{e},q_{1},n_{1})&a(\omega_{e},q_{2},n_{2})&\cdots&a(\omega_{e},q_{1},n_{m})\\ \end{pmatrix} (5.8.1)

has rank dd. If ω1,,ωe\omega_{1},\dots,\omega_{e} generate tCot0(Aκ¯)t^{*}\operatorname{Cot}_{0}(A_{\overline{\kappa}}), then the previous statement is an equivalence.

To apply Proposition 5.8, we will take κ\kappa to have odd characteristic and AA to be a rank 0 abelian variety; since torsion injects under reduction, the hypothesis t(J1(R))={0}t(J^{1}(R))=\{0\} is satisfied.

Our plan for N=65N=65 and 121121 is as follows. We first show, either via a brute force search or using the Hasse bound, that a cubic point on X1(N)X_{1}(N) will reduce modulo some prime p2Np\nmid 2N to a degree 3 divisor which is supported on cusps. Then, we know that the reduction modulo pp of a cubic point on X1(N)X_{1}(N) will map to a degree 3 cuspidal divisor on X0(N)X_{0}(N). Using Magma’s intrinsic Decomposition(JZero(NN)), we find a rank zero quotient AA of J0(N)J_{0}(N), and then we verify that the morphism X0(N)(3)AX_{0}(N)^{(3)}\to A is a formal immersion at these degree 3 divisors.

See Subsections 7.1 and 7.3 for greater detail.

6. Cubic points on modular curves with rank 0 Jacobians

In this section, we determine the cubic points on the modular curves X1(N)X_{1}(N) from Remark 1.5 which have Jacobians J1(N)J_{1}(N) of rank 0.

For claims about gonality of X1(N)X_{1}(N), see [derickx2014gonality, Table 1], for claims about the number and degrees of cusps on X1(N)X_{1}(N) or X0(N)X_{0}(N), see Lemma 2.8, and for claims about the torsion on J1(N)()J_{1}(N)(\mathbb{Q}) and J1(2,2N)()J_{1}(2,2N)(\mathbb{Q}) see Corollary 4.14, Proposition 4.16, and Tables 2 and 3.

6.1. Local methods — the cases X1(22)X_{1}(22) and X1(25)X_{1}(25).

The modular curve X1(22)X_{1}(22) is a genus 6 tetragonal curve with 10 rational cusps. Using Magma, we determine that modulo 3, X1(22)X_{1}(22) has 10 degree 1 places, 0 degree 2 places, and 0 degree 3 places. Since we found 10 rational points, we immediately conclude that X1(22)X_{1}(22) has no cubic points. An identical argument handles X1(25)X_{1}(25).

See the Magma files master-22.m and master-25.m for code verifying these claims.

6.2. Direct analysis over \mathbb{Q} — the case of X1(21)X_{1}(21)

The modular curve X1(21)X_{1}(21) has genus 5 and is tetragonal. We prove that the only \mathbb{Q}-rational points of X1(21)(3)X_{1}(21)^{(3)} arise from combinations of the 6 rational cusps, the 2 quadratic cusps, and the 2 cubic points D0D_{0} and D0D^{\prime}_{0} corresponding to the elliptic curve E/E/\mathbb{Q} with Cremona label 162b1, which has a point of exact order 2121 over (ζ9)+\mathbb{Q}(\zeta_{9})^{+}. We follow the method of Subsection 5.2.

From Example 4.2, we know that J1(21)()J_{1}(21)(\mathbb{Q}) is cyclic of order 364. Differences of the rational cusps only generate the subgroup of order 364/2, but D:=D03D:=D_{0}-3\infty has order 364364 (where \infty is any rational cusp).

Let

f3,3:X1(21)(3)()J1(21)(),EE3f_{3,3\infty}\colon X_{1}(21)^{(3)}(\mathbb{Q})\to J_{1}(21)(\mathbb{Q}),\,E\mapsto E-3\infty

be an Abel–Jacobi map. For each point nDJ1(21)()nD\in J_{1}(21)(\mathbb{Q}), nDnD is in the image of f3,3f_{3,3\infty} if and only if the linear system |nD+3||nD+3\infty| is nonempty, and the Magma intrinsic RiemannRochSpace will check whether this is true. Moreover, since X1(21)X_{1}(21) is tetragonal, dim|nD+3|0\dim|nD+3\infty|\leq 0, so if it is nonempty, it follows that |nD+3|={E}|nD+3\infty|=\{E\} for some effective divisor EE of degree 3; in Magma one can easily compute EE and its support. We find that nDnD is in the image of f3,3f_{3,3\infty} if and only if

n{0,1,5,8,12,14,16,22,38,40,42,58,60,64,65,67,76,84,91,92,94,101,104,111,118,121,123,138,145,147,167,172,183,188,190,200,201,202,204,206,214,226,228,230,234,241,246,248,250,252,254,268,272,274,278,280,282,284,289,291,292,294,297,298,306,308,315,318,326,328,332,335,338,352,360,362};n\in\{0,1,5,8,12,14,16,22,38,40,42,58,60,64,65,67,76,84,91,92,94,101,104,111,\\ 118,121,123,138,145,147,167,172,183,188,190,200,201,202,204,206,214,226,228,\\ 230,234,241,246,248,250,252,254,268,272,274,278,280,282,284,289,291,292,294,\\ 297,298,306,308,315,318,326,328,332,335,338,352,360,362\};

each of these correspond to combinations of known rational, quadratic, and cubic points, and in particular nDnD is the image of a cubic point (under f3,3f_{3,3\infty}) if and only if n=1,183n=1,183.

See the Magma file master-21.m for code verifying these claims.

Remark 6.3.

While direct analysis over \mathbb{Q} works in this example, it took over a week to complete. In contrast, we quickly verify our results via a direct analysis over 𝔽5\mathbb{F}_{5}, which took around 8 seconds.

6.4. Direct analysis over \mathbb{Q} on a non-trigonal curve quotient — the cases X1(N)X_{1}(N) for N{30,33,35,39}N\in\{30,33,35,39\}.

The curve X1(30)X_{1}(30) has genus 9 and is 6-gonal. While the genus is not prohibitively large, we instead work on the genus 3, hyperelliptic curve X0(30)X_{0}(30), which is not trigonal. Using the Magma intrinsic SmallModularCurve(30), we have the affine equation

X0(30):y2+(x4x3x2)y=3x7+19x6+60x5+110x4+121x3+79x2+28x+4.X_{0}(30)\colon y^{2}+(-x^{4}-x^{3}-x^{2})y=3x^{7}+19x^{6}+60x^{5}+110x^{4}+121x^{3}+79x^{2}+28x+4.

We note that the map X0(30)(3)J0(30)X_{0}(30)^{(3)}\to J_{0}(30) is not injective since X0(30)(2)X_{0}(30)^{(2)} contains a rational curve (see Remark 5.4). However, the map is still injective on cubic points of X0(30)X_{0}(30), which suffices for our purposes.

In Example 4.18, we proved that J0(30)()/2×/4×/24J_{0}(30)(\mathbb{Q})\cong\mathbb{Z}/2\mathbb{Z}\times\mathbb{Z}/4\mathbb{Z}\times\mathbb{Z}/24\mathbb{Z}, and so next we perform a direct analysis over \mathbb{Q} and find 4848 cubic points on X0(30)X_{0}(30), which are all necessarily non-cuspidal by Lemma 2.8(2). To conclude, we compute the jj-invariants of the cubic points and check to see if there is a 30-torsion point on a twist of the corresponding curve EE. To check this, it suffices to show that for some prime pp not dividing 30NEΔ30N_{E}\Delta, where NEN_{E} is the conductor of the elliptic curve EE and Δ\Delta the discriminant of the cubic number field where EE is defined, and for all primes 𝔭\mathfrak{p} above pp, EE modulo 𝔭\mathfrak{p} and its twists do not have an 𝔽𝔭\mathbb{F}_{\mathfrak{p}}-rational point of order 3030. For each of the 48 cubic points, we verify this, which tells us that these cubic points on X0(30)X_{0}(30) do not lift to X1(30)X_{1}(30), and thus, there are no cubic points on X1(30)X_{1}(30). We use a similar combination of methods to handle N=33N=33, 3535 and 3939.

See the Magma files master-30.m, master-33.m, master-35.m, and master-39.m for code.

6.5. Hecke bounds and direct analysis over 𝔽p\mathbb{F}_{p}— the cases X1(2,16)X_{1}(2,16), X1(24)X_{1}(24), and X1(2,18)X_{1}(2,18).

The modular curve X1(2,16)X_{1}(2,16) has genus 5 and is tetragonal. Using the model from Derickx–Sutherland [derickxS:quintic-sextic-torsion], we find 88 rational cusps and 22 quadratic cusps on X1(2,16)X_{1}(2,16). The local bound on the torsion is [2,2,20,20][2,2,20,20], and the Hecke bound improves this to [2,20,20][2,20,20]. The cusps generate a subgroup isomorphic to [2,20,20][2,20,20] and hence they generate all of the torsion on J1(2,16)()J_{1}(2,16)(\mathbb{Q}). We also know that these cusps give rise to 136136 rational points on X1(2,16)(3)X_{1}(2,16)^{(3)}.

To conclude, we compute the intersection of the image of an Abel–Jacobi with our known subgroup modulo 3. This gives 136 cubic divisors, and so by Subsection 5.1, we have completely determined the rational points on X1(2,16)(3)X_{1}(2,16)^{(3)}. A very similar argument handles X1(24)X_{1}(24) and X1(2,18)X_{1}(2,18).

See the Magma files master-2-16.m, master-24.m and master-2-18.m and the Sage file
torsionComputations.py for code verifying these claims.

6.6. Hecke bounds and direct analysis over 𝔽3\mathbb{F}_{3} — the cases X1(28)X_{1}(28) and X1(26)X_{1}(26).

The modular curve X1(28)X_{1}(28) has genus 10 and is 6-gonal. A direct analysis over \mathbb{Q} takes too much time, and we have no suitable quotient curve.

We know that J1(28)()J_{1}(28)(\mathbb{Q}) is cuspidal, but we would like to find explicit generators for J1(28)()J_{1}(28)(\mathbb{Q}). Via the results on modular units in Subsection 2.9, we determine that the cuspidal divisors on J1(28)()J_{1}(28)(\mathbb{Q}) are generated by the principal divisors together with the cuspidal divisors of degree 11, 22, and 33 (i.e., we do not need cuspidal divisors of degree 6 to generate the torsion). Following Subsection 5.2, we perform direct analysis over 𝔽3\mathbb{F}_{3} and determine that there is only one cubic point locally in the image of Abel–Jacobi, namely the known cubic cusp (strictly speaking: “Galois orbit of cubic cusps”). Therefore, we can conclude that the only cubic point on X1(28)X_{1}(28) is the known cubic cusp. A similar argument handles X1(26)X_{1}(26).

See the Magma files master-28.m and master-26.m, and the Sage file torsionComputations.py for code verifying these claims.

6.7. Hecke bounds and direct analysis over \mathbb{Q} on a non-trigonal curve quotient — the case of X1(45)X_{1}(45)

The curve X1(45)X_{1}(45) has genus 41, so we look for a quotient. It maps to X0(45)X_{0}(45), a non-hyperelliptic genus 3 curve, which is necessarily trigonal and thus has infinitely many cubic points. We work instead with the intermediate genus 5 non-hyperelliptic non-trigonal curve XH(45)X_{H}(45), where HH is the subgroup

H\displaystyle H ={(ab0c)Γ0(45):a is a square modulo 15}.\displaystyle=\left\{\begin{pmatrix}a&b\\ 0&c\end{pmatrix}\in\Gamma_{0}(45):a\text{ is a square modulo }15\right\}.

We use the algorithm from [DvHZ] to compute an equation P(x,y)=0P(x,y)=0 for XH(45)X_{H}(45) as follows. (See also [derickx2014gonality, Example 1].) We construct modular units x,y(X1(45))x,y\in\mathbb{Q}(X_{1}(45)) that are invariant under the diamond action 4\langle 4\rangle, and compute a relation P(x,y)=0P(x,y)=0; we then check its genus to verify that x,yx,y generate (XH(45))\mathbb{Q}(X_{H}(45)). We chose yy to be the image of xx under the diamond action 2\langle 2\rangle so that PP is symmetric, allowing it to be written as P(x,y)=Q(xy,x+y)=0P(x,y)=Q(xy,x+y)=0 for some QQ. Here

Q(u,v)=u3+(v2+7v+7)u2+(2v+3)(v2+5v+3)u+(v2+3v)2Q(u,v)=u^{3}+(v^{2}+7v+7)u^{2}+(2v+3)(v^{2}+5v+3)u+(v^{2}+3v)^{2}

is an equation for X0(45)X_{0}(45). As a check for correctness, we computed an alternative model of XH(45)X_{H}(45) using [zywina:ComputingActionsCuspForms], and checked in Magma that they are isomorphic.

Next, we determine the cubic points on XH(45)X_{H}(45). Since X1(45)X_{1}(45) dominates XH(45)X_{H}(45), JH(45)()J_{H}(45)(\mathbb{Q}) has rank 0, and local computations tell us that that JH(45)()J_{H}(45)(\mathbb{Q}) is isomorphic to a subgroup of

/2×/4×/24×/48.\mathbb{Z}/2\mathbb{Z}\times\mathbb{Z}/4\mathbb{Z}\times\mathbb{Z}/24\mathbb{Z}\times\mathbb{Z}/48\mathbb{Z}.

The Galois orbits of cusps generate a subgroup isomorphic to /2×/4×/48\mathbb{Z}/2\mathbb{Z}\times\mathbb{Z}/4\mathbb{Z}\times\mathbb{Z}/48\mathbb{Z}, and the Hecke bounds from Subsection 4.3 prove that this generates JH(45)()J_{H}(45)(\mathbb{Q}).

We fix a known rational point XH(45)()\infty\in X_{H}(45)(\mathbb{Q}). Via direct analysis over \mathbb{Q}, we determine that f3,3(XH(45)())=f3,3(XH(45)())JH(45)()f_{3,3\infty}(X_{H}(45)(\mathbb{Q}))=f_{3,3\infty}(X_{H}(45)(\mathbb{Q}))\cap J_{H}(45)(\mathbb{Q}) and find that there are 8 non-cuspidal cubic points on XH(45)X_{H}(45). Finally, we lift the cubic points back to X1(45)X_{1}(45) to verify that they do not come from degree 3 points on X1(45)X_{1}(45).

See the Maple files FindModel-XH-45-input and Lift-XY-back-to-X1-45-input, the Magma file master-45.m and the Sage file torsionComputations.py for code verifying these computations.

7. Cubic points on modular curves with positive rank Jacobians

In this section, we use the formal immersion criterion of Subsection 5.5 to determine the cubic points on modular curves X1(N)X_{1}(N) for N=65,121N=65,121.

7.1. The case of X1(121)X_{1}(121)

We will prove that the cubic points on X1(121)X_{1}(121) are cuspidal. Consider the morphism ϕ:X1(121)X0(121)\phi\colon X_{1}(121)\to X_{0}(121), and let ϕ(3):X1(121)(3)X0(121)(3)\phi^{(3)}\colon X_{1}(121)^{(3)}\to X_{0}(121)^{(3)}. The algorithm for X0(N)X_{0}(N) from Ozman–Siksek [ozman:quadraticpoints] provides the canonical model of X0(121)X_{0}(121). This genus 6 curve is not trigonal by [hasegawa1999trigonal, Theorem 3.3]. By Lemma 2.8(2), the cuspidal subscheme of X0(121)X_{0}(121) is isomorphic to the disjoint union of μ1\mu_{1}, μ1\mu_{1}, and (μ11)(\mu_{11})^{\prime}, where the prime notation refers to points of exact order 11, i.e., X0(121)X_{0}(121) has 12 cusps: 2 rational cusps {c0,c}\{c_{0},c_{\infty}\} and a Galois orbit of size 10 defined over (ζ11)\mathbb{Q}(\zeta_{11}).

To begin, we claim that for a cubic point xx on X1(121)X_{1}(121), ϕ(3)(x𝔽5)\phi^{(3)}(x_{\mathbb{F}_{5}}) is equal (as a divisor) to one of

3[c,𝔽5],2[c,𝔽5]+[c0,𝔽5],[c,𝔽5]+2[c0,𝔽5], or 3[c0,𝔽5].3[c_{\infty,\mathbb{F}_{5}}],2[c_{\infty,\mathbb{F}_{5}}]+[c_{0,\mathbb{F}_{5}}],[c_{\infty,\mathbb{F}_{5}}]+2[c_{0,\mathbb{F}_{5}}],\text{ or }3[c_{0,\mathbb{F}_{5}}].

Indeed, by a brute force search, we see that there are no elliptic curves over 𝔽5i\mathbb{F}_{5^{i}} for i=1,2,3i=1,2,3 with a rational 121 torsion point, and so a cubic point on X1(121)X_{1}(121) has bad reduction at each prime above 5, i.e., it must reduce to a sum of cusps (considered as a degree 3 divisor). (The Hasse bound at the prime 55 unfortunately does not a priori exclude the existence of such an elliptic curve.) The image ϕ(3)(x𝔽5)\phi^{(3)}(x_{\mathbb{F}_{5}}) is thus also a sum of cusps. The cuspidal subscheme of X0(121)X_{0}(121) further decomposes modulo 5: the prime 5 splits in (ζ11)\mathbb{Q}(\zeta_{11}) as 2 primes each with inertia degree 5, and so the component (μ11)𝔽5(\mu_{11})^{\prime}_{\mathbb{F}_{5}} splits as two copies of Spec𝔽55\operatorname{Spec}\mathbb{F}_{5^{5}}, i.e., the modulo 5 reduction of the Galois orbit of size 10 is defined over 𝔽55\mathbb{F}_{5^{5}}. Now, our claim follows since c0,𝔽5c_{0,\mathbb{F}_{5}} and c,𝔽5c_{\infty,\mathbb{F}_{5}} are the only cusps defined over 𝔽52\mathbb{F}_{5^{2}} or 𝔽53\mathbb{F}_{5^{3}}.

Using Magma’s intrinsic Decomposition(JZero(121)), we find that

J0(121)E1×E2×E3×E4×X0(11)×X1(11),J_{0}(121)\sim_{\mathbb{Q}}E_{1}\times E_{2}\times E_{3}\times E_{4}\times X_{0}(11)\times X_{1}(11),

and we compute that E1()E_{1}(\mathbb{Q}) has rank 1 and X0(11)(),X1(11)(),X_{0}(11)(\mathbb{Q}),X_{1}(11)(\mathbb{Q}), and Ei()E_{i}(\mathbb{Q}) have rank 0 for i=2,3,4i=2,3,4. Let A:=E2×E3×E4×X1(11)A:=E_{2}\times E_{3}\times E_{4}\times X_{1}(11). Let C1:=3[c],C2:=2[c]+[c0],C3:=[c]+2[c0],C_{1}:=3[c_{\infty}],C_{2}:=2[c_{\infty}]+[c_{0}],C_{3}:=[c_{\infty}]+2[c_{0}], and C4:=3[c0].C_{4}:=3[c_{0}]. Define for i=1,,4i=1,\dots,4, the morphisms μi:X0(121)(3)J0(121)\mu_{i}\colon X_{0}(121)^{(3)}\to J_{0}(121) given by z[zCi]z\mapsto[z-C_{i}] and t:J0(121)At\colon J_{0}(121)\to A, where the latter is projection. By our first claim, we know that ϕ(3)(x𝔽5)\phi^{(3)}(x_{\mathbb{F}_{5}}) is equal to one of C1,𝔽5,C2,𝔽5,C3,𝔽5C_{1,\mathbb{F}_{5}},C_{2,\mathbb{F}_{5}},C_{3,\mathbb{F}_{5}} or C4,𝔽5C_{4,\mathbb{F}_{5}}, and so for the appropriate ii, the image of (tμi)(ϕ(3)(x))(t\circ\mu_{i})(\phi^{(3)}(x)) belongs to the kernel of reduction A()A(𝔽5)A(\mathbb{Q})\to A(\mathbb{F}_{5}). However, since A()A(\mathbb{Q}) is torsion, the kernel of reduction is trivial [katz:galois-properties-of-torsion, Appendix], and so for each appropriate ii, (tμi)(ϕ(3)(x))=0(t\circ\mu_{i})(\phi^{(3)}(x))=0.

To conclude our analysis, we verify that the morphism

τμi:X0(121)(3)A\tau\circ\mu_{i}\colon X_{0}(121)^{(3)}\to A

is a formal immersion at the point Ci,𝔽5C_{i,\mathbb{F}_{5}} using Proposition 5.8. Via Magma’s intrinsic Newform, we can compute a basis for the 1-forms on AA, and so to verify the formal immersion criterion, we need to check that certain 4×34\times 3 matrices (see 5.8.1) have rank 33. We can compute the qq-expansion at cc_{\infty} in Magma, and since the Atkin–Lehner involution ω121\omega_{121} swaps c0c_{0} and cc_{\infty}, we can also directly compute the qq-expansion at c0c_{0} in Magma. We then observe the four matrices (modulo 5) defined in Proposition 5.8 all have rank 3, and thus, (tμi)(t\circ\mu_{i}) is a formal immersion at the above points. Finally, by Lemma 5.7, ϕ(3)(x)\phi^{(3)}(x) is equal to one of 3[c],2[c]+[c0],[c]+2[c0],3[c_{\infty}],2[c_{\infty}]+[c_{0}],[c_{\infty}]+2[c_{0}], or 3[c0]3[c_{0}], and therefore we can conclude that any cubic point on X1(121)X_{1}(121) must be cuspidal.

See the Magma file master-121.m for code verifying these claims.

Remark 7.2.

The argument that ϕ(3)(x𝔽5)\phi^{(3)}(x_{\mathbb{F}_{5}}) is a sum of reductions of rational cusps proceeded by brute force. Typically, one would work at a smaller prime (in this case, 3) and apply the Hasse bound as in Lemma 7.4 below. It turns out that X0(121)(3)AX_{0}(121)^{(3)}\to A is not a formal immersion modulo 3, forcing us to work at a larger prime.

7.3. The case of X1(65)X_{1}(65)

We will prove that the cubic points on X1(65)X_{1}(65) must be cuspidal. Consider the morphism ϕ:X1(65)X0(65)\phi\colon X_{1}(65)\to X_{0}(65), and let ϕ(3):X1(65)(3)X0(65)(3)\phi^{(3)}\colon X_{1}(65)^{(3)}\to X_{0}(65)^{(3)}. Using equations from Ozman–Siksek [ozman:quadraticpoints], we have a model for the genus 5 curve X0(65)X_{0}(65), which is not trigonal [hasegawa1999trigonal, Theorem 3.3]. By Lemma 2.8(2), we have that X0(65)X_{0}(65) has 4 cusps, c0c_{0}, cc_{\infty}, c1/5,c_{1/5}, and c1/13c_{1/13}, all of which are rational.

Unlike the case of N=121N=121, the map X0(65)(3)Je(65)X_{0}(65)^{(3)}\to J_{e}(65) to the winding quotient is not a formal immersion at all cuspidal divisors. Fortunately, the following lemma tells us that we only need to verify the formal immersion criterion at particular sums of cusps on X0(65)X_{0}(65).

Lemma 7.4.

Let ϕ(3):X1(65)(3)X0(65)(3)\phi^{(3)}\colon X_{1}(65)^{(3)}\to X_{0}(65)^{(3)}, and let xx be a cubic point on X1(65)X_{1}(65). Then, ϕ(3)(x𝔽3)\phi^{(3)}(x_{\mathbb{F}_{3}}) is equal to 3[c𝔽3]3[c_{{\mathbb{F}}_{3}}] for some rational cusp cX0(65)()c\in X_{0}(65)(\mathbb{Q}).

Proof.

First, we claim that x𝔽3x_{\mathbb{F}_{3}} must be supported on a sum of cusps using the following Hasse bound computation.

If EE is an elliptic curve over a cubic field KK and 𝔭{\mathfrak{p}} is a prime of KK over a rational prime pp, then, if its reduction modulo 𝔭{\mathfrak{p}} is smooth, it has (by the Hasse bound) at most

#E(𝔽q)2q+(q+1)\#E({\mathbb{F}}_{q})\leq 2{\sqrt{q}}+(q+1)

points, where qp3q\leq p^{3}. In our setting of N=65N=65, the Hasse bound tells us that there cannot exist an elliptic curve over 𝔽3i\mathbb{F}_{3^{i}} for i=1,2,3i=1,2,3 with a rational 6565-torsion point, and thus x𝔽3x_{\mathbb{F}_{3}} must be a degree 3 cuspidal divisor on X1(65)𝔽3X_{1}(65)_{\mathbb{F}_{3}}.

By Lemma 2.8(1), the cuspidal subscheme of X1(65)X_{1}(65) is isomorphic to

(/65)/[1](/5×μ13)/[1](μ5×/13)/[1](μ65)/[1],(\mathbb{Z}/65\mathbb{Z})^{\prime}/[-1]\,\sqcup\,(\mathbb{Z}/5\mathbb{Z}\times\mu_{13})^{\prime}/[-1]\,\sqcup\,(\mu_{5}\times\mathbb{Z}/13\mathbb{Z})^{\prime}/[-1]\,\sqcup\,(\mu_{65})^{\prime}/[-1],

where the prime notation means points of exact order 6565, and each piece reduces to a distinct rational cusp on X0(65)X_{0}(65). We need to analyze the reduction modulo 3 of each piece.

First, no cubic point can reduce to either of the last two pieces. The scheme (μ5×/13)(\mu_{5}\times\mathbb{Z}/13\mathbb{Z})^{\prime} is isomorphic to 12 copies of (μ5)\left(\mu_{5}\right)^{\prime}; the [1][-1] action identifies pairs of copies of (μ5)(\mu_{5})^{\prime}, thus the (μ5×/13)/[1](\mu_{5}\times\mathbb{Z}/13\mathbb{Z})^{\prime}/[-1] piece is isomorphic to 6 copies of μ5\mu_{5}^{\prime}. Since the 5-th cyclotomic polynomial is irreducible modulo 3, (μ5)𝔽3(\mu_{5})^{\prime}_{\mathbb{F}_{3}} is still irreducible; (μ5×/13)/[1](\mu_{5}\times\mathbb{Z}/13\mathbb{Z})^{\prime}/[-1] is thus a sum of quartic points, and our cubic point xx cannot reduce to this component. A similar argument shows that our cubic point xx cannot reduce to ((μ65)/[1])𝔽3((\mu_{65})^{\prime}/[-1])_{\mathbb{F}_{3}}.

Next, the (/5×μ13)/[1](\mathbb{Z}/5\mathbb{Z}\times\mu_{13})^{\prime}/[-1] part breaks up as 4 copies of μ13\mu_{13}^{\prime}, and the [1][-1] action identifies pairs of copies of (μ13)(\mu_{13})^{\prime}. The image of (/5×μ13)/[1](\mathbb{Z}/5\mathbb{Z}\times\mu_{13})^{\prime}/[-1] in X0(65)X_{0}(65) is supported at a single rational cusp. Since each component is isomorphic to μ13\mu_{13}^{\prime} and the 13-th cyclotomic polynomial factors into four cubics over 𝔽3\mathbb{F}_{3}, (μ13)𝔽3(\mu_{13})^{\prime}_{\mathbb{F}_{3}} breaks up into four pieces each defined over 𝔽33\mathbb{F}_{3^{3}}. The (/65)/[1](\mathbb{Z}/65\mathbb{Z})^{\prime}/[-1] part corresponds to the ϕ(65)/2=24\phi(65)/2=24 rational cusps on X1(65)X_{1}(65), and these cusps have the same image in X0(65)X_{0}(65). A cubic point (considered as a degree 3 divisor) thus reduces either to a sum of three 𝔽3{\mathbb{F}}_{3}-rational cusps, which have the same image in X0(65)X_{0}(65), or a single cubic cusp (again, considered as a degree 3 divisor), whose image in X0(65)(3)X_{0}(65)^{(3)} is 3[c𝔽3]3[c_{{\mathbb{F}}_{3}}] for a rational cusp cc. ∎

We now analyze the decomposition of J0(65)J_{0}(65). Using Magma’s intrinsic Decomposition(JZero(65)), we find that

JE×A1×A2,J\sim_{\mathbb{Q}}E\times A_{1}\times A_{2},

where E()E(\mathbb{Q}) has rank 1 and AiA_{i} are modular abelian surfaces with analytic rank 0. Let AA be the winding quotient. Let C1:=3[c]C_{1}:=3[c_{\infty}], C2:=3[c0]C_{2}:=3[c_{0}], C3:=3[c1/5]C_{3}:=3[c_{1/5}], and C4:=3[c1/13]C_{4}:=3[c_{1/13}]. For i=1,,4i=1,\dots,4, define the morphisms μi:X0(65)(3)J0(65)\mu_{i}\colon X_{0}(65)^{(3)}\to J_{0}(65) given by z[zCi]z\mapsto[z-C_{i}] and t:J0(65)At\colon J_{0}(65)\to A, where the latter is projection. Since ϕ(3)(x𝔽3)\phi^{(3)}(x_{\mathbb{F}_{3}}) is equal to Ci,𝔽3C_{i,\mathbb{F}_{3}} for some ii, the point (tμi)(ϕ(3)(x))A()(t\circ\mu_{i})(\phi^{(3)}(x))\in A(\mathbb{Q}) belongs to the kernel of reduction A()A(𝔽3)A(\mathbb{Q})\to A(\mathbb{F}_{3}) for the appropriate ii. However, since A()A(\mathbb{Q}) is torsion, the kernel of reduction is trivial [katz:galois-properties-of-torsion, Appendix], so (tμi)(ϕ(3)(x))=0(t\circ\mu_{i})(\phi^{(3)}(x))=0 for the appropriate ii.

To conclude, we verify that the morphism

τμi:X0(65)(3)A\tau\circ\mu_{i}\colon X_{0}(65)^{(3)}\to A

a formal immersion at Ci,𝔽3C_{i,\mathbb{F}_{3}} using Proposition 5.8. Via Magma’s intrinsic Newform, we can compute a basis {ω1,,ω5}\{\omega_{1},\dots,\omega_{5}\} for the 1-forms on J0(65)J_{0}(65). The initial basis has qq-expansions with non-integer coefficients; to find a basis with integer qq-expansions, we simultaneously diagonalize these 11-forms with respect to the action of the Hecke operators. The Hecke action also identifies the subspace pulled back from AA. To verify that the formal immersion criterion holds at Ci,𝔽3C_{i,\mathbb{F}_{3}}, we check that certain 4×34\times 3 matrices (see 5.8.1) have rank 33; we can compute the expansions of the ω\omega at the other cusps via the Atkin–Lehner involutions. Finally, Lemma 5.7 asserts that ϕ(3)(x)\phi^{(3)}(x) is equal to either 3[c0], 3[c], 3[c1/5],3[c_{0}],\,3[c_{\infty}],\,3[c_{1/5}], or 3[c1/13]3[c_{1/13}]; we conclude that xx is cuspidal.

See the Magma file master-65.m for code verifying these claims.

Remark 7.5.

Wang addresses the cases of N1=22,25,40,49N_{1}=22,25,40,49 in [wang:cyclictorsion2, Theorem 1.2], N1=55,65N_{1}=55,65 in [wang:cyclictorsion, Theorem 1.2], and N1=39N_{1}=39 in a recent preprint [wang:cyclictorsion3, Theorem 0.3]. However, his proofs of these cases are incorrect. For instance, [wang:cyclictorsion2, Lemma 3.5] claims that for N>4N>4 and a prime pNp\nmid N, when the gonality of X1(N)>dX_{1}(N)>d and J1(N)()J_{1}(N)(\mathbb{Q}) is finite, the moduli of each non-cuspidal, degree dd point of X1(N)X_{1}(N) has good reduction at every prime 𝔭\mathfrak{p} over pp. (The proofs for N1=22,25,40,49N_{1}=22,25,40,49 and 3939 rely on this claim.)

We provide a counter-example to this. By [derickx2014gonality, Table 1], the gonality of X1(31)X_{1}(31) is 1212. Over the number field [a]/(a114a10+9a915a8+21a721a6+17a58a4+3a23a+1)\mathbb{Q}[a]/(a^{11}-4a^{10}+9a^{9}-15a^{8}+21a^{7}-21a^{6}+17a^{5}-8a^{4}+3a^{2}-3a+1), the elliptic curve

y2+(a102a9+3a83a7+5a6+a5+a4+3a32a2+a1)xy\displaystyle y^{2}+(a^{10}-2a^{9}+3a^{8}-3a^{7}+5a^{6}+a^{5}+a^{4}+3a^{3}-2a^{2}+a-1)xy
+(a10+3a9+4a8+6a7+6a6+a5+6a46a3+2a23a1)(yx2)x3=0\displaystyle\mbox{}\hskip 10.55pt+(a^{10}+3a^{9}+4a^{8}+6a^{7}+6a^{6}+a^{5}+6a^{4}-6a^{3}+2a^{2}-3a-1)(y-x^{2})-x^{3}=0

has a point of order 31, namely (0,0)(0,0). However the norm of the jj-invariant of this elliptic curve has 311 in its denominator, and hence, this curve has multiplicative reduction for at least one prime above 311; it cannot have additive reduction because it has a point of order 31. Indeed the prime 311311 splits into three primes in this degree 11 number field, one of degree 9 and two of degree 1, and this elliptic curve has good reduction at the prime of degree 9 and one prime of degree 1, but multiplicative reduction at the other prime of degree 1.

Similarly, for N1=55,65N_{1}=55,65, while [wang:cyclictorsion, Lemma 3.6] is correct, its application to [wang:cyclictorsion, Theorem 3.7] contains an error. Wang applies loc. cit. Lemma 3.6 to conclude that the points (ωn(x1),,ωn(xd))(\omega_{n}(x_{1}),\dots,\omega_{n}(x_{d})) and (,,)(\infty,\dots,\infty) on the dd-th symmetric power X0(N1)(d)X_{0}(N_{1})^{(d)} of the modular curve X0(N1)X_{0}(N_{1}) reduce to the same point modulo pp, where ωn\omega_{n} is some Atkin–Lehner involution on X0(N1)X_{0}(N_{1}). The correct condition that one needs to check is that gcd(N1,32i1)=1\gcd(N_{1},3^{2i}-1)=1 for i=1,2i=1,2, which succeeds for N1=143,91,77N_{1}=143,91,77, however it fails for N1=55,65N_{1}=55,65. (In Lemma 7.4, we circumvent this issue for N1=65N_{1}=65 by verifying the formal immersion criteria at more general cuspidal divisors.)

Remark 7.6.

For each ii, the morphism (tμi)(t\circ\mu_{i}) in Section 7.3 is not a formal immersion at every cuspidal divisor on X0(65)(3)(𝔽3)X_{0}(65)^{(3)}(\mathbb{F}_{3}). The failure of the formal immersion criterion is due to the fact that the Atkin–Lehner quotient X0+(65)X_{0}^{+}(65) is a rank 1 elliptic curve and the quotient map X0(65)X0+(65)X_{0}(65)\to X_{0}^{+}(65) has degree 2. Since X0(65)X_{0}(65) has rational points (the 4 rational cusps), there are 4 “copies” of X0(65)(2)()X_{0}(65)^{(2)}(\mathbb{Q}) lying on X0(65)(3)()X_{0}(65)^{(3)}(\mathbb{Q}). In particular, X0(65)(3)()X_{0}(65)^{(3)}(\mathbb{Q}) is infinite, but only finitely many rational points of X0(65)(3)X_{0}(65)^{(3)} do not lie on one of the copies of X0(65)(2)()X_{0}(65)^{(2)}(\mathbb{Q}). Moreover, the reduction modulo 3 of these copies of X0(65)(2)()X_{0}(65)^{(2)}(\mathbb{Q}) correspond to some of the points on X0(65)(3)(𝔽3)X_{0}(65)^{(3)}(\mathbb{F}_{3}) where the formal immersion criterion fails. In fact, even more is true: one can compute (in Magma, via cotangent spaces) that the map X0(65)(3)Je(65)X_{0}(65)^{(3)}\to J_{e}(65) to the winding quotient (which has dimension 4) is not an immersion at all points of X0(65)(3)X_{0}(65)^{(3)}.

Acknowledgements

We thank Lea Beneish, Nils Bruin, John Duncan, Bjorn Poonen, Jeremy Rouse, Andrew Sutherland, and Bianca Viray for helpful discussions. We also thank Lea Beneish, Abbey Bourdon, Bas Edixhoven, Álvaro Lozano-Robledo, and Filip Najman for useful comments on an earlier draft, and we are thankful to Koji Matsuda, Maleeha Khawaja and Samir Siksek to pointing out an error in Theorem 3.1. The third author was supported by NSF grant 1618657. The last author was partially supported by NSF grant DMS-1555048. Finally, we thank the anonymous referee for their comments.

References