return

Celebratio Mathematica

Andrew Pollard Ogg

Ogg’s conjectures over function fields

by Cécile Armana, Sheng-Yang Kevin Ho and Mihran Papikian

1. Introduction

In the early 1970s, An­drew Ogg made sev­er­al con­jec­tures about the ra­tion­al tor­sion points of el­lipt­ic curves over \( \mathbb{Q} \) and the Jac­obi­ans of mod­u­lar curves; see [1], [2], [3]. In par­tic­u­lar, he pro­posed a fi­nite list of groups that can oc­cur as ra­tion­al tor­sion sub­groups of el­lipt­ic curves over \( \mathbb{Q} \). These con­jec­tures were proved shortly after by Barry Mazur [e13], [e14] as a con­sequence of his fun­da­ment­al study of the arith­met­ic prop­er­ties of mod­u­lar curves and Hecke al­geb­ras. The power­ful tech­niques in­tro­duced by Mazur had tre­mend­ous im­pact on later de­vel­op­ments in arith­met­ic geo­metry; for ex­ample, they have been in­stru­ment­al in the proof of the main con­jec­ture of Iwas­awa the­ory [e22] and the proof of Fer­mat’s Last The­or­em [e37], [e38], [e30].

At around the same time, Vladi­mir Drin­feld [e10] in­tro­duced cer­tain func­tion field ana­logues of el­lipt­ic curves, which he called el­lipt­ic mod­ules, nowadays known as Drin­feld mod­ules. Drin­feld’s mo­tiv­a­tion was to con­struct func­tion field ana­logues of clas­sic­al mod­u­lar curves clas­si­fy­ing el­lipt­ic curves with some ad­di­tion­al data, and to use these to re­late auto­morph­ic forms and Galois rep­res­ent­a­tions, in line with the pro­gram en­vi­sioned by Lang­lands. The the­ory of Drin­feld mod­ules and their gen­er­al­iz­a­tions, called shtu­kas, has since lead to a suc­cess­ful res­ol­u­tion of the Lang­lands cor­res­pond­ence over func­tion fields (see [e15], [e35], [e57], [e93]), and it con­tin­ues to play a cent­ral role in num­ber the­ory be­cause of its ap­plic­a­tions to many oth­er im­port­ant prob­lems, such as the Birch and Swin­ner­ton-Dyer con­jec­ture; see [e91].

Due to the close sim­il­ar­ity between el­lipt­ic curves and Drin­feld mod­ules, Ogg con­jec­tures have nat­ur­al ana­logues in the func­tion field set­ting, as was already sug­ges­ted by Mazur in [e13]. As in the clas­sic­al case, a com­pre­hens­ive ap­proach to these con­jec­tures re­lies heav­ily on the the­ory of (Drin­feld) mod­u­lar forms and mod­u­lar vari­et­ies. Since some of the ne­ces­sary as­pects of this the­ory were de­veloped only in the 1990s, the ana­logues of Ogg’s con­jec­tures were stated and proved in cer­tain cases only in the 2000s; see [e58], [e66], [e78].

In this pa­per, we re­view the func­tion field ana­logues of Ogg’s con­jec­tures, their cur­rent status, and the meth­ods that have been ap­plied to prove some of these con­jec­tures. The meth­ods are based on the ideas of Mazur and Ogg, but there are in­ter­est­ing dif­fer­ences and tech­nic­al com­plic­a­tions that arise in the func­tion field set­ting, as well as in­triguing pos­sible new dir­ec­tions for gen­er­al­iz­a­tions.

The con­tents of the pa­per are as fol­lows. In Sec­tion 2, we re­call the state­ments of Ogg’s ori­gin­al con­jec­tures, some of their gen­er­al­iz­a­tions, and what is cur­rently known about these con­jec­tures (Ogg’s con­jec­tures have been gen­er­al­ized in many dif­fer­ent dir­ec­tions, and our ex­pos­i­tion of these gen­er­al­iz­a­tions is by no means ex­haust­ive). In Sec­tion 3, we re­view the ba­sic the­ory of Drin­feld mod­ules, put­ting the em­phas­is on those as­pects of the the­ory that are rel­ev­ant for Ogg’s con­jec­tures. In Sec­tion 4, we re­view the the­ory of Drin­feld mod­u­lar forms, which is ex­tens­ively used in the proofs of the ana­logues of Ogg’s con­jec­tures. In Sec­tions 5 and 6, we give the state­ments of Ogg’s con­jec­tures in the set­ting of Drin­feld’s the­ory, the cur­rent status of these con­jec­tures, a brief sum­mary of the ideas that go in­to the proofs, and some of the non­trivi­al com­plic­a­tions that arise in this set­ting.

2. Ogg’s conjectures

2.1. Torsion of elliptic curves
Let \( N \) be a pos­it­ive in­teger and let \( \Gamma_0(N) \) (resp. \( \Gamma_1(N) \)) be the con­gru­ence sub­group of \( \mathrm{SL}_2(\mathbb{Z}) \) con­sist­ing of matrices that are up­per-tri­an­gu­lar (resp. uni­po­tent) mod­ulo \( N \). Let \( Y_i(N)(\mathbb{C}):=\Gamma_i(N)\setminus \mathcal{H} \), \( i=0,1 \), be the af­fine mod­u­lar curve, where \( \mathcal{H} \) is the com­plex up­per half-plane and \( \mathrm{SL}_2(\mathbb{Z}) \) acts on \( \mathcal{H} \) by lin­ear frac­tion­al trans­form­a­tions. De­note by \( X_i(N)(\mathbb{C}) \) the com­pac­ti­fic­a­tion of \( Y_i(N)(\mathbb{C}) \) ob­tained by ad­join­ing fi­nitely many points, called cusps. The curve \( X_i(N)(\mathbb{C}) \) has a ca­non­ic­al nonsin­gu­lar pro­ject­ive mod­el \( X_i(N) \) defined over \( \mathbb{Q} \), and in this mod­el the set of cusps is in­vari­ant un­der the ac­tion of \( \operatorname{Gal}(\overline{\mathbb{Q}}/\mathbb{Q}) \).

The first, and prob­ably most fam­ous, con­jec­ture in Ogg’s pa­per [3] gives a clas­si­fic­a­tion of pos­sible tor­sion sub­groups \( E(\mathbb{Q})_\mathrm{tor} \) of el­lipt­ic curves over \( \mathbb{Q} \). This con­jec­ture was es­sen­tially for­mu­lated by Beppo Levi in 1908, al­though Ogg was not aware of this as Levi’s work on the arith­met­ic of el­lipt­ic curves did not re­ceive the at­ten­tion it de­served. In any case, Ogg’s pa­pers were in­stru­ment­al in spelling out the close con­nec­tion between this prob­lem and the the­ory of mod­u­lar curves and in pop­ular­iz­ing the con­jec­ture.

Con­jec­ture TEC: If \( E \) is an el­lipt­ic curve over \( \mathbb{Q} \) then its ra­tion­al tor­sion sub­group \( E(\mathbb{Q})_\mathrm{tor} \) is one of the fol­low­ing fif­teen groups: \[ \mathbb{Z}/N\mathbb{Z}, \quad 1\leq N\leq 10 \text{ or } N=12; \] \[ \mathbb{Z}/2\mathbb{Z}\times \mathbb{Z}/2N\mathbb{Z}, \quad 1\leq N\leq 4. \] This is es­sen­tially equi­val­ent to say­ing that the mod­u­lar curve \( X_1(N) \) has no \( \mathbb{Q} \)-ra­tion­al points be­sides cusps, un­less \( 1\leq N\leq 10 \) and \( N=12 \).

A re­lated con­jec­ture, stated in [3] as a prob­lem, con­cerns ra­tion­al cyc­lic sub­groups of el­lipt­ic curves over \( \mathbb{Q} \). If \( C \) is a fi­nite sub­group of an el­lipt­ic curve \( E \) over \( \mathbb{Q} \), then \( C \) is said to be ra­tion­al if \( \sigma(C)=C \) for all \( \sigma\in \operatorname{Gal}(\overline{\mathbb{Q}}/\mathbb{Q}) \).

Con­jec­ture TEC\( ^+ \): A ra­tion­al cyc­lic sub­group of an el­lipt­ic curve over \( \mathbb{Q} \) has or­der \( 1\leq N\leq 19 \), or \( N=21, 25, 27, 37, 43, 67, 163 \). Equi­val­ently, the mod­u­lar curve \( X_0(N) \) has no \( \mathbb{Q} \)-ra­tion­al points be­sides cusps, un­less \( N \) is one of the lis­ted val­ues.

The philo­sophy be­hind Con­jec­tures TEC and TEC\( ^+ \) is that “mod­u­lar curves only have ra­tion­al points for which there is a reas­on”, the reas­on be­ing geo­met­ric ([3], p. 17). More spe­cific­ally, \( Y_1(N) \) has a \( \mathbb{Q} \)-ra­tion­al point if and only if \( X_1(N) \) has genus 0. The genus of \( X_1(N) \) is 0 ex­actly for \( 1\leq N\leq 10 \) and \( N=12 \). Moreover, for these val­ues \( X_1(N) \) has in­fin­itely many \( \mathbb{Q} \)-ra­tion­al points, so the groups lis­ted in Con­jec­tures TEC oc­cur as the ra­tion­al tor­sion sub­groups of in­fin­itely many non­i­so­morph­ic el­lipt­ic curves over \( \mathbb{Q} \). Sim­il­arly, \( X_0(N)(\mathbb{Q}) \) is in­fin­ite if and only if its genus is 0, which hap­pens ex­actly for \( 1\leq N\leq 10 \) and \( N=12,13,16,18,25 \). The curve \( X_0(N) \) has genus 1, and \( Y_0(N)(\mathbb{Q}) \) is fi­nite and known for \( N=11,14,15,17,19,21,27 \). The cases when \( X_0(N) \) has genus \( \geq 2 \) and \( Y_0(N)(\mathbb{Q}) \) is nonempty are \( N=37, 43, 67, 163 \). For \( N=43, 67, 163 \), the curve \( Y_0(N) \) has one \( \mathbb{Q} \)-ra­tion­al point which cor­res­ponds to an el­lipt­ic curve over \( \mathbb{Q} \) with CM by the ring of in­tegers of \( \mathbb{Q}(\sqrt{-N}) \) (note that in these cases the class num­ber of \( \mathbb{Q}(\sqrt{-N}) \) is one). The case \( N=37 \) is some­what spe­cial: \( Y_0(37) \) has two \( \mathbb{Q} \)-ra­tion­al points, whose ex­ist­ence is re­lated to the fact that the hy­per­el­lipt­ic in­vol­u­tion of \( X_0(37) \) is not the Atkin–Lehner in­vol­u­tion.

Con­jec­tures TEC and TEC\( ^+ \) were proved by Mazur in [e13], [e14], where he de­veloped an in­tric­ate arith­met­ic the­ory of mod­u­lar curves, Ei­s­en­stein ideals in Hecke al­geb­ras, and iso­geny char­ac­ters of el­lipt­ic curves.

By ex­tend­ing Mazur’s tech­niques, one can at­tack the prob­lem of clas­si­fy­ing the points on \( X_1(N) \) and \( X_0(N) \) that are ra­tion­al over num­ber fields of giv­en de­gree; see [e34], [e108]. One can also con­sider ra­tion­al points on Shimura curves clas­si­fy­ing abeli­an sur­faces with qua­ternion­ic mul­ti­plic­a­tion; see [e25], [e82].

The Uni­form Bounded­ness Con­jec­ture (UBC) ex­tends Con­jec­ture TEC to all num­ber fields, but without giv­ing an ex­pli­cit clas­si­fic­a­tion. This con­jec­ture is men­tioned in [3] as a “folk­lore con­jec­ture”.

Con­jec­ture UBC: The or­ders of \( K \)-ra­tion­al tor­sion sub­groups of el­lipt­ic curves over a num­ber field \( K \) are uni­formly bounded by a con­stant de­pend­ing only on the de­gree \( [K:\mathbb{Q}] \).

The UBC was proved by Mer­el [e45], again build­ing upon Mazur’s work and sub­sequent re­fine­ments by Kami­enny, and com­bin­ing them with the pro­gress, re­cent at the time, on the Birch and Swin­ner­ton-Dyer con­jec­ture.

Fi­nally, one can ex­tend Con­jec­tures TEC and TEC\( ^+ \) to mod­u­lar curves clas­si­fy­ing el­lipt­ic curves with oth­er level struc­tures, such as split or non­split Cartan sub­groups. These con­jec­tures are closely re­lated to Serre’s Uni­form­ity Ques­tion (SUQ):

Con­jec­ture SUQ: The Galois rep­res­ent­a­tion \( \rho_{E, p}\colon \operatorname{Gal}(\overline{\mathbb{Q}}/\mathbb{Q})\to \operatorname{Aut}(E[p])\cong \mathrm{GL}_2(\mathbb{F}_p) \), arising from the ac­tion of the Galois group on the \( p \)-tor­sion sub­group of an el­lipt­ic curve over \( \mathbb{Q} \) without com­plex mul­ti­plic­a­tion, is sur­ject­ive once \( p > 37 \).

Al­though the SUQ re­mains open, sub­stan­tial pro­gress has been made: it is known that for \( p > 37 \) the rep­res­ent­a­tion \( \rho_{E, p} \) is either sur­ject­ive or its im­age is con­tained in the nor­mal­izer of a non­split Cartan sub­group; see [e73], [e79], [e95].

2.2. Torsion of the Jacobian variety of \( X_0(N) \)
Let \( J_0(N) \) be the Jac­obi­an vari­ety of the mod­u­lar curve \( X_0(N) \). Let \( \mathcal{C}_N \) be the sub­group of \( J_0(N) \) gen­er­ated by the di­visor classes \( [c]-[c^{\prime}] \), where \( [c] \) and \( [c^{\prime}] \) run over the cusps of \( X_0(N) \). By a res­ult of Man­in and Drin­feld [e7], \( \mathcal{C}_N \) is a fi­nite group. De­note by \( \mathcal{C}_N(\mathbb{Q}) \) the sub­group of \( \mathcal{C}_N \) fixed by the Galois ac­tion, so \( \mathcal{C}_N(\mathbb{Q})\subseteq J_0(N)(\mathbb{Q})_\mathrm{tor} \). Let \( \mathcal{C}(N) \) be the ra­tion­al cuspid­al di­visor class group of \( X_0(N) \); this is the sub­group of \( \mathcal{C}_N(\mathbb{Q}) \) gen­er­ated by the lin­ear equi­val­ence classes of de­gree 0 ra­tion­al cuspid­al di­visors on \( X_0(N) \).

When \( N=p \) is prime, there are only two cusps on \( X_0(p) \), usu­ally de­noted by \( [0] \) and \( [\infty] \). In [2], Ogg com­puted that the di­visor class of \( [0]-[\infty] \) in \( J_0(p) \) has or­der \[ n=(p-1)/\gcd(p-1, 12). \] Both cusps are ra­tion­al, so \( \mathcal{C}(p)=\mathcal{C}_p(\mathbb{Q})=\mathcal{C}_p \). Based on nu­mer­ic­al ex­per­i­ment­a­tions [3], Ogg con­jec­tured the fol­low­ing.

Con­jec­ture CJ-p: The cyc­lic group \( \mathcal{C}(p) \) is the full tor­sion sub­group of \( J_0(p)(\mathbb{Q}) \).

Mazur proved this con­jec­ture in [e13] us­ing his the­ory of the Ei­s­en­stein ideal of the Hecke al­gebra of level \( p \). There is a nat­ur­al gen­er­al­iz­a­tion of this con­jec­ture to ar­bit­rary \( N \), nowadays called the gen­er­al­ized Ogg con­jec­ture.

Con­jec­ture CJ-N: For ar­bit­rary \( N\geq 1 \), we have \( \mathcal{C}(N)=\mathcal{C}_N(\mathbb{Q})=J_0(N)(\mathbb{Q})_\mathrm{tor} \).

Note that Con­jec­ture CJ-N is a com­bin­a­tion of two con­jec­tures. The first is the equal­ity \( \mathcal{C}(N)=\mathcal{C}_N(\mathbb{Q}) \), which con­cerns only the cuspid­al di­visor group; this was ex­pli­citly con­jec­tured by Hwa­jong Yoo in [e107] in re­sponse to a ques­tion of Ribet. The second is the equal­ity \( \mathcal{C}_N(\mathbb{Q})=J_0(N)(\mathbb{Q})_\mathrm{tor} \), which pre­dicts that all \( \mathbb{Q} \)-ra­tion­al tor­sion points on \( J_0(N) \) arise from cuspid­al di­visors, so there are no “un­ex­pec­ted” \( \mathbb{Q} \)-ra­tion­al tor­sion points on \( J_0(N) \).

De­term­in­ing the struc­ture of the cuspid­al di­visor group for gen­er­al \( N \) is quite com­plic­ated. Re­cently, Yoo [e107] was able to de­term­ine the struc­ture of \( \mathcal{C}(N) \) com­pletely by ex­tend­ing earli­er work of Loren­zini, Ling, Tak­agi, and oth­ers. Yoo’s proof com­bines a care­ful study of mod­u­lar units with ex­pli­cit con­struc­tion of ap­pro­pri­ate ra­tion­al cuspid­al di­visors. Con­sequently, if Con­jec­ture CJ-N is true, then the struc­ture of \( J_0(N)(\mathbb{Q})_\mathrm{tor} \) is de­term­ined as well.

The con­jec­ture \( \mathcal{C}(N)=\mathcal{C}_N(\mathbb{Q}) \) is known in some cases, but not in gen­er­al. Note that for square-free \( N \), we trivi­ally have \( \mathcal{C}(N)=\mathcal{C}_N(\mathbb{Q})=\mathcal{C}_N \) be­cause all the cusps are ra­tion­al. The known non­trivi­al cases are es­sen­tially those \( N \) whose prime de­com­pos­i­tion con­tains at most two prime factors with high­er powers; the in­ter­ested read­er may con­sult the in­tro­duc­tion of [e107] for a list of known cases.

Giv­en a fi­nite abeli­an group \( G \) and a prime \( \ell \), we de­note by \( G_\ell \) the \( \ell \)-primary sub­group of \( G \). Note that the equal­ity \begin{equation}\label{eqC=J} \mathcal{C}(N)_\ell = (J_0(N)(\mathbb{Q})_\mathrm{tor})_\ell \tag{2.1} \end{equation} im­plies \( \mathcal{C}(N)_\ell=\mathcal{C}_N(\mathbb{Q})_\ell \). The strongest res­ult to­wards the equal­ity \eqref{eqC=J} is the fol­low­ing re­cent res­ult of Yoo [e110].

The­or­em 2.1: Let \( N \) be a pos­it­ive in­teger. Let \( \ell \) be an odd prime whose square does not di­vide \( N \). If \( \ell \geq 5 \), then \eqref{eqC=J} holds. If \( \ell=3 \), then \eqref{eqC=J} holds un­der the ad­di­tion­al as­sump­tion that either \( 3\nmid N \), or \( N \) has a prime di­visor con­gru­ent to \( -1 \) mod­ulo 3.

Yoo’s proof is based on the the­ory of Ei­s­en­stein ideals, with im­port­ant re­fine­ments in­tro­duced by Ohta in [e83], where The­or­em 2.1 is proved for square-free \( N \).

Re­mark 2.2: (1)  When \( N = p^r \) is a prime power, \eqref{eqC=J} holds for all odd \( \ell\neq p \), so the technical conditions on \( \ell=3 \) can be dropped; see ([e110], Equation (1.1)). Moreover, in this case there is a different geometric approach to this problem due to Lorenzini. Let \( \mathcal{J}_0(N) \) denote the Néron model of \( J_0(N) \) over \( \operatorname{Spec}(\mathbb{Z}) \). Let \( \mathcal{J}_0(N)_{\mathbb{F}_p} \) denote the fiber of \( \mathcal{J}_0(N) \) at \( p \), and let \( \mathcal{J}_0(N)_{\mathbb{F}_p}^0 \) be its connected component of the identity. Let \( \Phi_{N,p}=\mathcal{J}_0(N)_{\mathbb{F}_p}/\mathcal{J}_0(N)_{\mathbb{F}_p}^0 \) be the group of connected components of \( \mathcal{J}_0(N) \) at \( p \). There is a canonical reduction map \( J_0(N)(\mathbb{Q})_\mathrm{tor}\to \mathcal{J}_0(N)_{\mathbb{F}_p} \), which is in­ject­ive; see the Appendix in [e19]. Composing this reduction map with \( \mathcal{J}_0(N)_{\mathbb{F}_p} \to \Phi_{N,p} \), one obtains a canonical homomorphism \( \pi_p\colon J_0(N)(\mathbb{Q})_\mathrm{tor}\to \Phi_{N,p} \). The idea of Lorenzini’s approach [e40] is that, for certain \( \ell \), the map \( \pi_p\colon (\mathcal{C}(N))_\ell\to (\Phi_{N,p})_\ell \) is surjective, whereas \( \pi_p\colon (J_0(N)(\mathbb{Q})_\mathrm{tor})_\ell\to (\Phi_{N,p})_\ell \) is injective. This is based on complicated computations with \( (\Phi_{N,p})_\ell \) and some inductive arguments with \( r \). We note that for prime \( N=p \), Mazur already proved in [e13] that \( \pi_p \) is an isomorphism.

(2)  The proofs of spe­cial cases of Con­jec­ture CJ-N by Loren­zini, Mazur, Ohta, and Yoo im­pli­citly rely on the know­ledge of the or­der of \( \mathcal{C}(N) \). In [e105], Ribet and Wake prove the equal­ity \eqref{eqC=J} for square-free \( N \) and \( \ell\nmid 6N \) by show­ing that both sides are iso­morph­ic mod­ules for the Hecke al­gebra act­ing on \( J_0(N) \), without ac­tu­ally com­put­ing either side ex­pli­citly.

(3)  The meth­ods used so far to prove \eqref{eqC=J} in cer­tain cases fall short of prov­ing this equal­ity for \( \ell=2 \). Thus, \eqref{eqC=J} re­mains largely open for the 2-primary part, ex­cept when \( N=p \) is prime. We note that the most tech­nic­al ar­gu­ments in Mazur’s pa­per [e13] prov­ing Con­jec­ture CJ-p are ac­tu­ally con­cerned with the 2-primary tor­sion.

(4)  In [e81], Ohta proves that \( (J_1(p)(\mathbb{Q})_\mathrm{tor})_\ell \) is gen­er­ated by cuspid­al di­visors for any odd prime \( \ell \), as was con­jec­tured by Con­rad, Edix­hoven and Stein [e60]. The proof again re­lies on the Ei­s­en­stein ideal ma­chinery.

Con­jec­ture CJ-p has a “dual” ver­sion con­cern­ing the largest \( \mu \)-type sub­group \( \mathcal{M}(N) \) of \( J_0(N) \); a com­mut­at­ive group scheme is \( \mu \)-type if it is fi­nite, flat, and its Carti­er dual is a con­stant group scheme. The nat­ur­al morph­ism \( X_1(N)\to X_0(N) \) in­duces, by Pi­card func­tori­al­ity, a morph­ism of their Jac­obi­ans \( J_0(N)\to J_1(N) \). The ker­nel of this last morph­ism is called the Shimura sub­group of \( J_0(N) \), and will be de­noted by \( \mathcal{S}(N) \). When \( N=p \) is prime, \( \mathcal{S}(p) \) is cyc­lic of the same or­der as \( \mathcal{C}(p) \). In [3], Ogg con­jec­tured:

Con­jec­ture SJ-p: For any prime \( p \), we have \( \mathcal{S}(p) = \mathcal{M}(p) \).

This con­jec­ture was proved by Mazur in [e13]. Des­pite the dual ap­pear­ance of Con­jec­tures CJ-p and SJ-p, the proof of Con­jec­ture SJ-p lies deep­er than the proof of Con­jec­ture CJ-p since its proof re­lies on the fact that the com­ple­tion of the Hecke al­gebra at any prime ideal in the sup­port of Ei­s­en­stein ideal is Goren­stein. For gen­er­al \( N \), the or­der of \( \mathcal{S}(N) \), the ac­tion of the Hecke op­er­at­ors and the ac­tion of \( \operatorname{Gal}(\overline{\mathbb{Q}}/\mathbb{Q}) \) on \( \mathcal{S}(N) \) are com­puted in [e33]. In par­tic­u­lar, \( \mathcal{S}(N)\subseteq \mathcal{M}(N) \). The fol­low­ing gen­er­al­iz­a­tion of Mazur’s res­ult was proved by Vat­sal [e65].

The­or­em 2.3: Let \( \ell \) be an odd prime whose square does not di­vide \( N \). Then \( \mathcal{S}(N)_\ell=\mathcal{M}(N)_\ell \).

Vat­sal’s proof is very dif­fer­ent from Mazur’s. It com­bines a the­or­em of Ihara about un­rami­fied cov­er­ings of mod­u­lar curves over fi­nite fields with some deep res­ults about the su­per­sin­gu­lar re­duc­tions of CM points on \( X_0(N) \) and spe­cial val­ues of \( L \)-func­tions.

Re­mark 2.4: By definition, \( \mathcal{S}(N) \) is the kernel of a natural homomorphism \( J_0(N)\to J_1(N) \). One might wonder whether the quotients of \( J_0(N) \) by subgroups of \( \mathcal{C}(N) \) also give arithmetically interesting abelian varieties. Incidentally, there is another conjecture of Ogg in that direction. Assume \( N \) is square-free and divisible by an even number of primes. Let \( \mathcal{O}_N \) be a maximal order in the indefinite quaternion algebra over \( \mathbb{Q} \) with discriminant \( N \), and let \( \mathcal{O}_N^1 \) be the subgroup of \( \mathcal{O}_N \) consisting of elements of reduced norm 1. Then \( \mathcal{O}_N^1 \) is isomorphic to a subgroup of \( \mathrm{SL}_2(\mathbb{R}) \), so it acts on the upper half-plane \( \mathcal{H} \). The quotient \( X^N:= \mathcal{O}_N^1\setminus \mathcal{H} \) is a Shimura curve and has a canonical model over \( \mathbb{Q} \). In [e18], Ribet proved the existence of an isogeny defined over \( \mathbb{Q} \) between the “new” part \( J_0(N)^\mathrm{new} \) of \( J_0(N) \) and the Jacobian \( J^N \) of \( X^N \). Unfortunately, Ribet’s proof provides no information about this isogeny beyond its existence. In [4], by studying the component groups of \( J_0(N) \) and \( J^N \), Ogg made an explicit conjecture about the kernel of Ribet’s isogeny when \( N=pq \) is a product of two distinct primes and \( p = 2,3,5,7,13 \). The conjecture predicts that there is an isogeny \( J_0(N)^\mathrm{new}\to J^N \) of minimal degree whose kernel is a specific subgroup of \( \mathcal{C}_N \). Although this conjecture is false in general [e98], it seems to be correct when \( J_0(N)^\mathrm{new}=J_0(N) \); see [e94].

3. Review of Drinfeld modules

Drin­feld mod­ules can be defined for sub­rings of func­tion fields of gen­er­al curves over a fi­nite field \( \mathbb{F}_q \) with \( q \) ele­ments. On the oth­er hand, the ana­logue of \( \mathbb{Q} \) in this set­ting is the ra­tion­al func­tion field \( F=\mathbb{F}_q(T) \), where \( T \) is an in­de­term­in­ate. Since Ogg’s con­jec­tures are ex­pli­cit clas­si­fic­a­tions of ra­tion­al tor­sion points of el­lipt­ic curves over \( \mathbb{Q} \) or Jac­obi­ans of mod­u­lar curves, their ana­logues over func­tion fields are also stated mostly over \( F \). Thus, in this sec­tion, we dis­cuss Drin­feld mod­ules only for the poly­no­mi­al ring \( A=\mathbb{F}_q[T] \). This as­sump­tion also sim­pli­fies some of the non­es­sen­tial tech­nic­al­it­ies of the the­ory. Most of the res­ults in this sec­tion are due to Drin­feld [e10]. De­tailed proofs of these state­ments can be found in [e48] or [e109].
3.1. Analytic theory
We start with the ana­lyt­ic the­ory of Drin­feld mod­ules, which par­al­lels closely the the­ory of roots of unity \( e^{2\pi i/n} \) and also ana­lyt­ic uni­form­iz­a­tion of el­lipt­ic curves as quo­tients of \( \mathbb{C} \) by lat­tices.

First, we in­tro­duce the ana­logues of real and com­plex num­bers in the func­tion field set­ting. The de­gree func­tion \( \deg=\deg_T\colon A\to \mathbb{Z}_{\geq 0}\cup \{-\infty\} \), which as­signs to \( 0\neq a\in A \) its de­gree as a poly­no­mi­al in \( T \) and \( \deg_T(0)=-\infty \), ex­tends to \( F \) by \( \deg(a/b)=\deg(a)-\deg(b) \). The map \( -\deg \) is a valu­ation on \( F \); the cor­res­pond­ing place of \( F \) is usu­ally de­noted by \( \infty \). Let \( \lvert\,\cdot\,\rvert \) de­note the cor­res­pond­ing ab­so­lute value on \( F \) nor­mal­ized by \( \lvert T \rvert=q \). The com­ple­tion \( F_\infty \) of \( F \) with re­spect to this ab­so­lute value is iso­morph­ic to the field \( \mathbb{F}_q(\!({1/T})\!) \) of Laurent series in \( 1/T \). Fi­nally, let \( \mathbb{C}_\infty \) be the com­ple­tion of an al­geb­ra­ic clos­ure of \( F_\infty \). The ab­so­lute value \( \lvert\,\cdot\,\rvert \) has a unique ex­ten­sion, also de­noted by \( \lvert\,\cdot\,\rvert \), to \( \mathbb{C}_\infty \).

An \( A \)-lat­tice \( \Lambda\subset \mathbb{C}_\infty \) of rank \( r\geq 1 \) is a dis­crete \( A \)-sub­mod­ule of \( \mathbb{C}_\infty \) of rank \( r \), where “dis­crete” means that for any \( N > 0 \) the set \( \{\lambda\in \Lambda: \lvert\lambda\rvert\leq N\} \) is fi­nite. One shows that any \( A \)-lat­tice is of the form \( \Lambda=A\omega_1+\cdots +A\omega_r \), where \( \omega_1, \dots, \omega_r\in \mathbb{C}_\infty \) are lin­early in­de­pend­ent over \( F_\infty \). Since the de­gree of \( \mathbb{C}_\infty \) over \( F_\infty \) is in­fin­ite, there are \( A \)-lat­tices of ar­bit­rar­ily large ranks (un­like \( \mathbb{Z} \)-lat­tices in \( \mathbb{C} \)).

The ex­po­nen­tial func­tion of \( \Lambda \) is \[ e_\Lambda(x)=x\prod_{0\neq \lambda\in \Lambda} \left(1-\frac{x}{\lambda}\right). \] Us­ing the dis­crete­ness of \( \Lambda \), it is not hard to show that the func­tion \( e_\Lambda(x) \) is en­tire, i.e., con­verges every­where on \( \mathbb{C}_\infty \). Be­cause of the non­archimedean set­ting, this im­plies that \( e_\Lambda(x)\colon \mathbb{C}_\infty\to \mathbb{C}_\infty \) is sur­ject­ive. The set of zer­os of \( e_\Lambda \) is ex­actly \( \Lambda \). Fi­nally, be­cause \( \Lambda \) is an \( \mathbb{F}_q \)-vec­tor space, \( e_\Lambda(x) \) sat­is­fies \( e_\Lambda(x+y)=e_\Lambda(x)+e_\Lambda(y) \) and \( e_\Lambda(\alpha x)=\alpha e_\Lambda(x) \) for all \( \alpha\in \mathbb{F}_q \). In oth­er words, \( e_\Lambda(x) \) is an \( \mathbb{F}_q \)-lin­ear func­tion. Thus, the power series ex­pan­sion of \( e_\Lambda(x) \) is of the form \[ e_\Lambda(x)=\sum_{n\geq 0} e_n(\Lambda)x^{q^n}. \] There are re­curs­ive for­mu­las for the coef­fi­cients of \( e_\Lambda \) in terms of Ei­s­en­stein series: If we put \begin{equation}\label{eqEisSer} E_n(\Lambda)=\sum_{0\neq \lambda\in \Lambda} \frac{1}{\lambda^n},\tag{3.1} \end{equation} then \begin{equation}\label{eqCoefExp} e_n(\Lambda) = E_{q^n-1}(\Lambda)+\sum_{i=1}^{n-1}e_i(\Lambda) E_{q^{n-i}-1}(\Lambda)^{q^i}. \tag{3.2} \end{equation} Let \[ \mathbb{C}_\infty\{x\}=\{a_0x+a_1x^q+\cdots+a_n x^{q^n}\mid n\geq 0, a_0, \dots, a_n\in \mathbb{C}_\infty\} \] be the non­com­mut­at­ive ring of \( \mathbb{F}_q \)-lin­ear poly­no­mi­als with usu­al ad­di­tion of poly­no­mi­als but where mul­ti­plic­a­tion is defined via the com­pos­i­tion of poly­no­mi­als. For ex­ample, \[ (Tx+x^q)\circ (x+Tx^{q^2})=T(x+Tx^{q^2})+(x+Tx^{q^2})^q= Tx+x^q+T^2x^{q^2}+T^qx^{q^3}. \] Giv­en \( f(x)=a_0x+a_1x^q+\cdots+a_n x^{q^n} \) in \( \mathbb{C}_\infty\{x\} \), we de­note \( \partial f=\frac{\mathrm{d}}{\mathrm{d} x}f(x)=a_0 \).

An im­port­ant prop­erty of \( e_\Lambda(x) \) is the func­tion­al equa­tion \[ e_\Lambda(T x) = \phi_T^\Lambda(e_\Lambda(x)), \] where \[ \phi_T^\Lambda(x)=Tx+g_1(\Lambda)x^q+\cdots+ g_r(\Lambda)x^{q^r}\in \mathbb{C}_\infty\{x\},\quad g_r(\Lambda)\neq 0. \] Be­cause \( e_\Lambda(x) \) is \( \mathbb{F}_q \)-lin­ear, this func­tion­al equa­tion ex­tends to all \( a\in A \): for each \( a\in A \), there is \( \phi_a^\Lambda(x)\in \mathbb{C}_\infty\{x\} \) such that \( \deg_x \phi_a^\Lambda(x)=\lvert a \rvert^r \), \( \partial \phi_a^\Lambda=a \), and \( e_\Lambda(a x)= \phi_a^\Lambda(e_\Lambda(x)) \). Moreover, the map \begin{align*} \phi^\Lambda\colon A &\longrightarrow \mathbb{C}_\infty\{x\}, \\ a &\longmapsto \phi_a^\Lambda(x) \end{align*} is an \( \mathbb{F}_q \)-al­gebra ho­mo­morph­ism, called the Drin­feld mod­ule of rank \( r \) as­so­ci­ated to \( \Lambda \).

Con­versely, a Drin­feld \( A \)-mod­ule of rank \( r \) over \( \mathbb{C}_\infty \) is an \( \mathbb{F}_q \)-al­gebra ho­mo­morph­ism \( \phi\colon A\to \mathbb{C}_\infty\{x\} \), \( a\mapsto \phi_a(x) \), defined by \( \phi_T(x)=Tx+g_1x^q+\cdots +g_r x^{q^r} \) with \( g_r\neq 0 \). One con­structs an en­tire \( \mathbb{F}_q \)-lin­ear func­tion \( e_\phi(x) \) sat­is­fy­ing \( e_\phi(Tx)=\phi_T(e_\phi(x)) \) as fol­lows. Put \[ e_\phi(x) = e_0x+e_1x^q+e_2x^{q^2}+\cdots, \] where \( e_0, e_1, \dots \) are to be de­term­ined. The func­tion­al equa­tion \( e_\phi(Tx)=\phi_T(e_\phi(x)) \) leads to a sys­tem of equa­tions \[ (T^{q^n}-T)e_n=e_{n-1}^qg_1+e_{n-2}^{q^2}g_2+\cdots+ e_{n-r}^{q^r}g_r, \quad n\geq 0, \] where \( e_i=0 \) for \( i < 0 \). If we put \( e_0=1 \), then every oth­er \( e_n \) is uniquely de­term­ined from the above re­curs­ive for­mu­las. It is not hard to show that the res­ult­ing func­tion \( e_\phi(x) \) is en­tire and the set of zer­os \( \Lambda_\phi \) of \( e_\phi \) is an \( A \)-lat­tice of rank \( r \) such that \( \phi=\phi^{\Lambda_\phi} \).

A morph­ism \( u\colon \phi\to \psi \) of Drin­feld mod­ules is a poly­no­mi­al \( u(x)\in \mathbb{C}_\infty\{x\} \) such that \( u(\phi_a(x))=\psi_a(u(x)) \) for all \( a\in A \). A morph­ism \( u\colon \phi\to \psi \) is an iso­morph­ism if \( u \) is in­vert­ible, i.e., \( u=c\in \mathbb{C}_\infty^\times \). Note that since \( T \) gen­er­ates the \( \mathbb{F}_q \)-al­gebra \( A \), the com­mut­a­tion \begin{equation}\label{eq-defmor} u(\phi_T(x))=\psi_T(u(x))\tag{3.3} \end{equation} is suf­fi­cient to en­sure \( u(\phi_a(x))=\psi_a(u(x)) \) for all \( a\in A \). Com­par­ing the de­grees of both sides of \eqref{eq-defmor}, we see that nonzero morph­isms can ex­ist only between Drin­feld mod­ules of the same rank. We de­note the set of all morph­isms \( \phi\to \psi \) by \( \operatorname{Hom}(\phi, \psi) \). A morph­ism of lat­tices \( \Lambda\to \Lambda^{\prime} \) is an ele­ment \( c\in \mathbb{C}_\infty \) such that \( c\Lambda\subseteq \Lambda^{\prime} \). The set of all morph­isms \( \Lambda\to \Lambda^{\prime} \) is de­noted \( \operatorname{Hom}(\Lambda, \Lambda^{\prime}) \). Both \( \operatorname{Hom}(\phi, \psi) \) and \( \operatorname{Hom}(\Lambda, \Lambda^{\prime}) \) are nat­ur­ally \( A \)-mod­ules. One shows that there is an iso­morph­ism of \( A \)-mod­ules \begin{align*} \operatorname{Hom}(\phi, \psi) &\overset{\sim}{\longrightarrow} \operatorname{Hom}(\Lambda_\phi, \Lambda_\psi)\\ u &\longmapsto \partial u. \end{align*} From these con­struc­tions we get:

The­or­em 3.1: The cat­egory of Drin­feld mod­ules of rank \( r \) over \( \mathbb{C}_\infty \) and the cat­egory of \( A \)-lat­tices of rank \( r \) in \( \mathbb{C}_\infty \) are equi­val­ent.
Ex­ample 3.2: The Carl­itz mod­ule is the Drinfeld module defined by \( \phi_T=Tx+x^q \). This is the simplest possible Drinfeld module. We will distinguish the Carlitz module among all other Drinfeld modules by denoting it by \( C \), i.e., \( C_T=Tx+x^q \). The rank of \( C \) is 1. It is easy to show that the coefficients of \( e_C(x)=\sum_{n\geq 0} e_n x^{q^n} \) are given by the formula \[ e_0=1 \quad \text{and}\quad e_n=(T^{q^n}-T)(T^{q^{n}}-T^q)\cdots (T^{q^n}-T^{q^{n-1}})\quad \text{for}\quad n\geq 1. \] The lattice \( \Lambda_C \) of \( C \) has rank 1, so \( \Lambda=\pi_C A \) for some \( \pi_C\in \mathbb{C}_\infty^\times \). The generator \( \pi_C \), which is well-defined up to an \( \mathbb{F}_q^\times \)-multiple, is called the Carl­itz peri­od; it plays the role of \( 2\pi i\in \mathbb{C} \). There are various explicit formulas for the Carlitz period, one of which is the following: \begin{equation}\label{eqCarlitzFormua} \sum_{\substack{a\in A \\ a\,\textup{ monic}}} \frac{1}{a^{q-1}}=-\frac{\pi_C^{q-1}}{T^{q}-T}.\tag{3.4} \end{equation} This is the analogue of Euler’s formula \( \sum_{n\geq 1} 1/n^2 =\pi^2/6 \). Wade [e3] proved that \( \pi_C \) is transcendental over \( F \), just like \( \pi \) is transcendental over \( \mathbb{Q} \).
Re­mark 3.3: The Carlitz module was originally introduced by Carlitz in [e1], where he defined \( e_C(x) \) and proved \eqref{eqCarlitzFormua}. Carlitz also showed in [e2] that \( C \) gives rise to the correct analogue of cyclotomic polynomials over \( F \). Unfortunately, [e1] and [e2] did not receive the attention they deserved from the larger mathematical community and were mostly forgotten until the 1970s.

Giv­en a Drin­feld mod­ule \( \phi \), we equip \( \mathbb{C}_\infty \) with a new \( A \)-mod­ule struc­ture \( a\circ z =\phi_a(z) \) de­noted \( {^\phi}\mathbb{C}_\infty \). This gives an ex­act se­quence of \( A \)-mod­ules \begin{equation}\label{eqDMUnif} 0\longrightarrow \Lambda_\phi \longrightarrow \mathbb{C}_\infty \overset{e_\phi}{\longrightarrow} {^{\phi}}\mathbb{C}_\infty\longrightarrow 0, \tag{3.5} \end{equation} which can be in­ter­preted as the ana­logue of ana­lyt­ic uni­form­iz­a­tion \( \mathbb{C}/\Lambda\overset{\sim}{\longrightarrow}E(\mathbb{C}) \), \( z\mapsto (\wp(z), \wp^{\prime}(z)) \) of an el­lipt­ic curve \( E \) over \( \mathbb{C} \) (here \( \wp \) is the Wei­er­strass \( \wp \)-func­tion as­so­ci­ated to the lat­tice \( \Lambda \)).

For nonzero \( a\in A \), the \( a \)-tor­sion points of \( \phi \) are the roots of \( \phi_a(x) \). The set of these roots, de­noted \( \phi[a] \), is nat­ur­ally an \( A \)-mod­ule via \( b\circ \alpha = \phi_b(\alpha) \), where \( b\in A \) and \( \alpha\in \phi[a] \) (to see that \( b\circ \alpha \) is in \( \phi[a] \), com­pute \( \phi_a(b\circ \alpha)=\phi_a(\phi_b(\alpha))=\phi_b(\phi_a(\alpha))=\phi_b(0)=0 \)). Ap­ply­ing the snake lemma to the ex­act se­quence \eqref{eqDMUnif}, we get an iso­morph­ism of \( A \)-mod­ules: \[ \phi[a]\cong \Lambda_\phi/a\Lambda_\phi\cong (A/aA)^r. \] A morph­ism \( u\colon \phi\to \psi \) in­duces a ho­mo­morph­ism \( {^\phi}\mathbb{C}_\infty\to {^\psi}\mathbb{C}_\infty \) of \( A \)-mod­ules. In par­tic­u­lar, \( u \) in­duces a ho­mo­morph­ism \( \phi[a]\to \psi[a] \), \( \alpha\mapsto u(\alpha) \).

It is easy to show that two Drin­feld mod­ules \( \phi \) and \( \psi \) are iso­morph­ic if and only if their cor­res­pond­ing lat­tices are ho­mothet­ic: \( \Lambda_\phi= c\Lambda_\psi \) for \( c\in\mathbb{C}_\infty^\times \). Thus, to clas­si­fy Drin­feld mod­ules of rank \( r \) up to iso­morph­ism, it is enough to clas­si­fy \( A \)-lat­tices in \( \mathbb{C}_\infty \) of rank \( r \) up to ho­mothety. This is trivi­al if \( r=1 \). For \( r\geq 2 \) the key ob­ject for our task is the Drin­feld sym­met­ric space \[ \Omega^r= \mathbb{P}^{r-1}(\mathbb{C}_\infty)-\bigcup_{F_\infty\mathrm{-rational }\, H} H, \] where the uni­on is over the \( F_\infty \)-ra­tion­al hy­per­planes; equi­val­ently, \( \Omega^r \) is the set of all points \( (z_1, \dots, z_r) \) of \( \mathbb{P}^{r-1}(\mathbb{C}_\infty) \) such that \( z_1, \dots, z_r \) are lin­early in­de­pend­ent over \( F_\infty \). To the point \( (z_1, \dots, z_r) \) we as­so­ci­ate the ho­mothety class of the lat­tice \( Az_1+\cdots+Az_r \). The ac­tion of \( \mathrm{GL}_r(F_\infty) \) on \( \mathbb{P}^{r-1}(\mathbb{C}_\infty) \) pre­serves \( \Omega^r \).

Ex­ample 3.4: Suppose \( r=2 \). In this case, \( \mathbb{P}^{r-1}(\mathbb{C}_\infty)=\mathbb{P}^1(\mathbb{C}_\infty) \) consists of \( [1, 0] \) and \( [z, 1] \), \( z\in \mathbb{C}_\infty \). It is easy to see that \begin{align*}\Omega^2 &=\{[z, 1]\mid z\not\in F_\infty\}\\ &=\mathbb{C}_\infty-F_\infty. \end{align*} After identifying \( \Omega^2 \) with \( \mathbb{C}_\infty-F_\infty \), \( \mathrm{GL}_2(F_\infty) \) acts on \( \Omega^2 \) by linear fractional transformations \[ \begin{pmatrix} a & b \\ c & d\end{pmatrix} z = \frac{az+b}{cz+d}. \]

From the bijec­tion between lat­tices and Drin­feld mod­ules, one de­duces that the set of iso­morph­ism classes of Drin­feld mod­ules of rank \( r \) over \( \mathbb{C}_\infty \) is in nat­ur­al bijec­tion with the set of or­bits \[ \mathrm{GL}_r(A)\setminus\Omega^r. \]

Each nonzero ideal \( \mathfrak{n}\lhd A \) has a unique mon­ic gen­er­at­or, which, by ab­use of nota­tion, we will also de­note by \( \mathfrak{n} \). Define the sub­groups \( \Gamma(\mathfrak{n})\subseteq \Gamma_1(\mathfrak{n})\subseteq \Gamma_0(\mathfrak{n})\subseteq \mathrm{GL}_r(A) \) as \[ \Gamma(\mathfrak{n})=\left\{\gamma\in \mathrm{GL}_r(A)\mid \gamma\equiv 1 (\operatorname{mod} \mathfrak{n})\right\}, \] \[ \Gamma_1(\mathfrak{n})= \left\{\begin{pmatrix} c_{ij}\end{pmatrix}\in \mathrm{GL}_r(A)\mid \begin{pmatrix} c_{11}\\ c_{21}\\ \vdots\\ c_{r1} \end{pmatrix}\equiv \begin{pmatrix} 1\\ 0\\ \vdots\\ 0 \end{pmatrix} (\operatorname{mod}\mathfrak{n})\right\}, \] \[ \Gamma_0(\mathfrak{n})= \left\{\begin{pmatrix} c_{ij}\end{pmatrix}\in \mathrm{GL}_r(A)\mid \begin{pmatrix} c_{11}\\ c_{21}\\ \vdots\\ c_{r1} \end{pmatrix}\equiv \begin{pmatrix} \ast\\ 0\\ \vdots\\ 0 \end{pmatrix} (\operatorname{mod}\mathfrak{n})\right\}. \] Ex­tend­ing the above bijec­tion, one shows that the or­bits \( Y^r_0(\mathfrak{n})(\mathbb{C}_\infty):= \Gamma_0(\mathfrak{n})\setminus \Omega^r \) are in bijec­tion with the iso­morph­ism classes of pairs \( (\phi, G_\mathfrak{n}) \), where \( \phi \) is a Drin­feld mod­ule of rank \( r \) and \( G_\mathfrak{n}\subset \phi[\mathfrak{n}] \) is an \( A \)-sub­mod­ule iso­morph­ic to \( A/\mathfrak{n} \) (two pairs \( (\phi, G_\mathfrak{n}) \) and \( (\phi^{\prime}, G_\mathfrak{n}^{\prime}) \) are iso­morph­ic if there is an iso­morph­ism \( u\colon \phi\overset{\sim}{\to}\phi^{\prime} \) such that \( u(G_n)=G_n^{\prime} \)). Sim­il­arly, \[ Y^r_1(\mathfrak{n})(\mathbb{C}_\infty):=\Gamma_1(\mathfrak{n})\setminus \Omega^r \longleftrightarrow \{\text{isomorphism classes of pairs } (\phi, P_\mathfrak{n})\}, \] where \( \phi \) is a Drin­feld mod­ule of rank \( r \) and \( P_\mathfrak{n}\in \phi[\mathfrak{n}] \) is a tor­sion point which gen­er­ates an \( A \)-sub­mod­ule iso­morph­ic to \( A/\mathfrak{n} \). Fi­nally, \( Y^r(\mathfrak{n})(\mathbb{C}_\infty):=\Gamma(\mathfrak{n})\setminus \Omega^r \) clas­si­fies the iso­morph­ism classes of pairs \( (\phi, \tilde{\iota}) \), where \( \tilde{\iota} \) is a choice of an \( A/\mathfrak{n} \)-basis of \( \phi[\mathfrak{n}] \) such that the Weil pair­ing ([e109], p. 298) on this basis takes a cer­tain value in \( (A/\mathfrak{n})^\times/\mathbb{F}_q^\times \).

The quo­tients \( Y=Y^r(\mathfrak{n}), Y^r_1(\mathfrak{n}), Y^r_0(\mathfrak{n}) \) are much more than just sets. In [e10], Drin­feld showed that \( \Omega^r \) has a nat­ur­al struc­ture of a smooth ri­gid-ana­lyt­ic man­i­fold over \( F_\infty \), and that the group \( \mathrm{GL}_r(A) \) acts dis­con­tinu­ously on \( \Omega^r \), so \( Y \) has a struc­ture of an ana­lyt­ic man­i­fold over \( F_\infty \). Moreover, he proved that \( Y \) is al­geb­ra­iz­able, in the sense that it is the ana­lyt­ic space cor­res­pond­ing to an af­fine al­geb­ra­ic vari­ety over \( F_\infty \). This point of view will be dis­cussed at the end of Sec­tion 3.2. For more about ana­lyt­ic geo­metry over non­archimedean fields, the read­er might con­sult [e61].

3.2. Algebraic theory
An \( A \)-field is a field \( K \) equipped with an \( \mathbb{F}_q \)-al­gebra ho­mo­morph­ism \( \gamma\colon A\to K \). We will de­note \( t=\gamma(T) \). We call a nonzero prime ideal \( \mathfrak{p}\lhd A \) a prime of \( A \); the primes of \( A \) are in bijec­tion with the mon­ic ir­re­du­cible poly­no­mi­als of \( A \) of pos­it­ive de­gree. Since \( K \) is a field, there are two pos­sib­il­it­ies for the ker­nel of \( \gamma \): either \( \ker(\gamma)=0 \) or \( \ker(\gamma)=\mathfrak{p} \) is a prime of \( A \). We call \( \operatorname{char}_A(K):= \ker(\gamma) \) the \( A \)-char­ac­ter­ist­ic of \( K \).

Let \( K\{x\} \) be the non­com­mut­at­ive ring of \( \mathbb{F}_q \)-lin­ear poly­no­mi­als with coef­fi­cients in \( K \), defined as earli­er for \( K=\mathbb{C}_\infty \). The ring \( K\{x\} \) can be iden­ti­fied with the ring of \( \mathbb{F}_q \)-lin­ear en­do­morph­isms of the ad­dit­ive group scheme \( \mathbb{G}_{a, K} \) over \( K \). A Drin­feld mod­ule of rank \( r\geq 1 \) over \( K \) is an \( \mathbb{F}_q \)-al­gebra ho­mo­morph­ism \( \phi\colon A \to K\{x\} \), \( a\mapsto \phi_a(x) \), such that \[ \phi_T(x)=tx+g_1x^q+\cdots+g_rx^{q^r}, \quad g_r\neq 0. \] Note that the ho­mo­morph­ism \( \phi \) is al­ways in­ject­ive, even if \( \operatorname{char}_A(K)\neq 0 \), so \( \phi \) gives an em­bed­ding of the com­mut­at­ive ring \( A \) in­to the non­com­mut­at­ive ring \( K\{x\} \). From the defin­i­tion it fol­lows that \( \partial \phi_a=\gamma(a) \) for all \( a\in A \). Hence, \( \phi \) equips \( \mathbb{G}_{a, K} \) with a new ac­tion of \( A \) such that on the tan­gent space of \( \mathbb{G}_{a, K} \) around the ori­gin the in­duced ac­tion of \( A \) is via the struc­ture morph­ism \( \gamma \).

Giv­en a poly­no­mi­al \( f(x)=a_0x+a_1x^q +\cdots+ a_dx^{q^d}\in K\{x\} \), the smal­lest in­dex \( 0\leq h\leq d \) such that \( a_h\neq 0 \) is the height of \( f \), de­noted \( \operatorname{ht}(f) \). If \( \operatorname{char}_A(K)=\mathfrak{p}\neq 0 \), then \( \gamma \) factors through the quo­tient \( A\to A/\mathfrak{p} \) and \( \operatorname{ht}(\phi_\mathfrak{p}) > 0 \). Us­ing the com­mut­a­tion \( \phi_a(\phi_\mathfrak{p}(x))=\phi_\mathfrak{p}(\phi_a(x)) \), it is not hard to show that there is an in­teger \( 1\leq H(\phi)\leq r \), called the height of \( \phi \), such that for all nonzero \( a\in A \) we have \[ \operatorname{ht}(\phi_a)=H(\phi)\cdot \operatorname{ord}_\mathfrak{p}(a)\cdot \deg(\mathfrak{p}), \] where \( \operatorname{ord}_\mathfrak{p}(a) \) is the power with which \( \mathfrak{p} \) di­vides \( a \).

Giv­en a Drin­feld mod­ule \( \phi \) over \( K \) of rank \( r \) and \( 0\neq a\in A \), let \( \phi[a] \) be the set of roots of \( \phi_a(x) \) in an al­geb­ra­ic clos­ure \( \overline{K} \) of \( K \) without re­pe­ti­tions. As over \( \mathbb{C}_\infty \), the set \( \phi[a] \) is nat­ur­ally an \( A \)-mod­ule. De­com­pos­ing \( a=\mathfrak{p}_1^{s_1}\cdots\mathfrak{p}_m^{s_m} \) in­to dis­tinct prime powers, we get an iso­morph­ism of \( A \)-mod­ules \[ \phi[a]\cong \phi[\mathfrak{p}_1^{s_1}]\times \cdots \times \phi[\mathfrak{p}_m^{s_m}]. \] Moreover, for a prime \( \mathfrak{p} \) we have \[ \phi[\mathfrak{p}^n]\cong \begin{cases} (A/\mathfrak{p}^n)^r, & \text{if } \mathfrak{p}\neq \operatorname{char}_A(K); \\ (A/\mathfrak{p}^n)^{r-H(\phi)}, & \text{if } \mathfrak{p} = \operatorname{char}_A(K). \end{cases} \]

As­sume that \( \operatorname{char}_A(K)\nmid a \). In this case the poly­no­mi­al \( \phi_a(x)\in K[x] \) is sep­ar­able, so the \( A \)-mod­ule \( \phi[a] \) is nat­ur­ally equipped with an ac­tion of the ab­so­lute Galois group \( G_K:=\operatorname{Gal}(K^{\operatorname{sep}}/K) \) of \( K \). Since this ac­tion com­mutes with the ac­tion of \( A \), we ob­tain a rep­res­ent­a­tion \begin{equation}\label{eqResRep} \rho_{\phi, a}: G_K\longrightarrow \operatorname{Aut}_A(\phi[a])\cong \mathrm{GL}_r(A/aA). \tag{3.6} \end{equation}

Ex­ample 3.5: Let \( K=F \) and \( \gamma\colon A\to F \) be the natural embedding of \( A \) into its field of fractions. Consider the Carlitz module defined by \( C_T=Tx+x^q \) as a Drinfeld module of rank 1 over \( F \). The extensions \( F(C[a])/F \) are the analogues of cyclotomic extensions of \( \mathbb{Q} \); see [e11]. For example, \( F(C[a])/F \) is unramified at the primes of \( A \) not dividing \( a \) and \( \operatorname{Gal}(F(C[a])/F) \cong (A/aA)^\times \), where the isomorphism is given by mapping the Frobenius at \( \mathfrak{p} \) to the residue of \( \mathfrak{p} \) modulo \( a \).

Re­peat­ing an earli­er defin­i­tion, a morph­ism \( u\colon \phi\to \psi \) is an ele­ment \( u\in K\{x\} \) such that \( u\circ \phi_T=\psi_T\circ u \) (and thus, \( u\circ \phi_a=\psi_a\circ u \) for all \( a\in A \)). The set \( \operatorname{Hom}_K(\phi, \psi) \) of all morph­isms \( \phi\to \psi \) is nat­ur­ally an \( A \)-mod­ule, with the ac­tion of \( a\in A \) on \( u\in \operatorname{Hom}_K(\phi, \psi) \) defined by \( a * u=u\circ \phi_a=\psi_a\circ u \). A ba­sic fact of the the­ory of Drin­feld mod­ules is that if \( \phi \) and \( \psi \) have rank \( r \), then \( \operatorname{Hom}_K(\phi, \psi) \) is a free \( A \)-mod­ule of rank \( \leq r^2 \). De­note \( \operatorname{End}_K(\phi):=\operatorname{Hom}_K(\phi, \phi) \); this is the cent­ral­izer of \( \phi(A) \) in \( K\{x\} \).

Ex­ample 3.6: Let \( K= \mathbb{F}_{q^2} \) and \( \gamma\colon A\to A/(T)\cong \mathbb{F}_q \hookrightarrow K \). Let \( \phi\colon A\to K\{x\} \) be the Drinfeld module of rank 2 defined by \( \phi_T=tx+x^{q^2}=x^{q^2} \). In this case we have \( \operatorname{End}_K(\phi)=K\{x\} \), since \( \phi_T \) is in the center of \( K\{x\} \). Now it is not hard to see that \( \operatorname{End}_K(\phi) \) has rank 4 over \( A \). But one can say more: \( \operatorname{End}_K(\phi) \) is a maximal \( A \)-order in the quaternion division algebra over \( F \) ramified at \( T \) and \( \infty \). To see this, assume for simplicity that \( q \) is odd. Fix a nonsquare \( \alpha \) in \( \mathbb{F}_q^\times \) and let \( j\in \mathbb{F}_{q^2} \) be such that \( j^2=\alpha \). The conjugate of \( j \) over \( \mathbb{F}_q \) is \( -j=j^q \). Thus, \( \mathbb{F}_{q^2}\cong \mathbb{F}_q(j) \), and \( x^q\circ j=j^q\circ x^q=-j\circ x^q \). If we denote \( i=x^q \), then \( i\,\circ\,i=\phi_T \), and we see that \( K\{\tau\}\cong A[i, j] \), where \( i^2=T \), \( j^2=\alpha \), \( ij=-ji \).
Ex­ample 3.7: If \( K=\mathbb{C}_\infty \), then \[ \operatorname{End}_{\mathbb{C}_\infty}(\phi) \overset{\sim}{\longrightarrow}\{c\in \mathbb{C}_\infty\mid c\Lambda_\phi\subseteq \Lambda_\phi\}. \] This can be used to show that \( \operatorname{End}_{\mathbb{C}_\infty}(\phi) \) is commutative and the rank of \( \operatorname{End}_{\mathbb{C}_\infty}(\phi) \) as an \( A \)-module divides \( r \). To construct a Drinfeld module with large endomorphism ring in this setting, let \( L/F \) be a field extension such that there is a unique place in \( L \) over \( \infty \). Let \( B \) be the integral closure of \( A \) in \( L \). Let \( 1=z_1, z_2, \dots, z_r\in B \) be a basis of \( L \) as a vector space over \( F \). Put \[ \Lambda=Az_1+Az_2+\cdots +Az_r. \] Then \( \Lambda\subset \mathbb{C}_\infty \) is an \( A \)-lattice of rank \( r \) and \( \operatorname{End}_{\mathbb{C}_\infty}(\phi^\Lambda)\cong B \).

The ac­tion of \( u\in \operatorname{End}_K(\phi) \) on \( \phi[a] \) com­mutes with the ac­tion of \( G_K \). If \( \mathfrak{p}\neq \operatorname{char}_A(\phi) \), then there is an in­ject­ive ho­mo­morph­ism \[ \operatorname{End}_K(\phi)\otimes_A A/aA\longrightarrow \operatorname{End}_{A[G_K]}(\phi[a]). \] Thus, the im­age of \( G_K \) in \( \operatorname{Aut}(\phi[a])\cong \mathrm{GL}_r(A/aA) \) is pro­por­tion­ally smal­ler to the size of \( \operatorname{End}_K(\phi) \) (lar­ger the en­do­morph­ism ring, smal­ler the im­age of \( G_K \) will be).

Defin­i­tion 3.8: A morphism \( u\colon \phi\to \psi \) over \( K \) is an iso­morph­ism if it has an inverse in \( K\{x\} \). Hence, an isomorphism is given by a nonzero constant \( c\in K \) such that \( \phi_T(x)=c\psi_T(c^{-1}x) \). If \[ \phi_T=tx+g_1x^q+\cdots+g_rx^{q^r}\quad\text{ and }\quad \psi_T=tx+h_1x^q+\cdots+h_rx^{q^r}, \] then this is equivalent to \( g_i=h_i c^{q^i-1} \) for all \( 1\leq i\leq r \).
Ex­ample 3.9: Suppose \( \phi \) and \( \psi \) have rank 1. Let \( c \) be a root of \( x^{q-1}=g_1/h_1 \). Then \( c\phi_Tc^{-1}=\psi_T \), so \( \phi \) and \( \psi \) are isomorphic over \( K(\sqrt[q-1]{g_1/h_1}\,) \). This implies that, up to isomorphism, the Carlitz module \( C_T=tx+x^q \) is the only Drinfeld module of rank 1 over \( K^{\operatorname{sep}} \). But note that \( x^{q-1}=g_1/h_1 \) might not have roots in \( K \), so \( \phi \) and \( \psi \) might not be isomorphic over \( K \).

Now sup­pose \( \phi \) and \( \psi \) have rank 2. The \( j \)-in­vari­ant of \( \phi \) is \[ j(\phi):=g_1^{q+1}/g_2. \] It is not hard to check that \( \phi \) and \( \psi \) are iso­morph­ic over \( \overline{K} \) if and only if \( j(\phi)=j(\psi) \).

For a Drin­feld mod­ule \( \phi \) of rank \( \geq 3 \) there is a fi­nite col­lec­tion of \( j \)-in­vari­ants, which are ra­tion­al func­tions in the coef­fi­cients of \( \phi \) and which dis­tin­guish the iso­morph­ism class of \( \phi \) over \( \overline{K} \). These were con­struc­ted by Po­temine [e53]. These \( j \)-in­vari­ants are not al­geb­ra­ic­ally in­de­pend­ent, which is re­flec­ted in the fact that the mod­uli space of Drin­feld mod­ules of rank \( r\geq 3 \) is not iso­morph­ic to the af­fine space \( \mathbb{A}^{r-1} \).

A \( \Gamma(\mathfrak{n}) \)-struc­ture on a Drin­feld mod­ule \( \phi \) of rank \( r \) over \( K \), if \( \operatorname{char}_A(K)\nmid \mathfrak{n} \), is an iso­morph­ism of \( A \)-mod­ules \[ \iota\colon (A/\mathfrak{n})^r \longrightarrow \phi[\mathfrak{n}](K), \] where \( \phi[\mathfrak{n}](K) \) is the set of \( K \)-ra­tion­al \( \mathfrak{n} \)-tor­sion points of \( \phi \). A \( \Gamma_1(\mathfrak{n}) \)-struc­ture is an in­ject­ive ho­mo­morph­ism \( A/\mathfrak{n} \to \phi[\mathfrak{n}](K) \), and a \( \Gamma_0(\mathfrak{n}) \)-struc­ture is a morph­ism \( \phi\to \psi \) over \( K \) whose ker­nel is iso­morph­ic to \( A/\mathfrak{n} \). The no­tion of Drin­feld mod­ule and level struc­tures can be ex­ten­ded to an ar­bit­rary \( A \)-scheme \( S \) (in­stead of \( S=\operatorname{Spec}(K) \)): a Drin­feld mod­ule over \( S \) is a ho­mo­morph­ism \( \phi\colon A\to \operatorname{End}(\mathcal{L}) \) from \( A \) to the ring of \( S \)-en­do­morph­isms of the group scheme un­der­ly­ing a line bundle \( \mathcal{L} \) on \( S \) sat­is­fy­ing cer­tain con­di­tions, and \( \Gamma(\mathfrak{n}) \)-struc­ture is a ho­mo­morph­ism of \( A \)-mod­ules \( \iota: (A/\mathfrak{n})^r \longrightarrow \phi[\mathfrak{n}](S) \), which in­duces an equal­ity of Carti­er di­visors \( \phi[\mathfrak{n}](S) = \sum_{\alpha\in (A/\mathfrak{n})^r} \iota(\alpha) \). There res­ults a mod­uli func­tor \( \mathcal{M}^r(\mathfrak{n}) \) from the cat­egory of \( A \)-schemes to the cat­egory of sets which to an \( A \)-scheme \( S \) as­so­ci­ates the set \( \mathcal{M}^r(\mathfrak{n})(S) \) of iso­morph­ism classes of Drin­feld mod­ules over \( S \) of rank \( r \) equipped with a \( \Gamma(\mathfrak{n}) \)-struc­ture. Un­der the as­sump­tion that \( \mathfrak{n} \) is di­vis­ible by at least two primes, Drin­feld proved in [e10] that \( \mathcal{M}^r(\mathfrak{n}) \) is rep­res­ent­able by an af­fine flat \( A \)-scheme \( M^r(\mathfrak{n}) \) of di­men­sion \( r-1 \) whose fibers over \( \operatorname{Spec}(A) \) are smooth away from \( \mathfrak{n} \). Tak­ing the quo­tients of \( M^r(\mathfrak{n}) \) by fi­nite groups one con­structs (coarse) mod­uli schemes \( Y^r_1(\mathfrak{n}) \) and \( Y^r_0(\mathfrak{n}) \) clas­si­fy­ing Drin­feld mod­ules with level \( \Gamma_1(\mathfrak{n}) \) and \( \Gamma_0(\mathfrak{n}) \)-struc­tures. One of the con­nec­ted com­pon­ents of \( M^r(\mathfrak{n})(\mathbb{C}_\infty) \) is \( Y^r(\mathfrak{n})(\mathbb{C}_\infty) \), con­struc­ted earli­er as a quo­tient of \( \Omega^r \) by the group \( \Gamma(\mathfrak{n}) \) (and sim­il­arly, the \( \mathbb{C}_\infty \)-val­ued points on \( Y^r_1(\mathfrak{n}) \) and \( Y^r_0(\mathfrak{n}) \) are in bijec­tion with the quo­tients of \( \Omega^r \) un­der the ac­tion of \( \Gamma_1(\mathfrak{n}) \) and \( \Gamma_0(\mathfrak{n}) \), re­spect­ively).

As­sume \( r=2 \). Let \( Y_0(\mathfrak{n}):= Y_0^2(\mathfrak{n}) \) and de­note by \( X_0(\mathfrak{n}) \) the unique pro­ject­ive smooth geo­met­ric­ally con­nec­ted curve con­tain­ing \( Y_0(\mathfrak{n}) \) as an af­fine sub­vari­ety. The set of points \( X_0(\mathfrak{n})-Y_0(\mathfrak{n}) \) are the cusps of \( X_0(\mathfrak{n}) \).

Ex­ample 3.10: Over \( \mathbb{C}_\infty \) the \( j \)-invariant of Drinfeld modules gives an isomorphism between the projective line \( \mathbb{P}^1_{\mathbb{C}_\infty} \) and \( X_0(1) \). Let \begin{equation}\label{eqUDM} \phi_T(x)=Tx+x^q+j^{-1}x^{q^2}, \tag{3.7} \end{equation} be the “universal” Drinfeld module of rank 2 with \( j \)-invariant \( j \), where we consider \( j \) as a variable. Let \( \mathcal{C}\cong A/TA \) be a cyclic submodule of \( \phi \). Then \( \mathcal{C} \) is the set of roots of an \( \mathbb{F}_q \)-linear polynomial of the form \( f_{\mathcal{C}}(x)=x+\alpha x^q \), where \( \alpha \) is another variable (we may assume that the coefficient of \( x \) in \( f_{\mathcal{C}}(x) \) is 1 because the polynomial is separable and the set of zeros does not change if we multiply \( f_{\mathcal{C}} \) by a nonzero constant). Since \( \mathcal{C}\subset \phi[T] \), we have \[ \phi_T(x)= (Tx+\tilde{\alpha}x^q)\circ (x+\alpha x^q) \] for some \( \tilde{\alpha} \). Thus, \( j^{-1}=\tilde{\alpha}\alpha^q \) and \( \alpha T+\tilde{\alpha}=1 \). This leads to \( j^{-1}=(1-\alpha T)\alpha^q \). Substituting \( \alpha\mapsto -1/\alpha \), we obtain \[ T+\alpha+j^{-1}\alpha^{q+1} = 0. \] Note that \( \mathcal{C} \) is automatically a \( \phi(A) \)-module, i.e., \( f_{\mathcal{C}}\circ \phi_T=g\circ f_{\mathcal{C}} \) for some \( g\in \mathbb{C}_\infty\{x\} \), since any root of \( f_{\mathcal{C}}(x) \) maps to 0 under the action of \( \phi_T \). Thus, \( X_0(T) \) is defined by \[ j=-\frac{\alpha^{q+1}}{\alpha+T}. \] Similarly, \( X_0(T+1) \) is defined by \( j=-\beta^{q+1}/(\beta+(T+1)) \) for another variable \( \beta \). A cyclic \( T(T+1) \)-submodule of \( \phi \) decomposes into a direct product of cyclic \( T \) and \( T+1 \) submodules. Changing the notation for the variables \( \alpha \) and \( \beta \) to the more conventional \( x \) and \( y \), we obtain the following as an equation of the curve \( X_0(T(T+1)) \) in the affine plane \( \operatorname{Spec}(\mathbb{C}_\infty[x, y]) \): \[ \frac{x^{q+1}}{T+x} = \frac{y^{q+1}}{(T+1)+y}. \]

4. Modular forms over \( \mathbb{F}_q(T) \)

In the con­text of Drin­feld mod­u­lar vari­et­ies there are two dif­fer­ent con­cepts that gen­er­al­ize the clas­sic­al mod­u­lar forms. The first are the Drin­feld mod­u­lar forms, which are \( \mathbb{C}_\infty \)-val­ued holo­morph­ic func­tions on \( \Omega^r \). The second are the Drin­feld auto­morph­ic forms, which are \( \mathbb{C} \)-val­ued func­tions on cer­tain ad­ele groups; these auto­morph­ic forms have a com­bin­at­or­i­al in­ter­pret­a­tion as har­mon­ic co­chains on the Bruhat–Tits build­ing \( \mathcal{B}^r \) of \( \operatorname{PGL}_r(F_\infty) \). Both of these con­cepts are im­port­ant in the study of Drin­feld mod­u­lar vari­et­ies.
4.1. Drinfeld modular forms
Our main ref­er­ences are [e17], [e26], [e29], [e92], [e116]. Let \( r\geq 2 \) and let \( \Gamma \) be a con­gru­ence sub­group of \( \mathrm{GL}_r(A) \), i.e., \( \Gamma(\mathfrak{n})\subseteq \Gamma\subseteq \mathrm{GL}_r(A) \) for some \( \mathfrak{n}\neq 0 \). If we nor­mal­ize the pro­ject­ive co­ordin­ates of \( \boldsymbol{z}=(z_1, \dots, z_r)\in \Omega^r \) by \( z_r=1 \), then \( \gamma=(\gamma_{m,n})\in \Gamma \) acts on \( \Omega^r \) as \[ \gamma \boldsymbol{z} = j(\gamma, \boldsymbol{z})^{-1} (\,\dots, \sum_n \gamma_{m, n} z_n, \dots ), \] where \( j(\gamma, \boldsymbol{z}) = \sum_{n=1}^r \gamma_{r, n}z_n \). For ex­ample, if \( r=2 \) and we identi­fy \( \Omega^2 \) with \( \mathbb{C}_\infty - F_\infty \), then for \[ \gamma = \begin{pmatrix}a&b\\c&d\end{pmatrix}\in\Gamma \] we have \[ j(\gamma, \boldsymbol{z})=j(\gamma,z) = cz+d\quad\text{ and }\quad \gamma \boldsymbol{z} = \gamma z = \dfrac{az+b}{cz+d}. \]
Defin­i­tion 4.1: A Drin­feld mod­u­lar form for \( \Gamma \) of weight \( k\in \mathbb{Z}_{\geq 0} \) and type \( m\in \mathbb{Z}/(q-1)\mathbb{Z} \) is a holo­morph­ic func­tion \[ f\colon \Omega^r\longrightarrow \mathbb{C}_\infty \] such that
  1. \( f(\gamma \boldsymbol{z}) =(\det \gamma)^{-m} j(\gamma, \boldsymbol{z})^k f(\boldsymbol{z}) \) for all \( \gamma\in \Gamma \), and
  2. \( f(\boldsymbol{z}) \) is holo­morph­ic at the cusps of \( \Gamma \).

De­note the space of such func­tions by \( M^r_{k, m}(\Gamma) \). (It is shown in [e116] that \( M^r_{k, m}(\Gamma) \) is fi­nite di­men­sion­al over \( \mathbb{C}_\infty \).)

Con­di­tion (ii) is tech­nic­ally com­plic­ated to ex­plain when \( r\geq 3 \), so we will ex­plain it only for \( r=2 \), and refer to [e116] for \( r\geq 2 \). Be­cause \( \Gamma \) is a con­gru­ence group, it con­tains the sub­group \( U_b=\left(\begin{smallmatrix} 1 & bA \\ 0 & 1\end{smallmatrix}\right) \) for some nonzero \( b\in A \). Con­di­tion (i) im­plies that \( f(z+b)=f(z) \), which it­self im­plies that \( f(z) \) can be ex­pan­ded as \[ f(z)=\sum_{n\in \mathbb{Z}} a_n (1/e_{bA}(z))^n, \qquad a_n\in \mathbb{C}_\infty, \] as­sum­ing \( \Im(z):= \inf_{\alpha\in F_\infty}\lvert z-\alpha \rvert\gg 0 \) (for sim­pli­city, we will omit this con­di­tion in what fol­lows). We say that \( f(z) \) is holo­morph­ic at the cusp \( [\infty] \) if in the above ex­pan­sion \( a_n=0 \) for all \( n < 0 \) (this van­ish­ing of coef­fi­cients with neg­at­ive in­dices does not de­pend on the choice of \( b \)). Next, for \( g\in \mathrm{GL}_2(A) \), put \( f|_g (z) = (\det g)^m j(g,z)^{-k}f(g z) \). This \( f|_g \) sat­is­fies (i) for any \( \gamma \in g^{-1}\Gamma g \), which is again a con­gru­ence group. Con­di­tion (ii) means that \( f|_g \) is holo­morph­ic at \( [\infty] \) for all \( g\in \mathrm{GL}_2(A) \). (Note that \( f|_g=f \) for \( g\in \Gamma \), so for this last con­di­tion to hold it suf­fices that \( f|_g \) is holo­morph­ic at \( [\infty] \) for left coset rep­res­ent­at­ives of \( \Gamma \) in \( \mathrm{GL}_2(A) \).)

Ex­ample 4.2: Important examples of modular forms for \( \mathrm{GL}_r(A) \) arise as “coefficient forms”. For \( \boldsymbol{z}=(z_1, \dots, z_r)\in \Omega^r \), we define the rank-\( r \) lattice \( \Lambda_{\boldsymbol{z}}=Az_1+Az_2+\cdots+Az_r\subset \mathbb{C}_\infty \). Denote the Drinfeld module of rank \( r \) associated to \( \Lambda_{\boldsymbol{z}} \) by \( \phi^{\boldsymbol{z}} \). It is determined by \begin{equation}\label{eqCoeffForms} \phi^{\boldsymbol{z}}_T=Tx+g_1(\boldsymbol{z})x^q+\cdots+g_r(\boldsymbol{z})x^{q^r}. \tag{4.1} \end{equation} The functions \( g_i(\boldsymbol{z}) \), \( 1\leq i\leq r \), are Drinfeld modular forms for \( \mathrm{GL}_r(A) \) of type 0 and weights \( q^i-1 \). The function \( \Delta_r(\boldsymbol{z}):= g_r(\boldsymbol{z}) \), called the Drin­feld dis­crim­in­ant func­tion, plays an especially important role in the study of the cuspidal divisor group in this context; it is the analogue of Ramanujan’s \( \Delta \) function. Note that \( \Delta_r(\boldsymbol{z}) \) is nonvanishing on \( \Omega^r \) since \( g_r(\Lambda)\neq 0 \) for any \( A \)-lattice \( \Lambda \) of rank \( r \).

The Ei­s­en­stein series \( E_{q^n-1}(\boldsymbol{z}):= E_{q^n-1}(\Lambda_{\boldsymbol{z}}) \) defined in (3.1) and the coef­fi­cients \( e_{n}(\boldsymbol{z}):= e_{n}(\Lambda_{\boldsymbol{z}}) \) defined in (3.2) are Drin­feld mod­u­lar forms of weight \( q^n-1 \). It can be shown that the \( \mathbb{C}_\infty \)-al­gebra of all Drin­feld mod­u­lar forms of type 0 is a poly­no­mi­al ring: \begin{align*} \smash[b]{\bigoplus_{k\geq 1} M^r_{k, 0}}(\mathrm{GL}_r(A)) &= \mathbb{C}_\infty[g_1, \dots, g_r] \\ & = \mathbb{C}_\infty[e_1, \dots, e_r]\\ & = \mathbb{C}_\infty[E_{q-1}, E_{q^2-1}, \dots, E_{q^r-1}]. \end{align*}

Ex­ample 4.3: The \( j \)-func­tion on \( \Omega^2 \) is \( j(z):= g_1(z)^{q+1}/g_2(z) \). This function is holomorphic on \( \Omega \) but has a pole at the cusp \( [\infty] \). Since \( j(\gamma z)=j(z) \) for all \( \gamma\in \mathrm{GL}_2(A) \), it defines a rational function on \( X_0(1) \). In fact, \( j(z) \) generates the field of rational functions on \( X_0(1)\cong \mathbb{P}^1_{\mathbb{C}_\infty} \).

Now we spe­cial­ize to \( r=2 \) and \( \Gamma=\Gamma_0(\mathfrak{n}) \). To sim­pli­fy the nota­tion, we will omit the su­per­script \( r \), so for ex­ample \( \Omega:=\Omega^2 \) and \( M_{k,m}(\mathfrak{n}) := M^2_{k,m}(\Gamma_0(\mathfrak{n})) \). Since \( \Gamma_0(\mathfrak{n}) \) con­tains the group of scal­ar matrices \( \left(\begin{smallmatrix} \alpha & 0 \\ 0 & \alpha\end{smallmatrix}\right) \), \( \alpha\in \mathbb{F}_q^\times \), ap­ply­ing con­di­tion (i) to such matrices one con­cludes that if \( M_{k, m}(\mathfrak{n})\neq 0 \), then \( k\equiv 2m(\operatorname{mod} q-1) \). Hence, if \( q \) is odd and \( M_{k, m}(\mathfrak{n})\neq 0 \), then \( k \) is ne­ces­sar­ily even and \( m=k/2 \) or \( m=k/2+(q-1)/2 \) mod­ulo \( q-1 \). Next, a simple cal­cu­la­tion shows that the dif­fer­en­tial \( \mathrm{d} z \) on \( \Omega \) sat­is­fies \[ \mathrm{d}(\gamma z) = \frac{\det(\gamma)}{(cz+d)^2}\,\mathrm{d} z \quad \text{for all }\gamma\in \mathrm{GL}_2(F_\infty). \] Hence, if \( f(z)\in M_{2k, k}(\mathfrak{n}) \), then \( f(z)(\mathrm{d} z)^k \) can be iden­ti­fied with a \( k \)-fold dif­fer­en­tial form on the Drin­feld mod­u­lar curve \( X_0(\mathfrak{n}) \).

Since \( \Gamma_0(\mathfrak{n}) \) con­tains the sub­group \( U_1 \), the ex­pan­sion of a Drin­feld mod­u­lar form \( f \) at \( [\infty] \) is \( f(z)=\sum_{n\geq 0} a_n (1/e_A(z))^n \). In­stead we will now use the para­met­er \[ t(z) = \frac{1}{\pi_C e_A(z)} = \frac{1}{e_C(\pi_C z)} = \frac{1}{\pi_C} \sum_{a\in A} \frac{1}{z+a}, \] which plays the role of the clas­sic­al para­met­er \( q = e^{2i\pi z} \) (this renor­mal­iz­a­tion is chosen so that the ex­pan­sions of the mod­u­lar forms \( \pi_C^{1-q} g_1 \) and \( \pi_{C}^{1-q^2} g_2 \) at \( [\infty] \) have coef­fi­cients in \( A \)). Then we have \( f = \sum_{n\geq 0} a^{\prime}_n t^n \) with \( a^{\prime}_n \in \mathbb{C}_\infty \). Since \( t(\alpha z)=\alpha^{-1}t(z) \) for any \( \alpha \in \mathbb{F}_q^\times \), the coef­fi­cients \( a^{\prime}_n \) are zero un­less \( n\equiv m (\operatorname{mod}q-1) \) so the ex­pan­sion of \( f \) is of the form \[ f = \sum_{i\geq 0} b_i t^{m+i(q-1)}. \] A Drin­feld mod­u­lar form is said to be cuspid­al (resp. doubly cuspid­al) at the cusp \( [\infty] \) if \( a^{\prime}_0=0 \) (resp. \( a^{\prime}_0 = 0 \) and \( a^{\prime}_1=0 \)). If for all \( g\in \mathrm{GL}_2(A) \), \( f|_g \) is cuspid­al (resp. doubly cuspid­al) at \( [\infty] \), we say that the Drin­feld mod­u­lar form \( f \) is cuspid­al (resp. doubly cuspid­al). Let \( M_{k,m}^0(\mathfrak{n}) \) (resp. \( M_{k,m}^{0,0}(\mathfrak{n}) \)) de­note the sub­spaces of such func­tions. Since \( \mathrm{d} e_A(z) / \mathrm{d} z=1 \), we have \( \mathrm{d} t = -\pi_C t^2 \mathrm{d} z \) so doubly cuspid­al Drin­feld mod­u­lar forms play a role sim­il­ar to clas­sic­al cusp forms. Namely the map \( f(z) \mapsto f(z) \mathrm{d} z \) is an iso­morph­ism between \( M_{2,1}^{0,0}(\mathfrak{n}) \) and the space of holo­morph­ic dif­fer­en­tial forms on the curve \( X_0(\mathfrak{n}) \). In par­tic­u­lar the di­men­sion of \( M_{2,1}^{0,0}(\mathfrak{n}) \) is equal to the genus of \( X_0(\mathfrak{n}) \).

Hecke op­er­at­ors on Drin­feld mod­u­lar forms can be defined as double coset op­er­at­ors for \( \Gamma_0(\mathfrak{n}) \). Let \( \mathfrak{m} \lhd A \) be a nonzero ideal of \( A \). For the double coset \( \Gamma_0(\mathfrak{n}) \setminus (\Gamma_0(\mathfrak{n}) \left(\begin{smallmatrix}\mathfrak{m}&0\\0&1\end{smallmatrix}\right) \Gamma_0(\mathfrak{n})) \), a set of rep­res­ent­at­ives is \begin{equation}\label{eq-Sm} S_{\mathfrak{m}}=\bigl\{ \left(\begin{smallmatrix} a & b \\ 0 & d \end{smallmatrix}\right) \in M_2(A) : a,d \text{ monic}, (ad) = \mathfrak{m}, (a)+\mathfrak{n} = A, \deg b < \deg d\bigr\}.\tag{4.2} \end{equation} For \( f \in M_{k,m}(\mathfrak{n}) \), we define \( f|T_\mathfrak{m} = \sum_{g\in S_{\mathfrak{m}}} f|_g \). In more con­crete terms \[ f|T_{\mathfrak{m}}\, (z) = \frac{1}{\mathfrak{m}^{k-m}} \sum_{\left(\begin{smallmatrix}a&b\\0&d\end{smallmatrix}\right) \in S_\mathfrak{m}} f\left( \frac{az+b}{d} \right). \] The \( \mathfrak{m} \)-th Hecke op­er­at­or \( T_\mathfrak{m} \) is a \( \mathbb{C}_\infty \)-lin­ear trans­form­a­tion of the space \( M_{k,m}(\mathfrak{n}) \). It sta­bil­izes the cuspid­al and doubly-cuspid­al sub­spaces. The Hecke op­er­at­ors \( (T_\mathfrak{m})_{\mathfrak{m}\lhd A} \) gen­er­ate a com­mut­at­ive \( \mathbb{C}_\infty \)-sub­al­gebra of \( M_{k,m}(\mathfrak{n}) \), called the Hecke al­gebra for \( M_{k,m}(\mathfrak{n}) \). On Drin­feld mod­u­lar forms, Hecke op­er­at­ors are com­pletely mul­ti­plic­at­ive, i.e., for all \( \mathfrak{m},\mathfrak{m}^{\prime} \lhd A \), we have \( T_\mathfrak{m} T_{\mathfrak{m}^{\prime}} = T_{\mathfrak{m} \mathfrak{m}^{\prime}}=T_{\mathfrak{m}^{\prime}} T_\mathfrak{m} \). This prop­erty dis­tin­guishes them from Hecke op­er­at­ors on clas­sic­al mod­u­lar forms, where they are only mul­ti­plic­at­ive.

An­oth­er im­port­ant dif­fer­ence is that, the char­ac­ter­ist­ic be­ing pos­it­ive, the space \( M_{k,m}^{0,0}(\mathfrak{n}) \) has no in­ner product. Hence there is no guar­an­tee that \( T_\mathfrak{m} \) is di­ag­on­al­iz­able. Goss was the first to point out this prob­lem [e16]. Since then, ex­amples of non­di­ag­on­al­iz­able Hecke op­er­at­ors have been ob­tained in spe­cial cases (see [e70] for the sub­groups \( \Gamma_1(T) \) and \( \Gamma(T) \)) but in gen­er­al, the ques­tion is still wide open. In par­tic­u­lar we know no nat­ur­al bases of Drin­feld mod­u­lar forms which are sim­ul­tan­eously ei­gen­forms for the Hecke op­er­at­ors.

4.2. Harmonic cochains
Let \( r\geq 2 \) and let \( V \) be an \( r \)-di­men­sion­al vec­tor space over \( F_\infty \). De­note by \( \mathcal{O}_\infty \) the ring of in­tegers of \( F_\infty \) and let \( \pi_\infty \) be a uni­form­izer of \( \mathcal{O}_\infty \). An \( \mathcal{O}_\infty \)-lat­tice in \( V \) is a free \( \mathcal{O}_\infty \)-mod­ule of rank \( r \) which con­tains a basis of \( V \). Two lat­tices \( L_1 \) and \( L_2 \) are ho­mothet­ic if there ex­ists \( \alpha\in F_\infty^\times \) with \( \alpha\cdot L_1=L_2 \); this defines an equi­val­ence re­la­tion on the set of lat­tices in \( V \). We de­note the equi­val­ence class of \( L \) by \( [L] \). The Bruhat–Tits build­ing of \( \operatorname{PGL}_r(F_\infty) \) is the \( (r-1) \)-di­men­sion­al sim­pli­cial com­plex \( \mathscr{B}^r \) with set of ver­tices \[ \mathrm{Ver}(\mathscr{B}^r)=\{[L]\mid L \text{ is a lattice in }V\}, \] in which the ver­tices \( [L_0], \dots, [L_n] \) form an \( n \)-sim­plex if and only if there is \( L_i^{\prime}\in [L_i] \), \( 1\leq i\leq n \), such that \[ L_0^{\prime}\supsetneq L_1^{\prime}\supsetneq \cdots \supsetneq L_n^{\prime}\supsetneq \pi_\infty L_0^{\prime}. \] Since \( \mathrm{GL}_r(F_\infty) \) acts trans­it­ively on \( \mathrm{Ver}(\mathscr{B}^r) \) and the sta­bil­izer of a ver­tex is con­jug­ate to \( F_\infty^\times \mathrm{GL}_r(\mathcal{O}_\infty) \), we have a bijec­tion \[ \mathrm{Ver}(\mathscr{B}^r) \cong \mathrm{GL}_r(F_\infty)/F_\infty^\times \mathrm{GL}_r(\mathcal{O}_\infty). \] Sim­il­ar bijec­tions ex­ist between the sets of high­er di­men­sion­al sim­plices of \( \mathscr{B}^r \) of vari­ous types and left cosets of parahor­ic sub­groups of \( \mathrm{GL}_r(\mathcal{O}_\infty) \). When \( r=2 \), \( \mathscr{B}^2 \) is an in­fin­ite tree in which every ver­tex is ad­ja­cent to ex­actly \( q+1 \) oth­er ver­tices.

Har­mon­ic co­chains on sim­pli­cial com­plexes nat­ur­ally arise in the study of com­bin­at­or­i­al lapla­cians; see [e8]. A vari­ant of har­mon­ic \( i \)-co­chains on \( \mathscr{B}^r \) was defined by de Shalit [e54]: these are func­tions on poin­ted \( i \)-sim­plices of \( \mathscr{B}^r \), \( 1\leq i\leq r-1 \), sat­is­fy­ing cer­tain con­di­tions (in [e54] the build­ing \( \mathscr{B}^r \) is defined over an ar­bit­rary loc­al field). The sig­ni­fic­ance of the group \( \operatorname{Har}^i(\mathscr{B}^r, \mathbb{Q}_\ell) \) of \( \mathbb{Q}_\ell \)-val­ued har­mon­ic \( i \)-co­chains is that it is iso­morph­ic to the \( \ell \)-ad­ic co­homo­logy group \( H^i_{\mathrm{et}}(\Omega^r, \mathbb{Q}_\ell) \) of \( \Omega^r \); see [e32], [e54]. Also, \( \operatorname{Har}^i(\mathscr{B}^r, \mathbb{Q}_\ell) \) can be in­ter­preted as a space of auto­morph­ic forms on the ad­ele group \( \mathrm{GL}_r(\mathbb{A}_F) \) with spe­cial re­stric­tion at \( \infty \); see [e68] (one trans­forms func­tions on \( \mathrm{GL}_r(\mathbb{A}_F) \) to func­tions on \( \mathrm{GL}_r(F_\infty) \) us­ing the strong ap­prox­im­a­tion the­or­em and uses the bijec­tions between the sets of sim­plices of \( \mathscr{B}^r \) and cosets in \( \operatorname{PGL}_r(F_\infty) \) men­tioned earli­er).

The most rel­ev­ant for our pur­poses are the har­mon­ic 1-co­chains, which for \( r=2 \) were already defined by van der Put in [e21].

Defin­i­tion 4.4: Let \( \mathrm{Ed}(\mathscr{B}^r) \) be the set of oriented 1-simplices of \( \mathscr{B}^r \). Let \( R \) be a commutative ring with unity. For \( r=2 \), an \( R \)-valued harmonic 1-cochain on \( \mathscr{B}^2 \) is a function \( f\colon \mathrm{Ed}(\mathscr{B}^2)\rightarrow R \) that satisfies
  1. \( f(e)+f(\overline{e})=0\quad\text{ for all }\,e\in \mathrm{Ed}(\mathscr{B}^2). \)
  2. \( \sum_{\substack{e\in \mathrm{Ed}(\mathscr{B}^2)\\o(e)=v}}f(e)=0\quad\text{ for all }\,v\in \mathrm{Ver}(\mathscr{B}^2). \)

Here, \( o(e) \) is the ori­gin of \( e \) and \( \bar{e} \) is the edge \( e \) with op­pos­ite ori­ent­a­tion.

When \( r\geq 3 \), two of the con­di­tions that define har­mon­ic 1-co­chains are sim­il­ar to (1) and (2), with (2) re­fined by the “types” of edges, where \( \mathrm{type}([L_0], [L_1])=\dim_{\mathbb{F}_q}L_0/L_1 \). But there are two ad­di­tion­al con­di­tions. One of these con­di­tions says that the sum of val­ues of \( f \) over the edges of a closed path in \( \mathscr{B}^r \) is 0, and the oth­er that the val­ues of \( f \) are uniquely de­term­ined by its val­ues on edges of type 1. We de­note the space of \( R \)-val­ued har­mon­ic 1-co­chains by \( \operatorname{Har}^1(\mathscr{B}^r, R) \).

From an­oth­er per­spect­ive, \( \mathscr{B}^r \) is a com­bin­at­or­i­al “skel­et­on” of \( \Omega^r \). In fact, there is an ad­miss­ible cov­er­ing of \( \Omega^r \) through af­fin­oid sub­spaces and a \( \mathrm{GL}_r(F_\infty) \)-equivari­ant map \( \Omega^r\to \mathscr{B}^r \) that sends af­fin­oids to sim­plices; see [e10]. The fol­low­ing im­port­ant res­ult relates the group of holo­morph­ic in­vert­ible func­tions \( \mathcal{O}(\Omega^r)^\times \) on \( \Omega^r \) to \( \mathbb{Z} \)-val­ued har­mon­ic 1-co­chains on \( \mathscr{B}^r \): \begin{equation}\label{eqGvdP} 0 \longrightarrow \mathbb{C}_\infty^\times \longrightarrow \mathcal{O}(\Omega^r)^\times \xrightarrow{\mathrm{dlog}} \mathrm{Har}^1(\mathscr{B}^r, \mathbb{Z})\longrightarrow 0, \tag{4.3} \end{equation} where \( \mathrm{dlog} \) is some sort of a log­ar­ithmic de­riv­at­ive. The ex­ist­ence of \eqref{eqGvdP} was proved by van der Put for \( r=2 \) and ex­ten­ded to ar­bit­rary \( r\geq 2 \) by Gekel­er [e99].

From now on, the dis­cus­sion will fo­cus mostly on har­mon­ic 1-co­chains on the Bruhat–Tits tree \( \mathscr{B}^2 \). To sim­pli­fy the nota­tion, we de­note \( \mathscr{T}=\mathscr{B}^2 \). The group \( \mathrm{GL}_2(F_\infty) \) acts on \( \mathscr{T} \) from the left. Let \[ \mathcal{H}(\mathfrak{n}, R):= \operatorname{Har}^1(\mathscr{T}, R)^{\Gamma_0(\mathfrak{n})} \] be the sub­mod­ule of \( R \)-val­ued har­mon­ic 1-co­chains such that \( f(\gamma e)=f(e) \) for all \( \gamma\in \Gamma_0(\mathfrak{n}) \) and \( e\in \mathrm{Ed}(\mathscr{T}) \). The mod­ule of \( R \)-val­ued \( \Gamma_0(\mathfrak{n}) \)-in­vari­ant cuspid­al har­mon­ic co­chains, de­noted \( \mathcal{H}_0(\mathfrak{n}, R) \), is the sub­mod­ule of \( \mathcal{H}(\mathfrak{n}, R) \) con­sist­ing of func­tions which have com­pact sup­port as func­tions on \( \Gamma_0(\mathfrak{n})\setminus \mathscr{T} \), i.e., func­tions which have value 0 on all but fi­nitely many edges of \( \Gamma_0(\mathfrak{n})\setminus \mathscr{T} \). We de­note by \( \mathcal{H}_{00}(\mathfrak{n}, R) \) the im­age of \( \mathcal{H}_0(\mathfrak{n},\mathbb{Z})\otimes_{\mathbb{Z}} R \) in \( \mathcal{H}_0(\mathfrak{n}, R) \). (It is easy to con­struct ex­amples where the in­clu­sion \( \mathcal{H}_{00}(\mathfrak{n}, R)\subseteq \mathcal{H}_0(\mathfrak{n}, R) \) is strict; see ([e85], Sec­tion 1.1).) It is known that the quo­tient graph \( \Gamma_0(\mathfrak{n})\setminus \mathscr{T} \) is the edge dis­joint uni­on \[ \Gamma_0(\mathfrak{n})\setminus \mathscr{T} = (\Gamma_0(\mathfrak{n})\setminus \mathscr{T})^0\cup \bigcup_{s\in \Gamma_0(\mathfrak{n})\setminus \mathbb{P}^1(F)} h_s \] of a fi­nite graph \( (\Gamma_0(\mathfrak{n})\setminus \mathscr{T})^0 \) with a fi­nite num­ber of half-lines \( h_s \), called cusps. The cusps are in bijec­tion with the or­bits of the nat­ur­al (left) ac­tion of \( \Gamma_0(\mathfrak{n}) \) on \( \mathbb{P}^1(F) \). It is clear that \( f\in \mathcal{H}(\mathfrak{n}, R) \) is cuspid­al if and only if it even­tu­ally van­ishes on each \( h_s \). One can show that \( \mathcal{H}_0(\mathfrak{n}, \mathbb{Z}) \) and \( \mathcal{H}(\mathfrak{n}, \mathbb{Z}) \) are fi­nitely gen­er­ated free \( \mathbb{Z} \)-mod­ules of rank \( g(\mathfrak{n}) \) and \( g(\mathfrak{n})+c(\mathfrak{n})-1 \), re­spect­ively, where \( g(\mathfrak{n}) \) is the genus of the curve \( X_0(\mathfrak{n}) \) and \( c(\mathfrak{n}) \) is the num­ber of its cusps.

Re­mark 4.5: Drinfeld modular forms are related to \( \mathbb{C}_\infty \)-valued harmonic cochains. More precisely, using residues of differential forms on \( \Omega \), Teitelbaum [e31] established an isomorphism between \( M_{2,1}^0(\mathfrak{n}) \) and \( \mathcal{H}_0(\mathfrak{n}, \mathbb{C}_\infty) \). This isomorphism restricts to an isomorphism \[ M_{2,1}^{00}(\mathfrak{n}) \overset{\sim}{\longrightarrow} \mathcal{H}_{00}(\mathfrak{n}, \mathbb{C}_\infty); \] see ([e47], (6.5)).

Har­mon­ic 1-co­chains in­vari­ant un­der con­gru­ence groups have Four­i­er ex­pan­sions. This the­ory for \( r=2 \) was first de­veloped by Weil in ad­el­ic lan­guage, and re­cast in­to more ex­pli­cit for­mu­las by Gekel­er [e43] and Pál [e66]. The the­ory was ex­ten­ded to ar­bit­rary \( r\geq 2 \) in [e112]. We briefly dis­cuss the \( r=2 \) case.

The edges of \( \mathscr{T} \) are in bijec­tion with \( \mathrm{GL}_2(F_\infty)/F_\infty^\times \mathcal{I}_\infty \), where \( \mathcal{I}_\infty \) is the Iwahori group: \[ \mathcal{I}_\infty:=\left\{\begin{pmatrix} a & b\\ c & d\end{pmatrix}\in \mathrm{GL}_2(\mathcal{O}_\infty)\ \bigg|\ c\in \pi_\infty\mathcal{O}_\infty\right\}. \] Let \[ \mathrm{Ed}(\mathscr{T})^+:=\left\{\begin{pmatrix} \pi_\infty^k & u \\ 0 & 1\end{pmatrix}\ \bigg|\ \begin{matrix} k\in \mathbb{Z}\\ u\in F_\infty,\ u\mod{\pi_\infty^k\mathcal{O}_\infty}\end{matrix}\right\}. \] It is not hard to show that \[ \mathrm{Ed}(\mathscr{T})=\mathrm{Ed}(\mathscr{T})^+\sqcup \mathrm{Ed}(\mathscr{T})^+ \begin{pmatrix} 0 & 1\\ \pi_\infty & 0\end{pmatrix}. \]

As­sume that \( R \) is a ring such that \( p\in R^\times \) and \( R \) con­tains the \( p \)-th roots of unity. A func­tion \( f\in \mathcal{H}(\mathfrak{n}, R) \) has a Four­i­er ex­pan­sion \[ f \left(\begin{pmatrix} \pi_\infty^k & u \\ 0 &1\end{pmatrix}\right) = f^0(\pi_\infty^k)+ \sum_{0\leq j\leq k-2} q^{-k+2+j}\sum_{\substack{\mathfrak{m}\in A\text{, monic}\\ \deg(\mathfrak{m})=j}}f^\ast(\mathfrak{m})\, \nu(\mathfrak{m} u), \] where \( \nu(x):=-1 \) if \( x \) has a term of or­der \( \pi_\infty \) in its \( \pi_\infty \)-ex­pan­sion; \( \nu(x):=q-1 \) oth­er­wise. Here the Four­i­er coef­fi­cients \( f^0(\pi_\infty^k) \) and \( f^\ast(\mathfrak{m}) \) are cer­tain fi­nite sums of val­ues of \( f \) twis­ted by a char­ac­ter \( \chi\colon F_\infty\to \mathbb{C}^\times \) tak­ing val­ues in the \( p \)-th roots of unity. Moreover, the con­stant Four­i­er coef­fi­cient \( f^0(\pi_\infty^k) \) is equal to 0 for cuspid­al har­mon­ic co­chains.

Fol­low­ing Weil, one can at­tach a Di­rich­let \( L \)-series to a cuspid­al har­mon­ic co­chain \( f \in \mathcal{H}_0(\mathfrak{n},\mathbb{C}) \). Put \[ L (f,s) = \sum_{\mathfrak{m}} f^*(\mathfrak{m}) |\mathfrak{m} |^{-s-1}, \] where the sum is over all non­neg­at­ive di­visors on \( F \). This \( L \)-series is defined for \( s\in\mathbb{C} \), is a poly­no­mi­al in \( q^{-s} \), and sat­is­fies a func­tion­al equa­tion when sub­sti­tut­ing \( s \mapsto 2-s \).

Fix a nonzero ideal \( \mathfrak{n}\lhd A \). Giv­en a nonzero ideal \( \mathfrak{m}\lhd A \), define an \( R \)-lin­ear trans­form­a­tion of the space of \( R \)-val­ued func­tions on \( \mathrm{Ed}(\mathscr{T}) \) by \begin{equation}\label{eqDefTm} f|T_\mathfrak{m}:=\sum_{g\in S_{\mathfrak{m}}} f|_g, \tag{4.4} \end{equation} where the set \( S_{\mathfrak{m}} \) has been defined in \eqref{eq-Sm}. This trans­form­a­tion is the Hecke op­er­at­or \( T_\mathfrak{m} \). Fol­low­ing a com­mon con­ven­tion, for a prime di­visor \( \mathfrak{p} \) of \( \mathfrak{n} \) one some­times writes \( U_\mathfrak{p} \) in­stead of \( T_\mathfrak{p} \).

The group-the­or­et­ic proofs of the prop­er­ties of Hecke op­er­at­ors act­ing on clas­sic­al mod­u­lar forms also work in this set­ting. In par­tic­u­lar, the Hecke op­er­at­ors pre­serve \( \mathcal{H}_0(\mathfrak{n}, R) \), and sat­is­fy the re­curs­ive for­mu­las \begin{align*} T_{\mathfrak{m}\mathfrak{m}^{\prime}}&= T_\mathfrak{m} T_{\mathfrak{m}^{\prime}}&\quad \text{if}\quad \mathfrak{m}+\mathfrak{m}^{\prime}=A,\\ T_{\mathfrak{p}^i} &= T_{\mathfrak{p}^{i-1}}T_\mathfrak{p}-|\mathfrak{p}|T_{\mathfrak{p}^{i-2}}\quad \text{if}\quad \mathfrak{p}\nmid \mathfrak{n}, \\ T_{\mathfrak{p}^i} &= T_\mathfrak{p}^i\quad \text{if}\quad \mathfrak{p}| \mathfrak{n}. \end{align*} Let \( \mathbb{T}(\mathfrak{n}) \) be the com­mut­at­ive \( \mathbb{Z} \)-sub­al­gebra of \( \operatorname{End}_\mathbb{Z}(\mathcal{H}_0(\mathfrak{n}, \mathbb{Z})) \) gen­er­ated by all Hecke op­er­at­ors.

An ex­pli­cit for­mula can be giv­en for the ac­tion of \( T_\mathfrak{m} \) on the Four­i­er ex­pan­sion of \( f\in \mathcal{H}_0(\mathfrak{n}, R) \). This for­mula im­plies that \begin{equation}\label{eqfTm} (f|T_\mathfrak{m})^\ast(1)=|\mathfrak{m}|f^\ast(\mathfrak{m}). \tag{4.5} \end{equation} Thus, the pair­ing \begin{equation}\label{eqPairing} \begin{array}{ccc} \mathbb{T}(\mathfrak{n})\times \mathcal{H}_0(\mathfrak{n}, \mathbb{Z}) & \longrightarrow & \mathbb{Z} \\ ( T, f) & \longmapsto & (f|T_\mathfrak{m})^\ast(1) \end{array}\tag{4.6} \end{equation} is nonde­gen­er­ate and be­comes per­fect after tensor­ing with \( \mathbb{Z}[p^{-1}] \).

Figure 1. \( \Gamma_0(\mathfrak{p})\setminus \mathscr{T} \), \( \mathfrak{p} \) is a prime of degree 3.
Ex­ample 4.6: The Gekeler–van der Put map \eqref{eqGvdP}, when applied to modular units arising from the Drinfeld discriminant function, produces Eisenstein harmonic 1-cochains. This is a reflection of the Kronecker limit formula in this context; see [e66], [e97]. We give an example of a special case, where the Eisenstein harmonic cochain can be described quite explicitly.

Let \( \mathfrak{p}\lhd A \) be a prime of de­gree 3. In this case, the quo­tient graph \( \Gamma_0(\mathfrak{p})\setminus \mathscr{T} \) looks like the graph in Fig­ure 1; see ([e87], Sec­tion 4.1).

The mat­rix rep­res­ent­at­ives of the edges are the fol­low­ing.

  1. The dashed edges \[ s_\infty=\begin{pmatrix} \pi_\infty & 0 \\0&1\end{pmatrix}, \quad s_1=\begin{pmatrix} \pi_\infty^3 & 0 \\0&1\end{pmatrix} \] in­dic­ate that they are the first edges on a half-line cor­res­pond­ing to the cusps \( [\infty] \) and \( [0] \), re­spect­ively;
  2. \[ a_\infty=\begin{pmatrix} \pi_\infty^2 & \pi_\infty \\0&1\end{pmatrix}, \quad a_1=\begin{pmatrix} \pi_\infty^3 & \pi_\infty^2 \\0&1\end{pmatrix}, \quad d_\infty=\begin{pmatrix} \pi_\infty^2 & 0 \\0&1\end{pmatrix}; \]
  3. There are \( q \) edges \[ b_u = \begin{pmatrix} \pi_\infty^3 & \pi_\infty + u \pi_\infty^2 \\0&1\end{pmatrix}, \quad u \in \mathbb{F}_q. \]

Let \( E:= \mathrm{dlog}(\Delta/\Delta_{\mathfrak{p}})\in \mathcal{H}(\mathfrak{p}, \mathbb{Z}) \), where \( \Delta = \Delta_2 \) and \( \Delta_{\mathfrak{p}}(z) :=\Delta(\mathfrak{p} z) \). One can com­pute the val­ues of \( E \) on the edges of \( \Gamma_0(\mathfrak{p})\setminus \mathscr{T} \) us­ing ([e49], Co­rol­lary 2.9) and ([e114], Lemma 2.4):

  1. \( E(s_\infty) = E(s_1) = (q^2+q+1)(q-1)^2 \).
  2. \( E(a_\infty) = E(\overline{a_1}) = q(q-1)^2 \) and \( E(d_\infty) = (2q+1)(q-1)^2 \).
  3. \( E(b_u) = (q-1)^2 \) for \( u\in \mathbb{F}_q \).

By ([e49], Co­rol­lary 2.11) and ([e115], Lemma 4.4), we have \[ E|U_{\mathfrak{p}}=E \quad \text{and}\quad E|T_\mathfrak{q} = (|\mathfrak{q}|+1)E\quad \text{for all prime }\, \mathfrak{q}\neq \mathfrak{p}. \] By us­ing this and ap­ply­ing ([e49], (2.5) and Co­rol­lary 2.8) and [e115], Lemma 3.6), one can com­pute the Four­i­er coef­fi­cients of \( E \):

  1. \( E^0(\pi_\infty^k) = q^{1-k} (q^2+q+1)(q-1)^2 \).
  2. \( E^\ast(1) = q^{-1}(q+1)(q-1)^2 \).
  3. \( E^\ast(\mathfrak{m}) = \frac{(q+1)(q-1)^2}{q\, |\mathfrak{m}|} \, \prod_{i=1}^s \frac{|\mathfrak{q}_i|^{k_i+1}-1}{|\mathfrak{q}_i|-1}\, \) for \( \,\mathfrak{m} = \mathfrak{p}^k \, \prod_{i=1}^s {\mathfrak{q}_i}^{k_i}\lhd A \).

Note that, sim­il­ar to the ap­pear­ance of the di­visor func­tion in the Four­i­er ex­pan­sions of clas­sic­al Ei­s­en­stein series ([e67], p. 100), the factor \( (\lvert\mathfrak{q}\rvert^{k+1}-1)/(\lvert\mathfrak{q}\rvert-1)=\lvert\mathfrak{q}\rvert^k+\cdots+\lvert\mathfrak{q}\rvert+1 \) in the ex­pres­sion for \( E^\ast(\mathfrak{m}) \) is a ver­sion of the di­visor func­tion for \( \mathfrak{q}^{k+1} \). We also point out that the term \( (q^2+q+1) \) in \( E^0(\pi_\infty^k) \) is the or­der of the cuspid­al di­visor group of \( X_0(\mathfrak{p}) \), and this type of re­la­tion is im­port­ant in the proofs in Sec­tion 6.

The ana­logue of the Petersson in­ner product in this set­ting is the pair­ing on \( \mathcal{H}_0(\mathfrak{n}, \mathbb{C}) \) defined by \[ (f, g) =\sum_{e\in \mathrm{Ed}(\Gamma_0(\mathfrak{n})\setminus \mathscr{T})} f(e)\overline{g(e)} \mu(e)^{-1}, \] where \( \mu(e)=\frac{q-1}{2}\# \operatorname{Stab}_{\Gamma_0(\mathfrak{n})}(\tilde{e}) \) and \( \tilde{e} \) is a preim­age of \( e \) in \( \mathscr{T} \). (A Haar meas­ure on \( \mathrm{GL}_2(F_\infty) \) in­duces a push-for­ward meas­ure on \( \mathrm{Ed}(\Gamma_0(\mathfrak{n})\setminus \mathscr{T}) \), which, up to a scal­ar mul­tiple, is equal to \( \mu(e) \).) The Hecke op­er­at­ors \( T_\mathfrak{m} \), \( (\mathfrak{n}, \mathfrak{m})=1 \), are self-ad­joint with re­spect to this pair­ing. Hence, the usu­al con­clu­sions about Hecke op­er­at­ors be­ing di­ag­on­al­iz­able and their ei­gen­val­ues be­ing real are also val­id in this set­ting.

The Hecke op­er­at­ors may also be defined us­ing cor­res­pond­ences on \( X_0(\mathfrak{n}) \). Re­call the mod­uli in­ter­pret­a­tion of \( Y_0(\mathfrak{n}) \): it is the gen­er­ic fiber of the coarse mod­uli scheme for pairs \( (\phi,G_\mathfrak{n}) \) con­sist­ing of a Drin­feld \( A \)-mod­ule \( \phi \) of rank 2 over \( F \) with an \( F \)-ra­tion­al \( A \)-sub­mod­ule \( G_\mathfrak{n} \) of \( \phi[\mathfrak{n}] \) iso­morph­ic to \( A/\mathfrak{n} \) (\( F \)-ra­tion­al means that \( \sigma(G_\mathfrak{n})=G_\mathfrak{n} \) for all \( \sigma\in \operatorname{Gal}(F^{\mathrm{alg}}/F) \)). For \( \mathfrak{m} \lhd A \), the Hecke op­er­at­or \( T_\mathfrak{m} \) is defined as the cor­res­pond­ence on \( Y_0(\mathfrak{n}) \) giv­en by \[ T_\mathfrak{m} : (\phi,G_\mathfrak{n}) \mapsto \sum_{G_\mathfrak{m} \cap G_\mathfrak{n} = \{ 0 \} } (\phi/G_\mathfrak{m} , (G_\mathfrak{n} + G_\mathfrak{m})/G_\mathfrak{m}). \] It uniquely ex­tends to \( X_0(\mathfrak{n}) \). The res­ult­ing cor­res­pond­ence in­duces an en­do­morph­ism of the Jac­obi­an vari­ety \( J_0(\mathfrak{n}) \) of \( X_0(\mathfrak{n}) \), also de­noted \( T_\mathfrak{m} \). The \( \mathbb{Z} \)-sub­al­gebra of \( \operatorname{End}(J_0(\mathfrak{n})) \) gen­er­ated by \( (T_{\mathfrak{m}})_{\mathfrak{m} \lhd A} \) is ca­non­ic­ally iso­morph­ic to \( \mathbb{T}(\mathfrak{n}) \); this is a con­sequence of Drin­feld’s Re­cipro­city Law ([e10], The­or­em 2). Hav­ing this fact, one can use the clas­sic­al Shimura con­struc­tion to as­so­ci­ate to a \( \mathbb{T}(\mathfrak{n}) \)-ei­gen­form \( f\in \mathcal{H}_0(\mathfrak{n}, \mathbb{C}) \) an abeli­an vari­ety \( B_f \) whose num­ber of points over \( \mathbb{F}_\mathfrak{p} \), \( \mathfrak{p}\nmid \mathfrak{n} \), is com­puted us­ing the ei­gen­value of \( T_\mathfrak{p} \) act­ing on \( f \). In par­tic­u­lar, if \( f \) has ra­tion­al ei­gen­val­ues, then \( B_f \) is an el­lipt­ic curve over \( F \). Com­bin­ing this with some deep res­ults of Grothen­dieck and De­ligne, one can de­duce the fol­low­ing ana­logue of the Mod­u­lar­ity The­or­em (see ([e47], (8.3))).

The­or­em 4.7: Let \( E \) be an el­lipt­ic curve over \( F \) with split mul­ti­plic­at­ive re­duc­tion at \( \infty \). Then there is a non­con­stant morph­ism \( X_0(\mathfrak{n})\to E \) defined over \( F \), where \( \mathfrak{n} \) is the fi­nite part of the con­duct­or of \( E \).
Defin­i­tion 4.8: The Ei­s­en­stein ideal \( \mathfrak{E}(\mathfrak{n}) \) of \( \mathbb{T}(\mathfrak{n}) \) is the ideal generated by the elements \( T_\mathfrak{p}-(|\mathfrak{p}|+1) \) for all primes \( \mathfrak{p}\nmid \mathfrak{n} \).

The quo­tient \( \mathbb{T}(\mathfrak{n})/\mathfrak{E}(\mathfrak{n}) \) is a fi­nite ring. In­deed, oth­er­wise the per­fect­ness of the pair­ing \eqref{eqPairing} im­plies that there is a nonzero \( f\in \mathcal{H}_{0}(\mathfrak{n}, \mathbb{Q}) \) an­ni­hil­ated by \( \mathfrak{E}(\mathfrak{n}) \), which con­tra­dicts Weil’s bounds us­ing the Shimura con­struc­tion men­tioned earli­er.

Re­mark 4.9: (1)  The Eisenstein ideal in the context of Drinfeld modular curves was first defined by Tamagawa [e41] as the kernel of \( \mathbb{T}(\mathfrak{p})\to \operatorname{End}_\mathbb{Z}(\mathcal{C}(\mathfrak{p})) \), where \( \mathfrak{p} \) is prime and \( \mathcal{C}(\mathfrak{p}) \) is the cuspidal divisor group of \( J_0(\mathfrak{p}) \). It can be shown that this is equivalent to Definition 4.8 when \( \mathfrak{n}=\mathfrak{p} \) is prime; see [e87]. Moreover, \( U_\mathfrak{p}-1\in \mathfrak{E}(\mathfrak{p}) \), so \( \mathfrak{E}(\mathfrak{p}) \) is the analogue of Mazur’s definition of the Eisenstein ideal in [e13].

(2)  De­pend­ing on the prob­lem where it is used, the defin­i­tion of Ei­s­en­stein ideal might be mod­i­fied to in­clude ele­ments of the form \( U_\mathfrak{p}+a \) or \( W_\mathfrak{p}+a \), where \( \mathfrak{p}\mid \mathfrak{n} \), \( a\in \mathbb{Z} \), and \( W_\mathfrak{p} \) is an Atkin–Lehner in­vol­u­tion; see [e87], [e90], [e114].

5. Torsion of Drinfeld modules

5.1. Analogue of Ogg’s conjecture
Let \( \phi \) be a Drin­feld \( A \)-mod­ule of rank 2 over \( F \). Its tor­sion \( A \)-mod­ule \( (^\phi F)_\mathrm{tor}:= \bigcup_{a\in A} \phi[a](F) \) is fi­nite, as can be proved, for in­stance, by an ar­gu­ment in­volving the re­duc­tions of \( \phi \); see Re­mark 5.1 (2). Moreover, \( (^\phi F)_\mathrm{tor} \) can be gen­er­ated by two ele­ments \[ (^\phi F)_\mathrm{tor} \cong A/\mathfrak{m} \oplus A/\mathfrak{n}, \] where \( \mathfrak{m} \) and \( \mathfrak{n} \) are nonzero ideals of \( A \) and \( \mathfrak{m} \) di­vides \( \mathfrak{n} \). For rank-2 Drin­feld mod­ules we have a coun­ter­part of Con­jec­ture TEC.
Con­jec­ture TDM: ([e58], Conjecture 1) Let \( \phi \) be a rank-2 Drin­feld \( A \)-mod­ule over \( F \). If we write \[ (^\phi F)_\mathrm{tor}\cong A/\mathfrak{m}\oplus A/\mathfrak{n}\,\text{ with }\,\mathfrak{m}|\mathfrak{n}, \] then \( \deg \mathfrak{m}+\deg \mathfrak{n}\leq 2 \). This is es­sen­tially equi­val­ent to say­ing that if \( \mathfrak{p} \) is a prime of \( A \) of de­gree \( \geq 3 \), then the Drin­feld mod­u­lar curve \( Y_1(\mathfrak{p}) \) has no \( F \)-ra­tion­al points.
Re­mark 5.1: (1)  The finite \( A \)-modules listed in Conjecture TDM as possible \( F \)-rational torsion submodules of rank-2 Drinfeld modules correspond exactly to the levels of those Drinfeld modular curves \( X_1(\mathfrak{n}) \) and \( X(\mathfrak{n}) \) which are isomorphic to \( \mathbb{P}^1_F \). Thus, these finite modules occur as the torsion submodules of infinitely many nonisomorphic Drinfeld modules.

(2)  As of 2024, Con­jec­ture TDM re­mains largely open. On the oth­er hand, un­der an ex­tra as­sump­tion, it is easy to prove. We say that the Drin­feld mod­ule \( \phi \) defined over \( F \) by \( \phi_T(x)=Tx+g_1x^q+\cdots+g_r x^{q^r} \) has good re­duc­tion at the prime \( \mathfrak{l}\lhd A \) if \( \operatorname{ord}_{\mathfrak{l}}(g_i)\geq 0 \) for \( 1\leq i\leq r \) and \( \operatorname{ord}_{\mathfrak{l}}(g_r)=0 \). If \( \phi \) has good re­duc­tion at \( \mathfrak{l} \), then re­du­cing the coef­fi­cients of \( \phi_T \) mod­ulo \( \mathfrak{l} \), one ob­tains a Drin­feld mod­ule \( \bar{\phi} \) over \( \mathbb{F}_\mathfrak{l} \) of the same rank. It is not hard to show that the re­duc­tion mod­ulo \( \mathfrak{l} \) in­duces an in­jec­tion \( ({^\phi}F)_\mathrm{tor}\hookrightarrow ({^{\bar{\phi}}}\mathbb{F}_\mathfrak{l})_\mathrm{tor} \); see ([e109], The­or­em 6.5.10). Thus, \( \# ({^\phi}F)_\mathrm{tor}\leq \lvert\mathfrak{l}\rvert \). In par­tic­u­lar, if after re­pla­cing \( \phi \) by an iso­morph­ic Drin­feld mod­ule it ac­quires good re­duc­tion at a prime of de­gree \( \leq 2 \), then Con­jec­ture TDM holds for \( \phi \).

(3)  It seems like an in­ter­est­ing and im­port­ant prob­lem to give an ex­pli­cit con­jec­tur­al clas­si­fic­a­tion of pos­sible \( F \)-ra­tion­al tor­sion sub­mod­ules of Drin­feld mod­ules of rank 3, pos­sibly us­ing the geo­metry of Drin­feld mod­u­lar sur­faces as a guide.

Fol­low­ing Ogg’s philo­sophy on the ex­ist­ence of ra­tion­al points on mod­u­lar curves, we may also for­mu­late the fol­low­ing con­jec­ture which is an ana­logue of TEC\( ^+ \). It is sup­por­ted by the fact that the curve \( X_0(\mathfrak{p}) \) has nonzero genus if and only if \( \deg \mathfrak{p} \geq 3 \) (and this genus is then at least 2), and by the list of all ideals \( \mathfrak{n} \) of \( A \) with \( \deg \mathfrak{n}\geq 3 \) such that \( Y_0(\mathfrak{n}) \) has a \( F \)-ra­tion­al CM point; see [e58].

Con­jec­ture TDM\( ^+ \): Let \( C = 3 \) if \( q\neq 3 \), and \( C = 4 \) if \( q = 3 \). If \( \mathfrak{p} \) is a prime of \( A \) of de­gree \( \geq C \), there does not ex­ist a Drin­feld \( A \)-mod­ule \( \phi \) of rank 2 over \( F \) with an \( F \)-ra­tion­al \( A \)-sub­mod­ule of \( \phi[\mathfrak{p}] \) iso­morph­ic to \( A/\mathfrak{p} \). Equi­val­ently, if \( \mathfrak{p} \) is a prime of \( A \) of de­gree \( \geq C \), the Drin­feld mod­u­lar curve \( Y_0(\mathfrak{p}) \) has no \( F \)-ra­tion­al points.

There is also an ana­logue of Con­jec­ture UBC for Drin­feld mod­ules of ar­bit­rary rank \( r \). If \( r\geq 2 \), it is re­min­is­cent of the uni­form bounded­ness con­jec­ture for the tor­sion of abeli­an vari­et­ies over a num­ber field.

Con­jec­ture UBC-DM: ([e51], Conjecture 2) Fix \( q \), as well as \( r\geq 1 \) and \( d\geq 1 \). There is a uni­form bound on \( \#(^\phi L)_\text{tors} \) as \( L \) ranges over the ex­ten­sions of \( F \) of de­gree \( \leq d \), and \( \phi \) ranges over rank-\( r \) Drin­feld \( A \)-mod­ules over \( L \). In the rank \( r = 2 \) case, this is equi­val­ent to say­ing that there ex­ists a con­stant \( C > 0 \) such that if \( \mathfrak{n} \) is an ideal of \( A \) with \( \deg\mathfrak{n} \geq C \) and \( L \) is an ex­ten­sion of \( F \) of de­gree \( \leq d \), then the Drin­feld mod­u­lar curve \( Y_1(\mathfrak{n}) \) has no \( L \)-ra­tion­al points.
Re­mark 5.2: (1)  The dependence of the universal bound on the rank \( r \) in Conjecture UBC-DM cannot be avoided. Indeed, let \( V\subset F \) be an arbitrary \( r \)-dimensional \( \mathbb{F}_q \)-vector subspace of \( F \), and let \[ \phi_T(x)=Tx\prod_{0\neq v\in V}(1-x/v). \] Since \( \phi_T(x) \) is an \( \mathbb{F}_q \)-linear polynomial of degree \( q^r \), it defines a Drinfeld module over \( F \) of rank \( r \). On the other hand, by construction, \( \phi[T]\cong (A/TA)^r \) is rational over \( F \).

(2)  This con­jec­ture in [e51] is stated for more gen­er­al rings \( A \), namely those of reg­u­lar func­tions on the af­fine curve ob­tained by re­mov­ing a closed point from a nonsin­gu­lar pro­ject­ive curve over \( \mathbb{F}_q \). This more gen­er­al con­jec­ture re­duces to UBC-DM for \( \mathbb{F}_q[T] \) since any Drin­feld \( A \)-mod­ule can be con­sidered as a Drin­feld \( \mathbb{F}_q[T] \)-mod­ule of high­er rank. An earli­er for­mu­la­tion of the con­jec­ture can be found in Problème 3 of [e44].

(3)  For rank-1 Drin­feld mod­ules, Poon­en [e51] gave sev­er­al proofs of this con­jec­ture, in­clud­ing an ex­pli­cit bound on the size of the tor­sion. (For ex­ample, \( \# ({^\phi}F)_\mathrm{tor}\leq q \) if \( \phi \) has rank 1 and \( q\neq 2 \).) See also [e59] and [e58] for oth­er proofs or im­prove­ments in this case. A fact es­sen­tial for the proofs is that any rank-1 Drin­feld mod­ule defined over \( L \) be­comes iso­morph­ic to the Carl­itz mod­ule over an ex­ten­sion of \( L \) of de­gree \( \leq q-1 \).

(4)  It is pos­sible to re­for­mu­late Con­jec­tures TDM and UBC-DM in an equi­val­ent, ele­ment­ary form, which makes no ref­er­ence to Drin­feld mod­ules; see [e52], [e63]. Us­ing ter­min­o­logy from arith­met­ic dy­nam­ics, we say that \( \alpha\in L \) is a preperi­od­ic point for the poly­no­mi­al \( f(x)\in L[x] \) if the set \[ \{\alpha, f(\alpha), f(f(\alpha)), f(f(f(\alpha))),\dots\} \] is fi­nite. Note that for a Drin­feld mod­ule \( \phi \) over \( L \), the set of preperi­od­ic points of \( \phi_T(x) \) is ex­actly \( ({^\phi}L)_\mathrm{tor} \). Thus, Con­jec­ture TDM can be stated as say­ing that for any \( f(x)=Tx+gx^q+\Delta x^{q^2}\in F[x] \), with \( \Delta\neq 0 \), the set of preperi­od­ic points in \( F \) has car­din­al­ity at most \( q^2 \). Also, In­gram [e76] has pro­posed a per­spect­ive on Con­jec­ture UBC-DM via the ad­el­ic filled Ju­lia set at­tached to the Drin­feld mod­ule when viewed as a dy­nam­ic­al sys­tem, with ap­plic­a­tions to cer­tain fam­il­ies of rank-\( r \) Drin­feld mod­ules.

(5)  A ver­sion of Con­jec­ture UBC for el­lipt­ic curves over func­tion fields is re­l­at­ively easy to prove, and was already done by Lev­in [e4] in 1968. Sim­il­arly, let \( L \) be a field of tran­scend­ence de­gree 1 over \( F^\mathrm{alg} \), and let \( \phi \) be a Drin­feld mod­ule of rank 2 over \( L \) such that \( j(\phi)\not\in F^\mathrm{alg} \). Sch­weizer proved in [e62] that \[ \# ({^\phi}L)_\mathrm{tor}\leq (\gamma^2(q^2+1)(q+1))^{q/(q-1)}, \] where \( \gamma \) is the gon­al­ity of \( L \).

In rank 2, Con­jec­ture UBC-DM is cur­rently open in gen­er­al but pro­gress has been made. Re­call that a first and im­port­ant step for the proof of Con­jec­ture UBC was Man­in’s res­ult [e5], namely a uni­form bound on the \( p \)-primary tor­sion of el­lipt­ic curves over a num­ber field \( L \), de­pend­ing on \( p \) and \( L \). Its ana­logue for Drin­feld \( A \)-mod­ules of rank 2 has been es­tab­lished by Poon­en [e51]. The fol­low­ing the­or­em of Sch­weizer im­proves this res­ult by mak­ing the con­stant de­pend only on the prime of \( A \) and the de­gree of \( L \) (see also [e84] for a gen­er­al­iz­a­tion to more gen­er­al func­tion fields than \( F \)).

The­or­em 5.3: ([e58], Theorem 2.4) Fix \( q \), as well as \( d\geq 1 \) and a prime \( \mathfrak{p} \) of \( A \). As \( L \) ranges over all ex­ten­sions of \( F \) with \( [L:F] \leq d \) and \( \phi \) ranges over all rank-2 Drin­feld \( A \)-mod­ules over \( L \), there is a uni­form bound on the size of the \( \mathfrak{p} \)-primary part of \( (^\phi L)_\mathrm{tor} \).

As a con­sequence, Con­jec­ture UBC-DM in the rank-2 case is re­duced to the prob­lem of bound­ing uni­formly the num­ber of primes \( \mathfrak{p} \) in the primary de­com­pos­i­tion of the tor­sion, which is equi­val­ent to prov­ing that for any \( d\geq 1 \), there ex­ists a con­stant \( C > 0 \) such that if \( \mathfrak{p} \) is a prime of \( A \) with \( \deg\mathfrak{p} \geq C \) and \( L \) is an ex­ten­sion of \( F \) of de­gree \( \leq d \), then \( Y_1(\mathfrak{p}) \) has no \( L \)-ra­tion­al points.

Re­cently and for ar­bit­rary rank, Ishii proved a ver­sion of Man­in’s uni­form­ity res­ult on the \( \mathfrak{p} \)-primary tor­sion for one-di­men­sion­al fam­il­ies of Drin­feld mod­ules over a fi­nitely gen­er­ated ex­ten­sion of \( F \), which is the ana­logue of a res­ult of Cadoret and Tamagawa [e77] for 1-di­men­sion­al fam­il­ies of abeli­an vari­et­ies.

The­or­em 5.4: ([e113], Theorem 1.1) Let \( \mathfrak{p} \) be a prime of \( A \), \( L \) a fi­nitely gen­er­ated ex­ten­sion of \( F \), \( S \) a one-di­men­sion­al scheme which is of fi­nite type over \( L \), and \( \phi \) a Drin­feld \( A \)-mod­ule over \( S \). Then there ex­ists an in­teger \( N := N(\phi, S, L, \mathfrak{p})\geq 0\, \) such that \( \,\phi_s[\mathfrak{p}^\infty](L)\subset \phi_s[\mathfrak{p}^N](L)\, \) for every \( s\in S(L) \).

Be­sides this res­ult, not much is cur­rently known to­wards Con­jec­ture UBC-DM in ranks \( r\geq 3 \).

We make a few com­ments about the ana­logue of Con­jec­ture SUQ. In a series of pa­pers cul­min­at­ing in [e71], Richard Pink and his col­lab­or­at­ors proved the fol­low­ing ana­logue of Serre’s Open Im­age The­or­em.

The­or­em 5.5: (Open Image Theorem) Let \( \phi \) be a Drin­feld \( A \)-mod­ule of rank \( r \) over a fi­nite ex­ten­sion \( L \) of \( F \). As­sume \( \operatorname{End}_{L^{\operatorname{sep}}}(\phi)=A \). Then there is a con­stant \( N(\phi, L) \) de­pend­ing only of \( \phi \) and \( L \) such that \[ [\mathrm{GL}_r(A/\mathfrak{n}):\rho_{\phi, \mathfrak{n}}(G_L)]\leq N(\phi, L) \quad \text{for all nonzero}\, \mathfrak{n}\lhd A. \]

The ana­logue of Con­jec­ture SUQ is the state­ment that the bound \( N(\phi, L) \) in The­or­em 5.5 can be made uni­form, i.e., in­de­pend­ent of \( \phi \). For \( r=1 \) this not hard to prove, but it re­mains a ma­jor open ques­tion for \( r\geq 2 \).

Re­mark 5.6: (1)  Let \( \phi \) be the rank-2 Drinfeld module over \( F \) defined by \[ \phi_T=Tx+x^q-\mathfrak{p} x^{q^2}. \] Using the Weil pairing for Drinfeld modules, it is possible to show that \( \rho_{\phi, \mathfrak{p}} \) is not surjective if \( \mathfrak{p} \) is a prime of odd degree and \( q\equiv 1(\operatorname{mod} 4) \), so the direct analogue of Conjecture SUQ does not hold; see ([e103], Section 4.2).

(2)  In [e103], ex­tend­ing a con­struc­tion of Zy­wina for \( r=2 \), Chen proved that for the rank-\( r \) Drin­feld mod­ule over \( F \) defined by \[ \phi_T=Tx+x^{q^{r-1}}+T^{q-1}x^{q^r} \] the rep­res­ent­a­tions \( \rho_{\phi, \mathfrak{n}}\colon G_F\to \mathrm{GL}_r(A/\mathfrak{n}) \) are sur­ject­ive for all \( \mathfrak{n}\neq 0 \), as­sum­ing \( r \) is prime, \( q\equiv 1(\operatorname{mod} r) \), and the char­ac­ter­ist­ic of \( F \) is suf­fi­ciently large com­pared to \( r \).

We now turn to known res­ults on Con­jec­tures TDM and TDM\( ^+ \), some of which have con­sequences for Con­jec­ture UBC-DM. In 2010, Pál made ma­jor pro­gress by prov­ing Con­jec­ture TDM\( ^+ \) for the case \( q=2 \), i.e., the field \( F = \mathbb{F}_2(T) \).

The­or­em 5.7: ([e72], Theorem 1.2) As­sume that \( q=2 \). If \( \mathfrak{p} \) is any prime of \( A \) with \( \deg \mathfrak{p}\geq 3 \) then \( Y_0(\mathfrak{p}) \) has no \( F \)-ra­tion­al points.

The pre­vi­ous the­or­em im­plies that Con­jec­ture TDM and Con­jec­ture UBC-DM hold for \( q = 2 \) and \( L = F = \mathbb{F}_2(T) \); see The­or­ems 1.4 and 1.6 in [e72]. In a dif­fer­ent dir­ec­tion and more re­cently, Ishii made more pro­gress to­wards Con­jec­ture TDM\( ^+ \) for ar­bit­rary \( q \) and levels of small de­gree.

The­or­em 5.8: ([e102], Theorem 0.4) Let \( \mathfrak{p} \) be a prime of \( A \) of de­gree 4. Then \( Y_0(\mathfrak{p}) \) has no \( F \)-ra­tion­al points.
Re­mark 5.9: Even though Conjecture TDM\( ^+ \) is not formulated for composite levels, we mention additional results of Schweizer for small degree: the curve \( Y_0(\mathfrak{n}) \) has no \( F \)-rational points when
  1. \( \mathfrak{n} \in \{T(T^2 + T + 1), T^3, T^2(T + 1)\} \) in \( \mathbb{F}_2[T] \),
  2. \( \mathfrak{n} \in \{T(T - 1)(T + 1), T^2(T - 1)\} \) in \( \mathbb{F}_3[T] \), and
  3. when \( \mathfrak{n} \) is the product of three dis­tinct lin­ear factors in \( \mathbb{F}_4[T] \); see [e62], [e74].

For Con­jec­ture TDM, Sch­weizer proved that we al­ways have \( \deg \mathfrak{m} \leq 1 \) in \( (^\phi F)_\mathrm{tor}\cong A/\mathfrak{m}\oplus A/\mathfrak{n} \); see Pro­pos­i­tion 4.4 in [e58]. Moreover since any \( F \)-ra­tion­al point on \( Y_1(\mathfrak{p}) \) nat­ur­ally provides an \( F \)-ra­tion­al point on \( Y_0(\mathfrak{p}) \), The­or­ems 5.7 and 5.8 re­main val­id for the curve \( Y_1(\mathfrak{p}) \).

Re­mark 5.10: If \( q=2 \), it is also known that the curve \( Y_1((T^2+T+1)^2) \) with composite level has no \( F \)-rational points; see ([e72], Theorem 10.8).

For fi­nite ex­ten­sions of \( F \), Ar­mana has ob­tained par­tial or con­di­tion­al res­ults to­wards Con­jec­tures TDM and UBC-DM. The first one fo­cuses on levels of small de­gree.

The­or­em 5.11: ([e78], Theorem 7.5) Let \( \mathfrak{p} \) be a prime of \( A \).
  1. If \( \mathfrak{p} \) is of de­gree 3, the curve \( Y_1(\mathfrak{p}) \) has no \( L \)-ra­tion­al points for any ex­ten­sion \( L/F \) of de­gree \( \leq 2 \).
  2. Sup­pose \( \mathfrak{p} \) is of de­gree 4. Let \( d = 1 \) if \( q < 5 \), \( d = 2 \) if \( q = 5 \), and \( d = 3 \) if \( q \geq 7 \). Then the curve \( Y_1(\mathfrak{p}) \) has no \( L \)-ra­tion­al points for any ex­ten­sion \( L/F \) of de­gree \( \leq d \).
  3. Let \( \mathfrak{p} \) be a prime of \( A \) such that for any nor­mal­ized Hecke ei­gen­form \( f \in \mathcal{H}_{0}(\mathfrak{p},\mathbb{C}) \), we have \( \operatorname{ord}_{s=1}L(f,s) \leq 1 \). Then the curve \( Y_1(\mathfrak{p}) \) has no \( L \)-ra­tion­al points for any ex­ten­sion \( L/F \) of de­gree \( < \min (\deg \mathfrak{p}, |\mathfrak{p}|/(2(q^2+1)(q+1))) \).

Giv­en a prime \( \mathfrak{p} \), the as­sump­tion in (iii) on the van­ish­ing or­der of \( L \)-series at the cent­ral value is not al­ways sat­is­fied. However, if one be­lieves in the philo­sophy that el­lipt­ic curves over \( \mathbb{Q} \) with rank great­er than 1 are ex­pec­ted to be rare and in its coun­ter­part for el­lipt­ic curves over \( F \), we should get many ex­amples of pairs \( (q,\mathfrak{p} \)) to which state­ment (iii) ap­plies. For in­stance it ap­plies to curves \( Y_1(\mathfrak{p}) \) for all prime levels \( \mathfrak{p} \) of de­gree 5 in \( \mathbb{F}_2[T] \) with \( L=F \), and for \( \mathfrak{p}=T^5-T^4-T^2-1 \in \mathbb{F}_3[T] \) with \( [L:F]\leq 3 \). There is also a sim­il­ar res­ult for \( F \)-ra­tion­al points on \( Y_0(\mathfrak{p}) \) when \( q\geq 5 \), un­der the same hy­po­thes­is as in (iii); see ([e78], The­or­em 7.8).

The second res­ult is con­di­tion­al but with a more gen­er­al con­clu­sion to­wards Con­jec­ture TDM. To state it we need some nota­tion. Sim­il­arly to clas­sic­al mod­u­lar forms, it is pos­sible to de­vel­op a the­ory of al­geb­ra­ic Drin­feld mod­u­lar forms of weight 2 us­ing sec­tions of the sheaf of re­l­at­ive dif­fer­en­tials on \( X_0(\mathfrak{p}) \). Let \( \mathcal{M}_{\mathfrak{p}} \) be the \( A[1/\mathfrak{p}] \)-mod­ule of doubly cuspid­al al­geb­ra­ic Drin­feld mod­u­lar forms of weight 2 and type 1 for \( \Gamma_0(\mathfrak{p}) \) and let \( \mathbb{T}_{\mathfrak{p}} \) be the Hecke al­gebra gen­er­ated by all Hecke op­er­at­ors act­ing on \( \mathcal{M}_{\mathfrak{p}} \). For a prime \( \mathfrak{l} \) of \( A \), let \( \mathbb{F}_\mathfrak{l} = A/\mathfrak{l} \) and \( \mathcal{M}_{\mathfrak{p}}(\mathbb{F}_\mathfrak{l}) = \mathcal{M}_\mathfrak{p}\otimes_{A[1/\mathfrak{p}]} \mathbb{F}_\mathfrak{l} \). By ex­tend­ing the res­ults of Sec­tion 4.1, any \( f \in \mathcal{M}_{\mathfrak{p}}(\mathbb{F}_\mathfrak{l}) \) has an ex­pan­sion at the cusp \( [\infty] \) of the form \( \sum_{i\geq 0} b_i(f) t^{1+i(q-1)} \) with coef­fi­cients \( b_i(f) \in \mathbb{F}_\mathfrak{l} \). In the Hecke al­gebra \( \mathbb{T}_{\mathfrak{p}} \), let \( I_e \) (resp. \( \widetilde{I_e} \)) be the the ideal which an­ni­hil­ates the ana­logue of the wind­ing ele­ment (resp. the wind­ing ele­ment mod­ulo the char­ac­ter­ist­ic of \( F \)); see [e78] for the defin­i­tions. Let \( \mathcal{M}_{\mathfrak{p}}(\mathbb{F}_\mathfrak{l})[\widetilde{I_e}] \) be the sub­space of \( \mathcal{M}_{\mathfrak{p}}(\mathbb{F}_\mathfrak{l}) \) an­ni­hil­ated by \( \widetilde{I_e} \).

The­or­em 5.12: ([e78], Theorem 1.5) Let \( \mathfrak{p} \) be a prime of de­gree \( \geq 3 \). Sup­pose there ex­ist:
  1. a sat­ur­ated ideal \( I \) of \( \mathbb{T}_{\mathfrak{p}} \), with an­ni­hil­at­or de­noted \( \hat{I} \), sat­is­fy­ing \( I_e \subset I \subset \tilde{I_e} \) and \( \hat{I} + \tilde{I_e} = \mathbb{T}_{\mathfrak{p}} \);
  2. a prime \( \mathfrak{l} \) of \( A \) of de­gree 1 such that the \( \mathbb{F}_\mathfrak{l} \)-lin­ear map \[ (\mathbb{T}_{\mathfrak{p}}/\tilde{I_e}) \otimes_{\mathbb{Z}} \mathbb{F}_\mathfrak{l} \longrightarrow \operatorname{Hom}(\mathcal{M}_{\mathfrak{p}}(\mathbb{F}_\mathfrak{l})[\tilde{I_e}], \mathbb{F}_\mathfrak{l}), \] which maps \( u \in \mathbb{T}_{\mathfrak{p}} \) to the lin­ear form \( f \mapsto b_1(f|u) \), is an iso­morph­ism.

Then:

  1. If \( \deg \mathfrak{p} \geq \max(q + 1, 5) \), the curve \( Y_1(\mathfrak{p}) \) has no \( L \)-ra­tion­al points for any ex­ten­sion \( L/F \) of de­gree less than or equal to \( q \).
  2. Con­jec­ture TDM is true for the ideal \( \mathfrak{p} \), namely \( Y_1(\mathfrak{p}) \) has no \( F \)-ra­tion­al points.

The as­sump­tion in (i) on the ex­ist­ence of the ideal \( I \) is likely of a tech­nic­al nature. The as­sump­tion in (ii) is deep­er: it is re­lated to com­plic­a­tions that arise when ad­apt­ing Mazur’s form­al im­mer­sion ar­gu­ment to the Drin­feld set­ting. This will be dis­cussed in Sec­tion 5.2.

5.2. Outline of the proofs
We re­view the main ideas be­hind the proofs of The­or­ems 5.3, 5.7, 5.8, 5.11 and 5.12, which are mostly in­spired by the work of Man­in and Mazur, with a fo­cus on the dif­fer­ences from the clas­sic­al set­ting and com­plic­a­tions that arise.

Sch­weizer’s uni­form bound for the \( \mathfrak{p} \)-primary tor­sion, The­or­em 5.3, is an ana­logue of a the­or­em of Kami­enny and Mazur [e36], it­self a stronger ver­sion of Man­in’s the­or­em [e5] for the \( p \)-primary part of the tor­sion of el­lipt­ic curves over \( \mathbb{Q} \). The proof of Sch­weizer, after Poon­en [e51], fol­lows the same ap­proach. The key in­gredi­ent es­sen­tially states that for any curve \( C \) over a glob­al field \( K \), the \( d \)-fold sym­met­ric power \( C^{(d)} \) has only fi­nitely many \( K \)-ra­tion­al points if \( C \) does not ad­mit a \( K \)-ra­tion­al cov­er­ing of \( \mathbb{P}^1_{K} \) of de­gree \( \leq 2d \). When \( K \) is a num­ber field, this is a the­or­em of Frey which de­rives from Falt­ings’ the­or­em on ra­tion­al points of sub­vari­et­ies of abeli­an vari­et­ies. When \( K \) is a func­tion field, Sch­weizer has ob­tained a sim­il­ar cri­terion from the Mor­dell–Lang con­jec­ture for abeli­an vari­et­ies over func­tion fields, proved by Hrushovski [e46]. He ap­plied it to the Drin­feld mod­u­lar curves \( X_0(\mathfrak{p}^e) \) for primes \( \mathfrak{p} \) to ob­tain The­or­em 5.3.

To dis­cuss the proofs of The­or­ems 5.7, 5.8, 5.11 and 5.12, we need to re­call Mazur’s ap­proach from [e14] for the study of \( \mathbb{Q} \)-ra­tion­al points on the clas­sic­al mod­u­lar curve \( X_0(p) \), \( p \) prime (see also [e106] in this volume and the over­view [e39] for ad­di­tion­al de­tails). Hand­ling points defined over a num­ber field \( K \) of de­gree \( d \) re­quires Kami­enny’s gen­er­al­ized ap­proach us­ing the \( d \)-fold sym­met­ric power of \( X_0(p) \). For sim­pli­city we fo­cus here only on \( d=1 \).

Let \( P=(E,C_p) \in X_0(p)(\mathbb{Q}) \) be a non­cuspid­al point for a prime num­ber \( p \). As­sume that the genus of \( X_0(p) \) is pos­it­ive. The ob­ject­ive of Mazur’s form­al im­mer­sion ar­gu­ment is to prove that the el­lipt­ic curve \( E \) has po­ten­tially good re­duc­tion at all primes \( \ell\neq \{2, p\} \); see Co­rol­lary 4.4 in [e14]. Con­sider the Abel–Jac­obi map \[ \begin{array}{rcl} X_0(p) & \longrightarrow & J_0(p) \\ Q & \longmapsto & (Q)-(\infty)\end{array} \] where \( J_0(p) \) is the Jac­obi­an vari­ety of \( X_0(p) \). The main in­gredi­ent is an op­tim­al quo­tient \( A \) of \( J_0(p) \) defined over \( \mathbb{Q} \) which sat­is­fies the fol­low­ing two prop­er­ties:

  1. The Mor­dell–Weil group \( A(\mathbb{Q}) \) is fi­nite.
  2. Let \( \varphi:X_0(p) \to J_0(p) \to A \), which ex­tends over \( \mathbb{Z} \) to \( \varphi:X_0(p)_{\mathrm{sm}} \to \mathcal{A} \), where \( X_0(p)_{\mathrm{sm}} \) is the largest open sub­set of \( X_0(p) \) smooth over \( \mathbb{Z} \) and \( \mathcal{A} \) is the Néron mod­el of \( A \) over \( \mathbb{Z} \). When look­ing at the fibers at any prime \( \ell \) such that \( \ell \nmid 2p \), the map \( \varphi_{\ell} : X_0(p)_{\mathbb{F}_\ell} \to \mathcal{A}_{\mathbb{F}_\ell} \) is a form­al im­mer­sion along the cuspid­al sec­tion \( [\infty] \).

Op­tim­al quo­tients sat­is­fy­ing (1) ex­ist: the idea is to con­struct them with the prop­erty that their Hasse–Weil \( L \)-func­tion does not van­ish at the cent­ral value and de­duce that their Mor­dell–Weil group is fi­nite, in the spir­it of the Birch and Swin­ner­ton-Dyer con­jec­ture. For \( A \), one may take Mazur’s Ei­s­en­stein quo­tient [e13] or Mer­el’s wind­ing quo­tient [e45] (the lat­ter one be­ing the largest quo­tient of \( J_0(p) \) sat­is­fy­ing this prop­erty).

Con­di­tion (2) is equi­val­ent to the sur­jectiv­ity of the in­duced map \( \varphi_{\ell}^*:\mathrm{Cot}_0(A_{\mathbb{F}_\mathfrak{l}}) \to \mathrm{Cot}_\infty(X_0(p)_{\mathbb{F}_\ell}) \) on the co­tan­gent spaces, which in this case is equi­val­ent to \( \varphi_{\ell}^* \) be­ing nonzero. It is in­struct­ive to re­for­mu­late this con­di­tion for the Abel–Jac­obi map \( X_0(p)_{\mathrm{sm}} \to \mathcal{J}_0(p) \) over \( \mathbb{Z} \) in terms of \( q \)-ex­pan­sions of cusps forms: on the co­tan­gent spaces at \( [\infty] \), it is simply the map \[ \begin{array}{ccc} S_2(\Gamma_0(p),\mathbb{Z}) & \longrightarrow & \mathbb{Z} \\ \sum_{n\geq 1} a_n q^n & \longmapsto & a_1 \end{array} \] where \( S_2(\Gamma_0(p),\mathbb{Z}) \) is the space of cusp forms of weight 2 with coef­fi­cients in \( \mathbb{Z} \). This map is nonzero, as a con­sequence of the clas­sic­al for­mula \begin{equation}\label{eq-a1Tnf} a_n(f)=a_1(f|T_n),\quad \text{for all } n\geq 1, \tag{5.1} \end{equation} for the ac­tion of the Hecke op­er­at­ors \( (T_n)_{n\geq 1} \) on cusp forms. For the map \( \varphi_{\ell}^* \), there is a sim­il­ar de­scrip­tion in­volving cusp forms an­ni­hil­ated by the Ei­s­en­stein ideal, resp. the wind­ing ideal of the Hecke al­gebra. One then uses a Hecke ei­gen­vector in \( \mathrm{Cot}_0(A_{\mathbb{F}_\mathfrak{l}}) \), which ne­ces­sar­ily has \( a_1\neq 0 \). This en­sures Con­di­tion (2) is sat­is­fied.

Us­ing (1) and (2), Mazur was able to prove that the el­lipt­ic curve \( E \) ne­ces­sar­ily has po­ten­tially good re­duc­tion at all primes \( \ell \neq \{2,p\} \) as fol­lows. If \( E \) has po­ten­tially mul­ti­plic­at­ive re­duc­tion at \( \ell \), the point \( P \) will spe­cial­ize on \( X_0(p) \) at \( \ell \) to one of the cusps. By ap­ply­ing the Atkin–Lehner in­vol­u­tion \( w_p \), we may as­sume that it spe­cial­izes to \( [\infty] \), there­fore the point \( \varphi(P) \) spe­cial­izes to 0 at \( \ell \). By the form­al im­mer­sion prop­erty (2) for \( \varphi \) at \( \ell \), we also have \( \varphi(P)\neq 0 \). By (1) we also know that \( \varphi(P) \) is a tor­sion point. This con­tra­dicts a spe­cial­iz­a­tion lemma for points of fi­nite or­der in group schemes.

The last part of Mazur’s proofs of Con­jec­tures TEC and TEC\( ^+ \) is to dis­card the cases where \( E \) does not have po­ten­tially good re­duc­tion at \( \ell \) when \( p \) is large enough. If the point \( P \) on \( X_0(p) \) comes from a tor­sion point of or­der \( p \) on \( E \), this fol­lows from known bounds for the or­der of the spe­cial­iz­a­tion of a tor­sion point at a prime \( \ell \) with \( p \neq \ell \), when \( E \) has good or ad­dit­ive re­duc­tion at \( \ell \). If \( P \) comes from a ra­tion­al cyc­lic sub­group \( C_p \) of or­der \( p \), Mazur stud­ies the iso­geny char­ac­ter, i.e., the rep­res­ent­a­tion \( \operatorname{Gal}(\mathbb{Q}^\mathrm{alg}/\mathbb{Q}) \to \mathrm{GL}_1(\mathbb{Z}/p\mathbb{Z}) \) com­ing from the nat­ur­al Galois ac­tion on \( C_p \). Con­straints com­ing from the ex­ist­ence of this iso­geny char­ac­ter with the Riemann hy­po­thes­is for the re­duc­tion of \( E \) mod­ulo \( \ell \notin\{ 2,p \} \) ul­ti­mately provide a fi­nite list of pos­sible val­ues for \( p \).

In this strategy, most of the steps ad­apt more or less eas­ily to the Drin­feld mod­u­lar curve \( X_0(\mathfrak{p}) \). As an op­tim­al quo­tient \( A \) of the Jac­obi­an \( J_0(\mathfrak{p}) \), one may take the Ei­s­en­stein quo­tient in­tro­duced by Tamagawa us­ing the Ei­s­en­stein ideal \( \mathfrak{E}(\mathfrak{p}) \), see Defin­i­tion 4.8. An­oth­er pos­sib­il­ity is the wind­ing quo­tient of [e72] or [e78]. In both cases their Mor­dell–Weil groups are fi­nite, by Schneider’s in­equal­ity in the Birch and Swin­ner­ton-Dyer con­jec­ture for abeli­an vari­et­ies over func­tion fields [e20], so there is an ana­logue of con­di­tion (1). The main obstacle, however, resides in the func­tion field coun­ter­part of the form­al im­mer­sion prop­erty in con­di­tion (2). In­deed, the ac­tion of Hecke op­er­at­ors on the ex­pan­sion of Drin­feld mod­u­lar forms is not well un­der­stood and a for­mula sim­il­ar to \eqref{eq-a1Tnf} is lack­ing (on this top­ic see also [e75]). This is the reas­on why The­or­ems 5.7, 5.8, 5.11 and 5.12 do not provide com­plete an­swers to the ana­logues of Ogg’s con­jec­tures. This also ex­plains as­sump­tion (ii) in The­or­em 5.12: if one re­moves the ideal \( \widetilde{I_e} \) and the prime place \( \mathfrak{l} \) from its for­mu­la­tion, the as­sump­tion be­comes equi­val­ent to the per­fect­ness of the \( \mathbb{C}_\infty \)-pair­ing between the space of Drin­feld mod­u­lar forms \( M_{2,1}^{0,0}(\mathfrak{n}) \) and its Hecke al­gebra, defined by \( (f,u) \mapsto b_1(f|u) \). If true, it would im­ply a “mul­ti­pli­city one” res­ult in \( M_{2,1}^{0,0}(\mathfrak{n}) \), which at the mo­ment is open in gen­er­al. (It is known to fail for oth­er con­gru­ence sub­groups; see Böckle ([e56], Ex­ample 15.4) for \( \Gamma_1(T) \) and Hat­tori ([e104], The­or­em 1) for \( \Gamma_1(T^n) \). These are strong in­dic­a­tions that mul­ti­pli­city one could fail for \( \Gamma_0(\mathfrak{n}) \).) Moreover Ar­mana has re­cently es­tab­lished that this pair­ing is not per­fect in a quite gen­er­al case, namely when \( \mathfrak{n} \) is prime of de­gree \( \geq 5 \); see [e111]. Al­though it is not enough to show that the ver­sion of this per­fect­ness stated in as­sump­tion (ii) is not sat­is­fied, it con­firms the sever­ity of the ob­struc­tion that arises when ad­apt­ing Mazur’s meth­od.

Be­cause of this ob­struc­tion, the un­con­di­tion­al state­ments that we cur­rently know, such as The­or­ems 5.7, 5.8 and 5.11, em­ploy work­arounds or vari­ants of the form­al im­mer­sion prop­erty at \( \ell \). They are es­sen­tial of two types:

  • Fol­low­ing an idea of Mer­el and Par­ent for clas­sic­al mod­u­lar curves, Pál es­tab­lishes a vari­ant of the form­al im­mer­sion prop­erty at the place \( \infty \) of \( F \), in­stead of a fi­nite place \( \mathfrak{l} \) of \( A \); see Pro­pos­i­tion 7.14 in [e72]. To prove this vari­ant, he con­structs a reg­u­lar mod­el of the curve \( X_0(\mathfrak{p}) \) at \( \infty \), uses the in­cid­ence graph of the fiber of this mod­el and har­mon­ic co­chains on this graph. The study of iso­geny char­ac­ters is re­placed with an ad hoc ar­gu­ment. Pál’s strategy uses ex­tens­ively the hy­po­thes­is \( q=2 \) and ul­ti­mately provides The­or­em 5.7.
  • If \( \mathfrak{p} \) has de­gree 3 or 4, or if the prime \( \mathfrak{p} \) is such that for any nor­mal­ized Hecke ei­gen­form \( f \in \mathcal{H}_{0}(\mathfrak{p},\mathbb{C}) \), we have \( \operatorname{ord}_{s=1}L(f,s) \leq 1 \), the situ­ation be­comes sim­pler. The wind­ing quo­tient is then \( J_0(\mathfrak{p}) \) or iso­gen­ous to \( J_0(\mathfrak{p})/(1+w_\mathfrak{p}) J_0(\mathfrak{p}) \) where \( w_\mathfrak{p} \) de­notes the Atkin–Lehner in­vol­u­tion. In this situ­ation, Ar­mana is able to avoid the form­al im­mer­sion prop­erty thanks to the fact that the cusp \( [\infty] \) is not a Wei­er­strass point on \( X_0(\mathfrak{p}) \) and to a lower bound of Sch­weizer on the gon­al­ity of \( X_0(\mathfrak{p}) \); see ([e78], Pro­pos­i­tion 7.6). These ar­gu­ments lead to The­or­em 5.11. Ishii’s work [e102], which is a gen­er­al study of iso­geny char­ac­ters com­ing from rank-2 Drin­feld mod­ules as in Mazur [e14], is also based on this work­around. Us­ing con­gru­ences com­ing from these char­ac­ters, he ob­tains sev­er­al fam­il­ies of con­di­tions which en­sure that \( Y_0(\mathfrak{p}) \) has no \( F \)-ra­tion­al points; see ([e102], The­or­ems 0.1, 0.2, 0.3). The con­di­tions he com­bines are of three dif­fer­ent types: on \( q \), on the prime \( \mathfrak{p} \) and on the set of fi­nite places where the Drin­feld mod­ule has po­ten­tially good re­duc­tion. Here again, it is not pos­sible to reach the full po­ten­tial of Mazur’s ap­proach be­cause the form­al im­mer­sion ar­gu­ment is miss­ing in gen­er­al. Ishii ob­tains The­or­em 5.8 for ra­tion­al points on \( Y_0(\mathfrak{p}) \) if \( \mathfrak{p} \) has de­gree 4 by com­bin­ing his work on iso­geny char­ac­ters and Ar­mana’s re­place­ment for the form­al im­mer­sion prop­erty in this case.

6. Torsion of the Jacobian of \( X_0(\mathfrak{n}) \)

6.1. Cuspidal divisor group
For a nonzero ideal \( \mathfrak{n}\lhd A \), let \( X_0(\mathfrak{n}) \) de­note the Drin­feld mod­u­lar curve in­tro­duced in Sec­tion 3. Its set of cusps is \( X_0(\mathfrak{n})(\mathbb{C}_\infty)-Y_0(\mathfrak{n})(\mathbb{C}_\infty) \). Let \( J_0(\mathfrak{n}) \) be the Jac­obi­an vari­ety of \( X_0(\mathfrak{n}) \). Sim­il­ar to \( \mathcal{C}_N \) ap­pear­ing in Con­jec­ture CJ-N, one defines the cuspid­al di­visor group \( \mathcal{C}_{\mathfrak{n}} \) as the sub­group of \( J_0(\mathfrak{n}) \) gen­er­ated by the di­visor classes \( [c]-[c^{\prime}] \) of dif­fer­ences of all cusps.

One in­ter­est­ing fea­ture of the the­ory of Drin­feld mod­ules is that Drin­feld mod­ules of rank \( r\geq 3 \) do not really have clas­sic­al ana­logues (Drin­feld mod­ules of rank 1 are sim­il­ar to the mul­ti­plic­at­ive group of a field and Drin­feld mod­ules of rank 2 are sim­il­ar to el­lipt­ic curves). Such Drin­feld mod­ules can be defined and stud­ied us­ing a single equa­tion \( \phi_T \) like el­lipt­ic curves, but the Galois rep­res­ent­a­tions arising from their tor­sion points have im­ages in \( \mathrm{GL}_r \) and their mod­u­lar vari­et­ies are \( (r-1) \)-di­men­sion­al. Hence, Drin­feld mod­ules of rank \( r\geq 3 \) are more com­plic­ated than el­lipt­ic curves but sim­pler than abeli­an vari­et­ies. This presents the in­triguing pos­sib­il­ity of prov­ing res­ults for these Drin­feld mod­ules which are known for el­lipt­ic curves but per­haps un­known or very hard for abeli­an vari­et­ies. One such res­ult is the fi­nite­ness of the cuspid­al di­visor group of their mod­u­lar vari­et­ies.

For \( r\geq 3 \), there are dif­fer­ent pos­sible com­pac­ti­fic­a­tions of \( Y_0^r(\mathfrak{n}) \) with de­sir­able prop­er­ties; see [e28] and [e96]. The Satake com­pac­ti­fic­a­tions of Drin­feld mod­u­lar vari­et­ies were con­struc­ted (at dif­fer­ent levels of gen­er­al­ity and de­tails of proof) by Gekel­er [e27], [e101], Kapran­ov [e28], Pink [e80], and Häberli [e100]. The con­struc­tions by Gekel­er, Häberli, and Kapran­ov are ri­gid-ana­lyt­ic, where­as Pink’s con­struc­tion is al­gebro-geo­met­ric. Häberli also proved that the ana­lyt­ic and al­geb­ra­ic Satake com­pac­ti­fic­a­tions give the same vari­ety. The Satake com­pac­ti­fic­a­tion can be de­scribed ana­lyt­ic­ally as fol­lows. For \[ \boldsymbol{z}=(z_1, \dots, z_r)\in \mathbb{P}^{r-1}(\mathbb{C}_\infty), \] define \[ d(\boldsymbol{z})=\dim_F(Fz_1+\cdots+Fz_r) \quad \text{and}\quad d_\infty(\boldsymbol{z})=\dim_{F_\infty}(F_\infty z_1+\cdots+F_\infty z_r). \] Then \[ 1\leq d_\infty(\boldsymbol{z})\leq d(\boldsymbol{z})\leq r\quad\text{ and }\quad \Omega^r=\{\boldsymbol{z}\mid d_\infty(\boldsymbol{z})=r\}. \] More gen­er­ally, for \( 1\leq i\leq r \), put \[ \Omega^{i, r} = \{\boldsymbol{z}\mid d_\infty(\boldsymbol{z})=d(\boldsymbol{z})=i\} \quad\text{ and }\quad \overline{\Omega^r}=\cup_{1\leq i\leq r} \Omega^{i, r}. \] The space \( \overline{\Omega^r} \) is in­vari­ant un­der the ac­tion of \( \Gamma_0(\mathfrak{n}) \) by lin­ear frac­tion­al trans­form­a­tions. The quo­tient \( X_0^r(\mathfrak{n})(\mathbb{C}_\infty)=\Gamma_0(\mathfrak{n})\setminus \overline{\Omega^r} \) is a pro­ject­ive con­nec­ted nor­mal vari­ety over \( \mathbb{C}_\infty \) of di­men­sion \( r-1 \) con­tain­ing \( Y_0^r(\mathfrak{n}) \) as an open sub­vari­ety, which has a ca­non­ic­al mod­el over \( F \). The cusps of \( X_0^r(\mathfrak{n}) \) are the (geo­met­ric­ally) ir­re­du­cible com­pon­ents of \( X_0^r(\mathfrak{n})(\mathbb{C}_\infty)-Y_0^r(\mathfrak{n})(\mathbb{C}_\infty) \) of di­men­sion \( r-2 \). The cuspid­al di­visor group \( \mathcal{C}^r_\mathfrak{n} \) of \( X^r_0(\mathfrak{n}) \) is the sub­group of the di­visor class group of \( X^r_0(\mathfrak{n}) \) gen­er­ated by the Weil di­visors \( [c]-[c^{\prime}] \), where \( c \) and \( c^{\prime} \) run over the cusps of \( X^r_0(\mathfrak{n}) \). The ana­logue of the res­ult of Man­in and Drin­feld in this set­ting is the fol­low­ing.

The­or­em 6.1: \( \mathcal{C}_{\mathfrak{n}}^r \) is a fi­nite group.

Proof. For \( r=2 \), this was proved by Gekel­er in [e23]. For \( r\geq 3 \) this is a res­ult of Kapran­ov [e28]; see also ([e112], The­or­em 7.8) and ([e117], The­or­em 10.7). The gen­er­al idea of the proof is to con­struct suf­fi­ciently many mod­u­lar units on \( X_0^r(\mathfrak{n}) \), i.e., func­tions whose di­visors are sup­por­ted on the cusps, so that the sub­group of de­gree-0 di­visors of these func­tions has fi­nite in­dex in the group gen­er­ated by all \( [c]-[c^{\prime}] \). A nat­ur­al meth­od for con­struct­ing mod­u­lar units uses the Drin­feld dis­crim­in­ant func­tions \( \Delta_{r}(z_1, \dots, z_r) \). Re­call that \( \Delta_{r} \) does not van­ish on \( \Omega^r \). One eas­ily shows that \( \Delta_{r}(z_1, \dots, z_r)/\Delta_{r}(\mathfrak{m} z_1, \dots, z_r) \) is in­vari­ant un­der \( \Gamma_0^r(\mathfrak{n}) \) for any \( \mathfrak{m}\mid \mathfrak{n} \), and hence that it defines a mod­u­lar unit on \( X_0^r(\mathfrak{n}) \) whose di­visor can be com­puted from the or­der of van­ish­ing of \( \Delta_{r}(\boldsymbol{z}) \) at the cusps. en­d­proof

Com­put­ing the group struc­ture of \( \mathcal{C}_{\mathfrak{n}}^r \) is much harder than prov­ing that this group is fi­nite. For clas­sic­al mod­u­lar curves an im­port­ant tool for solv­ing this prob­lem is a res­ult of Ligoz­at ([e12], Pro­pos­i­tion 3.2.1), which gives ne­ces­sary and suf­fi­cient con­di­tions for a func­tion of the form \( \prod_{m\mid N}\eta(m z)^{s_m} \), \( s_m\in \mathbb{Z} \), to be a mod­u­lar unit on \( X_0(N) \), where \( \eta(z) \) is the Dede­kind eta func­tion. (Re­call that \( \eta(z) \) is the 24th root of the clas­sic­al dis­crim­in­ant func­tion.) However, such a strong res­ult does not seem to hold over func­tion fields; for ex­ample, \( \Delta_r \) only has a \( (q-1) \)-th root in \( \mathcal{O}(\Omega^r)^\times \), but the ana­logue of 24 in this con­text is \( (q-1)(q^2-1) \).

In­stead, one fol­lows a dif­fer­ent strategy, spe­cif­ic to func­tion fields, which was ini­ti­ated by Gekel­er in [e49]. The key res­ult here is the ex­act se­quence (4.3). Giv­en a mod­u­lar unit \( u(\boldsymbol{z}) \) on \( X_0^r(\mathfrak{n}) \), to de­term­ine the max­im­al root that can be ex­trac­ted from \( u(\boldsymbol{z}) \) (and thus to de­term­ine the or­der of the un­der­ly­ing cuspid­al di­visor) one ap­plies \( \mathrm{dlog} \) to \( u(\boldsymbol{z}) \) to get a com­bin­at­or­i­al \( \mathbb{Z} \)-val­ued func­tion \( \mathrm{dlog}(u)\in \operatorname{Har}^1(\mathscr{B}^r, \mathbb{Z}) \). If one is able to com­pute the value of \( \mathrm{dlog}(u) \) on a few well-chosen edges of \( \mathscr{B}^r \), then \( u \) has an \( m \)-th root only if \( m \) di­vides the gcd of these val­ues. Gekel­er util­ized this strategy in [e49] to com­pute the or­der of the di­visor \( [0] - [\infty] \) on \( X_0^2(\mathfrak{n}) \), where \( \mathfrak{n} \) is either a prime or a square of a prime. Since \( [0] \) and \( [\infty] \) are the only cusps when \( \mathfrak{n}=\mathfrak{p} \) is prime, Gekel­er’s res­ult im­plies that \( \mathcal{C}^2_{\mathfrak{p}} \) is cyc­lic of or­der \begin{equation}\label{eqGekCp} \frac{\lvert\mathfrak{p}\rvert-1}{\gcd(q^2-1, \lvert\mathfrak{p}\rvert-1)}. \tag{6.1} \end{equation} Ho [e114] ex­ten­ded Gekel­er’s com­pu­ta­tion to ar­bit­rary prime power levels \( \mathfrak{p}^s \). We note that the for­mula for the or­der of \( [0] - [\infty] \) on \( X_0^2(\mathfrak{p}^s) \) when \( s \geq 3 \) does not spe­cial­ize to the cases \( s = 1 \) or \( s = 2 \).

In high­er ranks the only res­ult so far is the fol­low­ing gen­er­al­iz­a­tion of \eqref{eqGekCp} proved in [e112]: \( \mathcal{C}^r_{\mathfrak{p}} \) is cyc­lic of or­der \begin{equation}\label{eqPapWei} \frac{\lvert\mathfrak{p}\rvert^{r-1}-1}{\gcd(q^r-1, \lvert\mathfrak{p}\rvert-1)}. \tag{6.2} \end{equation}

In ana­logy with \( \mathcal{C}(N) \), define the ra­tion­al cuspid­al di­visor class group \( \mathcal{C}(\mathfrak{n}) := \mathcal{C}^2(\mathfrak{n}) \) of \( X_0(\mathfrak{n}) \) to be the sub­group of \( J_0(\mathfrak{n}) \) gen­er­ated by the lin­ear equi­val­ence classes of the de­gree 0 ra­tion­al cuspid­al di­visors on \( X_0(\mathfrak{n}) \). The group \( \mathcal{C}(\mathfrak{n}) \) is ex­pli­citly com­puted in the fol­low­ing cases:

  • \( \deg(\mathfrak{n})=3 \); see [e87].
  • \( \mathfrak{n}=\mathfrak{p}_1\mathfrak{p}_2 \) is a product of two dis­tinct primes; see ([e85], Sec­tion 6).
  • \( \mathfrak{n}=\mathfrak{p}_1\cdots \mathfrak{p}_s \) is square-free, but \( \mathcal{C}(\mathfrak{n})_\ell \) is de­term­ined only for \( \ell\nmid q-1 \); see [e90]. In this case, \( \mathcal{C}(\mathfrak{n})_p=0 \), where \( p \) is the char­ac­ter­ist­ic of \( F \). Moreover, \( \mathcal{C}(\mathfrak{n})_\ell \) nat­ur­ally de­com­poses in­to a dir­ect sum of \( 2^{s-1} \) cyc­lic sub­groups each of which is an ei­gen­space with re­spect to the Atkin–Lehner in­vol­u­tions of \( J_0(\mathfrak{n}) \).
  • \( \mathfrak{n}=\mathfrak{p}^s \) is a prime power; see [e114]. An in­ter­est­ing fact in this case is that \( \mathcal{C}(\mathfrak{n}) \) is “mostly” \( p \)-primary.
Re­mark 6.2: When \( \deg(\mathfrak{n})=3 \) or \( \mathfrak{n} \) is square-free, all the cusps of \( X_0(\mathfrak{n}) \) are \( F \)-rational by ([e55], Proposition 6.7) and ([e87], Lemma 3.1), so \( \mathcal{C}(\mathfrak{n}) = \mathcal{C}_{\mathfrak{n}} \), but generally \( \mathcal{C}(\mathfrak{n}) \) can be strictly smaller than \( \mathcal{C}_{\mathfrak{n}} \).
6.2. Analogue of Ogg’s conjecture
Let \( J_0(\mathfrak{n})(F)_\mathrm{tor} \) be the ra­tion­al tor­sion sub­group of the Jac­obi­an vari­ety \( J_0(\mathfrak{n}) \). By the Lang–Néron the­or­em, \( J_0(\mathfrak{n})(F)_\mathrm{tor} \) is a fi­nite abeli­an group. As earli­er, let \( \mathcal{C}_{\mathfrak{n}} \) be the cuspid­al sub­group of \( J_0(\mathfrak{n}) \), \( \mathcal{C}_{\mathfrak{n}}(F) \) be the sub­group of ra­tion­al points on \( \mathcal{C}_{\mathfrak{n}} \), and \( \mathcal{C}(\mathfrak{n}) \) be the ra­tion­al cuspid­al di­visor class group of \( X_0(\mathfrak{n}) \). We have \[ \mathcal{C}(\mathfrak{n}) \subseteq \mathcal{C}_{\mathfrak{n}}(F) \subseteq J_0(\mathfrak{n})(F)_\mathrm{tor}. \] The ana­logue of Con­jec­ture CJ-N in this set­ting is the fol­low­ing.
Con­jec­ture CJD-\( \mathfrak{n} \): For any nonzero \( \mathfrak{n} \in A \), \[ \mathcal{C}(\mathfrak{n})= \mathcal{C}_{\mathfrak{n}}(F)= J_0(\mathfrak{n})(F)_\mathrm{tor}. \]
Re­mark 6.3: In view of the formula \eqref{eqPapWei}, which is valid for all \( r\geq 2 \), one might wonder whether Ogg’s conjecture extends to higher rank Drinfeld modular varieties: Is the \( F \)-rational torsion subgroup of the Picard group of \( X_0^r(\mathfrak{n}) \) cuspidal, i.e., generated by the \( F \)-rational elements of \( C^r_\mathfrak{n} \)?

There is a nat­ur­al morph­ism \( X_1(\mathfrak{n})\to X_0(\mathfrak{n}) \), which, by Pi­card func­tori­al­ity, in­duces a morph­ism \( J_0(\mathfrak{n})\to J_1(\mathfrak{n}) \). The ker­nel of this morph­ism, de­noted \( \mathcal{S}(\mathfrak{n}) \), is fi­nite; see ([e85], Sec­tion 8). This is the Shimura sub­group in this con­text. Moreover, when \( \mathfrak{n} \) is square-free, \( \mathcal{S}(\mathfrak{n}) \) is a \( \mu \)-type étale group-scheme over \( F \), so its or­der is coprime to \( p \). For the proof of this claim, as well as the cal­cu­la­tion of the group struc­ture of \( \mathcal{S}(\mathfrak{n}) \), we refer to ([e85], Sec­tion 8). The ana­logue of Con­jec­ture SJ-p is the fol­low­ing.

Con­jec­ture SJD-\( \mathfrak{p} \): For a prime \( \mathfrak{p}\in A \), the max­im­al \( \mu \)-type sub­group \( \mathcal{M}(\mathfrak{p}) \) of \( J_0(\mathfrak{p}) \) is \( \mathcal{S}(\mathfrak{p}) \).
Re­mark 6.4: For general square-free \( \mathfrak{n} \), the maximal \( \mu \)-type subgroup \( \mathcal{M}(\mathfrak{n}) \) of \( J_0(\mathfrak{n}) \) can be strictly larger than \( \mathcal{S}(\mathfrak{n}) \); we refer to ([e85], Section 8) for an explicit example. Also, if \( \mathfrak{n} \) is not square-free, then \( \mathcal{S}(\mathfrak{n}) \) has a nontrivial connected subgroup-scheme. Thus, Conjecture SJD-\( \mathfrak{p} \) does not extend to general \( \mathfrak{n} \). But we expect, in analogy with Vatsal’s result, that \( \mathcal{M}(\mathfrak{n})_\ell = \mathcal{S}(\mathfrak{n})_\ell \) for any square-free \( \mathfrak{n} \) and any prime \( \ell\nmid (q-1) \).

Con­jec­ture CJD-\( \mathfrak{n} \) for prime \( \mathfrak{n}=\mathfrak{p} \), as well as Con­jec­ture SJD-\( \mathfrak{p} \), were proved by Pál in [e66]. Over­all, Pál’s ap­proach is modeled on Mazur’s, but with some in­ter­est­ing dif­fer­ences that we will ex­plain be­low. Com­bin­ing Pál’s ap­proach with some ideas from [e83] and [e110], one can prove par­tial res­ults to­wards Con­jec­ture CJD-\( \mathfrak{n} \):

  • When \( \mathfrak{n} \) is square-free, we have \( \mathcal{C}(\mathfrak{n})_{\ell} = (J_0(\mathfrak{n})(F)_\mathrm{tor})_{\ell} \) for any prime \( \ell\nmid q(q-1) \); see [e90].
  • When \( \mathfrak{n}=\mathfrak{p}^s \) is a prime power, we have \( \mathcal{C}(\mathfrak{p}^s)_{\ell} = (J_0(\mathfrak{p}^s)(F)_\mathrm{tor})_{\ell} \) for any prime \( \ell\nmid q(q-1) \); see [e115].
6.3. Outline of the proofs
It is in­struct­ive to first re­call Mazur’s strategy for prov­ing Con­jec­ture CJ-p, up to 2-primary tor­sion. Let \( \mathbb{T}(p) \) be the Hecke al­gebra act­ing on \( J_0(p) \) and \( \mathfrak{E}(p)\subset \mathbb{T}(p) \) be the Ei­s­en­stein ideal gen­er­ated by the ele­ments \( 1+\ell-T_\ell \) for all prime \( \ell\neq p \) and by \( w_p+1 \), where \( w_p \) is the Atkin–Lehner in­vol­u­tion. Us­ing prop­er­ties of mod­u­lar forms, Mazur proved that \( \mathbb{T}(p)/\mathfrak{E}(p)\cong \mathbb{Z}/n\mathbb{Z} \), where \( n=(p-1)/\gcd(12, p-1) \). Let \( \mathcal{J} \) de­note the Néron mod­el of \( J_0(p) \) over \( \mathbb{Z} \), and \( \mathcal{J}_{\mathbb{F}_\ell} \) de­note its fiber at \( \ell \). Let \( \mathcal{G}_\ell:= \mathcal{J}_{\mathbb{F}_\ell}(\overline{\mathbb{F}}_\ell)[\ell^\infty] \) be the étale part of the \( \ell \)-di­vis­ible group of \( \mathcal{J}_{\mathbb{F}_\ell} \). The Hecke al­gebra \( \mathbb{T}(p) \) acts on \( \mathcal{J}_{\mathbb{F}_\ell} \). By a the­or­em of Carti­er and Serre, for odd \( \ell\neq p \) there is an in­jec­tion \[ \mathcal{G}_\ell[\ell]\otimes_{\mathbb{F}_\ell} \overline{\mathbb{F}}_\ell\hookrightarrow H^0(X_0(p)_{\mathbb{F}_\ell}, \Omega^1_{X_0(p)_{\mathbb{F}_\ell}}) \] com­pat­ible with the ac­tion of \( \mathbb{T}(p) \). Sup­pose \( \ell \) di­vides \( n \), and de­note \( \mathfrak{P}_\ell\lhd \mathbb{T}(p) \) the Ei­s­en­stein max­im­al ideal \( (\ell, \mathfrak{E}(p)) \). From the above in­jec­tion, one de­duces that the ker­nel \( \mathcal{G}_\ell[\mathfrak{P}_\ell] \) is con­tained in \( H^0(X_0(p)_{\mathbb{F}_\ell}, \Omega^1_{X_0(p)_{\mathbb{F}_\ell}})[\mathfrak{P}_\ell] \). Now us­ing the du­al­ity between the Hecke al­gebra and the space of cusp forms, one de­duces that this lat­ter space is one di­men­sion­al over \( \mathbb{T}(\mathfrak{p})/\mathfrak{P}_\ell\cong \mathbb{F}_\ell \), thus the di­men­sion of \( \mathcal{G}_\ell[\mathfrak{P}_\ell] \) is also at most one over \( \mathbb{T}(\mathfrak{p})/\mathfrak{P}_\ell\cong \mathbb{F}_\ell \). This can be ex­ten­ded to show that \( \mathcal{G}_\ell[\mathfrak{E}(p)] \) is a cyc­lic mod­ule over \( \mathbb{T}(p)\otimes_{\mathbb{Z}}\mathbb{Z}_\ell \), so \( \mathcal{G}_\ell[\mathfrak{E}(p)] \) is iso­morph­ic to a sub­group of \( \mathbb{Z}_\ell/n\mathbb{Z}_\ell \). Let \( \mathcal{T}:= J_0(p)(\mathbb{Q})_\mathrm{tor} \). The Eichler–Shimura con­gru­ence re­la­tions im­ply that \( \mathfrak{E}(p) \) an­ni­hil­ates \( \mathcal{T} \). Hence, if \( \ell\mid \#\mathcal{T} \), then \( \ell\mid n \). From the Néron map­ping prop­erty, we ob­tain an in­jec­tion \( \mathcal{T}_\ell\hookrightarrow \mathcal{G}_\ell[\mathfrak{E}(p)] \). Thus, \( \# \mathcal{T}_\ell\leq \mathbb{Z}_\ell/n\mathbb{Z}_\ell \). On the oth­er hand, \( \mathbb{Z}_\ell/n\mathbb{Z}_\ell\cong \mathcal{C}(p)_\ell\hookrightarrow \mathcal{T}_\ell \), so we con­clude that \( \mathcal{T}_\ell= \mathcal{C}(p)_\ell \).

big­break Be­fore pro­ceed­ing to the proof in the func­tion field case, we need to dis­cuss a few more facts about the Jac­obi­ans of Drin­feld mod­u­lar curves. The Jac­obi­an \( J_0(\mathfrak{n}) \) has a ri­gid-ana­lyt­ic uni­form­iz­a­tion over \( F_\infty \) as a quo­tient of a mul­ti­plic­at­ive tor­us by a dis­crete lat­tice; this is closely re­lated to the fact that \( Y_0(\mathfrak{n}) \) has a ri­gid-ana­lyt­ic uni­form­iz­a­tion \( \Gamma_0(\mathfrak{n})\setminus \Omega^2 \), so \( X_0(\mathfrak{n}) \) is a Mum­ford curve. The con­nec­tion between the ana­lyt­ic uni­form­iz­a­tion of Mum­ford curves and their Jac­obi­ans was first ex­plic­ated by Man­in and Drin­feld [e9]. In the set­ting of Drin­feld mod­u­lar curves this was done by Gekel­er and Re­ver­sat in [e47]. De­note by \( \overline{\Gamma_0(\mathfrak{n})} \) the max­im­al tor­sion-free abeli­an quo­tient of \( \Gamma_0(\mathfrak{n}) \). Us­ing ana­lyt­ic theta func­tions, Gekel­er and Re­ver­sat con­struct a pair­ing \begin{equation}\label{eqGRpair} \langle\, \cdot\,, \cdot\,\rangle \colon \overline{\Gamma_0(\mathfrak{n})}\times \overline{\Gamma_0(\mathfrak{n})}\to F_\infty\tag{6.3} \end{equation} and show that it in­duces an ex­act se­quence \begin{equation}\label{eqGRs} 0\to\overline{\Gamma_0(\mathfrak{n})}\xrightarrow{\alpha\mapsto \langle \alpha,\, \cdot\,\rangle}\operatorname{Hom}(\overline{\Gamma_0(\mathfrak{n})}, \mathbb{C}_\infty^\times)\to J_0(\mathfrak{n})(\mathbb{C}_\infty)\to 0. \tag{6.4} \end{equation} One can define Hecke op­er­at­ors \( T_\mathfrak{p} \) as en­do­morph­isms of \( \overline{\Gamma_0(\mathfrak{n})} \) in purely group-the­or­et­ic­al terms as some sort of Ver­la­ger­ung (see ([e47], (9.3))). The above ex­act se­quence then be­comes com­pat­ible with the ac­tion of \( \mathbb{T}(\mathfrak{n}) \) on its three terms. Also, by ([e47], (3.3.3)) and [e42], there is a ca­non­ic­al iso­morph­ism \[ \overline{\Gamma_0(\mathfrak{n})}\xrightarrow{\sim}\mathcal{H}_0(\mathfrak{n},\mathbb{Z}) \] com­pat­ible with the ac­tion of Hecke op­er­at­ors, so in \eqref{eqGRs} one can re­place \( \overline{\Gamma_0(\mathfrak{n})} \) with \( \mathcal{H}_0(\mathfrak{n},\mathbb{Z}) \).

Let \( \mathcal{J}_0(\mathfrak{n}) \) be the Néron mod­el of \( J_0(\mathfrak{n}) \) over \( \mathbb{P}^1_{\mathbb{F}_q} \). Let \( \mathcal{J}_0(\mathfrak{n})_{\mathbb{F}_\infty} \) be the fiber of \( \mathcal{J}_0(\mathfrak{n}) \) over the closed point \( \infty \) of \( \mathbb{P}^1_{\mathbb{F}_q} \), let \( \mathcal{J}_0(\mathfrak{n})_{\mathbb{F}_\infty}^0 \) be its con­nec­ted com­pon­ent of the iden­tity, and let \[ \Phi_\infty(\mathfrak{n}):= \mathcal{J}_0(\mathfrak{n})_{\overline{\mathbb{F}}_\infty}/\mathcal{J}_0(\mathfrak{n})_{\overline{\mathbb{F}}_\infty}^0 \] be the group of con­nec­ted com­pon­ents of \( \mathcal{J}_0(\mathfrak{n})_{\mathbb{F}_\infty} \).

The valu­ation of the pair­ing \eqref{eqGRpair}, \[ \operatorname{ord}_\infty \langle\, \cdot\,, \cdot\,\rangle \colon\ \mathcal{H}_0(\mathfrak{n},\mathbb{Z}) \times \mathcal{H}_0(\mathfrak{n},\mathbb{Z})\longrightarrow \mathbb{Z}, \] is a weighted ver­sion of the cycle pair­ing on \( \mathcal{H}_0(\mathfrak{n},\mathbb{Z}) \) (see ([e47], The­or­em 5.7.1)), and the uni­form­iz­a­tion se­quence \eqref{eqGRs} in­duces the ex­act se­quence \begin{equation}\label{eqGpairPhi} 0\to \mathcal{H}_0(\mathfrak{n},\mathbb{Z})\xrightarrow{f \mapsto \operatorname{ord}_\infty \langle f,\, \cdot\,\rangle}\operatorname{Hom}(\mathcal{H}_0(\mathfrak{n},\mathbb{Z}), \mathbb{Z})\to \Phi_\infty(\mathfrak{n})\to 0,\tag{6.5} \end{equation} com­pat­ible with the ac­tion of \( \mathbb{T}(\mathfrak{n}) \) on its three terms; see ([e43], Co­rol­lary 2.11). (The pair­ing \( \operatorname{ord}_\infty \langle\, \cdot\, , \cdot\, \rangle \) can also be iden­ti­fied with Grothen­dieck’s “mono­dromy pair­ing”, so the se­quence \eqref{eqGpairPhi} is a spe­cial case of a the­or­em of Grothen­dieck from Ex­posé IX in SGA 7 [e6].)

With these pre­lim­in­ar­ies out of the way, we can now ex­plain Pál’s ap­proach to Con­jec­ture CJD-\( \mathfrak{n} \) for prime \( \mathfrak{n}=\mathfrak{p} \), which was ex­ten­ded to square-free \( \mathfrak{n} \) and prime power \( \mathfrak{n} \) in [e90] and [e114], re­spect­ively. Let \( \mathcal{T}(\mathfrak{n}):= J_0(\mathfrak{n})(F)_\mathrm{tor} \). Let \( \ell \) be a prime not equal to \( p \). The Eichler–Shimura con­gru­ence re­la­tions im­ply that \( \mathfrak{E}(\mathfrak{n}) \) an­ni­hil­ates \( \mathcal{T}(\mathfrak{n})_\ell \). On the oth­er hand, by the Néron map­ping prop­erty, there is a ca­non­ic­al in­ject­ive morph­ism \( \mathcal{T}(\mathfrak{n})_\ell \hookrightarrow \mathcal{J}_0(\mathfrak{n})_{\mathbb{F}_\infty}(\mathbb{F}_\infty) \). Since \( \mathcal{J}_0(\mathfrak{n})^0_{\mathbb{F}_\infty} \) is a split tor­us, we have \( \mathcal{J}_0(\mathfrak{n})^0_{\mathbb{F}_\infty}(\mathbb{F}_\infty)\cong (\mathbb{F}_q^\times)^g \), where \( g=\dim(J_0(\mathfrak{n})) \). Hence, if we as­sume that \( \ell\nmid (q-1) \), then we get an in­ject­ive ho­mo­morph­ism \( \mathcal{T}(\mathfrak{n})_\ell\hookrightarrow \Phi_\infty(\mathfrak{n})_\ell[\mathfrak{E}(\mathfrak{n})] \), where the lat­ter group is the sub­group of \( \Phi_\infty(\mathfrak{n})_\ell \) an­ni­hil­ated by the Ei­s­en­stein ideal. Fix some \( n \) large enough so that \( \ell^n \) an­ni­hil­ates \( \Phi_\infty(\mathfrak{n})_\ell \). Mul­tiply­ing the se­quence \eqref{eqGpairPhi} by \( \ell^n \) and ap­ply­ing the snake lemma, we get an in­jec­tion \[ \mathcal{T}(\mathfrak{n})_\ell\hookrightarrow \mathcal{E}_{00}(\mathfrak{n}, \mathbb{Z}/\ell^n\mathbb{Z}), \] where \( \mathcal{E}_{00}(\mathfrak{n}, \mathbb{Z}/\ell^n\mathbb{Z}) \) is the sub­mod­ule of \( \mathcal{H}_{00}(\mathfrak{n}, \mathbb{Z}/\ell^n\mathbb{Z}) \) an­ni­hil­ated by \( \mathfrak{E}(\mathfrak{n}) \). Us­ing the Four­i­er ex­pan­sions of har­mon­ic co­chains, one shows that if \( f\in \mathcal{H}_{00}(\mathfrak{n}, \mathbb{Z}/\ell^n\mathbb{Z}) \) is an­ni­hil­ated by the Ei­s­en­stein ideal, then it is a scal­ar mul­tiple of the re­duc­tion mod­ulo \( \ell^n \) of a \( \mathbb{Z} \)-val­ued Ei­s­en­stein series (this is a “mul­ti­pli­city-one” state­ment). From this one de­duces that \( \mathcal{E}_{00}(\mathfrak{n}, \mathbb{Z}/\ell^n\mathbb{Z})\cong \mathbb{Z}_\ell/N(\mathfrak{n})\mathbb{Z}_\ell \), where \( N(\mathfrak{n}) \) is es­sen­tially the con­stant Four­i­er coef­fi­cient of an Ei­s­en­stein series. The num­ber \( N(\mathfrak{n}) \) can be ex­pli­citly com­puted and it matches the size of \( \mathcal{C}(\mathfrak{n})_\ell \); see Ex­ample 4.6. Since \( \mathcal{C}(\mathfrak{n})_\ell\subseteq \mathcal{T}(\mathfrak{n})_\ell \subseteq \mathcal{E}_{00}(\mathfrak{n}, \mathbb{Z}/\ell^n\mathbb{Z})\cong \mathbb{Z}_\ell/N(\mathfrak{n})\mathbb{Z}_\ell \), we con­clude that \( \mathcal{C}(\mathfrak{n})_\ell\cong \mathcal{T}(\mathfrak{n})_\ell \). Thus, over­all, this ar­gu­ment is sim­il­ar to Mazur’s ar­gu­ment but in­stead of spe­cial­iz­ing \( \mathcal{T}(\mathfrak{n}) \) in­to the fibers of \( \mathcal{J}(\mathfrak{n}) \) at fi­nite primes, one uses the fiber over \( \infty \) (which does not have a dir­ect ana­logue over \( \mathbb{Q} \)).

For primes \( \ell \) di­vid­ing \( q-1 \), prov­ing \( \mathcal{C}(\mathfrak{p})_\ell=\mathcal{T}(\mathfrak{p})_\ell \) is much harder and is closely linked to the fact that \( \mathbb{T}(\mathfrak{p}) \) is loc­ally Goren­stein at the prime ideals con­tain­ing \( \mathfrak{E}(\mathfrak{p}) \). This is also a key fact used in the proof of Con­jec­ture SJD-\( \mathfrak{p} \) in [e66]. The Goren­stein prop­erty was proved by Pál in [e66] by ad­apt­ing Mazur’s Ei­s­en­stein des­cent ar­gu­ment. This prop­erty im­plies that \( \mathcal{T}(\mathfrak{p})_\ell \) and \( \mathcal{M}(\mathfrak{p})_\ell \) are dual to each oth­er for \( \ell\nmid (q-1) \), and \( J_0(\mathfrak{p})[\mathfrak{E}(\mathfrak{p})]_\ell=\mathcal{C}(\mathfrak{p})_\ell\oplus \mathcal{S}(\mathfrak{p})_\ell \). When \( \ell\mid (q-1) \), the groups \( \mathcal{C}(\mathfrak{p})_\ell \) and \( \mathcal{S}(\mathfrak{p})_\ell \) in­ter­sect in \( J_0(\mathfrak{p}) \) and the proof re­quires the con­struc­tion of an aux­il­i­ary “di­hed­ral sub­group”. We will not dis­cuss the de­tails of these ar­gu­ments due to their more tech­nic­al nature.

Re­mark 6.5: (1)  Pál also gave a second proof of the Gorenstein property of the localizations of \( \mathbb{T}(\mathfrak{p}) \) at Eisenstein primes in [e69] by adapting an argument of Calegari and Emerton [e64]. In this approach \( \mathbb{T}(\mathfrak{p}) \) is identified with a universal deformation ring \( R(\mathfrak{p}) \), which is then shown to be generated by a single element over \( \mathbb{Z}_\ell \) using cohomological methods.

(2)  If \( N=p_1p_2 \) is a product of two dis­tinct primes, then the clas­sic­al Hecke al­gebra \( \mathbb{T}(N) \) is gen­er­ally not loc­ally Goren­stein at the Ei­s­en­stein primes; see [e88]. The same is most likely true for \( \mathbb{T}(\mathfrak{n}) \). Hence, some genu­inely new ideas might be needed to prove \( \mathcal{C}(\mathfrak{n})_\ell=\mathcal{T}(\mathfrak{n})_\ell \) when \( \mathfrak{n} \) is not prime and \( \ell\mid q-1 \).

(3)  The em­bed­ding \( \mathcal{T}(\mathfrak{n})_\ell\hookrightarrow \Phi_\infty(\mathfrak{n})[\mathfrak{E}(\mathfrak{n})] \) plays an im­port­ant role in the ar­gu­ment out­lined earli­er. It is a well known fact due to Ribet and Edix­hoven that the com­pon­ent groups of the clas­sic­al mod­u­lar Jac­obi­an \( J_0(N) \) are an­ni­hil­ated by \( T_\ell-(\ell+1) \) for all \( \ell\nmid N \). Thus, one might won­der wheth­er \( \Phi_\infty(\mathfrak{n})[\mathfrak{E}(\mathfrak{n})]=\Phi_\infty(\mathfrak{n}) \) and try to es­tim­ate the size of \( \Phi_\infty(\mathfrak{n}) \). However, it turns out that \( \Phi_\infty(\mathfrak{n}) \) is ex­po­nen­tially lar­ger than its sub­group \( \Phi_\infty(\mathfrak{n})[\mathfrak{E}(\mathfrak{n})] \); see [e89].

Fi­nally, we dis­cuss the \( p \)-primary ra­tion­al tor­sion sub­group \( \mathcal{T}(\mathfrak{n})_p \) of \( J_0(\mathfrak{n}) \). Let \( P\in \mathcal{T}(\mathfrak{n})_p \) be an ele­ment of or­der \( p \), and let \( G\cong \mathbb{Z}/p\mathbb{Z} \) be the con­stant sub­group-scheme of \( J_0(\mathfrak{n}) \) gen­er­ated by \( P \). By the ex­ten­sion prop­erty for étale points of Néron mod­els, \( G \) ex­tends to a fi­nite flat sub­group scheme \( \mathcal{G} \) of \( \mathcal{J}_0(\mathfrak{n}) \). For a closed point \( x \) of \( \mathbb{P}^1_{\mathbb{F}_q} \), de­note the spe­cial fiber of \( \mathcal{G} \) at \( x \) by \( \mathcal{G}_{\mathbb{F}_x} \). Sup­pose \( J_0(\mathfrak{n}) \) has purely tor­ic re­duc­tion at \( x \), i.e., \( \mathcal{J}^0_{\overline{\mathbb{F}}_x} \) is iso­morph­ic to a product of cop­ies of the mul­ti­plic­at­ive group \( \mathbb{G}_{m, \overline{\mathbb{F}}_x} \). In that case, if \( \mathcal{G}_{\mathbb{F}_x} \) is a sub­group scheme of \( \mathcal{J}^0_{\overline{\mathbb{F}}_x} \), then \( \mathcal{G}_{\mathbb{F}_x}\cong \mu_p \). Since the Carti­er dual of \( \mu_p \) is \( \mathbb{Z}/p\mathbb{Z} \) and vice versa, the Carti­er dual of \( \mathcal{G} \) would have con­nec­ted gen­er­ic fiber but étale closed fiber, which im­possible. There­fore, \( \mathcal{G}_{\mathbb{F}_x} \) is iso­morph­ic to \( \mathbb{Z}/p\mathbb{Z} \) and is a sub­group of \( \Phi_x(\mathfrak{n}) \).

Pro­pos­i­tion 6.6: ([e66], Corollary 7.15) If \( \mathfrak{p} \) is prime, then \( \mathcal{T}(\mathfrak{p})_p=0 \).

Proof. \( J_0(\mathfrak{p}) \) has purely tor­ic re­duc­tion at \( \mathfrak{p} \), as is shown in [e24] us­ing De­ligne–Ra­po­port type ar­gu­ments. Moreover, us­ing the struc­ture of \( X_0(\mathfrak{p})_{\mathbb{F}_\mathfrak{p}} \) and Raynaud’s meth­ods, Gekel­er com­puted in [e24] that \[ \Phi_\mathfrak{p}(\mathfrak{p})\cong (\lvert\mathfrak{p}\rvert-1)/\gcd(\lvert\mathfrak{p}\rvert-1, q^2-1). \] In par­tic­u­lar, \( \Phi_\mathfrak{p}(\mathfrak{p}) \) has no \( p \)-tor­sion. The claim now fol­lows from our earli­er ar­gu­ment. □

The Jac­obi­an \( J_0(\mathfrak{n}) \) has purely tor­ic re­duc­tion at \( \mathfrak{p}\mid \mathfrak{n} \) in a few oth­er cases when \( \mathfrak{n} \) is the product of \( \mathfrak{p} \) and a square-free poly­no­mi­al of de­gree \( \leq 2 \). The proof of Pro­pos­i­tion 6.6 ex­tends to these cases to show that \( \mathcal{T}(\mathfrak{n})_p=0 \). Gen­er­ally, we ex­pect that \( \mathcal{T}(\mathfrak{n})_p=0 \) when \( \mathfrak{n} \) is square-free, but this is cur­rently open. In the oth­er ex­treme, when \( \mathfrak{n}=\mathfrak{p}^s \) is a prime power with \( s\geq 3 \), we ex­pect that \( \mathcal{T}(\mathfrak{n}) \) is “mostly” \( p \)-primary, as this fol­lows from Con­jec­ture CJD-\( \mathfrak{n} \) and the com­pu­ta­tion of \( \mathcal{C}(\mathfrak{p}^s)_p \) in [e114]; for ex­ample, \( \mathcal{T}(T^s) \) must be a \( p \)-primary group.

Re­mark 6.7: \( J_0(\mathfrak{n}) \) has purely toric reduction at \( \infty \) for any \( \mathfrak{n} \). Unfortunately, the group \( \Phi_\infty(\mathfrak{n}) \) usually has nontrivial \( p \)-torsion, even when \( \mathfrak{n}=\mathfrak{p} \) is prime, so \( \Phi_\infty(\mathfrak{n}) \) cannot be used to show that \( \mathcal{T}(\mathfrak{n})_p=0 \) for square-free \( \mathfrak{n} \). Moreover, the structure of \( \Phi_\infty(\mathfrak{n}) \) depends on the actual prime divisors of \( \mathfrak{n} \), and not only on their degrees, so there is no uniform way of describing this group.
Defin­i­tion 6.8: We say that \( \mathcal{T}(\mathfrak{n})_p \) is Ei­s­en­stein if it is annihilated by \( \eta_\mathfrak{p}:= T_\mathfrak{p}-(\lvert\mathfrak{p}\rvert+1) \) for all prime \( \mathfrak{p}\nmid \mathfrak{n} \).
Lemma 6.9: If \( \mathfrak{p} \) is a prime of good or­din­ary re­duc­tion of \( J_0(\mathfrak{n}) \) then \( \eta_\mathfrak{p} \) an­ni­hil­ates \( \mathcal{T}(\mathfrak{n})_p \).

Proof: By as­sump­tion, \( \mathcal{J}_0(\mathfrak{n})_{\mathbb{F}_\mathfrak{p}}(\overline{\mathbb{F}}_\mathfrak{p})[p^s] \cong (\mathbb{Z}/p^s\mathbb{Z})^g \) for all \( s\geq 1 \), where \( g=\dim J_0(\mathfrak{n}) \). This im­plies that the re­duc­tion \( \mathcal{T}(\mathfrak{n})_p \to \mathcal{J}_0(\mathfrak{n})_{\mathbb{F}_\mathfrak{p}}(\mathbb{F}_\mathfrak{p}) \) is in­ject­ive. Let \( \operatorname{Frob}_\mathfrak{p} \) de­note the Frobeni­us en­do­morph­ism of the abeli­an vari­ety \( \mathcal{J}_0(\mathfrak{n})_{\mathbb{F}_\mathfrak{p}} \). Since \( \mathfrak{p}\nmid \mathfrak{n} \), the Hecke op­er­at­or \( T_\mathfrak{p} \) sat­is­fies the Eichler–Shimura re­la­tion, \[ \operatorname{Frob}_\mathfrak{p}^2-T_\mathfrak{p}\cdot \operatorname{Frob}_\mathfrak{p}+\lvert\mathfrak{p}\rvert =0, \] in the en­do­morph­ism ring of \( \mathcal{J}_0(\mathfrak{n})_{\mathbb{F}_\mathfrak{p}} \). Since \( \operatorname{Frob}_\mathfrak{p} \) fixes the re­duc­tion of \( \mathcal{T}(\mathfrak{n})_p \), the en­do­morph­ism \( 1-T_\mathfrak{p}+\lvert\mathfrak{p}\rvert \) an­ni­hil­ates this group. As the re­duc­tion map com­mutes with the ac­tion of the Hecke al­gebra, we get that \( \eta_\mathfrak{p} \) an­ni­hil­ates \( \mathcal{T}(\mathfrak{n})_p \). □

Pro­pos­i­tion 6.10: \( J_0(\mathfrak{n}) \) has good or­din­ary re­duc­tion at all but fi­nitely many places of \( F \).

Proof: As we dis­cussed, \( J_0(\mathfrak{n})(\mathbb{C}_\infty)\cong (\mathbb{C}_\infty^\times)^g/\Lambda \), where \( \Lambda \) is a lat­tice of rank \( g \). Thus, \[ \dim_{\mathbb{F}_p} J_0(\mathfrak{n})(\mathbb{C}_\infty)[p] = \dim_{\mathbb{F}_p}\Lambda/p\Lambda=g, \] so \( J_0(\mathfrak{n}) \) is an or­din­ary abeli­an vari­ety over \( F \). Now an ar­gu­ment us­ing the Hasse–Witt mat­rix im­plies that \( \mathcal{J}_0(\mathfrak{n}) \) has good or­din­ary re­duc­tion at all but fi­nitely many places of \( F \); see ([e86], Pro­pos­i­tion 8.3). □

Ex­ample 6.11: (1)  \( J_0(T^s) \) has good ordinary reduction at all primes of \( A \) different from \( T \); see ([e50], Proposition 3.5). Thus, \( \mathcal{T}(T^s)_p \) is Eisenstein, so can be studied using the machinery of the Eisenstein ideal. Unfortunately, the standard method for computing the index of the Eisenstein ideal in \( \mathbb{T}(\mathfrak{n}) \) relies on the Fourier expansions of harmonic cochains, which requires inverting \( p \).

(2)  Let \( q=2 \) and \( \mathfrak{n}=T(T^2+T+1) \). In this case, \( J_0(\mathfrak{n}) \) is iso­gen­ous to a product of two el­lipt­ic curves which can be ana­lyzed us­ing their ex­pli­cit equa­tions. One con­cludes that \( J_0(\mathfrak{n}) \) has su­per­sin­gu­lar re­duc­tion at \( T+1 \), and good or­din­ary re­duc­tion at any prime of \( A \) not equal to \( T \), \( T+1 \), or \( T^2+T+1 \). In this case, \( \mathcal{T}(\mathfrak{n})\cong \mathcal{C}l(\mathfrak{n})\cong \mathbb{T}(\mathfrak{n})/\mathfrak{E}(\mathfrak{n})\cong\mathbb{Z}/15\mathbb{Z} \); see [e87].

Let \( J_0(\mathfrak{n})^\mathrm{new} \) be the “new” quo­tient of \( J_0(\mathfrak{n}) \), i.e., the quo­tient of \( J_0(\mathfrak{n}) \) by its abeli­an sub­vari­ety gen­er­ated by the im­ages \( J_0(\mathfrak{m})\to J_0(\mathfrak{n}) \) for all \( \mathfrak{m}\supsetneq \mathfrak{n} \) un­der the morph­isms in­duces by the de­gen­er­acy maps \( X_0(\mathfrak{n})\to X_0(\mathfrak{m}) \).

Pro­pos­i­tion 6.12: Sup­pose \( \mathfrak{p} \) is a place of good nonor­din­ary re­duc­tion for \( J_0(\mathfrak{n})^\mathrm{new} \). If the im­age of \( \mathcal{T}(\mathfrak{n})_p \) in \( J_0(\mathfrak{n})^\mathrm{new} \) is nonzero, then \( \mathcal{T}(\mathfrak{n})_p \) is not Ei­s­en­stein.

Proof: If \( \mathcal{T}(\mathfrak{n})_p \) is Ei­s­en­stein and its im­age in \( J_0(\mathfrak{n})^\mathrm{new} \) is nonzero, then in the quo­tient \( \mathbb{T}(\mathfrak{n})^\mathrm{new} \) by which \( \mathbb{T}(\mathfrak{n}) \) acts on \( J_0(\mathfrak{n})^\mathrm{new} \) there is a prop­er max­im­al ideal \( \mathfrak{M} \) con­tain­ing \( p \) and the im­age of \( \mathfrak{E}(\mathfrak{n}) \). In par­tic­u­lar, \( \eta_\mathfrak{p}\in \mathfrak{M} \). On the oth­er hand, be­cause of the nonor­din­ary re­duc­tion as­sump­tion, \( T_\mathfrak{p}\in \mathfrak{M} \) by ([e86], The­or­em B.13 (b)). This im­plies that \( \lvert\mathfrak{p}\rvert+1\in \mathfrak{M} \). Since \( p\in \mathfrak{M} \), we get \( 1\in \mathfrak{M} \), a con­tra­dic­tion. (This type of ar­gu­ments also ap­pear in the proof of The­or­em 1.2 in [e69].) □

Cécile Ar­mana is an as­so­ci­ate pro­fess­or at the Uni­versité Mar­ie et Louis Pas­teur, Be­s­ançon, France. She works on mod­u­lar forms and arith­met­ic geo­metry over func­tion fields.

Sheng-Yang Kev­in Ho is cur­rently a gradu­ate stu­dent at the Pennsylvania State Uni­versity un­der the su­per­vi­sion of Mihran Papiki­an.

Mihran Papiki­an is a pro­fess­or at the Pennsylvania State Uni­versity. His re­search in­terests are in al­geb­ra­ic num­ber the­ory, es­pe­cially in the the­ory of Drin­feld mod­ules and re­lated top­ics.

Works

[1] A. P. Ogg: “Ra­tion­al points of fi­nite or­der on el­lipt­ic curves,” In­vent. Math. 12 (1971), pp. 105–​111. MR 291084 Zbl 0216.​05602 article

[2] A. P. Ogg: “Ra­tion­al points on cer­tain el­lipt­ic mod­u­lar curves,” pp. 221–​231 in Ana­lyt­ic num­ber the­ory (St. Louis Univ., 1972). Edi­ted by H. G. Dia­mond. Proc. Sym­pos. Pure Math. XXIV. Amer. Math. Soc., 1973. MR 337974 Zbl 0273.​14008 inproceedings

[3] A. P. Ogg: “Di­o­phant­ine equa­tions and mod­u­lar forms,” Bull. Amer. Math. Soc. 81 (1975), pp. 14–​27. MR 354675 Zbl 0316.​14012 article

[4] A. P. Ogg: “Mauvaise réduc­tion des courbes de Shimura,” pp. 199–​217 in Sémin­aire de théorie des nombres, Par­is 1983–84. Edi­ted by C. Gold­stein. Pro­gr. Math. 59. Birkhäuser (Bo­ston), 1985. MR 902833 Zbl 0581.​14024 incollection