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

A modular proof of the properness of the Coleman-Mazur eigencurve

Lynnelle Ye Department of Mathematics, Building 380, Stanford, CA 94305 ([email protected])
Abstract

We give a new proof of the properness of the Coleman-Mazur eigencurve. The question of whether the eigencurve satisfies the valuative criterion for properness was first asked by Coleman and Mazur in 1998 and settled by Diao and Liu in 2016 using deep, powerful Hodge- and Galois- theoretic machinery. Our proof is short and explicit and uses no Galois theory. Instead we adapt an earlier method of Buzzard and Calegari based on elementary properties of overconvergent modular forms. To facilitate this, we extend Pilloni’s geometric construction of overconvergent forms of arbitrary weight farther into the supersingular locus. Along the way, we show that the Hecke operator UpU_{p} is injective on spaces of forms of large overconvergence radius of any analytic weight.

1 Introduction

1.1 Background

The Coleman-Mazur eigencurve \mathscr{E} is a rigid analytic space parametrizing overconvergent pp-adic modular Hecke eigenforms with nonzero UpU_{p}-eigenvalues. It has a projection map w:𝒲w:\mathscr{E}\to\mathscr{W}, where 𝒲\mathscr{W} is the weight space parametrizing continuous characters of p×\mathbb{Z}_{p}^{\times}. It was constructed by Coleman and Mazur in [9] for p>2p>2 and tame level 11, and generalized to all primes and tame levels by Buzzard in [4].

In their original 1998 paper, Coleman and Mazur asked whether there exist pp-adic analytic families of overconvergent eigenforms of finite slope parametrized by a punctured disc, converging at the puncture to an overconvergent eigenform of infinite slope. Buzzard and Calegari suggested in [6] that this can be reframed as the question of whether w:𝒲w:\mathscr{E}\to\mathscr{W} satisfies the valuative criterion for properness, and proved by explicit means that the eigencurve is indeed “proper” in this sense for p=2p=2 and tame level 11. (Note that w:𝒲w:\mathscr{E}\to\mathscr{W} has infinite degree, so is not proper in the “usual” sense.) Calegari in [7] generalized the Buzzard-Calegari argument to show that the eigencurve is proper at algebraic weights for all primes and tame levels.

The question was eventually resolved by Diao and Liu, who proved in 2014 ([12]) that the eigencurve is indeed proper. Their proof is completely different from the method of Buzzard-Calegari and Calegari. It proceeds by analyzing families of Galois representations over the eigencurve, relying on deep, powerful Galois- and Hodge- theoretic machinery developed by Berger-Colmez, Kedlaya-Liu, Bellovin, Kedlaya-Pottharst-Xiao, Liu, and others.

In this paper, we give a new proof of the theorem of Diao-Liu based on the original method of Buzzard and Calegari. Our proof is short and explicit, and it uses no Galois theory, instead relying only on elementary properties of overconvergent modular forms. A precise statement of the theorem follows.

Theorem 1.1.1 (Originally proved by Diao-Liu).

Let \mathscr{E} be the pp-adic Coleman-Mazur eigencurve of tame level NN. Let DD be the closed unit disc, and write D×D^{\times} for DD with the origin removed. Suppose that h:D×h:D^{\times}\to\mathscr{E} is a morphism of rigid spaces such that whw\circ h extends to DD. Then hh extends to a morphism h~:D\tilde{h}:D\to\mathscr{E} compatible with whw\circ h.

1.2 Method of proof

We now describe our method. It suffices to prove the statement over the center disc of 𝒲\mathscr{W}, consisting of analytic weights (i.e. characters of p×\mathbb{Z}_{p}^{\times} that are analytic on l+ppl+p\mathbb{Z}_{p} for any integer ll). This is because the geometry of \mathscr{E} over the remainder (“boundary annulus”) of 𝒲\mathscr{W} is relatively simple. In particular, by Theorem 1.2.1 of [20], in that region, \mathscr{E} decomposes into a countable disjoint union of pieces that are finite over 𝒲\mathscr{W}. (For philosophical completeness, we remark that Ren and Zhao’s proof of this theorem is also based on concrete computation with no use of Galois theory.) On the other hand, the geometry of \mathscr{E} over the center disc of 𝒲\mathscr{W} is expected to be very complicated.

Now the argument of Buzzard-Calegari and Calegari for algebraic weights can be summarized as follows. First, it is standard that a finite-slope eigenform of any overconvergence radius analytically continues to a finite-slope eigenform of “large” overconvergence radius. Next, it is not too hard to check that the family h:D×h:D^{\times}\to\mathscr{E} of finite-slope overconvergent eigenforms extends to an overconvergent eigenform h~(0)\tilde{h}(0) at the puncture, also of “large” overconvergence radius; the question is whether h~(0)\tilde{h}(0) is also finite-slope. Finally, one shows that UpU_{p} is in fact injective on the space of forms of “large” overconvergence radius, so indeed Uph~(0)0U_{p}\tilde{h}(0)\neq 0.

The limitation of this argument is its reliance on Katz’s geometric definition of an overconvergent modular form of algebraic weight kk as a section of the line bundle ωk\omega^{k} on a certain rigid subset of the modular curve X:=X1(N)X:=X_{1}(N), along with Coleman’s ad hoc definition of an overconvergent modular form of general weight as a Katz overconvergent form of weight 0 multiplied by the pp-deprived Eisenstein series of the desired weight. In order to generalize it, we need to be able to write overconvergent modular forms of arbitrary weight as sections of line bundles on modular curves.

Fortunately, Pilloni provides such a definition in [19], constructing a weight-ww line bundle ωw\omega^{w} on a suitable rigid subset of X1(N)X_{1}(N) for every weight ww. (Andreatta, Iovita, and Stevens independently constructed similar sheaves using a different method in [1].) However, the subset on which Pilloni defines ωw\omega^{w} is not quite large enough: for ww analytic, ωw\omega^{w} is defined on the locus where the Hasse invariant has valuation less than p1p\frac{p-1}{p}, whereas Buzzard and Calegari require the forms they work with to be well-defined even on elliptic curves that are “too supersingular” (meaning they have Hasse invariant of valuation at least pp+1\frac{p}{p+1}).

In light of this, we begin by constructing extensions of Pilloni’s line bundles on higher-level modular curves of the form X0(pm):=X(Γ0(pm)Γ1(N))X_{0}(p^{m}):=X(\Gamma_{0}(p^{m})\cap\Gamma_{1}(N)) (which parametrize tuples (E,ψN,Cm)(E,\psi_{N},C^{m}) where EE is an elliptic curve, ψN\psi_{N} is a tame level structure, and CmC^{m} is a cyclic subgroup of EE of order pmp^{m}). These extensions are well-defined in particular over the closure of the locus where Cm[p]C^{m}[p] is canonical, which is sufficient for the application to properness, but in fact go even farther. For v[0,1]v\in[0,1], one can define rigid subsets X0(pm)[0,v]X_{0}(p^{m})[0,v] of X0(pm)X_{0}(p^{m}) such that X0(pm)[0,v]X0(pm)[0,v]X_{0}(p^{m})[0,v]\subset X_{0}(p^{m})[0,v^{\prime}] if v<vv<v^{\prime}, X0(pm)[0,v]X_{0}(p^{m^{\prime}})[0,v] is the preimage of X0(pm)[0,v]X_{0}(p^{m})[0,v] under the forgetful projection if m>mm^{\prime}>m, and v<1X0(pm)[0,v]\bigcup_{v<1}X_{0}(p^{m})[0,v] contains, among other things, the entire locus where EE is supersingular. Then we have the following.

Theorem 1.2.1 (Proposition 2.4.3).

For a fixed weight ww, there are cutoffs vm(0,1)v_{m}\in(0,1) such that ωw\omega^{w} is well-defined on X0(pm)[0,vm]X_{0}(p^{m})[0,v_{m}] for each mm and vm1v_{m}\to 1 as mm\to\infty.

While we do not need such aggressive extensions to prove properness, we hope that they may be of independent interest for other applications. For algebraic weights, Buzzard has used analytic continuations of overconvergent forms over the supersingular locus to establish modularity of Galois representations ([3]).

Once we can work on the closure of the canonical locus of X0(pm)X_{0}(p^{m}), we can use Buzzard and Calegari’s method; in fact, in Pilloni’s setup, the argument becomes even cleaner than theirs was originally. In particular, we have the following result about the injectivity of UpU_{p}.

Theorem 1.2.2 (Proposition 3.1.1).

Let w:p×p×w:\mathbb{Z}_{p}^{\times}\to\mathbb{C}_{p}^{\times} be an analytic weight. If fH0(X0(pm)[0,v],ωw)f\in H^{0}(X_{0}(p^{m})[0,v],\omega^{w}) for some v1p+1v\geq\frac{1}{p+1} and Upf0U_{p}f\equiv 0, then f0f\equiv 0.

Understanding these linear-algebraic properties of UpU_{p} on spaces of forms of large overconvergence radius is useful for many reasons. The Newton polygon upper bounds in Liu-Wan-Xiao’s proof of the spectral halo conjecture for definite quaternion algebras over \mathbb{Q} ([17]) and the author’s analysis of slopes in eigenvarieties for definite unitary groups of arbitrary rank ([21]) rely on checking that UpU_{p} is nonzero on certain classical subspaces. Calegari has conjectured (see [8]) the much stronger statement that in some cases UpU_{p} admits a convergent spectral expansion (this has only been proven for p=2p=2, N=1N=1, and w=0w=0, by Loeffler in [18]).

We hope that our new proof of the properness of the eigencurve along with its various intermediate results may expand the range of available techniques for analyzing the geometry of eigenvarieties. Note that Hattori [14] has successfully applied the Buzzard-Calegari method to some Hilbert modular eigenvarieties over algebraic weights. The author hopes to extend these methods to the question of properness for these or other higher-dimensional eigenvarieties in the future.

Verifying properness in other situations would likely have interesting additional implications for the structure of pp-adic eigenforms. One has to be cautious here; it is not true, for example, that limits of sequences of finite-slope modular forms must be finite-slope—counterexamples are constructed by Coleman and Stein in [10]. However, properness can be used to show, as Hattori and Newton do in [15], that an irreducible component of the Coleman-Mazur eigencurve of finite degree over weight space must actually be finite.

1.3 Organization

In Section 2, we construct our extended sheaves of overconvergent modular forms. These sheaves are given by functions on spaces of images of Hodge-Tate maps on Tate modules of elliptic curves. In Section 2.1, we compute the images of these Hodge-Tate maps. In Section 2.2, we construct the rigid spaces corresponding to these images. In Section 2.3, we define and describe the admissible open subsets over which our invertible sheaves will be defined. In Section 2.4, we construct our sheaves and show that they are invertible. In Section 2.5, we construct the UpU_{p} operator on our extended overconvergent forms.

In Section 3, we prove Theorem 1.1.1. In Section 3.1, we prove that infinite-slope forms must have relatively small overconvergence radius, whereas finite-slope forms must have relatively large overconvergence radius. In Section 3.2, we complete the proof by showing that families of finite-slope forms of large overconvergence radius over a punctured disc extend to a form of large overconvergence radius over the puncture.

Acknowledgments

I am very grateful to Richard Taylor for regular discussions and guidance on this topic, and to Vincent Pilloni for suggesting the method of extending sheaves of overconvergent modular forms by renormalizing the Hodge-Tate map. I would also like to thank Rebecca Bellovin, George Boxer, Kevin Buzzard, Frank Calegari, Mark Kisin, Joe Kramer-Miller, James Newton, James Upton, and Daqing Wan for valuable conversations about the Coleman-Mazur eigencurve.

This work was done during the support of the National Science Foundation Mathematical Sciences Postdoctoral Research Fellowship.

2 Sheaves of overconvergent modular forms on X0(pm)X_{0}(p^{m})

Our construction of overconvergent forms follows that of Pilloni, whose insight was as follows. As in the introduction, let N5N\geq 5 be an integer not divisible by pp, and write, for short, X=X(Γ1(N))X=X(\Gamma_{1}(N)) and X0(pm)=X(Γ1(N)Γ0(pm))X_{0}(p^{m})=X(\Gamma_{1}(N)\cap\Gamma_{0}(p^{m})). Let \mathcal{E} be the universal semi-abelian scheme over X0(pm)X_{0}(p^{m}), e:X0(pm)e:X_{0}(p^{m})\to\mathcal{E} be the identity section, and ω=eΩ/X0(pm)1\omega_{\mathcal{E}}=e^{*}\Omega^{1}_{\mathcal{E}/X_{0}(p^{m})}; more generally, any time we have a group GG over X0(pm)X_{0}(p^{m}), we will write e:X0(pm)Ge:X_{0}(p^{m})\to G for the identity section and ωG\omega_{G} for eΩG/X0(pm)1e^{*}\Omega_{G/X_{0}(p^{m})}^{1}. Let 𝒯=spec(Symω)\mathcal{T}=\operatorname{spec}(\operatorname{Sym}^{\bullet}\omega_{\mathcal{E}}^{\vee}) be the total space of ω\omega_{\mathcal{E}} over X0(pm)X_{0}(p^{m}), and 𝒯×\mathcal{T}^{\times} the open subset of nonzero sections.

Classical forms (resp. Katz overconvergent forms) of algebraic weight kk on XX are sections of ωk\omega_{\mathcal{E}}^{k} over XX (resp. a rigid subset of XX excluding small supersingular discs). Equivalently, they are functions on 𝒯\mathcal{T} whose restriction to any fiber of 𝒯\mathcal{T} over, say, a p\mathbb{C}_{p}-point of XX transforms by weight kk under multiplication by p×\mathbb{Z}_{p}^{\times}. If we work instead with a locally analytic weight w:p×p×w:\mathbb{Z}_{p}^{\times}\to\mathbb{C}_{p}^{\times}, there is no such thing as a function on p\mathbb{C}_{p} that transforms by weight ww, since ww does not converge on all of p\mathbb{C}_{p}. However, ww does converge on a union of balls in p\mathbb{C}_{p}. Thus we could define forms of weight ww if we replaced 𝒯\mathcal{T} by a smaller rigid space whose fibers over points of XX were unions of balls on which ww converged.

Pilloni shows that computing Hodge-Tate maps over points of XX naturally gives rise to a rigid space with the desired geometry. To extend his work, we begin by making a more detailed series of Hodge-Tate calculations.

2.1 Hodge-Tate maps on Tate modules of elliptic curves

Let EE be an elliptic curve over 𝒪p\mathcal{O}_{\mathbb{C}_{p}}, e:spec𝒪pEe:\operatorname{spec}\mathcal{O}_{\mathbb{C}_{p}}\to E its identity, ωE=eΩE/𝒪p1\omega_{E}=e^{*}\Omega_{E/\mathcal{O}_{\mathbb{C}_{p}}}^{1} its space of invariant differentials, and TpE=limnE[pn](𝒪p)T_{p}E=\varprojlim_{n}E[p^{n}](\mathcal{O}_{\mathbb{C}_{p}}) its Tate module (note that E[pn](𝒪p)=E[pn](p)E[p^{n}](\mathcal{O}_{\mathbb{C}_{p}})=E[p^{n}](\mathbb{C}_{p}) since E[pn]E[p^{n}] is finite flat over spec𝒪p\operatorname{spec}\mathcal{O}_{\mathbb{C}_{p}}). Let HT:TpEωEHT:T_{p}E\to\omega_{E} be the Hodge-Tate map, defined as follows: if x=(,xn,xn1,,x1)TpEx=(\dotsc,x_{n},x_{n-1},\dotsc,x_{1})\in T_{p}E (where xnE[pn](𝒪p)x_{n}\in E[p^{n}](\mathcal{O}_{\mathbb{C}_{p}}) and xn1=pxnx_{n-1}=px_{n}), let x:E[p]μpx^{\vee}:E[p^{\infty}]\to\mu_{p^{\infty}} be the map taking PE[pn]P\in E[p^{n}] to the Weil pairing of xnx_{n} and PP (also defined over 𝒪p\mathcal{O}_{\mathbb{C}_{p}}); then

HT(x)=(x)dTT.HT(x)=(x^{\vee})^{*}\frac{dT}{T}.

(The reader is warned that we have inserted a dualization into the usual definition of this map, because for our purposes it looks cleaner this way.)

Let vv be the standard valuation on p\mathbb{C}_{p} (so v(p)=1v(p)=1). We will compute v(HT(x))v(HT(x)) for any xTpEpTpEx\in T_{p}E\setminus pT_{p}E. Let h(E)h(E) be the valuation of the Hasse invariant of EE if that valuation lies in [0,1][0,1], and 11 otherwise. If h(E)>0h(E)>0, let n(E)0n(E)\in\mathbb{Z}_{\geq 0} be such that 1pn(E)1(p+1)h(E)<1pn(E)2(p+1)\frac{1}{p^{n(E)-1}(p+1)}\leq h(E)<\frac{1}{p^{n(E)-2}(p+1)}; if h(E)=0h(E)=0, let n(E)=n(E)=\infty.

Let n=n(E)n=n(E); recall that if n1n\geq 1, then EE has canonical subgroups Hcan1HcannH_{can}^{1}\subset\dotsc\subset H_{can}^{n}, where HcaniH_{can}^{i} has order pip^{i}, defined as follows. Hcan1H_{can}^{1} is the subgroup of order pp defined by Lubin and Katz in Chapter 3 of [16], and once Hcani1H_{can}^{i-1} has been defined, HcaniH_{can}^{i} is the preimage of Hcani1H_{can}^{i-1} under the isogeny EE/Hcan1E\mapsto E/H_{can}^{1}. When EE is ordinary, Hcan1H_{can}^{1} is the unique lift of the kernel of Frobenius on the reduction of EE, and HcannH_{can}^{n} exists for all nn.

Proposition 2.1.1.

Let xTpEpTpEx\in T_{p}E\setminus pT_{p}E. Let h=h(E)h=h(E) and n=n(E)n=n(E). Let 0an0\leq a\leq n be maximal such that ker(x)[pa]Hcana\ker(x^{\vee})[p^{a}]\subset H_{can}^{a}, if it exists; otherwise let a=a=\infty. Then

v(HT(x))={a+hp10h<pp+1,a<n<n+1p1pn+pn11p1h0h<pp+1,a=n<h=0,a=n=p(p1)(p+1)hpp+1.v(HT(x))=\begin{cases}a+\frac{h}{p-1}&0\leq h<\frac{p}{p+1},\qquad a<n<\infty\\ n+\frac{1}{p-1}-\frac{p^{n}+p^{n-1}-1}{p-1}h&0\leq h<\frac{p}{p+1},\qquad a=n<\infty\\ \infty&h=0,\qquad a=n=\infty\\ \frac{p}{(p-1)(p+1)}&h\geq\frac{p}{p+1}.\end{cases}
Remark 2.1.2.

Note that in the second case, plugging in 1pn1(p+1)h<1pn2(p+1)\frac{1}{p^{n-1}(p+1)}\leq h<\frac{1}{p^{n-2}(p+1)} gives

n1+hp1<n+1p1pn+pn11p1hn+hp1.n-1+\frac{h}{p-1}<n+\frac{1}{p-1}-\frac{p^{n}+p^{n-1}-1}{p-1}h\leq n+\frac{h}{p-1}.

In order to make this computation, we first need to understand the pp-torsion points of the completion E^\widehat{E} of EE at the identity. Recall that E^\widehat{E} is associated to a formal group over 𝒪p\mathcal{O}_{\mathbb{C}_{p}} of height 22 if EE is supersingular and 11 if EE is ordinary, and that we have an embedding E^[p](𝒪p)E[p](𝒪p)\widehat{E}[p^{\infty}](\mathcal{O}_{\mathbb{C}_{p}})\hookrightarrow E[p^{\infty}](\mathcal{O}_{\mathbb{C}_{p}}) which is an isomorphism if EE is supersingular and an isomorphism onto the canonical subgroup if EE is ordinary.

We will compute the valuation v(P)v(P) of a point PE^[p](𝒪p)P\in\widehat{E}[p^{\infty}](\mathcal{O}_{\mathbb{C}_{p}}), as defined in Section 1.2.1 of [13]: one chooses an isomorphism E^Spf(𝒪pT)\widehat{E}\xrightarrow{\sim}\operatorname{Spf}(\mathcal{O}_{\mathbb{C}_{p}}\llbracket T\rrbracket) taking the identity section of E^\widehat{E} to T=0T=0, and defines v(P)v(P) to be the valuation of the TT-coordinate T(P)T(P) of the image of PP under this isomorphism. The choice of isomorphism does not matter because a ring automorphism of 𝒪pT\mathcal{O}_{\mathbb{C}_{p}}\llbracket T\rrbracket fixing the ideal (T)(T) must fix valuations of TT-coordinates.

When convenient, we will also use PP to refer to the image of PP in E[p](𝒪p)E[p^{\infty}](\mathcal{O}_{\mathbb{C}_{p}}).

Lemma 2.1.3.

Let PE^[pk]E^[pk1]P\in\widehat{E}[p^{k}]\setminus\widehat{E}[p^{k-1}]. Let h=h(E)h=h(E) and n=n(E)n=n(E).

  1. 1.

    Assume 0<h<pp+10<h<\frac{p}{p+1}.

    1. (a)

      If knk\leq n, then

      v(P)={1pk1hpk1(p1)PHcankHcank1hp(p1)not the above, but pPHcank1Hcank2hp2k2a1(p1)not the above, but pkaPHcanaHcana1hp2k1(p1)pk1PHcan1.v(P)=\begin{cases}\frac{1-p^{k-1}h}{p^{k-1}(p-1)}&P\in H_{can}^{k}\setminus H_{can}^{k-1}\\ \frac{h}{p(p-1)}&\text{not the above, but }pP\in H_{can}^{k-1}\setminus H_{can}^{k-2}\\ \vdots&\vdots\\ \frac{h}{p^{2k-2a-1}(p-1)}&\text{not the above, but }p^{k-a}P\in H_{can}^{a}\setminus H_{can}^{a-1}\\ \vdots&\vdots\\ \frac{h}{p^{2k-1}(p-1)}&p^{k-1}P\notin H_{can}^{1}.\end{cases}
    2. (b)

      If kn+1k\geq n+1, then

      v(P)={1pn1hp2kn1(p1)pknPHcannHcann1hp1+2k2n(p1)not the above, but pkn+1PHcann1Hcann2hp2k2a1(p1)not the above, but pkaPHcanaHcana1hp2k1(p1)pk1PHcan1.v(P)=\begin{cases}\frac{1-p^{n-1}h}{p^{2k-n-1}(p-1)}&p^{k-n}P\in H_{can}^{n}\setminus H_{can}^{n-1}\\ \frac{h}{p^{1+2k-2n}(p-1)}&\text{not the above, but }p^{k-n+1}P\in H_{can}^{n-1}\setminus H_{can}^{n-2}\\ \vdots&\vdots\\ \frac{h}{p^{2k-2a-1}(p-1)}&\text{not the above, but }p^{k-a}P\in H_{can}^{a}\setminus H_{can}^{a-1}\\ \vdots&\vdots\\ \frac{h}{p^{2k-1}(p-1)}&p^{k-1}P\notin H_{can}^{1}.\end{cases}
  2. 2.

    Assume hpp+1h\geq\frac{p}{p+1}. Then v(P)=1p2k2(p21).v(P)=\frac{1}{p^{2k-2}(p^{2}-1)}.

  3. 3.

    Assume h=0h=0. Then necessarily PHcankHcank1P\in H_{can}^{k}\setminus H_{can}^{k-1}, and v(P)=1pk1(p1)v(P)=\frac{1}{p^{k-1}(p-1)}.

Proof.
  1. 1.

    We will check this by induction, simultaneously with the following statements:

    —if knk\leq n and pPHcank1Hcank2pP\in H_{can}^{k-1}\setminus H_{can}^{k-2}, then the possible valuations of PP are given by the Newton polygon with vertices (0,1pk1hpk1(p1)),(p,h),(p2,0)\left(0,\frac{1-p^{k-1}h}{p^{k-1}(p-1)}\right),(p,h),(p^{2},0).

    —if Hcank1H_{can}^{k-1} does not exist or pPHcank1pP\notin H_{can}^{k-1}, then the possible valuations of PP are given by the Newton polygon (line segment) with vertices (0,v(pP)),(p2,0)(0,v(pP)),(p^{2},0).

    First we check the base case k=1k=1. From the proof of Theorem 3.10.7 of [16], we see that the Newton polygon of the power series [p](T)[p](T) associated to multiplication by pp on E^\widehat{E} is given by the points (0,),(1,1),(p,h),(p2,0)(0,\infty),(1,1),(p,h),(p^{2},0). By definition, the elements of E^[p]\widehat{E}[p] are the roots of [p](T)[p](T), so for PE^[p]P\in\widehat{E}[p], we have

    v(P)={P=01hp1PHcan1{0}hp(p1)PHcan1.v(P)=\begin{cases}\infty&P=0\\ \frac{1-h}{p-1}&P\in H_{can}^{1}\setminus\{0\}\\ \frac{h}{p(p-1)}&P\notin H_{can}^{1}.\end{cases}

    More generally, as at the beginning of Section 1.3 of [13], if pPE^(𝒪p)pP\in\widehat{E}(\mathcal{O}_{\mathbb{C}_{p}}), the possible valuations of PP are given by the Newton polygon obtained as the convex hull of (0,v(pP)),(1,1),(p,h),(p2,0)(0,v(pP)),(1,1),(p,h),(p^{2},0) (since the possible choices of PP are the roots of the power series [p](T)T(pP)[p](T)-T(pP)).

    Now assume the statements for k1k-1. First suppose knk\leq n. Then the possible Newton polygons for [p](T)T(pP)[p](T)-T(pP) are given by the convex hulls of each of

    (0,1pk2hpk2(p1)),(0,hp(p1)),,(0,hp2k3p1)\left(0,\frac{1-p^{k-2}h}{p^{k-2}(p-1)}\right),\left(0,\frac{h}{p(p-1)}\right),\dotsc,\left(0,\frac{h}{p^{2k-3}{p-1}}\right)

    together with (p,h),(p2,0)(p,h),(p^{2},0). The slope from (0,1pk2hpk2(p1))\left(0,\frac{1-p^{k-2}h}{p^{k-2}(p-1)}\right) to (p,h)(p,h) is

    1pk2hpk2(p1)hp=1pk2hpk2(p1)hpk1(p1)=1pk1hpk1(p1).\frac{\frac{1-p^{k-2}h}{p^{k-2}(p-1)}-h}{p}=\frac{1-p^{k-2}h-p^{k-2}(p-1)h}{p^{k-1}(p-1)}=\frac{1-p^{k-1}h}{p^{k-1}(p-1)}.

    This is greater than the slope from (p,h)(p,h) to (p2,0)(p^{2},0) if and only if

    1pk1hpk1(p1)\displaystyle\frac{1-p^{k-1}h}{p^{k-1}(p-1)} >hp(p1)\displaystyle>\frac{h}{p(p-1)}
    1pk1h\displaystyle 1-p^{k-1}h >pk2h\displaystyle>p^{k-2}h
    1pk2(p+1)=1pk1+pk2\displaystyle\frac{1}{p^{k-2}(p+1)}=\frac{1}{p^{k-1}+p^{k-2}} >h\displaystyle>h

    which is true because knk\leq n. This gives us the statement about the Newton polygon in the case pPHcank1pP\in H_{can}^{k-1}, and thus also the first two cases for v(P)v(P) (that is, PHcankP\in H_{can}^{k} and PHcankP\notin H_{can}^{k} but pPHcank1pP\in H_{can}^{k-1}). The slope from (0,hp(p1))\left(0,\frac{h}{p(p-1)}\right) to (p2,0)(p^{2},0) is hp3(p1)\frac{h}{p^{3}(p-1)}, which is always less than hp(p1)\frac{h}{p(p-1)}, the slope from (p,h)(p,h) to (p2,0)(p^{2},0); this gives us the statement about the Newton polygon in the case pPHcank1pP\notin H_{can}^{k-1}. Therefore the remaining cases for v(P)v(P) can be obtained by taking each of hp(p1),,hp2k3(p1)\frac{h}{p(p-1)},\dotsc,\frac{h}{p^{2k-3}(p-1)} and dividing by p2p^{2}.

    Now suppose kn+1k\geq n+1. Then everything is the same except that even in the case pknPHcannp^{k-n}P\in H_{can}^{n}, the Newton polygon for PP given pPpP is a single line segment, with slope 1pn1hp2kn1(p1)\frac{1-p^{n-1}h}{p^{2k-n-1}(p-1)}.

  2. 2.

    If hpp+1h\geq\frac{p}{p+1}, the Newton polygon of [p](T)[p](T) has vertices (0,),(1,1),(p2,0)(0,\infty),(1,1),(p^{2},0). The claim follows from induction simultaneously with the statement that if PE^[pk]E^[pk1]P\in\widehat{E}[p^{k}]\setminus\widehat{E}[p^{k-1}] for k2k\geq 2, the Newton polygon of [p](T)T(pP)[p](T)-T(pP) is the single line segment from (0,v(pP))(0,v(pP)) to (p2,0)(p^{2},0).

  3. 3.

    If h=0h=0, the Newton polygon of [p](T)[p](T) has vertices (0,),(1,1),(p,0)(0,\infty),(1,1),(p,0). The claim follows from induction simultaneously with the statement that if PE^[pk]E^[pk1]P\in\widehat{E}[p^{k}]\setminus\widehat{E}[p^{k-1}] for k2k\geq 2, the Newton polygon of [p](T)X(pP)[p](T)-X(pP) is the single line segment from (0,1pk2(p1))\left(0,\frac{1}{p^{k-2}(p-1)}\right) to (p,0)(p,0).

With this calculation in hand, we may apply formulas from [13] to compute the values of v(HT)v(HT) asserted in Proposition 2.1.1.

Proof of Proposition 2.1.1.

First assume 0<h<pp+10<h<\frac{p}{p+1}. Choose any kn+1k\geq n+1 and let PE[pk]E[pk1]P\in E[p^{k}]\setminus E[p^{k-1}] be an 𝒪p\mathcal{O}_{\mathbb{C}_{p}}-point in the kernel of x:E[p]μpx^{\vee}:E[p^{\infty}]\to\mu_{p^{\infty}}. Suppose aa is maximal such that pkaPHcanap^{k-a}P\in H_{can}^{a}. Identifying PP with its preimage in E^[pk]\widehat{E}[p^{k}], Theorem 1.11 of [13] says that

v(HT(x))\displaystyle v(HT(x)) =pp1QP{0}v(Q)1p1QpP{0}v(Q)\displaystyle=\frac{p}{p-1}\sum_{Q\in\langle P\rangle\setminus\{0\}}v(Q)-\frac{1}{p-1}\sum_{Q\in\langle pP\rangle\setminus\{0\}}v(Q)
=pp1QPpPv(Q)+QpP{0}v(Q).\displaystyle=\frac{p}{p-1}\sum_{Q\in\langle P\rangle\setminus\langle pP\rangle}v(Q)+\sum_{Q\in\langle pP\rangle\setminus\{0\}}v(Q).

Split the second sum further to get

v(HT(x))=pp1QPpPv(Q)+QpPpkaPv(Q)+QpkaP{0}v(Q).v(HT(x))=\frac{p}{p-1}\sum_{Q\in\langle P\rangle\setminus\langle pP\rangle}v(Q)+\sum_{Q\in\langle pP\rangle\setminus\langle p^{k-a}P\rangle}v(Q)+\sum_{Q\in\langle p^{k-a}P\rangle\setminus\{0\}}v(Q).

First assume a<na<n. Then the first sum is

pp1(pkpk1)hp2k12a(p1)=hpk12a(p1),\frac{p}{p-1}\cdot(p^{k}-p^{k-1})\cdot\frac{h}{p^{2k-1-2a}(p-1)}=\frac{h}{p^{k-1-2a}(p-1)},

the second is

(pk1pk2)hp2k32a(p1)++(pa+1pa)hp(p1)=hpk12a++hp1a(p^{k-1}-p^{k-2})\cdot\frac{h}{p^{2k-3-2a}(p-1)}+\dotsb+(p^{a+1}-p^{a})\cdot\frac{h}{p(p-1)}=\frac{h}{p^{k-1-2a}}+\dotsb+\frac{h}{p^{1-a}}
=hpk12a(1++pka2)=hpk12apk1a1p1,=\frac{h}{p^{k-1-2a}}(1+\dotsb+p^{k-a-2})=\frac{h}{p^{k-1-2a}}\cdot\frac{p^{k-1-a}-1}{p-1},

and the third is

(papa1)1pa1hpa1(p1)+(pa1pa2)1pa2hpa2(p1)++(p1)1hp1(p^{a}-p^{a-1})\cdot\frac{1-p^{a-1}h}{p^{a-1}(p-1)}+(p^{a-1}-p^{a-2})\cdot\frac{1-p^{a-2}h}{p^{a-2}(p-1)}+\dotsb+(p-1)\cdot\frac{1-h}{p-1}
=(1pa1h)+(1pa2h)++(1h)=a(pa1++1)h=apa1p1h.=(1-p^{a-1}h)+(1-p^{a-2}h)+\dotsb+(1-h)=a-(p^{a-1}+\dotsb+1)h=a-\frac{p^{a}-1}{p-1}h.

This gives a total of

hpk12a(p1)+hpk12apk1a1p1+apa1p1h=pk1ahpk12a(p1)+apa1p1h\frac{h}{p^{k-1-2a}(p-1)}+\frac{h}{p^{k-1-2a}}\cdot\frac{p^{k-1-a}-1}{p-1}+a-\frac{p^{a}-1}{p-1}h=\frac{p^{k-1-a}h}{p^{k-1-2a}(p-1)}+a-\frac{p^{a}-1}{p-1}h
=pahp1+apa1p1h=a+hp1.=\frac{p^{a}h}{p-1}+a-\frac{p^{a}-1}{p-1}h=a+\frac{h}{p-1}.

On the other hand if a=na=n, we have

pp1QPpPv(Q)=pp1(pkpk1)1pn1hp2kn1(p1)=1pn1hpkn1(p1)\frac{p}{p-1}\sum_{Q\in\langle P\rangle\setminus\langle pP\rangle}v(Q)=\frac{p}{p-1}\cdot(p^{k}-p^{k-1})\cdot\frac{1-p^{n-1}h}{p^{2k-n-1}(p-1)}=\frac{1-p^{n-1}h}{p^{k-n-1}(p-1)}

and

QpPpkaPv(Q)=(pk1pk2)1pn1hp2k3n(p1)++(pa+1pa)1pn1hpn+1(p1)\sum_{Q\in\langle pP\rangle\setminus\langle p^{k-a}P\rangle}v(Q)=(p^{k-1}-p^{k-2})\cdot\frac{1-p^{n-1}h}{p^{2k-3-n}(p-1)}+\dotsb+(p^{a+1}-p^{a})\cdot\frac{1-p^{n-1}h}{p^{n+1}(p-1)}
=1pn1hpkn1++1pn1hp=1pn1hpkn1(1++pk2n)=(1pn1h)(pk1n1)pkn1(p1)=\frac{1-p^{n-1}h}{p^{k-n-1}}+\dotsb+\frac{1-p^{n-1}h}{p}=\frac{1-p^{n-1}h}{p^{k-n-1}}(1+\dotsb+p^{k-2-n})=\frac{(1-p^{n-1}h)(p^{k-1-n}-1)}{p^{k-n-1}(p-1)}

giving a total of

1pn1hpkn1(p1)+(1pn1h)(pkn11)pkn1(p1)+npn1p1h\frac{1-p^{n-1}h}{p^{k-n-1}(p-1)}+\frac{(1-p^{n-1}h)(p^{k-n-1}-1)}{p^{k-n-1}(p-1)}+n-\frac{p^{n}-1}{p-1}h
=1pn1hp1+npn1p1h=n+1p1pn+pn11p1h.=\frac{1-p^{n-1}h}{p-1}+n-\frac{p^{n}-1}{p-1}h=n+\frac{1}{p-1}-\frac{p^{n}+p^{n-1}-1}{p-1}h.

These give the asserted values of v(HT(x))v(HT(x)) for 0<h<pp+10<h<\frac{p}{p+1}. Next, if hpp+1h\geq\frac{p}{p+1}, then for any kk and any PE[pk]P\in E[p^{k}] as before, we have

pp1QPpPv(Q)=pp1(pkpk1)1p2k2(p21)=1pk2(p21)\frac{p}{p-1}\sum_{Q\in\langle P\rangle\setminus\langle pP\rangle}v(Q)=\frac{p}{p-1}\cdot(p^{k}-p^{k-1})\cdot\frac{1}{p^{2k-2}(p^{2}-1)}=\frac{1}{p^{k-2}(p^{2}-1)}
QpP{0}v(Q)=(pk1pk2)1p2k4(p21)++(p1)1p21\sum_{Q\in\langle pP\rangle\setminus\{0\}}v(Q)=(p^{k-1}-p^{k-2})\frac{1}{p^{2k-4}(p^{2}-1)}+\dotsb+(p-1)\frac{1}{p^{2}-1}
=1pk2(p+1)++1p+1=1pk2(p+1)(1++pk2)=pk11pk2(p21)=\frac{1}{p^{k-2}(p+1)}+\dotsb+\frac{1}{p+1}=\frac{1}{p^{k-2}(p+1)}(1+\dotsb+p^{k-2})=\frac{p^{k-1}-1}{p^{k-2}(p^{2}-1)}

giving us a total of

1pk2(p21)+pk11pk2(p21)=pk1pk2(p21)=pp21\frac{1}{p^{k-2}(p^{2}-1)}+\frac{p^{k-1}-1}{p^{k-2}(p^{2}-1)}=\frac{p^{k-1}}{p^{k-2}(p^{2}-1)}=\frac{p}{p^{2}-1}

as claimed.

Finally, we consider the case h=0h=0. We cannot use Fargues’s formula directly in this case, because it only applies to pp-divisible groups coming from formal groups (i.e. connected ones), which E[p]E[p^{\infty}] is not. However, we can follow most of the same calculations as in Section 1.2.2 of [13]. In particular, if PE[pk]E^[pk]P\in E[p^{k}]\setminus\widehat{E}[p^{k}], define v(P)=0v(P)=0. Then Proposition 1.2 of [13] still holds, since if DE[p]D\subset E[p^{\infty}] is a finite locally free subgroup, ωD\omega_{D} depends only on the connected component of the identity in DD.

With this definition of v(P)v(P), using the same calculations as those leading up to Proposition 1.6 of [13], we conclude the following: if there exists an integer kk such that for an 𝒪p\mathcal{O}_{\mathbb{C}_{p}}-point PE[pk]E[pk1]P\in E[p^{k}]\setminus E[p^{k-1}] in the kernel of ker(x:E[p]μp)\ker(x^{\vee}:E[p^{\infty}]\to\mu_{p^{\infty}}), we have

QPpPv(Q)<p1p,\sum_{Q\in\langle P\rangle\setminus\langle pP\rangle}v(Q)<\frac{p-1}{p},

then Fargues’s formula for v(HT(x))v(HT(x)) holds; otherwise, HT(x)HT(x) must be 0 in ωE/pkωE\omega_{E}/p^{k}\omega_{E} for all kk, and hence must be 0 in ωE\omega_{E}. When aa exists, the first situation happens, and the same calculation as when 0<h<pp+10<h<\frac{p}{p+1} holds. When a=a=\infty, the second situation happens, and we have v(HT(x))=v(HT(x))=\infty. To avoid a long, irrelevant digression, details are left as an exercise to the reader. ∎

Let ^\widehat{\mathcal{E}} be a formal model for EE over 𝒪p\mathcal{O}_{\mathbb{C}_{p}}. Let ωEint\omega_{E}^{int} be the pullback of the sheaf of integral differential forms on ^\widehat{\mathcal{E}}, so that ωEint𝒪p\omega_{E}^{int}\cong\mathcal{O}_{\mathbb{C}_{p}}. Let ωE+\omega_{E}^{+} be the 𝒪p\mathcal{O}_{\mathbb{C}_{p}}-submodule of ωEint\omega_{E}^{int} generated by the image of HT:TpEωEintHT:T_{p}E\to\omega_{E}^{int}. It will be useful to work with ωE+\omega_{E}^{+} while constructing the fiber bundle of unions of balls on which our forms will be defined.

Proposition 2.1.4.

We have

ωE+={php1ωEinth<pp+1pp(p1)(p+1)ωEinthpp+1.\omega_{E}^{+}=\begin{cases}p^{\frac{h}{p-1}}\omega_{E}^{int}&h<\frac{p}{p+1}\\ p^{\frac{p}{(p-1)(p+1)}}\omega_{E}^{int}&h\geq\frac{p}{p+1}.\end{cases}
Proof.

If h<pp+1h<\frac{p}{p+1}, our formula for v(HT(x))v(HT(x)) (Proposition 2.1.1) is minimized when a=0a=0, in which case v(HT(x))=hp1v(HT(x))=\frac{h}{p-1}. (In particular, by Remark 2.1.2, we have

hp1n1+hp1<n+1p1pn+pn11p1h\frac{h}{p-1}\leq n-1+\frac{h}{p-1}<n+\frac{1}{p-1}-\frac{p^{n}+p^{n-1}-1}{p-1}h

so a=0a=0 gives a lower valuation than a=na=n.) ∎

Remark 2.1.5.

One can also directly check that ωE+\omega_{E}^{+} is generated by HT(P)HT(P) for any PHcan1{0}P\in H_{can}^{1}\setminus\{0\} without using most of Proposition 2.1.1. This is because the argument in Section 1.2.2 of [13] shows that for PE[p]P\in E[p], if

pp1QP{0}v(Q)<1,\frac{p}{p-1}\sum_{Q\in\langle P^{\vee}\rangle\setminus\{0\}}v(Q)<1,

then v(HT(P))v(HT(P)) equals the LHS, and otherwise v(HT(P))1v(HT(P))\geq 1. But the LHS can also be written as pp1(p1)v(P)=pv(P)\frac{p}{p-1}\cdot(p-1)v(P^{\vee})=pv(P^{\vee}), which is clearly minimized when PHcan1{0}P\in H_{can}^{1}\setminus\{0\}, because Lubin and Katz actually define Hcan1H_{can}^{1} in [16] as the subset of E[p]E[p] of largest slope. In this case pv(P)=php(p1)=hp1<1pv(P^{\vee})=p\cdot\frac{h}{p(p-1)}=\frac{h}{p-1}<1 as in the base case of Lemma 2.1.3, so we are done.

For any mm, since the map HT:TpEωE+HT:T_{p}E\to\omega_{E}^{+} is p\mathbb{Z}_{p}-linear, it descends to a map E[pm]=TpE/pmTpEωE+/pmωE+E[p^{m}]=T_{p}E/p^{m}T_{p}E\to\omega_{E}^{+}/p^{m}\omega_{E}^{+} which we will also refer to by HTHT. This “renormalized” Hodge-Tate map, a lift of the standard map E[pm]ωE/pmωEE[p^{m}]\to\omega_{E}/p^{m}\omega_{E}, is key to constructing a fibration of unions of balls small enough to extend our modular sheaves. We are grateful to Vincent Pilloni for suggesting this renormalization.

2.2 Spaces of Hodge-Tate images over X0(pm)X_{0}(p^{m})

Let \mathcal{E} be the universal semi-abelian scheme over X0(pm)X_{0}(p^{m}), 𝒞m\mathcal{C}^{m} the universal cyclic subgroup of order pmp^{m}, and 𝒞m,×\mathcal{C}^{m,\times} the subset of generating elements. Let ^\widehat{\mathcal{E}}, 𝒳^0(pm)\widehat{\mathcal{X}}_{0}(p^{m}), 𝒞^m\widehat{\mathcal{C}}^{m}, and 𝒞^m,×\widehat{\mathcal{C}}^{m,\times} be their formal completions along their special fibers.

Let 𝒯^\widehat{\mathcal{T}} and 𝒯^×\widehat{\mathcal{T}}^{\times} be the formal completions of 𝒯\mathcal{T} and 𝒯×\mathcal{T}^{\times} along their special fibers, 𝒯rig\mathcal{T}_{rig} and 𝒯rig×\mathcal{T}_{rig}^{\times} the rigid generic fibers of 𝒯^\widehat{\mathcal{T}} and 𝒯^×\widehat{\mathcal{T}}^{\times}, and 𝒯an\mathcal{T}_{an} and 𝒯an×\mathcal{T}_{an}^{\times} the analytifications of 𝒯\mathcal{T} and 𝒯×\mathcal{T}^{\times}. (Recall that the fibers of 𝒯rig\mathcal{T}_{rig} over X0(pm)X_{0}(p^{m}) are unit balls, with natural inclusions into the fibers of 𝒯an\mathcal{T}_{an} over X0(pm)X_{0}(p^{m}), which are affine lines; see e.g. Section 3.3 of [11].)

Proposition 2.2.1.

There exists a unique rigid open subset Im𝒞m×X0(pm)𝒯rigI^{m}\hookrightarrow\mathcal{C}^{m}\times_{X_{0}(p^{m})}\mathcal{T}_{rig} such that for each finite extension KK of p\mathbb{Q}_{p} and each point (E,ψN,Cm)X0(pm)(K)(E,\psi_{N},C^{m})\in X_{0}(p^{m})(K), where EE is a semi-abelian scheme of dimension 11 over 𝒪K\mathcal{O}_{K}, ψN\psi_{N} is a tame level structure on EE, and CmC^{m} is a cyclic order-pmp^{m} subgroup of EE, we have

Im|(E,ψN,Cm)(p)={(P,s)Cm(p)×eΩE/𝒪p1HT(P)=s|Cm}.I^{m}|_{(E,\psi_{N},C^{m})}(\mathbb{C}_{p})=\{(P,s)\in C^{m}(\mathbb{C}_{p})\times e^{*}\Omega_{E/\mathcal{O}_{\mathbb{C}_{p}}}^{1}\mid HT(P)=s|_{C^{m}}\}.
Proof.

We follow the proof of Theorem 3.1 of [19]. Let Spf(A)\operatorname{Spf}(A) be a formal open affine of 𝒳^0(pm)\widehat{\mathcal{X}}_{0}(p^{m}) small enough to trivialize 𝒯^\widehat{\mathcal{T}}. Let ss be a generating section of 𝒯^\widehat{\mathcal{T}} over Spf(A)\operatorname{Spf}(A), so that ω^=As\omega_{\widehat{\mathcal{E}}}=As. The conormal exact sequence

eΩ(^/𝒞^m)/𝒳^0(pm)1eΩ^/𝒳^0(pm)1eΩ𝒞^m/𝒳^0(pm)10e^{*}\Omega_{(\widehat{\mathcal{E}}/\widehat{\mathcal{C}}^{m})/\widehat{\mathcal{X}}_{0}(p^{m})}^{1}\to e^{*}\Omega_{\widehat{\mathcal{E}}/\widehat{\mathcal{X}}_{0}(p^{m})}^{1}\to e^{*}\Omega_{\widehat{\mathcal{C}}^{m}/\widehat{\mathcal{X}}_{0}(p^{m})}^{1}\to 0

gives a surjection ω^ω𝒞^m\omega_{\widehat{\mathcal{E}}}\to\omega_{\widehat{\mathcal{C}}^{m}}, so over Spf(A)\operatorname{Spf}(A), we have ω𝒞^m=(A/aA)s\omega_{\widehat{\mathcal{C}}^{m}}=(A/aA)s for some aAa\in A. Let RR be the AA-algebra giving rise to the map 𝒞^mSpf(A)\widehat{\mathcal{C}}^{m}\to\operatorname{Spf}(A). Let PunivP_{univ} be the universal order-pmp^{m} point over 𝒞^m\widehat{\mathcal{C}}^{m}, and HTuniv=HT(Puniv)eΩ𝒞^m×𝒞^m/𝒞^m1HT_{univ}=HT(P_{univ})\in e^{*}\Omega_{\widehat{\mathcal{C}}^{m}\times\widehat{\mathcal{C}}^{m}/\widehat{\mathcal{C}}^{m}}^{1}. Then eΩ𝒞^m×𝒞^m/𝒞^m1=(R/aR)se^{*}\Omega_{\widehat{\mathcal{C}}^{m}\times\widehat{\mathcal{C}}^{m}/\widehat{\mathcal{C}}^{m}}^{1}=(R/aR)s, and HTunivHT_{univ} is an element of (R/aR)s(R/aR)s. The total space of eΩ^×𝒞^m/𝒞^m1e^{*}\Omega_{\widehat{\mathcal{E}}\times\widehat{\mathcal{C}}^{m}/\widehat{\mathcal{C}}^{m}}^{1} can be written as SpfRT1\operatorname{Spf}R\langle T_{1}\rangle where the section T1=1T_{1}=1 corresponds to ss. Then the rigid generic fiber of Spf(RT1,T2/(T1HTunivaT2))\operatorname{Spf}(R\langle T_{1},T_{2}\rangle/(T_{1}-HT_{univ}-aT_{2})) satisfies the desired property of ImI^{m} over the rigid generic fiber of Spf(A)\operatorname{Spf}(A). ∎

Definition 2.2.2.

Let Im,×=Im×𝒞m(𝒞m)×I^{m,\times}=I^{m}\times_{\mathcal{C}^{m}}(\mathcal{C}^{m})^{\times}. Let πX:Im,×X0(pm)\pi_{X}:I^{m,\times}\to X_{0}(p^{m}) and π𝒯:Im,×𝒯rig\pi_{\mathcal{T}}:I^{m,\times}\to\mathcal{T}_{rig} be the natural projections.

2.3 Admissible open subsets of X0(pm)X_{0}(p^{m})

We now set up some notation for the admissible open subsets of X0(pm)X_{0}(p^{m}) over which our sheaves of modular forms will be defined. Recall the following function on X0(p)X_{0}(p) from [3].

Definition 2.3.1 (Buzzard).

Let (E,ψN,C)X0(p)(E,\psi_{N},C)\in X_{0}(p). Let

v(E,ψN,C)={0h=0 and C=Hcan1hh<pp+1 and C=Hcan1pp+1hpp+11hp=1h(E/C)h<pp+1 and CHcan11h=0 and CHcan1.v(E,\psi_{N},C)=\begin{cases}0&h=0\text{ and }C=H_{can}^{1}\\ h&h<\frac{p}{p+1}\text{ and }C=H_{can}^{1}\\ \frac{p}{p+1}&h\geq\frac{p}{p+1}\\ 1-\frac{h}{p}=1-h(E/C)&h<\frac{p}{p+1}\text{ and }C\neq H_{can}^{1}\\ 1&h=0\text{ and }C\neq H_{can}^{1}.\end{cases}

Recall that X0(p)X_{0}(p) can be constructed geometrically by taking two copies of the locus XordX^{ord} of XX where EE is ordinary, one corresponding to the locus X0(p)ord,canX_{0}(p)^{ord,can} of X0(p)X_{0}(p) where EE is ordinary and CC is canonical, the other corresponding to the locus X0(p)ord,etX_{0}(p)^{ord,et} of X0(p)X_{0}(p) where EE is ordinary and CC is not canonical, and connecting them along their missing supersingular discs with supersingular “tubes”; then Buzzard’s function vv measures the location of a supersingular point along a supersingular tube, with vv increasing from 0 to 11 as one moves away from X0(p)ord,canX_{0}(p)^{ord,can} and toward X0(p)ord,etX_{0}(p)^{ord,et}.

For an interval I[0,1]I\subset[0,1], let X0(p)(I)X_{0}(p)(I) be the subset of X0(p)X_{0}(p) on which v(E,ψN,C)Iv(E,\psi_{N},C)\in I. Then, for example, X0(p){0}=X0(p)ord,canX_{0}(p)\{0\}=X_{0}(p)^{ord,can}; X0(p)[0,p/(p+1)]X_{0}(p)[0,p/(p+1)] is the locus where either CC is canonical, or EE has no canonical subgroup and CC is arbitrary; X0(p)[p/(p+1),1]X_{0}(p)[p/(p+1),1] is the locus where CC is not canonical; and X0(p){1}=X0(p)ord,etX_{0}(p)\{1\}=X_{0}(p)^{ord,et}. See [3] for more details.

Let X0(pm)(I)X_{0}(p^{m})(I) be the preimage of X0(p)(I)X_{0}(p)(I) under the projection map X0(pm)X0(p)X_{0}(p^{m})\to X_{0}(p) taking (E,ψN,Cm)(E,\psi_{N},C^{m}) to (E,ψN,Cm[p])(E,\psi_{N},C^{m}[p]). Then, for example, if v>pp+1v>\frac{p}{p+1}, X0(pm)[0,v]X_{0}(p^{m})[0,v] is the locus of (E,ψN,Cm)(E,\psi_{N},C^{m}) such that either Cm[p]C^{m}[p] is canonical, or Cm[p]C^{m}[p] is not canonical but h(E)p(1v)h(E)\geq p(1-v).

2.4 Invertible sheaves of locally analytic weight

We are almost ready to define our sheaves of modular forms of locally analytic weight ww on X0(pm)X_{0}(p^{m}). In order to do so, we have to make sure that ww converges on the image of Im,×I^{m,\times} in 𝒯rig\mathcal{T}_{rig} under π𝒯\pi_{\mathcal{T}}.

For each c>0c\in\mathbb{Z}_{>0}, choose a section (xlc)l(/pc)×(x_{l}^{c})_{l\in(\mathbb{Z}/p^{c}\mathbb{Z})^{\times}} of p×(/pc)×\mathbb{Z}_{p}^{\times}\to(\mathbb{Z}/p^{c}\mathbb{Z})^{\times}. For short, we will write v(HT(Cm))v(HT(C^{m})) for the valuation of HTHT of a generating element of CmC^{m}.

Proposition 2.4.1.

Let (E,ψN,Cm)X0(pm)(𝒪p)(E,\psi_{N},C^{m})\in X_{0}(p^{m})(\mathcal{O}_{\mathbb{C}_{p}}), where EE is an elliptic curve, ψN\psi_{N} a tame level structure, and CmC^{m} a subgroup of EE of order pmp^{m}. Choose a trivialization 𝒯rig|(E,ψN,Cm)𝒪p\mathcal{T}_{rig}|_{(E,\psi_{N},C^{m})}\cong\mathcal{O}_{\mathbb{C}_{p}}.

  1. 1.

    If (E,ψN,Cm)X0(pm)[0,v](E,\psi_{N},C^{m})\in X_{0}(p^{m})[0,v] for some v<11pm(p+1)v<1-\frac{1}{p^{m}(p+1)}, the image of π𝒯:Im,×𝒯rig\pi_{\mathcal{T}}:I^{m,\times}\to\mathcal{T}_{rig} over (E,ψN,Cm)(E,\psi_{N},C^{m}) does not contain 0.

  2. 2.

    If (E,ψN,Cm)X0(pm)[0,v](E,\psi_{N},C^{m})\in X_{0}(p^{m})[0,v] for some v<11pmc+1(p+1)v<1-\frac{1}{p^{m-c+1}(p+1)}, the image of π𝒯:Im,×𝒯rig\pi_{\mathcal{T}}:I^{m,\times}\to\mathcal{T}_{rig} over (E,ψN,Cm)(E,\psi_{N},C^{m}) is contained in

    l(/pc)×pv(HT(Cm))(xlc+pu𝒪p)\coprod_{l\in(\mathbb{Z}/p^{c}\mathbb{Z})^{\times}}p^{v(HT(C^{m}))}(x_{l}^{c}+p^{u}\mathcal{O}_{\mathbb{C}_{p}})

    for some u>c1u>c-1.

  3. 3.

    If (E,ψN,Cm)X0(pm)[0,pp+1](E,\psi_{N},C^{m})\in X_{0}(p^{m})\left[0,\frac{p}{p+1}\right], the image of π𝒯:Im,×𝒯rig\pi_{\mathcal{T}}:I^{m,\times}\to\mathcal{T}_{rig} over (E,ψN,Cm)(E,\psi_{N},C^{m}) can be written in the form

    l(/pm)×php1(xlm+pm𝒪p).\coprod_{l\in(\mathbb{Z}/p^{m}\mathbb{Z})^{\times}}p^{\frac{h}{p-1}}(x_{l}^{m}+p^{m}\mathcal{O}_{\mathbb{C}_{p}}).
Proof.

Let h=h(E)h=h(E) and n=n(E)n=n(E).

  1. 1.

    By definition, the image of π𝒯\pi_{\mathcal{T}} over (E,ψN,Cm)(E,\psi_{N},C^{m}) is the preimage of HT(Cm,×)ωE+/pmωE+HT(C^{m,\times})\subset\omega_{E}^{+}/p^{m}\omega_{E}^{+} inside ωE+ωEint=𝒯rig|(E,ψN,Cm)\omega_{E}^{+}\subset\omega_{E}^{int}=\mathcal{T}_{rig}|_{(E,\psi_{N},C^{m})}. So we want to check that if (E,ψN,Cm)X0(pm)[0,v](E,\psi_{N},C^{m})\in X_{0}(p^{m})[0,v], then HT(Cm)HT(C^{m}) is nonzero in ωE+/pmωE+\omega_{E}^{+}/p^{m}\omega_{E}^{+} (so that the preimage of HT(Cm,×)HT(C^{m,\times}) does not contain 0). As stated in the previous section, either Cm[p]C^{m}[p] is canonical or Cm[p]C^{m}[p] is not canonical and h>1pm1(p+1)h>\frac{1}{p^{m-1}(p+1)}.

    If Cm[p]C^{m}[p] is canonical then (Cm)(C^{m})^{\vee} is étale, so v(HT(Cm))=hp1v(HT(C^{m}))=\frac{h}{p-1} and HT(Cm)HT(C^{m}) generates ωE+\omega_{E}^{+}.

    If h>1pm1(p+1)h>\frac{1}{p^{m-1}(p+1)}, then nmn\leq m. Then either

    • an1<ma\leq n-1<m, so that v(HT(Cm))=a+hp1<m+hp1v(HT(C^{m}))=a+\frac{h}{p-1}<m+\frac{h}{p-1},

    • a=n<ma=n<m, so that by Remark 2.1.2 we have v(HT(Cm))n+hp1<m+hp1v(HT(C^{m}))\leq n+\frac{h}{p-1}<m+\frac{h}{p-1}, or

    • a=n=ma=n=m, so that

      v(HT(Cm))\displaystyle v(HT(C^{m})) =m+1p1pm+pm11p1h\displaystyle=m+\frac{1}{p-1}-\frac{p^{m}+p^{m-1}-1}{p-1}h
      <m+1p1pm+pm1p11pm1(p+1)+hp1=m+hp1.\displaystyle<m+\frac{1}{p-1}-\frac{p^{m}+p^{m-1}}{p-1}\cdot\frac{1}{p^{m-1}(p+1)}+\frac{h}{p-1}=m+\frac{h}{p-1}.

    In all cases, HT(Cm)HT(C^{m}) is nonzero mod pmωE+p^{m}\omega_{E}^{+}.

  2. 2.

    This image can be written in the form

    l(/pm)×(pv(HT(Cm))xlm+pm+hp1𝒪p).\bigcup_{l\in(\mathbb{Z}/p^{m}\mathbb{Z})^{\times}}\left(p^{v(HT(C^{m}))}x_{l}^{m}+p^{m+\frac{h}{p-1}}\mathcal{O}_{\mathbb{C}_{p}}\right).

    Each ball in the union is pv(HT(Cm))p^{v(HT(C^{m}))} times a ball centered at some xlmx_{l}^{m} whose radius has valuation m+hp1v(HT(Cm))m+\frac{h}{p-1}-v(HT(C^{m})). The proposition claims that if xlmx_{l}^{m} and xlmx_{l^{\prime}}^{m} are in distinct residue classes mod pcp^{c}, the balls centered at xlmx_{l}^{m} and xlmx_{l^{\prime}}^{m} do not overlap. Since in this case v(xlmxlm)c1v(x_{l}^{m}-x_{l^{\prime}}^{m})\leq c-1, it is necessary and sufficient to have

    m+hp1v(HT(Cm))>c1m+\frac{h}{p-1}-v(HT(C^{m}))>c-1

    or equivalently

    v(HT(Cm))<mc+1+hp1.v(HT(C^{m}))<m-c+1+\frac{h}{p-1}.

    The analysis then proceeds exactly as in Part 1.

  3. 3.

    This is because in our expression for the image in the proof of Part 2, we have v(HT(Cm))=hp1v(HT(C^{m}))=\frac{h}{p-1} over X0(pm)[0,pp+1]X_{0}(p^{m})\left[0,\frac{p}{p+1}\right].

/pm\mathbb{Z}/p^{m}\mathbb{Z} acts on both 𝒞m\mathcal{C}^{m} and ω𝒞m\omega_{\mathcal{C}^{m}} by scalar multiplication. The two actions are compatible under HTHT: if P𝒞mP\in\mathcal{C}^{m}, and PHom(𝒞m,μpm)P^{\vee}\in\text{Hom}(\mathcal{C}^{m},\mu_{p^{m}}) is the map given by taking the Weil pairing with PP, then l/pml\in\mathbb{Z}/p^{m}\mathbb{Z} takes PP^{\vee} to ()lP(\cdot)^{l}\circ P^{\vee}, hence (P)dT/T(P^{\vee})^{*}dT/T to

(()lP)dTT=(P)(()l)dTT=(P)dTlTl=(P)lTl1dTTl=(P)ldTT=l(P)dTT.((\cdot)^{l}\circ P^{\vee})^{*}\frac{dT}{T}=(P^{\vee})^{*}((\cdot)^{l})^{*}\frac{dT}{T}=(P^{\vee})^{*}\frac{dT^{l}}{T^{l}}=(P^{\vee})^{*}\frac{lT^{l-1}dT}{T^{l}}=(P^{\vee})^{*}\frac{ldT}{T}=l(P^{\vee})^{*}\frac{dT}{T}.

Thus the action of p×\mathbb{Z}_{p}^{\times} on (𝒞m)××X0(pm)𝒯rig(\mathcal{C}^{m})^{\times}\times_{X_{0}(p^{m})}\mathcal{T}_{rig} via reduction to (/pm)×(\mathbb{Z}/p^{m}\mathbb{Z})^{\times} on the first factor and scalar multiplication on the second factor preserves Im,×I^{m,\times}.

We say that a weight w:p×p×w:\mathbb{Z}_{p}^{\times}\to\mathbb{C}_{p}^{\times} is uu-locally analytic if it is analytic on any ball of the form a+pu𝒪pa+p^{u}\mathcal{O}_{\mathbb{C}_{p}} where ap×a\in\mathbb{Z}_{p}^{\times}. We say that ww is analytic if it is 11-locally analytic.

Definition 2.4.2.

The sheaf ωw\omega^{w} on X0(pm)X_{0}(p^{m}) is the subsheaf of (πX)𝒪Im,×(\pi_{X})_{*}\mathscr{O}_{I^{m,\times}} of sections that are homogeneous of weight ww under the action of p×\mathbb{Z}_{p}^{\times}.

That is, a section ff of ωw\omega^{w} over X0(pm)X_{0}(p^{m}) is an analytic function on points (E,ψN,Cm,P,s)Im,×(E,\psi_{N},C^{m},P,s)\in I^{m,\times} (where EE is an elliptic curve, ψN\psi_{N} a tame level structure, CmC^{m} is a subgroup of EE of order pmp^{m}, P(Cm,)×P\in(C^{m,\vee})^{\times}, and sωEs\in\omega_{E} are such that (P,s)Im,×|(E,ψN,Cm)(P,s)\in I^{m,\times}|_{(E,\psi_{N},C^{m})}) satisfying

z.f(E,ψN,Cm,P,s):=f(E,ψN,Cm,z1P,z1s)=w(z)f(E,ψN,Cm,P,s)z.f(E,\psi_{N},C^{m},P,s):=f(E,\psi_{N},C^{m},z^{-1}P,z^{-1}s)=w(z)f(E,\psi_{N},C^{m},P,s)

for all zp×z\in\mathbb{Z}_{p}^{\times}.

Proposition 2.4.3.

Suppose that ww is uu-locally analytic and vv is such that over any 𝒪p\mathcal{O}_{\mathbb{C}_{p}}-point (E,ψN,Cm)(E,\psi_{N},C^{m}) of X0(pm)[0,v]X_{0}(p^{m})[0,v], the image of π𝒯:Im,×𝒯rig\pi_{\mathcal{T}}:I^{m,\times}\to\mathcal{T}_{rig} is contained in

l(/pc)×pv(HT(Cm))(xlc+pu𝒪p)\coprod_{l\in(\mathbb{Z}/p^{c}\mathbb{Z})^{\times}}p^{v(HT(C^{m}))}(x_{l}^{c}+p^{u}\mathcal{O}_{\mathbb{C}_{p}})

for some c<u+1c<u+1. Then ωw\omega^{w} is an invertible sheaf on X0(pm)[0,v]X_{0}(p^{m})[0,v].

Proof.

By Lemma 2.1 of [19], for any u>uu^{\prime}>u, any analytic function on p×(1+pu𝒪p)\mathbb{Z}_{p}^{\times}(1+p^{u^{\prime}}\mathcal{O}_{\mathbb{C}_{p}}) that is homogeneous of weight ww for the action of the group p×\mathbb{Z}_{p}^{\times} acting by translation is of the form CwCw for some CpC\in\mathbb{C}_{p}. Therefore, under the given assumption on the fiber of Im,×I^{m,\times} over (E,ψN,Cm)(E,\psi_{N},C^{m}), the space of analytic functions on Im,×|(E,ψN,Cm)I^{m,\times}|_{(E,\psi_{N},C^{m})} homogeneous of weight ww—that is to say, ωw|(E,ψN,Cm)\omega^{w}|_{(E,\psi_{N},C^{m})}—is a 11-dimensional vector space over p\mathbb{C}_{p}, as desired. ∎

By Propositions 2.4.1 and 2.4.3, we may conclude the following.

Corollary 2.4.4.
  1. 1.

    If ww is uu-locally analytic for all u>cu>c, then ωw\omega^{w} is an invertible sheaf over X0(pm)[0,v]X_{0}(p^{m})[0,v] for any v<11pmc(p+1)v<1-\frac{1}{p^{m-c}(p+1)}.

  2. 2.

    If ww is analytic, then ωw\omega^{w} is an invertible sheaf over X0(pm)[0,pp+1]X_{0}(p^{m})\left[0,\frac{p}{p+1}\right].

2.5 UpU_{p} operators

As usual, we can define an operator UpU_{p} which acts on sections of the sheaves ωw\omega^{w}. The following notation will help us keep track of how much UpU_{p} increases overconvergence radius.

Definition 2.5.1.

For v(0,1)v\in(0,1), let

succ(v)={pv0<v<pp+111pn(p+1)11pn1(p+1)v<1pn(p+1) for some n1.\operatorname{succ}(v)=\begin{cases}pv&0<v<\frac{p}{p+1}\\ 1-\frac{1}{p^{n}(p+1)}&1-\frac{1}{p^{n-1}(p+1)}\leq v<\frac{1}{p^{n}(p+1)}\text{ for some }n\geq 1.\end{cases}

Note that for all v(0,1)v\in(0,1), succ(v)>v\operatorname{succ}(v)>v, and succn(v)1\operatorname{succ}^{n}(v)\to 1 as nn\to\infty.

Definition 2.5.2.

Let w:p×p×w:\mathbb{Z}_{p}^{\times}\to\mathbb{C}_{p}^{\times} be a uu-locally analytic weight. Suppose

fH0(X0(pm)[0,v],ωw)f\in H^{0}(X_{0}(p^{m})[0,v],\omega^{w})

for some vv such that ωw\omega^{w} is invertible over X0(pm)[0,v]X_{0}(p^{m})[0,v]. Interpreting ff as a function on points (E,ψN,Cm,P,s)Im,×(E,\psi_{N},C^{m},P,s)\in I^{m,\times}, we define

Upf(E,ψN,Cm,P,s)=1pDCm[p]f(E/D,ψ¯N,Cm¯,P¯,(πE,D)1s)U_{p}f(E,\psi_{N},C^{m},P,s)=\frac{1}{p}\sum_{D\neq C^{m}[p]}f(E/D,\overline{\psi}_{N},\overline{C^{m}},\overline{P},(\pi_{E,D}^{*})^{-1}s)

where DD ranges over order-pp subgroups of EE different from Cm[p]C^{m}[p] and πE,D\pi_{E,D} is the projection EE/DE\to E/D.

Proposition 2.5.3.

If fH0(X0(pm)[0,v],ωw)f\in H^{0}(X_{0}(p^{m})[0,v],\omega^{w}), and ωw\omega^{w} is invertible on X0(pm)[0,succ(v)]X_{0}(p^{m})[0,\operatorname{succ}(v)], then UpfU_{p}f is a well-defined element of H0(X0(pm)[0,succ(v)],ωw)H^{0}(X_{0}(p^{m})[0,\operatorname{succ}(v)],\omega^{w}).

Proof.

By Part 2 of Lemma 4.2 of [3], if (E,ψN,Cm)X0(pm)[0,succ(v)](E,\psi_{N},C^{m})\in X_{0}(p^{m})[0,\operatorname{succ}(v)], then
(E/D,ψ¯N,Cm¯)X0(pm)[0,v](E/D,\overline{\psi}_{N},\overline{C^{m}})\in X_{0}(p^{m})[0,v]. It is not necessarily the case, on the other hand, that (P¯,(πE,D)1s)Im,×|(E/D,ψ¯N,Cm¯)(\overline{P},(\pi_{E,D}^{*})^{-1}s)\in I^{m,\times}|_{(E/D,\overline{\psi}_{N},\overline{C^{m}})} (though one can check that this is true on X0(pm)[0,pp+1]X_{0}(p^{m})\left[0,\frac{p}{p+1}\right]). So

f(E/D,ψ¯N,Cm¯,P¯,(πE,D)1s)f(E/D,\overline{\psi}_{N},\overline{C^{m}},\overline{P},(\pi_{E,D}^{*})^{-1}s)

may not be initially defined. However, by the proof of Proposition 2.4.3, f|(E/D,ψ¯N,Cm¯)f|_{(E/D,\overline{\psi}_{N},\overline{C^{m}})} has a unique analytic extension to the image of Im,×|(E,ψN,Cm)I^{m,\times}|_{(E,\psi_{N},C^{m})} under (πE,D)1(\pi_{E,D}^{*})^{-1} as long as ww is still analytic on the image, that is, the image is still contained in l(/pc)×(xlc+pu𝒪p)\coprod_{l\in(\mathbb{Z}/p^{c}\mathbb{Z})^{\times}}(x_{l}^{c}+p^{u}\mathcal{O}_{\mathbb{C}_{p}}) times a power of pp. Since (πE,D)1(\pi_{E,D}^{*})^{-1} is an isomorphism, this is always true. ∎

3 Properness

We may now use the method of Buzzard and Calegari to prove Theorem 1.1.1.

3.1 Overconvergence radius of eigenforms of finite vs. infinite slope

Let w:p×p×w:\mathbb{Z}_{p}^{\times}\to\mathbb{C}_{p}^{\times} be an analytic weight. We can now show that an infinite-slope eigenform of weight ww cannot overconverge to radius 1p+1\frac{1}{p+1}.

Proposition 3.1.1.

If fH0(X0(p)[0,v],ωw)f\in H^{0}(X_{0}(p)[0,v],\omega^{w}) for some v1p+1v\geq\frac{1}{p+1} and Upf0U_{p}f\equiv 0, then f0f\equiv 0.

Proof.

Plugging in any EE with hpp+1h\geq\frac{p}{p+1} and CC varying over all p+1p+1 subgroups of EE of order pp (i.e. so that v(E,ψN,C)=pp+1v(E,\psi_{N},C)=\frac{p}{p+1}), we conclude that

Upf(E,ψN,C,P,s)=DCf(E/D,ψ¯N,C¯=E[p]/D,P¯,(πE,D)1s)=0U_{p}f(E,\psi_{N},C,P,s)=\sum_{D\neq C}f(E/D,\overline{\psi}_{N},\overline{C}=E[p]/D,\overline{P},(\pi_{E,D}^{*})^{-1}s)=0

for every CC. Summing the p+1p+1 resulting equations and dividing by pp, we find that

Df(E/D,ψ¯N,E[p]/D,P¯,(πE,D)1s)=0.\sum_{D}f(E/D,\overline{\psi}_{N},E[p]/D,\overline{P},(\pi_{E,D}^{*})^{-1}s)=0.

Subtracting the first equation from the second, we find that

f(E/C,ψN,E[p]/C,P¯,(πE,C)1s)=0f(E/C,\psi_{N},E[p]/C,\overline{P},(\pi_{E,C}^{*})^{-1}s)=0

for every such EE and CC. Since E/CE/C ranges over the entire circle X0(p){1p+1}X_{0}(p)\left\{\frac{1}{p+1}\right\} (See Theorem 3.3 of [3]), we conclude that f0f\equiv 0 on this circle. Since ff is an analytic function on a connected rigid analytic space, ff must be 0 everywhere. ∎

On the other hand, it is a standard fact, as we check below for completeness, that finite-slope eigenforms overconverge “as much as possible” for the weight ww; in this case, in particular to radius pp+1\frac{p}{p+1}.

Proposition 3.1.2.

If fH0(X0(p)[0,v],ωw)f\in H^{0}(X_{0}(p)[0,v],\omega^{w}) for some v>0v>0 and Upf=λfU_{p}f=\lambda f for some λ0\lambda\neq 0, then ff can be (uniquely) extended to an element of H0(X0(p)[0,pp+1],ωw)H^{0}\left(X_{0}(p)\left[0,\frac{p}{p+1}\right],\omega^{w}\right).

Proof.

On X0(p)[0,v]X_{0}(p)[0,v] we have f=1λnUpnff=\frac{1}{\lambda^{n}}U_{p}^{n}f for all nn. But UpnfU_{p}^{n}f is defined on X0(p)[0,succn(v)]X_{0}(p)[0,\operatorname{succ}^{n}(v)] as long as ωw\omega^{w} is. By Corollary 2.4.4, ωw\omega^{w} is well-defined on X0(p)[0,pp+1]X_{0}(p)\left[0,\frac{p}{p+1}\right], so choosing nn such that pp+1succn(v)\frac{p}{p+1}\leq\operatorname{succ}^{n}(v) gives the desired extension. ∎

3.2 Filling in the puncture

We now prove Theorem 1.1.1. As in the statement, let \mathscr{E} be the pp-adic Coleman-Mazur eigencurve of tame level NN, D×D^{\times} the punctured closed unit disc, and w:𝒲w:\mathscr{E}\to\mathscr{W} the projection map to weight space. We are given a map h:D×h:D^{\times}\to\mathscr{E} such that whw\circ h extends to DD. We may assume that h(D×)h(D^{\times}) is contained in the locus of \mathscr{E} corresponding to cuspidal overconvergent eigenforms, since the Eisenstein locus is finite over weight space and hence proper (for details on the construction of the cuspidal and Eisenstein loci, see Section 7 of [5], the unabridged version of [6]).

Then h(D×)h(D^{\times}) corresponds to a normalized qq-expansion n=1anqn𝒪(D×)q\sum_{n=1}^{\infty}a_{n}q^{n}\in\mathscr{O}(D^{\times})\llbracket q\rrbracket, where a1=1a_{1}=1, and n=1an(d)qn\sum_{n=1}^{\infty}a_{n}(d)q^{n} is the qq-expansion of an overconvergent finite-slope eigenform of weight w(h(d))w(h(d)) for each dD×d\in D^{\times}. We have supdD×|an(d)|1\sup_{d\in D^{\times}}|a_{n}(d)|\leq 1 for all nn, because Hecke operators on spaces of overconvergent modular forms have integral eigenvalues by Lemma 7.1 and Remark 7.6 of [5]. This bound implies that ana_{n} extends to the closed unit disc DD, giving a formal qq-expansion nan(0)qn𝒪pq\sum_{n}a_{n}(0)q^{n}\in\mathcal{O}_{\mathbb{C}_{p}}\llbracket q\rrbracket which, under the action of Hecke operators on formal qq-expansions, is a normalized Hecke eigenform of weight w(h(0))w(h(0)) (nontrivial, since a1(0)=1a_{1}(0)=1). We wish to show that n=1an(0)qn\sum_{n=1}^{\infty}a_{n}(0)q^{n} is overconvergent and finite-slope.

As discussed in the introduction, by Theorem 1.2.1 of [20], over the locus of w𝒲w\in\mathscr{W} such that v(w(exp(p))1)<1v(w(\exp(p))-1)<1, \mathscr{E} decomposes into a countable disjoint union of pieces that are finite over 𝒲\mathscr{W}, so is evidently proper. Therefore we may assume that v(w(exp(p))1)1v(w(\exp(p))-1)\geq 1, in which case ww is analytic (with power series expansion w(z)=w(exp(p))1plogzw(z)=w(\exp(p))^{\frac{1}{p}\log z} for z1+p𝒪pz\in 1+p\mathcal{O}_{\mathbb{C}_{p}}). Shrinking DD if necessary, we may assume that all of w(h(D))w(h(D)) is analytic.

Then for dD×d\in D^{\times}, the eigenform corresponding to n=1an(d)qn\sum_{n=1}^{\infty}a_{n}(d)q^{n} is a section of ωw(h(d))\omega^{w(h(d))} over X0(p)[0,v]X_{0}(p)[0,v] for some v>0v>0; furthermore, by Proposition 3.1.2, this section extends uniquely over X0(p)[0,pp+1]X_{0}(p)\left[0,\frac{p}{p+1}\right], and we may interpret it as a function on I1,×[0,pp+1]I^{1,\times}\left[0,\frac{p}{p+1}\right].

Now n=1an(0)qn\sum_{n=1}^{\infty}a_{n}(0)q^{n} is a nontrivial section of ωw(h(0))\omega^{w(h(0))} over a small disc around a cusp in X0(p){0}X_{0}(p)\{0\} on which qq is well-defined. By Proposition 3.1.1, it suffices to show that nan(0)qn\sum_{n}a_{n}(0)q^{n} also extends to X0(p)[0,pp+1]X_{0}(p)\left[0,\frac{p}{p+1}\right], since then n=1an(0)qn\sum_{n=1}^{\infty}a_{n}(0)q^{n} cannot be in the kernel of UpU_{p}. For this, we use the following lemma of Buzzard-Calegari.

Lemma 3.2.1 (Lemma 7.1 of [6]).

Let YY be a connected affinoid variety, VV a nonempty admissible open affinoid subdomain of YY, D=Sp(pT)D=\operatorname{Sp}(\mathbb{C}_{p}\langle T\rangle), and A=Sp(pT,T1)A=\operatorname{Sp}(\mathbb{C}_{p}\langle T,T^{-1}\rangle). If ff is a function on V×DV\times D and the restriction of ff to V×AV\times A extends to a function on Y×AY\times A, then ff extends to a function on Y×DY\times D.

For completeness, we include their proof.

Proof.

Since YY is connected, we have 𝒪(Y)𝒪(V)\mathscr{O}(Y)\subseteq\mathscr{O}(V). Since ff is defined on V×DV\times D, we have f𝒪(V)Tf\in\mathscr{O}(V)\langle T\rangle. Since f|V×Af|_{V\times A} extends to Y×AY\times A, we also have f𝒪(Y)T,T1f\in\mathscr{O}(Y)\langle T,T^{-1}\rangle. But the intersection of 𝒪(V)T\mathscr{O}(V)\langle T\rangle and 𝒪(Y)T,T1\mathscr{O}(Y)\langle T,T^{-1}\rangle is 𝒪(Y)T\mathscr{O}(Y)\langle T\rangle, which is the space of functions on Y×DY\times D. ∎

We can now show that n=1an(0)qn\sum_{n=1}^{\infty}a_{n}(0)q^{n} extends to X0(p)[0,pp+1]X_{0}(p)\left[0,\frac{p}{p+1}\right] by combining Proposition 3.1.2 with the following Proposition, the same way as in the proof of Theorem 7.2 of [6].

Proposition 3.2.2.

If nan(d)qn\sum_{n}a_{n}(d)q^{n} is vv-overconvergent for all dD×d\in D^{\times}, then n=1an(0)qn\sum_{n=1}^{\infty}a_{n}(0)q^{n} is also vv-overconvergent.

Proof.

Apply Lemma 3.2.1 with Y=I1,×[0,pp+1]Y=I^{1,\times}\left[0,\frac{p}{p+1}\right], VV the preimage of I1,×I^{1,\times} over a small disc around a cusp in X0(p){0}X_{0}(p)\{0\} on which qq is well-defined, and

f(y,d)=n=1an(d)q(y)nf(y,d)=\sum_{n=1}^{\infty}a_{n}(d)q(y)^{n}

for yVy\in V and dDd\in D. Then ff is a function on V×DV\times D which is vv-overconvergent on D×D^{\times}, so its restriction to V×D×V\times D^{\times} extends to Y×D×Y\times D^{\times}, so it extends to Y×DY\times D, so f(y,0)f(y,0) is also vv-overconvergent. ∎

This completes the proof of Theorem 1.1.1.

References

  • [1] Fabrizio Andreatta, Adrian Iovita, and Glenn Stevens. Overconvergent modular sheaves and modular forms for GL2/FGL_{2/F}. Israel Journal of Mathematics, 201(1):299–359, 2014.
  • [2] George Boxer and Vincent Pilloni. Higher Hida and Coleman theories on the modular curve. arXiv preprint arXiv:2002.06845, 2020.
  • [3] Kevin Buzzard. Analytic continuation of overconvergent eigenforms. Journal of the American Mathematical Society, 16(1):29–55, 2003.
  • [4] Kevin Buzzard. Eigenvarieties. LL-functions and Galois representations, London Math. Soc. Lecture Notes, 320:59–120, 2007.
  • [5] Kevin Buzzard and Frank Calegari. The 2-adic eigencurve is proper. arXiv preprint math/0503362, 2005.
  • [6] Kevin Buzzard and Frank Calegari. The 2-adic eigencurve is proper. Doc. Math., (Extra Vol.):211–232, 2006.
  • [7] Frank Calegari. The Coleman–Mazur eigencurve is proper at integral weights. Algebra & Number Theory, 2(2):209–215, 2008.
  • [8] Frank Calegari. Arizona Winter School notes. available at swc. math. arizona. edu, 2013.
  • [9] Robert Coleman and Barry Mazur. The eigencurve. London Mathematical Society Lecture Note Series, pages 1–114, 1998.
  • [10] Robert F Coleman and William A Stein. Approximation of eigenforms of infinite slope by eigenforms of finite slope. Geometric aspects of Dwork theory, 1:437–449, 2003.
  • [11] Brian Conrad. Arizona Winter School notes. available at swc. math. arizona. edu, 2013.
  • [12] Hansheng Diao and Ruochuan Liu. The eigencurve is proper. Duke Mathematical Journal, 165(7):1381–1395, 2016.
  • [13] Laurent Fargues. Application de Hodge-Tate duale d’un groupe de Lubin-Tate, immeuble de Bruhat-Tits du groupe linéaire et filtrations de ramification. Duke Mathematical Journal, 140(3):499–590, 2007.
  • [14] Shin Hattori. On a properness of the Hilbert eigenvariety at integral weights: the case of quadratic residue fields. arXiv preprint arXiv:1601.00775, 2016.
  • [15] Shin Hattori and James Newton. Irreducible components of the eigencurve of finite degree are finite over the weight space. Journal für die reine und angewandte Mathematik, 2020(763):251–269, 2020.
  • [16] Nicholas M Katz. P-adic properties of modular schemes and modular forms. In Modular functions of one variable III, pages 69–190. Springer, 1973.
  • [17] Ruochuan Liu, Daqing Wan, and Liang Xiao. The eigencurve over the boundary of weight space. Duke Math. J., 166(9):1739–1787, June 2017.
  • [18] David Loeffler. Spectral expansions of overconvergent modular functions. International Mathematics Research Notices, 2007(9):rnm050–rnm050, 2007.
  • [19] Vincent Pilloni. Formes modulaires surconvergentes. Ann. Inst. Fourier (Grenoble), 63:219–239, 2013.
  • [20] Rufei Ren and Bin Zhao. Spectral halo for Hilbert modular forms. arXiv preprint arXiv:2005.14267, 2020.
  • [21] Lynnelle Ye. Slopes in eigenvarieties for definite unitary groups. arXiv preprint arXiv:2004.12490, 2020.