A modular proof of the properness of the Coleman-Mazur eigencurve
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 is injective on spaces of forms of large overconvergence radius of any analytic weight.
1 Introduction
1.1 Background
The Coleman-Mazur eigencurve is a rigid analytic space parametrizing overconvergent -adic modular Hecke eigenforms with nonzero -eigenvalues. It has a projection map , where is the weight space parametrizing continuous characters of . It was constructed by Coleman and Mazur in [9] for and tame level , and generalized to all primes and tame levels by Buzzard in [4].
In their original 1998 paper, Coleman and Mazur asked whether there exist -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 satisfies the valuative criterion for properness, and proved by explicit means that the eigencurve is indeed “proper” in this sense for and tame level . (Note that 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 be the -adic Coleman-Mazur eigencurve of tame level . Let be the closed unit disc, and write for with the origin removed. Suppose that is a morphism of rigid spaces such that extends to . Then extends to a morphism compatible with .
1.2 Method of proof
We now describe our method. It suffices to prove the statement over the center disc of , consisting of analytic weights (i.e. characters of that are analytic on for any integer ). This is because the geometry of over the remainder (“boundary annulus”) of is relatively simple. In particular, by Theorem 1.2.1 of [20], in that region, decomposes into a countable disjoint union of pieces that are finite over . (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 over the center disc of 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 of finite-slope overconvergent eigenforms extends to an overconvergent eigenform at the puncture, also of “large” overconvergence radius; the question is whether is also finite-slope. Finally, one shows that is in fact injective on the space of forms of “large” overconvergence radius, so indeed .
The limitation of this argument is its reliance on Katz’s geometric definition of an overconvergent modular form of algebraic weight as a section of the line bundle on a certain rigid subset of the modular curve , along with Coleman’s ad hoc definition of an overconvergent modular form of general weight as a Katz overconvergent form of weight multiplied by the -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- line bundle on a suitable rigid subset of for every weight . (Andreatta, Iovita, and Stevens independently constructed similar sheaves using a different method in [1].) However, the subset on which Pilloni defines is not quite large enough: for analytic, is defined on the locus where the Hasse invariant has valuation less than , 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 ).
In light of this, we begin by constructing extensions of Pilloni’s line bundles on higher-level modular curves of the form (which parametrize tuples where is an elliptic curve, is a tame level structure, and is a cyclic subgroup of of order ). These extensions are well-defined in particular over the closure of the locus where is canonical, which is sufficient for the application to properness, but in fact go even farther. For , one can define rigid subsets of such that if , is the preimage of under the forgetful projection if , and contains, among other things, the entire locus where is supersingular. Then we have the following.
Theorem 1.2.1 (Proposition 2.4.3).
For a fixed weight , there are cutoffs such that is well-defined on for each and as .
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 , 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 .
Theorem 1.2.2 (Proposition 3.1.1).
Let be an analytic weight. If for some and , then .
Understanding these linear-algebraic properties of 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 ([17]) and the author’s analysis of slopes in eigenvarieties for definite unitary groups of arbitrary rank ([21]) rely on checking that is nonzero on certain classical subspaces. Calegari has conjectured (see [8]) the much stronger statement that in some cases admits a convergent spectral expansion (this has only been proven for , , and , 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 -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 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
Our construction of overconvergent forms follows that of Pilloni, whose insight was as follows. As in the introduction, let be an integer not divisible by , and write, for short, and . Let be the universal semi-abelian scheme over , be the identity section, and ; more generally, any time we have a group over , we will write for the identity section and for . Let be the total space of over , and the open subset of nonzero sections.
Classical forms (resp. Katz overconvergent forms) of algebraic weight on are sections of over (resp. a rigid subset of excluding small supersingular discs). Equivalently, they are functions on whose restriction to any fiber of over, say, a -point of transforms by weight under multiplication by . If we work instead with a locally analytic weight , there is no such thing as a function on that transforms by weight , since does not converge on all of . However, does converge on a union of balls in . Thus we could define forms of weight if we replaced by a smaller rigid space whose fibers over points of were unions of balls on which converged.
Pilloni shows that computing Hodge-Tate maps over points of 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 be an elliptic curve over , its identity, its space of invariant differentials, and its Tate module (note that since is finite flat over ). Let be the Hodge-Tate map, defined as follows: if (where and ), let be the map taking to the Weil pairing of and (also defined over ); then
(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 be the standard valuation on (so ). We will compute for any . Let be the valuation of the Hasse invariant of if that valuation lies in , and otherwise. If , let be such that ; if , let .
Let ; recall that if , then has canonical subgroups , where has order , defined as follows. is the subgroup of order defined by Lubin and Katz in Chapter 3 of [16], and once has been defined, is the preimage of under the isogeny . When is ordinary, is the unique lift of the kernel of Frobenius on the reduction of , and exists for all .
Proposition 2.1.1.
Let . Let and . Let be maximal such that , if it exists; otherwise let . Then
Remark 2.1.2.
Note that in the second case, plugging in gives
In order to make this computation, we first need to understand the -torsion points of the completion of at the identity. Recall that is associated to a formal group over of height if is supersingular and if is ordinary, and that we have an embedding which is an isomorphism if is supersingular and an isomorphism onto the canonical subgroup if is ordinary.
We will compute the valuation of a point , as defined in Section 1.2.1 of [13]: one chooses an isomorphism taking the identity section of to , and defines to be the valuation of the -coordinate of the image of under this isomorphism. The choice of isomorphism does not matter because a ring automorphism of fixing the ideal must fix valuations of -coordinates.
When convenient, we will also use to refer to the image of in .
Lemma 2.1.3.
Let . Let and .
-
1.
Assume .
-
(a)
If , then
-
(b)
If , then
-
(a)
-
2.
Assume . Then
-
3.
Assume . Then necessarily , and .
Proof.
-
1.
We will check this by induction, simultaneously with the following statements:
—if and , then the possible valuations of are given by the Newton polygon with vertices .
—if does not exist or , then the possible valuations of are given by the Newton polygon (line segment) with vertices .
First we check the base case . From the proof of Theorem 3.10.7 of [16], we see that the Newton polygon of the power series associated to multiplication by on is given by the points . By definition, the elements of are the roots of , so for , we have
More generally, as at the beginning of Section 1.3 of [13], if , the possible valuations of are given by the Newton polygon obtained as the convex hull of (since the possible choices of are the roots of the power series ).
Now assume the statements for . First suppose . Then the possible Newton polygons for are given by the convex hulls of each of
together with . The slope from to is
This is greater than the slope from to if and only if
which is true because . This gives us the statement about the Newton polygon in the case , and thus also the first two cases for (that is, and but ). The slope from to is , which is always less than , the slope from to ; this gives us the statement about the Newton polygon in the case . Therefore the remaining cases for can be obtained by taking each of and dividing by .
Now suppose . Then everything is the same except that even in the case , the Newton polygon for given is a single line segment, with slope .
-
2.
If , the Newton polygon of has vertices . The claim follows from induction simultaneously with the statement that if for , the Newton polygon of is the single line segment from to .
-
3.
If , the Newton polygon of has vertices . The claim follows from induction simultaneously with the statement that if for , the Newton polygon of is the single line segment from to .
∎
With this calculation in hand, we may apply formulas from [13] to compute the values of asserted in Proposition 2.1.1.
Proof of Proposition 2.1.1.
First assume . Choose any and let be an -point in the kernel of . Suppose is maximal such that . Identifying with its preimage in , Theorem 1.11 of [13] says that
Split the second sum further to get
First assume . Then the first sum is
the second is
and the third is
This gives a total of
On the other hand if , we have
and
giving a total of
These give the asserted values of for . Next, if , then for any and any as before, we have
giving us a total of
as claimed.
Finally, we consider the case . We cannot use Fargues’s formula directly in this case, because it only applies to -divisible groups coming from formal groups (i.e. connected ones), which is not. However, we can follow most of the same calculations as in Section 1.2.2 of [13]. In particular, if , define . Then Proposition 1.2 of [13] still holds, since if is a finite locally free subgroup, depends only on the connected component of the identity in .
With this definition of , using the same calculations as those leading up to Proposition 1.6 of [13], we conclude the following: if there exists an integer such that for an -point in the kernel of , we have
then Fargues’s formula for holds; otherwise, must be in for all , and hence must be in . When exists, the first situation happens, and the same calculation as when holds. When , the second situation happens, and we have . To avoid a long, irrelevant digression, details are left as an exercise to the reader. ∎
Let be a formal model for over . Let be the pullback of the sheaf of integral differential forms on , so that . Let be the -submodule of generated by the image of . It will be useful to work with while constructing the fiber bundle of unions of balls on which our forms will be defined.
Proposition 2.1.4.
We have
Proof.
Remark 2.1.5.
One can also directly check that is generated by for any without using most of Proposition 2.1.1. This is because the argument in Section 1.2.2 of [13] shows that for , if
then equals the LHS, and otherwise . But the LHS can also be written as , which is clearly minimized when , because Lubin and Katz actually define in [16] as the subset of of largest slope. In this case as in the base case of Lemma 2.1.3, so we are done.
For any , since the map is -linear, it descends to a map which we will also refer to by . This “renormalized” Hodge-Tate map, a lift of the standard map , 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
Let be the universal semi-abelian scheme over , the universal cyclic subgroup of order , and the subset of generating elements. Let , , , and be their formal completions along their special fibers.
Let and be the formal completions of and along their special fibers, and the rigid generic fibers of and , and and the analytifications of and . (Recall that the fibers of over are unit balls, with natural inclusions into the fibers of over , which are affine lines; see e.g. Section 3.3 of [11].)
Proposition 2.2.1.
There exists a unique rigid open subset such that for each finite extension of and each point , where is a semi-abelian scheme of dimension over , is a tame level structure on , and is a cyclic order- subgroup of , we have
Proof.
We follow the proof of Theorem 3.1 of [19]. Let be a formal open affine of small enough to trivialize . Let be a generating section of over , so that . The conormal exact sequence
gives a surjection , so over , we have for some . Let be the -algebra giving rise to the map . Let be the universal order- point over , and . Then , and is an element of . The total space of can be written as where the section corresponds to . Then the rigid generic fiber of satisfies the desired property of over the rigid generic fiber of . ∎
Definition 2.2.2.
Let . Let and be the natural projections.
2.3 Admissible open subsets of
We now set up some notation for the admissible open subsets of over which our sheaves of modular forms will be defined. Recall the following function on from [3].
Definition 2.3.1 (Buzzard).
Let . Let
Recall that can be constructed geometrically by taking two copies of the locus of where is ordinary, one corresponding to the locus of where is ordinary and is canonical, the other corresponding to the locus of where is ordinary and is not canonical, and connecting them along their missing supersingular discs with supersingular “tubes”; then Buzzard’s function measures the location of a supersingular point along a supersingular tube, with increasing from to as one moves away from and toward .
For an interval , let be the subset of on which . Then, for example, ; is the locus where either is canonical, or has no canonical subgroup and is arbitrary; is the locus where is not canonical; and . See [3] for more details.
Let be the preimage of under the projection map taking to . Then, for example, if , is the locus of such that either is canonical, or is not canonical but .
2.4 Invertible sheaves of locally analytic weight
We are almost ready to define our sheaves of modular forms of locally analytic weight on . In order to do so, we have to make sure that converges on the image of in under .
For each , choose a section of . For short, we will write for the valuation of of a generating element of .
Proposition 2.4.1.
Let , where is an elliptic curve, a tame level structure, and a subgroup of of order . Choose a trivialization .
-
1.
If for some , the image of over does not contain .
-
2.
If for some , the image of over is contained in
for some .
-
3.
If , the image of over can be written in the form
Proof.
Let and .
-
1.
By definition, the image of over is the preimage of inside . So we want to check that if , then is nonzero in (so that the preimage of does not contain ). As stated in the previous section, either is canonical or is not canonical and .
If is canonical then is étale, so and generates .
If , then . Then either
-
•
, so that ,
-
•
, so that by Remark 2.1.2 we have , or
-
•
, so that
In all cases, is nonzero mod .
-
•
-
2.
This image can be written in the form
Each ball in the union is times a ball centered at some whose radius has valuation . The proposition claims that if and are in distinct residue classes mod , the balls centered at and do not overlap. Since in this case , it is necessary and sufficient to have
or equivalently
The analysis then proceeds exactly as in Part 1.
-
3.
This is because in our expression for the image in the proof of Part 2, we have over .
∎
acts on both and by scalar multiplication. The two actions are compatible under : if , and is the map given by taking the Weil pairing with , then takes to , hence to
Thus the action of on via reduction to on the first factor and scalar multiplication on the second factor preserves .
We say that a weight is -locally analytic if it is analytic on any ball of the form where . We say that is analytic if it is -locally analytic.
Definition 2.4.2.
The sheaf on is the subsheaf of of sections that are homogeneous of weight under the action of .
That is, a section of over is an analytic function on points (where is an elliptic curve, a tame level structure, is a subgroup of of order , , and are such that ) satisfying
for all .
Proposition 2.4.3.
Suppose that is -locally analytic and is such that over any -point of , the image of is contained in
for some . Then is an invertible sheaf on .
Proof.
By Lemma 2.1 of [19], for any , any analytic function on that is homogeneous of weight for the action of the group acting by translation is of the form for some . Therefore, under the given assumption on the fiber of over , the space of analytic functions on homogeneous of weight —that is to say, —is a -dimensional vector space over , as desired. ∎
Corollary 2.4.4.
-
1.
If is -locally analytic for all , then is an invertible sheaf over for any .
-
2.
If is analytic, then is an invertible sheaf over .
2.5 operators
As usual, we can define an operator which acts on sections of the sheaves . The following notation will help us keep track of how much increases overconvergence radius.
Definition 2.5.1.
For , let
Note that for all , , and as .
Definition 2.5.2.
Let be a -locally analytic weight. Suppose
for some such that is invertible over . Interpreting as a function on points , we define
where ranges over order- subgroups of different from and is the projection .
Proposition 2.5.3.
If , and is invertible on , then is a well-defined element of .
Proof.
By Part 2 of Lemma 4.2 of [3], if , then
. It is not necessarily the case, on the other hand, that (though one can check that this is true on ). So
may not be initially defined. However, by the proof of Proposition 2.4.3, has a unique analytic extension to the image of under as long as is still analytic on the image, that is, the image is still contained in times a power of . Since 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 be an analytic weight. We can now show that an infinite-slope eigenform of weight cannot overconverge to radius .
Proposition 3.1.1.
If for some and , then .
Proof.
Plugging in any with and varying over all subgroups of of order (i.e. so that ), we conclude that
for every . Summing the resulting equations and dividing by , we find that
Subtracting the first equation from the second, we find that
for every such and . Since ranges over the entire circle (See Theorem 3.3 of [3]), we conclude that on this circle. Since is an analytic function on a connected rigid analytic space, must be 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 ; in this case, in particular to radius .
Proposition 3.1.2.
If for some and for some , then can be (uniquely) extended to an element of .
Proof.
On we have for all . But is defined on as long as is. By Corollary 2.4.4, is well-defined on , so choosing such that gives the desired extension. ∎
3.2 Filling in the puncture
We now prove Theorem 1.1.1. As in the statement, let be the -adic Coleman-Mazur eigencurve of tame level , the punctured closed unit disc, and the projection map to weight space. We are given a map such that extends to . We may assume that is contained in the locus of 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 corresponds to a normalized -expansion , where , and is the -expansion of an overconvergent finite-slope eigenform of weight for each . We have for all , 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 extends to the closed unit disc , giving a formal -expansion which, under the action of Hecke operators on formal -expansions, is a normalized Hecke eigenform of weight (nontrivial, since ). We wish to show that is overconvergent and finite-slope.
As discussed in the introduction, by Theorem 1.2.1 of [20], over the locus of such that , decomposes into a countable disjoint union of pieces that are finite over , so is evidently proper. Therefore we may assume that , in which case is analytic (with power series expansion for ). Shrinking if necessary, we may assume that all of is analytic.
Then for , the eigenform corresponding to is a section of over for some ; furthermore, by Proposition 3.1.2, this section extends uniquely over , and we may interpret it as a function on .
Now is a nontrivial section of over a small disc around a cusp in on which is well-defined. By Proposition 3.1.1, it suffices to show that also extends to , since then cannot be in the kernel of . For this, we use the following lemma of Buzzard-Calegari.
Lemma 3.2.1 (Lemma 7.1 of [6]).
Let be a connected affinoid variety, a nonempty admissible open affinoid subdomain of , , and . If is a function on and the restriction of to extends to a function on , then extends to a function on .
For completeness, we include their proof.
Proof.
Since is connected, we have . Since is defined on , we have . Since extends to , we also have . But the intersection of and is , which is the space of functions on . ∎
We can now show that extends to 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 is -overconvergent for all , then is also -overconvergent.
Proof.
Apply Lemma 3.2.1 with , the preimage of over a small disc around a cusp in on which is well-defined, and
for and . Then is a function on which is -overconvergent on , so its restriction to extends to , so it extends to , so is also -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 . 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. -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.