Slopes of modular forms and reducible Galois representations, an oversight in the ghost conjecture

By John Bergdall and Robert Pollack

Abstract

The ghost conjecture, formulated by this article’s authors, predicts the list of -adic valuations of the non-zero -eigenvalues (“slopes”) for overconvergent -adic modular eigenforms in terms of the Newton polygon of an easy-to-describe power series (the “ghost series”). The prediction is restricted to eigenforms whose Galois representation modulo is reducible on a decomposition group at . It has been discovered, however, that the conjecture is not formulated correctly. Here we explain the issue and propose a salvage.

1. Introduction

Let be a prime number. The authors previously made a conjecture on the -adic slopes of modular eigenforms with a fixed Galois representation modulo subject to that representation being locally reducible at . See Conjecture 1.1 and also Reference 6, Conjecture 7.1. The local reducibility condition has turned out to be too strong — one reducible situation (and its twists) must also be excluded. This article will explain the exclusion and how to salvage the conjecture. We provide both computational and theoretical evidence for the corrected conjecture.

It is important to correct false conjectures. In this case, however, it is specifically important because there is ongoing progress toward a proof of the conjecture, for Galois representations that are locally reducible with sufficiently generic Serre weights, by Liu, Truong, Xiao, and Zhao. A paper explaining their strategy is available as a preprint Reference 19. The proof itself has also been announced Reference 18.

A general proof, however, will remain out of reach until faulty hypotheses are removed! Addressing the prior misconception further allows us the opportunity to address, in writing, the special case of eigenforms that are twists of modulo . We had previously not tested that case out of convenience.

In Reference 6, the authors called the reducibility assumption “Buzzard regular”, based on Buzzard’s work Reference 11. It is unfortunate that Buzzard’s name was attached, by us, to the faulty condition. Our strongest apologies.

1.1. Recollection of the ghost conjecture

Fix algebraic closures of , of , and of . Let and be an integer that is co-prime to . We then consider continuous, semi-simple, representations that are modular of level . Recall that if is a cuspidal, normalized, eigenform of level with coefficients in , then there is a continuous, semi-simple, Galois representation such that if is a prime, then is unramified at and the trace of a Frobenius element at is equal to , the -th Fourier coefficient of . Saying is modular of level means for such an .

For we define to be the -span of the eigenforms of weight and level such that . So, is modular of level exactly when for some . Define and then in a similar fashion. We write the dimensions of these spaces as:

Set . (The function counts the eigenforms that are -new.)

The ghost conjecture is based on the definition of a sequence of monic polynomials as follows:

(1)

The zeros of are , for the finite list of integers such that

(2)

Fix now. If , , …, is the list of consecutive indices for which (the satisfying the prior inequalities), then the multiplicities of as a zero are: for , the multiplicity is one; for , the multiplicity is two; and so on.

in other words, the multiplicity of as a zero of for , , , …has the form

where the first string of zeros ends at index and there are -many non-zero numbers.

Having defined the polynomials , we set

The ghost conjecture predicts that this series models the characteristic power series of the -operator acting on overconvergent -adic cuspforms whose eigensystem modulo matches .

More precisely, let be a decomposition group at and be the inertia subgroup. Write for the cyclotomic character modulo . Then, denote by the unique integer such that .

We consider the rigid analytic space whose -points are the -adic weights whose action on the torsion subgroup of is raising to the -th power. Given , write and for the -isotypic component of the space of overconvergent -adic cuspforms of level and weight . The zeros defined in (1) occur only at integer weights that lie in . (Recall, , so if then .)

The ghost conjecture for stated in Reference 6 is:

Conjecture 1.1 (False).

Assume and for any . If is reducible, then, for each , the Newton polygon of is the same as the Newton polygon of the characteristic series of the -operator acting on .

1.2. Counter-examples

We give two counter-examples here.

1.2.1. Detailed example

Let and . We choose where is the quadratic Dirichlet character of conductor 3. The dimension functions , , and can be determined using a computer algebra system with built-in libraries for modular forms, such as Magma Reference 9 or Sage Reference 21. The result is compiled in Table 1. From the dimensions, we see that

Since has -adic valuation 1, the least slope on the Newton polygon of is 1 or less. On the other hand, any weight 7 overconvergent -adic eigenform with slope at most 1 is classical by Coleman’s classicality theorem Reference 15, Theorem 6.1. Yet, the slopes of acting on are (calculated using Magma). This is the contradiction.

Remark 1.2.

Strictly speaking, neither Magma nor Sage currently has intrinsic commands to determine the slopes in . The authors provide code in the github repository Reference 4.

1.2.2. Another example

The reader might wonder if the fundamental issue in the prior example is that either is Eisenstein or , both somehow global phenomena. This is not the case.

Explicitly, continue to let and also let be the quadratic character modulo . Then, in there is a 20-dimensional Galois orbit of newforms in which the -adic slope occurs twice, once for an absolutely irreducible and once more for (which is a new representation). Further, is 1-dimensional and thus 2 is the lowest slope occurring for . Yet, as in Section 1.2.1, one can check that the ghost series formalism would predict the lowest slope in is 1.

In fact, lifts to an ordinary eigenform of weight , and thus is locally reducible at . By Proposition 2.2 and point (ii) in the proof of Theorem 2.5 below, we even have

where . (For unfamiliar notations or definitions, see Section 2.) The extension Equation 1.1 is even non-split! If it were split, then by Reference 16, Theorem 4.5 the representation would arise from a weight one form over of level 43 and character . We verified such a form does not exist using a computer package called Weight1, which calculates weight one eigenforms modulo , written by Wiese. (The code, which requires Magma, is currently available on Wiese’s website Reference 24.)

1.2.3. The general phenomenon

The common link between the examples is indicated by equation Equation 1.1. In both cases, is reducible at and its semi-simplification is an unramified representation for which a Frobenius element acts with trace zero. Generalizing these examples, the fundamental gap we found in the ghost conjecture occurs for that are twists of representations which locally at are reducible, with semi-simplification that is unramified with Frobenius trace zero. Indeed, in Proposition 3.8 we will show that the ghost conjecture will always fail in such examples. However, as long as we exclude such representations (which we do in Section 2), we believe that the ghost conjecture will hold for remaining ’s.

1.3. A brief history

We first encountered the example in Section 1.2.1 in 2015-16, when we learned by explicit computation that the ghost series formalism failed to predict the 5-adic slopes of eigenforms of level (with no restriction on the Galois representations modulo 5). We apparently overlooked that computation when formulating the -version of our conjecture. Later, in 2018, James Newton asked us about the -conjecture for examples where is unramified at (up to a twist). We mistakenly missed the counter-examples once more. We realized our mistake only in early 2021 after our attention was diverted toward questions on reductions of crystalline Galois representations.

1.4. Plan for the remainder of the article

In Section 2 we revisit the reducible versus irreducible dichotomy and replace it with a new one, which we name regular versus irregular. We study this new dichotomy in both local and global terms, relating it to crystalline lifts on the one hand and slopes of modular forms in low weight on the other. The salvaged conjecture and evidence is given Section 3. Concluding, and open-ended, remarks occupy Section 4.

2. Salvaging the ghost conjecture: Redefining regularity

To salvage Conjecture 1.1, we redefine regularity, replacing the reducible vs. irreducible dichotomy with a slightly different one. In this section we let be any prime, unless noted.

We first fix notations for local Galois representations modulo . Let be the unramified quadratic extension of and write for a niveau 2 fundamental character. Let and . We then write where is the quadratic character of . Thus has determinant and . Given we also write for the character on that is unramified and whose value on a Frobenius element is . In this notation, .

The continuous, semi-simple, representations are completely classified. Up to a twist by a character, any is isomorphic to either for some or with and . The irreducible occur for twists of with . The representation is reducible when .

Definition 2.1.

A continuous, semi-simple, representation is called irregular if is a twist of for some .

Naturally, we say is regular when it is not irregular. Note that regular representations are always reducible, but the regularity of depends on the entire action of , whereas the reducibility depends only on .

Proposition 2.2 illustrates the regular versus irregular dichotomy in terms of crystalline lifts with small Hodge–Tate weights. The reader who is unfamiliar with -adic Hodge theory may prefer to skip to Theorem 2.5 or reference the well-written introduction to Reference 13. The notation is as follows. For such that and , denotes the unique two-dimensional, irreducible, crystalline representation of whose Hodge–Tate weights are and such that the characteristic polynomial of the crystalline Frobenius acting on is . Write for the semi-simplification of the reduction of modulo .

Proposition 2.2.

Let be a continuous, semi-simple, representation. The following conditions are equivalent (the (a)’s and (b)’s are individually equivalent):

(1)

The representation is irregular:

(a)

is a twist of with , or

(b)

is a twist of with .

(2)

Either:

(a)

is irreducible, or

(b)

is a twist of an unramified representation whose Frobenius trace is zero.

(3)

Either:

(a)

is a twist of for some and any , or

(b)

is a twist of for any .

Proof.

The equivalence of (1) and (2) follows directly from the discussion prior to Definition 2.1. The equivalence of (1) and (3) follows from calculations of when is small. Namely, if then (which is, in particular, irreducible and thus irregular). When , we have:

Thus is irregular when . See Reference 7, Théorème 3.2.1 for these results. From them, we plainly see (3) implies (1) and (1) implies (3) since any irregular is a twist of for some .

Remark 2.3.

The trivial representation mod 2 is irregular by Proposition 2.2(2), since it is unramified with Frobenius trace . See Section 4.1.

We now turn toward global considerations. Given a representation , write for the semi-simplification. (This depends on the group that is representing, but the context will be clear any time we use the notation.) Since the decomposition groups at are all conjugate and isomorphic to , we can speak unambiguously about a global representation being regular on or not:

Definition 2.4.

We say is regular if is regular.

The action of is reducible for every regular , so our new regular condition implies the condition in Conjecture 1.1. The difference is that we have further excluded one particular twist-stable family of reducible representations.

We describe now a mechanism in terms of the arithmetic of modular forms to determine whether or not a given is regular. Write

The space is stable under the -operator. Recall, if is an eigenform then . We say is ordinary if .

Theorem 2.5.

Consider the following two conditions:

(1)

The representation is regular.

(2)

For all , the eigenforms in are all ordinary, and the eigenforms in are either ordinary or satisfy .

Then, (1) always implies (2). If , then (2) implies (1).

More precisely:

(a)

If is irreducible, then there exists a non-ordinary eigenform of weight between and such that is a cyclotomic twist of .

(b)

If and is irregular but is reducible, then there exists a such that and for all .

Proof.

To start, we recall two fundamental facts. For any eigenform , we let

be the corresponding global -adic Galois representation. Then:

(i)

If is ordinary, then is reducible and .

(ii)

If is non-ordinary, then is isomorphic to a twist of .

It seems the first result was first proven by Deligne in the 1970’s but never published. A common reference is Reference 25, Theorem 2. For the second, see Reference 10, Théorème 6.5.

Now we prove (1) implies (2). Suppose that is regular and is an eigenform of weight such that . Then, either is ordinary or is regular by (ii). By Proposition 2.2, and the classification of in equation Equation 2.1, the latter can only happen if and . This proves (1) implies (2).

Now we will prove (2) implies (1). By Reference 16, Theorem 3.4 there exists an eigenform of weight where , with the caveat that may be an Eisenstein series or, in the case , may be only a mod modular form. Furthermore, the theory of -cycles (see Reference 16, Proposition 3.3) predicts the least positive weight in which lifts to a modular eigenform.

Now consider (a), where is irreducible. Then, the possibility of cannot actually occur. Indeed, if has weight one then is unramified by theorems of Gross, Coleman–Voloch, and Wiese. See Reference 23, Corollary 1.3. The theory of -cycles ( and in Reference 16, Proposition 3.3 cited above) then implies there exists a non-ordinary eigenform of weight between and such that for some . This completes the proof of (a).

In case (b), we assume that and after twisting that with reducible, yet irregular and has weight . We first claim that either and is ordinary, or . Further, in both cases .

To see this, note first that if then, as above, is unramified at and thus . If , we first check that is ordinary. Indeed, if is Eisenstein, then it is ordinary because . If were cuspidal non-ordinary, then (ii) and Proposition 2.2 would imply is irreducible, which it is not. So, is ordinary. Then, by (i) we see . But is irregular, and so Proposition 2.2 implies , which implies since .

Finally, regardless of whether and is ordinary, or , the theory of -cycles guarantees there exists a modular eigenform of weight such that . For such , we have and thus is non-ordinary by (i). Here we use to know . In particular we’ve shown that . We also see from (ii) that is reducible and irregular for any . Thus, by Proposition 2.2, for all . This completes the proof of (b).

Remark 2.6.

A key step in the proof of part (b) of Theorem 2.5 is that if is an ordinary eigenform of weight , then lifts to an eigenform of weight . This statement also holds when is non-ordinary, as long as . For instance, one could directly argue using modular symbols modulo as in Reference 3, Theorem 3.4(a,b). (That reference assumes throughout.)

For , though, there exist weight 3 forms where doesn’t lift to weight 5. For instance, if is the quadratic character modulo 7, then is spanned by a unique eigenform Reference 22, Newform 7.3.b.a whose -expansion begins , whereas is spanned by a unique eigenform Reference 22, Newform 7.5.b.a whose -expansion is . These two eigenforms are not twists of each other modulo 3.

3. The salvaged conjecture and evidence

We now state for the record our salvaged -ghost conjecture. We place into the conjecture the hypotheses we need, as a reminder.

Conjecture 3.1 (The salvaged -ghost conjecture).

Let . If is regular in the sense of Definition 2.4, then, for each , the Newton polygon of is equal to the Newton polygon of the characteristic series of the -operator acting on .

In the next subsections, we provide computational and theoretical evidence for our salvaged conjecture.

Remark 3.2.

Notice that Conjecture 3.1 includes , in contrast with Conjecture 1.1. See Section 3.1.

Remark 3.3.

Condition (2) in Theorem 2.5 appears in Buzzard’s slope conjectures when (see Reference 11, Definition 1.3). For odd primes, it was irrelevant in Buzzard’s work, since restricting to level removes concerns about odd weights and .

3.1. Galois multiplicities

In order to construct the series , and thus test Conjecture 3.1, one must understand the functions and . In fact, each of these functions is determined after a finite computation. Specifically, either can be determined from the values with and along with multiplicities of Eisenstein series (if is Eisenstein). This is explained in Reference 6, Section 6. The technique involves modular symbols modulo , which is the source of the assumption . We do not currently know dimension formulas when . For an alternative method via the trace formula, which is adaptable to , see forthcoming work of Anni, Ghitza, and Medvedovsky Reference 1.

There is one caveat: the exposition in Reference 6 ignores the case (“twists of ”) for convenience. These are precisely the Galois representations modulo that contribute to torsion in ’s, which creates complications. However, the complications are luckily limited to exposition. The actual result, that low weight calculations for and all its twists will determine -multiplicities in all weights, are still valid for twists of .

So, while omitting more specific details, we used this principle in order to carry out the numerical testing explained in the next section, regardless of whether is a twist of or not. (But still, .) For the interested reader, we also created computer code that implements the natural process to calculate from low weight values. See Reference 4.

3.2. Numerical evidence

In Reference 6, Section 7.2, we gave significant numerical evidence for Conjecture 3.1. Here we report more numerical evidence, especially highlighting the following two cases:

(1)

Representations with a twist of unramified with non-zero Frobenius trace (Examples 3.4-3.5).

(2)

Twists of (Example 3.6).

The latter was a case not numerically tested in Reference 6, and the former is meant to ensure the issue with the counter-examples in Section 1.2 has been accurately diagnosed.

In each example, we describe a single and explain why it is regular. Then, when we write that we “verified Conjecture 3.1 in this case for weights …” we mean: for each we calculated the -slopes on and checked that they matched the first -many slopes on . That is, we checked (using the code referenced in Remark 1.2) the ghost series accurately predicts the classical slopes for and each of its twists (for a range of weights).

Example 3.4.

Let denote the quadratic character of conductor thought of as taking values in and let . If then is unramified at and it is regular if and only if . For each of and for all primes for which is regular, we verified Conjecture 3.1 in this case for weights respectively.

Example 3.5.

Let , , and be the quadratic character modulo 23. Then, supports a unique whose Frobenius trace at is equal to 2. Specifically, if is a member of the newform orbit Reference 22, Newform 23.5.b.b and is the unique prime above 5 with inertia degree 2 for the number field generated by ’s Hecke eigenvalues, then is the Galois representation associated with modulo . The eigenform is -ordinary and arises from a weight 7 form with slope 1. By Theorem 2.5, is regular. Unlike Example 3.4, this is globally irreducible. We verified Conjecture 3.1 in this case for weights .

Example 3.6.

Let which is regular. For both of and and for all primes with , we verified Conjecture 3.1 in this case for weights respectively.

Remark 3.7.

Gouvêa and Mazur once conjectured that if is a rational number and , then the multiplicity of the (-)slope in weight will be the same as the multiplicity in weight provided (see Reference 17, Conjecture 1). Buzzard and Calegari found many counter-examples, all occurring in situations where is locally irreducible at Reference 12. Those are, in particular, irregular. In the final paragraphs of loc. cit., however, it is mentioned that Clay noticed issues in a globally reducible, and in fact regular, situation.

Specifically, let where is the quadratic character modulo 4, which is regular for because (compare with Example 3.4). The violation of Gouvêa–Mazur, which Clay noticed, occurs because the -adic -slopes in weight at level are (two times), whereas in weight they are (four times). We highlight this because this violation of the Gouvêa–Mazur conjecture is consistent with Conjecture 3.1. That is, the ghost series predicts exactly the same slope pattern.

The data collected in Example 3.4 gives rather systematic counter-examples to the Gouvêa–Mazur conjecture (on -subspaces), all of which are consistent with Conjecture 3.1. The issue is also a local one, as opposed to a global one, it seems. We found the same phenomenon in the setting of Example 3.5, which is globally irreducible. Specifically, in Example 3.5, in weight the slope sequence is , in weight 27 it is . Conjecture 3.1 predicts these slopes.

3.3. Theoretical evidence

For theoretical evidence, we show that must be regular if the ghost series correctly calculates slopes of modular forms.

Proposition 3.8.

Let . If is irregular, then the ghost series of some twist of does not correctly compute slopes of modular forms. That is, there is some and some weight such that the slopes of the Newton polygon of do not match the slopes of acting on .

Proof.

First assume that is irreducible. By Theorem 2.5, after replacing by a twist, we may assume that arises from a non-ordinary cuspform of weight between and . We will show that the ghost series for does not correctly compute the slopes of in weight .

To this end, write . The smallest such that is possibly a zero of any is . From the definition of the ghost series, if is a zero is for then

Since , we have that is increasing by Reference 6, Proposition 6.12. So, and thus the first slopes of the Newton polygon of this ghost series (at any weight!) are all 0. In particular, if the slopes attached to this series equaled the true slopes of , then the (-)slopes in would be all be 0. But this contradicts the fact that arises from a non-ordinary form of weight .

Now we consider the case where is reducible (but still irregular). By part (b) of Theorem 2.5, after replacing by a twist, we may assume that and for all . We will show that the ghost series for does not correctly compute slopes in weight .

To this end, again write . Since is non-zero, we have that . Thus, as argued above, , and because the multiplicity pattern described on pg. 432 begins , , …. Thus, the lowest slope of the Newton polygon of is at most 1. But we’ve also assumed every slope in is larger than one. This is a contradiction.

4. Conclusion

We have presented a new, salvaged, ghost conjecture for Galois representations modulo , along with theoretical and computational evidence. We end by briefly describing two questions raised by our work (separate from whether or not the conjecture is true).

4.1. Small primes

Let . According to the definitions given here, the trivial representation is irregular. This is rather confounding. The -adic slopes at level , where the only Galois representation modulo 2 is the trivial one, are the most carefully studied of all examples. It is a major motivating example for both Buzzard’s work and the work of the authors. The ghost formalism is believed to predict the -adic slopes at level , yet it does not (nor does Buzzard’s work) correctly predict the -adic -slopes at level where is still the trivial representation.

Of course, the second condition of Theorem 2.5 holds for the trivial representation when and , so perhaps that condition is the correct one for which the ghost series will predict slopes (regardless of being small). This is in-line with Buzzard’s ad hoc definition of -regular when in Reference 11. Note however that even in the -regular situation, the -adic ghost conjecture requires a modification as in Reference 5, Section 5.

The authors have not seen a reason to exclude from Conjecture 3.1. Yet, we want to be careful. We have not done extensive testing for because the derivations of Galois multiplicities in Section 3.1 rely on Reference 3, which assumes that . The same methods work immediately when as long as is a group without -torsion, but stating a conjecture for under such a restriction would be rather prosaic.

4.2. Integral slopes and local Galois representations

It is conjectured that if is odd, is even, and then is irreducible. See Reference 14, Conjecture 4.1.1. The possibility of such a result was discussed “in emails between Breuil, Buzzard, and Emerton” in the early 2000’s according to loc. cit. To give it a neutral name we’ll call it the integral slope conjecture. We’ve been told by Breuil and Buzzard that the inspiration for discussing the integral slope conjecture was numerical calculations of global slopes and explicit local calculations in small weights. In light of this article’s emphasis on reducibility being a faulty lens through which to predict global slopes, we offer the following comments.

When is odd and is even, the irregular in the sense of Definition 2.1 are all irreducible. So, the integral slope conjecture might as well say if is even and then is irregular. This re-phrasing opens up the possibility of stating a more full conjecture. For instance, is not included in the original integral slope conjecture because when and , then is reducible and hence so is for any (see Reference 14, Remark 4.1.6). Yet, these examples are in fact all irregular. Separate objections are raised when is odd (see Reference 14, Remark 4.1.5), each of which is rendered moot if instead the conjecture begins with the assumption that and ends with the conclusion that is irregular. (One might also note that in Proposition 2.2 we could additionally add the condition (d): is a twist of for some .)

So, we know no counter-examples or counter-arguments for the statement:

In fact, Equation  is supported by theorems of Buzzard and Gee Reference 13, Theorem 1.6 and Bhattacharya and Ghate Reference 8, Theorem 1.1. An example theorem derived from the latter work is that , as long as and and . This nicely complements the discussion in the previous paragraph, further illustrating why “irregular” replacing “irreducible” is necessary for any generalization of the integral slope conjecture. We note, finally, that Equation  is also consistent with a theorem of Nagel and Pande Reference 20, Theorem 0.1 focused on , and work of Arsovski Reference 2, Theorem 1, which covers a more sporadic range of slopes.

This article began by correcting a false conjecture. So, here, it is best to wait for a more conceptual explanation before declaring Equation  as an outright conjecture. Even the integral slope conjecture, limited to even weights and odd , is based on a combination of computational evidence and theoretical evidence restricted to low weights and slopes. It has not, for instance, been contextualized within a wider context of all local Galois representations modulo nor, except in the above paragraphs, has issues with odd or been really confronted. Does replacing reducibility by regularity offer any conceptual clarity? There’s certainly a fascinating opportunity here for further explanation.

Acknowledgments

We thank Liang Xiao for keeping us updated on the progress of his collaboration with Liu, Truong, and Zhao, which aims to prove the ghost conjecture. Both authors also thank the Max Planck Institute for Mathematics in Bonn, Germany, for its hospitality in July 2021, when this paper was largely written.

Figures

Table 1.

Dimensions of -isotypic components in and , for

3022
7142
11262
15284

Mathematical Fragments

Conjecture 1.1 (False).

Assume and for any . If is reducible, then, for each , the Newton polygon of is the same as the Newton polygon of the characteristic series of the -operator acting on .

Remark 1.2.

Strictly speaking, neither Magma nor Sage currently has intrinsic commands to determine the slopes in . The authors provide code in the github repository Reference 4.

Equation (1.1)
Definition 2.1.

A continuous, semi-simple, representation is called irregular if is a twist of for some .

Proposition 2.2.

Let be a continuous, semi-simple, representation. The following conditions are equivalent (the (a)’s and (b)’s are individually equivalent):

(1)

The representation is irregular:

(a)

is a twist of with , or

(b)

is a twist of with .

(2)

Either:

(a)

is irreducible, or

(b)

is a twist of an unramified representation whose Frobenius trace is zero.

(3)

Either:

(a)

is a twist of for some and any , or

(b)

is a twist of for any .

Equation (2.1)
Definition 2.4.

We say is regular if is regular.

Theorem 2.5.

Consider the following two conditions:

(1)

The representation is regular.

(2)

For all , the eigenforms in are all ordinary, and the eigenforms in are either ordinary or satisfy .

Then, (1) always implies (2). If , then (2) implies (1).

More precisely:

(a)

If is irreducible, then there exists a non-ordinary eigenform of weight between and such that is a cyclotomic twist of .

(b)

If and is irregular but is reducible, then there exists a such that and for all .

Conjecture 3.1 (The salvaged -ghost conjecture).

Let . If is regular in the sense of Definition 2.4, then, for each , the Newton polygon of is equal to the Newton polygon of the characteristic series of the -operator acting on .

Example 3.4.

Let denote the quadratic character of conductor thought of as taking values in and let . If then is unramified at and it is regular if and only if . For each of and for all primes for which is regular, we verified Conjecture 3.1 in this case for weights respectively.

Example 3.5.

Let , , and be the quadratic character modulo 23. Then, supports a unique whose Frobenius trace at is equal to 2. Specifically, if is a member of the newform orbit Reference 22, Newform 23.5.b.b and is the unique prime above 5 with inertia degree 2 for the number field generated by ’s Hecke eigenvalues, then is the Galois representation associated with modulo . The eigenform is -ordinary and arises from a weight 7 form with slope 1. By Theorem 2.5, is regular. Unlike Example 3.4, this is globally irreducible. We verified Conjecture 3.1 in this case for weights .

Example 3.6.

Let which is regular. For both of and and for all primes with , we verified Conjecture 3.1 in this case for weights respectively.

Proposition 3.8.

Let . If is irregular, then the ghost series of some twist of does not correctly compute slopes of modular forms. That is, there is some and some weight such that the slopes of the Newton polygon of do not match the slopes of acting on .

Equation ()

References

Reference [1]
S. Anni, A. Ghitza, and A. Medvedovsky, Refined dimensions of Atkin–Lehner eigenspaces, In preparation.
Reference [2]
B. Arsovski, On the reductions of certain two-dimensional crystabelline representations, Res. Math. Sci. 8 (2021), no. 1, Paper No. 12, 50, DOI 10.1007/s40687-020-00231-6. MR4215364,
Show rawAMSref \bib{Arsovski-Reductions1}{article}{ author={Arsovski, Bodan}, title={On the reductions of certain two-dimensional crystabelline representations}, journal={Res. Math. Sci.}, volume={8}, date={2021}, number={1}, pages={Paper No. 12, 50}, issn={2522-0144}, review={\MR {4215364}}, doi={10.1007/s40687-020-00231-6}, }
Reference [3]
A. Ash and G. Stevens, Modular forms in characteristic and special values of their -functions, Duke Math. J. 53 (1986), no. 3, 849–868, DOI 10.1215/S0012-7094-86-05346-9. MR860675,
Show rawAMSref \bib{AshStevens-Duke}{article}{ author={Ash, Avner}, author={Stevens, Glenn}, title={Modular forms in characteristic $l$ and special values of their $L$-functions}, journal={Duke Math. J.}, volume={53}, date={1986}, number={3}, pages={849--868}, issn={0012-7094}, review={\MR {860675}}, doi={10.1215/S0012-7094-86-05346-9}, }
Reference [4]
J. Bergdall and P. Pollack, Computer code for calculating slopes of modular forms and the ghost series, https://github.com/rpollack9974/Slopes-of-modular-forms.
Reference [5]
J. Bergdall and R. Pollack, Slopes of modular forms and the ghost conjecture, Int. Math. Res. Not. IMRN 4 (2019), 1125–1144, DOI 10.1093/imrn/rnx141. MR3915298,
Show rawAMSref \bib{BergdallPollack-GhostPaperShort}{article}{ author={Bergdall, John}, author={Pollack, Robert}, title={Slopes of modular forms and the ghost conjecture}, journal={Int. Math. Res. Not. IMRN}, date={2019}, number={4}, pages={1125--1144}, issn={1073-7928}, review={\MR {3915298}}, doi={10.1093/imrn/rnx141}, }
Reference [6]
J. Bergdall and R. Pollack, Slopes of modular forms and the ghost conjecture, II, Trans. Amer. Math. Soc. 372 (2019), no. 1, 357–388, DOI 10.1090/tran/7549. MR3968772,
Show rawAMSref \bib{BergdallPollack-GhostPaper-II}{article}{ author={Bergdall, John}, author={Pollack, Robert}, title={Slopes of modular forms and the ghost conjecture, II}, journal={Trans. Amer. Math. Soc.}, volume={372}, date={2019}, number={1}, pages={357--388}, issn={0002-9947}, review={\MR {3968772}}, doi={10.1090/tran/7549}, }
Reference [7]
L. Berger, Représentations modulaires de et représentations galoisiennes de dimension 2 (French, with English and French summaries), Astérisque 330 (2010), 263–279. MR2642408,
Show rawAMSref \bib{Berger-Modp-compat}{article}{ author={Berger, Laurent}, title={Repr\'{e}sentations modulaires de $\operatorname {GL}_2(\mathbf Q_p)$ et repr\'{e}sentations galoisiennes de dimension 2}, language={French, with English and French summaries}, journal={Ast\'{e}risque}, number={330}, date={2010}, pages={263--279}, issn={0303-1179}, isbn={978-2-85629-281-5}, review={\MR {2642408}}, }
Reference [8]
S. Bhattacharya and E. Ghate, Reductions of Galois representations for slopes in , Doc. Math. 20 (2015), 943–987. MR3404215,
Show rawAMSref \bib{BhattacharyaGhate-Slope12}{article}{ author={Bhattacharya, Shalini}, author={Ghate, Eknath}, title={Reductions of Galois representations for slopes in $(1,2)$}, journal={Doc. Math.}, volume={20}, date={2015}, pages={943--987}, issn={1431-0635}, review={\MR {3404215}}, }
Reference [9]
W. Bosma, J. Cannon, and C. Playoust, The Magma algebra system. I. The user language, J. Symbolic Comput. 24 (1997), no. 3-4, 235–265, DOI 10.1006/jsco.1996.0125. Computational algebra and number theory (London, 1993). MR1484478,
Show rawAMSref \bib{MagmaCite}{article}{ author={Bosma, Wieb}, author={Cannon, John}, author={Playoust, Catherine}, title={The Magma algebra system. I. The user language}, note={Computational algebra and number theory (London, 1993)}, journal={J. Symbolic Comput.}, volume={24}, date={1997}, number={3-4}, pages={235--265}, issn={0747-7171}, review={\MR {1484478}}, doi={10.1006/jsco.1996.0125}, }
Reference [10]
C. Breuil, Sur quelques représentations modulaires et -adiques de . II (French, with French summary), J. Inst. Math. Jussieu 2 (2003), no. 1, 23–58, DOI 10.1017/S1474748003000021. MR1955206,
Show rawAMSref \bib{Breuil-SomeRepresentations2}{article}{ author={Breuil, Christophe}, title={Sur quelques repr\'{e}sentations modulaires et $p$-adiques de $\operatorname {GL}_2(\mathbf Q_p)$. II}, language={French, with French summary}, journal={J. Inst. Math. Jussieu}, volume={2}, date={2003}, number={1}, pages={23--58}, issn={1474-7480}, review={\MR {1955206}}, doi={10.1017/S1474748003000021}, }
Reference [11]
K. Buzzard, Questions about slopes of modular forms (English, with English and French summaries), Astérisque 298 (2005), 1–15. Automorphic forms. I. MR2141701,
Show rawAMSref \bib{Buzzard-SlopeQuestions}{article}{ author={Buzzard, Kevin}, title={Questions about slopes of modular forms}, language={English, with English and French summaries}, note={Automorphic forms. I}, journal={Ast\'{e}risque}, number={298}, date={2005}, pages={1--15}, issn={0303-1179}, review={\MR {2141701}}, }
Reference [12]
K. Buzzard and F. Calegari, A counterexample to the Gouvêa-Mazur conjecture (English, with English and French summaries), C. R. Math. Acad. Sci. Paris 338 (2004), no. 10, 751–753, DOI 10.1016/j.crma.2004.03.016. MR2059481,
Show rawAMSref \bib{BuzzardCalegari-GouveaMazur}{article}{ author={Buzzard, Kevin}, author={Calegari, Frank}, title={A counterexample to the Gouv\^{e}a-Mazur conjecture}, language={English, with English and French summaries}, journal={C. R. Math. Acad. Sci. Paris}, volume={338}, date={2004}, number={10}, pages={751--753}, issn={1631-073X}, review={\MR {2059481}}, doi={10.1016/j.crma.2004.03.016}, }
Reference [13]
K. Buzzard and T. Gee, Explicit reduction modulo of certain two-dimensional crystalline representations, Int. Math. Res. Not. IMRN 12 (2009), 2303–2317, DOI 10.1093/imrn/rnp017. MR2511912,
Show rawAMSref \bib{BuzzardGee-SmallSlope}{article}{ author={Buzzard, Kevin}, author={Gee, Toby}, title={Explicit reduction modulo $p$ of certain two-dimensional crystalline representations}, journal={Int. Math. Res. Not. IMRN}, date={2009}, number={12}, pages={2303--2317}, issn={1073-7928}, review={\MR {2511912}}, doi={10.1093/imrn/rnp017}, }
Reference [14]
K. Buzzard and T. Gee, Slopes of modular forms, Families of automorphic forms and the trace formula, Simons Symp., Springer, [Cham], 2016, pp. 93–109. MR3675164,
Show rawAMSref \bib{BuzzardGee-Slopes}{article}{ author={Buzzard, Kevin}, author={Gee, Toby}, title={Slopes of modular forms}, conference={ title={Families of automorphic forms and the trace formula}, }, book={ series={Simons Symp.}, publisher={Springer, [Cham]}, }, date={2016}, pages={93--109}, review={\MR {3675164}}, }
Reference [15]
R. F. Coleman, Classical and overconvergent modular forms, Invent. Math. 124 (1996), no. 1-3, 215–241, DOI 10.1007/s002220050051. MR1369416,
Show rawAMSref \bib{Coleman-ClassicalandOverconvergent}{article}{ author={Coleman, Robert F.}, title={Classical and overconvergent modular forms}, journal={Invent. Math.}, volume={124}, date={1996}, number={1-3}, pages={215--241}, issn={0020-9910}, review={\MR {1369416}}, doi={10.1007/s002220050051}, }
Reference [16]
B. Edixhoven, The weight in Serre’s conjectures on modular forms, Invent. Math. 109 (1992), no. 3, 563–594, DOI 10.1007/BF01232041. MR1176206,
Show rawAMSref \bib{Edixhoven-Weights}{article}{ author={Edixhoven, Bas}, title={The weight in Serre's conjectures on modular forms}, journal={Invent. Math.}, volume={109}, date={1992}, number={3}, pages={563--594}, issn={0020-9910}, review={\MR {1176206}}, doi={10.1007/BF01232041}, }
Reference [17]
F. Gouvêa and B. Mazur, Families of modular eigenforms, Math. Comp. 58 (1992), no. 198, 793–805, DOI 10.2307/2153218. MR1122070,
Show rawAMSref \bib{GouveaMazur-FamiliesEigenforms}{article}{ author={Gouv\^{e}a, F.}, author={Mazur, B.}, title={Families of modular eigenforms}, journal={Math. Comp.}, volume={58}, date={1992}, number={198}, pages={793--805}, issn={0025-5718}, review={\MR {1122070}}, doi={10.2307/2153218}, }
Reference [18]
R. Liu, N. X. Truong, L. Xiao, and B. Zhao, Slopes of modular forms and geometry of eigencurves, In preparation.
Reference [19]
R. Liu, N. X. Truong, L. Xiao, and B. Zhao, A local analogue of the ghost conjecture of Bergdall–Pollack, Preprint, arXiv:2206.15372, 2022.
Reference [20]
E. Nagel and A. Pande, Reductions of modular Galois representations of slope , Preprint, arXiv:1801.08820, 2019.
Reference [21]
W. Stein, et al., Sage Mathematics software, The Sage Development Team, http://www.sagemath.org.
Reference [22]
The LMFDB Collaboration, The -functions and modular forms database, http://www.lmfdb.org, 2021, accessed 22 September 2021.
Reference [23]
G. Wiese, On Galois representations of weight one, Doc. Math. 19 (2014), 689–707. MR3247800,
Show rawAMSref \bib{Wiese-WeightOne}{article}{ author={Wiese, Gabor}, title={On Galois representations of weight one}, journal={Doc. Math.}, volume={19}, date={2014}, pages={689--707}, issn={1431-0635}, review={\MR {3247800}}, }
Reference [24]
G. Wiese, MAGMA package Weight1, https://math.uni.lu/wiese/programs/Weight1/index.html, Accessed October 2021.
Reference [25]
A. Wiles, On ordinary -adic representations associated to modular forms, Invent. Math. 94 (1988), no. 3, 529–573, DOI 10.1007/BF01394275. MR969243,
Show rawAMSref \bib{Wiles-OrdinaryModular}{article}{ author={Wiles, A.}, title={On ordinary $\lambda $-adic representations associated to modular forms}, journal={Invent. Math.}, volume={94}, date={1988}, number={3}, pages={529--573}, issn={0020-9910}, review={\MR {969243}}, doi={10.1007/BF01394275}, }

Article Information

MSC 2020
Primary: 11F33 (Congruences for modular and -adic modular forms)
Secondary: 11F85 (-adic theory, local fields)
Author Information
John Bergdall
University of Arkansas, Fayetteville, Arkansas, 72703
bergdall@uark.edu
ORCID
MathSciNet
Robert Pollack
Department of Mathematics and Statistics, Boston University, 111 Cummington Mall, Boston, Massachusetts 02215
rpollack@math.bu.edu
Homepage
MathSciNet
Additional Notes

The research was supported by a Simons Collaboration Grant (PI: first author Award #713782) and an NSF grant (PI: second author DMS-1702178).

Communicated by
Romyar T. Sharifi
Journal Information
Proceedings of the American Mathematical Society, Series B, Volume 9, Issue 40, ISSN 2330-1511, published by the American Mathematical Society, Providence, Rhode Island.
Publication History
This article was received on , revised on , and published on .
Copyright Information
Copyright 2022 by the authors under Creative Commons Attribution-NonCommercial 3.0 License (CC BY NC 3.0)
Article References
  • Permalink
  • Permalink (PDF)
  • DOI 10.1090/bproc/136
  • MathSciNet Review: 4504234
  • Show rawAMSref \bib{4504234}{article}{ author={Bergdall, John}, author={Pollack, Robert}, title={Slopes of modular forms and reducible Galois representations, an oversight in the ghost conjecture}, journal={Proc. Amer. Math. Soc. Ser. B}, volume={9}, number={40}, date={2022}, pages={432-444}, issn={2330-1511}, review={4504234}, doi={10.1090/bproc/136}, }

Settings

Change font size
Resize article panel
Enable equation enrichment

Note. To explore an equation, focus it (e.g., by clicking on it) and use the arrow keys to navigate its structure. Screenreader users should be advised that enabling speech synthesis will lead to duplicate aural rendering.

For more information please visit the AMS MathViewer documentation.