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

\dedication

To Professor Richard Taylor with gratitude

Ordinary primes in Hilbert modular varieties

Junecue Suh [email protected] Department of Mathematics
1156 High St
University of California, Santa Cruz
Santa Cruz CA 95064
Abstract

A well-known conjecture, often attributed to Serre, asserts that any motive over any number field has infinitely many ordinary reductions (in the sense that the Newton polygon coincides with the Hodge polygon). In the case of Hilbert modular cuspforms ff of parallel weight (2,,2)(2,\cdots,2), we show how to produce more ordinary primes by using the Sato-Tate equidistribution and combining it with the Galois theory of the Hecke field. Under the assumption of stronger forms of Sato-Tate equidistribution, we get stronger (but conditional) results.

In the case of higher weights, we formulate the ordinariness conjecture for submotives of the intersection cohomology of proper algebraic varieties with motivic coefficients, and verify it for the motives whose \ell-adic Galois realisations are abelian on a finite index subgroup. We get some results for Hilbert cuspforms of weight (3,,3)(3,\cdots,3), weaker than those for (2,,2)(2,\cdots,2).

:
11F41 (primary), 14G35, 11F30, 11G18 (secondary).
keywords:
Hilbert modular form, ordinary reduction, Sato-Tate equidistribution.

1 Introduction

Let 𝒳\mathcal{X} be a projective smooth scheme over the ring 𝒪F,S\mathcal{O}_{F,S} of SS-integers in a number field FF, and let X/FX/F be the generic fibre. For each integer ii, one defines the Hodge polygon HPi(X)HP^{i}(X) from the Hodge filtration on HdRi(X/F)H^{i}_{dR}(X/F): it is the convex planar graph emanating from the origin in which the line segment of slope aa appears with the multiplicity equal to the Hodge number ha,ia=dimFHia(X,Ωa)h^{a,i-a}=\dim_{F}H^{i-a}(X,\Omega^{a}). For every 𝔭\mathfrak{p} outside the finite set SS, one also forms the Newton polygon NP(Frob𝔭,Hcrisi((𝒳𝒪Fk(𝔭))/W(k(𝔭))))\mathrm{NP}(\operatorname{Frob}_{\mathfrak{p}},H^{i}_{\mathrm{cris}}((\mathcal{X}\otimes_{\mathcal{O}_{F}}k(\mathfrak{p}))/W(k(\mathfrak{p})))) using the pp-adic slopes of the crystalline Frobenius. (See §3.4.)

Katz [K71, Conjecture 2.9] conjectured, and Mazur proved [Ma73], that the latter lies above the former:

NP(Frob𝔭,Hcrisi((𝒳𝒪Fk(𝔭))/W(k(𝔭)))HPi(X).\mathrm{NP}(\operatorname{Frob}_{\mathfrak{p}},H^{i}_{\mathrm{cris}}((\mathcal{X}\otimes_{\mathcal{O}_{F}}k(\mathfrak{p}))/W(k(\mathfrak{p})))\geq HP^{i}(X).

When the two coincide, we say that 𝔭\mathfrak{p} is an ordinary prime (in degree ii for XX), following Mazur. The following appears to have been first considered by Serre [Se13, no. 133] for abelian varieties XX (in degree 11):

Conjecture 1.1 ((Ordinariness Conjecture)).

There exists an infinite set of primes 𝔭\mathfrak{p} such that

NP(Frob𝔭,Hcrisi((𝒳𝒪Fk(𝔭))/W(k(𝔭)))=HPi(X).\mathrm{NP}(\operatorname{Frob}_{\mathfrak{p}},H^{i}_{\mathrm{cris}}((\mathcal{X}\otimes_{\mathcal{O}_{F}}k(\mathfrak{p}))/W(k(\mathfrak{p})))=HP^{i}(X).

It is known to be true for elliptic curves and abelian surfaces by arguments of Katz and of Ogus [Og82, Prop. 2.7], and for abelian varieties whose endomorphism ring is \mathbb{Z} and whose algebraic monodromy group satisfies a condition: See Pink [Pi98, §7]. It is also known for all CM abelian varieties. [\ast\ast]

In this article, we investigate Conjecture 1.1 for the factors of modular Jacobians cut out by cusp forms of weight 22, and provide several methods for finding ordinary primes in them.

More generally, we consider the parts of the intersection cohomology of the Hilbert modular varieties attached to totally real number fields FF of degree d=[F:]d=[F:\mathbb{Q}], cut out by new normalised cusp forms ff of parallel weight (2,,2)(2,\cdots,2). The ordinariness in this context first appeared as an assumption in the construction of Galois representations by Wiles [Wi88], which was later removed by Taylor [T89] and by Blasius and Rogawski [BR93]. However, many of the best results and constructions in Iwasawa and Hida theory at present depend on the existence of ordinary primes in a crucial way: See among others the works of Emerton, Pollack and Weston [EPW06]; of Ochiai [Oc06]; of Nekovář [N01]; of Dimitrov [Di13]; of Skinner and Urban [SkUr14]; and of Wan [Wn15].

Since, at the moment, we lack a satisfactory ‘crystalline’ theory of perverse sheaves or intersection cohomology, we first formulate in §2 analogues of the Katz Conjecture and Conjecture 1.1 for the \ell-adic étale intersection cohomology of any projective variety XX, where \ell is any auxiliary prime. Here we form the Hodge-Tate polygon attached to the Galois representation in place of the Hodge polygon, and the Newton polygon by using the \ell-adic Frobenius.

These conjectures (Conjectures 2.4 and 2.5) satisfy a basic consistency, in that (a) the Hodge-Tate polygon is independent of the choice of a prime λ\lambda of FF lying over \ell; and (b) the conjectures are independent of the auxiliary prime \ell. We prove these statements by using theorems of Gabber (on the independence of \ell in the intersection cohomology of complete varieties) [Fu00], of Katz and Laumon (on the constructibility properties of certain constructions in derived categories) [KL85], of André (his theory of motivated cycles) [A96] and of de Cataldo and Migliorini (on the motivated nature of the decomposition theorem in intersection cohomology) [dCM15].

Let us return to the Hilbert modular varieties and the forms ff. By using either (i) the theorem of de Cataldo and Migliorini in op. cit. and its rational extension due to Patrikis [Pa16] or (ii) the recent motivic constructions of Ivorra and Morel [IM19], we construct an intersection motive of XX which in realisations give the intersection cohomology. Then by lifting the action of the Hecke correspondences on the intersection cohomology to one on the intersection motive (see Proposition 2.2.6), we construct an André motive M(f)M(f). The conjectures make sense for these submotives, and the consistency mentioned above also hold for them. We denote by KfK_{f} the Hecke field of ff.

We say that a subset Σ\Sigma of MaxSpec(𝒪F)\mathrm{MaxSpec}(\mathcal{O}_{F}) is abundant if Σ\Sigma has lower (natural) density >0>0, and that Σ\Sigma is principally abundant if there exists a finite extension F/FF^{\prime}/F such that the inverse image of Σ\Sigma in MaxSpec(𝒪F)\mathrm{MaxSpec}(\mathcal{O}_{F^{\prime}}) has density =1=1 in FF^{\prime}. In the previous cases where Conjecture 1.1 has been established, in fact a principally abundant set of ordinary primes was found.

By using the construction of the Galois representation attached to ff and the purity of IHIH, we first show that M(f)M(f) satisfies the analogue of the Katz Conjecture and that we can push the Newton polygon ‘half way’ to ordinariness in a quantifiable sense, for a principally abundant set of primes. However, in the attempt to push just beyond the half-way threshold, we face an obstruction of “geometry of numbers” type (Minkowski). We show how to overcome it (for a principally abundant set of primes) by using a stronger form of Sato-Tate equidistribution (see §3.3), but this last form remains unknown in general.

In order to go further and to obtain unconditional results, we look into (a) ‘multivariate’ variants of the Sato-Tate Conjecture in §3.3 and (b) the interaction between FF and KfK_{f} in §3.5. For the latter, we define an invariant, the slope σF(K)[0,1]\sigma_{F}(K)\in[0,1] of a coefficient number field KK over a ground number field FF (see Definition 3.5.4), by using the action of Gal(¯/F)\operatorname{Gal}(\overline{\mathbb{Q}}/F) on the set Hom(K,¯)\operatorname{Hom}(K,\overline{\mathbb{Q}}) of field embeddings. The slope σF(K)\sigma_{F}(K) is 0 if, for example, (1) [K:][K:\mathbb{Q}] is a prime not dividing [F:][F:\mathbb{Q}]; or (2) the Galois group of K/K/\mathbb{Q} is the (full) symmetric group on [K:][K:\mathbb{Q}] letters111this ‘Maeda-like’ condition appears to be often satisfied for the Hecke fields K=KfK=K_{f} in practice, but not always. See Section 5 for examples. and KK and FF have coprime discriminants.

Here is a collection of results which follow from the Main Theorem 4.1.1:

Theorem.

Let the notation be as above. Then M(f)M(f) has an abundant set of ordinary primes if at least one of the following conditions is satisfied:

  1. (a)

    [Kf:]2[K_{f}^{\circ}:\mathbb{Q}]\leq 2; where KfK_{f}^{\circ} is the smallest Frobenius field of ff in the sense of Ribet (see §3.2 for the precise definition);

  2. (b)

    ff is of CM type;

  3. (c)

    the slope σF~(Kf)\sigma_{\widetilde{F}}(K_{f}) is equal to 0, where F~\widetilde{F} is the Galois closure of FF over \mathbb{Q}; or

  4. (d)

    an element of Gal(¯/F~)\operatorname{Gal}(\overline{\mathbb{Q}}/\widetilde{F}) has exactly 22 orbits in Hom(Kf,¯)\operatorname{Hom}(K_{f},\overline{\mathbb{Q}}), the orbits have the same size and ff satisfies a strong form of Sato-Tate equidistribution named (RST) in §3.3.

The statement in (c) follows from the quantitative part (4) of Theorem 4.1.1: The smaller the slope σ\sigma, the closer to ordinariness we can push the Newton polygons.

In Section 5, we look at the forms ff with low levels for 44 number fields FF of degree d4d\leq 4, and show that for most ff under consideration, (a), (b) and (c) provide an abundant set of ordinary primes unconditionally, and that (d) complements them under the strong Sato-Tate condition. We give descriptions of the conditional cases as well as some cases where our methods fall short of yielding abundance of ordinary primes.

Plan. Here are some main ingredients and ideas in the text.

In Section 2, we show (Theorem 2.2.1) that for André motives, the Hodge polygon made of the Hodge numbers in its (transcendental) Betti realisations coincide with the Hodge-Tate polygon made of the pp-adic Hodge-Tate weights in the (algebraic) pp-adic étale realisations. This applies in particular to the Hecke isotypic components of the intersection cohomology (motive) of the Hilbert modular varieties.

In Section 3, we introduce a few things in preparation for the Main Theorem in Section 4. (a) In §3.4 we study the partially ordered semiring of Newton polygons; this will be useful in dealing with the tensor induction in Theorem 4.1.1. (b) In §3.5, we define the notion of slope and bisection in the interaction between the ground field FF and the coefficient field KK; very roughly speaking, they measure the sizes and shapes of ‘large’ orbits in the Hecke and Frobenius fields. In Theorem 4.1.1, these will be combined with the Chebotarev density theorem and strong forms of the Sato-Tate Conjecture.

In Section 4 we prove the Main Theorem 4.1.1. Here we make a connection (which to the author’s knowledge is new) between two conjectures on the eigenvalues of the Frobenius elements: Namely, the Sato-Tate Conjecture on the archimedean properties of the eigenvalues (which in turn is closely linked to the Langlands Program) on the one hand and the Ordinariness Conjecture on the pp-adic properties (with varying pp) of the eigenvalues on the other hand.

In the final Section 6, we formulate analogues of the Katz and the ordinariness conjectures for submotives of the intersection cohomology of more general motivic coefficients, following a suggestion of Katz. In cases where we have good crystalline realisations compatible with the \ell-adic realisations (which include the nonconstant motivic coefficients on Hilbert modular varieties), we verify the Katz conjecture by using Mazur’s theorem. In case the submotive has potentially abelian \ell-adic realisation, we also verify the ordinariness conjecture by using Serre’s theory [Se98]. Finally, we provide some methods to deal with the parallel motivic weight (3,,3)(3,\cdots,3) in the Hilbert modular case.

2 Formulation of Conjectures for IHIH

2.1 Polygons

Let FF be a number field with algebraic closure FsF^{s}, \ell a prime number, VV a \mathbb{Q}_{\ell}-vectorspace of dimension m<m<\infty, and

ρ:Gal(Fs/F)Aut(V)GLm()\rho:\operatorname{Gal}(F^{s}/F)\longrightarrow\operatorname{Aut}(V)\simeq GL_{m}(\mathbb{Q}_{\ell})

a continuous representation that is unramified outside a finite set SS of maximal ideals of 𝒪F\mathcal{O}_{F}.

Definition 2.1.1.

Assume that ρ\rho is \mathbb{Q}-rational in the sense of Serre [Se98].

Then for each maximal ideal 𝔭\mathfrak{p} of 𝒪F\mathcal{O}_{F} outside SS and with residue characteristic pp\neq\ell, we define the Newton polygon NP(Frob𝔭,ρ)=NP(Frob𝔭|V)\mathrm{NP}(\operatorname{Frob}_{\mathfrak{p}},\rho)=\mathrm{NP}(\operatorname{Frob}_{\mathfrak{p}}|_{V}) as the Newton polygon of the characteristic polynomial

det(TFrob𝔭:V)[T]\det(T-\operatorname{Frob}_{\mathfrak{p}}:V)\in\mathbb{Q}[T]

with respect to the pp-adic valuation v𝔭v_{\mathfrak{p}} on \mathbb{Q} normalised by v𝔭(𝔭)=1v_{\mathfrak{p}}(\mathbb{N}\mathfrak{p})=1.

Equivalently: Choose an isomorphism ¯¯p\overline{\mathbb{Q}}_{\ell}\simeq\overline{\mathbb{Q}}_{p} and let x1,,xm¯x_{1},\cdots,x_{m}\in\overline{\mathbb{Q}}_{\ell} be the eigenvalues of Frob𝔭\operatorname{Frob}_{\mathfrak{p}} on VV. Then, the multiset of slopes

{vp(x1)vp(𝔭),,vp(xm)vp(𝔭)}\left\{\frac{v_{p}(x_{1})}{v_{p}(\mathbb{N}\mathfrak{p})},\cdots,\frac{v_{p}(x_{m})}{v_{p}(\mathbb{N}\mathfrak{p})}\right\}

gives the Newton polygon. It is independent of the chosen isomorphism.

Definition 2.1.2.

Assume that ρ\rho is Hodge-Tate at every prime λ\lambda of FF lying over ()(\ell), and that the set with multiplicities of the Hodge-Tate weights at λ\lambda is independent of λ|\lambda|\ell.

Then we define the Hodge-Tate polygon HTP(ρ)=HTP(V)\mathrm{HTP}(\rho)=\mathrm{HTP}(V) as the convex planar polygon starting from (0,0)(0,0) in which the slope ii appears as many times as the Hodge-Tate weight ii appears in ρ\rho.

There seem to be competing sign conventions for the Hodge-Tate weights. We take the ‘geometric’ one, so that H1H^{1} of an elliptic curve has Hodge-Tate weights {0,1}\{0,1\}.

2.2 Independence of Hodge-Tate weights

First, we show that the Hodge-Tate polygon we have defined coincides with the classical Hodge polygon for all André motives, thereby extending a theorem of Faltings [Fa88], [Fa89]. For this we will crucially rely on the theory of motivated cycles and the resulting category of André motives, given in [A96].

Theorem 2.2.1.

Let MM be an André motive over a finite extension field KK of p\mathbb{Q}_{p}, and let σ:K\sigma:K\longrightarrow\mathbb{C} be a complex embedding. Denote by MpM_{p} its pp-adic étale realisation, and by MσM_{\sigma} its Betti realisation via σ\sigma.

Then the set with multiplicities of the Hodge-Tate weights of MpM_{p} coincides with that of the complex Hodge numbers of MσM_{\sigma}.

Proof.

We may and will assume that MM is simple, and that there exist a projective smooth variety YY of dimension dd over KK, an integer nn, and an André motivated cycle

ξAmotd(Y×KY)\xi\in A^{d}_{\mathrm{mot}}(Y\times_{K}Y)

such that ξ\xi acts as the idempotent cutting out MM in 𝔥n(Y)\mathfrak{h}^{n}(Y). Let epe_{p} (resp. eσe_{\sigma}, resp. edRe_{dR}) be the image of ξ\xi in the pp-adic étale (resp. σ\sigma-Betti, resp. de Rham) realisation:

epH2d((Y×Y)KK¯,p)(d),eσH2d(σ(Y×Y),(d)),edRHdR2d(Y×Y)(d).e_{p}\in H^{2d}((Y\times Y)\otimes_{K}\overline{K},\mathbb{Q}_{p})(d),\,\,e_{\sigma}\in H^{2d}(\sigma(Y\times Y),\mathbb{Q}(d)),\,\,e_{dR}\in H^{2d}_{dR}(Y\times Y)(d).

We have the following diagram

Amotd(Y×Y)\textstyle{A^{d}_{\mathrm{mot}}(Y\times Y)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}cldR\scriptstyle{\mathrm{cl}_{dR}}clσ\scriptstyle{\mathrm{cl}_{\sigma}}clp\scriptstyle{\mathrm{cl}_{p}}HdR\textstyle{H_{dR}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Hσ\textstyle{H_{\sigma}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Hp\textstyle{H_{p}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Art\scriptstyle{\mathrm{Art}}\scriptstyle{\sim}HdRK\textstyle{H_{dR}\otimes_{K}\mathbb{C}}Hσ\textstyle{H_{\sigma}\otimes\mathbb{C}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}comp\scriptstyle{\mathrm{comp}_{\mathbb{C}}}\scriptstyle{\sim}Hσp\textstyle{H_{\sigma}\otimes\mathbb{Q}_{p}}HdRKBdR\textstyle{H_{dR}\otimes_{K}B_{dR}}HppBdR\textstyle{H_{p}\otimes_{\mathbb{Q}_{p}}B_{dR}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\scriptstyle{\sim}compdR\scriptstyle{\mathrm{comp}_{dR}}

Here we suppressed Y×YY\times Y in the argument for all cohomology theories as well as the degree 2d2d and the Tate twist (d)(d). Undecorated arrows are extensions of scalars and Art denotes Artin’s comparison isomorphism.

Main Point: The diagram is commutative. In particular, the image of ξ\xi in any group in the diagram is the same, no matter which path emanating from Amotd(Y×Y)A^{d}_{\mathrm{mot}}(Y\times Y) is followed.

This follows from the definition of André motivated cycles, together with the fact that the comparison isomorphisms in display are isomorphisms between Weil cohomology theories (and as such compatible with pullback, pushforward, cup product, cycle class, and Poincaré duality, that are involved in the definition of André motivated cycles). See André [A96, §2.3 and §2.4].

(In contrast, it is not clear whether the similar diagram would be commutative, if we replace the apex AmotA_{\mathrm{mot}} with the larger space of the absolute Hodge cycles.)

Now applying the idempotents obtained from ξ\xi to the similar diagram without apex:

HdRn(Y)\textstyle{H^{n}_{dR}(Y)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Hn(σ(Y),)\textstyle{H^{n}(\sigma(Y),\mathbb{Q})\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Hn(YKK¯,p)\textstyle{H^{n}(Y\otimes_{K}\overline{K},\mathbb{Q}_{p})\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Art\scriptstyle{\mathrm{Art}}\scriptstyle{\sim}HdRn(Y)K\textstyle{H^{n}_{dR}(Y)\otimes_{K}\mathbb{C}}Hn(σ(Y),)\textstyle{H^{n}(\sigma(Y),\mathbb{Q})\otimes\mathbb{C}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}comp\scriptstyle{\mathrm{comp}_{\mathbb{C}}}\scriptstyle{\sim}Hn(σ(Y),)p\textstyle{H^{n}(\sigma(Y),\mathbb{Q})\otimes\mathbb{Q}_{p}}HdRn(Y)KBdR\textstyle{H^{n}_{dR}(Y)\otimes_{K}B_{dR}}Hn(XKK¯,p)pBdR\textstyle{H^{n}(X\otimes_{K}\overline{K},\mathbb{Q}_{p})\otimes_{\mathbb{Q}_{p}}B_{dR}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\scriptstyle{\sim}compdR\scriptstyle{\mathrm{comp}_{dR}}

we get the diagram

MdR\textstyle{M_{dR}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Mσ\textstyle{M_{\sigma}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Mp\textstyle{M_{p}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Art\scriptstyle{\mathrm{Art}}\scriptstyle{\sim}MdRK\textstyle{M_{dR}\otimes_{K}\mathbb{C}}Mσ\textstyle{M_{\sigma}\otimes\mathbb{C}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}comp\scriptstyle{\mathrm{comp}_{\mathbb{C}}}\scriptstyle{\sim}Mσp\textstyle{M_{\sigma}\otimes\mathbb{Q}_{p}}MdRKBdR\textstyle{M_{dR}\otimes_{K}B_{dR}}MppBdR\textstyle{M_{p}\otimes_{\mathbb{Q}_{p}}B_{dR}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\scriptstyle{\sim}compdR\scriptstyle{\mathrm{comp}_{dR}}

Now, on the one hand, the Hodge-Tate weights of MpM_{p} can be read off from the filtration on

(MppBdR)Gal(K¯/K)MdR (filtered isomorphism) .(M_{p}\otimes_{\mathbb{Q}_{p}}B_{dR})^{\operatorname{Gal}(\overline{K}/K)}\simeq M_{dR}\,\,\,\,\,\mbox{ (filtered isomorphism) }.

On the other hand, the Hodge numbers of MσM_{\sigma} can also be read off from the (algebraic) Hodge filtration on MdRM_{dR} in a similar manner.

Corollary 2.2.2.

Let XX be any projective variety over any finite extension KK of p\mathbb{Q}_{p}, and let σ:K\sigma:K\longrightarrow\mathbb{C} be any complex embedding. Then for any integer nn, the Hodge-Tate weights of IHn(XKK¯,p)IH^{n}(X\otimes_{K}\overline{K},\mathbb{Q}_{p}) coincide with the Hodge numbers of IHn(σ(X),)IH^{n}(\sigma(X),\mathbb{Q}).

Proof.

Let π:YX\pi:Y\longrightarrow X be any resolution of singularities. We use the main theorem of de Cataldo and Migliorini [dCM15], strengthened in KK-rationality by Patrikis [Pa16, §8] to deduce the existence of an André motivated cycle

ξ:=ξnAmotd(Y×Y)\xi:=\xi_{n}\in A^{d}_{\mathrm{mot}}(Y\times Y)

such that the Betti realisation

eσ:=clσ(ξ)H2d(σ(Y×Y),(d))e_{\sigma}:=\mathrm{cl}_{\sigma}(\xi)\in H^{2d}(\sigma(Y\times Y),\mathbb{Q}(d))

defines as a correspondence the idempotent for the direct summand

IHn(σ(X),)Hn(σ(Y),)IH^{n}(\sigma(X),\mathbb{Q})\subseteq H^{n}(\sigma(Y),\mathbb{Q})

and similarly for the pp-adic étale realisation

ep:=clp(ξ)H2d((Y×Y)KK¯,p(d)).e_{p}:=\mathrm{cl}_{\mathbb{Q}_{p}}(\xi)\in H^{2d}((Y\times Y)\otimes_{K}\overline{K},\mathbb{Q}_{p}(d)).

Using this, de Cataldo-Migliorini and Patrikis define the KK-rational intersection de Rham cohomology of XX:

IHdRn(X/K)IH^{n}_{dR}(X/K)

as the image of the idempotent

edR:=cldR(ξ)H2d(Y×Y)(d)e_{dR}:=\mathrm{cl}_{dR}(\xi)\in H^{2d}(Y\times Y)(d)

acting on HdRn(Y)H^{n}_{dR}(Y). Clearly, there is a comparison isomorphism of IHdRn(X/K)KIH^{n}_{dR}(X/K)\otimes_{K}\mathbb{C} with the (transcendental) Hodge structure on the Betti realisation IHn(σ(X),)IH^{n}(\sigma(X),\mathbb{Q})\otimes\mathbb{C}. Moreover, this last Hodge structure coincides with the Hodge structure contsructed by Morihiko Saito, see de Cataldo [dC12, Th. 4.3.5].

Apply Theorem 2.2.1 to this situation. ∎

Corollary 2.2.3.

Let MM be an André motive over a number field KK, 𝔓\mathfrak{P} any maximal ideal of 𝒪K\mathcal{O}_{K} of residue characteristic pp, and σ:K\sigma:K\longrightarrow\mathbb{C} any complex embedding.

Then the Hodge-Tate weights of MpM_{p} at 𝔓\mathfrak{P} coincide with the Hodge numbers of MσM_{\sigma}. In particular, the Hodge-Tate weights are independent of 𝔓MaxSpec(𝒪K)\mathfrak{P}\in\mathrm{MaxSpec}(\mathcal{O}_{K}).

This applies, for example, to the André motives cut out by algebraic cycles from the cohomology of a projective smooth variety over KK.

Corollary 2.2.4.

Let XX be a projective variety defined over a number field KK, 𝔓\mathfrak{P} any maximal ideal of 𝒪K\mathcal{O}_{K} of residue characteristic pp, and σ:K\sigma:K\longrightarrow\mathbb{C} any complex embedding.

Then for every integer nn, the Hodge-Tate weights of IHn(XKK¯,p)IH^{n}(X\otimes_{K}\overline{K},\mathbb{Q}_{p}) at 𝔓\mathfrak{P} coincide with the Hodge numbers of IHn(σ(X),)IH^{n}(\sigma(X),\mathbb{Q}).

In particular, the Hodge-Tate weights are independent of 𝔓MaxSpec(𝒪K)\mathfrak{P}\in\mathrm{MaxSpec}(\mathcal{O}_{K}).

Remark 2.2.5.

Strictly speaking, the results on the intersection cohomology can be proven by using the de Cataldo-Migliorini theorem [dCM15] only (and not using the KK-rational version [Pa16]). To see this, note that the Hodge numbers and the Hodge-Tate weights are insensitive to the base change to a finite extension of KK (in both local and global cases), and the construction of [dCM15] yields the necessary André motivated cycle over a finite extension of KK.

In case X=XBBX=X^{BB} is the Baily-Borel compactification of a Shimura variety XX^{\circ} (we refer to Ash-Mumford-Rapoport-Tang [AMRT] and Pink [Pi98] in general, and Brylinski-Labesse [BL84] and Rapoport [Ra78] in the special case of Hilbert modular varieties), one further decomposes the intersection cohomology of XX into the Hecke-isotypic components: the Hecke correspondences act on the intersection cohomology of XX, and span a \mathbb{Q}-subalgebra X,\mathcal{H}_{X,\mathbb{Q}} in the finite dimensional \mathbb{Q}-algebra EndHS(IHBd(X,))\operatorname{End}_{HS}(IH^{d}_{B}(X,\mathbb{Q})). By decomposing X,\mathcal{H}_{X,\mathbb{Q}} into a product of \mathbb{Q}-algebras, we obtain the Hecke isotypic components.

Proposition 2.2.6.

The Hecke isotypic components come from André motives over the reflex field EE.

As a consequence, Theorem 2.2.1 applies to these components.

Proof.

This boils down to first finding an André (pure Nori) motive 𝔦𝔥(X)\mathfrak{ih}(X) whose \ell-adic and Betti realisations give the \ell-adic and Betti intersection cohomology of XX; and then lifting the action of the Hecke correspondences to one on 𝔦𝔥(X)\mathfrak{ih}(X).

While the first step can be done as above in an ad hoc fashion — using a (noncanonical) resolution of singularities and using the theorems of de Cataldo, Migliorini and Patrikis — the second step is done most systematically (in our opinion) by using the theory of weight filtration. We learned the argument from S. Morel (cf. [Mo08, §5]), which uses the more recent motivic constructions of Ivorra and Morel [IM19].

(Before the details, let us stress the main point and indicate where the innovations are. The intersection complex ICIC (as a perverse sheaf in the derived category of constructible sheaves) was originally constructed as an iterated application of the 22-step procedure: taking the direct image RjRj_{\ast} under open immersions jj, and then truncating with respect to a topological stratification and a function called ‘perversity’. See the explicit formula [BBD, Prop. 2.1.11].

One of the innovations in [Mo08] was to realise ICIC (for the middle perversity) over finite fields as the truncation with respect to the weight filtration: See [Mo08, Th. 3.1.4].

This renders the extension of the Hecke operators to ICIC deceptively easy: One needs neither to worry about singularities in the boundary (which can be bad) — to which one must pay heed if one uses the original (topological) definition — nor to rely on the toroidal compactifications — of which there is no canonical choice, and many are needed to extend the Hecke correspondences. This is why we adopt the idea.

In [IM19], Ivorra and Morel construct the four operations of Grothendieck (namely ff^{\ast}_{\mathcal{M}}, f!f^{!}_{\mathcal{M}}, ff_{\ast}^{\mathcal{M}}, f!f_{!}^{\mathcal{M}} for morphisms ff between quasiprojective varieties) and the weight filtrations on the derived category of ‘perverse mixed motives’; this last abelian category is moreover shown to be equivalent to the Nori motives.

This allows one to formally apply the algorithm described in [Mo08, §5], but this time applied in the motivic derived category of [IM19], rather than in the derived category of constructible sheaves as in [Mo08]. Then the Betti and \ell-adic realisations of these constructions give rise to the Hecke operators constructed previously on the level of complexes of sheaves.)

Now let us turn to a more detailed argument, and indicate which constructions in [IM19] replace those in the parallel argument from [Mo08, §5].

Let j:XX=XBBj:X^{\circ}\hookrightarrow X=X^{BB} be the open immersion of the Shimura variety and c1,c2:YXc_{1},c_{2}:Y^{\circ}\longrightarrow X^{\circ} be the two finite étale maps that give rise to a given Hecke correspondence. With j:YYj^{\prime}:Y^{\circ}\hookrightarrow Y denoting the open immersion into the Baily-Borel compactification, the cic_{i} extend canonically to ci¯:YX\overline{c_{i}}:Y\longrightarrow X. We start with the identity correspondence

u=1:c1Xc2!X.u=1:c_{1}^{\ast}\mathbb{Q}_{X^{\circ}}\longrightarrow c_{2}^{!}\mathbb{Q}_{X^{\circ}}.

arising from the natural isomorphisms c1XYc_{1}^{\ast}\mathbb{Q}_{X^{\circ}}\simeq\mathbb{Q}_{Y^{\circ}} and c2!XYc_{2}^{!}\mathbb{Q}_{X^{\circ}}\simeq\mathbb{Q}_{Y^{\circ}}.

Here and below, the functors ff^{\ast} and f!f^{!}, etc., refer to the ones in [IM19, Th. 5.1], where the notations ff^{\ast}_{\mathcal{M}}, f!f^{!}_{\mathcal{M}}, etc., are used.

Just as in [Mo08, §5.1], but using the motivic constructions of the 44 functors and the base change morphisms, stated [IM19, Th. 5.1], we take Rj=jRj^{\prime}_{\ast}={j^{\prime}}_{\ast}^{\mathcal{M}} and use the base change morphisms

c1¯RjXRjc1X\textstyle{\overline{c_{1}}^{\ast}Rj_{\ast}\mathbb{Q}_{X^{\circ}}\longrightarrow Rj^{\prime}_{\ast}c_{1}^{\ast}\mathbb{Q}_{X^{\circ}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}u\scriptstyle{u}Rjc2!Xc2¯!RjX.\textstyle{Rj^{\prime}_{\ast}c_{2}^{!}\mathbb{Q}_{X^{\circ}}\longrightarrow\overline{c_{2}}^{!}Rj_{\ast}\mathbb{Q}_{X^{\circ}}.}

We then use the fact that the lowest weight filtration of RjXRj_{\ast}\mathbb{Q}_{X^{\circ}} is canonically isomorphic to the intersection complex j!(X[d])[d]j_{!\ast}(\mathbb{Q}_{X^{\circ}}[d])[-d]. For this, it is enough to show that the (motivic) weight filtration has \ell-adic realisation equal to the (\ell-adic) weight filtration. But this follows from the definitions and construction: [IM19, Def. 6.12, Def. 6.13, and Prop. 6.16].

Then we proceed as in [Mo08, Lem. 5.1.4] to obtain

u¯:c1¯ICXc2¯!ICX,\overline{u}:\overline{c_{1}}^{\ast}IC_{X}\longrightarrow\overline{c_{2}}^{!}IC_{X},

lifting the cohomological correspondence in realisations. Again the point is that all the arrows in display in [Mo08, Lem. 5.1.4] are constructed using (only) the functoriality of RjRj_{\ast}, the base change morphisms, and the weight filtration. In our (motivic) context, we use the main theorem [IM19, Th. 5.1] for the first two and the weight filtration constructed in [IM19, Prop. 6.16] for the last. ∎

This applies in particular to the Hilbert modular varieties and the submotive M(f)M(f) of 𝔦𝔥d(XBB)\mathfrak{ih}^{d}(X^{BB}) cut out by any new cuspform ff of parallel weight (2,,2)(2,\cdots,2) and its conjugates, and Corollary 2.2.3 applies to it.

2.3 Conjectures

Now we can formulate the analogue of Katz’s Conjecture:

Conjecture 2.4.

Let XX be a projective variety over a number field FF, and let nn be an integer.

Then there exists a finite set S=S(X,n)S=S(X,n) of maximal ideals of 𝒪F\mathcal{O}_{F} such that for every prime number \ell and every maximal ideal 𝔭\mathfrak{p} of 𝒪F\mathcal{O}_{F} outside SS and with residue characteristic \neq\ell, we have

NP(Frob𝔭|IHn(XFFs,))HTP(IHn(XFFs,)).\mathrm{NP}(\operatorname{Frob}_{\mathfrak{p}}|_{IH^{n}(X\otimes_{F}F^{s},\mathbb{Q}_{\ell})})\geq\mathrm{HTP}(IH^{n}(X\otimes_{F}F^{s},\mathbb{Q}_{\ell})).

In case XX is also smooth, this is known to be true, by theorems of Katz and Messing [KM74] (comparing the \ell-adic Frobenius at 𝔭\mathfrak{p} with the crystalline Frobenius), of Mazur [Ma73] (showing that the Newton polygon lies on or above the Hodge polygon for the crystalline cohomology), and of Faltings [Fa88] (showing that the Hodge-Tate polygon is the same as the Hodge polygon).

And we formulate the analogue of the ‘ordinariness’ conjecture:

Conjecture 2.5.

Let XX be a projective variety over a number field FF, and let nn be an integer.

For every prime number \ell, there exists an infinite set of maximal ideals 𝔭\mathfrak{p} of 𝒪F\mathcal{O}_{F} with residue characteristic \neq\ell such that

NP(Frob𝔭|IHn(XFFs,))=HTP(IHn(XFFs,)).\mathrm{NP}(\operatorname{Frob}_{\mathfrak{p}}|_{IH^{n}(X\otimes_{F}F^{s},\mathbb{Q}_{\ell})})=\mathrm{HTP}(IH^{n}(X\otimes_{F}F^{s},\mathbb{Q}_{\ell})).

We note that the right hand side of the conjectures is independent of λ\lambda or \ell by Corollary 2.2.4, and that the left hand side is independent of \ell because the IHn(X,)IH^{n}(X,\mathbb{Q}_{\ell}) form a strictly compatible system by Gabber [Fu00] and Katz-Laumon [KL85, Th. 3.1.2]. Therefore Conjectures 2.4 and 2.5 are independent of the auxiliary prime \ell.

In case XX is also smooth, Conjecture 2.5 is equivalent to Conjecture 1.1 recalled in §1, since (1) the Newton polygons of the \ell-adic and crystalline Frobenius endomorphisms are the same by Katz and Messing [KM74], and (2) the Hodge-Tate polygon coincides with the Hodge polygon by Faltings [Fa88].

3 Preparation

3.1 Notation

From this point on, F¯F\subseteq\overline{\mathbb{Q}} denotes a totally real number field of degree d=[F:]d=[F:\mathbb{Q}] and discriminant disc(F)\mathrm{disc}(F); F~\widetilde{F} is the Galois closure of F/F/\mathbb{Q}, of degree d~=[F~:]\widetilde{d}=[\widetilde{F}:\mathbb{Q}].

Let ff be a new normalised Hilbert eigencuspform of parallel weight (2,,2)(2,\cdots,2) of level 𝔫𝒪F\mathfrak{n}\subseteq\mathcal{O}_{F}. The Fourier coefficients of ff generate the number field:

Kf:=({a𝔭}𝔭),K_{f}:=\mathbb{Q}\left(\left\{a_{\mathfrak{p}}\right\}_{\mathfrak{p}}\right),

where 𝔭\mathfrak{p} ranges over the primes of 𝒪F\mathcal{O}_{F} not dividing 𝔫\mathfrak{n}. It is either a totally real number field or a CM field, and we let kf:=[Kf:]k_{f}:=[K_{f}:\mathbb{Q}].

We note that the ordinariness in this context is equivalent to the following simple condition: 𝔭\mathfrak{p} is an ordinary prime (for ff) if and only if a𝔭a_{\mathfrak{p}} is nonzero and does not belong to any prime ideal \wp of 𝒪Kf\mathcal{O}_{K_{f}} lying over (p)=𝔭(p)=\mathfrak{p}\cap\mathbb{Z}.

We fix once and for all a rational prime \ell that splits completely222we choose a split prime just for simplifying the exposition a little bit. The obvious analogues of Conjectures 2.4 and 2.5 are independent of \ell for M(f)M(f) defined below, since we still have a strictly compatible system of Galois representations. in KfK_{f}. For every nonarchimedean place λ\lambda of KfK_{f} dividing \ell, we denote by

ρ=ρf,λ:Gal(¯/F)GL2(Kf,λ)\rho=\rho_{f,\lambda}:\operatorname{Gal}(\overline{\mathbb{Q}}/F)\longrightarrow GL_{2}(K_{f,\lambda})

the associated semisimple, KfK_{f}-rational and integral Galois representation: See Deligne [D71], Ohta [Oh83], Carayol [C86], Wiles [Wi88], Taylor [T89], Blasius and Rogawski [BR93] and the references therein. The Tate twist of its determinant det(ρ)(1)\det(\rho)(1) is a character of finite order.

Let G=Gf,λG=G_{f,\lambda} be the Zariski closure of the image of ρf,λ\rho_{f,\lambda} in GL2GL_{2} over Kf,λK_{f,\lambda}. Since we assume ρ\rho to be semisimple, the derived group of the connected component (G)der=[G,G](G^{\circ})^{\mathrm{der}}=[G^{\circ},G^{\circ}] is a semisimple algebraic subgroup of SL2SL_{2}, that is, either SL2SL_{2} or trivial. If the reductive group Gf,λG^{\circ}_{f,\lambda} is a torus for some λ\lambda, we say that ff is of CM type.333The notion is independent of λ\lambda and \ell by a theorem of Serre, cf. the argument in the proof of Theorem 4.2.1.

The product

ρf,:=λ|ρf,λ:Gal(¯/F)λ|GL2(Kf,λ)=(ResKfGL2)()\rho_{f,\ell}:=\prod_{\lambda|\ell}\rho_{f,\lambda}:\operatorname{Gal}(\overline{\mathbb{Q}}/F)\longrightarrow\prod_{\lambda|\ell}GL_{2}(K_{f,\lambda})=(\operatorname{Res}^{K_{f}}_{\mathbb{Q}}GL_{2})(\mathbb{Q}_{\ell})

is \mathbb{Q}-rational and integral in the sense of Serre. We denote by Gf,G_{f,\ell} the Zariski closure over \mathbb{Q}_{\ell} of its image, and Gf,G_{f,\ell}^{\circ} its connected component.

3.2 Frobenius field and Ribet’s argument

Following Ribet [Ri76], for every finite extension FF^{\prime} of FF, we consider the Frobenius field:

Tr(ρf,λ,F):=({Trρf,λ(Frob𝔭)}𝔭)Kf\mathrm{Tr}(\rho_{f,\lambda},F^{\prime}):=\mathbb{Q}\left(\left\{\mathrm{Tr}\rho_{f,\lambda}(\operatorname{Frob}_{\mathfrak{p}^{\prime}})\right\}_{\mathfrak{p}^{\prime}}\right)\leq K_{f}

where 𝔭\mathfrak{p}^{\prime} ranges over the primes of 𝒪F\mathcal{O}_{F^{\prime}} coprime to disc(F)𝔫\mathrm{disc}(F)\mathfrak{n}\cdot\ell; it is independent of λ\lambda, since the ρf,λ\rho_{f,\lambda} form a strictly compatible system. Since kf=[Kf:]<k_{f}=[K_{f}:\mathbb{Q}]<\infty, there is the smallest Frobenius field of ff

KfKf and kf:=[Kf:].K_{f}^{\circ}\leq K_{f}\,\,\,\,\mbox{ and }\,\,\,\,k_{f}^{\circ}:=[K_{f}^{\circ}:\mathbb{Q}].

It is totally real, since det(ρf,λ)(1)\det(\rho_{f,\lambda})(1) has finite order and the eigenvalues of ρf,λ(Frob𝔭)\rho_{f,\lambda}(\operatorname{Frob}_{\mathfrak{p^{\prime}}}) are Weil integers (see Lemma 4.1.3).

We have thus the following algebraic groups:

(ResKfSL2)\textstyle{(\operatorname{Res}^{K_{f}^{\circ}}_{\mathbb{Q}}SL_{2})_{\mathbb{Q}_{\ell}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}((ResKfGL2)det×)\textstyle{((\operatorname{Res}^{K_{f}^{\circ}}_{\mathbb{Q}}GL_{2})^{\det\subseteq\mathbb{Q}^{\times}})_{\mathbb{Q}_{\ell}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}(ResKfGL2)GL2kf,\textstyle{(\operatorname{Res}^{K_{f}^{\circ}}_{\mathbb{Q}}GL_{2})_{\mathbb{Q}_{\ell}}\leq GL_{2k_{f}^{\circ},\mathbb{Q}_{\ell}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}(ResKfSL2)\textstyle{(\operatorname{Res}^{K_{f}}_{\mathbb{Q}}SL_{2})_{\mathbb{Q}_{\ell}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}((ResKfGL2)det×)\textstyle{((\operatorname{Res}^{K_{f}}_{\mathbb{Q}}GL_{2})^{\det\subseteq\mathbb{Q}^{\times}})_{\mathbb{Q}_{\ell}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}(ResKfGL2)GL2kf,\textstyle{(\operatorname{Res}^{K_{f}}_{\mathbb{Q}}GL_{2})_{\mathbb{Q}_{\ell}}\leq GL_{2k_{f},\mathbb{Q}_{\ell}}}

Here by the \mathbb{Q}-algebraic group (ResKfGL2)det×(\operatorname{Res}^{K_{f}}_{\mathbb{Q}}GL_{2})^{\det\subseteq\mathbb{Q}^{\times}}, we mean the following fibred product, which is often denoted by GG^{\ast} in the literature:

G\textstyle{G^{\ast}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝔾m,\textstyle{\mathbb{G}_{m,\mathbb{Q}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ResKfGL2\textstyle{\operatorname{Res}^{K_{f}}_{\mathbb{Q}}GL_{2}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}det\scriptstyle{\det}ResKf𝔾m,Kf,\textstyle{\operatorname{Res}^{K_{f}}_{\mathbb{Q}}\mathbb{G}_{m,K_{f}},}

and similarly with KfK_{f} replaced with KfK_{f}^{\circ}.

Proposition 3.2.1 ((Ribet)).

Suppose that ff is not of CM type. Then

Gf,=((ResKfGL2)det×).G_{f,\ell}^{\circ}=((\operatorname{Res}^{K_{f}^{\circ}}_{\mathbb{Q}}GL_{2})^{\det\subseteq\mathbb{Q}^{\times}})_{\mathbb{Q}_{\ell}}.
Proof.

This amounts to showing that the (algebraic) Lie algebra 𝔤\mathfrak{g} of Gf,G_{f,\ell}^{\circ} is equal to that of the right hand side.

The containment \subseteq follows from the fact that, if FF^{\prime} is any sufficiently large finite extension of FF, then ρf,λ(Frob𝔭)\rho_{f,\lambda}(\operatorname{Frob}_{\mathfrak{p^{\prime}}}) has trace and determinant in KfK_{f}^{\circ} for any prime 𝔭\mathfrak{p}^{\prime} of FF^{\prime} coprime to disc(F)𝔫\mathrm{disc}(F)\mathfrak{n}\ell, so that if λ\lambda and λ\lambda^{\prime} are any 22 primes of KfK_{f} that lie over the same prime of KfK_{f}^{\circ} over ()(\ell), then Gf,G^{\circ}_{f,\ell} is contained in the partial diagonal of (ResKfGL2)(\operatorname{Res}^{K_{f}}_{\mathbb{Q}}GL_{2})_{\mathbb{Q}_{\ell}} where the λ\lambda- and the λ\lambda^{\prime}-components are equal.

To prove the containment \supseteq, we first note that since ff is not of CM type, 𝔤\mathfrak{g} surjects onto each factor 𝔤𝔩2,λ\mathfrak{gl}_{2,\lambda^{\circ}}, which contains 𝔰𝔩2,λ\mathfrak{sl}_{2,\lambda^{\circ}}. If λ1\lambda_{1}^{\circ} and λ2\lambda_{2}^{\circ} are distinct primes of KfK_{f}^{\circ} lying over \ell, then the representations of 𝔤\mathfrak{g} in the 22 factors are nonisomorphic, since the representations of (germs of) Gal(¯/F)\operatorname{Gal}(\overline{\mathbb{Q}}/F) have different traces, and the image of 𝔤\mathfrak{g} in 𝔤𝔩2,λ1×𝔤𝔩2,λ2\mathfrak{gl}_{2,\lambda_{1}^{\circ}}\times\mathfrak{gl}_{2,\lambda_{2}^{\circ}} contains 𝔰𝔩2×𝔰𝔩2\mathfrak{sl}_{2}\times\mathfrak{sl}_{2}.

Then by Goursat’s Lemma (in the middle of the proof of Ribet [Ri76, Th. 4.4.10]) the image of 𝔤\mathfrak{g} contains λ|𝔰𝔩2,λ\prod_{\lambda^{\circ}|\ell}\mathfrak{sl}_{2,\lambda^{\circ}}, where λ\lambda^{\circ} ranges over the primes of KfK_{f}^{\circ} lying over ()(\ell). Since the determinant on Gf,λG_{f,\lambda^{\circ}}^{\circ} is a dominant map onto 𝐆m\mathbb{\mathbf{G}}_{m}, we get the desired equality. ∎

Definition 3.2.2.

Let FF^{\circ} be the Galois extension of FF cut out by two representations with finite image:

Gal(¯/F)=ker(ρf,:Gal(¯/F)Gf,()/Gf,()))ker(det(ρf,)(1)).\operatorname{Gal}(\overline{\mathbb{Q}}/F^{\circ})=\ker(\rho_{f,\ell}:\operatorname{Gal}(\overline{\mathbb{Q}}/F)\longrightarrow G_{f,\ell}(\mathbb{Q}_{\ell})/G^{\circ}_{f,\ell}(\mathbb{Q}_{\ell})))\cap\ker(\det(\rho_{f,\ell})(1)).

and let F~\widetilde{F^{\circ}} be the compositum FF~F^{\circ}\widetilde{F}.

3.3 Variants of Sato-Tate equidistribution

Let i1,,ikfi_{1},\cdots,i_{k_{f}^{\circ}} denote the complete set of embeddings of KfK_{f}^{\circ} into \mathbb{R}. For each maximal ideal 𝔭\mathfrak{p} of 𝒪F~\mathcal{O}_{\widetilde{F^{\circ}}} coprime to disc(F)𝔫\mathrm{disc}(F)\cdot\mathfrak{n}\cdot\ell, we let

a𝔭=Trρf,λ(Frob𝔭)a_{\mathfrak{p}}=\mathrm{Tr}\rho_{f,\lambda}(\operatorname{Frob}_{\mathfrak{p}})

and consider the set of vectors in kf\mathbb{R}^{k_{f}^{\circ}}:

A(f)={(i1(a𝔭)𝔭,,ikf(a𝔭)𝔭)}𝔭.A(f)=\left\{\left(\frac{i_{1}(a_{\mathfrak{p}})}{\sqrt{\mathbb{N}\mathfrak{p}}},\cdots,\frac{i_{k_{f}^{\circ}}(a_{\mathfrak{p}})}{\sqrt{\mathbb{N}\mathfrak{p}}}\right)\right\}_{\mathfrak{p}}.
Definition 3.3.1.

We say that ff satisfies (SST) if A(f)A(f) is equidistributed in the kfk_{f}^{\circ}-fold product of the Sato-Tate (half-circle) measure on [2,2][-2,2]; say that ff satisfies (RST) if A(f)A(f) is equidistributed in a measure φ(𝐱)dμL(𝐱)\varphi(\mathbf{x})d\mu_{L}(\mathbf{x}), where φ:[2,2]kf0\varphi:[-2,2]^{k_{f}^{\circ}}\longrightarrow\mathbb{R}_{\geq 0} is a continuous function and dμLd\mu_{L} is the Lebesgue measure.

For an integer t[1,kf]t\in[1,k_{f}^{\circ}], we say that ff satisfies (tt-ST’) if there exists a sequence 1j1<j2<<jtkf1\leq j_{1}<j_{2}<\cdots<j_{t}\leq k_{f}^{\circ} such that the projection prj1,,jt(A(f))[2,2]t\mathrm{pr}_{j_{1},\cdots,j_{t}}(A(f))\subseteq[-2,2]^{t} is equidistributed in the tt-fold product of the Sato-Tate measure on [2,2][-2,2]; we say that ff satisfies (tt-ST) if for all sequences 𝐣\mathbf{j} of length tt, pr𝐣(A(f))\mathrm{pr}_{\mathbf{j}}(A(f)) is equidistributed in the tt-fold product of the Sato-Tate measure.

We expect the strongest (SST) to be true; it fits into Serre’s general framework of Sato-Tate equidistribution [Se12, Chpt. 8], for almost all primes \ell (depending on ff). 444The construction in §8.3 of op. cit., as stated, deals only with representations that come from the \ell-adic cohomology of algebraic varieties, but appears to use the condition only to the extent that they are rational and Hodge-Tate. In case dd is odd or the automorphic representation corresponding to ff has a discrete series at some finite prime, the Hodge-Tate condition (even the de Rham condition) for all \ell follows from the motivic nature of the available constructions and theorems of Faltings [Fa89], see Blasius and Rogawski [BR93]. In the general case, the Hodge-Tate condition is known for all but finitely many \ell: See Taylor [T95], where they are shown to be (even) crystalline. Namely, the compact Lie group attached to ρf,|F~\rho_{f,\ell}|_{\widetilde{F^{\circ}}} by Serre is the product of kfk_{f}^{\circ} copies of SU2SU_{2} and the axioms (A1)(A_{1}) and (A2)(A_{2}) should hold.

When t<kft<k_{f}^{\circ}, the mere conjunction of (RST) and (tt-ST) does not imply (SST).

Remark 3.3.2.

The condition (SST) is stronger than the (usual) Sato-Tate equidistribution theorems available at the moment (see [HSBT10], [BLGHT11], [BLGG11]).

In order to prove (SST) in the manner that the aforementioned results were obtained, one would need to control the LL-functions not only of the symmetric powers:

Symmj(ij(f))\mathrm{Sym}^{m_{j}}(i_{j}(f))

(through potential automorphy), but also of their tensor products:

Symm1(i1(f))Symmkf(ikf(f)).\mathrm{Sym}^{m_{1}}(i_{1}(f))\otimes\cdots\otimes\mathrm{Sym}^{m_{k_{f}^{\circ}}}(i_{k_{f}^{\circ}}(f)).

for all tuples (m1,,mkf)(m_{1},\cdots,m_{k_{f}^{\circ}}).

The case of (tt-ST), for t2t\leq 2, looks accessible, see Harris [Ha09].

3.4 Multisets and Newton Polygons

Definition 3.4.6, the operations \otimes, \oplus and the partial order will be used in Theorem 4.1.1.

We consider finite subsets with positive finite multiplicities, or simply multisets, of \mathbb{Q}. For example, {1/3,2/3}\{1/3,2/3\} (each with multiplicity 11) and {1/2,1/2}\{1/2,1/2\} (with multiplicity 22).

Definition 3.4.1.

Let S={s1,,sm}S=\{s_{1},\cdots,s_{m}\} and T={t1,,tn}T=\{t_{1},\cdots,t_{n}\} be multisets. Define the sum

ST={s1,,sm,t1,,tn},S\oplus T=\{s_{1},\cdots,s_{m},t_{1},\cdots,t_{n}\},

the product

ST={si+tj}1im,  1jn,S\otimes T=\{s_{i}+t_{j}\}_{1\leq i\leq m,\,\,1\leq j\leq n},

and the dual

S={s1,,sm}.S^{\vee}=\{-s_{1},\cdots,-s_{m}\}.

Also, for k>0k>0, we write Sk=SSS^{\oplus k}=S\oplus\cdots\oplus S and Sk=SSS^{\otimes k}=S\otimes\cdots\otimes S, repeated kk times.

The cardinality is denoted by |S||S| or rkS\operatorname{rk}S (which is mm for the SS as above) and

S:=s1++sm\int S:=s_{1}+\cdots+s_{m}
Proposition 3.4.2.

Multisets form a commutative semiring with involution, in which the empty set is the additive neutral element and {0}\{0\} the multiplicative identity element.

The map S|S|S\mapsto|S| is a semiring homomorphism into the natural numbers and

(ST)=S+T and (ST)=|T|S+|S|T.\int(S\oplus T)=\int S+\int T\,\,\,\mbox{ and }\,\,\,\int(S\otimes T)=|T|\int S+|S|\int T.

Given a multiset consisting of a1ana_{1}\leq\cdots\leq a_{n} we form its Newton polygon emanating from (0,0)(0,0) with the slopes a1,,ana_{1},\cdots,a_{n} (in this order). Conversely, any finite Newton polygon emanating from the origin and with rational slopes uniquely determines a multiset of \mathbb{Q}.

From this point on, we will thus identify multisets with Newton polygons. This allows us to impose a partial order on the class of multisets:

ST if and only if |S|=|T| and NP(S)NP(T);S\leq T\,\,\,\,\mbox{ if and only if }\,\,\,\,|S|=|T|\,\,\mbox{ and }\,\,\mathrm{NP}(S)\leq\mathrm{NP}(T);

the last meaning that NP(T)\mathrm{NP}(T) lies on or above NP(S)\mathrm{NP}(S).

Proposition 3.4.3.

Let SSS\leq S^{\prime} and TT be three multisets. Then (1) STSTS\oplus T\leq S^{\prime}\oplus T and (2) STSTS\otimes T\leq S^{\prime}\otimes T. If, in addition, SS and SS^{\prime} end at the same point, then (3) SSS^{\vee}\leq S^{\prime\vee}.

Proof.

(1) By induction on |T||T|, we are reduced to the case where TT consists of 11 element, say T={t}T=\{t\}. Twisting by t-t (i.e. taking {t}\otimes\{-t\}) allows us to assume that t=0t=0. Enumerate SS and SS^{\prime} in the order:

s1s2sm and s1s2sms_{1}\leq s_{2}\leq\cdots\leq s_{m}\,\,\,\mbox{ and }\,\,\,s^{\prime}_{1}\leq s^{\prime}_{2}\leq\cdots\leq s^{\prime}_{m}

and let aa and bb be such that:

sa<0sa+1 and sb<0sb+1;s_{a}<0\leq s_{a+1}\,\,\,\mbox{ and }\,\,\,s^{\prime}_{b}<0\leq s^{\prime}_{b+1};

if all the sis_{i} are 0\geq 0 (resp. <0<0), then we let a:=0a:=0 (resp. a:=ma:=m), and similarly for bb.

If we define ΣS(i):=s1++si\Sigma_{S}(i):=s_{1}+\cdots+s_{i} for i[0,m]i\in[0,m], the condition SSS\leq S^{\prime} becomes

ΣS(i)ΣS(i) for all i[0,m].\Sigma_{S}(i)\leq\Sigma_{S^{\prime}}(i)\,\,\,\mbox{ for all }\,\,i\in[0,m].

We need to prove

ΣS{0}(i)ΣS{0}(i) for all i[0,m+1].\Sigma_{S\oplus\{0\}}(i)\leq\Sigma_{S^{\prime}\oplus\{0\}}(i)\,\,\,\mbox{ for all }\,\,i\in[0,m+1]. (3.4.3.1)

(1a) Suppose that a<ba<b. Then for i[0,a][b+1,m+1]i\in[0,a]\cup[b+1,m+1], (3.4.3.1) is clearly satisfied. For i[a+1,b]i\in[a+1,b], since 0 is inserted into SS^{\prime} at the (b+1)(b+1)-th place, we have

ΣS{0}(i)\displaystyle\Sigma_{S^{\prime}\oplus\{0\}}(i) =\displaystyle= ΣS{0}(b+1)(si+1++sb+0)\displaystyle\Sigma_{S^{\prime}\oplus\{0\}}(b+1)-(s^{\prime}_{i+1}+\cdots+s^{\prime}_{b}+0)
=\displaystyle= ΣS(b)(si+1++sb)ΣS(b)\displaystyle\Sigma_{S^{\prime}}(b)-(s^{\prime}_{i+1}+\cdots+s^{\prime}_{b})\geq\Sigma_{S^{\prime}}(b)

while, since 0 is inserted into SS at the (a+1)(a+1)-th place, we have

ΣS{0}(i)\displaystyle\Sigma_{S\oplus\{0\}}(i) =\displaystyle= ΣS(i1)=ΣS(b)(si++sb)ΣS(b).\displaystyle\Sigma_{S}(i-1)=\Sigma_{S}(b)-(s_{i}+\cdots+s_{b})\leq\Sigma_{S}(b).

and we get (3.4.3.1).

(1b) Suppose that aba\geq b. Then again for i[0,b][a+1,m+1]i\in[0,b]\cup[a+1,m+1], (3.4.3.1) is trivially satisfied. For i[b+1,a]i\in[b+1,a], we have this time:

ΣS{0}(i)\displaystyle\Sigma_{S^{\prime}\oplus\{0\}}(i) =\displaystyle= ΣS(i1)=ΣS(b)+(sb+1++si1)ΣS(b)\displaystyle\Sigma_{S^{\prime}}(i-1)=\Sigma_{S^{\prime}}(b)+(s^{\prime}_{b+1}+\cdots+s^{\prime}_{i-1})\geq\Sigma_{S^{\prime}}(b)
ΣS{0}(i)\displaystyle\Sigma_{S\oplus\{0\}}(i) =\displaystyle= ΣS(i)=ΣS(b)+(sb+1++si)ΣS(b)\displaystyle\Sigma_{S}(i)=\Sigma_{S}(b)+(s_{b+1}+\cdots+s_{i})\leq\Sigma_{S^{\prime}}(b)

This completes the proof of (1).

(2) By decomposing TT into singletons and using the distributive law, we deduce (2) from (1).

(3) The duals SS^{\vee} and SS^{\prime\vee} are enumerated:

smsm1s1 and smsm1s1.-s_{m}\leq-s_{m-1}\leq\cdots\leq-s_{1}\,\,\,\mbox{ and }\,\,\,-s^{\prime}_{m}\leq-s^{\prime}_{m-1}\leq\cdots\leq-s^{\prime}_{1}.

The assumption that SS and SS^{\prime} end at the same point means that ΣS(m)=ΣS(m)\Sigma_{S}(m)=\Sigma_{S^{\prime}}(m). Thus

ΣS(i)=ΣS(mi)ΣS(m)ΣS(mi)ΣS(m)=ΣS(i)\Sigma_{S^{\vee}}(i)=\Sigma_{S}(m-i)-\Sigma_{S}(m)\leq\Sigma_{S^{\prime}}(m-i)-\Sigma_{S^{\prime}}(m)=\Sigma_{S^{\prime\vee}}(i)

for all i[1,m]i\in[1,m], and this completes the proof of the Proposition. ∎

Remark 3.4.4.

In view of (3), one may want to consider the more restrictive partial order:

ST if and only if |S|=|T|,S=T, and NP(S)NP(T),S\leq^{\prime}T\,\,\,\mbox{ if and only if }\,\,\,|S|=|T|,\int S=\int T,\mbox{ and }\mathrm{NP}(S)\leq\mathrm{NP}(T),

so as to make the involution SSS\mapsto S^{\vee} order-preserving. Below, we use the partial order \leq only in the case where \leq^{\prime} also applies.

We are particularly interested in:

Definition 3.4.5.

By the partially ordered semiring of integral Newton polygons, we mean the subsemiring of multisets whose Newton polygons have integral breaking points.

The following polygons appear in the statement of Theorem 4.1.1.

Definition 3.4.6.

Let d1d\geq 1, k1k\geq 1 and i[0,k]i\in[0,k] be integers. We define the multiset (and the corresponding Newton polygon):

P(d;k,i):=({0,1}d)(ki)({1/2,1/2}d)iP(d;k,i):=\left(\{0,1\}^{\otimes d}\right)^{\oplus(k-i)}\oplus\left(\{1/2,1/2\}^{\otimes d}\right)^{\oplus i}

The Newton polygon of P(d;k,i)P(d;k,i) has integral breaking points. By Proposition 3.4.3, we have

P(d;k,i)P(d;k,j) if and only if ij.P(d;k,i)\leq P(d;k,j)\,\,\,\,\mbox{ if and only if }\,\,\,\,i\leq j.

3.5 Interaction of FF and KK: Slope and Bisection

Let GG be a group acting on a finite set XX.

Definition 3.5.1.

By the length of maximal parts of gGg\in G on XX, which we denote by λ(g,X)\lambda(g,X), we mean the largest of the cardinalities of the gg-orbits in XX. We define λ(G,X)\lambda(G,X) as the supremum of λ(g,X)\lambda(g,X), as gg ranges over GG.

Definition 3.5.2.

We say that an element gGg\in G bisects XX if gg has exactly 22 orbits in XX and the orbits have the same number of elements.

Let FF be a (ground) number field, KK a (coefficient) number field, and FsF^{s} an algebraic closure of FF. The Galois group G:=Gal(Fs/F)G:=\operatorname{Gal}(F^{s}/F) acts continuously on the discrete set

X:=Hom(K,Fs)X:=\operatorname{Hom}(K,F^{s})

of field embeddings of KK into FsF^{s}.

Definition 3.5.3.

We define

λF(K):=λ(Gal(Fs/F),Hom(K,Fs)).\lambda_{F}(K):=\lambda(\operatorname{Gal}(F^{s}/F),\operatorname{Hom}(K,F^{s})).

When F=F=\mathbb{Q}, we drop FF from the notation and write λ(K)\lambda(K).

In more concrete terms: When F=F=\mathbb{Q}, the Galois group of the normal closure K~\widetilde{K} of K/K/\mathbb{Q} determines λ(K)\lambda(K). For example, if the group is the full symmetric group of degree [K:][K:\mathbb{Q}] or the cyclic group of [K:][K:\mathbb{Q}] elements, then λ(K)=[K:]\lambda(K)=[K:\mathbb{Q}]. If the group is the alternating group, then λ(K)=[K:]\lambda(K)=[K:\mathbb{Q}] (resp. =[K:]1=[K:\mathbb{Q}]-1) if [K:][K:\mathbb{Q}] is odd (resp. even).

The notion of bisection is also determined by the Galois group; for example, the alternating group of even degree and the Klein 44-group acting on itself by translation contain bisecting elements.

For general FF, one needs to look at the action of the subgroup Gal(¯/F)Gal(¯/).\operatorname{Gal}(\overline{\mathbb{Q}}/F)\leq\operatorname{Gal}(\overline{\mathbb{Q}}/\mathbb{Q}).

Definition 3.5.4.

Given two number fields FF and KK, we define the slope of KK over FF:

σF(K)=1λF(K)[K:][0,1).\sigma_{F}(K)=1-\frac{\lambda_{F}(K)}{[K:\mathbb{Q}]}\,\,\,\,\in[0,1)\cap\mathbb{Q}.

When F=F=\mathbb{Q}, we write σ(K)=σ(K)\sigma(K)=\sigma_{\mathbb{Q}}(K).

We call σ\sigma the slope, in view of the following ‘semistability’ property, formally analagous to that of Harder and Narasimhan for vector bundles on curves, in the variable KK:

Proposition 3.5.5.

Let KK^{\prime} be a subfield of KK and let n=[K:K]n=[K:K^{\prime}]. Then

nλ(g,Hom(K,Fs))λ(g,Hom(K,Fs))n\cdot\lambda(g,\operatorname{Hom}(K^{\prime},F^{s}))\geq\lambda(g,\operatorname{Hom}(K,F^{s}))

for all gG=Gal(Fs/F)g\in G=\operatorname{Gal}(F^{s}/F) and therefore

σF(K)σF(K).\sigma_{F}(K^{\prime})\leq\sigma_{F}(K).
Proof.

Let X=Hom(K,Fs)X=\operatorname{Hom}(K,F^{s}) and X:=Hom(K,Fs)X^{\prime}:=\operatorname{Hom}(K^{\prime},F^{s}). Then we have a surjective map of GG-sets:

XXX\longrightarrow X^{\prime}

obtained by restriction. Since K/KK/K^{\prime} is separable, each fibre has exactly nn elements.

The first inequality follows from inspecting the images of the gg-orbits in XX, and the second follows from the first by definition. ∎

In the variable FF, we trivially have

σF(K)σF(K) if FF.\sigma_{F}(K)\leq\sigma_{F^{\prime}}(K)\,\,\,\,\mbox{ if }\,\,\,\,F\subseteq F^{\prime}.

The following is useful in computing σ\sigma and finding bisecting elements in practice.

Proposition 3.5.6.

Let FF and KK be two number fields, with the respective normal closures F~\widetilde{F} and K~\widetilde{K} over \mathbb{Q}.

  1. (1)

    If [K:][K:\mathbb{Q}] is a prime number not dividing [F:][F:\mathbb{Q}], then σF(K)=0\sigma_{F}(K)=0.

  2. (2)

    If Gal(K~/)\operatorname{Gal}(\widetilde{K}/\mathbb{Q}) is the symmetric group of degree [K:][K:\mathbb{Q}] and if [F~:][\widetilde{F}:\mathbb{Q}] is odd, then σF~(K)=0\sigma_{\widetilde{F}}(K)=0.

  3. (3)

    Suppose that K~\widetilde{K} is linearly disjoint from F~\widetilde{F} over \mathbb{Q} (which is the case, for example, if FF and KK have coprime discriminants). Then σF(K)=σ(K)\sigma_{F}(K)=\sigma_{\mathbb{Q}}(K), and Gal(¯/F)\operatorname{Gal}(\overline{\mathbb{Q}}/F) possesses an element bisecting Hom(K,¯)\operatorname{Hom}(K,\overline{\mathbb{Q}}) exactly when Gal(¯/)\operatorname{Gal}(\overline{\mathbb{Q}}/\mathbb{Q}) does.

Proof.

(1) Recall that Gal(K~/)\operatorname{Gal}(\widetilde{K}/\mathbb{Q}) acts transitively on Hom(K,)\operatorname{Hom}(K,\mathbb{Q}), and therefore has order divisible by p=[K:]p=[K:\mathbb{Q}]. Since the image of Gal(¯/F)\operatorname{Gal}(\overline{\mathbb{Q}}/F) in Gal(K~/)\operatorname{Gal}(\widetilde{K}/\mathbb{Q}) (via Gal(¯/)\operatorname{Gal}(\overline{\mathbb{Q}}/\mathbb{Q})) has index dividing [F:][F:\mathbb{Q}], the image also has order divisible by pp. Therefore the image contains an element gg of exact order pp, which has λ(g,Hom(K,))=p=[K:]\lambda(g,\operatorname{Hom}(K,\mathbb{Q}))=p=[K:\mathbb{Q}].

(2) Use the fact that a symmetric group has no proper normal subgroup of odd index to deduce that the image of Gal(¯/F~)\operatorname{Gal}(\overline{\mathbb{Q}}/\widetilde{F}) in Gal(K~/)\operatorname{Gal}(\widetilde{K}/\mathbb{Q}) is the full symmetric group.

(3) By assumption, the natural map

(ϕ1,ϕ2):Gal(¯/)Gal(F~/)×Gal(K~/)(\phi_{1},\phi_{2}):\operatorname{Gal}(\overline{\mathbb{Q}}/\mathbb{Q})\longrightarrow\operatorname{Gal}(\widetilde{F}/\mathbb{Q})\times\operatorname{Gal}(\widetilde{K}/\mathbb{Q})

is surjective and by definition Gal(¯/F)Gal(¯/F~)=ker(ϕ1)\operatorname{Gal}(\overline{\mathbb{Q}}/F)\supseteq\operatorname{Gal}(\overline{\mathbb{Q}}/\widetilde{F})=\ker(\phi_{1}). Therefore Gal(¯/F)\operatorname{Gal}(\overline{\mathbb{Q}}/F) surjects onto Gal(K~/)\operatorname{Gal}(\widetilde{K}/\mathbb{Q}). The statements about the slope and bisecting elements follow from this. ∎

3.6 Zariski density and Haar density (Serre)

We will use the following in the proof of Theorem 4.1.1.

Lemma 3.6.1.

Let EE be a finite extension of \mathbb{Q}_{\ell}, GG a connected algebraic group over EE, ΓG(E)\Gamma\leq G(E) a compact and Zariski dense subgroup, and

φ:G𝔸E1\varphi:G\longrightarrow\mathbb{A}^{1}_{E}

a regular morphism of algebraic varieties that is constant on the conjugacy classes.

Then for any finite subset SS of E=𝔸1(E)E=\mathbb{A}^{1}(E) that does not contain φ(1G)\varphi(1_{G}), the subset Γφ1(S)\Gamma\cap\varphi^{-1}(S) has Haar measure 0 in Γ\Gamma.

In particular, if Γ\Gamma is the image of a continuous representation ρ\rho of Gal(¯/F)\operatorname{Gal}(\overline{\mathbb{Q}}/F) unramified outside a finite set of primes of the number field FF, then the set of primes 𝔭\mathfrak{p} in FF such that φ(ρ(Frob𝔭))S\varphi(\rho(\operatorname{Frob}_{\mathfrak{p}}))\in S has (natural) density 0.

Proof.

This follows from Serre [Se12, Prop.5.12] applied to Z:=φ1(S)Z:=\varphi^{-1}(S), which is a proper algebraic subset of GG by the assumptions and has Zariski density 0 by definition. ∎

4 Main Theorems

4.1 Non CM case

The notation and terminology in the following theorem are explained in the previous preparatory section. References to the precise subsections are provided as they arise.

Theorem 4.1.1.

Let ff be a new normalised Hilbert eigencuspform of level 𝔫𝒪F\mathfrak{n}\subseteq\mathcal{O}_{F} and parallel weight (2,,2)(2,\cdots,2), and suppose that it is not of CM type (§3.1).

Denote by M(f)M(f) the André motive (see Proposition 2.2.6), whose realisations give the part of the intersection cohomology of the Hilbert modular variety corresponding to {σ(f)}σ\{\sigma(f)\}_{\sigma}, where σ\sigma ranges over all the embeddings of KfK_{f} into ¯\overline{\mathbb{Q}}.

  1. (1)

    (Analogue of the Katz Conjecture) For all rational primes pp coprime to disc(F)𝔫\mathrm{disc}(F)\cdot\mathfrak{n}\cdot\ell, we have

    NP(Frobp|M(f))HTP(M(f)).\mathrm{NP}(\operatorname{Frob}_{p}|_{M(f)})\geq\mathrm{HTP}(M(f)).

    Moreover, if pp splits completely in FF (equivalently in F~\widetilde{F}) and pp is unramified in KfK_{f}, then there exists an integer k(p)[0,kf]k(p)\in[0,k_{f}] such that

    NP(Frobp|M(f))=P(d;kf,k(p)).\mathrm{NP}(\operatorname{Frob}_{p}|_{M(f)})=P(d;k_{f},k(p)).

    (Here kf=[Kf:]k_{f}=[K_{f}:\mathbb{Q}] and we refer to Definition 3.4.6 for the right hand side.)

    In the remaining parts, we only consider the primes splitting completely in FF and unramified in KfK_{f}.

  2. (2)

    For a principally abundant set of primes pp, we have

    k(p)kf2.k(p)\leq\frac{k_{f}}{2}.
  3. (3)

    If kf=[Kf:]2k_{f}^{\circ}=[K_{f}^{\circ}:\mathbb{Q}]\leq 2 (KfK_{f}^{\circ} is defined in §3.2), then for a principally abundant set of primes pp, we have k(p)=0k(p)=0, that is, the Newton and Hodge-Tate polygons coincide.

  4. (4)

    For an abundant set of primes pp, we have (σ\sigma defined in §3.5)

    k(p)kfmin{1/2,σF~(Kf)}k(p)\leq k_{f}\cdot\mathrm{min}\left\{1/2,\sigma_{\widetilde{F}}(K_{f})\right\}
  5. (4bis)

    For an abundant set of primes pp, we have (F~\widetilde{F^{\circ}} defined in §3.2)

    k(p)kfmin{1/2,σF~(Kf)}k(p)\leq k_{f}\cdot\mathrm{min}\left\{1/2,\sigma_{\widetilde{F^{\circ}}}(K_{f}^{\circ})\right\}
  6. (5)

    If kfk_{f}^{\circ} is even, suppose that ff satisfies (RST) (resp. (tt-ST’ for an integer t1t\geq 1), as defined in §3.3. Then for a principally abundant set of primes pp (resp. for an abundant set of primes pp), we have

    k(p)kfkfkf12;k(p)\leq\frac{k_{f}}{k_{f}^{\circ}}\left\lfloor\frac{k_{f}^{\circ}-1}{2}\right\rfloor;

    in particular, k(p)<kf/2k(p)<k_{f}/2.

  7. (6)

    Suppose that ff satisfies (RST) and that an element of Gal(¯/F~)\operatorname{Gal}(\overline{\mathbb{Q}}/\widetilde{F^{\circ}}) bisects Hom(Kf,¯)\operatorname{Hom}(K_{f}^{\circ},\overline{\mathbb{Q}})3.5). Then for an abundant set of primes pp, we have k(p)=0k(p)=0, i.e., the Newton and Hodge-Tate polygons coincide.

Proof.

We have already fixed a rational prime \ell that splits completely in KfK_{f}. Now for all rational primes pp\neq\ell, we fix once and for all an isomorphism ¯\textstyle{\overline{\mathbb{Q}}_{\ell}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\scriptstyle{\sim}¯p,\textstyle{\overline{\mathbb{Q}}_{p},} and pull back the pp-adic valuation on the target to get a rank-11 (discontinuous) valuation vpv_{p} on ¯\overline{\mathbb{Q}}_{\ell}, normalised by vp(p)=1v_{p}(p)=1.

To prove (1), we may pass to F~\widetilde{F}, and consider Frob𝔭\operatorname{Frob}_{\mathfrak{p}} for any prime 𝔭\mathfrak{p} lying over pp, because doing so does not change the Newton polygon.

The key fact (from Brylinski and Labesse [BL84]) that we use from the constructions (see Deligne [D71], Ohta [Oh83], Carayol [C86], Wiles [Wi88], Taylor [T89] and Blasius and Rogawski [BR93] and the references therein) of the Galois representations associated with the {σ(f)}\{\sigma(f)\} is the following: the \ell-adic étale realisation M(f)M(f)_{\ell} of M(f)M(f) is the direct sum of the tensor inductions (see Curtis and Reiner [CR, §80C]):

σHom(Kf,)-IndGal(¯/F)Gal(¯/)(ρσ(f),λ).\bigoplus_{\sigma\in\operatorname{Hom}(K_{f},\mathbb{Q}_{\ell})}\otimes\mbox{-}\operatorname{Ind}^{\operatorname{Gal}(\overline{\mathbb{Q}}/\mathbb{Q})}_{\operatorname{Gal}(\overline{\mathbb{Q}}/F)}(\rho_{\sigma(f),\lambda}). (4.1.1.1)

This implies that for transversals (coset representatives) g1,,gdGal(¯/)g_{1},\cdots,g_{d}\in\operatorname{Gal}(\overline{\mathbb{Q}}/\mathbb{Q}) of Gal(¯/)\operatorname{Gal}(\overline{\mathbb{Q}}/\mathbb{Q}) modulo Gal(¯/F)\operatorname{Gal}(\overline{\mathbb{Q}}/F), we have: For any element γ\gamma in the finite index normal subgroup Gal(¯/F~)\operatorname{Gal}(\overline{\mathbb{Q}}/\widetilde{F}) of Gal(¯/F)\operatorname{Gal}(\overline{\mathbb{Q}}/F), the action of γ\gamma on M(f)M(f)_{\ell} is given by

σ(ρσ(f),λ(g1γg11)ρσ(f),λ(g2γg21)ρσ(f),λ(gdγgd1))\bigoplus_{\sigma}\left(\rho_{\sigma(f),\lambda}(g_{1}\gamma g_{1}^{-1})\otimes\rho_{\sigma(f),\lambda}(g_{2}\gamma g_{2}^{-1})\otimes\cdots\otimes\rho_{\sigma(f),\lambda}(g_{d}\gamma g_{d}^{-1})\right) (4.1.1.2)

Let K:=Kf(Frob𝔭)K^{\prime}:=K_{f}(\operatorname{Frob}_{\mathfrak{p}}) be the splitting field over KfK_{f} of the polynomial

X2Tr(ρ(Frob𝔭))X+det(ρ(Frob𝔭)),X^{2}-\mathrm{Tr}(\rho(\operatorname{Frob}_{\mathfrak{p}}))X+\det(\rho(\operatorname{Frob}_{\mathfrak{p}})), (4.1.1.3)

and let R𝔭={α𝔭,β𝔭}KR_{\mathfrak{p}}=\{\alpha_{\mathfrak{p}},\beta_{\mathfrak{p}}\}\subset K^{\prime} be the roots. Each embedding σ:Kf\sigma:K_{f}\longrightarrow\mathbb{Q}_{\ell} extends to (at most 22) embeddings σ:K¯\sigma^{\prime}:K^{\prime}\longrightarrow\overline{\mathbb{Q}}_{\ell}, and the image σ(R𝔭)¯\sigma^{\prime}(R_{\mathfrak{p}})\subset\overline{\mathbb{Q}}_{\ell} is independent of the choice σ\sigma^{\prime}. Therefore we can unambiguously form the multiset of slopes555this may not have integral breaking points:

S𝔭,σ={vp(σ(α𝔭))vp(𝔭),vp(σ(β𝔭))vp(𝔭)}.S_{\mathfrak{p},\sigma}=\left\{\frac{v_{p}(\sigma^{\prime}(\alpha_{\mathfrak{p}}))}{v_{p}(\mathbb{N}\mathfrak{p})},\frac{v_{p}(\sigma^{\prime}(\beta_{\mathfrak{p}}))}{v_{p}(\mathbb{N}\mathfrak{p})}\right\}.

Since Tr(ρ(Frob𝔭))\mathrm{Tr}(\rho(\operatorname{Frob}_{\mathfrak{p}})) is an algebraic integer, its pp-adic valuation is 0\geq 0. Also, det(ρ(Frob𝔭))\det(\rho(\operatorname{Frob}_{\mathfrak{p}})) is 𝔭\mathbb{N}\mathfrak{p} times a root of unity, so its pp-adic valuation is equal to vp(𝔭)v_{p}(\mathbb{N}\mathfrak{p}). These two facts imply the inequalities on the σ\sigma-slopes of (4.1.1.3) for all σ\sigma:

{0,1}S𝔭,σ{1/2,1/2}\{0,1\}\leq S_{\mathfrak{p},\sigma}\leq\{1/2,1/2\} (4.1.1.4)

(see §3.4 for the partial order by Newton polygon). Moreover, if 𝔭=p\mathbb{N}\mathfrak{p}=p (in particular, if pp splits completely in FF) and pp is unramified in KfK_{f}, then one of the two inequalities must be an equality.

In view of the description of cohomology in terms of tensor induction (4.1.1.1) and (4.1.1.2), we have

NP(Frob𝔭|M(f))=σHom(Kf,)S𝔭,σd,\mathrm{NP}(\operatorname{Frob}_{\mathfrak{p}|_{M(f)}})=\bigoplus_{\sigma\in\operatorname{Hom}(K_{f},\mathbb{Q}_{\ell})}S_{\mathfrak{p},\sigma}^{\otimes d},

and since

HTP(M(f))=HP(M(f))=P(d;kf,0)=({0,1}d)kf,\mathrm{HTP}(M(f))=HP(M(f))=P(d;k_{f},0)=(\{0,1\}^{\otimes d})^{\oplus k_{f}},

we have proven (1).

Remark 4.1.2.

We also get a bound for the denominators of the slopes: they are divisors of integers in the interval [1,d~][1,\widetilde{d}], or equal to 22.

In order to proceed further, we first recall the following known fact (generalised Ramanujan-Petersson conjecture, see Taylor [T95] and Blasius[B06]):

Lemma 4.1.3.

The roots α𝔭\alpha_{\mathfrak{p}} and β𝔭\beta_{\mathfrak{p}} are 𝔭\mathbb{N}\mathfrak{p}-Weil integers of weight 11.

Proof of Lemma. In case dd is odd or the automorphic representation πf\pi_{f} corresponding to ff is a discrete series representation at some finite prime, this follows from the essentially motivic nature of some of the constructions, see Blasius and Rogawski [BR93], together with Deligne’s proof of the Weil conjectures [D74].

This can be proved in the general case, and only with the a priori non motivic constructions of Wiles [Wi88] and Taylor [T89]. Note that by the description (4.1.1.1), the algebraic integers σ(α𝔭d) and σ(β𝔭d)\sigma^{\prime}(\alpha_{\mathfrak{p}}^{d})\,\mbox{ and }\,\sigma^{\prime}(\beta_{\mathfrak{p}}^{d}) for any embedding σ\sigma^{\prime} of KK^{\prime} into ¯\overline{\mathbb{Q}}_{\ell}, are eigenvalues of Frob𝔭\operatorname{Frob}_{\mathfrak{p}} acting on the IHdIH^{d} of the Baily-Borel compactification XBB(𝔫)X^{BB}(\mathfrak{n}) of the Hilbert modular variety.

Now this last variety admits a surjective, generically finite map from a projective smooth toroidal compactification over (p)\mathbb{Z}_{(p)} [Ra78].

By the decomposition theorem for perverse sheaves [BBD], the 22 algebraic integers therefore appear as eigenvalues of Frob𝔭\operatorname{Frob}_{\mathfrak{p}} in the HdH^{d} of the projective smooth variety. Then by Deligne’s proof of the Weil conjectures, they have all the archimedean absolute values =(𝔭)d/2=(\mathbb{N}\mathfrak{p})^{d/2}. By taking the dd-th root, we get the Lemma. \square

From this point on, we assume that 𝔭\mathfrak{p} is a prime of absolute degree 11 over (p)(p) (in addition to being coprime to disc(F)𝔫\mathrm{disc}(F)\cdot\mathfrak{n}\cdot\ell). We also assume that pp is unramified in KfK_{f}.

(2) Now we look more closely at

a𝔭:=Tr(ρf,λ(Frob𝔭)).a_{\mathfrak{p}}:=\mathrm{Tr}(\rho_{f,\lambda}(\operatorname{Frob}_{\mathfrak{p}})).

By the assumption that ff is not of CM type, the image of Gal(/F~)\operatorname{Gal}(\mathbb{Q}/\widetilde{F}) under ρf,λ\rho_{f,\lambda} is Zariski dense in GL2GL_{2} over Kf,λK_{f,\lambda}. Since Tr\mathrm{Tr} is a regular morphism of the algebraic variety GL2GL_{2} into 𝔸1\mathbb{A}^{1} and takes value 202\neq 0 at the identity element I2I_{2}, the set of primes 𝔭\mathfrak{p} of F~\widetilde{F} such that a𝔭=0a_{\mathfrak{p}}=0 has density zero by Lemma 3.6.1. We exclude them from this point on.

Let 1,,m\wp_{1},\cdots,\wp_{m} be the primes of KfK_{f} lying over (p)=𝔭(p)=\mathfrak{p}\cap\mathbb{Z}, and write the ideal factorisation

a𝔭𝒪Kf=1e1kemI,a_{\mathfrak{p}}\cdot\mathcal{O}_{K_{f}}=\wp_{1}^{e_{1}}\cdots\wp_{k}^{e_{m}}\cdot I, (4.1.3.1)

where II is an integral ideal coprime to pp, and we carry on with the argument preceeding the Lemma.

For each embedding σ:Kf\sigma:K_{f}\longrightarrow\mathbb{Q}_{\ell}, let i(σ)\wp_{i(\sigma)} be the inverse image in 𝒪Kf\mathcal{O}_{K_{f}} of the maximal ideal of the integral closure p¯¯p\overline{\mathbb{Z}_{p}}\subset\overline{\mathbb{Q}}_{p} of p\mathbb{Z}_{p} under the composite of σ\sigma with the fixed ¯¯p\overline{\mathbb{Q}}_{\ell}\simeq\overline{\mathbb{Q}}_{p}:

𝒪KfKf\textstyle{\mathcal{O}_{K_{f}}\subset K_{f}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}σ\scriptstyle{\sigma}¯¯p𝔪p¯.\textstyle{\overline{\mathbb{Q}}_{\ell}\simeq\overline{\mathbb{Q}}_{p}\supset\mathfrak{m}_{\overline{\mathbb{Z}_{p}}}.}

Then S𝔭,σS_{\mathfrak{p},\sigma} is the Newton polygon of the polynomial (obtained by applying σ\sigma to (4.1.1.3)) with respect to vpv_{p} on \mathbb{Q}_{\ell}:

X2σ(a𝔭)X+σ(det(ρ(Frob𝔭)))X^{2}-\sigma(a_{\mathfrak{p}})X+\sigma(\det(\rho(\operatorname{Frob}_{\mathfrak{p}})))

and as such is equal to:

S𝔭,σ={{0,1} if ei(σ)=0{1/2,1/2} if ei(σ)>0S_{\mathfrak{p},\sigma}=\begin{cases}\{0,1\}&\mbox{ if }e_{i(\sigma)}=0\\ \{1/2,1/2\}&\mbox{ if }e_{i(\sigma)}>0\end{cases}

Let f1,,fmf_{1},\cdots,f_{m} be the degrees of the residue class extensions:

fi:=dim𝔽p𝒪Kf/i.f_{i}:=\dim_{\mathbb{F}_{p}}\mathcal{O}_{K_{f}}/\wp_{i}.

Then we have

NP(Frob𝔭|M(f))=P(d;kf,k(p)),\mathrm{NP}(\operatorname{Frob}_{\mathfrak{p}}|_{M(f)})=P(d;k_{f},k(p)),

where k(p)k(p) is the sum of those fif_{i} for which ei>0e_{i}>0.

Since a𝔭0a_{\mathfrak{p}}\neq 0, we may apply the product formula. By the Lemma, we have

v|a𝔭v(2p)kf,\prod_{v|\infty}\|a_{\mathfrak{p}}\|_{v}\leq(2\sqrt{p})^{k_{f}},

while by the factorisation (4.1.3.1):

v|pa𝔭v=pi=1meifi\prod_{v|p}\|a_{\mathfrak{p}}\|_{v}=p^{-\sum_{i=1}^{m}e_{i}f_{i}}

and

vp,a𝔭v=vp,Iv1.\prod_{v\nmid p,\infty}\|a_{\mathfrak{p}}\|_{v}=\prod_{v\nmid p,\infty}\|I\|_{v}\leq 1.

Therefore

1=va𝔭2kfpkf2ieifi,1=\prod_{v}\|a_{\mathfrak{p}}\|\leq 2^{k_{f}}p^{\frac{k_{f}}{2}-\sum_{i}e_{i}f_{i}},

which implies for all p>22kfp>2^{2k_{f}}

kf2i=1meifii:ei>0fi=k(p).\frac{k_{f}}{2}\geq\sum_{i=1}^{m}e_{i}f_{i}\geq\sum_{i:e_{i}>0}f_{i}=k(p).

This proves (2).

For (3), we assume in addition that pp splits completely in F~\widetilde{F^{\circ}}, and choose a prime 𝔭\mathfrak{p} of F~\widetilde{F^{\circ}} lying over pp, so that by definition

ρf,(Frob𝔭)Gf,() and a𝔭𝒪Kf.\rho_{f,\ell}(\operatorname{Frob}_{\mathfrak{p}})\in G_{f,\ell}^{\circ}(\mathbb{Q}_{\ell})\,\,\,\mbox{ and }\,\,\,a_{\mathfrak{p}}\in\mathcal{O}_{K_{f}^{\circ}}.

If kf=1k_{f}^{\circ}=1, that is, if Kf=K_{f}^{\circ}=\mathbb{Q}, then a𝔭a_{\mathfrak{p}}\in\mathbb{Z} and |a𝔭|<2p|a_{\mathfrak{p}}|_{\infty}<2\sqrt{p}. As soon as p5p\geq 5, the only way p|a𝔭p|a_{\mathfrak{p}} is then a𝔭=0a_{\mathfrak{p}}=0, which we have excluded above.

Suppose therefore that kf=2k_{f}^{\circ}=2, and consider the homomorphisms

Gal(/F~)\textstyle{\operatorname{Gal}(\mathbb{Q}/\widetilde{F^{\circ}})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ρ\scriptstyle{\rho^{\circ}}((ResKfGL2)det×)()\textstyle{\left((\operatorname{Res}^{K_{f}^{\circ}}_{\mathbb{Q}}GL_{2})^{\det\subseteq\mathbb{Q}^{\times}}\right)(\mathbb{Q}_{\ell})\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ι\scriptstyle{\iota}det\scriptstyle{\det}GL4()\textstyle{GL_{4}(\mathbb{Q}_{\ell})}×\textstyle{\mathbb{Q}_{\ell}^{\times}}

and the regular map of algebraic varieties

Tr(2(ι)det1):((ResKfGL2)det×)𝔸1,\mathrm{Tr}(\wedge^{2}(\iota)\otimes{\det}^{-1}):((\operatorname{Res}^{K_{f}^{\circ}}_{\mathbb{Q}}GL_{2})^{\det\subseteq\mathbb{Q}^{\times}})\otimes\mathbb{Q}_{\ell}\longrightarrow\mathbb{A}^{1}_{\mathbb{Q}_{\ell}}, (4.1.3.2)

where 2(ι)\wedge^{2}(\iota) takes values in GL6GL_{6} and det\det takes values in 𝐆m\mathbb{\mathbf{G}}_{m}.

In order to prove (3), we find a set of primes 𝔭\mathfrak{p} of F~\widetilde{F^{\circ}} of density =1=1 such that a𝔭a_{\mathfrak{p}} is not divisible by any prime of KfK_{f}^{\circ} lying over (p)(p). If pp is inert in KfK_{f}^{\circ}, the bound (2) suffices, and we exclude the finitely many primes that are ramified in KfK_{f}^{\circ}, so we assume that pp splits:

p𝒪Kf=12, where 12 are primes.p\cdot\mathcal{O}_{K_{f}^{\circ}}=\wp_{1}\wp_{2},\,\,\,\,\mbox{ where }\wp_{1}\neq\wp_{2}\mbox{ are primes.} (4.1.3.3)

By (2), we may assume that at most one of the 22 primes can divide a𝔭a_{\mathfrak{p}}, say a𝔭1a_{\mathfrak{p}}\in\wp_{1} but a𝔭2a_{\mathfrak{p}}\not\in\wp_{2}. Let ϵ\epsilon be the nontrivial field automorphism of KfK_{f}^{\circ}, and let α𝔭\alpha_{\mathfrak{p}} be a root of the polynomial X2a𝔭X+p=0X^{2}-a_{\mathfrak{p}}X+p=0 and let α𝔭\alpha^{\prime}_{\mathfrak{p}} be a root of X2ϵ(a𝔭)X+p=0.X^{2}-\epsilon(a_{\mathfrak{p}})X+p=0.

Then the 44 eigenvalues of (ιρ)(Frob𝔭)(\iota\circ{\rho^{\circ}})(\operatorname{Frob}_{\mathfrak{p}}) are {α𝔭,pα𝔭,α𝔭,pα𝔭}\left\{\alpha_{\mathfrak{p}},\frac{p}{\alpha_{\mathfrak{p}}},\alpha^{\prime}_{\mathfrak{p}},\frac{p}{\alpha_{\mathfrak{p}}}\right\}, and the 66 eigenvalues of (2ιρ)(Frob𝔭)(\wedge^{2}\iota\circ{\rho^{\circ}})(\operatorname{Frob}_{\mathfrak{p}}) are {p,α𝔭α𝔭,α𝔭pα𝔭,pα𝔭α𝔭,pα𝔭pα𝔭,p}.\left\{p,\,\alpha_{\mathfrak{p}}\alpha^{\prime}_{\mathfrak{p}},\,\alpha_{\mathfrak{p}}\frac{p}{\alpha^{\prime}_{\mathfrak{p}}},\,\frac{p}{\alpha_{\mathfrak{p}}}\alpha^{\prime}_{\mathfrak{p}},\,\frac{p}{\alpha_{\mathfrak{p}}}\frac{p}{\alpha^{\prime}_{\mathfrak{p}}},\,p\right\}. Therefore the value of (4.1.3.2) at Frob𝔭\operatorname{Frob}_{\mathfrak{p}} is

1p(2p+(α𝔭+pα𝔭)(α𝔭+pα𝔭))=2+a𝔭ϵ(a𝔭)p\frac{1}{p}\left(2p+(\alpha_{\mathfrak{p}}+\frac{p}{\alpha_{\mathfrak{p}}})(\alpha^{\prime}_{\mathfrak{p}}+\frac{p}{\alpha^{\prime}_{\mathfrak{p}}})\right)=2+\frac{a_{\mathfrak{p}}\epsilon(a_{\mathfrak{p}})}{p}\in\mathbb{Z}

Now by the Lemma 4.1.3, which implies

|a𝔭|,|ϵ(a𝔭)|2p|a_{\mathfrak{p}}|_{\infty},|\epsilon(a_{\mathfrak{p}})|_{\infty}\leq 2\sqrt{p}

and by our assumption that pp is unramified in KfK_{f}^{\circ}, which implies that the inequalities are strict, we have

Tr(2(ι)det1)(Frob𝔭)[1,5].\mathrm{Tr}(\wedge^{2}(\iota)\otimes{\det}^{-1})(\operatorname{Frob}_{\mathfrak{p}})\in[-1,5]\cap\mathbb{Z}. (4.1.3.4)

Therefore a𝔭a_{\mathfrak{p}} does not belong to any prime of KfK_{f}^{\circ} lying over (p)(p), as soon as we avoid (4.1.3.4). But since

Tr(2(ι)det1)(1)=6\mathrm{Tr}(\wedge^{2}(\iota)\otimes{\det}^{-1})(1)=6

(here 11 denotes the identity element in the group Gf,G^{\circ}_{f,\ell}), the set of 𝔭\mathfrak{p} for which (4.1.3.4) holds has density 0 by Lemma 3.6.1. Now if 𝔭\mathfrak{p} avoids (4.1.3.4), then

a𝔭𝒪Kf=(a𝔭𝒪Kf)𝒪Kfa_{\mathfrak{p}}\mathcal{O}_{K_{f}}=(a_{\mathfrak{p}}\mathcal{O}_{K_{f}^{\circ}})\mathcal{O}_{K_{f}}

is coprime to (p)(p), and we have k(p)=0k(p)=0. This completes the proof of (3).

For the sake of continuity in exposition, we treat the conditional (5) before the unconditional (4). By an argument similar to that for (2), but applied to the restriction ρ:Gal(¯/F~)GL2(Kf,λ)\rho^{\circ}:\operatorname{Gal}(\overline{\mathbb{Q}}/\widetilde{F^{\circ}})\longrightarrow GL_{2}(K^{\circ}_{f,\lambda^{\circ}}) (where λ=λKf\lambda^{\circ}=\lambda\cap K^{\circ}_{f}), for a prime 𝔭\mathfrak{p} of density 11 in F~\widetilde{F^{\circ}}, if we write

a𝔭𝒪Kf=1e1memIa_{\mathfrak{p}}\mathcal{O}_{K^{\circ}_{f}}=\wp_{1}^{e_{1}}\cdots\wp_{m^{\prime}}^{e_{m^{\prime}}}I

where 1,,m\wp_{1},\cdots,\wp_{m^{\prime}} are the primes of KfK^{\circ}_{f} lying over pp and II is coprime to pp, then we have

i:ei>0dim𝔽p(𝒪Kf/i)kf2,\sum_{i:e_{i}>0}\dim_{\mathbb{F}_{p}}(\mathcal{O}_{K^{\circ}_{f}}/\wp_{i})\leq\frac{k_{f}^{\circ}}{2}, (4.1.3.5)

and

k(p)=kfkfi:ei>0dim𝔽p(𝒪Kf/i).k(p)=\frac{k_{f}}{k_{f}^{\circ}}\sum_{i:e_{i}>0}\dim_{\mathbb{F}_{p}}(\mathcal{O}_{K^{\circ}_{f}}/\wp_{i}).

If kfk_{f}^{\circ} is odd, then (4.1.3.5) trivially implies (5), so we assume that kfk_{f}^{\circ} is even.

Now if the equality holds in (4.1.3.5), then we necessarily have ei=1e_{i}=1 whenever ei>0e_{i}>0, and by Lemma 4.1.3 and the product formula

(2p)kfv|a𝔭v=(v|pa𝔭vvpa𝔭v)1pkf/2.(2\sqrt{p})^{k_{f}^{\circ}}\geq\prod_{v|\infty}\|a_{\mathfrak{p}}\|_{v}=\left(\prod_{v|p}\|a_{\mathfrak{p}}\|_{v}\cdot\prod_{v\nmid p\infty}\|a_{\mathfrak{p}}\|_{v}\right)^{-1}\in p^{k_{f}^{\circ}/2}\mathbb{Z}.

In other words, if i1,,ikfi_{1},\cdots,i_{k_{f}^{\circ}} are the real embeddings of KfK_{f}^{\circ}, then we have

j=1kfij(a𝔭)p[2kf,2kf].\prod_{j=1}^{k_{f}^{\circ}}\frac{i_{j}(a_{\mathfrak{p}})}{\sqrt{p}}\in\mathbb{Z}\cap[-2^{k_{f}^{\circ}},2^{k_{f}^{\circ}}].

In kf\mathbb{R}^{k_{f}^{\circ}}, consider the nowhere dense real analytic subsets

Bj={(x1,,xkf):a=1kfxa=j}B_{j}=\left\{(x_{1},\cdots,x_{k_{f}^{\circ}}):\prod_{a=1}^{k_{f}^{\circ}}x_{a}=j\right\}

for nonzero j[2kf,2kf]j\in\mathbb{Z}\cap[-2^{k_{f}^{\circ}},2^{k_{f}^{\circ}}], B0={(0,,0)}B_{0}=\{(0,\cdots,0)\}, and let BB be their union.

If ff satisfies (RST), then the set A(f)A(f) in §3.3 is equidistributed in φ(𝐱)dμL(𝐱)\varphi(\mathbf{x})d\mu_{L}(\mathbf{x}), where φ:[2,2]kf0\varphi:[-2,2]^{k_{f}^{\circ}}\longrightarrow\mathbb{R}_{\geq 0} is a continuous function and dμLd\mu_{L} is the Lebesgue measure. Therefore the set of primes 𝔭\mathfrak{p} of F~\widetilde{F^{\circ}} such that

(i1(a𝔭)𝔭,,ikf(a𝔭)𝔭)B\left(\frac{i_{1}(a_{\mathfrak{p}})}{\sqrt{\mathbb{N}\mathfrak{p}}},\cdots,\frac{i_{k_{f}^{\circ}}(a_{\mathfrak{p}})}{\sqrt{\mathbb{N}\mathfrak{p}}}\right)\in B

has density 0, since Bφ𝑑μL=0\int_{B}\varphi d\mu_{L}=0.

If ff satisfies (tt-ST’), then there exists a sequence 1j1<<jtkf1\leq j_{1}<\cdots<j_{t}\leq k_{f}^{\circ} such that prj1,,jt(A(f))\mathrm{pr}_{j_{1},\cdots,j_{t}}(A(f)) is equidistributed in the tt-fold product of the Sato-Tate measure. Then the set

C={(x1,,xkf):a=1t|xja|<2(kft)}C=\left\{(x_{1},\cdots,x_{k_{f}^{\circ}}):\prod_{a=1}^{t}|x_{j_{a}}|<2^{-(k_{f}^{\circ}-t)}\right\}

is disjoint from BjB_{j} for all nonzero j[2kf,2kf]j\in\mathbb{Z}\cap[-2^{k_{f}^{\circ}},2^{k_{f}^{\circ}}]. Since the set of 𝔭\mathfrak{p} such that a𝔭=0a_{\mathfrak{p}}=0 has density 0 as noted above, the lower natural density of the set of primes 𝔭\mathfrak{p} such that (i1(a𝔭)𝔭,,ikf(a𝔭)𝔭)B\displaystyle{\left(\frac{i_{1}(a_{\mathfrak{p}})}{\sqrt{\mathbb{N}\mathfrak{p}}},\cdots,\frac{i_{k_{f}^{\circ}}(a_{\mathfrak{p}})}{\sqrt{\mathbb{N}\mathfrak{p}}}\right)}\not\in B is bounded from below by

c(kf,t):=|y1||yt|<2tkf𝑑μST(y1)𝑑μST(yt)>0 for 1t<kfc(k_{f}^{\circ},t):=\int_{|y_{1}|\cdots|y_{t}|<2^{t-k_{f}^{\circ}}}d\mu_{ST}(y_{1})\cdots d\mu_{ST}(y_{t})>0\,\,\,\,\mbox{ for }1\leq t<k_{f}^{\circ} (4.1.3.6)

where dμST(y)=12π4y2dyd\mu_{ST}(y)=\frac{1}{2\pi}\sqrt{4-y^{2}}dy concentrated in [2,2][-2,2].

(4) There is nothing to prove if σF~(Kf)1/2\sigma_{\widetilde{F}}(K_{f})\geq 1/2, so we assume that

λ0:=λF~(Kf)>kf/2.\lambda_{0}:=\lambda_{\widetilde{F}}(K_{f})>k_{f}/2.

We return to the primes 𝔭\mathfrak{p} of F~\widetilde{F} of absolute degree 11 over \mathbb{Q}. The conjugacy class of Frob𝔭\operatorname{Frob}_{\mathfrak{p}} in Gal(¯S/F~)\operatorname{Gal}(\overline{\mathbb{Q}}_{S}/\widetilde{F}) maps into the conjugacy class of Frobp\operatorname{Frob}_{p} in Gal(¯S/)\operatorname{Gal}(\overline{\mathbb{Q}}_{S}/\mathbb{Q}). (Here ¯S\overline{\mathbb{Q}}_{S} is the maximal subfield of ¯\overline{\mathbb{Q}} unramified outside disc(F)disc(Kf)(𝔫)\mathrm{disc}(F)\cdot\mathrm{disc}(K_{f})\cdot\ell\cdot\mathbb{N}(\mathfrak{n}).) The following diagram exhibits the interaction of F~\widetilde{F} and KfK_{f}3.5) and the Galois representation at hand:

Frob𝔭\textstyle{\operatorname{Frob}_{\mathfrak{p}}^{\sharp}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Gal(¯S/F~)\textstyle{\operatorname{Gal}(\overline{\mathbb{Q}}_{S}/\widetilde{F})\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ρf,λ\scriptstyle{\rho_{f,\lambda}}GL2(Kf,λ)\textstyle{GL_{2}(K_{f,\lambda})}Frobp\textstyle{\operatorname{Frob}_{p}^{\sharp}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Gal(¯S/)\textstyle{\operatorname{Gal}(\overline{\mathbb{Q}}_{S}/\mathbb{Q})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Aut(Hom(Kf,¯))H\textstyle{\operatorname{Aut}(\operatorname{Hom}(K_{f},\overline{\mathbb{Q}}))\supseteq H^{\sharp}}

Let Γ\Gamma be the image of Gal(¯S/F~)\operatorname{Gal}(\overline{\mathbb{Q}}_{S}/\widetilde{F}) in Aut(Hom(Kf,¯))\operatorname{Aut}(\operatorname{Hom}(K_{f},\overline{\mathbb{Q}})), and let HΓH^{\sharp}\subseteq\Gamma be the nonempty subset of elements hh such that λ(h,Hom(Kf,¯))=λ0\lambda(h,\operatorname{Hom}(K_{f},\overline{\mathbb{Q}}))=\lambda_{0}; one sees easily that HH^{\sharp} is stable under conjugation by Γ\Gamma. Then if Frob𝔭\operatorname{Frob}_{\mathfrak{p}} maps into HH^{\sharp}, there exists a prime ideal \wp of KfK_{f} lying over (p)(p) with

dim𝔽p𝒪Kf/=λ0.\dim_{\mathbb{F}_{p}}\mathcal{O}_{K_{f}}/\wp=\lambda_{0}.

Now the bound in (2) prevents \wp from occuring in the ideal factorisation of a𝔭a_{\mathfrak{p}} (4.1.3.1) with multiplicity >0>0. Therefore k(p)k(p) is at most the sum of the degrees of the residue class field extensions at the other primes of KfK_{f} lying over (p)(p), and

k(p)kfλ0=kfσF~(Kf).k(p)\leq k_{f}-\lambda_{0}=k_{f}\cdot\sigma_{\widetilde{F}}(K_{f}).

The density of such primes 𝔭\mathfrak{p} (i.e. not dividing disc(F)disc(Kf)(𝔫)\mathrm{disc}(F)\cdot\mathrm{disc}(K_{f})\cdot\ell\cdot\mathbb{N}(\mathfrak{n}), having a𝔭0a_{\mathfrak{p}}\neq 0, and whose Frob𝔭\operatorname{Frob}_{\mathfrak{p}} mapping into HH^{\sharp}) in F~\widetilde{F} is, by the Chebotarev density theorem, equal to |H|/|Γ|>0|H^{\sharp}|/|\Gamma|>0, and we get (4).

The proof of (4bis) is parallel to that of (4), except we consider the degree-11 primes of F~\widetilde{F^{\circ}} and a𝔭𝒪Kfa_{\mathfrak{p}}\in\mathcal{O}_{K_{f}^{\circ}}, and we omit it.

(6) Consider a similar diagram (with F~\widetilde{F} replaced with F~\widetilde{F^{\circ}}) to the one in (4), and let Γ\Gamma^{\circ} be the image of Gal(¯S/F~)\operatorname{Gal}(\overline{\mathbb{Q}}_{S}/\widetilde{F^{\circ}}) in Aut(Hom(Kf,¯))\operatorname{Aut}(\operatorname{Hom}(K_{f}^{\circ},\overline{\mathbb{Q}})). This time, choose HH^{\sharp} to consist of those elements of Γ\Gamma^{\circ} that bisect Hom(Kf,¯)\operatorname{Hom}(K_{f}^{\circ},\overline{\mathbb{Q}}). Again, HH^{\sharp} is stable under conjugation by Γ\Gamma^{\circ}.

Then for any prime 𝔭\mathfrak{p} of F~\widetilde{F^{\circ}} such that Frob𝔭\operatorname{Frob}_{\mathfrak{p}} maps into HH^{\sharp} and pp is unramified in KfK_{f}^{\circ}, there are exactly 22 primes of KfK_{f}^{\circ} lying over (p)(p) with the same degree of residue class extension (namely =kf/2=k_{f}^{\circ}/2). The density of such primes 𝔭\mathfrak{p} in F~\widetilde{F^{\circ}} is |H|/|Γ|>0|H^{\sharp}|/|\Gamma^{\circ}|>0 by the Chebotarev density theorem.

Now the bound in (5), which is in effect because we assume (RST), keeps either of the 22 primes from appearing in the ideal decomposition of a𝔭𝒪Kfa_{\mathfrak{p}}\mathcal{O}_{K_{f}^{\circ}} with multiplicity >0>0, except for a set of primes 𝔭\mathfrak{p} with density 0. ∎

Remark 4.1.4.

The constant c(k,t)c(k,t) for t=1t=1 (where k=kfk=k^{\circ}_{f}) can be expressed:

c(k,1)=2π(12k1122k+arcsin(12k)) for k2c(k,1)=\frac{2}{\pi}\left(\frac{1}{2^{k}}\sqrt{1-\frac{1}{2^{2k}}}+\arcsin(\frac{1}{2^{k}})\right)\,\,\,\,\mbox{ for }k\geq 2

and is asymptotically 1/(π2k2)1/(\pi 2^{k-2}) as kk\rightarrow\infty. Here are approximate values of c(k,t)c(k,t) for 1tk61\leq t\leq k\leq 6:

kk || tt 1 2 3 4 5 6
1 1
2 0.315 1
3 0.159 0.501 1
4 0.0795 0.320 0.62 1
5 0.0398 0.195 0.45 0.71 1
6 0.0199 0.115 0.31 0.56 0.8 1

4.2 CM case

Theorem 4.2.1.

Let ff be a new normalised Hilbert eigencuspform of level 𝔫𝒪F\mathfrak{n}\subseteq\mathcal{O}_{F} and parallel weight (2,,2)(2,\cdots,2). Suppose that ff is of CM type (§3.1).

Denote by M(f)M(f) the André motive, whose realisations give the part of the intersection cohomology of the Hilbert modular variety corresponding to {σ(f)}\{\sigma(f)\}, where σ\sigma ranges over all the embeddings of KfK_{f} into ¯\overline{\mathbb{Q}}. Then:

  1. (1)

    For all rational primes pp coprime to disc(F)𝔫\mathrm{disc}(F)\cdot\mathfrak{n}\cdot\ell, we have

    NP(Frobp|M(f))HTP(M(f)).\mathrm{NP}(\operatorname{Frob}_{p}|_{M(f)})\geq\mathrm{HTP}(M(f)).
  2. (2)

    For a principally abundant set of primes pp, we have

    NP(Frobp|M(f))=HTP(M(f)).\mathrm{NP}(\operatorname{Frob}_{p}|_{M(f)})=\mathrm{HTP}(M(f)).
Proof.

Let λ\lambda be a prime of KfK_{f} lying over ()(\ell) such that the connected component GG^{\circ} of the Zariski closure G=Gf,λG=G_{f,\lambda} of the image of ρf,λ\rho_{f,\lambda} is a torus. The argument employed in proving part (1) of Theorem 4.1.1 goes through without change: The non CM condition was not used. This way we get the inequality (1) for primes pp coprime to disc(F)𝔫\mathrm{disc}(F)\cdot\mathfrak{n}\cdot\ell, and for those splitting completely in FF and unramified in KfK_{f} in addition, an integer k(p)[0,kf]k(p)\in[0,k_{f}] such that

NP(Frobp|M(f))=P(d;kf,k(p)).\mathrm{NP}(\operatorname{Frob}_{p}|_{M(f)})=P(d;k_{f},k(p)).

Let FF^{\circ} be the Galois extension of FF cut out by the two representations with finite image:

Gal(¯/F)=ker(Gal(¯/F)G(Kf,λ)/G(Kf,λ))ker(det(ρf,λ)(1))\operatorname{Gal}(\overline{\mathbb{Q}}/F^{\circ})=\ker(\operatorname{Gal}(\overline{\mathbb{Q}}/F)\longrightarrow G(K_{f,\lambda})/G^{\circ}(K_{f,\lambda}))\cap\ker(\det(\rho_{f,\lambda})(1))

and let F~\widetilde{F^{\circ}} be the compositum of FF^{\circ} and F~\widetilde{F}.

Since the restriction of ρf,λ\rho_{f,\lambda} to F~\widetilde{F^{\circ}} is then abelian and KfK_{f}-rational, by a theorem of Serre [Se98, Chpt. III, §3], augmented with a transcendence result of Waldschmidt [Ws81] (see Henniart [He82]), this restriction is locally algebraic (and semisimple by assumption). Then by a theorem of Ribet [Ri76, §1.6] (which extends that of Serre [Se98, §III.2.3]), there exist (i) a 22-dimensional KfK_{f}-rational vector subspace V0V_{0} of Kf,λ2K_{f,\lambda}^{\oplus 2}, (ii) a modulus 𝔪\mathfrak{m} of F~\widetilde{F^{\circ}}, and (iii) a rational representation

ϕ0:S𝔪KfGLV0\phi_{0}:S_{\mathfrak{m}}\otimes_{\mathbb{Q}}K_{f}\longrightarrow GL_{V_{0}}

such that ρ|F~\rho|_{\widetilde{F^{\circ}}} is isomorphic to the λ\lambda-adic representation associated with ϕ0\phi_{0}.

The image of ϕ0\phi_{0} is a maximal algebraic torus of GLV0GL_{V_{0}}, since detϕ0\det\phi_{0} gives the Tate structure (1)\mathbb{Q}_{\ell}(-1) on Gal(¯/F)\operatorname{Gal}(\overline{\mathbb{Q}}/F^{\circ}), and the cyclotomic character is not divisible by 22 as the character of any number field.

Let KK^{\prime} be the splitting field over KfK_{f} of this algebraic torus, so that [K:Kf]2[K^{\prime}:K_{f}]\leq 2 and [Gf,λ:Gf,λ]2[G_{f,\lambda}:G_{f,\lambda}^{\circ}]\leq 2. (It is worth clarifying that unlike the KK^{\prime} introduced in the proof of Theorem 4.1.1 in a similar context, this KK^{\prime} depends only on ρf,λ\rho_{f,\lambda} and is independent of 𝔭\mathfrak{p}.)

For every prime 𝔭\mathfrak{p} of F~\widetilde{F^{\circ}} of absolute degree 11 and coprime to disc(F)𝔫\mathrm{disc}(F)\cdot\mathfrak{n}\cdot\ell, let {α𝔭,β𝔭}K\{\alpha_{\mathfrak{p}},\beta_{\mathfrak{p}}\}\subset K^{\prime} be the 22 eigenvalues of ρf,λ(Frob𝔭)\rho_{f,\lambda}(\operatorname{Frob}_{\mathfrak{p}}). Since they are Weil pp-integers by Lemma 4.1.3 and det(ρ(Frob𝔭))=p\det(\rho(\operatorname{Frob}_{\mathfrak{p}}))=p, we have

β𝔭=pα𝔭=α𝔭¯\beta_{\mathfrak{p}}=\frac{p}{\alpha_{\mathfrak{p}}}=\overline{\alpha_{\mathfrak{p}}} (4.2.1.1)

where the bar denotes the complex conjugation on (α𝔭)=(β𝔭)\mathbb{Q}(\alpha_{\mathfrak{p}})=\mathbb{Q}(\beta_{\mathfrak{p}}).

Now consider only those 𝔭\mathfrak{p} such that, in addition, (p):=𝔭(p):=\mathfrak{p}\cap\mathbb{Z} splits completely in KK^{\prime}, a fortiori also in (α𝔭)\mathbb{Q}(\alpha_{\mathfrak{p}}); the resulting set is clearly principally abundant. Then, since (p)(p) is unramified in (α𝔭)\mathbb{Q}(\alpha_{\mathfrak{p}}), α𝔭\alpha_{\mathfrak{p}} cannot be totally real, and generates a CM field. Let {1,,m,1¯,,m¯}\{\wp_{1},\cdots,\wp_{m},\overline{\wp_{1}},\cdots,\overline{\wp_{m}}\} be the set of primes of (α𝔭)\mathbb{Q}(\alpha_{\mathfrak{p}}) lying over (p)(p), where 2m=[(α𝔭):]2m=[\mathbb{Q}(\alpha_{\mathfrak{p}}):\mathbb{Q}]. The equation (4.2.1.1) further shows that, perhaps after renaming the primes, we get

α𝔭𝒪(α𝔭)=1m and β𝔭𝒪(α𝔭)=1¯m¯.\alpha_{\mathfrak{p}}\cdot\mathcal{O}_{\mathbb{Q}(\alpha_{\mathfrak{p}})}=\wp_{1}\cdots\wp_{m}\,\,\mbox{ and }\,\,\beta_{\mathfrak{p}}\cdot\mathcal{O}_{\mathbb{Q}(\alpha_{\mathfrak{p}})}=\overline{\wp_{1}}\cdots\overline{\wp_{m}}.

It follows that

Tr(ρf,λ(Frob𝔭))=α𝔭+β𝔭\mathrm{Tr}(\rho_{f,\lambda}(\operatorname{Frob}_{\mathfrak{p}}))=\alpha_{\mathfrak{p}}+\beta_{\mathfrak{p}}

does not belong to any prime ideal of 𝒪K\mathcal{O}_{K^{\prime}} lying over (p)(p). Since Tr(ρf,λ(Frob𝔭))𝒪Kf\mathrm{Tr}(\rho_{f,\lambda}(\operatorname{Frob}_{\mathfrak{p}}))\in\mathcal{O}_{K_{f}}, it belongs to no prime of 𝒪Kf\mathcal{O}_{K_{f}} lying over (p)(p), either. This proves that k(p)=0k(p)=0 and completes the proof of Theorem 4.2.1. ∎

5 Examples

For the dimensions of and the Hecke orbits in the spaces of newforms, we rely on the information published in “The LL-functions and modular forms database” http://www.lmfdb.org/. We compute the slope σ\sigma by using the polynomials given in LMFDB generating KfK_{f}; sometimes the Galois group of Kf~/\widetilde{K_{f}}/\mathbb{Q} and the discriminant of KfK_{f} are also given in LMFDB, in which case we utilise the information also.

All the computations are for Γ0(𝔫)\Gamma_{0}(\mathfrak{n}) (trivial Nebentypus). Recall that we say two normalised eigencuspforms ff and gg with complex coefficients are conjugate (and that they belong to the same conjugacy class) if there is σAut()\sigma\in\operatorname{Aut}(\mathbb{C}) such that fσ=gf^{\sigma}=g.

5.1 F=F=\mathbb{Q}

The number of new normalised eigencuspforms ff of weight 22 and level N300N\leq 300 is:

N=1300dimS2new(Γ0(N),)=2074.\sum_{N=1}^{300}\dim_{\mathbb{C}}S_{2}^{\mathrm{new}}(\Gamma_{0}(N),\mathbb{C})=2074.

The degree of the field KfK_{f} in this range takes the following values:

kf=[Kf:]{1,2,3,4,5,6,7,8,9,10,11,12,14,16,17}k_{f}=[K_{f}:\mathbb{Q}]\in\{1,2,3,4,5,6,7,8,9,10,11,12,14,16,17\}

For 20702070 of the 20742074 forms ff, parts (3) and (4) of Theorem 4.1.1 and Theorem 4.2.1 show that M(f)M(f) has an abundant set of ordinary primes. The 44 exceptions:

  • There is 11 conjugacy class of 44 forms of level 275275 without CM, under the name 275.2.1.h in LMFDB, such that

    Kf=(3,11),K_{f}=\mathbb{Q}(\sqrt{3},\sqrt{11}),

    which is Galois with the Klein 44-group. There is a bisecting element, and σ(Kf)=1/2\sigma_{\mathbb{Q}}(K_{f})=1/2.

    If KfKfK_{f}^{\circ}\neq K_{f}, then part (2) of Theorem 4.1.1 provides a principally abundant set of ordinary primes. In case Kf=KfK_{f}^{\circ}=K_{f}, part (5) and the univariate (i.e. t=1t=1) Sato-Tate equidistribution (proven in [HSBT10] and [BLGHT11]) gives an abundant set of primes pp (of lower density 0.0794\geq 0.0794) such that k(p){0,1}k(p)\in\{0,1\}; if in addition ff satisfies (RST), then part (6) will imply the abundance of ordinary (k(p)=0k(p)=0) primes.

5.2 F=(2)F=\mathbb{Q}(\sqrt{2})

This quadratic field has discriminant 88 and class number 11.

LMFDB lists 10471047 new normalised eigencuspforms ff of parallel weight (2,2)(2,2) of level 𝔫\mathfrak{n} with (𝔫)350\mathbb{N}(\mathfrak{n})\leq 350, and the degree of KfK_{f} takes the following values:

kf{1,2,3,4,5,6,7,8,9,11,13}k_{f}\in\{1,2,3,4,5,6,7,8,9,11,13\}

For 10311031 of the 10471047, parts (3) and (4) of Theorem 4.1.1 and Theorem 4.2.1 show that M(f)M(f) has an abundant set of ordinary primes. The 1616 exceptions:

  1. (a)

    For the 88 forms ff in the classes 161.2-c and 161.3-c, we have Kf=(3,11)K_{f}=\mathbb{Q}(\sqrt{3},\sqrt{11}), which is Galois with the Klein 44-group. As KfK_{f} is linearly disjoint from F~=F=(2)\widetilde{F}=F=\mathbb{Q}(\sqrt{2}), by Proposition 3.5.6 (3), σF~(Kf)=σ(Kf)=1/2\sigma_{\widetilde{F}}(K_{f})=\sigma_{\mathbb{Q}}(K_{f})=1/2 and there is an element of Gal(¯/F)\operatorname{Gal}(\overline{\mathbb{Q}}/F) bisecting Hom(Kf,¯)\operatorname{Hom}(K_{f},\overline{\mathbb{Q}}).

  2. (b)

    For the 88 forms ff in 329.2-c and 329.3-c, [Kf:]=4[K_{f}:\mathbb{Q}]=4, Kf~/\widetilde{K_{f}}/\mathbb{Q} has group D8D_{8}, the image of Gal(¯/F)\operatorname{Gal}(\overline{\mathbb{Q}}/F) is a Klein subgroup, and σF(Kf)=1/2\sigma_{F}(K_{f})=1/2.

In both cases: If KfKfK_{f}^{\circ}\neq K_{f}, then part (3) of Theorem 4.1.1 gives a principally abundant set of ordinary primes. In case Kf=KfK_{f}^{\circ}=K_{f}, part (5) gives (unconditionally) an abundant set of primes pp such that k(p){0,1}k(p)\in\{0,1\}; if ff satisfies (RST) in addition, then part (6) will imply the abundance of ordinary primes.

5.3 F=(cos(2π/7))F=\mathbb{Q}(\cos(2\pi/7))

This is the largest totally real subfield of the cyclotomic field (e2πi/7)\mathbb{Q}(e^{2\pi i/7}). It is Galois over \mathbb{Q} with group /3\mathbb{Z}/3 and has discriminant 4949 and class number 11.

LMFDB lists 10751075 new normalised eigencuspforms ff of parallel weight (2,2,2)(2,2,2) and level 𝔫\mathfrak{n} with (𝔫)800\mathbb{N}(\mathfrak{n})\leq 800, and the degree of KfK_{f} takes the following values:

kf{1,2,3,4,5,6,7,8}k_{f}\in\{1,2,3,4,5,6,7,8\}

For 10481048 of the 10751075, parts (3) and (4) of Theorem 4.1.1 and Theorem 4.2.1 show that M(f)M(f) has an abundant set of ordinary primes. The 2727 exceptions:

  • For ff in the classes

    448.1-a, 547-1-c, 547.2-c, 547.3-c, 729.1-c, 729.1-d, 743.1-a, 743.2-a and 743.3-a

    we have F=KfF=K_{f} and σF(Kf)=2/3\sigma_{F}(K_{f})=2/3.

    For each of these ff: If KfKfK_{f}^{\circ}\neq K_{f}, in which case Kf=K_{f}^{\circ}=\mathbb{Q}, then part (3) of Theorem 4.1.1 would give a principally abundant set of ordinary primes. If Kf=KfK_{f}^{\circ}=K_{f}, then the theorem only provides a principally abundant set of primes pp such that k(p){0,1}k(p)\in\{0,1\}.

5.4 F=(cos(π/8))F=\mathbb{Q}(\cos(\pi/8))

This largest totally real subfield of (e2πi/16)\mathbb{Q}(e^{2\pi i/16}) has discriminant 2048=2112048=2^{11} and class number 11, and is Galois over \mathbb{Q} with group /4\mathbb{Z}/4. The nontrivial proper subgroup of /4\mathbb{Z}/4 allows a richer array of examples in which Theorems 4.1.1 and 4.2.1 fall short.

LMFDB lists 61856185 new normalised eigencuspforms ff of parallel weight (2,2,2,2)(2,2,2,2) and level 𝔫\mathfrak{n} with (𝔫)607\mathbb{N}(\mathfrak{n})\leq 607, and the degree of KfK_{f} takes the following values:

kf{1,2,,12}{14,15,16,17,18,19,20,22,24,25,26,27,28,30,33,39,42}.k_{f}\in\{1,2,\cdots,12\}\cup\{14,15,16,17,18,19,20,22,24,25,26,27,28,30,33,39,42\}.

For 60376037 of the 61856185, parts (3) and (4) of Theorem 4.1.1 and Theorem 4.2.1 show that M(f)M(f) has an abundant set of ordinary primes. The 136136 confirmed exceptions and 1212 possible exceptions:

  • (a)

    For the 2424 forms ff in the classes 392.1-f, 392.2-f, 544.1-\ell, 544.2-\ell, 544.3-\ell and 544.4-\ell, KfK_{f} is Galois over \mathbb{Q} with the Klein 44-group and linearly disjoint from FF over \mathbb{Q}. There is a bisecting element and σF(Kf)=σ(Kf)=1/2\sigma_{F}(K_{f})=\sigma_{\mathbb{Q}}(K_{f})=1/2. (cf. exceptions in F=F=\mathbb{Q} and (a) in F=(2)F=\mathbb{Q}(\sqrt{2}).)

  • (b)

    For the 2020 forms ff in the classes 81.1-c, 289.1-f, 289.4-f, 578.1-h and 578.4-h, kf=4k_{f}=4 and Kf~\widetilde{K_{f}} is Galois over \mathbb{Q} with group D8D_{8}, so σ(Kf)=0\sigma_{\mathbb{Q}}(K_{f})=0. However, the image of Gal(¯/F)\operatorname{Gal}(\overline{\mathbb{Q}}/F) is a Klein 44-group, σF(Kf)=1/2\sigma_{F}(K_{f})=1/2, and there is a bisecting element. (cf. exceptions (b) in F=(2)F=\mathbb{Q}(\sqrt{2}).)

  • (c)

    For the 2424 forms ff in 289.7-k, 289.8-k, 289.9-k and 289.10-k, KfK_{f} is Galois with group /6\mathbb{Z}/6\mathbb{Z}, and σ(Kf)=0\sigma_{\mathbb{Q}}(K_{f})=0. However, the image of Gal(¯/F)\operatorname{Gal}(\overline{\mathbb{Q}}/F) is the subgroup 2/62\mathbb{Z}/6\mathbb{Z}, σF(Kf)=1/2\sigma_{F}(K_{f})=1/2, and there is a bisecting element.

  • (d)

    For the 6868 forms ff in 1717 classes (8 in level norm 289, 1 in 324 and 8 in 578), we have Kf=FK_{f}=F. Thus σF(Kf)=3/4\sigma_{F}(K_{f})=3/4 and there is no bisecting element. (cf. exceptions in F=(cos(2π/7))F=\mathbb{Q}(\cos(2\pi/7)).)

  • (e)

    For the 1212 forms ff in 392.1-g and 392.2-g, KfK_{f} has degree 66 but is not cyclic over \mathbb{Q} (hence qualitatively different from (c)). So far we have observed: σ(Kf)=0\sigma_{\mathbb{Q}}(K_{f})=0, σF(Kf)1/2\sigma_{F}(K_{f})\leq 1/2, and there is a bisecting element.

For the ff in (d), part (5) of Theorem 4.1.1 provides an abundant set of primes pp such that k(p){0,1}k(p)\in\{0,1\} unconditionally.

In the remaining cases, we have k(p)kf/2k(p)\leq k_{f}/2 (resp. k(p)<kf/2k(p)<k_{f}/2) for a principally abundant (resp. abundant) set of primes pp by part (2) (resp. by part (5)), unconditionally. If ff satisfies (RST) in addition, then part (6) will provide an abundant set of ordinary (k(p)=0k(p)=0) primes.

6 General motivic coefficients

6.1 Conjectures in a general setting

Let XX be a projective variety of dimension dd over a number field FF, j:UXj:U\hookrightarrow X the inclusion of a smooth dense open subset, and π:𝒴U\pi:\mathcal{Y}\longrightarrow U a projective smooth scheme. For each integer ii and every prime number \ell, form the local system on UU,

i=Riπ()\mathcal{L}^{i}_{\ell}=R^{i}\pi_{\ast}(\mathbb{Q}_{\ell})

and the intermediate extension

¯i=j!(i[d])[d].\overline{\mathcal{L}}^{i}_{\ell}=j_{!\ast}(\mathcal{L}^{i}_{\ell}[d])[-d].
Conjecture 6.2.

Let the notation be as above, and let kk be any integer.

  1. (a)

    There exists a pure Grothendieck homological motive 𝔐=𝔐k,i\mathfrak{M}=\mathfrak{M}^{k,i} whose \ell-adic étale realisation 𝔐\mathfrak{M}_{\ell} is isomorphic to Hk(XFFs,¯i)H^{k}(X\otimes_{F}F^{s},\overline{\mathcal{L}}^{i}_{\ell}) for every \ell. 666In other words, there exists a projective smooth variety ZkZ_{k} over FF and an idempotent algebraic cycle (modulo homological equivalence) ϵk\epsilon_{k} on Zk×FZkZ_{k}\times_{F}Z_{k} such that ϵk𝔥(Zk)\epsilon_{k}\mathfrak{h}(Z_{k}) has \ell-adic realisation isomorphic to Hk(XFFs,¯i)H^{k}(X\otimes_{F}F^{s},\overline{\mathcal{L}}^{i}_{\ell}).

  2. (a’)

    There exists an André motive M=Mk,iM=M^{k,i} such that MHk(XFFs,¯i)M_{\ell}\simeq H^{k}(X\otimes_{F}F^{s},\overline{\mathcal{L}}^{i}_{\ell}) for every \ell.

For the following statements, we assume that (a’) is true.777It appears that (a’) follows from the construction of pure Nori motives realising the \ell-adic intersection cohomology groups, due to Ivorra and Morel [IM19, §6].

Let ee be an idempotent endomorphism of MM in the category of André motives (with \mathbb{Q}-coefficients) and let RR be the direct summand of MM cut out by ee, with the \ell-adic étale realisation RR_{\ell}.

  1. (b)

    The \ell-adic Galois representations RR_{\ell} form a strictly compatible system.

  2. (c)

    There exists a finite set S=S(π,i,k,e)S=S(\pi,i,k,e) of primes of FF such that, for every prime \ell and 𝔭\mathfrak{p} outside SS and not dividing \ell, we have

    NP(Frob𝔭,R)HTP(R).\mathrm{NP}(\operatorname{Frob}_{\mathfrak{p}},R_{\ell})\geq\mathrm{HTP}(R_{\ell}).
  3. (d)

    For infinitely many primes 𝔭\mathfrak{p} of FF and every prime number \ell, we have

    NP(Frob𝔭,R)=HTP(R).\mathrm{NP}(\operatorname{Frob}_{\mathfrak{p}},R_{\ell})=\mathrm{HTP}(R_{\ell}).

We note that, by Corollary 2.2.3, the Hodge-Tate polygon of RR_{\ell} at λ\lambda on the right hand sides is independent of the \ell-adic place λ\lambda of FF.

Proposition 6.2.1.

Assume that part (a) of Conjecture 6.2 is true and let MM be the André motive of 𝔐\mathfrak{M}, and that the idempotent ee is an algebraic cycle, and let \mathfrak{R} be the Grothendieck motive cut out by ee from 𝔐\mathfrak{M}. Then:

  1. (1)

    Parts (b) and (c) of the conjecture are also true for the André motive RR of \mathfrak{R}.

  2. (2)

    If, in addition, there exists a finite extension FF^{\prime} of FF such that RR_{\ell} restricts to an abelian Galois representation of FF^{\prime} for some (equivalently every) prime \ell, then for a principally abundant set of primes 𝔭\mathfrak{p} of FF, we have

    NP(Frob𝔭,R)=HTP(R),\mathrm{NP}(\operatorname{Frob}_{\mathfrak{p}},R_{\ell})=\mathrm{HTP}(R_{\ell}),

    and, in particular, part (d) of the conjecture is also true for RR.

Proof.

(1) The key point is that under the assumptions, we can use the crystalline realisation to compute the two polygons in part (c). Namely, for almost all 𝔭\mathfrak{p}, we have the free W(k(𝔭))W(k(\mathfrak{p}))-module cris,𝔭\mathfrak{R}_{\mathrm{cris},\mathfrak{p}}, equipped with the Hodge filtration and the crystalline Frobenius ϕcris,𝔭\phi_{\mathrm{cris},\mathfrak{p}} (induced from those on 𝔐\mathfrak{M}).

Then, on the one hand, by Katz and Messing [KM74, Th. 2], ϕcris,𝔭[k(𝔭):𝔽p]\phi_{\mathrm{cris},\mathfrak{p}}^{[k(\mathfrak{p}):\mathbb{F}_{p}]} has the same (multiset of) eigenvalues as the \ell-adic Frobenius Frob𝔭\operatorname{Frob}_{\mathfrak{p}} on {\mathfrak{R}}_{\ell}. Therefore they have the same Newton polygons. This also proves (b).

On the other hand, by Corollary 2.2.3, the Hodge-Tate polygon of R=R_{\ell}=\mathfrak{R}_{\ell} also coincides with the Hodge polygon of cris,𝔭\mathfrak{R}_{\mathrm{cris},\mathfrak{p}}, which by definition is equal to the Hodge polygon of the de Rham realisation dR\mathfrak{R}_{dR}.

Now the statement (c) follows from Mazur’s theorem [Ma73] applied to cris,𝔭\mathfrak{R}_{\mathrm{cris},\mathfrak{p}}. In summary:

NP(Frob𝔭,R)=NP(ϕcris,𝔭|cris,𝔭)HP(cris,𝔭)=HP(dR)=HTP(R).\mathrm{NP}(\operatorname{Frob}_{\mathfrak{p}},R_{\ell})=\mathrm{NP}(\phi_{\mathrm{cris},\mathfrak{p}}|_{\mathfrak{R}_{\mathrm{cris},\mathfrak{p}}})\geq HP(\mathfrak{R}_{\mathrm{cris},\mathfrak{p}})=HP(\mathfrak{R}_{dR})=\mathrm{HTP}(R_{\ell}).

For (2), let F~\widetilde{F^{\prime}} be the normal closure of FF^{\prime} over \mathbb{Q} and replace FF with F~\widetilde{F^{\prime}}, so that ρ\rho_{\ell} is abelian. Since ρ\rho_{\ell} is \mathbb{Q}-rational and Hodge-Tate, by a theorem of Serre [Se98, §III.2.3], it is associated with a \mathbb{Q}-rational representation ϕ0:S𝔪GLV0\phi_{0}:S_{\mathfrak{m}}\longrightarrow GL_{V_{0}}, where V0V_{0} is a \mathbb{Q}-form of RR_{\ell} and 𝔪\mathfrak{m} is a modulus of FF.

The restriction of ρ0\rho_{0} to T𝔪S𝔪T_{\mathfrak{m}}\subseteq S_{\mathfrak{m}} can then be diagonalised: ρ0|T𝔪¯=χ1χN\rho_{0}|_{T_{\mathfrak{m}}}\otimes\overline{\mathbb{Q}}=\chi_{1}\oplus\cdots\oplus\chi_{N} and

χi=[σ]nσ(i)[σ],\chi_{i}=\sum_{[\sigma]}n_{\sigma}(i)[\sigma],

where [σ][\sigma] ranges over the characters of T𝔪T_{\mathfrak{m}} arising from the embeddings σ\sigma of FF into ¯\overline{\mathbb{Q}}.

Since we already know part (b) of the conjecture, we can choose a rational prime \ell that splits completely in FF. Then we can identify the σ\sigma with the embeddings of FF into \mathbb{Q}_{\ell}, once an embedding ι:¯¯{\iota}_{\ell}:\overline{\mathbb{Q}}\longrightarrow\overline{\mathbb{Q}}_{\ell} has been fixed. Under this identification, the multiset of the Hodge-Tate weights of ρ\rho_{\ell} at any \ell-adic place λ:F¯\lambda:F\longrightarrow\overline{\mathbb{Q}}_{\ell} is {nσ0,λ(i)}i=1,,N\displaystyle{\left\{n_{\sigma_{0,\lambda}(i)}\right\}_{i=1,\cdots,N}}, where σ0,λ\sigma_{0,\lambda} is the (unique) embedding such that ισ0,λ=λ\iota_{\ell}\circ\sigma_{0,\lambda}=\lambda, by [Se98, Prop. 2, §III.1.1].

On the other hand, let 𝔭\mathfrak{p} be any prime of FF lying over any rational prime pp\neq\ell that splits completely in FF. We can also identify the embeddings σ:F¯\sigma:F\longrightarrow\overline{\mathbb{Q}} with the embeddings into p\mathbb{Q}_{p}, once we fix a pp-adic place of ιp:¯¯p\iota_{p}:\overline{\mathbb{Q}}\longrightarrow\overline{\mathbb{Q}}_{p}. Then the multiset of the pp-adic valuations of ρ(Frob𝔭)\rho_{\ell}(\operatorname{Frob}_{\mathfrak{p}}) is given by {nσ1,𝔭(i)}i=1,,N\displaystyle{\left\{n_{\sigma_{1,\mathfrak{p}}(i)}\right\}_{i=1,\cdots,N}}, where ιpσ1,𝔭\iota_{p}\circ\sigma_{1,\mathfrak{p}} is the pp-adic place 𝔭\mathfrak{p}: See [Se98, Cor. 2., §II.3.4].

Since ρ0\rho_{0} is \mathbb{Q}-rational, the two multisets are independent of λ|\lambda|\ell and 𝔭|p\mathfrak{p}|p, respectively, and are equal to each other. This proves (2). ∎

6.3 Hilbert modular forms of motivic weights

Let us specialise to the Baily-Borel compactification XX of the Hilbert modular variety UU defined over \mathbb{Q} (of some level 𝔫\mathfrak{n}) for the totally real field FF. For 𝒴\mathcal{Y}, we take the universal abelian scheme π:𝒜U\pi:\mathcal{A}\longrightarrow U and the fibred product 𝒜×U×U𝒜\mathcal{A}\times_{U}\cdots\times_{U}\mathcal{A} over UU.

Recall that a motivic weight k=(k(τ))τ:Fk=(k(\tau))_{\tau:F\longrightarrow\mathbb{R}} is a collection of integers k(τ)2k(\tau)\geq 2 of the same parity, for each real embedding τ\tau of FF.

Proposition 6.3.1.

Let ff be a new cusp form of any motivic weight k(2,,2)k\neq(2,\cdots,2). Then

  1. (1)

    The part M(f)M(f) of the intersection cohomology of XX cut out by all the conjugates of ff satisfies parts (a), (b) and (c) of Conjecture 6.2.

  2. (2)

    If, in addition, ff is of CM type, then M(f)M(f) satisfies part (d) of Conjecture 6.2 also.

Proof.

The first part follows immediately from Proposition 6.2.1 and the motivic construction of Galois representations (see Blasius-Rogawski [BR93] and the references therein). The second part follows from Proposition 6.2.1. ∎

In case 𝒴=𝒜\mathcal{Y}=\mathcal{A} is the universal abelian scheme, the cohomology decomposes into the parts cut out by ff of parallel weight (3,,3)(3,\cdots,3).

Definition 6.3.2.

Let GG be a group acting on a finite set XX. For gGg\in G, we define λ(g,X)\lambda^{\prime}(g,X) to be the smallest of the cardinalities of the gg-orbits in XX; we denote by λ(G,X)\lambda^{\prime}(G,X) the supremum of λ(g,X)\lambda^{\prime}(g,X) as gg ranges over GG.

Given two number fields FF and KK, we define

σF(K):=1λ(Gal(¯/F),Hom(K,¯))[K:][0,1]\sigma^{\prime}_{F}(K):=1-\frac{\lambda^{\prime}(\operatorname{Gal}(\overline{\mathbb{Q}}/F),\operatorname{Hom}(K,\overline{\mathbb{Q}}))}{[K:\mathbb{Q}]}\in\mathbb{Q}\cap[0,1]

The proof of the following is similar to that of Proposition 3.5.5, and is omitted:

Proposition 6.3.3.

If KK^{\prime} is a subfield of KK, then σF(K)σF(K)\sigma^{\prime}_{F}(K^{\prime})\leq\sigma^{\prime}_{F}(K).

Definition 6.3.4.

Let d1d\geq 1, k1k\geq 1 and i[0,k]i\in[0,k] be integers. We define the multiset (and the corresponding Newton polygon):

P(d;k,i):=({0,2}d)(ki)({1,1}d)iP^{\prime}(d;k,i):=\left(\{0,2\}^{\otimes d}\right)^{\oplus(k-i)}\oplus\left(\{1,1\}^{\otimes d}\right)^{\oplus i}

This is the polygon obtained by vertically stretching P(d;k,i)P(d;k,i) by a factor of 22.

Proposition 6.3.5.

Let ff be a new normalised Hilbert eigencuspform of level 𝔫𝒪F\mathfrak{n}\subseteq\mathcal{O}_{F} and parallel weight (3,,3)(3,\cdots,3). Assume that ff is not of CM type, and denote by F~\widetilde{F}, FF^{\circ}, F~\widetilde{F^{\circ}}, KfK_{f} and KfK_{f}^{\circ} the number fields defined in the manner of §3.1 and §3.2; λ\lambda is a prime of KfK_{f} lying over a rational prime \ell splitting completely in KfK_{f}.

Denote by M(f)M(f) the André motive (see Remark 2.2.6), whose realisations give the part of the intersection cohomology of the Hilbert modular variety corresponding to {σ(f)}σ\{\sigma(f)\}_{\sigma}, where σ\sigma ranges over all the embeddings of KfK_{f} into ¯\overline{\mathbb{Q}}.

  1. (1)

    For all rational primes pp that splits completely in FF (equivalently in F~\widetilde{F}) and pp is unramified in KfK_{f}, then there exists an integer k(p)[0,kf]k(p)\in[0,k_{f}] such that

    NP(Frobp|M(f))=P(d;kf,k(p)).\mathrm{NP}(\operatorname{Frob}_{p}|_{M(f)})=P^{\prime}(d;k_{f},k(p)).

    (Here kf=[Kf:]k_{f}=[K_{f}:\mathbb{Q}] and we refer to Definition 3.4.6 for the right hand side.)

    For the following parts, we only consider the primes splitting completely in FF and unramified in KfK_{f}.

  2. (2)

    For a principally abundant set of primes pp, we have k(p)kf(kf/kf)k(p)\leq k_{f}-(k_{f}/k_{f}^{\circ}).

  3. (3)

    For an abundant set of primes pp, we have k(p)kfmin(σF~(Kf),σF~(Kf))k(p)\leq k_{f}\cdot\mathrm{min}(\sigma^{\prime}_{\widetilde{F}}(K_{f}),\sigma^{\prime}_{\widetilde{F^{\circ}}}(K_{f}^{\circ})).

Proof.

(1) The proof is similar to that of part (1) of Theorem 4.1.1. The only difference is that the linear (resp. constant) coefficient of the polynomial (cf. (4.1.1.3))

X2Tr(ρ(Frob𝔭))X+det(ρ(Frob𝔭)),X^{2}-\mathrm{Tr}(\rho(\operatorname{Frob}_{\mathfrak{p}}))X+\det(\rho(\operatorname{Frob}_{\mathfrak{p}})),

has pp-adic valuation =2=2 (resp. an integer 0\geq 0 or \infty) for those pp considered.

(2) We find a set of primes 𝔭\mathfrak{p} of F~\widetilde{F^{\circ}} (resp. of F~\widetilde{F}) of density =1=1 such that a𝔭=Tr(ρλ(Frob𝔭))a_{\mathfrak{p}}=\mathrm{Tr}(\rho_{\lambda}(\operatorname{Frob}_{\mathfrak{p}})) is not divisible by at least one pp-adic prime \wp of F~\widetilde{F^{\circ}} (resp. of F~\widetilde{F}), where (p)=𝔭(p)=\mathfrak{p}\cap\mathbb{Z}.

If a𝔭a_{\mathfrak{p}} is divisible by all the pp-adic places, then it belongs to p𝒪Kfp\cdot\mathcal{O}_{K_{f}^{\circ}} (resp. to p𝒪Kfp\cdot\mathcal{O}_{K_{f}}. Since M(f)M(f) has pure motivic weight 22, for any archimedean place v|v|\infty, we have |a𝔭|v2p|a_{\mathfrak{p}}|_{v}\leq 2p, hence the algebraic integer |a𝔭/p|v2|a_{\mathfrak{p}}/p|_{v}\leq 2, and we form the finite set

S={α2:|α|v2 for all archimedean v of Kf (resp. Kf)}.S=\{\alpha^{2}:|\alpha|_{v}\leq 2\,\mbox{ for all archimedean }v\,\mbox{ of }\,K_{f}^{\circ}\mbox{ (resp. }K_{f}\mbox{)}\}.

By the assumption that ff is not of CM type, the connected algebraic monodromy group Gf,λG^{\circ}_{f,\lambda} is the full GL2GL_{2} over (Kf)λ)(K_{f})_{\lambda}), and the regular map of algebraic varieties

Tr(ρf,λ2det(ρf,λ)1):GL2𝔸1\mathrm{Tr}\left(\rho_{f,\lambda^{\circ}}^{\otimes 2}\otimes\det(\rho_{f,\lambda^{\circ}})^{-1}\right):GL_{2}\longrightarrow\mathbb{A}^{1}

(resp. with λ\lambda^{\circ} replaced with λ\lambda) is nonconstant, as ρf,λ\rho_{f,\lambda^{\circ}} has Hodge-Tate weights both 0 and 22 at all \ell-adic places. Therefore the inverse image of S(Kf)λS\subseteq(K^{\circ}_{f})_{\lambda} (resp. S(Kf)λS\subseteq(K_{f})_{\lambda}) has Haar measure 0. This proves the density =1=1 statement.

(3) With (2), we can now proceed with the Chebotarev type argument, similar to the one in parts (4) and (4bis) of Theorem 4.1.1. We omit the details. ∎

Acknowledgements.
The author has benefited from discussions with many mathematicians. He thanks particularly R. Taylor for his comments on the Sato-Tate equidistributions (known and conjectural) and his suggestion of using the uni- and bivariate distribution (which are known or accessible) in order to go beyond the Weil-Ramanujan-Petersson (“square-root”) bound. He also thanks J.-P. Serre for pointing out that our (SST) fits within his theoretical framework developed in [Se12]; M. Harris for his comments on the bivariate Sato-Tate conjecture; S. Morel and Y. André for discussions around André motives; L. Illusie and F. Orgogozo about constructibility theorems; R. Boltje for a discussion about tensor inductions; and N. Katz and B. Mazur for various suggestions. Finally, he thanks the anonymous referees for their suggestions for improving the exposition. Addendum in response to a message in October 2024 from N. Katz The references to the literature regarding ordinary reductions of abelian varieties in this article (written mostly in the academic year 2017/18) were incomplete. I missed the following work : W. Sawin, Ordinary primes for Abelian surfaces, Comptes Rendus Mathémathique, Volume 354, Issue 6, June 2016, pages 566–568, which would have been listed in the paragraph marked [\ast\ast] on the first page. I thank N. Katz for pointing this out, and make this necessary addendum.

References

  • [A96] Y. André. Pour une théorie inconditionnelle des motifs. Inst. Hautes Études Sci. Publ. Math., (83):5–49, 1996.
  • [AMRT] A. Mumford D. Rapoport M. Ash and Y. Tai. Smooth compactification of locally symmetric varieties, volume IV of Lie Groups: History, Frontiers and Applications. Math. Sci. Press, Brookline, Mass., first edition edition, 1975.
  • [BLGG11] T. Barnet-Lamb, T. Gee, and D. Geraghty. The Sato-Tate conjecture for Hilbert modular forms. J. Amer. Math. Soc., 24(2):411–469, 2011.
  • [BLGHT11] T. Barnet-Lamb, D. Geraghty, M. Harris, and R. Taylor. A family of Calabi-Yau varieties and potential automorphy II. Publ. Res. Inst. Math. Sci., 47(1):29–98, 2011.
  • [BBD] A. A. Beĭlinson, J. Bernstein, and P. Deligne. Faisceaux pervers. In Analysis and topology on singular spaces, I (Luminy, 1981), volume 100 of Astérisque, pages 5–171. Soc. Math. France, Paris, 1982.
  • [BR93] D. Blasius and J. Rogawski. Motives for Hilbert modular forms. Invent. Math., 114(1):55–87, 1993.
  • [B06] Don Blasius. Hilbert modular forms and the Ramanujan conjecture. In Noncommutative geometry and number theory, Aspects Math., E37, pages 35–56. Friedr. Vieweg, Wiesbaden, 2006.
  • [BL84] J.-L. Brylinski and J.-P. Labesse. Cohomologie d’intersection et fonctions LL de certaines variétés de Shimura. Ann. Sci. École Norm. Sup. (4), 17(3):361–412, 1984.
  • [C86] H. Carayol. Sur les représentations ll-adiques associées aux formes modulaires de Hilbert. Ann. Sci. École Norm. Sup. (4), 19(3):409–468, 1986.
  • [CR] C. Curtis and I. Reiner. Methods of representation theory. Vol. II. Pure and Applied Mathematics (New York). John Wiley & Sons, Inc., New York, 1987. With applications to finite groups and orders, A Wiley-Interscience Publication.
  • [dC12] M. A. de Cataldo. The perverse filtration and the Lefschetz hyperplane theorem, II. J. Algebraic Geom., 21(2):305–345, 2012.
  • [dCM15] M.A. de Cataldo and L. Migliorini. The projectors of the decomposition theorem are motivated. Math. Res. Lett., 22(4):1061–1088, 2015.
  • [D71] P. Deligne. Formes modulaires et représentations \ell-adiques. In Séminaire Bourbaki. Vol. 1968/69: Exposés 347–363, volume 175 of Lecture Notes in Math., pages Exp. No. 355, 139–172. Springer, Berlin, 1971.
  • [D74] P. Deligne. La conjecture de Weil. I. Inst. Hautes Études Sci. Publ. Math., (43):273–307, 1974.
  • [Di13] M. Dimitrov. Automorphic symbols, pp-adic LL-functions and ordinary cohomology of Hilbert modular varieties. Amer. J. Math., 135(4):1117–1155, 2013.
  • [EPW06] M. Emerton, R. Pollack, and T. Weston. Variation of Iwasawa invariants in Hida families. Invent. Math., 163(3):523–580, 2006.
  • [Fa88] G. Faltings. pp-adic Hodge theory. J. Amer. Math. Soc., 1(1):255–299, 1988.
  • [Fa89] G. Faltings. Crystalline cohomology and pp-adic Galois-representations. In Algebraic analysis, geometry, and number theory (Baltimore, MD, 1988), pages 25–80. Johns Hopkins Univ. Press, Baltimore, MD, 1989.
  • [Fu00] K. Fujiwara. Independence of \ell for intersection cohomology (after Gabber). In Algebraic geometry 2000, Azumino (Hotaka), volume 36 of Adv. Stud. Pure Math., pages 145–151. Math. Soc. Japan, Tokyo, 2002.
  • [Ha09] M. Harris. Potential automorphy of odd-dimensional symmetric powers of elliptic curves and applications. In Algebra, arithmetic, and geometry: in honor of Yu. I. Manin. Vol. II, volume 270 of Progr. Math., pages 1–21. Birkhäuser Boston, Inc., Boston, MA, 2009.
  • [HSBT10] M. Harris, N. Shepherd-Barron, and R. Taylor. A family of Calabi-Yau varieties and potential automorphy. Ann. of Math. (2), 171(2):779–813, 2010.
  • [He82] G. Henniart. Représentations \ell-adiques abéliennes. In Seminar on Number Theory, Paris 1980-81 (Paris, 1980/1981), volume 22 of Progr. Math., pages 107–126. Birkhäuser Boston, Boston, MA, 1982.
  • [IM19] F. Ivorra and S. Morel. The four operations on perverse motives (arxiv:1901.02096v1).
  • [K71] N.M. Katz. On a theorem of Ax. Amer. J. Math., 93:485–499, 1971.
  • [KL85] N.M. Katz and G. Laumon. Transformation de Fourier et majoration de sommes exponentielles. Inst. Hautes Études Sci. Publ. Math., (62):361–418, 1985.
  • [KM74] N.M. Katz and W. Messing. Some consequences of the Riemann hypothesis for varieties over finite fields. Invent. Math., 23:73–77, 1974.
  • [Ma73] B. Mazur. Frobenius and the Hodge filtration (estimates). Ann. of Math. (2), 98:58–95, 1973.
  • [Mo08] S. Morel. Complexes pondérés sur les compactifications de Baily-Borel: le cas des variétés de Siegel. J. Amer. Math. Soc., 21(1):23–61, 2008.
  • [N01] J. Nekovář. On the parity of ranks of Selmer groups. II. C. R. Acad. Sci. Paris Sér. I Math., 332(2):99–104, 2001.
  • [Oc06] T. Ochiai. On the two-variable Iwasawa main conjecture. Compos. Math., 142(5):1157–1200, 2006.
  • [Og82] A. Ogus. Hodge cycles and crystalline cohomology. In Hodge cycles, Motives and Shimura varieties (LNM 900), pages 357–414. Springer-Verlag, 1982.
  • [Oh83] M. Ohta. On the zeta function of an abelian scheme over the Shimura curve. Japan. J. Math. (N.S.), 9(1):1–25, 1983.
  • [Pa16] S. Patrikis. Generalized Kuga-Satake theory and rigid local systems, II: rigid Hecke eigensheaves. Algebra Number Theory, 10(7):1477–1526, 2016.
  • [Pi98] R. Pink. \ell-adic algebraic monodromy groups, cocharacters, and the Mumford-Tate conjecture. J. Reine Angew. Math., 495:187–237, 1998.
  • [Ra78] M. Rapoport. Compactifications de l’espace de modules de Hilbert-Blumenthal. Compositio Math., 36(3):255–335, 1978.
  • [Ri76] K. Ribet. Galois action on division points of Abelian varieties with real multiplications. Amer. J. Math., 98(3):751–804, 1976.
  • [Se98] J.-P. Serre. Abelian \ell-adic representations and elliptic curves, volume 7 of Research Notes in Mathematics. A K Peters, Ltd., Wellesley, MA, 1998. With the collaboration of Willem Kuyk and John Labute, Revised reprint of the 1968 original.
  • [Se12] J.-P. Serre. Lectures on NX(p)N_{X}(p), volume 11 of Chapman & Hall/CRC Research Notes in Mathematics. CRC Press, Boca Raton, FL, 2012.
  • [Se13] J.-P. Serre. Oeuvres/Collected papers. IV. 1985–1998. Springer Collected Works in Mathematics. Springer, Heidelberg, 2013. Reprint of the 2000 edition [MR1730973].
  • [SkUr14] Ch. Skinner and E. Urban. The Iwasawa main conjectures for GL2\mathrm{G}L_{2}. Invent. Math., 195(1):1–277, 2014.
  • [T89] R. Taylor. On Galois representations associated to Hilbert modular forms. Invent. Math., 98(2):265–280, 1989.
  • [T95] R. Taylor. On Galois representations associated to Hilbert modular forms. II. In Elliptic curves, modular forms, & Fermat’s last theorem (Hong Kong, 1993), Ser. Number Theory, I, pages 185–191. Int. Press, Cambridge, MA, 1995.
  • [T95] R. Taylor. Representations of Galois groups associated to modular forms. In Proceedings of the International Congress of Mathematicians, Vol. 1, 2 (Zürich, 1994), pages 435–442. Birkhäuser, Basel, 1995.
  • [Ws81] M. Waldschmidt. Transcendance et exponentielles en plusieurs variables. Invent. Math., 63(1):97–127, 1981.
  • [Wn15] X. Wan. The Iwasawa main conjecture for Hilbert modular forms. Forum Math. Sigma, 3:e18, 95, 2015.
  • [Wi88] A. Wiles. On ordinary λ\lambda-adic representations associated to modular forms. Invent. Math., 94(3):529–573, 1988.