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

Harder’s denominator problem for SL2()\mathrm{SL}_{2}(\mathbb{Z}) and its applications

Hohto Bekki Hohto Bekki
Max Planck Institute for Mathematics
Vivatsgasse 7
53111 Bonn
Germany
[email protected]
 and  Ryotaro Sakamoto Ryotaro Sakamoto
Department of Mathematics
University of Tsukuba
1-1-1 Tennodai
Tsukuba
Ibaraki 305-8571
Japan
[email protected]
Abstract.

The aim of this paper is to give a full detail of the proof given by Harder of a theorem on the denominator of the Eisenstein class for SL2()\mathrm{SL}_{2}(\mathbb{Z}) and to show that the theorem has some interesting applications including the proof of a recent conjecture by Duke on the integrality of the higher Rademacher symbols. We also present a sharp universal upper bound for the denominators of the values of partial zeta functions associated with narrow ideal classes of real quadratic fields in terms of the denominator of the values of the Riemann zeta function.

Key words and phrases:
Eisenstein class, special values of partial zeta functions, real quadratic fields, Duke’s conjecture
2020 Mathematics Subject Classification:
11F75, 11R42, 11F11, 11F67, 11R23

1. Introduction

1.1. The denominator of Eisenstein classes for SL2()\mathrm{SL}_{2}(\mathbb{Z})

Let Γ:=SL2()\Gamma:=\mathrm{SL}_{2}(\mathbb{Z}) be the modular group, and let :={zIm(z)>0}\mathbb{H}:=\{z\in\mathbb{C}\mid\operatorname{Im}(z)>0\} be the upper half plane, on which Γ\Gamma acts by the linear fractional transformation. We denote by Y:=Γ\Y:=\Gamma\backslash\mathbb{H} the modular curve of level Γ\Gamma, by YBSY^{\mathrm{BS}} its Borel–Serre compactification, and by YBS:=YBSY\partial Y^{\mathrm{BS}}:=Y^{\mathrm{BS}}\!-\!Y the Borel–Serre boundary of YY. For any even integer n2n\geq 2, we define a left Γ\Gamma-module n:=Symn(2)\mathcal{M}_{n}:=\mathrm{Sym}^{n}(\mathbb{Z}^{2}) to be the nn-th symmetric power of 2\mathbb{Z}^{2}, and define n:=Hom(n,)\mathcal{M}_{n}^{\vee}:=\operatorname{Hom}_{\mathbb{Z}}(\mathcal{M}_{n},\mathbb{Z}) to be its dual Γ\Gamma-module. Then the Γ\Gamma-module n\mathcal{M}_{n}^{\vee} naturally defines a sheaf on YBSY^{\mathrm{BS}} (which we also denote by n\mathcal{M}_{n}^{\vee}), and we can consider the cohomology groups H(YBS,n)H^{\bullet}(Y^{\mathrm{BS}},\mathcal{M}_{n}^{\vee}) and H(YBS,n)H^{\bullet}(\partial Y^{\mathrm{BS}},\mathcal{M}_{n}^{\vee}). Since n\mathcal{M}_{n}^{\vee} has a left action of M2()M_{2}(\mathbb{Z}), these cohomology groups carry the structure of Hecke modules.

The boundary YBS\partial Y^{\mathrm{BS}} is identified with Γ\\Gamma_{\infty}\backslash\mathbb{R}, where Γ:={(1a01)|a}\Gamma_{\infty}:=\left\{\begin{pmatrix}1&a\\ 0&1\end{pmatrix}\,\middle|\,a\in\mathbb{Z}\right\}. Hence it is easy to see that dimH1(YBS,n)=1\dim_{\mathbb{Q}}H^{1}(\partial Y^{\mathrm{BS}},\mathcal{M}_{n}^{\vee}\otimes\mathbb{Q})=1 and we have a natural generator ωnH1(YBS,n)/(torsion)\omega_{n}\in H^{1}(\partial Y^{\mathrm{BS}},\mathcal{M}_{n}^{\vee})/(\text{torsion}).

Harder considered in his book [Har] a unique Hecke-equivariant section

H1(YBS,n)H1(YBS,n)H^{1}(\partial Y^{\mathrm{BS}},\mathcal{M}_{n}^{\vee}\otimes\mathbb{Q})\longrightarrow H^{1}(Y^{\mathrm{BS}},\mathcal{M}_{n}^{\vee}\otimes\mathbb{Q})

of the canonical homomorphism H1(YBS,n)H1(YBS,n)H^{1}(Y^{\mathrm{BS}},\mathcal{M}_{n}^{\vee}\otimes\mathbb{Q})\longrightarrow H^{1}(\partial Y^{\mathrm{BS}},\mathcal{M}_{n}^{\vee}\otimes\mathbb{Q}) induced by the inclusion map YBS\longhookrightarrowYBS\partial Y^{\mathrm{BS}}\longhookrightarrow Y^{\mathrm{BS}}, and he defined the Eisenstein cohomology class

EisnH1(YBS,n)\mathrm{Eis}_{n}\in H^{1}(Y^{\mathrm{BS}},\mathcal{M}_{n}^{\vee}\otimes\mathbb{Q})

to be the image of ωn\omega_{n} under this section. Then Harder studied the denominator Δ(Eisn)\Delta(\mathrm{Eis}_{n}) of the Eisenstein cohomology class Eisn\mathrm{Eis}_{n}, that is, the smallest positive integer Δ(Eisn)\Delta(\mathrm{Eis}_{n}) such that

Δ(Eisn)EisnH1(YBS,n)/(torsion)H1(YBS,n).\Delta(\mathrm{Eis}_{n})\mathrm{Eis}_{n}\in H^{1}(Y^{\mathrm{BS}},\mathcal{M}_{n}^{\vee})/(\text{torsion})\subset H^{1}(Y^{\mathrm{BS}},\mathcal{M}_{n}^{\vee}\otimes\mathbb{Q}).

See also the dissertation [Wan89] of Wang. As a result, the following theorem was obtained by Harder (see [HP92, Staz 2, §1]).

Theorem 1.1 (Harder [Har, Theorem 5.1.2]).

For any even integer n2n\geq 2, we have

Δ(Eisn)=the numerator of ζ(1n),\Delta(\mathrm{Eis}_{n})=\textrm{the numerator of }\zeta(-1-n),

where ζ(s)\zeta(s) denotes the Riemann zeta function111In the present paper, the numerator and the denominator of a rational number are always defined to be positive integers..

Remark 1.2.

Haberland had proved a slightly weaker version of Theorem 1.1 in [Hab83, pp. 272–273]. More precisely, let pp be a prime number and assume that p>np>n. Then Haberland obtained

ordp(Δ(Eisn))ordp(the numerator of ζ(1n)).\mathrm{ord}_{p}(\Delta(\mathrm{Eis}_{n}))\leq\mathrm{ord}_{p}(\textrm{the numerator of }\zeta(-1-n)).

Moreover, if we further assume that pp divides ζ(1n)\zeta(-1-n) and that there exists ν{1,,n1}\nu\in\{1,\dots,n-1\} such that pζ(ν)ζ(νn)p\nmid\zeta(-\nu)\zeta(\nu-n), then he obtained the equality

ordp(Δ(Eisn))=ordp(the numerator of ζ(1n)).\mathrm{ord}_{p}(\Delta(\mathrm{Eis}_{n}))=\mathrm{ord}_{p}(\textrm{the numerator of }\zeta(-1-n)).
Remark 1.3.

For any prime numbers pp and \ell with p(p1)\ell\nmid p(p-1), the denominators over \mathbb{Z}_{\ell} of Eisenstein classes for Γ1(p)\Gamma_{1}(p) with a character was computed by Kaiser in the diploma thesis [Kai90] (see also the paper [Mah00] of Mahnkopf for the study of Eisenstein classes for Γ1(pe)\Gamma_{1}(p^{e}\ell)). Eisenstein classes for GL2\mathrm{GL}_{2} over totally real fields have been studied by Maennel in the dissertation [Mae93]. Eisenstein classes for GL2\mathrm{GL}_{2} over imaginary quadratic fields have been studied by Harder in [Har81, Har82], Weselmann in [Wes88], Berger in [Ber08, Ber09], and Branchereau in [Bra23].

The purpose of the present paper is to report some arithmetic applications of Theorem 1.1 especially to the special values of partial zeta functions of real quadratic fields. However, since the book [Har] (which is available on Harder’s web-page) is still under development, some of the important arguments and references in the proof of Theorem 1.1 are currently not given completely. Taking this situation into account, we decided to also give the detailed proof of Theorem 1.1, which is another main purpose of the present paper.

1.2. Reformulation in terms of the holomorphic Eisenstein series

In view of applications, we interpret the above definition of the Eisenstein class Eisn\mathrm{Eis}_{n} and Theorem 1.1 in terms of the holomorphic Eisenstein series and the Eichler–Shimura homomorphism.

In the following, let n2n\geq 2 be an even integer and let Mn+2(Γ)M_{n+2}(\Gamma) denotes the space of modular forms of level Γ=SL2()\Gamma=\mathrm{SL}_{2}(\mathbb{Z}) and weight n+2n+2. Then we have the Hecke-equivariant homomorphism

r:Mn+2(Γ)H1(Y,n)=H1(YBS,n)r\colon M_{n+2}(\Gamma)\longrightarrow H^{1}(Y,\mathcal{M}_{n}^{\vee}\otimes\mathbb{C})=H^{1}(Y^{\mathrm{BS}},\mathcal{M}_{n}^{\vee}\otimes\mathbb{C})

called the Eichler–Shimura homomorphism which is defined by certain path integrals on \mathbb{H} (see §2.6 for the precise definition of the Eichler–Shimura homomorphism). Let

En+2(z):=1+2ζ(1n)k=1σn+1(k)qkMn+2(Γ),q:=e2πizE_{n+2}(z):=1+\frac{2}{\zeta(-1-n)}\sum_{k=1}^{\infty}\sigma_{n+1}(k)q^{k}\in M_{n+2}(\Gamma),\quad q:=e^{2\pi iz}

denote the holomorphic Eisenstein series of weight n+2n+2. Then the following is the reformulation of Theorem 1.1 ([Har, Theorem 5.1.2]) which will be proved in the present paper.

Theorem 1.4 (Lemma 2.8, Proposition 2.11, and Theorem 2.13).
  1. (1)

    We have r(En+2)=Eisnr(E_{n+2})=\mathrm{Eis}_{n}, i.e., the class r(En+2)r(E_{n+2}) coincides with the Eisenstein class Eisn\mathrm{Eis}_{n}.

  2. (2)

    The denominator of r(En+2)r(E_{n+2}) is equal to the numerator of ζ(1n)\zeta(-1-n).

Remark 1.5.

From the qq-expansion of the Eisenstein series En+2E_{n+2}, we see that the denominator of En+2E_{n+2} with respect to the integral structure coming from the qq-expansion (the de Rham integral structure) is clearly the numerator of ζ(1n)\zeta(-1-n). An interesting and non-trivial point in Theorem 1.4 is that on the Eisenstein parts En+2\mathbb{Q}E_{n+2} and Eisn\mathbb{Q}\mathrm{Eis}_{n}, the Betti integral structure coincides with the de Rham integral structure under the Eichler–Shimura homomorphism. Cf. [Har21, §1.1] and Remark 2.14.

1.3. Strategy of the proof of Theorem 1.4

We review the strategy of the proof of Theorem 1.4, which is based on Harder’s argument in [Har]. First, note that for any prime number pp, we have

ordp(Δ(Eisn))=min{δ0pδEisnH1(YBS,n(p))}.\mathrm{ord}_{p}(\Delta(\mathrm{Eis}_{n}))=\min\{\delta\in\mathbb{Z}_{\geq 0}\mid p^{\delta}\mathrm{Eis}_{n}\in H^{1}(Y^{\mathrm{BS}},\mathcal{M}_{n}^{\vee}\otimes\mathbb{Z}_{(p)})\}.

Therefore, it suffices to prove that

ordp(Δ(Eisn))=ordp(the numerator of ζ(1n))\mathrm{ord}_{p}(\Delta(\mathrm{Eis}_{n}))=\mathrm{ord}_{p}(\text{the numerator of }\zeta(-1-n))

for each prime number pp. Then the proof consists roughly of the following four parts.

  1. (I)

    First, we construct in §3 a certain family of pp-adically integral homology classes

    Tpm(Cν(τ))~H1(YBS,n(p))\widetilde{T_{p}^{m}(C_{\nu}(\tau))}\in H_{1}(Y^{\mathrm{BS}},\mathcal{M}_{n}\otimes\mathbb{Z}_{(p)})

    for any τ\tau\in\mathbb{H}, sufficiently large integer mm, and 1νn11\leq\nu\leq n-1, where H1(YBS,n(p))H_{1}(Y^{\mathrm{BS}},\mathcal{M}_{n}\otimes\mathbb{Z}_{(p)}) is the cosheaf homology group associated with the Γ\Gamma-module n(p)\mathcal{M}_{n}\otimes\mathbb{Z}_{(p)}. Note that the homology class Tpm(Cν(τ))~\widetilde{T_{p}^{m}(C_{\nu}(\tau))} is independent of the choice of τ\tau\in\mathbb{H}. See Lemma 3.15.

  2. (II)

    Next, in §4, we compute the pp-adic limit of the value of the pairing

    limmEisn,Tpm!(Cν(τ))~,\lim_{m\to\infty}\langle\mathrm{Eis}_{n},\widetilde{T_{p}^{m!}(C_{\nu}(\tau))}\rangle,

    where ,:H1(YBS,n)×H1(YBS,n)\langle~{},~{}\rangle\colon H^{1}(Y^{\mathrm{BS}},\mathcal{M}_{n}^{\vee})\times H_{1}(Y^{\mathrm{BS}},\mathcal{M}_{n})\longrightarrow\mathbb{Z} is the pairing induced by n×n;(f,x)f(x)\mathcal{M}_{n}^{\vee}\times\mathcal{M}_{n}\longrightarrow\mathbb{Z};(f,x)\mapsto f(x). More precisely, we will show in Theorem 4.1 and Corollary 7.2 that this pp-adic limit can be described in terms of the Kubota–Leopoldt pp-adic LL-functions, namely, for any integer ν{1,,n1}\nu\in\{1,\ldots,n-1\} we obtain the following interesting formula

    limmEisn,Tpm!(Cν(τ))~=1pn+1(1pν)(1pnν)Dp(n,ν),\displaystyle\lim_{m\to\infty}\langle\mathrm{Eis}_{n},\widetilde{T_{p}^{m!}(C_{\nu}(\tau))}\rangle=\frac{1-p^{n+1}}{(1-p^{\nu})(1-p^{n-\nu})}D_{p}(n,\nu),

    where

    Dp(n,ν):=Lp(ν,ω1+ν)Lp(νn,ωnν+1)Lp(1n,ωn+2)Lp(ν,ω1+ν)Lp(νn,ωnν+1)D_{p}(n,\nu):=\frac{L_{p}(-\nu,\omega^{1+\nu})L_{p}(\nu-n,\omega^{n-\nu+1})}{L_{p}(-1-n,\omega^{n+2})}-L_{p}(-\nu,\omega^{1+\nu})-L_{p}(\nu-n,\omega^{n-\nu+1})

    and Lp(s,ωa)L_{p}(s,\omega^{a}) denotes the Kubota–Leopoldt pp-adic LL-function associated with the aa-th power of the Teichmüller character ω\omega.

  3. (III)

    In §5, we prove that the homology classes Tpm!(C1(τ))~,,Tpm!(Cn1(τ))~\widetilde{T_{p}^{m!}(C_{1}(\tau))},\ldots,\widetilde{T_{p}^{m!}(C_{n-1}(\tau))} give a generator of the ordinary part of the quotient group (H1(YBS,n)/H1(YBS,n))p(H_{1}(Y^{\mathrm{BS}},\mathcal{M}_{n})/H_{1}(\partial Y^{\mathrm{BS}},\mathcal{M}_{n}))\otimes\mathbb{Z}_{p} (see Proposition 5.7). Hence, since Eisn,H1(YBS,n)=\langle\mathrm{Eis}_{n},H_{1}(\partial Y^{\mathrm{BS}},\mathcal{M}_{n})\rangle=\mathbb{Z}, we obtain

    (1.1) ordp(Δ(Eisn))=min{a0paD(n,ν)p for any integer 1νn1}.\displaystyle\mathrm{ord}_{p}(\Delta(\mathrm{Eis}_{n}))=\min\{a\in\mathbb{Z}_{\geq 0}\mid p^{a}D(n,\nu)\in\mathbb{Z}_{p}\textrm{ for any integer }1\leq\nu\leq n-1\}.

    See Corollary 5.8, and also Proposition 8.1.

  4. (IV)

    In §8, we show that the right hand side of (1.1) is equal to the pp-adic valuation of the numerator of ζ(1n)\zeta(-1-n). We devote §6 and §7 to the preparation for proving this fact.

1.4. Applications to Duke’s conjecture and to the special values of the partial zeta functions of real quadratic fields

1.4.1. Duke’s conjecture

In the paper [Duk23], Duke defined a certain map

Ψk:Γ=SL2()\Psi_{k}:\Gamma=\mathrm{SL}_{2}(\mathbb{Z})\longrightarrow\mathbb{Q}

for each integer k2k\geq 2 called the higher Rademacher symbol which is a generalization of the classical Rademacher symbol, and he conjectured the integrality of the higher Rademacher symbol ([Duk23, Conjecture, p. 4]). As a first application of Theorem 1.4, we prove this conjecture.

Theorem 1.6 (Corollary 9.5).

Duke’s conjecture holds true, that is, for any integer k2k\geq 2 and matrix γΓ\gamma\in\Gamma, we have

Ψk(γ).\Psi_{k}(\gamma)\in\mathbb{Z}.

In fact, Duke proved in [Duk23, Lemma 6] that the higher Rademacher symbols can be written as the integral of the holomorphic Eisenstein series along a certain homology cycles (see Proposition 9.4). Therefore, we can derive Theorem 1.6 directly from Theorem 1.4.

Remark 1.7.

Duke’s conjecture is recently proved also by O’Sullivan in [O’S23] using a more direct method.

1.4.2. The denominators of the partial zeta functions of real quadratic fields

Next, we discuss the denominators of the partial zeta functions associated with narrow ideal classes of orders in real quadratic fields.

Let FF\subset\mathbb{R} be a real quadratic field, 𝒪F\mathcal{O}\subset F be an order in FF, and 𝒜Cl𝒪+\mathcal{A}\in Cl_{\mathcal{O}}^{+} be a narrow ideal class of 𝒪\mathcal{O}. Then we have the associated partial zeta function

ζ𝒪(𝒜,s)=𝔞𝒪,𝔞𝒜1N𝔞s,\zeta_{\mathcal{O}}(\mathcal{A},s)=\sum_{\mathfrak{a}\subset\mathcal{O},\mathfrak{a}\in\mathcal{A}}\frac{1}{N\mathfrak{a}^{s}},

which can be continued meromorphically to \mathbb{C}, and it is known that

ζ𝒪(𝒜,1k)\zeta_{\mathcal{O}}(\mathcal{A},1-k)\in\mathbb{Q}

for any integer k2k\geq 2. We also define the positive integer J2kJ_{2k} by

J2k:=the denominator of ζ(12k).J_{2k}:=\text{the denominator of }\zeta(1-2k).

Then in §9.2, we obtain the following as another consequence of Theorem 1.4.

Proposition 1.8 (Corollary 9.12).

Let k2k\geq 2 be an integer. Then the integer J2kJ_{2k} gives a universal upper bound for the denominator of ζ𝒪(𝒜,1k)\zeta_{\mathcal{O}}(\mathcal{A},1-k) with respect to orders 𝒪\mathcal{O} and narrow ideal classes 𝒜Cl𝒪+\mathcal{A}\in Cl_{\mathcal{O}}^{+}. In other words, we have

J2kζ𝒪(𝒜,1k)J_{2k}\zeta_{\mathcal{O}}(\mathcal{A},1-k)\in\mathbb{Z}

for all orders 𝒪\mathcal{O} in all real quadratic fields and narrow ideal classes 𝒜Cl𝒪+\mathcal{A}\in Cl_{\mathcal{O}}^{+}.

In fact, one can construct a natural map 𝔷𝒪,k:Cl𝒪+H1(YBS,2k2)\mathfrak{z}_{\mathcal{O},k}\colon Cl_{\mathcal{O}}^{+}\longrightarrow H_{1}(Y^{\mathrm{BS}},\mathcal{M}_{2k-2}) for any integer k2k\geq 2 (see Definition 9.6), and we show in Proposition 9.10 that

Eis2k2,𝔷𝒪,k(𝒜1)=(1)kζ𝒪(𝒜,1k)ζ(12k)=±J2kζ𝒪(𝒜,1k)N2k.\langle\mathrm{Eis}_{2k-2},\mathfrak{z}_{\mathcal{O},k}(\mathcal{A}^{-1})\rangle=(-1)^{k}\frac{\zeta_{\mathcal{O}}(\mathcal{A},1-k)}{\zeta(1-2k)}=\pm\frac{J_{2k}\zeta_{\mathcal{O}}(\mathcal{A},1-k)}{N_{2k}}.

Hence Proposition 1.8 follows from Theorem 1.4. See also Remark 9.13 for the relation between Duke’s conjecture and Proposition 1.8.

Next, we discuss the sharpness of the universal upper bound obtained in Proposition 1.8.

Theorem 1.9 (Corollary 9.16).

The universal upper bound in Proposition 1.8 is sharp, that is, we have

J2k=min{J>0|Jζ𝒪(𝒜,1k) for all orders 𝒪 in all real quadratic fieldsand narrow ideal classes 𝒜Cl𝒪+ }.J_{2k}=\min\left\{J\in\mathbb{Z}_{>0}\,\,\middle|\,\,\begin{aligned} J\zeta_{\mathcal{O}}(\mathcal{A},1-k)\in\mathbb{Z}\text{ for all orders $\mathcal{O}$ in all real quadratic fields}\\ \text{and narrow ideal classes $\mathcal{A}\in Cl_{\mathcal{O}}^{+}$ }\end{aligned}\,\right\}.

In order to derive Theorem 1.9 from Theorem 1.4, we need to show that the narrow ideal classes of orders in real quadratic fields produce sufficiently large submodule of the homology group H1(YBS,2k2)H_{1}(Y^{\mathrm{BS}},\mathcal{M}_{2k-2}), and this will be done in §9.3 using some techniques from Hida theory.

Remark 1.10.

As for the denominator or the integrality of the special values of partial zeta functions of real quadratic fields, or more generally of totally real fields, many works have been done by Coates and Sinnott [CS74a, CS74b, CS77], Deligne and Ribet [DR80], Cassou-Noguès [CN79], Charollois, Dasgupta, and Greenberg [CDG15], Beilinson, Kings, and Levin [BKL18], Bannai, Hagihara, Yamada, and Yamamoto [BHYY22], Bergeron, Charollois, and Garcia [BCG20] etc., by using variety of methods including Hilbert modular forms, Shintani zeta functions [Shi76], Sczech’s Eisenstein cocycles [Scz93], etc. Actually, when 𝒪=𝒪F\mathcal{O}=\mathcal{O}_{F}, the upper bound in Proposition 1.8 follows from these preceding works. More precisely, the results proved by Coates and Sinnott in [CS77] or Deligne and Ribet in [DR80] show that for any prime number pp, we have

21(1p2k)ζ𝒪F(𝒜,1k)[1/p],2^{-1}(1-p^{2k})\zeta_{\mathcal{O}_{F}}(\mathcal{A},1-k)\in\mathbb{Z}[1/p],

which implies that

J2kζ𝒪F(𝒜,1k)J_{2k}\zeta_{\mathcal{O}_{F}}(\mathcal{A},1-k)\in\mathbb{Z}

(see [Zag76, pp. 73, 75]).

One feature of the method in the present paper is that by using Theorem 1.4, we capture not only the upper bound for the denominators of the partial zeta functions associated with any orders, but also the sharpness of the upper bound.

Remark 1.11.

In the paper [Zag77], Zagier proved a certain formula [Zag77, p. 149, Corollaire] which explicitly computes the special values ζ𝒪(𝒜,1k)\zeta_{\mathcal{O}}(\mathcal{A},1-k) of partial zeta functions of orders of real quadratic fields at negative integers in a uniform way. Then by using this formula, he obtained a universal upper bound dkd_{k} of the denominators of the values ζ𝒪(𝒜,1k)\zeta_{\mathcal{O}}(\mathcal{A},1-k) and examined its sharpness briefly. More precisely, he observed that the upper bound dkd_{k} is not sharp and discussed how one can improve this upper bound when k=2,3k=2,3 (see [Zag77, pp. 149–150]). Theorem 1.9 can be seen as the complete answer to this problem of determining the sharp universal upper bound for the denominators of ζ𝒪(𝒜,1k)\zeta_{\mathcal{O}}(\mathcal{A},1-k).

Acknowledgements

We would like to express our deepest gratitude to Günter Harder who explained to us many beautiful ideas and showed us his notes and manuscripts on the proof of his theorem. We would also like to thank him for encouraging us to write the proof of his theorem in this paper. We are also grateful to Herbert Gangl, Christian Kaiser, and Don Zagier for the fruitful discussions and many valuable comments during the study. In particular, Herbert Gangl and Don Zagier suggested us to conduct a numerical experiment to test the sharpness of Proposition 1.8, and helped us to write PARI/GP programs for the experiment, which provided us with some important ideas to prove Theorem 1.9. Thanks are also due to Toshiki Matsusaka who drew our attention to Duke’s conjecture, which became one of the main motivations in this study. This research has been carried out during the first author’s stay at the Max Planck Institute for Mathematics in Bonn.

2. Preliminaries and the Eisenstein class

In this section, we give the definition of the Eisenstein class and explain Theorem 1.4 (see Theorem 2.13).

Throughout this paper, n2n\geq 2 denotes an even integer.

2.1. Definitions of Modular curve and Borel–Serre compactification

Let

:={zIm(z)>0}\mathbb{H}:=\{z\in\mathbb{C}\mid\operatorname{Im}(z)>0\}

denote the upper half plane, and let

BS:=r1()1(){r}\mathbb{H}^{\mathrm{BS}}:=\mathbb{H}\sqcup\bigsqcup_{r\in\mathbb{P}^{1}(\mathbb{Q})}\mathbb{P}^{1}(\mathbb{R})\setminus\{r\}

be the Borel–Serre compactification of \mathbb{H} (see [Gor05] or [Har, §1,2,7]). We set

Γ:=SL2().\Gamma:=\mathrm{SL}_{2}(\mathbb{Z}).

The group Γ\Gamma acts on \mathbb{H} and BS\mathbb{H}^{\mathrm{BS}} by the linear fractional transformation as usual. We denote by

Y:=Γ\ and YBS:=Γ\BS\displaystyle Y:=\Gamma\backslash\mathbb{H}\,\,\,\textrm{ and }\,\,\,Y^{\mathrm{BS}}:=\Gamma\backslash\mathbb{H}^{\mathrm{BS}}

the modular curve of level SL2()\mathrm{SL}_{2}(\mathbb{Z}) and its Borel–Serre compactification, respectively. Moreover, we denote by YBS:=YBSY\partial Y^{\mathrm{BS}}:=Y^{\mathrm{BS}}\!-\!Y the boundary of YBSY^{\mathrm{BS}}. The boundary YBS\partial Y^{\mathrm{BS}} is homeomorphic to the circle S1S^{1} and the fundamental group π1(YBS)\pi_{1}(\partial Y^{\mathrm{BS}}) can be identified with Γ:={(1a01)|a}\Gamma_{\infty}:=\left\{\begin{pmatrix}1&a\\ 0&1\end{pmatrix}\,\middle|\,a\in\mathbb{Z}\right\}.

In the following, ?\mathbb{H}^{?} (resp. Y?Y^{?}) means that it is either \mathbb{H} or BS\mathbb{H}^{\mathrm{BS}} (resp. YY or YBSY^{\mathrm{BS}}). Any left Γ\Gamma-module \mathcal{M} can be regarded as (co)sheaf on Y?Y^{?} in a natural way, and hence we can consider the homology groups

H(Y?,),H(YBS,),H(YBS,YBS,),\displaystyle H_{\bullet}(Y^{?},\mathcal{M}),\,\,\,H_{\bullet}(\partial Y^{\mathrm{BS}},\mathcal{M}),\,\,\,H_{\bullet}(Y^{\mathrm{BS}},\partial Y^{\mathrm{BS}},\mathcal{M}),

which fit into the long exact sequence

H1(YBS,)H1(YBS,)H1(YBS,YBS,)H0(YBS,).\displaystyle\cdots\rightarrow H_{1}(\partial Y^{\mathrm{BS}},\mathcal{M})\rightarrow H_{1}(Y^{\mathrm{BS}},\mathcal{M})\rightarrow H_{1}(Y^{\mathrm{BS}},\partial Y^{\mathrm{BS}},\mathcal{M})\rightarrow H_{0}(\partial Y^{\mathrm{BS}},\mathcal{M})\rightarrow\cdots.

Similarly, we have the cohomology groups

H(Y?,),H(YBS,),H(YBS,YBS,),\displaystyle H^{\bullet}(Y^{?},\mathcal{M}),\,\,\,H^{\bullet}(\partial Y^{\mathrm{BS}},\mathcal{M}),\,\,\,H^{\bullet}(Y^{\mathrm{BS}},\partial Y^{\mathrm{BS}},\mathcal{M}),

which fit into the long exact sequence

H1(YBS,YBS,)H1(YBS,)H1(YBS,)H2(YBS,YBS,).\displaystyle\cdots\rightarrow H^{1}(Y^{\mathrm{BS}},\partial Y^{\mathrm{BS}},\mathcal{M})\rightarrow H^{1}(Y^{\mathrm{BS}},\mathcal{M})\rightarrow H^{1}(\partial Y^{\mathrm{BS}},\mathcal{M})\rightarrow H^{2}(Y^{\mathrm{BS}},\partial Y^{\mathrm{BS}},\mathcal{M})\rightarrow\cdots.

We note that the inclusion map Y\longhookrightarrowYBSY\longhookrightarrow Y^{\mathrm{BS}} induces isomorphisms

H(Y,)H(YBS,) and H(YBS,)H(Y,).H_{\bullet}(Y,\mathcal{M})\stackrel{{\scriptstyle\sim}}{{\longrightarrow}}H_{\bullet}(Y^{\mathrm{BS}},\mathcal{M})\,\,\,\textrm{ and }\,\,\,H^{\bullet}(Y^{\mathrm{BS}},\mathcal{M})\stackrel{{\scriptstyle\sim}}{{\longrightarrow}}H^{\bullet}(Y,\mathcal{M}).

Moreover, if \mathcal{M} has an action of M2+():={γM2()det(γ)>0}M_{2}^{+}(\mathbb{Z}):=\{\gamma\in M_{2}(\mathbb{Z})\mid\det(\gamma)>0\}, then these homology groups (resp. cohomology groups) carry the structure of Hecke modules, namely, for each prime number pp we have a Hecke operator TpT_{p} (resp. TpT_{p}^{\prime}) on these homology groups (resp. cohomology groups), and the above long exact sequences are compatible with the Hecke operators.

In Section 2.2, we give a way to compute these (co)homology groups, and in Section 2.4, we give an explicit description of the Hecke operators.

Remark 2.1.

As a sheaf on YY, the stalk x\mathcal{M}_{x} at xYx\in Y coincides with Γx~\mathcal{M}^{\Gamma_{\tilde{x}}}, where x~\tilde{x}\in\mathbb{H} is a lift of xYx\in Y and Γx~:={γΓγx~=γ}\Gamma_{\tilde{x}}:=\{\gamma\in\Gamma\mid\gamma\tilde{x}=\gamma\}. This fact shows that a short exact sequence of left Γ\Gamma-modules does not give a short exact sequence of sheaves on YY in general, that is, the sheafification functor is not exact. However, a short exact sequence of left [1/6][Γ\mathbb{Z}[1/6][\Gamma]-modules induces a short exact sequence of sheaves on YBSY^{\mathrm{BS}} since the order of Γx~\Gamma_{\tilde{x}} divides 66.

2.2. Modular symbols and (co)homology

Let X{,BS,BS}X\in\{\mathbb{H},\mathbb{H}^{\mathrm{BS}},\partial\mathbb{H}^{\mathrm{BS}}\} and let (S(X),)(S_{\bullet}(X),\partial) denote the usual singular chain complex of XX, i.e., Sq(X)S_{q}(X) is the free abelian group generated by singular qq-simplices in XX and :Sq(X)Sq1(X)\partial\colon S_{q}(X)\longrightarrow S_{q-1}(X) is the boundary operator.

The left action of M2+()M_{2}^{+}(\mathbb{Z}) on XX induces a left action of M2+()M_{2}^{+}(\mathbb{Z}) on S(X)S_{\bullet}(X), and (S(X),)(S_{\bullet}(X),\partial) is actually an M2+()M_{2}^{+}(\mathbb{Z})-equivariant complex. Then it is known that for any left M2+()M_{2}^{+}(\mathbb{Z})-module \mathcal{M}, which is also seen as a (co)sheaf on Γ\X\Gamma\backslash X, we have natural isomorphisms

H(Γ\X,)\displaystyle H_{\bullet}(\Gamma\backslash X,\mathcal{M}) H((S(X))Γ),\displaystyle\cong H_{\bullet}((S_{\bullet}(X)\otimes\mathcal{M})_{\Gamma}),
H(YBS,YBS)\displaystyle H_{\bullet}(Y^{\mathrm{BS}},\partial Y^{\mathrm{BS}}\mathcal{M}) H(((S(BS)/S(BS))Γ),\displaystyle\cong H_{\bullet}(((S_{\bullet}(\mathbb{H}^{\mathrm{BS}})/S_{\bullet}(\partial\mathbb{H}^{\mathrm{BS}})\otimes\mathcal{M})_{\Gamma}),

where ()Γ(-)_{\Gamma} denotes the Γ\Gamma-coinvariant functor. Here the left M2+()M_{2}^{+}(\mathbb{Z})-action on S(X)S_{\bullet}(X)\otimes\mathcal{M} is defined by

γ(σm):=γσγm,\gamma\cdot(\sigma\otimes m):=\gamma\sigma\otimes\gamma m,

where σS(X)\sigma\in S_{\bullet}(X), mm\in\mathcal{M}, and γM2+()\gamma\in M_{2}^{+}(\mathbb{Z}). Set

𝒮(X):=coker(S2(X)S1(X)).\mathcal{MS}(X):=\operatorname{coker}(S_{2}(X)\stackrel{{\scriptstyle\partial}}{{\longrightarrow}}S_{1}(X)).

For any elements α,βX\alpha,\beta\in X, we denote the equivalence class of a path from α\alpha to β\beta in 𝒮(X)\mathcal{MS}(X) by

{α,β}𝒮(X).\{\alpha,\beta\}\in\mathcal{MS}(X).

The boundary map :S1(X)S0(X)\partial\colon S_{1}(X)\rightarrow S_{0}(X) induces a Γ\Gamma-homomorphism :𝒮(X)S0(X)\partial\colon\mathcal{MS}(X)\rightarrow S_{0}(X), and we have a natural isomorphism

H1(Γ\X,)ker((𝒮(X))Γ(S0(X))Γ).\displaystyle H_{1}(\Gamma\backslash X,\mathcal{M})\cong\ker((\mathcal{MS}(X)\otimes\mathcal{M})_{\Gamma}\overset{\partial}{\rightarrow}(S_{0}(X)\otimes\mathcal{M})_{\Gamma}).

Similarly, we also have natural isomorphisms

H(Γ\X,)\displaystyle H^{\bullet}(\Gamma\backslash X,\mathcal{M}) H(Hom(S(X),)Γ),\displaystyle\cong H^{\bullet}(\operatorname{Hom}_{\mathbb{Z}}(S_{\bullet}(X),\mathcal{M})^{\Gamma}),
H(YBS,YBS)\displaystyle H^{\bullet}(Y^{\mathrm{BS}},\partial Y^{\mathrm{BS}}\mathcal{M}) H(Hom(S(BS)/S(BS),)Γ),\displaystyle\cong H^{\bullet}(\operatorname{Hom}_{\mathbb{Z}}(S_{\bullet}(\mathbb{H}^{\mathrm{BS}})/S_{\bullet}(\partial\mathbb{H}^{\mathrm{BS}}),\mathcal{M})^{\Gamma}),

where ()Γ(-)^{\Gamma} denotes the Γ\Gamma-invariant functor. Here the left M2+()M_{2}^{+}(\mathbb{Z})-action on Hom(S(X),)\operatorname{Hom}_{\mathbb{Z}}(S_{\bullet}(X),\mathcal{M}) is defined by

(γϕ)(σ):=γ(ϕ(γ~σ)),(\gamma\phi)(\sigma):=\gamma(\phi(\tilde{\gamma}\sigma)),

where ϕHom(S(X),)\phi\in\operatorname{Hom}_{\mathbb{Z}}(S_{\bullet}(X),\mathcal{M}), σS(X)\sigma\in S_{\bullet}(X), and γ~\tilde{\gamma} is the adjugate of γM2+()\gamma\in M_{2}^{+}(\mathbb{Z}). Since

ker(Hom(S1(X),)Hom(S2(X),))=Hom(𝒮(X),),\ker\left(\operatorname{Hom}_{\mathbb{Z}}(S_{1}(X),\mathcal{M})\rightarrow\operatorname{Hom}_{\mathbb{Z}}(S_{2}(X),\mathcal{M})\right)=\operatorname{Hom}_{\mathbb{Z}}(\mathcal{MS}(X),\mathcal{M}),

we have a natural isomorphism

H1(Γ\X,)coker(Hom(S0(X),)ΓHom(𝒮(X),)Γ).H^{1}(\Gamma\backslash X,\mathcal{M})\cong\operatorname{coker}(\operatorname{Hom}_{\mathbb{Z}}(S_{0}(X),\mathcal{M})^{\Gamma}\rightarrow\operatorname{Hom}_{\mathbb{Z}}(\mathcal{MS}(X),\mathcal{M})^{\Gamma}).

2.3. M2+()M_{2}^{+}(\mathbb{Z})-modules n\mathcal{M}_{n} and n\mathcal{M}_{n}^{\flat}

For any 2×22\times 2 matrix γ=(abcd)\gamma=\begin{pmatrix}a&b\\ c&d\end{pmatrix}, we denote the adjugate of γ\gamma by

γ~:=(dbca).\widetilde{\gamma}:=\begin{pmatrix}d&-b\\ -c&a\end{pmatrix}.

Note that if γΓ\gamma\in\Gamma, then we have γ~=γ1\widetilde{\gamma}=\gamma^{-1}.

Let [X1,X2]\mathbb{Z}[X_{1},X_{2}] denote the ring of polynomials of two variables over \mathbb{Z}, and we equip [X1,X2]\mathbb{Z}[X_{1},X_{2}] with a left action of M2+()M_{2}^{+}(\mathbb{Z}) by

(γP)(X1,X2):=\displaystyle(\gamma P)(X_{1},X_{2}):= P(dX1bX2,cX1+aX2)\displaystyle P(dX_{1}-bX_{2},-cX_{1}+aX_{2})
=\displaystyle= P((X1,X2)γ~t),\displaystyle P\left((X_{1},X_{2})\cdot{}^{t}\widetilde{\gamma}\,\right),

where P[X1,X2]P\in\mathbb{Z}[X_{1},X_{2}] and γM2+()\gamma\in M_{2}^{+}(\mathbb{Z}). For each integer 0νn0\leq\nu\leq n, we set

eν:=X1νX2nν and eν:=(1)nν(nν)X1nνX2ν.\displaystyle e_{\nu}:=X_{1}^{\nu}X_{2}^{n-\nu}\,\,\,\textrm{ and }\,\,\,e_{\nu}^{\flat}:=(-1)^{n-\nu}\binom{n}{\nu}X_{1}^{n-\nu}X_{2}^{\nu}.

We then define submodules n\mathcal{M}_{n} and n\mathcal{M}_{n}^{\flat} of [X1,X2]\mathbb{Z}[X_{1},X_{2}] by

n:=ν=0neν and n:=ν=0neν.\displaystyle\mathcal{M}_{n}:=\bigoplus_{\nu=0}^{n}\mathbb{Z}e_{\nu}\,\,\,\textrm{ and }\,\,\,\mathcal{M}_{n}^{\flat}:=\bigoplus_{\nu=0}^{n}\mathbb{Z}e_{\nu}^{\flat}.

The \mathbb{Z}-modules n\mathcal{M}_{n} and n\mathcal{M}_{n}^{\flat} are closed under the left action of M2+()M_{2}^{+}(\mathbb{Z}) on [X1,X2]\mathbb{Z}[X_{1},X_{2}]. In particular, both n\mathcal{M}_{n} and n\mathcal{M}_{n}^{\flat} are left Γ\Gamma-modules. We also define the pairing

,:n×n\langle~{},~{}\rangle\colon\mathcal{M}_{n}^{\flat}\times\mathcal{M}_{n}\rightarrow\mathbb{Z}

by

eν,eμ=δν,μ,\langle e_{\nu}^{\flat},e_{\mu}\rangle=\delta_{\nu,\mu},

where δν,μ\delta_{\nu,\mu} is the Kronecker delta. The pairing ,\langle~{},~{}\rangle is perfect and M2+()M_{2}^{+}(\mathbb{Z})-equivariant in the sense that for any polynomials PnP\in\mathcal{M}_{n}^{\flat} and QnQ\in\mathcal{M}_{n} and matrix γM2+()\gamma\in M_{2}^{+}(\mathbb{Z}), we have

P,γQ=γ~P,Q.\langle P,\gamma Q\rangle=\langle\widetilde{\gamma}P,Q\rangle.

Hence the pairing ,\langle~{},~{}\rangle induces an M2+()M_{2}^{+}(\mathbb{Z})-equivariant isomorphism

nn:=Hom(n,),m(mm,m).\mathcal{M}_{n}^{\flat}\stackrel{{\scriptstyle\sim}}{{\longrightarrow}}\mathcal{M}_{n}^{\vee}:=\operatorname{Hom}_{\mathbb{Z}}(\mathcal{M}_{n},\mathbb{Z}),m^{\prime}\mapsto\left(m\mapsto\langle m^{\prime},m\rangle\right).

Here the left action of M2+()M_{2}^{+}(\mathbb{Z}) on n=Hom(n,)\mathcal{M}_{n}^{\vee}=\operatorname{Hom}_{\mathbb{Z}}(\mathcal{M}_{n},\mathbb{Z}) is given by

(γϕ)(Q)=ϕ(γ~Q),(\gamma\phi)(Q)=\phi(\tilde{\gamma}Q),

where ϕHom(n,)\phi\in\operatorname{Hom}_{\mathbb{Z}}(\mathcal{M}_{n},\mathbb{Z}), QnQ\in\mathcal{M}_{n}, and γM2+()\gamma\in M_{2}^{+}(\mathbb{Z}).

Remark 2.2.

The left actions of M2+()M_{2}^{+}(\mathbb{Z}) on n\mathcal{M}_{n} and n\mathcal{M}_{n}^{\flat} are slightly different from the left actions used in Harder’s book [Har, (1.57)]. However, since

(11)1(abcd)(11)=(dbca),\begin{pmatrix}&-1\\ 1&\end{pmatrix}^{-1}\begin{pmatrix}a&b\\ c&d\end{pmatrix}\begin{pmatrix}&-1\\ 1&\end{pmatrix}=\begin{pmatrix}d&-b\\ -c&a\end{pmatrix},

they are isomorphic as left M2+()M_{2}^{+}(\mathbb{Z})-modules. Therefore, there are no essential differences.

2.4. Hecke operators

Let X{,BS,}X\in\{\mathbb{H},\mathbb{H}^{\mathrm{BS}},\partial\mathbb{H}\} and let \mathcal{M} be a left M2+()M_{2}^{+}(\mathbb{Z})-module. In this subsection, we define the Hecke operators on H(Γ\X,)H_{\bullet}(\Gamma\backslash X,\mathcal{M}) and H(Γ\X,)H^{\bullet}(\Gamma\backslash X,\mathcal{M}), explicitly.

For each prime number pp, we have the following double coset decomposition:

Γ\Γ(p1)Γ=Γ(p1)j=0p1Γ(1jp).\Gamma\backslash\Gamma\begin{pmatrix}p&\\ &1\end{pmatrix}\Gamma=\Gamma\begin{pmatrix}p&\\ &1\end{pmatrix}\sqcup\bigsqcup_{j=0}^{p-1}\Gamma\begin{pmatrix}1&j\\ &p\end{pmatrix}.

Hence the endomorphism

;m(p001)m+j=0p1(1j0p)m\mathcal{M}\longrightarrow\mathcal{M};m\mapsto\begin{pmatrix}p&0\\ 0&1\end{pmatrix}m+\sum_{j=0}^{p-1}\begin{pmatrix}1&j\\ 0&p\end{pmatrix}m

induces an endomorphism of Γ\mathcal{M}_{\Gamma}. Similarly, the endomorphism

;m(p001)~m+j=0p1(1j0p)~m\mathcal{M}\longrightarrow\mathcal{M};m\mapsto\widetilde{\begin{pmatrix}p&0\\ 0&1\end{pmatrix}}m+\sum_{j=0}^{p-1}\widetilde{\begin{pmatrix}1&j\\ 0&p\end{pmatrix}}m

induces an endomorphism of Γ\mathcal{M}^{\Gamma}.

Definition 2.3.

Let pp be a prime number.

  1. (1)

    We define the Hecke operator TpT_{p} at pp on S(X)S_{\bullet}(X)\otimes\mathcal{M} by

    Tp(σP)\displaystyle T_{p}(\sigma\otimes P) :=(p001)σ(p001)P+j=0p1(1j0p)σ(1j0p)P\displaystyle:=\begin{pmatrix}p&0\\ 0&1\end{pmatrix}\sigma\otimes\begin{pmatrix}p&0\\ 0&1\end{pmatrix}P+\sum_{j=0}^{p-1}\begin{pmatrix}1&j\\ 0&p\end{pmatrix}\sigma\otimes\begin{pmatrix}1&j\\ 0&p\end{pmatrix}P

    for any simplex σS(X)\sigma\in S_{\bullet}(X) and element PP\in\mathcal{M}. The operator TpT_{p} induces operators on 𝒮(X)\mathcal{MS}(X)\otimes\mathcal{M} and H(Γ\X,)H_{\bullet}(\Gamma\backslash X,\mathcal{M}), etc., also written as TpT_{p}.

  2. (2)

    We define the Hecke operator TpT_{p}^{\prime} at pp on Hom(S(X),)\operatorname{Hom}_{\mathbb{Z}}(S_{\bullet}(X),\mathcal{M}) by

    (ϕ|Tp)(σ):=(p001)~(ϕ((p001)σ))+j=0p1(1j0p)~(ϕ((1j0p)σ))\displaystyle(\phi|T_{p}^{\prime})(\sigma):=\widetilde{\begin{pmatrix}p&0\\ 0&1\end{pmatrix}}\left(\phi(\begin{pmatrix}p&0\\ 0&1\end{pmatrix}\sigma)\right)+\sum_{j=0}^{p-1}\widetilde{\begin{pmatrix}1&j\\ 0&p\end{pmatrix}}\left(\phi(\begin{pmatrix}1&j\\ 0&p\end{pmatrix}\sigma)\right)

    for any homomorphism ϕHom(S(X),n)\phi\in\operatorname{Hom}_{\mathbb{Z}}(S_{\bullet}(X),\mathcal{M}_{n}^{\flat}) and simplex σS(X)\sigma\in S_{\bullet}(X). The operator TpT_{p}^{\prime} induces operators on Hom(𝒮(X),)\operatorname{Hom}_{\mathbb{Z}}(\mathcal{MS}(X),\mathcal{M}) and H(Γ\X,)H^{\bullet}(\Gamma\backslash X,\mathcal{M}), etc., also written as TpT_{p}^{\prime}.

For later use, we also define auxiliary operators UpU_{p} and VpV_{p} on S(X)S_{\bullet}(X)\otimes\mathcal{M} by

Vp(σP)\displaystyle V_{p}(\sigma\otimes P) :=(p001)σ(p001)P,\displaystyle:=\begin{pmatrix}p&0\\ 0&1\end{pmatrix}\sigma\otimes\begin{pmatrix}p&0\\ 0&1\end{pmatrix}P,
Up(σP)\displaystyle U_{p}(\sigma\otimes P) :=j=0p1(1j0p)σ(1j0p)P,\displaystyle:=\sum_{j=0}^{p-1}\begin{pmatrix}1&j\\ 0&p\end{pmatrix}\sigma\otimes\begin{pmatrix}1&j\\ 0&p\end{pmatrix}P,

so that Tp:=Vp+UpT_{p}:=V_{p}+U_{p}.

Lemma 2.4.

The composite VpUpV_{p}U_{p} acts on (S(X)n)Γ(S_{\bullet}(X)\otimes\mathcal{M}_{n})_{\Gamma}, and we have VpUp=pn+1V_{p}U_{p}=p^{n+1} as operators on (S(X)n)Γ(S_{\bullet}(X)\otimes\mathcal{M}_{n})_{\Gamma}.

Proof.

Since diagonal matrices act trivially on XX, we have

VpUp(σP)\displaystyle V_{p}U_{p}(\sigma\otimes P) =j=0p1(ppj0p)σ(ppj0p)P\displaystyle=\sum_{j=0}^{p-1}\begin{pmatrix}p&pj\\ 0&p\end{pmatrix}\sigma\otimes\begin{pmatrix}p&pj\\ 0&p\end{pmatrix}P
=pnj=0p1(1j01)(σP)\displaystyle=p^{n}\sum_{j=0}^{p-1}\begin{pmatrix}1&j\\ 0&1\end{pmatrix}\left(\sigma\otimes P\right)

for any simplex σS(X)\sigma\in S_{\bullet}(X) and polynomial PnP\in\mathcal{M}_{n}. Since (1j01)Γ\begin{pmatrix}1&j\\ 0&1\end{pmatrix}\in\Gamma for any integer jj, we obtain this lemma from this equality. ∎

2.5. Formal duality

Let X{,BS,BS}X\in\{\mathbb{H},\mathbb{H}^{\mathrm{BS}},\partial\mathbb{H}^{\mathrm{BS}}\}. As explained in §2.2, the homology and cohomology groups can be computed as

H(Γ\X,n)\displaystyle H_{\bullet}(\Gamma\backslash X,\mathcal{M}_{n}) H((S(X)n)Γ),\displaystyle\cong H_{\bullet}((S_{\bullet}(X)\otimes\mathcal{M}_{n})_{\Gamma}),
H(Γ\X,n)\displaystyle H^{\bullet}(\Gamma\backslash X,\mathcal{M}_{n}^{\flat}) H(Hom(S(X),n)Γ).\displaystyle\cong H_{\bullet}(\operatorname{Hom}_{\mathbb{Z}}(S_{\bullet}(X),\mathcal{M}_{n}^{\flat})^{\Gamma}).

The pairing ,:×\langle~{},~{}\rangle\colon\mathcal{M}^{\flat}\times\mathcal{M}\longrightarrow\mathbb{Z} induces a pairing

,:Hom(S(X),n)×S(X)n,\displaystyle\langle~{},~{}\rangle\colon\operatorname{Hom}_{\mathbb{Z}}(S_{\bullet}(X),\mathcal{M}_{n}^{\flat})\times S_{\bullet}(X)\otimes\mathcal{M}_{n}\longrightarrow\mathbb{Z},

which is computed as

ϕ,σP:=ϕ(σ),P.\big{\langle}\phi,\sigma\otimes P\big{\rangle}:=\langle\phi(\sigma),P\rangle.

Note that for any matrix γM2+()\gamma\in M_{2}^{+}(\mathbb{Z}), we have

γ~ϕ,σP=γ~ϕ(γσ),P=ϕ(γσ),γP=ϕ,γ(σP).\big{\langle}\widetilde{\gamma}\phi,\sigma\otimes P\big{\rangle}=\langle\widetilde{\gamma}\phi(\gamma\sigma),P\rangle=\langle\phi(\gamma\sigma),{\gamma}P\rangle=\big{\langle}\phi,{\gamma}(\sigma\otimes P)\big{\rangle}.

Therefore, we have

ϕ|Tp,σP=ϕ,Tp(σP).\big{\langle}\phi|T_{p}^{\prime},\sigma\otimes P\big{\rangle}=\big{\langle}\phi,T_{p}(\sigma\otimes P)\big{\rangle}.

In particular, we obtain a Hecke-equivariant pairing

,:H(Γ\X,n)×H(Γ\X,n),\langle~{},~{}\rangle\colon H^{\bullet}(\Gamma\backslash X,\mathcal{M}_{n}^{\flat})\times H_{\bullet}(\Gamma\backslash X,\mathcal{M}_{n})\longrightarrow\mathbb{Z},

which induces an isomorphism

H(Γ\X,n)/(torsion)Hom(H(Γ\X,n),).H^{\bullet}(\Gamma\backslash X,\mathcal{M}_{n}^{\flat})/\text{(torsion)}\overset{\sim}{\longrightarrow}\operatorname{Hom}_{\mathbb{Z}}(H_{\bullet}(\Gamma\backslash X,\mathcal{M}_{n}),\mathbb{Z}).

2.6. Eichler–Shimura homomorphism

Let Mn+2(Γ)M_{n+2}(\Gamma) denote the space of modular forms of weight n+2n+2 and level Γ=SL2()\Gamma=\mathrm{SL}_{2}(\mathbb{Z}). We define a homomorphism

r:Mn+2(Γ)Hom(𝒮(),n)\displaystyle r\colon M_{n+2}(\Gamma)\longrightarrow\operatorname{Hom}_{\mathbb{Z}}(\mathcal{MS}(\mathbb{H}),\mathcal{M}_{n}^{\flat}\otimes\mathbb{C})

by

r(f)({α,β}):=αβf(z)(X1zX2)n𝑑z.\displaystyle r(f)(\{\alpha,\beta\}):=\int_{\alpha}^{\beta}f(z)(X_{1}-zX_{2})^{n}\,dz.

for any modular form fMn+2(Γ)f\in M_{n+2}(\Gamma) and {α,β}𝒮()\{\alpha,\beta\}\in\mathcal{MS}(\mathbb{H}). It is well-known that

r(Mn+2(Γ))Hom(𝒮(),n)Γ,r(M_{n+2}(\Gamma))\subset\operatorname{Hom}_{\mathbb{Z}}(\mathcal{MS}(\mathbb{H}),\mathcal{M}_{n}^{\flat}\otimes\mathbb{C})^{\Gamma},

and the homomorphism rr induces an injective homomorphism (called Eichler–Shimura homomorphism)

r:Mn+2(Γ)\longhookrightarrowH1(Y,n)=H1(YBS,n).\displaystyle r\colon M_{n+2}(\Gamma)\longhookrightarrow H^{1}(Y,\mathcal{M}_{n}^{\flat})\otimes\mathbb{C}=H^{1}(Y^{\mathrm{BS}},\mathcal{M}_{n}^{\flat})\otimes\mathbb{C}.

See [Bel21, Section 5.3] for example.

Remark 2.5.

The definition of the Eichler–Shimura homomorphism shows that, for any element σ={α,β}P𝒮()n\sigma=\{\alpha,\beta\}\otimes P\in\mathcal{MS}(\mathbb{H})\otimes\mathcal{M}_{n} and modular form fMn+2(Γ)f\in M_{n+2}(\Gamma), the pairing r(f),σ\langle r(f),\sigma\rangle can be computed as

r(f),σ=αβf(z)P(z,1)𝑑z.\langle r(f),\sigma\rangle=\int_{\alpha}^{\beta}f(z)P(z,1)\,dz.
Remark 2.6.

For each prime number pp, the double coset operator Tp′′:=Γ(1p)ΓT_{p}^{\prime\prime}:=\Gamma\begin{pmatrix}1&\\ &p\end{pmatrix}\Gamma acts on the space Mn+2(Γ)M_{n+2}(\Gamma) of modular forms from the right by using the weight n+2n+2 slash operator |[]n+2|[~{}]_{n+2}. One can easily show that

r(f|[γ]n+2)=γ~r(f)r(f|[\gamma]_{n+2})=\widetilde{\gamma}\cdot r(f)

for any matrix γM2+()\gamma\in M_{2}^{+}(\mathbb{Z}). Hence the Eichler–Shimura homomorphism is Hecke-equivariant, that is, for all prime numbers pp, we have

r(f|Tp′′)=r(f)|Tp.r(f|T_{p}^{\prime\prime})=r(f)|T_{p}^{\prime}.

In other words, our Hecke operator TpT_{p}^{\prime} coincides with the usual one via the Eichler–Shimura homomorphism.

The following lemma is well-known (see [Bel21, Theorem 5.3.27] for example).

Lemma 2.7.

The Eichler–Shimura homomorphism induces a Hecke-equivariant isomorphism

Mn+2(Γ)/Sn+2(Γ)H1(YBS,n).M_{n+2}(\Gamma)/S_{n+2}(\Gamma)\stackrel{{\scriptstyle\sim}}{{\longrightarrow}}H^{1}(\partial Y^{\mathrm{BS}},\mathcal{M}_{n}^{\flat})\otimes\mathbb{C}.

Here Sn+2(Γ)S_{n+2}(\Gamma) denotes the space of cusp forms of weight n+2n+2 and level Γ\Gamma.

2.7. Definition of the Eisenstein class

In this subsection, we define the Eisenstein class Eisn\mathrm{Eis}_{n} and explain its basic properties.

We put i:=1i:=\sqrt{-1} and let σn+1(k)\sigma_{n+1}(k) denote the sum-of-positive-divisors function, namely, σn+1(k):=0<dkdn+1\sigma_{n+1}(k):=\sum_{0<d\mid k}d^{n+1}. Let

En+2(z):=1+2ζ(1n)k=1σn+1(k)e2πikzMn+2(Γ)\displaystyle E_{n+2}(z):=1+\frac{2}{\zeta(-1-n)}\sum_{k=1}^{\infty}\sigma_{n+1}(k)e^{2\pi ikz}\in M_{n+2}(\Gamma)

denote the normalized holomorphic Eisenstein series of weight n+2n+2.

Lemma 2.8.
  1. (1)

    For any element τ\tau\in\mathbb{H}, we have

    r(En+2),{τ,τ+1}e0=1.\langle r(E_{n+2}),\{\tau,\tau+1\}\otimes e_{0}\rangle=1.
  2. (2)

    For any prime number pp, we have

    r(En+2)|Tp=(1+pn+1)r(En+2).r(E_{n+2})|{T_{p}^{\prime}}=(1+p^{n+1})r(E_{n+2}).
Proof.

Claim (1) follows from the fact that the constant term of En+2E_{n+2} is 11, and claim (2) follows from the fact that En+2|Tp′′=(1+pn+1)En+2E_{n+2}|{T_{p}^{\prime\prime}}=(1+p^{n+1})E_{n+2} and the Eichler–Shimura homomorphism is Hecke-equivariant. ∎

Definition 2.9.

We define the Eisenstein class EisnH1(YBS,n)\mathrm{Eis}_{n}\in H^{1}(Y^{\mathrm{BS}},\mathcal{M}_{n}^{\flat})\otimes\mathbb{C} by

Eisn:=r(En+2).\mathrm{Eis}_{n}:=r\left(E_{n+2}\right).
Remark 2.10.

The method of defining the Eisenstein class Eisn\mathrm{Eis}_{n} in this paper differs from the method in Harder’s book [Har, §3.3.6 (3.74)]. However, thanks to Lemma 2.8, they coincide.

2.8. Main theorem

Proposition 2.11.

The Eisenstein class Eisn\mathrm{Eis}_{n} is rational, that is, EisnH1(YBS,n)\mathrm{Eis}_{n}\in H^{1}(Y^{\mathrm{BS}},\mathcal{M}_{n}^{\flat})\otimes\mathbb{Q}.

This proposition is proved in Corollary 4.18.

Definition 2.12.

For any Γ\Gamma-module \mathcal{M}, we define

Hint1(YBS,):=im(H1(YBS,)H1(YBS,)).H^{1}_{\mathrm{int}}(Y^{\mathrm{BS}},\mathcal{M}):=\operatorname{im}\left(H^{1}(Y^{\mathrm{BS}},\mathcal{M})\longrightarrow H^{1}(Y^{\mathrm{BS}},\mathcal{M})\otimes\mathbb{Q}\right).

Thanks to Proposition 2.11, we define the denominator Δ(Eisn)\Delta(\mathrm{Eis}_{n}) of the Eisenstein class Eisn\mathrm{Eis}_{n} with respect to the integral structure Hint1(YBS,n)H^{1}_{\mathrm{int}}(Y^{\mathrm{BS}},\mathcal{M}^{\flat}_{n}) by

Δ(Eisn):=min{Δ>0ΔEisnHint1(YBS,n)}.\Delta(\mathrm{Eis}_{n}):=\min\{\Delta\in\mathbb{Z}_{>0}\mid\Delta\mathrm{Eis}_{n}\in H^{1}_{\mathrm{int}}(Y^{\mathrm{BS}},\mathcal{M}_{n}^{\flat})\}.

Then the following is the main theorem which we want to prove in the present paper.

Theorem 2.13 ([Har, Theorem 5.1.2]).

The denominator Δ(Eisn)\Delta(\mathrm{Eis}_{n}) of the Eisenstein class Eisn\mathrm{Eis}_{n} is equal to the numerator of the special value ζ(1n)\zeta(-1-n) of the Riemann zeta function.

Remark 2.14.
  • (1)

    Since H1(YBS,n)=H1(YBS,n)H^{1}(Y^{\mathrm{BS}},\mathcal{M}_{n}^{\flat})\otimes\mathbb{Q}=H^{1}(Y^{\mathrm{BS}},\mathcal{M}_{n})\otimes\mathbb{Q}, we have another integral structure Hint1(YBS,n)H^{1}_{\mathrm{int}}(Y^{\mathrm{BS}},\mathcal{M}_{n}), and one can consider another denominator Δ(Eisn)\Delta^{\prime}(\mathrm{Eis}_{n}) of the Eisenstein class Eisn\mathrm{Eis}_{n}:

    Δ(Eisn):=min{Δ>0ΔEisnHint1(YBS,n)}.\Delta^{\prime}(\mathrm{Eis}_{n}):=\min\{\Delta\in\mathbb{Z}_{>0}\mid\Delta\mathrm{Eis}_{n}\in H^{1}_{\mathrm{int}}(Y^{\mathrm{BS}},\mathcal{M}_{n})\}.

    However, we show in Lemma 6.1 that Δ(Eisn)=Δ(Eisn)\Delta(\mathrm{Eis}_{n})=\Delta^{\prime}(\mathrm{Eis}_{n}).

  • (2)

    By using the qq-expansion at the cusp ii\infty, one can regard Mn+2(Γ)M_{n+2}(\Gamma) as a submodule of [[q]]\mathbb{C}[[q]], and we obtain the de Rham rational structure of Mn+2(Γ)M_{n+2}(\Gamma) by Mn+2(Γ)[[q]]M_{n+2}(\Gamma)\cap\mathbb{Q}[[q]]. The rationality of the critical values of the LL-function associated with a cusp form is obtained by studying the gap between the de Rham and Betti rational structures via the Eichler–Shimura homomorphism rr. However, Proposition 2.11 shows that the Eisenstein parts of the two rational structures coincide. Moreover, Theorem 2.13 says that the Eisenstein parts of the two integral structures Mn+2(Γ)[[q]]M_{n+2}(\Gamma)\cap\mathbb{Z}[[q]] and Hint1(YBS,n)H^{1}_{\mathrm{int}}(Y^{\mathrm{BS}},\mathcal{M}_{n}^{\flat}) coincide, namely,

    r(En+2(z)[[q]])=EisnHint1(YBS,n)r(\mathbb{Q}E_{n+2}(z)\cap\mathbb{Z}[[q]])=\mathbb{Q}\mathrm{Eis}_{n}\cap H^{1}_{\mathrm{int}}(Y^{\mathrm{BS}},\mathcal{M}_{n}^{\flat})

    since 22 is a regular prime. Cf. [Har21, §1.1].

3. Construction of the cycle Tpm(Cν(τ))~\widetilde{T_{p}^{m}(C_{\nu}(\tau))}

Fix a prime number pp. In this section, we construct a special homology cycle

Tpm(Cν(τ))~H1(YBS,n(p))\widetilde{T_{p}^{m}(C_{\nu}(\tau))}\in H_{1}(Y^{\mathrm{BS}},\mathcal{M}_{n}\otimes\mathbb{Z}_{(p)})

that is used to compute the pp-part of the denominator Δ(Eisn)\Delta(\mathrm{Eis}_{n}) of the Eisenstein class Eisn\mathrm{Eis}_{n}.

For any integer 1νn11\leq\nu\leq n-1 and element τBS\tau\in\mathbb{H}^{\mathrm{BS}}, we set

Cν(τ):={1τ,τ}eν𝒮(BS)n,C_{\nu}(\tau):=\left\{-\frac{1}{\tau},\tau\right\}\otimes e_{\nu}\in\mathcal{MS}(\mathbb{H}^{\mathrm{BS}})\otimes\mathcal{M}_{n},

where recall that eν=X1νX2nνe_{\nu}=X_{1}^{\nu}X_{2}^{n-\nu}.

3.1. Computation of Tpm(Cν(τ))T_{p}^{m}(C_{\nu}(\tau))

Recall also the operators

Vp(σP)\displaystyle V_{p}(\sigma\otimes P) :=(p001)σ(p001)P,\displaystyle:=\begin{pmatrix}p&0\\ 0&1\end{pmatrix}\sigma\otimes\begin{pmatrix}p&0\\ 0&1\end{pmatrix}P,
Up(σP)\displaystyle U_{p}(\sigma\otimes P) :=j=0p1(1j0p)σ(1j0p)P\displaystyle:=\sum_{j=0}^{p-1}\begin{pmatrix}1&j\\ 0&p\end{pmatrix}\sigma\otimes\begin{pmatrix}1&j\\ 0&p\end{pmatrix}P

on S(BS)nS_{\bullet}(\mathbb{H}^{\mathrm{BS}})\otimes\mathcal{M}_{n}. We have Tp=Vp+UpT_{p}=V_{p}+U_{p}. For each integer m0m\geq 0, set

Wm:=k=0mUpkVpmk.\displaystyle W_{m}:=\sum_{k=0}^{m}U_{p}^{k}V_{p}^{m-k}.

Note that W0W_{0} is the identity map. For any (commutative) ring RR and cycle CS(BS)(nR)C\in S_{\bullet}(\mathbb{H}^{\mathrm{BS}})\otimes(\mathcal{M}_{n}\otimes R), we denote by [C][C] the image of CC in (S(BS)(nR))Γ(S_{\bullet}(\mathbb{H}^{\mathrm{BS}})\otimes(\mathcal{M}_{n}\otimes R))_{\Gamma}.

Lemma 3.1.

Let m1m\geq 1 be an integer and CS(BS)nC\in S_{\bullet}(\mathbb{H}^{\mathrm{BS}})\otimes\mathcal{M}_{n}.

  • (1)

    We have Tp([Wm(C)])=[Wm+1(C)]+pn+1[Wm1(C)]T_{p}([W_{m}(C)])=[W_{m+1}(C)]+p^{n+1}[W_{m-1}(C)] and Tp([W0(C)])=[W1(C)]T_{p}([W_{0}(C)])=[W_{1}(C)].

  • (2)

    We have

Tpm([C])=A=0m/2C(mA,A)p(n+1)A[Wm2A(C)],\displaystyle T_{p}^{m}([C])=\sum_{A=0}^{\lfloor m/2\rfloor}C(m-A,A)p^{(n+1)A}[W_{m-2A}(C)],

where m/2\lfloor m/2\rfloor is the greatest integer less than or equal to m/2m/2 and

C(A,B):=(1BA+1)(A+BB)=(A+BB)(A+BB1).C(A,B):=\left(1-\frac{B}{A+1}\right)\binom{A+B}{B}=\binom{A+B}{B}-\binom{A+B}{B-1}\in\mathbb{Z}.

Here we assume (ab)=0\binom{a}{b}=0 if b<0b<0.

Proof.

Claim (1) follows from the fact that VpUp=pn+1V_{p}U_{p}=p^{n+1} proved in Lemma 2.4. Let us prove claim (2). For notational simplicity, we put

wm:=[Wm(C)].w_{m}:=[W_{m}(C)].

Then claim (1) shows that we can write

Tpm([C])=k=0mak(m)wmkT_{p}^{m}([C])=\sum_{k=0}^{m}a_{k}^{(m)}w_{m-k}

with ak(m)a_{k}^{(m)}\in\mathbb{Z} such that

a0(m)\displaystyle a_{0}^{(m)} =a0(m1)==a0(1)=a0(0)=1,\displaystyle=a_{0}^{(m-1)}=\cdots=a_{0}^{(1)}=a_{0}^{(0)}=1,
a1(m)\displaystyle a_{1}^{(m)} =a1(m1)==a1(1)=0,\displaystyle=a_{1}^{(m-1)}=\cdots=a_{1}^{(1)}=0,
ak(m)\displaystyle a_{k}^{(m)} =ak(m1)+pn+1ak2(m1),(2km1)\displaystyle=a_{k}^{(m-1)}+p^{n+1}a_{k-2}^{(m-1)},\,\,\,\quad(2\leq k\leq m-1)
am(m)\displaystyle a_{m}^{(m)} =pn+1am2(m1).\displaystyle=p^{n+1}a_{m-2}^{(m-1)}.

Therefore, we have a2k+1(m)=0a_{2k+1}^{(m)}=0 for any integer 0k(m1)/20\leq k\leq(m-1)/2, and hence

Tpm([C])=k=0m/2a2k(m)wm2k.T_{p}^{m}([C])=\sum_{k=0}^{\lfloor m/2\rfloor}a_{2k}^{(m)}w_{m-2k}.

Let us show that a2k(m)=C(mk,k)pk(n+1)a_{2k}^{(m)}=C(m-k,k)p^{k(n+1)} for any integer 0km/20\leq k\leq m/2 by induction on mm. When m=1m=1 or k=0k=0, this claim is clear. If m>1m>1 and 1k(m1)/21\leq k\leq(m-1)/2, then the induction hypothesis shows that

a2k(m)\displaystyle a_{2k}^{(m)} =a2k(m1)+pn+1a2k2(m1)\displaystyle=a_{2k}^{(m-1)}+p^{n+1}a_{2k-2}^{(m-1)}
=C(m1k,k)pk(n+1)+C(mk,k+1)pk(n+1)\displaystyle=C(m-1-k,k)p^{k(n+1)}+C(m-k,k+1)p^{k(n+1)}
=C(mk,k)pk(n+1).\displaystyle=C(m-k,k)p^{k(n+1)}.

Moreover, if mm is even, i.e., m=2tm=2t, then we have

am(m)\displaystyle a_{m}^{(m)} =pn+1am2(m1)=C(t,t1)pt(n+1)=1t+1(2t)!(t!)(t!)pt(n+1)=C(t,t)pt(n+1).\displaystyle=p^{n+1}a^{(m-1)}_{m-2}=C(t,t-1)p^{t(n+1)}=\frac{1}{t+1}\frac{(2t)!}{(t!)(t!)}p^{t(n+1)}=C(t,t)p^{t(n+1)}.

By definition, we have

Wm2A(Cν(τ))\displaystyle W_{m-2A}(C_{\nu}(\tau)) =k=0m2AUpkVpm2Ak(Cν(τ))\displaystyle=\sum_{k=0}^{m-2A}U_{p}^{k}V_{p}^{m-2A-k}(C_{\nu}(\tau))
=k=0m2AUpk(pm2Ak001)({1τ,τ}eν)\displaystyle=\sum_{k=0}^{m-2A}U_{p}^{k}\begin{pmatrix}p^{m-2A-k}&0\\ 0&1\end{pmatrix}\left(\left\{-\frac{1}{\tau},\tau\right\}\otimes e_{\nu}\right)
=k=0m2Ap(nν)(m2Ak)Upk({pm2Akτ,pm2Akτ}eν).\displaystyle=\sum_{k=0}^{m-2A}p^{(n-\nu)(m-2A-k)}U_{p}^{k}\left(\left\{-\frac{p^{m-2A-k}}{\tau},p^{m-2A-k}\tau\right\}\otimes e_{\nu}\right).
Definition 3.2.

Take elements τ0,τ1BS\tau_{0},\tau_{1}\in\mathbb{H}^{\mathrm{BS}}. For any integers ν\nu, jj, and kk satisfying 1νn11\leq\nu\leq n-1 and 0jpk10\leq j\leq p^{k}-1, we define

Cν,k,j(τ0,τ1)\displaystyle C_{\nu,k,j}(\tau_{0},\tau_{1}) :=(1j0pk){1τ0,τ1}(1j0pk)eν\displaystyle:=\begin{pmatrix}1&j\\ 0&p^{k}\end{pmatrix}\left\{-\frac{1}{\tau_{0}},\tau_{1}\right\}\otimes\begin{pmatrix}1&j\\ 0&p^{k}\end{pmatrix}e_{\nu}
={1τ0+jpk,τ1+jpk}(1j0pk)eν𝒮(BS)n.\displaystyle=\left\{\frac{-\frac{1}{\tau_{0}}+j}{p^{k}},\frac{\tau_{1}+j}{p^{k}}\right\}\otimes\begin{pmatrix}1&j\\ 0&p^{k}\end{pmatrix}e_{\nu}\in\mathcal{M}\mathcal{S}(\mathbb{H}^{\mathrm{BS}})\otimes\mathcal{M}_{n}.

Note that we have

Upk({1τ0,τ1}eν)=j=0pk1Cν,k,j(τ0,τ1).U_{p}^{k}\left(\left\{-\frac{1}{\tau_{0}},\tau_{1}\right\}\otimes e_{\nu}\right)=\sum_{j=0}^{p^{k}-1}C_{\nu,k,j}(\tau_{0},\tau_{1}).

To sum up, we obtain the following corollary.

Corollary 3.3.

We have

Tpm([Cν(τ)])\displaystyle T_{p}^{m}([C_{\nu}(\tau)])
=A=0m/2C(mA,A)p(n+1)Ak=0m2Ap(nν)(m2Ak)j=0pk1[Cν,k,j(τpm2Ak,pm2Akτ)].\displaystyle=\sum_{A=0}^{\lfloor m/2\rfloor}C(m-A,A)p^{(n+1)A}\sum_{k=0}^{m-2A}p^{(n-\nu)(m-2A-k)}\sum_{j=0}^{p^{k}-1}\left[C_{\nu,k,j}\left(\frac{\tau}{p^{m-2A-k}},p^{m-2A-k}\tau\right)\right].

3.2. Computation of the boundary Cν,k,j(τ0,τ1)\partial C_{\nu,k,j}\left(\tau_{0},\tau_{1}\right)

Next, we compute the boundary Cν,k,j(τ0,τ1)\partial C_{\nu,k,j}\left(\tau_{0},\tau_{1}\right).

Definition 3.4.
  1. (1)

    For any integers jj and NN with pjp\nmid j and N>0N>0, we denote by dN(j)d_{N}(j) and bN(j)b_{N}(j) the integers uniquely determined by

    1dN(j)<pN and jdN(j)pNbN(j)=1.1\leq d_{N}(j)<p^{N}\,\,\,\textrm{ and }\,\,\,jd_{N}(j)-p^{N}b_{N}(j)=1.

    We also put d0(j):=0d_{0}(j):=0 and b0(j):=1b_{0}(j):=-1 for any integer jj.

  2. (2)

    For any integers kk and jj, we set

    lk(j):=min{ordp(j),k}.l_{k}(j):=\min\{\mathrm{ord}_{p}(j),k\}.

    Note that lk(0)=kl_{k}(0)=k. We also put j:=j/plk(j)j^{\prime}:=j/p^{l_{k}(j)}.

In the following, for integers jj and kk with 0jpk10\leq j\leq p^{k}-1, we often write as

(3.1) l:=lk(j),j:=j/plk(j),d:=dklk(j)(j),b:=bklk(j)(j)\displaystyle\begin{split}l&:=l_{k}(j),\\ j^{\prime}&:=j/p^{l_{k}(j)},\\ d&:=d_{k-l_{k}(j)}(j^{\prime}),\\ b&:=b_{k-l_{k}(j)}(j^{\prime})\end{split}

for simplicity.

Definition 3.5.

For any integers ν\nu, jj, and kk with 1νn11\leq\nu\leq n-1 and 0jpk10\leq j\leq p^{k}-1, we define homogeneous polynomials Eν,k,j(1)E_{\nu,k,j}^{(1)} and Eν,k,j(0)E_{\nu,k,j}^{(0)} in n\mathcal{M}_{n} by

Eν,k,j(1)(X1,X2)\displaystyle E_{\nu,k,j}^{(1)}(X_{1},X_{2}) :=(pkX1jX2)νX2nν,\displaystyle:=(p^{k}X_{1}-jX_{2})^{\nu}X_{2}^{n-\nu},
Eν,k,j(0)(X1,X2)\displaystyle E_{\nu,k,j}^{(0)}(X_{1},X_{2}) :=(1)ν+1(plX2)ν(pklX1+dX2)nν.\displaystyle:=(-1)^{\nu+1}(p^{l}X_{2})^{\nu}(p^{k-l}X_{1}+dX_{2})^{n-\nu}.
Lemma 3.6.

We have

[Cν,k,j(τ0,τ1)]=[{τ1+jpk}Eν,k,j(1)+{plτ0dpkl}Eν,k,j(0)]\displaystyle\partial[C_{\nu,k,j}(\tau_{0},\tau_{1})]=\left[\left\{\frac{\tau_{1}+j}{p^{k}}\right\}\otimes E_{\nu,k,j}^{(1)}+\left\{\frac{p^{l}\tau_{0}-d}{p^{k-l}}\right\}\otimes E_{\nu,k,j}^{(0)}\right]

in (S0(BS)n)Γ(S_{0}(\mathbb{H}^{\mathrm{BS}})\otimes\mathcal{M}_{n})_{\Gamma}.

Proof.

By definition, we have

[Cν,k,j(τ0,τ1)]=[(1j0pk){τ1}(1j0pk)eν(1j0pk){1τ0}(1j0pk)eν].\displaystyle\partial[C_{\nu,k,j}(\tau_{0},\tau_{1})]=\left[\begin{pmatrix}1&j\\ 0&p^{k}\end{pmatrix}\left\{\tau_{1}\right\}\otimes\begin{pmatrix}1&j\\ 0&p^{k}\end{pmatrix}e_{\nu}-\begin{pmatrix}1&j\\ 0&p^{k}\end{pmatrix}\left\{-\frac{1}{\tau_{0}}\right\}\otimes\begin{pmatrix}1&j\\ 0&p^{k}\end{pmatrix}e_{\nu}\right].

The definition of Eν,k,j(1)E_{\nu,k,j}^{(1)} shows that

(1j0pk){τ1}(1j0pk)eν={τ1+jpk}Eν,k,j(1).\begin{pmatrix}1&j\\ 0&p^{k}\end{pmatrix}\left\{\tau_{1}\right\}\otimes\begin{pmatrix}1&j\\ 0&p^{k}\end{pmatrix}e_{\nu}=\left\{\frac{\tau_{1}+j}{p^{k}}\right\}\otimes E_{\nu,k,j}^{(1)}.

Moreover, we have

(1j0pk){1τ0}(1j0pk)eν\displaystyle\begin{pmatrix}1&j\\ 0&p^{k}\end{pmatrix}\left\{-\frac{1}{\tau_{0}}\right\}\otimes\begin{pmatrix}1&j\\ 0&p^{k}\end{pmatrix}e_{\nu} =(1j0pk)(0110){τ0}(1j0pk)eν\displaystyle=\begin{pmatrix}1&j\\ 0&p^{k}\end{pmatrix}\begin{pmatrix}0&-1\\ 1&0\end{pmatrix}\left\{\tau_{0}\right\}\otimes\begin{pmatrix}1&j\\ 0&p^{k}\end{pmatrix}e_{\nu}
=(jbpkld)(pld0pkl){τ0}(1j0pk)eν.\displaystyle=\begin{pmatrix}j^{\prime}&b\\ p^{k-l}&d\end{pmatrix}\begin{pmatrix}p^{l}&-d\\ 0&p^{k-l}\end{pmatrix}\left\{\tau_{0}\right\}\otimes\begin{pmatrix}1&j\\ 0&p^{k}\end{pmatrix}e_{\nu}.

Since

[(jbpkld)(pld0pkl){τ0}(1j0pk)eν]=[(pld0pkl){τ0}(dbpklj)(1j0pk)eν]\displaystyle\left[\begin{pmatrix}j^{\prime}&b\\ p^{k-l}&d\end{pmatrix}\begin{pmatrix}p^{l}&-d\\ 0&p^{k-l}\end{pmatrix}\left\{\tau_{0}\right\}\otimes\begin{pmatrix}1&j\\ 0&p^{k}\end{pmatrix}e_{\nu}\right]=\left[\begin{pmatrix}p^{l}&-d\\ 0&p^{k-l}\end{pmatrix}\left\{\tau_{0}\right\}\otimes\begin{pmatrix}d&-b\\ -p^{k-l}&j^{\prime}\end{pmatrix}\begin{pmatrix}1&j\\ 0&p^{k}\end{pmatrix}e_{\nu}\right]

and (dbpklj)(1j0pk)eν=Eν,k,j(0)\begin{pmatrix}d&-b\\ -p^{k-l}&j^{\prime}\end{pmatrix}\begin{pmatrix}1&j\\ 0&p^{k}\end{pmatrix}e_{\nu}=-E_{\nu,k,j}^{(0)}, we obtain

[(1j0pk){1τ0}(1j0pk)eν]=[{plτ0dpkl}Eν,k,j(0)].\displaystyle\left[\begin{pmatrix}1&j\\ 0&p^{k}\end{pmatrix}\left\{-\frac{1}{\tau_{0}}\right\}\otimes\begin{pmatrix}1&j\\ 0&p^{k}\end{pmatrix}e_{\nu}\right]=-\left[\left\{\frac{p^{l}\tau_{0}-d}{p^{k-l}}\right\}\otimes E_{\nu,k,j}^{(0)}\right].

3.3. A cycle Tpm(Cν(τ))~\widetilde{T_{p}^{m}(C_{\nu}(\tau))} in 𝒮(BS)(n)\mathcal{M}\mathcal{S}(\mathbb{H}^{\mathrm{BS}})\otimes(\mathcal{M}_{n}\otimes\mathbb{Q})

In this subsection, we construct a cycle Tpm(Cν(τ))~\widetilde{T_{p}^{m}(C_{\nu}(\tau))} in 𝒮(BS)(n)\mathcal{M}\mathcal{S}(\mathbb{H}^{\mathrm{BS}})\otimes(\mathcal{M}_{n}\otimes\mathbb{Q}) which is a lift of Tpm(Cν(τ))T_{p}^{m}(C_{\nu}(\tau)) and is pp-adically integral for any sufficiently large integer mm.

3.3.1. Bernoulli polynomials

Since a key tool for constructing the cycle Tpm(Cν(τ))~\widetilde{T_{p}^{m}(C_{\nu}(\tau))} is the Bernoulli polynomials, we briefly recall the basic properties of the Bernoulli polynomials.

Let tt be a non-negative integer. We denote by Bt(x)B_{t}(x) the tt-th Bernoulli polynomial and by

Bt:=Bt(0)B_{t}:=B_{t}(0)

the tt-th Bernoulli number. For notational simplicity, we put

B~t(x):=Bt(x)Btt.\widetilde{B}_{t}(x):=\frac{B_{t}(x)-B_{t}}{t}.

In this paper, we use the following well-known facts without any notice:

Bt(x)=μ=0t(tμ)Btμxμ,\displaystyle B_{t}(x)=\sum_{\mu=0}^{t}\binom{t}{\mu}B_{t-\mu}x^{\mu},
xx+1Bt(x)𝑑x=xt,\displaystyle\int_{x}^{x+1}B_{t}(x)\,dx=x^{t},
axBt(x)𝑑x=Bt+1(x)Bt+1(a)t+1,\displaystyle\int_{a}^{x}B_{t}(x)\,dx=\frac{B_{t+1}(x)-B_{t+1}(a)}{t+1},
ordp(Bt)1.\displaystyle\mathrm{ord}_{p}(B_{t})\geq-1.

The last fact is called the von Staudt–Clausen theorem. We note that the second and third facts imply that

j=0x1jt=B~t+1(x).\sum_{j=0}^{x-1}j^{t}=\widetilde{B}_{t+1}(x).

3.3.2. PP^{\dagger} and PP^{\ddagger}

Set

n,(p):=μ=0n(p)X1μX2nμ and n,(p):=μ=0n1(p)X1μX2nμ,\mathcal{M}_{n,(p)}:=\bigoplus_{\mu=0}^{n}\mathbb{Z}_{(p)}X_{1}^{\mu}X_{2}^{n-\mu}\,\,\,\textrm{ and }\,\,\,\mathcal{M}_{n,(p)}^{\circ}:=\bigoplus_{\mu=0}^{n-1}\mathbb{Z}_{(p)}X_{1}^{\mu}X_{2}^{n-\mu},

and let

:n,(p)n,(p);PP{\dagger}\colon\mathcal{M}_{n,(p)}^{\circ}\otimes\mathbb{Q}\rightarrow\mathcal{M}_{n,(p)}\otimes\mathbb{Q};P\mapsto P^{{\dagger}}

be the \mathbb{Q}-linear map defined by

(X1μX2nμ):=X2nBμ+1(X1/X2)Bμ+1μ+1=X2nB~μ+1(X1/X2).(X_{1}^{\mu}X_{2}^{n-\mu})^{{\dagger}}:=X_{2}^{n}\frac{B_{\mu+1}(X_{1}/X_{2})-B_{\mu+1}}{\mu+1}=X_{2}^{n}\widetilde{B}_{\mu+1}(X_{1}/X_{2}).

For any integer μ{0,,n1}\mu\in\{0,\ldots,n-1\}, we have 1+ordp(μ+1)n1+\mathrm{ord}_{p}(\mu+1)\leq n, and hence pnX2nB~μ+1(X1/X2)n,(p)p^{n}X_{2}^{n}\widetilde{B}_{\mu+1}(X_{1}/X_{2})\in\mathcal{M}_{n,(p)}. This fact shows that

(n,(p))1pnn,(p).{\dagger}(\mathcal{M}_{n,(p)}^{\circ})\subset\frac{1}{p^{n}}\mathcal{M}_{n,(p)}.

Similarly, let

:n,(p)n;PP{\ddagger}\colon\mathcal{M}_{n,(p)}^{\circ}\otimes\mathbb{Q}\rightarrow\mathcal{M}_{n}\otimes\mathbb{Q};P\mapsto P^{{\ddagger}}

be the \mathbb{Q}-linear map defined by

(X1μX2nμ):=X2n(X1/X2)μ+1Bμ+1μ+1.(X_{1}^{\mu}X_{2}^{n-\mu})^{{\ddagger}}:=X_{2}^{n}\frac{(X_{1}/X_{2})^{\mu+1}-B_{\mu+1}}{\mu+1}.

The following lemma follows from the definitions of PP^{\dagger} and PP^{\ddagger}.

Lemma 3.7.

For any polynomial P(X1,X2)nP(X_{1},X_{2})\in\mathcal{M}_{n}^{\circ}\otimes\mathbb{Q}, we have

P(X1+X2,X2)P(X1,X2)\displaystyle P^{{\dagger}}(X_{1}+X_{2},X_{2})-P^{{\dagger}}(X_{1},X_{2}) =P(X1,X2),\displaystyle=P(X_{1},X_{2}),
xx+1P(z,1)𝑑z\displaystyle\int_{x}^{x+1}P^{{\dagger}}(z,1)\,dz =P(x,1),\displaystyle=P^{{\ddagger}}(x,1),
x1x2P(z,1)𝑑z\displaystyle\int_{x_{1}}^{x_{2}}P(z,1)\,dz =P(x2,1)P(x1,1).\displaystyle=P^{{\ddagger}}(x_{2},1)-P^{{\ddagger}}(x_{1},1).

3.3.3. Definitions of polynomials Pν,k,j(1)P^{(1)}_{\nu,k,j} and Pν,k,j(0)P^{(0)}_{\nu,k,j}

Definition 3.8.

For any integers ν\nu, jj, and kk with 1νn11\leq\nu\leq n-1 and 0jpk10\leq j\leq p^{k}-1, we define polynomials Pν,k,j(1)P_{\nu,k,j}^{(1)} and Pν,k,j(0)P_{\nu,k,j}^{(0)} in n\mathcal{M}_{n}\otimes\mathbb{Q} by

Pν,k,j(1)\displaystyle P_{\nu,k,j}^{(1)} :=Eν,k,j(1),\displaystyle:=E_{\nu,k,j}^{(1){\dagger}},
Pν,k,j(0)\displaystyle P_{\nu,k,j}^{(0)} :=Eν,k,j(0).\displaystyle:=E_{\nu,k,j}^{(0){\dagger}}.
Lemma 3.9.

For each i{0,1}i\in\{0,1\}, we have Pν,k,j(i)pmin{0,kn}n,(p)P_{\nu,k,j}^{(i)}\in p^{\min\{0,k-n\}}\mathcal{M}_{n,(p)}.

Proof.

This lemma follows from the facts that (n,(p))pnn,(p){\dagger}(\mathcal{M}_{n,(p)}^{\circ})\subset p^{-n}\mathcal{M}_{n,(p)} and (X2n)=X1X2n1(X_{2}^{n})^{\dagger}=X_{1}X_{2}^{n-1} and that the coefficient of X1μX2nμX_{1}^{\mu}X_{2}^{n-\mu} in Eν,k,j(i)E_{\nu,k,j}^{(i)} is divided by pkp^{k} if μ1\mu\geq 1. ∎

3.3.4. Definition of the cycle Tpm(Cν(τ))~\widetilde{T_{p}^{m}(C_{\nu}(\tau))}

Now we can define C~ν,k,j(τ0,τ1)\widetilde{C}_{\nu,k,j}(\tau_{0},\tau_{1}) and Tpm(Cν(τ))~\widetilde{T_{p}^{m}(C_{\nu}(\tau))}.

Definition 3.10.

Let τ0,τ1BS\tau_{0},\tau_{1}\in\mathbb{H}^{\mathrm{BS}}. For any integers ν\nu, kk, and jj with 1νn11\leq\nu\leq n-1 and 0jpk10\leq j\leq p^{k}-1, we define an element C~ν,k,j(τ0,τ1)𝒮(BS)(n)\widetilde{C}_{\nu,k,j}(\tau_{0},\tau_{1})\in\mathcal{M}\mathcal{S}(\mathbb{H}^{\mathrm{BS}})\otimes(\mathcal{M}_{n}\otimes\mathbb{Q}) by

C~ν,k,j(τ0,τ1):=Cν,k,j(τ0,τ1){τ1+jpk,τ1+jpk+1}Pν,k,j(1){plτ0dpkl,plτ0dpkl+1}Pν,k,j(0).\displaystyle\widetilde{C}_{\nu,k,j}(\tau_{0},\tau_{1}):=C_{\nu,k,j}(\tau_{0},\tau_{1})-\left\{\frac{\tau_{1}+j}{p^{k}},\frac{\tau_{1}+j}{p^{k}}+1\right\}\otimes P_{\nu,k,j}^{(1)}-\left\{\frac{p^{l}\tau_{0}-d}{p^{k-l}},\frac{p^{l}\tau_{0}-d}{p^{k-l}}+1\right\}\otimes P_{\nu,k,j}^{(0)}.

We also put

C~ν,k,jint(τ0,τ1):=pmax{0,nk}C~ν,k,j(τ0,τ1).\widetilde{C}_{\nu,k,j}^{\rm int}(\tau_{0},\tau_{1}):=p^{\max\{0,n-k\}}\widetilde{C}_{\nu,k,j}(\tau_{0},\tau_{1}).

Note that C~ν,k,jint(τ0,τ1)𝒮(BS)n,(p)\widetilde{C}_{\nu,k,j}^{\rm int}(\tau_{0},\tau_{1})\in\mathcal{M}\mathcal{S}(\mathbb{H}^{\mathrm{BS}})\otimes\mathcal{M}_{n,(p)} by Lemma 3.9.

Lemma 3.11.

We have

[C~ν,k,jint(τ0,τ1)]=0 in (S0(BS)n,(p))Γ.\partial[\widetilde{C}_{\nu,k,j}^{\rm int}(\tau_{0},\tau_{1})]=0\,\,\,\textrm{ in }\,\,\,(S_{0}(\mathbb{H}^{\mathrm{BS}})\otimes\mathcal{M}_{n,(p)})_{\Gamma}.

In particular, [C~ν,k,jint(τ0,τ1)][\widetilde{C}_{\nu,k,j}^{\rm int}(\tau_{0},\tau_{1})] defines a homology class

[C~ν,k,jint(τ0,τ1)]H1(YBS,n,(p)).[\widetilde{C}_{\nu,k,j}^{\rm int}(\tau_{0},\tau_{1})]\in H_{1}(Y^{\mathrm{BS}},\mathcal{M}_{n,(p)}).
Proof.

This lemma follows from Definitions 3.5 and 3.8 and Lemmas 3.6 and 3.7. ∎

Lemma 3.12.

The homology class [C~ν,k,jint(τ0,τ1)][\widetilde{C}_{\nu,k,j}^{\rm int}(\tau_{0},\tau_{1})] is independent of the choices of τ0\tau_{0} and τ1\tau_{1}.

Proof.

Let τ0\tau_{0}^{\prime} and τ1\tau_{1}^{\prime} be another pair of points in BS\mathbb{H}^{\mathrm{BS}}. We will prove that

[C~ν,k,jint(τ0,τ1)]=[C~ν,k,jint(τ0,τ1)] in H1(YBS,n,(p)).[\widetilde{C}_{\nu,k,j}^{\rm int}(\tau_{0},\tau_{1})]=[\widetilde{C}_{\nu,k,j}^{\rm int}(\tau_{0}^{\prime},\tau_{1}^{\prime})]\,\,\,\text{ in }\,\,\,H_{1}(Y^{\mathrm{BS}},\mathcal{M}_{n,{(p)}}).

It suffices to construct an element hS2(BS)n,(p)h\in S_{2}(\mathbb{H}^{\mathrm{BS}})\otimes\mathcal{M}_{n,{(p)}} such that

[h]=[C~ν,k,jint(τ0,τ1)][C~ν,k,jint(τ0,τ1)] in (𝒮(BS)n,(p))Γ.\partial[h]={[\widetilde{C}_{\nu,k,j}^{\rm int}(\tau_{0},\tau_{1})]-[\widetilde{C}_{\nu,k,j}^{\rm int}(\tau_{0}^{\prime},\tau_{1}^{\prime})]}\,\,\,\text{ in }\,\,\,(\mathcal{MS}(\mathbb{H}^{\mathrm{BS}})\otimes\mathcal{M}_{n,{(p)}})_{\Gamma}.

For notational simplicity, set q:=pmax{0,nk}q:=p^{\max\{0,n-k\}}. First, since BS\mathbb{H}^{\mathrm{BS}} is simply connected, there exist elements h1,h2,h3S2(BS)h_{1},h_{2},h_{3}\in S_{2}(\mathbb{H}^{\mathrm{BS}}) such that

h1\displaystyle\partial h_{1} ={1τ0+jpk,τ1+jpk}{1τ0+jpk,τ1+jpk}{τ1+jpkτ1+jpk}+{1τ0+jpk,1τ0+jpk,},\displaystyle=\left\{\frac{-\frac{1}{\tau_{0}}+j}{p^{k}},\frac{\tau_{1}+j}{p^{k}}\right\}-\left\{\frac{-\frac{1}{\tau_{0}^{\prime}}+j}{p^{k}},\frac{\tau_{1}^{\prime}+j}{p^{k}}\right\}-\left\{\frac{\tau_{1}^{\prime}+j}{p^{k}}\frac{\tau_{1}+j}{p^{k}}\right\}+\left\{\frac{-\frac{1}{\tau_{0}^{\prime}}+j}{p^{k}},\frac{-\frac{1}{\tau_{0}}+j}{p^{k}},\right\},
h2\displaystyle\partial h_{2} ={τ1+jpk,τ1+jpk+1}{τ1+jpk,τ1+jpk+1}{τ1+jpk+1,τ1+jpk+1}+{τ1+jpk,τ1+jpk},\displaystyle=\left\{\frac{\tau_{1}+j}{p^{k}},\frac{\tau_{1}+j}{p^{k}}+1\right\}-\left\{\frac{\tau_{1}^{\prime}+j}{p^{k}},\frac{\tau_{1}^{\prime}+j}{p^{k}}+1\right\}-\left\{\frac{\tau_{1}^{\prime}+j}{p^{k}}+1,\frac{\tau_{1}+j}{p^{k}}+1\right\}+\left\{\frac{\tau_{1}^{\prime}+j}{p^{k}},\frac{\tau_{1}+j}{p^{k}}\right\},
h3\displaystyle\partial h_{3} ={plτ0dpkl,plτ0dpkl+1}{plτ0dpkl,plτ0dpkl+1}{plτ0dpkl+1,plτ0dpkl+1}+{plτ0dpkl,plτ0dpkl}.\displaystyle=\left\{\frac{p^{l}\tau_{0}-d}{p^{k-l}},\frac{p^{l}\tau_{0}-d}{p^{k-l}}+1\right\}-\left\{\frac{p^{l}\tau_{0}^{\prime}-d}{p^{k-l}},\frac{p^{l}\tau_{0}^{\prime}-d}{p^{k-l}}+1\right\}-\left\{\frac{p^{l}\tau_{0}^{\prime}-d}{p^{k-l}}+1,\frac{p^{l}\tau_{0}-d}{p^{k-l}}+1\right\}+\left\{\frac{p^{l}\tau_{0}^{\prime}-d}{p^{k-l}},\frac{p^{l}\tau_{0}-d}{p^{k-l}}\right\}.

Then we see that

h:=q(h1(1j0pk)eνh2Pν,k,j(1)h3Pν,k,j(0))h:={{q}\left(h_{1}\otimes\begin{pmatrix}1&j\\ 0&p^{k}\end{pmatrix}e_{\nu}-h_{2}\otimes P_{\nu,k,j}^{(1)}-h_{3}\otimes P_{\nu,k,j}^{(0)}\right)}

satisfies the desired property. Indeed, we have

h=q(C~ν,k,j(τ0,τ1)C~ν,k,j(τ0,τ1)B(1)B(0)),\displaystyle\partial h={{q}\left(\widetilde{C}_{\nu,k,j}(\tau_{0},\tau_{1})-\widetilde{C}_{\nu,k,j}(\tau_{0}^{\prime},\tau_{1}^{\prime})-B^{(1)}-B^{(0)}\right)},

where

B(1)\displaystyle B^{(1)} :={τ1+jpkτ1+jpk}(1j0pk)eν{τ1+jpk+1,τ1+jpk+1}Pν,k,j(1)+{τ1+jpk,τ1+jpk}Pν,k,j(1),\displaystyle:=\left\{\frac{\tau_{1}^{\prime}+j}{p^{k}}\frac{\tau_{1}+j}{p^{k}}\right\}\otimes\begin{pmatrix}1&j\\ 0&p^{k}\end{pmatrix}e_{\nu}-\left\{\frac{\tau_{1}^{\prime}+j}{p^{k}}+1,\frac{\tau_{1}+j}{p^{k}}+1\right\}\otimes P_{\nu,k,j}^{(1)}+\left\{\frac{\tau_{1}^{\prime}+j}{p^{k}},\frac{\tau_{1}+j}{p^{k}}\right\}\otimes P_{\nu,k,j}^{(1)},
B(0)\displaystyle B^{(0)} :={1τ0+jpk,1τ0+jpk,}(1j0pk)eν{plτ0dpkl+1,plτ0dpkl+1}Pν,k,j(0)+{plτ0dpkl,plτ0dpkl}Pν,k,j(0).\displaystyle:=-\left\{\frac{-\frac{1}{\tau_{0}^{\prime}}+j}{p^{k}},\frac{-\frac{1}{\tau_{0}}+j}{p^{k}},\right\}\otimes\begin{pmatrix}1&j\\ 0&p^{k}\end{pmatrix}e_{\nu}-\left\{\frac{p^{l}\tau_{0}^{\prime}-d}{p^{k-l}}+1,\frac{p^{l}\tau_{0}-d}{p^{k-l}}+1\right\}\otimes P_{\nu,k,j}^{(0)}+\left\{\frac{p^{l}\tau_{0}^{\prime}-d}{p^{k-l}},\frac{p^{l}\tau_{0}-d}{p^{k-l}}\right\}\otimes P_{\nu,k,j}^{(0)}.

Then the same computation as in the proof of Lemma 3.6 shows that

qB(1)\displaystyle{q}B^{(1)} {τ1+jpkτ1+jpk}q(Eν,k,j(1)(X1,X2)Pν,k,j(1)(X1+X2,X2)+Pν,k,j(1)(X1,X2))=0,\displaystyle\equiv\left\{\frac{\tau_{1}^{\prime}+j}{p^{k}}\frac{\tau_{1}+j}{p^{k}}\right\}\otimes{q}\left(E_{\nu,k,j}^{(1)}(X_{1},X_{2})-P_{\nu,k,j}^{(1)}(X_{1}+X_{2},X_{2})+P_{\nu,k,j}^{(1)}(X_{1},X_{2})\right)=0,
qB(0)\displaystyle{q}B^{(0)} {plτ0dpkl+1,plτ0dpkl+1}q(Eν,k,j(0)(X1,X2)Pν,k,j(0)(X1+X2,X2)+Pν,k,j(0)(X1,X2))=0,\displaystyle\equiv\left\{\frac{p^{l}\tau_{0}^{\prime}-d}{p^{k-l}}+1,\frac{p^{l}\tau_{0}-d}{p^{k-l}}+1\right\}\otimes{q}\left(E_{\nu,k,j}^{(0)}(X_{1},X_{2})-P_{\nu,k,j}^{(0)}(X_{1}+X_{2},X_{2})+P_{\nu,k,j}^{(0)}(X_{1},X_{2})\right)=0,

where \equiv means that it is an equality in the Γ\Gamma-coinvariant (𝒮(BS)n,(p))Γ(\mathcal{MS}(\mathbb{H}^{\mathrm{BS}})\otimes\mathcal{M}_{n,{(p)}})_{\Gamma}. Hence we obtain [h]=[C~ν,k,jint(τ0,τ1)][C~ν,k,jint(τ0,τ1)]\partial[h]=[\widetilde{C}_{\nu,k,j}^{\rm int}(\tau_{0},\tau_{1})]-[\widetilde{C}_{\nu,k,j}^{\rm int}(\tau_{0}^{\prime},\tau_{1}^{\prime})]. ∎

Definition 3.13.

For any integer m>0m>0 and element τBS\tau\in\mathbb{H}^{\mathrm{BS}}, we define a cycle Tpm(Cν(τ))~𝒮(BS)(n)\widetilde{T_{p}^{m}(C_{\nu}(\tau))}\in\mathcal{MS}(\mathbb{H}^{\mathrm{BS}})\otimes(\mathcal{M}_{n}\otimes\mathbb{Q}) by

Tpm(Cν(τ))~\displaystyle\widetilde{T_{p}^{m}(C_{\nu}(\tau))}
:=A=0m/2C(mA,A)p(n+1)Ak=0m2Ap(nν)(m2Ak)j=0pk1C~ν,k,j(τpm2Ak,pm2Akτ).\displaystyle:=\sum_{A=0}^{\lfloor m/2\rfloor}C(m-A,A)p^{(n+1)A}\sum_{k=0}^{m-2A}p^{(n-\nu)(m-2A-k)}\sum_{j=0}^{p^{k}-1}\widetilde{C}_{\nu,k,j}\left(\frac{\tau}{p^{m-2A-k}},p^{m-2A-k}\tau\right).
Definition 3.14.

For any integer ν{0,,n}\nu\in\{0,\ldots,n\}, we define a homology class

[Cν]H1(YBS,YBS,n)[C_{\nu}]\in H_{1}(Y^{\rm BS},\partial Y^{\rm BS},\mathcal{M}_{n})

to be the element represented by the cycle

Cν:={0,i}eν,C_{\nu}:=\{0,i\infty\}\otimes e_{\nu},

where iBSi\infty\in\partial\mathbb{H}^{\mathrm{BS}} is a point such that

i:=limt>0,tit.i\infty:=\lim_{t\in\mathbb{R}_{>0},t\to\infty}it.
Lemma 3.15.

Let 1νn11\leq\nu\leq n-1 be an integer.

  • (1)

    [Tpm(Cν(τ))~]=0\partial[\widetilde{T_{p}^{m}(C_{\nu}(\tau))}]=0 in (S0(BS)n)Γ(S_{0}(\mathbb{H}^{\mathrm{BS}})\otimes\mathcal{M}_{n}\otimes\mathbb{Q})_{\Gamma}.

  • (2)

    If mnm\geq n, then we have Tpm(Cν(τ))~𝒮()n,(p)\widetilde{T_{p}^{m}(C_{\nu}(\tau))}\in\mathcal{M}\mathcal{S}(\mathbb{H})\otimes\mathcal{M}_{n,(p)}. Hence Tpm(Cν(τ))~\widetilde{T_{p}^{m}(C_{\nu}(\tau))} defines a pp-integral homology class [Tpm(Cν(τ))~][\widetilde{T_{p}^{m}(C_{\nu}(\tau))}] in H1(YBS,n,(p))H_{1}(Y^{\mathrm{BS}},\mathcal{M}_{n,(p)}) which is independent of the choice of τ\tau.

  • (3)

    If mnm\geq n, then the image of the homology class [Tpm(Cν(τ))~][\widetilde{T_{p}^{m}(C_{\nu}(\tau))}] under the homomorphism H1(YBS,n,(p))H1(YBS,YBS,n,(p))H_{1}(Y^{\mathrm{BS}},\mathcal{M}_{n,(p)})\longrightarrow H_{1}(Y^{\rm BS},\partial Y^{\rm BS},\mathcal{M}_{n,(p)}) is Tpm([Cν])T_{p}^{m}([C_{\nu}]).

Proof.

Claim (1) follows from Lemma 3.11. Let us show claim (2). By Lemma 3.9, we have pmax{0,nk}C~ν,k,j(τ0,τ1)𝒮(BS)n,(p)p^{\max\{0,n-k\}}\widetilde{C}_{\nu,k,j}(\tau_{0},\tau_{1})\in\mathcal{M}\mathcal{S}(\mathbb{H}^{\mathrm{BS}})\otimes\mathcal{M}_{n,(p)}. Moreover, since 1νn11\leq\nu\leq n-1 and 2n2\leq n, we have

(n+1)A+(nν)(m2Ak)(n+1)A+m2Akmk(n+1)A+(n-\nu)(m-2A-k)\geq(n+1)A+m-2A-k\geq m-k

for any non-negative integer AA. This fact shows that Tpm(Cν(τ))~𝒮()n,(p)\widetilde{T_{p}^{m}(C_{\nu}(\tau))}\in\mathcal{M}\mathcal{S}(\mathbb{H})\otimes\mathcal{M}_{n,(p)} if mnm\geq n. It follows from Lemma 3.12 that the homology class [Tpm(Cν(τ))~][\widetilde{T_{p}^{m}(C_{\nu}(\tau))}] does not depend on the choice of τ\tau. Claim (3) follows from the definition of Tpm(Cν(τ))~\widetilde{T_{p}^{m}(C_{\nu}(\tau))} and Lemma 3.1.

4. Period

The aim of this section is to compute the value

Eisn,Tpm!(Cν(τ))~\displaystyle\langle\mathrm{Eis}_{n},\widetilde{T_{p}^{m!}(C_{\nu}(\tau))}\rangle

and its pp-adic limit as mm\to\infty. In this paper, we often consider the pp-adic limit, and hence the symbol limm\lim_{m\to\infty} will always mean the pp-adic limit. The following is the main result of this section.

Theorem 4.1.

For any integer ν{1,,n1}\nu\in\{1,\ldots,n-1\}, we have

limmEisn,Tpm!(Cν(τ))~=(1pn+1)(11pn+1ζ(ν)ζ(νn)ζ(1n)ζ(ν)1pnνζ(νn)1pν).\displaystyle\lim_{m\to\infty}\langle\mathrm{Eis}_{n},\widetilde{T_{p}^{m!}(C_{\nu}(\tau))}\rangle=(1-p^{n+1})\left(\frac{1}{1-p^{n+1}}\frac{\zeta(-\nu)\zeta(\nu-n)}{\zeta(-1-n)}-\frac{\zeta(-\nu)}{1-p^{n-\nu}}-\frac{\zeta(\nu-n)}{1-p^{\nu}}\right).

In fact, we show in §4.3 that

(4.1) limmk=0mp(nν)(mk)j=0pk1Eisn,C~ν,k,j(τpmk,pmkτ)=11pn+1ζ(ν)ζ(νn)ζ(1n)ζ(ν)1pnνζ(νn)1pν.\displaystyle\begin{split}\lim_{m\to\infty}\sum_{k=0}^{m}p^{(n-\nu)(m-k)}&\sum_{j=0}^{p^{k}-1}\langle\mathrm{Eis}_{n},\widetilde{C}_{\nu,k,j}\left(\frac{\tau}{p^{m-k}},p^{m-k}\tau\right)\rangle\\ &=\frac{1}{1-p^{n+1}}\frac{\zeta(-\nu)\zeta(\nu-n)}{\zeta(-1-n)}-\frac{\zeta(-\nu)}{1-p^{n-\nu}}-\frac{\zeta(\nu-n)}{1-p^{\nu}}.\end{split}

Hence Theorem 4.1 follows from (4.1) and the following lemma.

Lemma 4.2.

Suppose that the pp-adic limit

limmk=0mp(nν)(mk)j=0pk1Eisn,C~ν,k,j(τpmk,pmkτ)\lim_{m\to\infty}\sum_{k=0}^{m}p^{(n-\nu)(m-k)}\sum_{j=0}^{p^{k}-1}\langle\mathrm{Eis}_{n},\widetilde{C}_{\nu,k,j}\left(\frac{\tau}{p^{m-k}},p^{m-k}\tau\right)\rangle

exists. We then have

limmEisn,Tpm!(Cν(τ))~=(1pn+1)limmk=0mp(nν)(mk)j=0pk1Eisn,C~ν,k,j(τpmk,pmkτ).\displaystyle\lim_{m\to\infty}\langle\mathrm{Eis}_{n},\widetilde{T_{p}^{m!}(C_{\nu}(\tau))}\rangle=(1-p^{n+1})\lim_{m\to\infty}\sum_{k=0}^{m}p^{(n-\nu)(m-k)}\sum_{j=0}^{p^{k}-1}\langle\mathrm{Eis}_{n},\widetilde{C}_{\nu,k,j}\left(\frac{\tau}{p^{m-k}},p^{m-k}\tau\right)\rangle.
Proof.

For notational simplicity, we put

𝒲(m):=k=0mp(nν)(mk)j=0pk1Eisn,C~ν,k,j(τpmk,pmkτ) and 𝒲:=limm𝒲(m).\mathscr{W}^{(m)}:=\sum_{k=0}^{m}p^{(n-\nu)(m-k)}\sum_{j=0}^{p^{k}-1}\langle\mathrm{Eis}_{n},\widetilde{C}_{\nu,k,j}\left(\frac{\tau}{p^{m-k}},p^{m-k}\tau\right)\rangle\,\,\,\textrm{ and }\,\,\,\mathscr{W}:=\lim_{m\to\infty}\mathscr{W}^{(m)}.

We then have

Eisn,Tpm!(Cν(τ))~=A=0m!/2C(m!A,A)p(n+1)A𝒲(m!2A).\langle\mathrm{Eis}_{n},\widetilde{T_{p}^{m!}(C_{\nu}(\tau))}\rangle=\sum_{A=0}^{m!/2}C(m!-A,A)p^{(n+1)A}\mathscr{W}^{(m!-2A)}.

Take a positive integer QQ. Then there is a positive integer rr such that 𝒲(s)𝒲pQp\mathscr{W}^{(s)}-\mathscr{W}\in p^{Q}\mathbb{Z}_{p} for any integer srs\geq r. Hence we have

A=0m/2C(mA,A)p(n+1)A(𝒲(m2A)𝒲)\displaystyle\sum_{A=0}^{\lfloor m/2\rfloor}C(m-A,A)p^{(n+1)A}(\mathscr{W}^{(m-2A)}-\mathscr{W})
A=(mr)/2+1m/2C(mA,A)p(n+1)A(𝒲(m2A)𝒲)(modpQp).\displaystyle\equiv\sum_{A=\lfloor(m-r)/2\rfloor+1}^{\lfloor m/2\rfloor}C(m-A,A)p^{(n+1)A}(\mathscr{W}^{(m-2A)}-\mathscr{W})\pmod{p^{Q}\mathbb{Z}_{p}}.

The sequence (𝒲(m)𝒲)m=0(\mathscr{W}^{(m)}-\mathscr{W})_{m=0}^{\infty} is bounded in p\mathbb{Q}_{p}, and hence for any sufficiently large integer mm, we have

A=(mr)/2+1m/2C(mA,A)p(n+1)A(𝒲(m2A)𝒲)pQp.\sum_{A=\lfloor(m-r)/2\rfloor+1}^{\lfloor m/2\rfloor}C(m-A,A)p^{(n+1)A}(\mathscr{W}^{(m-2A)}-\mathscr{W})\in p^{Q}\mathbb{Z}_{p}.

This implies that

limmEisn,Tpm!(Cν(τ))~=limmA=0m!/2C(m!A,A)p(n+1)A𝒲.\lim_{m\to\infty}\langle\mathrm{Eis}_{n},\widetilde{T_{p}^{m!}(C_{\nu}(\tau))}\rangle=\lim_{m\to\infty}\sum_{A=0}^{m!/2}C(m!-A,A)p^{(n+1)A}\mathscr{W}.

Since C(m!A,A)=(m!A)(m!A1)C(m!-A,A)=\binom{m!}{A}-\binom{m!}{A-1} (note that (m1)=0\binom{m}{-1}=0), we have

A=0m!/2C(m!A,A)p(n+1)A\displaystyle\sum_{A=0}^{m!/2}C(m!-A,A)p^{(n+1)A} =A=0m!/2(m!A)p(n+1)AA=0m!/21(m!A)p(n+1)(A+1)\displaystyle=\sum_{A=0}^{m!/2}\binom{m!}{A}p^{(n+1)A}-\sum_{A=0}^{m!/2-1}\binom{m!}{A}p^{(n+1)(A+1)}
=(1pn+1)A=0m!/21(m!A)p(n+1)A+(m!m!/2)p(n+1)m!/2.\displaystyle=(1-p^{n+1})\sum_{A=0}^{m!/2-1}\binom{m!}{A}p^{(n+1)A}+\binom{m!}{m!/2}p^{(n+1)m!/2}.

Since

A=0m!/21(m!A)p(n+1)A\displaystyle\sum_{A=0}^{m!/2-1}\binom{m!}{A}p^{(n+1)A} A=0m!(m!A)p(n+1)A(modpm!/2)\displaystyle\equiv\sum_{A=0}^{m!}\binom{m!}{A}p^{(n+1)A}\pmod{p^{m!/2}}
=(1+pn+1)m!,\displaystyle=(1+p^{n+1})^{m!},

we obtain that

limmA=0m!/2C(m!A,A)p(n+1)A=(1pn+1)limm(1+pn+1)m!=1pn+1,\lim_{m\to\infty}\sum_{A=0}^{m!/2}C(m!-A,A)p^{(n+1)A}=(1-p^{n+1})\lim_{m\to\infty}(1+p^{n+1})^{m!}=1-p^{n+1},

which completes the proof. ∎

Therefore, it remains to prove (4.1), and this will be done in Proposition 4.13.

4.1. Eisn,C~ν,k,j(τ0,τ1)\langle\mathrm{Eis}_{n},\widetilde{C}_{\nu,k,j}\left(\tau_{0},\tau_{1}\right)\rangle

We start with computing the value Eisn,C~ν,k,j(τ0,τ1)\langle\mathrm{Eis}_{n},\widetilde{C}_{\nu,k,j}\left(\tau_{0},\tau_{1}\right)\rangle. In this subsection, we fix integers kk and jj with 0jpk10\leq j\leq p^{k}-1. Recall that

l:=lk(j) and d:=dklk(j)(j)=dklk(j)(j/plk(j))l:=l_{k}(j)\,\,\,\textrm{ and }\,\,\,d:=d_{k-l_{k}(j)}(j^{\prime})=d_{k-l_{k}(j)}(j/p^{l_{k}(j)})

are taken as in Definition 3.4 and (3.1) and that

C~ν,k,j(τ0,τ1)={1τ0+jpk,τ1+jpk}(pkX1jX2)νX2nν\displaystyle\widetilde{C}_{\nu,k,j}\left(\tau_{0},\tau_{1}\right)=\left\{\frac{-\frac{1}{\tau_{0}}+j}{p^{k}},\frac{\tau_{1}+j}{p^{k}}\right\}\otimes(p^{k}X_{1}-jX_{2})^{\nu}X_{2}^{n-\nu} {τ1+jpk,τ1+jpk+1}Pν,k,j(1)\displaystyle-\left\{\frac{\tau_{1}+j}{p^{k}},\frac{\tau_{1}+j}{p^{k}}+1\right\}\otimes P_{\nu,k,j}^{(1)}
{plτ0dpkl,plτ0dpkl+1}Pν,k,j(0).\displaystyle-\left\{\frac{p^{l}\tau_{0}-d}{p^{k-l}},\frac{p^{l}\tau_{0}-d}{p^{k-l}}+1\right\}\otimes P_{\nu,k,j}^{(0)}.

Hence we have

(4.2) Eisn,C~ν,k,j(τ0,τ1)=1τ0+jpkτ1+jpkEn+2(z)(pkzj)ν𝑑zτ1+jpkτ1+jpk+1En+2(z)Pν,k,j(1)(z,1)𝑑zplτ0dpklplτ0dpkl+1En+2(z)Pν,k,j(0)(z,1)𝑑z=1pk1τ0τ1En+2(z+jpk)zν𝑑zτ1+jpkτ1+jpk+1En+2(z)Pν,k,j(1)(z,1)𝑑zplτ0dpklplτ0dpkl+1En+2(z)Pν,k,j(0)(z,1)𝑑z.\displaystyle\begin{split}\langle\mathrm{Eis}_{n},\widetilde{C}_{\nu,k,j}\left(\tau_{0},\tau_{1}\right)\rangle&=\int_{\frac{-\frac{1}{\tau_{0}}+j}{p^{k}}}^{\frac{\tau_{1}+j}{p^{k}}}E_{n+2}(z)(p^{k}z-j)^{\nu}\,dz\\ &\quad-\int_{\frac{\tau_{1}+j}{p^{k}}}^{\frac{\tau_{1}+j}{p^{k}}+1}E_{n+2}(z)P_{\nu,k,j}^{(1)}(z,1)\,dz-\int_{\frac{p^{l}\tau_{0}-d}{p^{k-l}}}^{\frac{p^{l}\tau_{0}-d}{p^{k-l}}+1}E_{n+2}(z)P_{\nu,k,j}^{(0)}(z,1)\,dz\\ &=\frac{1}{p^{k}}\int_{-\frac{1}{\tau_{0}}}^{\tau_{1}}E_{n+2}\left(\frac{z+j}{p^{k}}\right)z^{\nu}\,dz\\ &\quad-\int_{\frac{\tau_{1}+j}{p^{k}}}^{\frac{\tau_{1}+j}{p^{k}}+1}E_{n+2}(z)P_{\nu,k,j}^{(1)}(z,1)\,dz-\int_{\frac{p^{l}\tau_{0}-d}{p^{k-l}}}^{\frac{p^{l}\tau_{0}-d}{p^{k-l}}+1}E_{n+2}(z)P_{\nu,k,j}^{(0)}(z,1)\,dz.\end{split}

The definition of the Eisenstein series En+2E_{n+2} shows that

En+2(z+jpk)1\displaystyle E_{n+2}\left(\frac{z+j}{p^{k}}\right)-1 =O(e2πIm(z)) for Im(z)1,\displaystyle=O(e^{-2\pi\operatorname{Im}(z)})\quad\text{ for }\operatorname{Im}(z)\geq 1,
En+2(z+jpk)pl(n+2)zn+2\displaystyle E_{n+2}\left(\frac{z+j}{p^{k}}\right)-\frac{p^{l(n+2)}}{z^{n+2}} =O(pl(n+2)|z|n+2e2πp2lk/Im(z)) for Im(z)1,\displaystyle=O\left(\frac{p^{l(n+2)}}{|z|^{n+2}}e^{-2\pi p^{2l-k}/\operatorname{Im}(z)}\right)\quad\text{ for }\operatorname{Im}(z)\leq 1,

where f(z)=O(g(z))f(z)=O(g(z)) means that there is a constant CC which does not depend on kk and jj such that |f(z)|Cg(z)|f(z)|\leq Cg(z). Set

k,j(s):=0(En+2(iy+jpk)1)ys𝑑y.\displaystyle\mathscr{L}_{k,j}(s):=\int_{0}^{\infty}\left(E_{n+2}\left(\frac{iy+j}{p^{k}}\right)-1\right)y^{s}dy.

We have the following Lemma 4.3 and Proposition 4.4, whose proofs will be given in §4.1.1 and §4.1.3, respectively.

Lemma 4.3.

The function k,j(s)\mathscr{L}_{k,j}(s) converges for Re(s)>n+1\operatorname{Re}(s)>n+1, and continued to a meromorphic function on \mathbb{C}. Moreover, it has at most simple poles at s=1s=-1 and s=n+1s=n+1. In particular, k,j(s)\mathscr{L}_{k,j}(s) is holomorphic at s=νs=\nu for any integer 1νn11\leq\nu\leq n-1.

Proposition 4.4.

We have

Eisn,C~ν,k,j(τ0,τ1)=iν+1pkk,j(ν)Eν,k,j(1)(jpk,1)Eν,k,j(0)(dpkl,1).\langle\mathrm{Eis}_{n},\widetilde{C}_{\nu,k,j}\left(\tau_{0},\tau_{1}\right)\rangle=\frac{i^{\nu+1}}{p^{k}}\mathscr{L}_{k,j}(\nu)-E_{\nu,k,j}^{(1){\ddagger}}\left(\frac{j}{p^{k}},1\right)-E_{\nu,k,j}^{(0){\ddagger}}\left(-\frac{d}{p^{k-l}},1\right).

Note that since [C~ν,k,j(τ0,τ1)]=0\partial[\widetilde{C}_{\nu,k,j}\left(\tau_{0},\tau_{1}\right)]=0 by Lemma 3.11, the value Eisn,C~ν,k,j(τ0,τ1)\langle\mathrm{Eis}_{n},\widetilde{C}_{\nu,k,j}\left(\tau_{0},\tau_{1}\right)\rangle does not depend on the choices of τ0\tau_{0} and τ1\tau_{1}. Therefore, in the following we take τ0=it0\tau_{0}=it_{0} and τ1=it1\tau_{1}=it_{1} for t0,t1>0t_{0},t_{1}\in\mathbb{R}_{>0}.

4.1.1. Computation of the first term of (4.2)

Here we compute the first term of (4.2):

1pkit0it1En+2(z+jpk)zν𝑑z.\frac{1}{p^{k}}\int_{\frac{i}{t_{0}}}^{it_{1}}E_{n+2}\left(\frac{z+j}{p^{k}}\right)z^{\nu}\,dz.

This integral is transformed as follows:

1pkit0it1En+2(z+jpk)zν𝑑z\displaystyle\frac{1}{p^{k}}\int_{\frac{i}{t_{0}}}^{it_{1}}E_{n+2}\left(\frac{z+j}{p^{k}}\right)z^{\nu}\,dz
=1pkit0(En+2(z+jpk)1)zν𝑑z+1pkit0it1zν𝑑z1pkit1(En+2(z+jpk)1)zν𝑑z\displaystyle=\frac{1}{p^{k}}\int_{\frac{i}{t_{0}}}^{\infty}\left(E_{n+2}\left(\frac{z+j}{p^{k}}\right)-1\right)z^{\nu}\,dz+\frac{1}{p^{k}}\int_{\frac{i}{t_{0}}}^{it_{1}}z^{\nu}\,dz-\frac{1}{p^{k}}\int_{it_{1}}^{\infty}\left(E_{n+2}\left(\frac{z+j}{p^{k}}\right)-1\right)z^{\nu}\,dz
=iν+1pk{1t0(En+2(iy+jpk)1)yν𝑑y+1t0t1yν𝑑yt1(En+2(iy+jpk)1)yν𝑑y}.\displaystyle=\frac{i^{\nu+1}}{p^{k}}\left\{\int_{\frac{1}{t_{0}}}^{\infty}\left(E_{n+2}\left(\frac{iy+j}{p^{k}}\right)-1\right)y^{\nu}\,dy+\int_{\frac{1}{t_{0}}}^{t_{1}}y^{\nu}\,dy-\int_{t_{1}}^{\infty}\left(E_{n+2}\left(\frac{iy+j}{p^{k}}\right)-1\right)y^{\nu}\,dy\right\}.

Set

𝒮k,j(t0,t1,s)\displaystyle\mathscr{S}_{k,j}(t_{0},t_{1},s) :=1t0(En+2(iy+jpk)1)ys𝑑y+1t0t1ys𝑑y,\displaystyle:=\int_{\frac{1}{t_{0}}}^{\infty}\left(E_{n+2}\left(\frac{iy+j}{p^{k}}\right)-1\right)y^{s}\,dy+\int_{\frac{1}{t_{0}}}^{t_{1}}y^{s}\,dy,
k,j(1)(t1,s)\displaystyle\mathscr{R}_{k,j}^{(1)}(t_{1},s) :=t1(En+2(iy+jpk)1)ys𝑑y,\displaystyle:=\int_{t_{1}}^{\infty}\left(E_{n+2}\left(\frac{iy+j}{p^{k}}\right)-1\right)y^{s}\,dy,
k,j(0)(t0,s)\displaystyle\mathscr{R}_{k,j}^{(0)}(t_{0},s) :=01t0(En+2(iy+jpk)pl(n+2)(iy)n+2)ys𝑑y.\displaystyle:=\int_{0}^{\frac{1}{t_{0}}}\left(E_{n+2}\left(\frac{iy+j}{p^{k}}\right)-\frac{p^{l(n+2)}}{(iy)^{n+2}}\right)y^{s}\,dy.

Then we have

1pkit0it1En+2(z+jpk)zν𝑑z=iν+1pk(𝒮k,j(t0,t1,ν)k,j(1)(t1,ν)).\displaystyle\frac{1}{p^{k}}\int_{\frac{i}{t_{0}}}^{it_{1}}E_{n+2}\left(\frac{z+j}{p^{k}}\right)z^{\nu}\,dz=\frac{i^{\nu+1}}{p^{k}}\left(\mathscr{S}_{k,j}(t_{0},t_{1},\nu)-\mathscr{R}_{k,j}^{(1)}(t_{1},\nu)\right).

Now, we see that

  • the first terms of 𝒮k,j(t0,t1,s)\mathscr{S}_{k,j}(t_{0},t_{1},s) and k,j(i)(ti,s)\mathscr{R}^{(i)}_{k,j}(t_{i},s) converge for all ss\in\mathbb{C},

  • the second term of 𝒮k,j(t0,t1,s)\mathscr{S}_{k,j}(t_{0},t_{1},s) is meromorphic and has at most simple pole at s=1s=-1.

In addition, we also see that

𝒮k,j(t0,t1,s)\displaystyle\mathscr{S}_{k,j}(t_{0},t_{1},s) =k,j(s)k,j(0)(t0,s)01t0pl(n+2)(iy)n+2ys𝑑y+0t1ys𝑑y\displaystyle=\mathscr{L}_{k,j}(s)-\mathscr{R}_{k,j}^{(0)}(t_{0},s)-\int_{0}^{\frac{1}{t_{0}}}\frac{p^{l(n+2)}}{(iy)^{n+2}}y^{s}\,dy+\int_{0}^{t_{1}}y^{s}\,dy
=k,j(s)k,j(0)(t0,s)pl(n+2)in+21sn11t0sn1+1s+1t1s+1.\displaystyle=\mathscr{L}_{k,j}(s)-\mathscr{R}_{k,j}^{(0)}(t_{0},s)-\frac{p^{l(n+2)}}{i^{n+2}}\frac{1}{s-n-1}\frac{1}{t_{0}^{s-n-1}}+\frac{1}{s+1}t_{1}^{s+1}.

In particular, all of these functions are meromorphically continued to ss\in\mathbb{C} and are holomorphic at s=νs=\nu. This proves Lemma 4.3, and moreover, we get the following.

Lemma 4.5.
1pkit0it1En+2(z+jpk)zν𝑑z\displaystyle\frac{1}{p^{k}}\int_{\frac{i}{t_{0}}}^{it_{1}}E_{n+2}\left(\frac{z+j}{p^{k}}\right)z^{\nu}\,dz
=iν+1pk{k,j(ν)+t1ν+1ν+1+pl(n+2)in+2t0nν+1nν+1k,j(1)(t1,ν)k,j(0)(t0,ν)}.\displaystyle=\frac{i^{\nu+1}}{p^{k}}\left\{\mathscr{L}_{k,j}(\nu)+\frac{t_{1}^{\nu+1}}{\nu+1}+\frac{p^{l(n+2)}}{i^{n+2}}\frac{t_{0}^{n-\nu+1}}{n-\nu+1}-\mathscr{R}_{k,j}^{(1)}(t_{1},\nu)-\mathscr{R}_{k,j}^{(0)}(t_{0},\nu)\right\}.

4.1.2. Computation of the second and the third terms of (4.2)

Here we compute

τ1+jpkτ1+jpk+1En+2(z)Pν,k,j(1)(z,1)𝑑z+plτ0dpklplτ0dpkl+1En+2(z)Pν,k,j(0)(z,1)𝑑z.\int_{\frac{\tau_{1}+j}{p^{k}}}^{\frac{\tau_{1}+j}{p^{k}}+1}E_{n+2}(z)P_{\nu,k,j}^{(1)}(z,1)\,dz+\int_{\frac{p^{l}\tau_{0}-d}{p^{k-l}}}^{\frac{p^{l}\tau_{0}-d}{p^{k-l}}+1}E_{n+2}(z)P_{\nu,k,j}^{(0)}(z,1)\,dz.

We put

𝒯ν,k,j(τ0,τ1):=τ1+jpkτ1+jpk+1(En+2(z)1)Pν,k,j(1)(z,1)𝑑z+plτ0dpklplτ0dpkl+1(En+2(z)1)Pν,k,j(0)(z,1)𝑑z.\displaystyle\mathscr{T}_{\nu,k,j}(\tau_{0},\tau_{1}):=\int_{\frac{\tau_{1}+j}{p^{k}}}^{\frac{\tau_{1}+j}{p^{k}}+1}\left(E_{n+2}(z)-1\right)P_{\nu,k,j}^{(1)}(z,1)\,dz+\int_{\frac{p^{l}\tau_{0}-d}{p^{k-l}}}^{\frac{p^{l}\tau_{0}-d}{p^{k-l}}+1}\left(E_{n+2}(z)-1\right)P_{\nu,k,j}^{(0)}(z,1)\,dz.

Note that 𝒯ν,k,j(τ0,τ1)0\mathscr{T}_{\nu,k,j}(\tau_{0},\tau_{1})\to 0 as τ0,τ1\tau_{0},\tau_{1}\to\infty. On the other hand by Lemma 3.7, we have

τ1+jpkτ1+jpk+1Pν,k,j(1)(z,1)𝑑z\displaystyle\int_{\frac{\tau_{1}+j}{p^{k}}}^{\frac{\tau_{1}+j}{p^{k}}+1}P_{\nu,k,j}^{(1)}(z,1)\,dz =Eν,k,j(1)(τ1+jpk,1)\displaystyle=E_{\nu,k,j}^{(1){\ddagger}}\left(\frac{\tau_{1}+j}{p^{k}},1\right)
=Eν,k,j(1)(jpk,1)+jpkτ1+jpkEν,k,j(1)(z,1)𝑑z\displaystyle=E_{\nu,k,j}^{(1){\ddagger}}\left(\frac{j}{p^{k}},1\right)+\int_{\frac{j}{p^{k}}}^{\frac{\tau_{1}+j}{p^{k}}}E_{\nu,k,j}^{(1)}(z,1)dz
=Eν,k,j(1)(jpk,1)+jpkτ1+jpk(pkzj)ν𝑑z\displaystyle=E_{\nu,k,j}^{(1){\ddagger}}\left(\frac{j}{p^{k}},1\right)+\int_{\frac{j}{p^{k}}}^{\frac{\tau_{1}+j}{p^{k}}}(p^{k}z-j)^{\nu}\,dz
=Eν,k,j(1)(jpk,1)+1pk(it1)ν+1ν+1.\displaystyle=E_{\nu,k,j}^{(1){\ddagger}}\left(\frac{j}{p^{k}},1\right)+\frac{1}{p^{k}}\frac{(it_{1})^{\nu+1}}{\nu+1}.

Similarly we have

plτ0dpklplτ0dpkl+1Pν,k,j(0)(z,1)𝑑z=Eν,k,j(0)(dpkl,1)+(1)ν+1pl(n+2)pk(it0)nν+1nν+1.\displaystyle\int_{\frac{p^{l}\tau_{0}-d}{p^{k-l}}}^{\frac{p^{l}\tau_{0}-d}{p^{k-l}}+1}P_{\nu,k,j}^{(0)}(z,1)\,dz=E_{\nu,k,j}^{(0){\ddagger}}\left(-\frac{d}{p^{k-l}},1\right)+(-1)^{\nu+1}\frac{p^{l(n+2)}}{p^{k}}\frac{(it_{0})^{n-\nu+1}}{n-\nu+1}.

4.1.3. Proof of Proposition 4.4

By combining the computations in subsections 4.1.1 and 4.1.2, we find

Eisn,C~ν,k,j(τ0,τ1)\displaystyle\langle\mathrm{Eis}_{n},\widetilde{C}_{\nu,k,j}\left(\tau_{0},\tau_{1}\right)\rangle
=iν+1pkν,k,j(ν)+1pk(it1)ν+1ν+1+(1)ν1pl(n+2)pk(it0)nν+1nν+1\displaystyle=\frac{i^{\nu+1}}{p^{k}}\mathscr{L}_{\nu,k,j}(\nu)+\frac{1}{p^{k}}\frac{(it_{1})^{\nu+1}}{\nu+1}+(-1)^{\nu-1}\frac{p^{l(n+2)}}{p^{k}}\frac{(it_{0})^{n-\nu+1}}{n-\nu+1}
Eν,k,j(1)(jpk,1)1pk(it1)ν+1ν+1Eν,k,j(0)(dpkl,1)(1)ν+1pl(n+2)pk(it0)nν+1nν+1\displaystyle\quad-E_{\nu,k,j}^{(1){\ddagger}}\left(\frac{j}{p^{k}},1\right)-\frac{1}{p^{k}}\frac{(it_{1})^{\nu+1}}{\nu+1}-E_{\nu,k,j}^{(0){\ddagger}}\left(-\frac{d}{p^{k-l}},1\right)-(-1)^{\nu+1}\frac{p^{l(n+2)}}{p^{k}}\frac{(it_{0})^{n-\nu+1}}{n-\nu+1}
iν+1pkk,j(1)(t1,ν)iν+1pkk,j(0)(t0,ν)𝒯k,j(τ0,τ1),\displaystyle\quad-\frac{i^{\nu+1}}{p^{k}}\mathscr{R}_{k,j}^{(1)}(t_{1},\nu)-\frac{i^{\nu+1}}{p^{k}}\mathscr{R}_{k,j}^{(0)}(t_{0},\nu)-\mathscr{T}_{k,j}(\tau_{0},\tau_{1}),

and hence

Eisn,C~ν,k,j(τ0,τ1)\displaystyle\langle\mathrm{Eis}_{n},\widetilde{C}_{\nu,k,j}\left(\tau_{0},\tau_{1}\right)\rangle =iν+1pkk,j(ν)Eν,k,j(1)(jpk,1)Eν,k,j(0)(dpkl,1)\displaystyle=\frac{i^{\nu+1}}{p^{k}}\mathscr{L}_{k,j}(\nu)-E_{\nu,k,j}^{(1){\ddagger}}\left(\frac{j}{p^{k}},1\right)-E_{\nu,k,j}^{(0){\ddagger}}\left(-\frac{d}{p^{k-l}},1\right)
iν+1pkk,j(1)(t1,ν)iν+1pkk,j(0)(t0,ν)𝒯k,j(τ0,τ1).\displaystyle\quad-\frac{i^{\nu+1}}{p^{k}}\mathscr{R}_{k,j}^{(1)}(t_{1},\nu)-\frac{i^{\nu+1}}{p^{k}}\mathscr{R}_{k,j}^{(0)}(t_{0},\nu)-\mathscr{T}_{k,j}(\tau_{0},\tau_{1}).

Since the value C~ν,k,j(τ0,τ1),Eisn\langle\widetilde{C}_{\nu,k,j}\left(\tau_{0},\tau_{1}\right),\mathrm{Eis}_{n}\rangle does not depend on t0t_{0} and t1t_{1}, we can take the limit t0,t1t_{0},t_{1}\to\infty. Then the last three terms vanish and we obtain the desired identity. ∎

4.2. Summation over jj

In this subsection, we compute the sum

j=0pk1Eisn,C~ν,k,j(τ0,τ1)=j=0pk1{iν+1pkk,j(ν)Eν,k,j(1)(jpk,1)Eν,k,j(0)(dpkl,1)}.\displaystyle\sum_{j=0}^{p^{k}-1}\langle\mathrm{Eis}_{n},\widetilde{C}_{\nu,k,j}\left(\tau_{0},\tau_{1}\right)\rangle=\sum_{j=0}^{p^{k}-1}\left\{\frac{i^{\nu+1}}{p^{k}}\mathscr{L}_{k,j}(\nu)-E_{\nu,k,j}^{(1){\ddagger}}\left(\frac{j}{p^{k}},1\right)-E_{\nu,k,j}^{(0){\ddagger}}\left(-\frac{d}{p^{k-l}},1\right)\right\}.

We keep using the abbreviation

l:=lk(j) and d:=dklk(j)(j)=dklk(j)(j/plk(j))l:=l_{k}(j)\,\,\,\textrm{ and }\,\,\,d:=d_{k-l_{k}(j)}(j^{\prime})=d_{k-l_{k}(j)}(j/p^{l_{k}(j)})

that are actually depending on kk and jj. Recall that B~t(x)=(Bt(x)Bt)/t\widetilde{B}_{t}(x)=(B_{t}(x)-B_{t})/t.

Lemma 4.6.
  1. (1)

    We have

    iν+1pkj=0pk1k,j(ν)=(1p(n+1)(k+1))(1p(n+1)k)pnν1pn+1ζ(ν)ζ(νn)ζ(1n).\displaystyle\frac{i^{\nu+1}}{p^{k}}\sum_{j=0}^{p^{k}-1}\mathscr{L}_{k,j}(\nu)=\frac{(1-p^{(n+1)(k+1)})-(1-p^{(n+1)k})p^{n-\nu}}{1-p^{n+1}}\frac{\zeta(-\nu)\zeta(\nu-n)}{\zeta(-1-n)}.
  2. (2)

    We have

    j=0pk1Eν,k,j(1)(jpk,1)=(1)ν(ν+1)1pkB~ν+2(pk)+j=0pk1Eν,k,j(1),(0,1).\displaystyle\sum_{j=0}^{p^{k}-1}E_{\nu,k,j}^{(1){\ddagger}}\left(\frac{j}{p^{k}},1\right)=\frac{(-1)^{\nu}}{(\nu+1)}\frac{1}{p^{k}}\widetilde{B}_{\nu+2}(p^{k})+\sum_{j=0}^{p^{k}-1}E_{\nu,k,j}^{(1),{\ddagger}}\left(0,1\right).
  3. (3)

    We have

    j=0pk1Eν,k,j(0)(dpkl,1)\displaystyle\sum_{j=0}^{p^{k}-1}E_{\nu,k,j}^{(0){\ddagger}}\left(-\frac{d}{p^{k-l}},1\right)
    =(1)νnν+11pkl=0k1pl(ν+1)(B~nν+2(pkl)pnν+1B~nν+2(pkl1))+j=0pk1Eν,k,j(0)(0,1).\displaystyle=\frac{(-1)^{\nu}}{n-\nu+1}\frac{1}{p^{k}}\sum_{l^{\prime}=0}^{k-1}p^{l^{\prime}(\nu+1)}\left(\widetilde{B}_{n-\nu+2}(p^{k-l^{\prime}})-p^{n-\nu+1}\widetilde{B}_{n-\nu+2}(p^{k-l^{\prime}-1})\right)+\sum_{j=0}^{p^{k}-1}E_{\nu,k,j}^{(0){\ddagger}}\left(0,1\right).
Proof.

Recall that

k,j(s)=0(En+2(iy+jpk)1)ys𝑑y.\displaystyle\mathscr{L}_{k,j}(s)=\int_{0}^{\infty}\left(E_{n+2}\left(\frac{iy+j}{p^{k}}\right)-1\right)y^{s}dy.

Hence we have

1pkj=0pk1k,j(s)\displaystyle\frac{1}{p^{k}}\sum_{j=0}^{p^{k}-1}\mathscr{L}_{k,j}(s) =1pkj=0pk12ζ(1n)μ=1σn+1(μ)e2πiμjpk0e2πμypkys+1dyy\displaystyle=\frac{1}{p^{k}}\sum_{j=0}^{p^{k}-1}\frac{2}{\zeta(-1-n)}\sum_{\mu=1}^{\infty}\sigma_{n+1}(\mu)e^{\frac{2\pi i\mu j}{p^{k}}}\int_{0}^{\infty}e^{-\frac{2\pi\mu y}{p^{k}}}y^{s+1}\frac{dy}{y}
=1pk2ζ(1n)j=0pk1μ=1σn+1(μ)e2πiμjpkpk(s+1)(2πμ)s+1Γ(s+1)\displaystyle=\frac{1}{p^{k}}\frac{2}{\zeta(-1-n)}\sum_{j=0}^{p^{k}-1}\sum_{\mu=1}^{\infty}\sigma_{n+1}(\mu)e^{\frac{2\pi i\mu j}{p^{k}}}\frac{p^{k(s+1)}}{(2\pi\mu)^{s+1}}\Gamma(s+1)
=2ζ(1n)Γ(s+1)pks(2π)s+1μ=1(j=0pk1e2πiμjpk)σn+1(μ)μs+1\displaystyle=\frac{2}{\zeta(-1-n)}\frac{\Gamma(s+1)p^{ks}}{(2\pi)^{s+1}}\sum_{\mu=1}^{\infty}\left(\sum_{j=0}^{p^{k}-1}e^{\frac{2\pi i\mu j}{p^{k}}}\right)\frac{\sigma_{n+1}(\mu)}{\mu^{s+1}}
=2Γ(s+1)ζ(1n)(2π)s+1μ=1σn+1(pkμ)μs+1.\displaystyle=\frac{2\Gamma(s+1)}{\zeta(-1-n)(2\pi)^{s+1}}\sum_{\mu=1}^{\infty}\frac{\sigma_{n+1}(p^{k}\mu)}{\mu^{s+1}}.

For notational simplicity, we put

p(s):=1ps.\mathcal{E}_{p}(s):=1-p^{-s}.

We then have

a=0σn+1(pk+a)pa(s+1)\displaystyle\sum_{a=0}^{\infty}\frac{\sigma_{n+1}(p^{k+a})}{p^{a(s+1)}} =a=01pa(s+1)1p(k+a+1)(n+1)1pn+1\displaystyle=\sum_{a=0}^{\infty}\frac{1}{p^{a(s+1)}}\frac{1-p^{(k+a+1)(n+1)}}{1-p^{n+1}}
=p(1+s)1p(k+1)(n+1)p(sn)11pn+1\displaystyle=\frac{\mathcal{E}_{p}(1+s)^{-1}-p^{(k+1)(n+1)}\mathcal{E}_{p}(s-n)^{-1}}{1-p^{n+1}}
=p(sn)p(k+1)(n+1)p(1+s)1pn+1p(1+s)1p(sn)1.\displaystyle=\frac{\mathcal{E}_{p}(s-n)-p^{(k+1)(n+1)}\mathcal{E}_{p}(1+s)}{1-p^{n+1}}\mathcal{E}_{p}(1+s)^{-1}\mathcal{E}_{p}(s-n)^{-1}.

Hence the well-known relation that ζ(1+s)ζ(sn)=a=1σn+1(a)a(s+1)\zeta(1+s)\zeta(s-n)=\sum_{a=1}^{\infty}\sigma_{n+1}(a)a^{-(s+1)} implies that

1pkj=0pk1k,j(s)\displaystyle\frac{1}{p^{k}}\sum_{j=0}^{p^{k}-1}\mathscr{L}_{k,j}(s) =2Γ(s+1)ζ(1n)(2π)s+1p(sn)p(k+1)(n+1)p(1+s)1pn+1ζ(s+1)ζ(sn).\displaystyle=\frac{2\Gamma(s+1)}{\zeta(-1-n)(2\pi)^{s+1}}\frac{\mathcal{E}_{p}(s-n)-p^{(k+1)(n+1)}\mathcal{E}_{p}(1+s)}{1-p^{n+1}}\zeta(s+1)\zeta(s-n).

By setting s=νs=\nu and using the functional equation of the Riemann zeta function (see [Hid93, p.29] for example), we find

iν+1pkj=0pk1k,j(ν)\displaystyle\frac{i^{\nu+1}}{p^{k}}\sum_{j=0}^{p^{k}-1}\mathscr{L}_{k,j}(\nu) =(p(νn)p(k+1)(n+1)p(ν+1))1pn+1ζ(ν)ζ(νn)ζ(1n)\displaystyle=\frac{(\mathcal{E}_{p}(\nu-n)-p^{(k+1)(n+1)}\mathcal{E}_{p}(\nu+1))}{1-p^{n+1}}\frac{\zeta(-\nu)\zeta(\nu-n)}{\zeta(-1-n)}
=(1p(n+1)(k+1))(1p(n+1)k)pnν1pn+1ζ(ν)ζ(νn)ζ(1n).\displaystyle=\frac{(1-p^{(n+1)(k+1)})-(1-p^{(n+1)k})p^{n-\nu}}{1-p^{n+1}}\frac{\zeta(-\nu)\zeta(\nu-n)}{\zeta(-1-n)}.

By using Lemma 3.7, we find

j=0pk1Eν,k,j(1),(jpk,1)\displaystyle\sum_{j=0}^{p^{k}-1}E_{\nu,k,j}^{(1),{\ddagger}}\left(\frac{j}{p^{k}},1\right) =j=0pk1(Eν,k,j(1),(jpk,1)Eν,k,j(1),(0,1))+j=0pk1Eν,k,j(1),(0,1)\displaystyle=\sum_{j=0}^{p^{k}-1}\left(E_{\nu,k,j}^{(1),{\ddagger}}\left(\frac{j}{p^{k}},1\right)-E_{\nu,k,j}^{(1),{\ddagger}}\left(0,1\right)\right)+\sum_{j=0}^{p^{k}-1}E_{\nu,k,j}^{(1),{\ddagger}}\left(0,1\right)
=j=0pk10jpkEν,k,j(1)(z,1)𝑑z+j=0pk1Eν,k,j(1),(0,1)\displaystyle=\sum_{j=0}^{p^{k}-1}\int_{0}^{\frac{j}{p^{k}}}E_{\nu,k,j}^{(1)}(z,1)dz+\sum_{j=0}^{p^{k}-1}E_{\nu,k,j}^{(1),{\ddagger}}\left(0,1\right)
=j=0pk10jpk(pkzj)ν𝑑z+j=0pk1Eν,k,j(1),(0,1)\displaystyle=\sum_{j=0}^{p^{k}-1}\int_{0}^{\frac{j}{p^{k}}}(p^{k}z-j)^{\nu}dz+\sum_{j=0}^{p^{k}-1}E_{\nu,k,j}^{(1),{\ddagger}}\left(0,1\right)
=(1)ν(ν+1)1pkj=0pk1jν+1+j=0pk1Eν,k,j(1),(0,1).\displaystyle=\frac{(-1)^{\nu}}{(\nu+1)}\frac{1}{p^{k}}\sum_{j=0}^{p^{k}-1}j^{\nu+1}+\sum_{j=0}^{p^{k}-1}E_{\nu,k,j}^{(1),{\ddagger}}\left(0,1\right).

Therefore, claim (2) follows from the fact that j=0pk1jν+1=B~ν+2(pk)\sum_{j=0}^{p^{k}-1}j^{\nu+1}=\widetilde{B}_{\nu+2}(p^{k}).

Lemma 3.7 shows that

Eν,k,j(0)(dpkl,1)\displaystyle E_{\nu,k,j}^{(0){\ddagger}}\left(-\frac{d}{p^{k-l}},1\right) =(1)νplν0dpkl(pklz+d)nν𝑑z+Eν,k,j(0)(0,1)\displaystyle=-(-1)^{\nu}p^{l\nu}\int_{0}^{-\frac{d}{p^{k-l}}}(p^{k-l}z+d)^{n-\nu}dz+E_{\nu,k,j}^{(0){\ddagger}}\left(0,1\right)
=(1)νnν+11pkpl(ν+1)dnν+1+Eν,k,j(0)(0,1).\displaystyle=\frac{(-1)^{\nu}}{n-\nu+1}\frac{1}{p^{k}}p^{l(\nu+1)}d^{n-\nu+1}+E_{\nu,k,j}^{(0){\ddagger}}\left(0,1\right).

Since d0(0)=0d_{0}(0)=0, nν+12n-\nu+1\geq 2, and the map adN(a)a\mapsto d_{N}(a) induces a permutation on the set {1a<pN(a,pN)=1}\{1\leq a<p^{N}\mid(a,p^{N})=1\}, we find

1pkj=0pk1pl(ν+1)dnν+1\displaystyle\frac{1}{p^{k}}\sum_{j=0}^{p^{k}-1}p^{l(\nu+1)}d^{n-\nu+1} =1pkl=0k1pl(ν+1)d=1(d,pkl)=1pkl1dnν+1\displaystyle=\frac{1}{p^{k}}\sum_{l^{\prime}=0}^{k-1}p^{l^{\prime}(\nu+1)}\sum_{\begin{subarray}{c}d^{\prime}=1\\ (d^{\prime},p^{k-l^{\prime}})=1\end{subarray}}^{p^{k-l^{\prime}}-1}d^{\prime n-\nu+1}
=1pkl=0k1pl(ν+1)(B~nν+2(pkl)pnν+1B~nν+2(pkl1)).\displaystyle=\frac{1}{p^{k}}\sum_{l^{\prime}=0}^{k-1}p^{l^{\prime}(\nu+1)}\left(\widetilde{B}_{n-\nu+2}(p^{k-l^{\prime}})-p^{n-\nu+1}\widetilde{B}_{n-\nu+2}(p^{k-l^{\prime}-1})\right).

Lemma 4.7.
  1. (1)

    We have

    j=0pk1Eν,k,j(1)(0,1)=(1)ν+1μ=0ν(νμ)(1)μpkμBμ+1μ+1B~νμ+1(pk).\displaystyle\sum_{j=0}^{p^{k}-1}E_{\nu,k,j}^{(1){\ddagger}}(0,1)=(-1)^{\nu+1}\sum_{\mu=0}^{\nu}\binom{\nu}{\mu}(-1)^{\mu}p^{k\mu}\frac{B_{\mu+1}}{\mu+1}\widetilde{B}_{\nu-\mu+1}(p^{k}).
  2. (2)

    We have

    j=0pk1Eν,k,j(0)(0,1)=(1)νpkνBnν+1nν+1+(1)νμ=0nν(nνμ)pkμBμ+1μ+1l=0k1pl(νμ)×(B~nνμ+1(pkl)pnνμB~nνμ+1(pkl1)).\sum_{j=0}^{p^{k}-1}E_{\nu,k,j}^{(0){\ddagger}}\left(0,1\right)=(-1)^{\nu}p^{k\nu}\frac{B_{n-\nu+1}}{n-\nu+1}+(-1)^{\nu}\sum_{\mu=0}^{n-\nu}\begin{pmatrix}n-\nu\\ \mu\end{pmatrix}p^{k\mu}\frac{B_{\mu+1}}{\mu+1}\sum_{l^{\prime}=0}^{k-1}p^{l^{\prime}(\nu-\mu)}\\ \times\left(\widetilde{B}_{n-\nu-\mu+1}(p^{k-l^{\prime}})-p^{n-\nu-\mu}\widetilde{B}_{n-\nu-\mu+1}(p^{k-l^{\prime}-1})\right).
Proof.

Note that (X1μX2nμ)(0,1)=Bμ+1/(μ+1)(X_{1}^{\mu}X_{2}^{n-\mu})^{{\ddagger}}(0,1)=-B_{\mu+1}/(\mu+1). Since Eν,k,j(1)(X1,X2)=(pkX1jX2)νX2nνE_{\nu,k,j}^{(1)}(X_{1},X_{2})=(p^{k}X_{1}-jX_{2})^{\nu}X_{2}^{n-\nu}, we have

Eν,k,j(1)(0,1)=μ=0ν(νμ)pkμ(Bμ+1μ+1)(j)νμ.E_{\nu,k,j}^{(1){\ddagger}}\left(0,1\right)=\sum_{\mu=0}^{\nu}\binom{\nu}{\mu}p^{k\mu}\left(-\frac{B_{\mu+1}}{\mu+1}\right)(-j)^{\nu-\mu}.

Hence we have

j=0pk1Eν,k,j(1)(0,1)=(1)ν+1μ=0ν(νμ)(1)μpkμBμ+1μ+1B~νμ+1(pk).\displaystyle\sum_{j=0}^{p^{k}-1}E_{\nu,k,j}^{(1){\ddagger}}(0,1)=(-1)^{\nu+1}\sum_{\mu=0}^{\nu}\binom{\nu}{\mu}(-1)^{\mu}p^{k\mu}\frac{B_{\mu+1}}{\mu+1}\widetilde{B}_{\nu-\mu+1}(p^{k}).

Since Eν,k,j(0)(X1,X2)=(1)ν+1(plX2)ν(pklX1+dX2)nνE_{\nu,k,j}^{(0)}(X_{1},X_{2})=(-1)^{\nu+1}(p^{l}X_{2})^{\nu}(p^{k-l}X_{1}+dX_{2})^{n-\nu}, we have

Eν,k,j(0)(0,1)=(1)ν+1plνμ=0nν(nνμ)pμ(kl)(Bμ+1μ+1)dnνμ.E_{\nu,k,j}^{(0){\ddagger}}(0,1)=(-1)^{\nu+1}p^{l\nu}\sum_{\mu=0}^{n-\nu}\binom{n-\nu}{\mu}p^{\mu(k-l)}\left(-\frac{B_{\mu+1}}{\mu+1}\right)d^{n-\nu-\mu}.

First note that in the case where j=0j=0, since d=d0(0)=0d=d_{0}(0)=0, we find

Eν,k,0(0)(0,1)=(1)νpkνBnν+1nν+1.E_{\nu,k,0}^{(0){\ddagger}}(0,1)=(-1)^{\nu}p^{k\nu}\frac{B_{n-\nu+1}}{n-\nu+1}.

Then by using the same argument as in the proof of Lemma 4.6(3), we also obtain

j=1pk1Eν,k,j(0)(0,1)\displaystyle\sum_{j=1}^{p^{k}-1}E_{\nu,k,j}^{(0){\ddagger}}(0,1) =(1)νl=0k1d=1(d,pkl)=1pkl1plνμ=0nν(nνμ)pμ(kl)Bμ+1μ+1dnνμ\displaystyle=(-1)^{\nu}\sum_{l^{\prime}=0}^{k-1}\sum_{\begin{subarray}{c}d^{\prime}=1\\ (d^{\prime},p^{k-l^{\prime}})=1\end{subarray}}^{p^{k-l^{\prime}}-1}p^{l^{\prime}\nu}\sum_{\mu=0}^{n-\nu}\binom{n-\nu}{\mu}p^{\mu(k-l^{\prime})}\frac{B_{\mu+1}}{\mu+1}d^{\prime n-\nu-\mu}
=(1)νl=0k1plνμ=0nν(nνμ)pμ(kl)Bμ+1μ+1\displaystyle=(-1)^{\nu}\sum_{l^{\prime}=0}^{k-1}p^{l^{\prime}\nu}\sum_{\mu=0}^{n-\nu}\binom{n-\nu}{\mu}p^{\mu(k-l^{\prime})}\frac{B_{\mu+1}}{\mu+1}
×(B~nνμ+1(pkl)pnνμB~nνμ+1(pkl1)).\displaystyle\quad\quad\quad\times\left(\widetilde{B}_{n-\nu-\mu+1}(p^{k-l^{\prime}})-p^{n-\nu-\mu}\widetilde{B}_{n-\nu-\mu+1}(p^{k-l^{\prime}-1})\right).

This completes the proof. ∎

4.3. Summation over kk and the pp-adic limits

In this subsection, we compute the value

𝒲(m):=k=0mp(nν)(mk)j=0pk1Eisn,C~ν,k,j(τpmk,pmkτ)\mathscr{W}^{(m)}:=\sum_{k=0}^{m}p^{(n-\nu)(m-k)}\sum_{j=0}^{p^{k}-1}\langle\mathrm{Eis}_{n},\widetilde{C}_{\nu,k,j}\left(\frac{\tau}{p^{m-k}},p^{m-k}\tau\right)\rangle

and its pp-adic limit as mm\to\infty. This enables us to complete the proof of Theorem 4.1.

We keep the notation in the previous sections. Proposition 4.4 shows that

Eisn,C~ν,k,j(τ0,τ1)=iν+1pkk,j(ν)Eν,k,j(1)(jpk,1)Eν,k,j(0)(dpkl,1),\langle\mathrm{Eis}_{n},\widetilde{C}_{\nu,k,j}\left(\tau_{0},\tau_{1}\right)\rangle=\frac{i^{\nu+1}}{p^{k}}\mathscr{L}_{k,j}(\nu)-E_{\nu,k,j}^{(1){\ddagger}}\left(\frac{j}{p^{k}},1\right)-E_{\nu,k,j}^{(0){\ddagger}}\left(-\frac{d}{p^{k-l}},1\right),

and hence

𝒲(m)=k=0mp(nν)(mk)j=0pk1(iν+1pkk,j(ν)Eν,k,j(1)(jpk,1)Eν,k,j(0)(dpkl,1)).\mathscr{W}^{(m)}=\sum_{k=0}^{m}p^{(n-\nu)(m-k)}\sum_{j=0}^{p^{k}-1}\left(\frac{i^{\nu+1}}{p^{k}}\mathscr{L}_{k,j}(\nu)-E_{\nu,k,j}^{(1){\ddagger}}\left(\frac{j}{p^{k}},1\right)-E_{\nu,k,j}^{(0){\ddagger}}\left(-\frac{d}{p^{k-l}},1\right)\right).

We set

𝒲1(m)\displaystyle\mathscr{W}_{1}^{(m)} :=k=0mp(nν)(mk)j=0pk1iν+1pkk,j(ν),\displaystyle:=\sum_{k=0}^{m}p^{(n-\nu)(m-k)}\sum_{j=0}^{p^{k}-1}\frac{i^{\nu+1}}{p^{k}}\mathscr{L}_{k,j}(\nu),
𝒲2(m)\displaystyle\mathscr{W}_{2}^{(m)} :=k=0mp(nν)(mk)j=0pk1Eν,k,j(1)(jpk,1),\displaystyle:=\sum_{k=0}^{m}p^{(n-\nu)(m-k)}\sum_{j=0}^{p^{k}-1}E_{\nu,k,j}^{(1){\ddagger}}\left(\frac{j}{p^{k}},1\right),
𝒲3(m)\displaystyle\mathscr{W}_{3}^{(m)} :=k=0mp(nν)(mk)j=0pk1Eν,k,j(0)(dpkl,1),\displaystyle:=\sum_{k=0}^{m}p^{(n-\nu)(m-k)}\sum_{j=0}^{p^{k}-1}E_{\nu,k,j}^{(0){\ddagger}}\left(-\frac{d}{p^{k-l}},1\right),

so that we have 𝒲(m)=𝒲1(m)𝒲2(m)𝒲3(m)\mathscr{W}^{(m)}=\mathscr{W}^{(m)}_{1}-\mathscr{W}^{(m)}_{2}-\mathscr{W}^{(m)}_{3}.

Lemma 4.8.

We have

limm𝒲1(m)=11pn+1ζ(ν)ζ(νn)ζ(1n).\lim_{m\to\infty}\mathscr{W}_{1}^{(m)}=\frac{1}{1-p^{n+1}}\frac{\zeta(-\nu)\zeta(\nu-n)}{\zeta(-1-n)}.
Proof.

Lemma 4.6(1) shows that

𝒲1(m)\displaystyle\mathscr{W}_{1}^{(m)} =11pn+1ζ(ν)ζ(νn)ζ(1n)k=0mp(nν)(mk)((1p(n+1)(k+1))(1p(n+1)k)pnν).\displaystyle=\frac{1}{1-p^{n+1}}\frac{\zeta(-\nu)\zeta(\nu-n)}{\zeta(-1-n)}\sum_{k=0}^{m}p^{(n-\nu)(m-k)}((1-p^{(n+1)(k+1)})-(1-p^{(n+1)k})p^{n-\nu}).

Moreover, we have

k=0mp(nν)(mk)\displaystyle\sum_{k=0}^{m}p^{(n-\nu)(m-k)} ((1p(n+1)(k+1))(1p(n+1)k)pnν)\displaystyle((1-p^{(n+1)(k+1)})-(1-p^{(n+1)k})p^{n-\nu})
=p(nν)m(1pnν)k=0mp(nν)kp(nν)m(pn+1pnν)k=0mp(ν+1)k\displaystyle=p^{(n-\nu)m}(1-p^{n-\nu})\sum_{k=0}^{m}p^{-(n-\nu)k}-p^{(n-\nu)m}(p^{n+1}-p^{n-\nu})\sum_{k=0}^{m}p^{(\nu+1)k}
=p(nν)m(1pnν)1p(nν)(m+1)1p(nν)p(nν)m(pn+1pnν)1p(ν+1)(m+1)1pν+1\displaystyle=p^{(n-\nu)m}(1-p^{n-\nu})\frac{1-p^{-(n-\nu)(m+1)}}{1-p^{-(n-\nu)}}-p^{(n-\nu)m}(p^{n+1}-p^{n-\nu})\frac{1-p^{(\nu+1)(m+1)}}{1-p^{\nu+1}}
=1p(nν)(m+1)p(nν)m(pn+1pnν)1p(ν+1)(m+1)1pν+1\displaystyle=1-p^{(n-\nu)(m+1)}-p^{(n-\nu)m}(p^{n+1}-p^{n-\nu})\frac{1-p^{(\nu+1)(m+1)}}{1-p^{\nu+1}}
1(m).\displaystyle\to 1\quad\quad(m\to\infty).

Lemma 4.9.

Let ss and tt be integers with t>0t>0.

  • (1)

    For any positive integer u>0u>0, we have

    limmk=0mpu(mk)pksB~t(pks)=Bt11pu.\lim_{m\to\infty}\sum_{k=0}^{m}\frac{p^{u(m-k)}}{p^{k-s}}\widetilde{B}_{t}(p^{k-s})=\frac{B_{t-1}}{1-p^{u}}.
  • (2)

    For any integer uu and any ε>0\varepsilon>0 with u+ε>0u+\varepsilon>0, we have

    limmpmεk=0mpu(mk)pksB~t(pks)=0.\lim_{m\to\infty}p^{m\varepsilon}\sum_{k=0}^{m}\frac{p^{u(m-k)}}{p^{k-s}}\widetilde{B}_{t}(p^{k-s})=0.
Proof.

We have

k=0mpu(mk)1pksB~t(pks)\displaystyle\sum_{k=0}^{m}p^{u(m-k)}\frac{1}{p^{k-s}}\widetilde{B}_{t}(p^{k-s}) =1tk=0mpu(mk)μ=1t(tμ)Btμp(μ1)(ks)\displaystyle=\frac{1}{t}\sum_{k=0}^{m}p^{u(m-k)}\sum_{\mu=1}^{t}\binom{t}{\mu}B_{t-\mu}p^{(\mu-1)(k-s)}
=1tμ=1t(tμ)Btμk=0mpu(mk)+(μ1)(ks)\displaystyle=\frac{1}{t}\sum_{\mu=1}^{t}\binom{t}{\mu}B_{t-\mu}\sum_{k=0}^{m}p^{u(m-k)+(\mu-1)(k-s)}
=1tμ=1t(tμ)Btμpum(μ1)s1p(μ1u)(m+1)1pμ1u\displaystyle=\frac{1}{t}\sum_{\mu=1}^{t}\binom{t}{\mu}B_{t-\mu}p^{um-(\mu-1)s}\frac{1-p^{(\mu-1-u)(m+1)}}{1-p^{\mu-1-u}}
=1tμ=1t(tμ)Btμpum(μ1)sp(μ1)s+(m+1)(μ1)u1pμ1u.\displaystyle=\frac{1}{t}\sum_{\mu=1}^{t}\binom{t}{\mu}B_{t-\mu}\frac{p^{um-(\mu-1)s}-p^{-(\mu-1)s+(m+1)(\mu-1)-u}}{1-p^{\mu-1-u}}.

Hence claim (2) is clear. Moreover if u>0u>0, then all the terms with μ2\mu\geq 2 vanish as mm\to\infty, and therefore,

limmk=0mpu(mk)pksB~t(pks)=Bt1pu1pu=Bt11pu.\lim_{m\to\infty}\sum_{k=0}^{m}\frac{p^{u(m-k)}}{p^{k-s}}\widetilde{B}_{t}(p^{k-s})=B_{t-1}\frac{-p^{-u}}{1-p^{-u}}=\frac{B_{t-1}}{1-p^{u}}.

Lemma 4.10.

We have

limm𝒲2(m)=(1)ν1pnνBν+1ν+1=11pnνζ(ν).\lim_{m\to\infty}\mathscr{W}_{2}^{(m)}=\frac{(-1)^{\nu}}{1-p^{n-\nu}}\frac{B_{\nu+1}}{\nu+1}=\frac{1}{1-p^{n-\nu}}\zeta(-\nu).
Proof.

Lemma 4.6(2) shows that

𝒲2(m)\displaystyle\mathscr{W}_{2}^{(m)} =k=0mp(nν)(mk)j=0pk1Eν,k,j(1)(jpk,1)\displaystyle=\sum_{k=0}^{m}p^{(n-\nu)(m-k)}\sum_{j=0}^{p^{k}-1}E_{\nu,k,j}^{(1){\ddagger}}\left(\frac{j}{p^{k}},1\right)
=k=0mp(nν)(mk)((1)ν(ν+1)1pkB~ν+2(pk)+j=0pk1Eν,k,j(1),(0,1)).\displaystyle=\sum_{k=0}^{m}p^{(n-\nu)(m-k)}\left(\frac{(-1)^{\nu}}{(\nu+1)}\frac{1}{p^{k}}\widetilde{B}_{\nu+2}(p^{k})+\sum_{j=0}^{p^{k}-1}E_{\nu,k,j}^{(1),{\ddagger}}\left(0,1\right)\right).

Hence Lemma 4.9(1) implies that

limm𝒲2(m)=(1)ν1pnνBν+1ν+1+limmk=0mp(nν)(mk)j=0pk1Eν,k,j(1),(0,1).\lim_{m\to\infty}\mathscr{W}_{2}^{(m)}=\frac{(-1)^{\nu}}{1-p^{n-\nu}}\frac{B_{\nu+1}}{\nu+1}+\lim_{m\to\infty}\sum_{k=0}^{m}p^{(n-\nu)(m-k)}\sum_{j=0}^{p^{k}-1}E_{\nu,k,j}^{(1),{\ddagger}}(0,1).

Moreover, by Lemma 4.7, we have

k=0mp(nν)(mk)j=0pk1Eν,k,j(1),(0,1)\displaystyle\sum_{k=0}^{m}p^{(n-\nu)(m-k)}\sum_{j=0}^{p^{k}-1}E_{\nu,k,j}^{(1),{\ddagger}}(0,1) =k=0mp(nν)(mk)μ=0ν(1)ν+μ+1(νμ)pk(μ+1)Bμ+1μ+1B~νμ+1(pk)pk\displaystyle=\sum_{k=0}^{m}p^{(n-\nu)(m-k)}\sum_{\mu=0}^{\nu}(-1)^{\nu+\mu+1}\binom{\nu}{\mu}p^{k(\mu+1)}\frac{B_{\mu+1}}{\mu+1}\frac{\widetilde{B}_{\nu-\mu+1}(p^{k})}{p^{k}}
=μ=0ν(1)ν+μ+1(νμ)Bμ+1μ+1pm(μ+1)k=0mp(nνμ1)(mk)B~νμ+1(pk)pk.\displaystyle=\sum_{\mu=0}^{\nu}(-1)^{\nu+\mu+1}\binom{\nu}{\mu}\frac{B_{\mu+1}}{\mu+1}p^{m(\mu+1)}\sum_{k=0}^{m}p^{(n-\nu-\mu-1)(m-k)}\frac{\widetilde{B}_{\nu-\mu+1}(p^{k})}{p^{k}}.

Since μ+1+nνμ1=nν>0\mu+1+n-\nu-\mu-1=n-\nu>0, Lemma 4.9(2) shows that

limmpm(μ+1)k=0mp(nνμ1)(mk)B~νμ+1(pk)pk=0.\lim_{m\to\infty}p^{m(\mu+1)}\sum_{k=0}^{m}p^{(n-\nu-\mu-1)(m-k)}\frac{\widetilde{B}_{\nu-\mu+1}(p^{k})}{p^{k}}=0.

Hence we conclude that

limm𝒲2(m)=(1)ν1pnνBν+1ν+1.\lim_{m\to\infty}\mathscr{W}_{2}^{(m)}=\frac{(-1)^{\nu}}{1-p^{n-\nu}}\frac{B_{\nu+1}}{\nu+1}.

Lemma 4.11.

Let bb, ss, tt, and uu be integers with s0s\geq 0 and t0t\geq 0.

  • (1)

    If s>0s>0 and u>0u>0, then

    limmk=0mpu(mk)l=0k1pls(B~t+1(pkl)pklpbB~t+1(pkl1)pkl1)=Bt1ps1pb1pu.\displaystyle\lim_{m\to\infty}\sum_{k=0}^{m}p^{u(m-k)}\sum_{l^{\prime}=0}^{k-1}p^{l^{\prime}s}\left(\frac{\widetilde{B}_{t+1}(p^{k-l^{\prime}})}{p^{k-l^{\prime}}}-p^{b}\frac{\widetilde{B}_{t+1}(p^{k-l^{\prime}-1})}{p^{k-l^{\prime}-1}}\right)=\frac{B_{t}}{1-p^{s}}\frac{1-p^{b}}{1-p^{u}}.
  • (2)

    For any ε>0\varepsilon>0 with u+ε>0u+\varepsilon>0, we have

    limmpεmk=0mpu(mk)l=0k1pls(B~t+1(pkl)pklpbB~t+1(pkl1)pkl1)=0.\displaystyle\lim_{m\to\infty}p^{\varepsilon m}\sum_{k=0}^{m}p^{u(m-k)}\sum_{l^{\prime}=0}^{k-1}p^{l^{\prime}s}\left(\frac{\widetilde{B}_{t+1}(p^{k-l^{\prime}})}{p^{k-l^{\prime}}}-p^{b}\frac{\widetilde{B}_{t+1}(p^{k-l^{\prime}-1})}{p^{k-l^{\prime}-1}}\right)=0.
Proof.

We have

k=0mpu(mk)l=0k1pls(B~t+1(pkl)pklpbB~t+1(pkl1)pkl1)\displaystyle\sum_{k=0}^{m}p^{u(m-k)}\sum_{l^{\prime}=0}^{k-1}p^{l^{\prime}s}\left(\frac{\widetilde{B}_{t+1}(p^{k-l^{\prime}})}{p^{k-l^{\prime}}}-p^{b}\frac{\widetilde{B}_{t+1}(p^{k-l^{\prime}-1})}{p^{k-l^{\prime}-1}}\right)
=1t+1k=0mpu(mk)l=0k1plsμ=1t+1(t+1μ)Bt+1μ×(p(kl)(μ1)p(kl1)(μ1)+b)\displaystyle=\frac{1}{t+1}\sum_{k=0}^{m}p^{u(m-k)}\sum_{l^{\prime}=0}^{k-1}p^{l^{\prime}s}\sum_{\mu=1}^{t+1}\binom{t+1}{\mu}B_{t+1-\mu}\times(p^{(k-l^{\prime})(\mu-1)}-p^{(k-l^{\prime}-1)(\mu-1)+b})
=1t+1μ=1t+1(t+1μ)Bt+1μ×(1pb(μ1))pumk=0mpk(μ1u)l=0k1pl(sμ+1)\displaystyle=\frac{1}{t+1}\sum_{\mu=1}^{t+1}\binom{t+1}{\mu}B_{t+1-\mu}\times(1-p^{b-(\mu-1)})p^{um}\sum_{k=0}^{m}p^{k(\mu-1-u)}\sum_{l^{\prime}=0}^{k-1}p^{l^{\prime}(s-\mu+1)}
=1t+1μ=1t+1(t+1μ)Bt+1μ×(1pb(μ1))pumk=0mpk(μ1u)1pk(sμ+1)1psμ+1\displaystyle=\frac{1}{t+1}\sum_{\mu=1}^{t+1}\binom{t+1}{\mu}B_{t+1-\mu}\times(1-p^{b-(\mu-1)})p^{um}\sum_{k=0}^{m}p^{k(\mu-1-u)}\frac{1-p^{k(s-\mu+1)}}{1-p^{s-\mu+1}}
=1t+1μ=1t+1(t+1μ)Bt+1μ1psμ+1×(1pb(μ1))(pump(m+1)(μ1)u1pμ1upump(m+1)su1psu),\displaystyle=\frac{1}{t+1}\sum_{\mu=1}^{t+1}\binom{t+1}{\mu}\frac{B_{t+1-\mu}}{1-p^{s-\mu+1}}\times(1-p^{b-(\mu-1)})\left(\frac{p^{um}-p^{(m+1)(\mu-1)-u}}{1-p^{\mu-1-u}}-\frac{p^{um}-p^{(m+1)s-u}}{1-p^{s-u}}\right),

which implies this lemma in the same way as in the proof of Lemma 4.9. ∎

Lemma 4.12.

We have

limm𝒲3(m)=(1)ν1pνBnν+1nν+1=11pνζ(νn).\lim_{m\to\infty}\mathscr{W}_{3}^{(m)}=\frac{(-1)^{\nu}}{1-p^{\nu}}\frac{B_{n-\nu+1}}{n-\nu+1}=\frac{1}{1-p^{\nu}}\zeta(\nu-n).
Proof.

Recall that in Lemma 4.6(3) we have shown that

j=0pk1Eν,k,j(0)(dpkl,1)=(1)νnν+11pkl=0k1pl(ν+1)(B~nν+2(pkl)pnν+1B~nν+2(pkl1))+j=0pk1Eν,k,j(0)(0,1).\sum_{j=0}^{p^{k}-1}E_{\nu,k,j}^{(0){\ddagger}}\left(-\frac{d}{p^{k-l}},1\right)\\ =\frac{(-1)^{\nu}}{n-\nu+1}\frac{1}{p^{k}}\sum_{l^{\prime}=0}^{k-1}p^{l^{\prime}(\nu+1)}\left(\widetilde{B}_{n-\nu+2}(p^{k-l^{\prime}})-p^{n-\nu+1}\widetilde{B}_{n-\nu+2}(p^{k-l^{\prime}-1})\right)+\sum_{j=0}^{p^{k}-1}E_{\nu,k,j}^{(0){\ddagger}}(0,1).

Since 1νn11\leq\nu\leq n-1, Lemma 4.11(1) shows that

k=0mp(nν)(mk)l=0k1plν(B~nν+2(pkl)pklpnνB~nν+2(pkl1)pkl1)\displaystyle\sum_{k=0}^{m}p^{(n-\nu)(m-k)}\sum_{l^{\prime}=0}^{k-1}p^{l^{\prime}\nu}\left(\frac{\widetilde{B}_{n-\nu+2}(p^{k-l^{\prime}})}{p^{k-l^{\prime}}}-p^{n-\nu}\frac{\widetilde{B}_{n-\nu+2}(p^{k-l^{\prime}-1})}{p^{k-l^{\prime}-1}}\right)
Bnν+11pν(m).\displaystyle\to\frac{B_{n-\nu+1}}{1-p^{\nu}}\quad\quad(m\to\infty).

By Lemma 4.7, we have

j=0pk1Eν,k,j(0)(0,1)=(1)νpkνBnν+1nν+1\displaystyle\sum_{j=0}^{p^{k}-1}E_{\nu,k,j}^{(0){\ddagger}}\left(0,1\right)=(-1)^{\nu}p^{k\nu}\frac{B_{n-\nu+1}}{n-\nu+1} +(1)νμ=0nν(nνμ)pkμBμ+1μ+1l=0k1pl(νμ)\displaystyle+(-1)^{\nu}\sum_{\mu=0}^{n-\nu}\begin{pmatrix}n-\nu\\ \mu\end{pmatrix}p^{k\mu}\frac{B_{\mu+1}}{\mu+1}\sum_{l^{\prime}=0}^{k-1}p^{l^{\prime}(\nu-\mu)}
×(B~nνμ+1(pkl)pnνμB~nνμ+1(pkl1)).\displaystyle\quad\times\left(\widetilde{B}_{n-\nu-\mu+1}(p^{k-l^{\prime}})-p^{n-\nu-\mu}\widetilde{B}_{n-\nu-\mu+1}(p^{k-l^{\prime}-1})\right).

Since (nν)(mk)+kμ=(nνμ)(mk)+mμ(n-\nu)(m-k)+k\mu=(n-\nu-\mu)(m-k)+m\mu and 0μnν0\leq\mu\leq n-\nu, all the terms with μ1\mu\geq 1 vanish when mm\to\infty, and hence

limmk=0mp(nν)(mk)j=0pk1Eν,k,j(0)(0,1)\displaystyle\lim_{m\to\infty}\sum_{k=0}^{m}p^{(n-\nu)(m-k)}\sum_{j=0}^{p^{k}-1}E_{\nu,k,j}^{(0){\ddagger}}(0,1)
=limmk=0m(1)νp(nν)(mk)+kνBnν+1nν+1\displaystyle=\lim_{m\to\infty}\sum_{k=0}^{m}(-1)^{\nu}p^{(n-\nu)(m-k)+k\nu}\frac{B_{n-\nu+1}}{n-\nu+1}
+(1)νB1limmk=0mp(nν)(mk)l=0k1plν(B~nν+1(pkl)pnνB~nν+1(pkl1)).\displaystyle+(-1)^{\nu}B_{1}\lim_{m\to\infty}\sum_{k=0}^{m}p^{(n-\nu)(m-k)}\sum_{l^{\prime}=0}^{k-1}p^{l^{\prime}\nu}\left(\widetilde{B}_{n-\nu+1}(p^{k-l^{\prime}})-p^{n-\nu}\widetilde{B}_{n-\nu+1}(p^{k-l^{\prime}-1})\right).

Since

limmk=0mp(nν)(mk)+kν=limmp(nν)(m+1)pν(m+1)pnνpν=0,\displaystyle\lim_{m\to\infty}\sum_{k=0}^{m}p^{(n-\nu)(m-k)+k\nu}=\lim_{m\to\infty}\frac{p^{(n-\nu)(m+1)}-p^{\nu(m+1)}}{p^{n-\nu}-p^{\nu}}=0,

the first limit vanishes. Moreover, Lemma 4.11(2) implies that

k=0mp(nν)(mk)l=0k1plν(B~nν+1(pkl)pnνB~nν+1(pkl1))\displaystyle\sum_{k=0}^{m}p^{(n-\nu)(m-k)}\sum_{l^{\prime}=0}^{k-1}p^{l^{\prime}\nu}\left(\widetilde{B}_{n-\nu+1}(p^{k-l^{\prime}})-p^{n-\nu}\widetilde{B}_{n-\nu+1}(p^{k-l^{\prime}-1})\right)
=pmk=0mp(nν1)(mk)l=0k1pl(ν1)(B~nν+1(pkl)pklpnν1B~nν+1(pkl1)pkl1)\displaystyle=p^{m}\sum_{k=0}^{m}p^{(n-\nu-1)(m-k)}\sum_{l^{\prime}=0}^{k-1}p^{l^{\prime}(\nu-1)}\left(\frac{\widetilde{B}_{n-\nu+1}(p^{k-l^{\prime}})}{p^{k-l^{\prime}}}-p^{n-\nu-1}\frac{\widetilde{B}_{n-\nu+1}(p^{k-l^{\prime}-1})}{p^{k-l^{\prime}-1}}\right)
0(m),\displaystyle\to 0\quad\quad(m\to\infty),

which implies this lemma. ∎

By Lemmas 4.8, 4.10, and 4.12, we obtain the following.

Proposition 4.13.

We have

limmk=0mp(nν)(mk)\displaystyle\lim_{m\to\infty}\sum_{k=0}^{m}p^{(n-\nu)(m-k)} j=0pk1Eisn,C~ν,k,j(τpmk,pmkτ)\displaystyle\sum_{j=0}^{p^{k}-1}\langle\mathrm{Eis}_{n},\widetilde{C}_{\nu,k,j}\left(\frac{\tau}{p^{m-k}},p^{m-k}\tau\right)\rangle
=11pn+1ζ(ν)ζ(νn)ζ(1n)ζ(ν)1pnνζ(νn)1pν.\displaystyle=\frac{1}{1-p^{n+1}}\frac{\zeta(-\nu)\zeta(\nu-n)}{\zeta(-1-n)}-\frac{\zeta(-\nu)}{1-p^{n-\nu}}-\frac{\zeta(\nu-n)}{1-p^{\nu}}.

By the argument in the beginning of §4, this completes the proof of Theorem 4.1.

4.4. Rationality of the Eisenstein class Eisn\mathrm{Eis}_{n}

The computations (in the proofs of) Proposition 4.4 and Lemmas 4.6 and 4.7 imply the following proposition as a special case.

Proposition 4.14.

For any integer ν{1,,n1}\nu\in\{1,\ldots,n-1\}, we have

Eisn,Cν(τ)~=ζ(ν)ζ(νn)ζ(1n)ζ(ν)ζ(νn).\displaystyle\langle\mathrm{Eis}_{n},\widetilde{C_{\nu}(\tau)}\rangle=\frac{\zeta(-\nu)\zeta(\nu-n)}{\zeta(-1-n)}-\zeta(-\nu)-\zeta(\nu-n)\in\mathbb{Q}.
Proof.

By Proposition 4.4, we have

Eisn,Cν(τ)~=Eisn,C~ν,0,0=iν+10,0(ν)Eν,0,0(1)(0,1)Eν,0,0(0)(0,1).\langle\mathrm{Eis}_{n},\widetilde{C_{\nu}(\tau)}\rangle=\langle\mathrm{Eis}_{n},\widetilde{C}_{\nu,0,0}\rangle=i^{\nu+1}\mathscr{L}_{0,0}(\nu)-E_{\nu,0,0}^{(1){\ddagger}}\left(0,1\right)-E_{\nu,0,0}^{(0){\ddagger}}\left(0,1\right).

By Lemma 4.6(1), we have

iν+10,0(ν)=ζ(ν)ζ(νn)ζ(1n).i^{\nu+1}\mathscr{L}_{0,0}(\nu)=\frac{\zeta(-\nu)\zeta(\nu-n)}{\zeta(-1-n)}.

Moreover, in the proof of Lemma 4.7, we showed that

Eν,0,0(1)(0,1)=ζ(ν) and Eν,0,0(0)(0,1)=ζ(νn).E_{\nu,0,0}^{(1){\ddagger}}\left(0,1\right)=\zeta(-\nu)\,\,\,\textrm{ and }\,\,\,E_{\nu,0,0}^{(0){\ddagger}}\left(0,1\right)=\zeta(\nu-n).

The following lemma will be well-known to experts. For instance, Harder mentioned in [Har, §5.1.3] that this is proved by Gebertz in her diploma thesis. Here we give a proof for the completeness of the paper.

Lemma 4.15.

The relative homology group H1(YBS,YBS,n)H_{1}(Y^{\mathrm{BS}},\partial Y^{\mathrm{BS}},\mathcal{M}_{n}) is generated by the set {[Cν]0νn}\{[C_{\nu}]\mid 0\leq\nu\leq n\}.

Proof.

Note that the relative homology group H1(YBS,YBS,n)H_{1}(Y^{\mathrm{BS}},\partial Y^{\mathrm{BS}},\mathcal{M}_{n}) can be computed as

H1(YBS,YBS,n)\displaystyle H_{1}(Y^{\mathrm{BS}},\partial Y^{\mathrm{BS}},\mathcal{M}_{n})
=ker(((𝒮(BS)/𝒮(BS))n)Γ((S0(BS)/S0(BS))n)Γ).\displaystyle=\ker\left(((\mathcal{MS}(\mathbb{H}^{\mathrm{BS}})/\mathcal{MS}(\partial\mathbb{H}^{\mathrm{BS}}))\otimes\mathcal{M}_{n})_{\Gamma}\overset{\partial}{\rightarrow}((S_{0}(\mathbb{H}^{\mathrm{BS}})/S_{0}(\partial\mathbb{H}^{\mathrm{BS}}))\otimes\mathcal{M}_{n})_{\Gamma}\right).

Let [σ]H1(YBS,YBS,n)[\sigma]\in H_{1}(Y^{\mathrm{BS}},\partial Y^{\mathrm{BS}},\mathcal{M}_{n}) be a class represented by a 11-chain

σ=j{aj,bj}Pj(𝒮(BS)/𝒮(BS))n,\sigma=\sum_{j}\{a_{j},b_{j}\}\otimes P_{j}\in(\mathcal{MS}(\mathbb{H}^{\mathrm{BS}})/\mathcal{MS}(\partial\mathbb{H}^{\mathrm{BS}}))\otimes\mathcal{M}_{n},

where aj,bjBSa_{j},b_{j}\in\mathbb{H}^{\mathrm{BS}} and PjnP_{j}\in\mathcal{M}_{n}. The condition that

σ=0 in ((S0(BS)/S0(BS))n)Γ\partial\sigma=0\,\,\,\text{ in }\,\,\,((S_{0}(\mathbb{H}^{\mathrm{BS}})/S_{0}(\partial\mathbb{H}^{\mathrm{BS}}))\otimes\mathcal{M}_{n})_{\Gamma}

implies that

(4.3) j{bj}Pj{aj}Pj=k(γk1)({dk}Qk)+l{cl}Rl\displaystyle\sum_{j}\{b_{j}\}\otimes P_{j}-\{a_{j}\}\otimes P_{j}=\sum_{k}(\gamma_{k}-1)(\{d_{k}\}\otimes Q_{k})+\sum_{l}\{c_{l}\}\otimes R_{l}

in S0(BS)nS_{0}(\mathbb{H}^{\mathrm{BS}})\otimes\mathcal{M}_{n} for some γkΓ,dkBS,clBS\gamma_{k}\in\Gamma,d_{k}\in\mathbb{H}^{\mathrm{BS}},c_{l}\in\partial\mathbb{H}^{\mathrm{BS}}, and Qk,RlnQ_{k},R_{l}\in\mathcal{M}_{n}. Then we can rewrite the identity (4.3) as

τBS{τ}(jbj=τPjjaj=τPjkγkdk=τγkQk+kdk=τQklcl=τRl)=0\displaystyle\sum_{\tau\in\mathbb{H}^{\mathrm{BS}}}\{\tau\}\otimes\left(\sum_{\begin{subarray}{c}j\\ b_{j}=\tau\end{subarray}}P_{j}-\sum_{\begin{subarray}{c}j\\ a_{j}=\tau\end{subarray}}P_{j}-\sum_{\begin{subarray}{c}k\\ \gamma_{k}d_{k}=\tau\end{subarray}}\gamma_{k}Q_{k}+\sum_{\begin{subarray}{c}k\\ d_{k}=\tau\end{subarray}}Q_{k}-\sum_{\begin{subarray}{c}l\\ c_{l}=\tau\end{subarray}}R_{l}\right)=0

in S0(BS)nS_{0}(\mathbb{H}^{\mathrm{BS}})\otimes\mathcal{M}_{n}. Since S0(BS)=τBS{τ}S_{0}(\mathbb{H}^{\mathrm{BS}})=\bigoplus_{\tau\in\mathbb{H}^{\mathrm{BS}}}\mathbb{Z}\{\tau\}, this shows that for any τBS\tau\in\mathbb{H}^{\mathrm{BS}}, we have

(4.4) jbj=τPjjaj=τPjkγkdk=τγkQk+kdk=τQklcl=τRl=0 in n.\displaystyle\sum_{\begin{subarray}{c}j\\ b_{j}=\tau\end{subarray}}P_{j}-\sum_{\begin{subarray}{c}j\\ a_{j}=\tau\end{subarray}}P_{j}-\sum_{\begin{subarray}{c}k\\ \gamma_{k}d_{k}=\tau\end{subarray}}\gamma_{k}Q_{k}+\sum_{\begin{subarray}{c}k\\ d_{k}=\tau\end{subarray}}Q_{k}-\sum_{\begin{subarray}{c}l\\ c_{l}=\tau\end{subarray}}R_{l}=0\,\,\,\textrm{ in }\,\,\,\mathcal{M}_{n}.

Now, let iBSi\infty\in\partial\mathbb{H}^{\mathrm{BS}} denote the point defined by

i:=limt>0,tit.i\infty:=\lim_{t\in\mathbb{R}_{>0},t\to\infty}it.

Then the identity (4.4) implies that in 𝒮(BS)n\mathcal{MS}(\mathbb{H}^{\mathrm{BS}})\otimes\mathcal{M}_{n}, we have

(4.5) τBS{i,τ}(jbj=τPjjaj=τPjkγkdk=τγkQk+kdk=τQklcl=τRl)=0.\displaystyle\sum_{\tau\in\mathbb{H}^{\mathrm{BS}}}\{i\infty,\tau\}\otimes\left(\sum_{\begin{subarray}{c}j\\ b_{j}=\tau\end{subarray}}P_{j}-\sum_{\begin{subarray}{c}j\\ a_{j}=\tau\end{subarray}}P_{j}-\sum_{\begin{subarray}{c}k\\ \gamma_{k}d_{k}=\tau\end{subarray}}\gamma_{k}Q_{k}+\sum_{\begin{subarray}{c}k\\ d_{k}=\tau\end{subarray}}Q_{k}-\sum_{\begin{subarray}{c}l\\ c_{l}=\tau\end{subarray}}R_{l}\right)=0.

Using the identity (4.5), in ((𝒮(BS)/𝒮(BS))n)Γ((\mathcal{MS}(\mathbb{H}^{\mathrm{BS}})/\mathcal{MS}(\partial\mathbb{H}^{\mathrm{BS}}))\otimes\mathcal{M}_{n})_{\Gamma}, we compute

[σ]\displaystyle[\sigma] =j[{i,bj}Pj][{i,aj}Pj]\displaystyle=\sum_{j}[\{i\infty,b_{j}\}\otimes P_{j}]-[\{i\infty,a_{j}\}\otimes P_{j}]
=k[{i,γkdk}γkQk]k[{i,dk}Qk]+l[{i,cl}Rl]\displaystyle=\sum_{k}[\{i\infty,\gamma_{k}d_{k}\}\otimes\gamma_{k}Q_{k}]-\sum_{k}[\{i\infty,d_{k}\}\otimes Q_{k}]+\sum_{l}[\{i\infty,c_{l}\}\otimes R_{l}]
=k[(γk1)({i,dk}Qk)]+k[{i,γki}γkQk]+l[{i,cl}Rl]\displaystyle=\sum_{k}[(\gamma_{k}-1)\left(\{i\infty,d_{k}\}\otimes Q_{k}\right)]+\sum_{k}[\{i\infty,\gamma_{k}i\infty\}\otimes\gamma_{k}Q_{k}]+\sum_{l}[\{i\infty,c_{l}\}\otimes R_{l}]
=k[{i,γki}γkQk]+l[{i,cl}Rl].\displaystyle=\sum_{k}[\{i\infty,\gamma_{k}i\infty\}\otimes\gamma_{k}Q_{k}]+\sum_{l}[\{i\infty,c_{l}\}\otimes R_{l}].

Moreover, (as we are considering the relative homology classes) we may replace clc_{l} with any point in the same connected component of BS\partial\mathbb{H}^{\mathrm{BS}} as clc_{l}, and in particular, we may replace clc_{l} by glig_{l}i\infty for some glΓg_{l}\in\Gamma. Thus we conclude that the class [σ]H1(YBS,YBS,n)[\sigma]\in H_{1}(Y^{\mathrm{BS}},\partial Y^{\mathrm{BS}},\mathcal{M}_{n}) is represented by a 11-chain of the form

σ=m{i,γmi}Rm\sigma^{\prime}=\sum_{m}\{i\infty,\gamma_{m}^{\prime}i\infty\}\otimes R_{m}^{\prime}

for some γmΓ\gamma_{m}^{\prime}\in\Gamma and RmnR_{m}^{\prime}\in\mathcal{M}_{n}. Now the lemma follows from the facts that the group Γ=SL2()\Gamma=\mathrm{SL}_{2}(\mathbb{Z}) is generated by matrices (0110)\begin{pmatrix}0&-1\\ 1&0\end{pmatrix} and (1101)\begin{pmatrix}1&1\\ 0&1\end{pmatrix} and that [{i,γ1γ2i}P]=[{i,γ1i}P]+[{i,γ2i}γ11P][\{i\infty,\gamma_{1}\gamma_{2}i\infty\}\otimes P]=[\{i\infty,\gamma_{1}i\infty\}\otimes P]+[\{i\infty,\gamma_{2}i\infty\}\otimes\gamma_{1}^{-1}P] for any γ1,γ2Γ\gamma_{1},\gamma_{2}\in\Gamma and PnP\in\mathcal{M}_{n}. ∎

Recall that Γ\Gamma_{\infty} is the subgroup of Γ\Gamma generated by (1101)\begin{pmatrix}1&1\\ 0&1\end{pmatrix}. Since 1()={}\mathbb{P}^{1}(\mathbb{R})=\mathbb{R}\cup\{\infty\}, the boundary YBS\partial Y^{\mathrm{BS}} can be identified with Γ\\Gamma_{\infty}\backslash\mathbb{R}. Hence we obtain the following lemma.

Lemma 4.16.

We have an identification H0(YBS,n)=(n)ΓH_{0}(\partial Y^{\mathrm{BS}},\mathcal{M}_{n})=(\mathcal{M}_{n})_{\Gamma_{\infty}}, and hence H0(YBS,n)H_{0}(\partial Y^{\mathrm{BS}},\mathcal{M}_{n})\otimes\mathbb{Q} is a 11-dimensional \mathbb{Q}-vector space generated by [en][e_{n}], where en=X1ne_{n}=X_{1}^{n}.

Lemma 4.17.

The kernel of the boundary homomorphism

:H1(YBS,YBSn)H0(YBS,n)\partial\colon H_{1}(Y^{\mathrm{BS}},\partial Y^{\mathrm{BS}}\mathcal{M}_{n})\otimes\mathbb{Q}\longrightarrow H_{0}(\partial Y^{\mathrm{BS}},\mathcal{M}_{n})\otimes\mathbb{Q}

is generated by the set {[Cν]1νn1}\{[C_{\nu}]\mid 1\leq\nu\leq n-1\}.

Proof.

Let σker(H1(YBS,YBSn)H0(YBS,n))\sigma\in\ker(H_{1}(Y^{\mathrm{BS}},\partial Y^{\mathrm{BS}}\mathcal{M}_{n})\otimes\mathbb{Q}\longrightarrow H_{0}(\partial Y^{\mathrm{BS}},\mathcal{M}_{n})\otimes\mathbb{Q}). By Lemma4.15, we can write as σ=ν=0naν[Cν]\sigma=\sum_{\nu=0}^{n}a_{\nu}[C_{\nu}] for some rational numbers a0,,ana_{0},\dots,a_{n}\in\mathbb{Q}. Then by using Lemma 4.16, we find that

0=σ=ν=0naν[{i}(eν(1)nνenν)]=(a0an)[{i}en].0=\partial\sigma=\sum_{\nu=0}^{n}a_{\nu}[\{i\infty\}\otimes(e_{\nu}-(-1)^{n-\nu}e_{n-\nu})]=-(a_{0}-a_{n})[\{i\infty\}\otimes e_{n}].

Therefore, Lemma 4.16 shows that a0=ana_{0}=a_{n}. On the other hand we see that

[C0]=[Cν] in H1(YBS,YBS,n).[C_{0}]=-[C_{\nu}]\,\,\,\textrm{ in }\,\,\,H_{1}(Y^{\mathrm{BS}},\partial Y^{\mathrm{BS}},\mathcal{M}_{n}).

This proves the lemma. ∎

Corollary 4.18.

We have EisnH1(YBS,n)\mathrm{Eis}_{n}\in H^{1}(Y^{\mathrm{BS}},\mathcal{M}_{n})\otimes\mathbb{Q}.

Proof.

By Lemma 4.17, the homology group H1(YBS,n)H_{1}(Y^{\mathrm{BS}},\mathcal{M}_{n})\otimes\mathbb{Q} is generated by the image of H1(YBS,n)H_{1}(\partial Y^{\mathrm{BS}},\mathcal{M}_{n}) and the set {[C~ν(τ)]1νn1}\{[\widetilde{C}_{\nu}(\tau)]\mid 1\leq\nu\leq n-1\}. Therefore, by Lemma 2.8 and Proposition 4.14 we have

Eisn,H1(YBS,n),\langle\mathrm{Eis}_{n},H_{1}(Y^{\mathrm{BS}},\mathcal{M}_{n})\otimes\mathbb{Q}\rangle\subset\mathbb{Q},

which implies EisnH1(YBS,n)\mathrm{Eis}_{n}\in H^{1}(Y^{\mathrm{BS}},\mathcal{M}_{n})\otimes\mathbb{Q}. ∎

5. Denominator of an ordinary cohomology class

In order to study the denominator Δ(Eisn)\Delta(\mathrm{Eis}_{n}) of the Eisenstein class Eisn\mathrm{Eis}_{n}, in this section, we interpret the denominator Δ(Eisn)\Delta(\mathrm{Eis}_{n}) in terms of the values Eisn,[TpmCν(τ)~]\langle\mathrm{Eis}_{n},[\widetilde{T_{p}^{m}C_{\nu}(\tau)}]\rangle of the pairing between the Eisenstein class Eisn\mathrm{Eis}_{n} and the cycles [TpmCν(τ)~]H1(YBS,n,(p))[\widetilde{T_{p}^{m}C_{\nu}(\tau)}]\in H_{1}(Y^{\mathrm{BS}},\mathcal{M}_{n,(p)}).

5.1. Definition of the ordinary part

Let pp be a prime number and MM a finitely generated p\mathbb{Z}_{p}-module with an endomorphism f:MMf\colon M\longrightarrow M. In this subsection, we introduce the notion of the ff-ordinary part of MM.

Since MM is a finitely generated p\mathbb{Z}_{p}-module, the pp-adic limit

ef:=limmfm!Endp(M)e_{f}:=\lim_{m\to\infty}f^{m!}\in\mathrm{End}_{\mathbb{Z}_{p}}(M)

always exists, and ef2=efe_{f}^{2}=e_{f}. We define the ff-ordinary part MordM_{\rm ord} of MM by

Mord:=efM,M_{\rm ord}:=e_{f}M,

and we say that mMm\in M is (ff-)ordinary if mMordm\in M_{\mathrm{ord}}, that is, efm=me_{f}m=m. We also put Mnilp:=(1ef)MM_{\textrm{nilp}}:=(1-e_{f})M. We then have M=MordMnilpM=M_{\mathrm{ord}}\oplus M_{\rm nilp}.

The following lemma follows from the fact that ef2=efe_{f}^{2}=e_{f},

Lemma 5.1.

The functor MMordM\mapsto M_{\mathrm{ord}} is exact.

5.2. Denominator of a cohomology class

Recall

Hint1(YBS,n)=im(H1(YBS,n)H1(YBS,n)).H^{1}_{\rm int}(Y^{\mathrm{BS}},\mathcal{M}_{n}^{\flat})=\operatorname{im}\left(H^{1}(Y^{\mathrm{BS}},\mathcal{M}_{n}^{\flat})\longrightarrow H^{1}(Y^{\mathrm{BS}},\mathcal{M}_{n}^{\flat})\otimes\mathbb{Q}\right).
Definition 5.2.

For any cohomology class cH1(YBS,n)c\in H^{1}(Y^{\mathrm{BS}},\mathcal{M}_{n}^{\flat})\otimes\mathbb{Q}, we define the denominator Δ(c)>0\Delta(c)\in\mathbb{Z}_{>0} of cc by

Δ(c):=min{Δ>0ΔcHint1(YBS,n)},\Delta(c):=\min\{\Delta\in\mathbb{Z}_{>0}\mid\Delta c\in H^{1}_{\rm int}(Y^{\mathrm{BS}},\mathcal{M}_{n}^{\flat})\},

and for each prime number pp, we set

δp(c):=ordp(Δ(c)) and Δp(c):=pδp(c).\delta_{p}(c):=\mathrm{ord}_{p}(\Delta(c))\,\,\,\textrm{ and }\,\,\,\Delta_{p}(c):=p^{\delta_{p}(c)}.
Lemma 5.3.

Let cH1(YBS,n)c\in H^{1}(Y^{\mathrm{BS}},\mathcal{M}_{n})\otimes\mathbb{Q} be a cohomology class. We have

Δ(c)=min{Δ>0Δc,H1(YBS,n)}.\Delta(c)=\min\{\Delta\in\mathbb{Z}_{>0}\mid\Delta\langle c,H_{1}(Y^{\mathrm{BS}},\mathcal{M}_{n})\rangle\subset\mathbb{Z}\}.

Moreover, for any prime number pp, we have

δp(c)=min{δ0pδc,H1(YBS,np)p}.\delta_{p}(c)=\min\{\delta\in\mathbb{Z}_{\geq 0}\mid p^{\delta}\langle c,H_{1}(Y^{\mathrm{BS}},\mathcal{M}_{n}\otimes\mathbb{Z}_{p})\rangle\subset\mathbb{Z}_{p}\}.
Proof.

This lemma follows immediately from the formal duality (see §2.5):

H(Γ\X,n)/(torsion)Hom(H(Γ\X,n),).H^{\bullet}(\Gamma\backslash X,\mathcal{M}_{n}^{\flat})/\text{(torsion)}\overset{\sim}{\longrightarrow}\operatorname{Hom}_{\mathbb{Z}}(H_{\bullet}(\Gamma\backslash X,\mathcal{M}_{n}),\mathbb{Z}).

5.3. Denominator of an ordinary cohomology class

In this subsection, we fix a prime number pp.

Definition 5.4.
  • (1)

    We put n,p:=np\mathcal{M}_{n,p}:=\mathcal{M}_{n}\otimes\mathbb{Z}_{p} and n,p:=np\mathcal{M}_{n,p}^{\flat}:=\mathcal{M}_{n}^{\flat}\otimes\mathbb{Z}_{p}.

  • (2)

    For any finitely generated p\mathbb{Z}_{p}-algebra, we put

    Hord(YBS,n,pR)\displaystyle H_{\bullet}^{\mathrm{ord}}(Y^{\mathrm{BS}},\mathcal{M}_{n,p}\otimes R) :=eTpH(YBS,n,pR),\displaystyle:=e_{T_{p}}H_{\bullet}(Y^{\mathrm{BS}},\mathcal{M}_{n,p}\otimes R),
    Hord(YBS,YBS,n,pR)\displaystyle H_{\bullet}^{\mathrm{ord}}(Y^{\mathrm{BS}},\partial Y^{\mathrm{BS}},\mathcal{M}_{n,p}\otimes R) :=eTpH(YBS,YBS,n,pR),\displaystyle:=e_{T_{p}}H_{\bullet}(Y^{\mathrm{BS}},\partial Y^{\mathrm{BS}},\mathcal{M}_{n,p}\otimes R),
    Hord(YBS,n,pR)\displaystyle H_{\bullet}^{\mathrm{ord}}(\partial Y^{\mathrm{BS}},\mathcal{M}_{n,p}\otimes R) :=eTpH(YBS,n,pR).\displaystyle:=e_{T_{p}}H_{\bullet}(\partial Y^{\mathrm{BS}},\mathcal{M}_{n,p}\otimes R).
  • (3)

    For any finitely generated p\mathbb{Z}_{p}-algebra RR, we put

    Hord(YBS,n,pR)\displaystyle H^{\bullet}_{\mathrm{ord}}(Y^{\mathrm{BS}},\mathcal{M}_{n,p}^{\flat}\otimes R) :=eTpH(YBS,n,pR),\displaystyle:=e_{T_{p}^{\prime}}H^{\bullet}(Y^{\mathrm{BS}},\mathcal{M}_{n,p}^{\flat}\otimes R),
    Hord(YBS,YBS,n,pR)\displaystyle H^{\bullet}_{\mathrm{ord}}(Y^{\mathrm{BS}},\partial Y^{\mathrm{BS}},\mathcal{M}_{n,p}^{\flat}\otimes R) :=eTpH(YBS,YBS,n,pR),\displaystyle:=e_{T_{p}^{\prime}}H^{\bullet}(Y^{\mathrm{BS}},\partial Y^{\mathrm{BS}},\mathcal{M}_{n,p}^{\flat}\otimes R),
    Hord(YBS,n,pR)\displaystyle H^{\bullet}_{\mathrm{ord}}(\partial Y^{\mathrm{BS}},\mathcal{M}_{n,p}^{\flat}\otimes R) :=eTpH(YBS,n,pR).\displaystyle:=e_{T_{p}^{\prime}}H^{\bullet}(\partial Y^{\mathrm{BS}},\mathcal{M}_{n,p}^{\flat}\otimes R).
Lemma 5.5.

Let cH1(YBS,n,p)pc\in H^{1}(Y^{\mathrm{BS}},\mathcal{M}_{n,p})\otimes\mathbb{Q}_{p} be a cohomology class. For any homology class CH1(YBS,n,p)pC\in H_{1}(Y^{\mathrm{BS}},\mathcal{M}_{n,p})\otimes\mathbb{Q}_{p}, we have

c,eTpC=limmc,Tpm!C=limmc|Tpm!,C=eTpc,C.\langle c,e_{T_{p}}C\rangle=\lim_{m\to\infty}\langle c,T_{p}^{m!}C\rangle=\lim_{m\to\infty}\langle c|T_{p}^{\prime m!},C\rangle=\langle e_{T_{p}^{\prime}}c,C\rangle.

In particular, if cc is ordinary, that is, eTpc=ce_{T_{p}^{\prime}}c=c, then

c,C=c,eTpC,\langle c,C\rangle=\langle c,e_{T_{p}}C\rangle,

and hence

δp(c)=min{δ0pδc,H1ord(YBS,n,p)p}.\delta_{p}(c)=\min\{\delta\in\mathbb{Z}_{\geq 0}\mid p^{\delta}\langle c,H_{1}^{\mathrm{ord}}(Y^{\mathrm{BS}},\mathcal{M}_{n,p})\rangle\subset\mathbb{Z}_{p}\}.
Proof.

This lemma follows from the facts that the pairing ,\langle~{},~{}\rangle is continuous and c|Tp,C=c,TpC\langle c|T_{p}^{\prime},C\rangle=\langle c,T_{p}C\rangle. ∎

Recall the identification H0(YBS,n,p)=(n,p)ΓH_{0}(\partial Y^{\mathrm{BS}},\mathcal{M}_{n,p})=(\mathcal{M}_{n,p})_{\Gamma_{\infty}} by Lemma 4.16.

Lemma 5.6.

The p\mathbb{Z}_{p}-module H0ord(YBS,n,p)=eTp(n,p)ΓH_{0}^{\mathrm{ord}}(\partial Y^{\mathrm{BS}},\mathcal{M}_{n,p})=e_{T_{p}}(\mathcal{M}_{n,p})_{\Gamma_{\infty}} is free of rank 11 and is generated by eTp[en]e_{T_{p}}[e_{n}].

Proof.

Let k{0,,n}k\in\{0,\ldots,n\} be an integer. Since ek=X1kX2nke_{k}=X_{1}^{k}X_{2}^{n-k}, we have

Tp([ek])=pnk[ek]+j=0p1k=0k(kk)pk(j)kk[ek].T_{p}([e_{k}])=p^{n-k}[e_{k}]+\sum_{j=0}^{p-1}\sum_{k^{\prime}=0}^{k}\binom{k}{k^{\prime}}p^{k^{\prime}}(-j)^{k-k^{\prime}}[e_{k^{\prime}}].

In particular, we have Tp([e0])=(pn+p)[e0]T_{p}([e_{0}])=(p^{n}+p)[e_{0}], and hence eTp[e0]=0e_{T_{p}}[e_{0}]=0. Therefore, inductively, we obtain that eTp[ek]=0e_{T_{p}}[e_{k}]=0 for any integer k{0,,n1}k\in\{0,\ldots,n-1\} and that eTp[en]=[en]e_{T_{p}}[e_{n}]=[e_{n}], which implies that the p\mathbb{Z}_{p}-module eTp(n,p)Γe_{T_{p}}(\mathcal{M}_{n,p})_{\Gamma_{\infty}} is generated by eTp[en]e_{T_{p}}[e_{n}]. Hence by Lemma 4.16, eTp(n,p)Γe_{T_{p}}(\mathcal{M}_{n,p})_{\Gamma_{\infty}} is a free p\mathbb{Z}_{p}-module of rank 11. ∎

Recall [Cν]=[{0,i}eν]H1(YBS,YBS,n)[C_{\nu}]=[\{0,i\infty\}\otimes e_{\nu}]\in H_{1}(Y^{\mathrm{BS}},\partial Y^{\mathrm{BS}},\mathcal{M}_{n}).

Proposition 5.7.

For any integer m0m\geq 0, the p\mathbb{Z}_{p}-module H1ord(YBS,n,p)H_{1}^{\mathrm{ord}}(Y^{\mathrm{BS}},\mathcal{M}_{n,p}) is generated by (the image of) H1ord(YBS,n,p)H_{1}^{\mathrm{ord}}(\partial Y^{\mathrm{BS}},\mathcal{M}_{n,p}) and a set of lifts of eTpTpm[Cν]e_{T_{p}}T_{p}^{m}[C_{\nu}] (1νn1)(1\leq\nu\leq n-1).

Proof.

By definition, we have an exact sequence of Hecke modules

0H1(YBS,\displaystyle 0\longrightarrow H_{1}(\partial Y^{\mathrm{BS}},\, n,p)H1(Y,n,p)H1(YBS,YBS,n,p)H0(YBS,n,p)0.\displaystyle\mathcal{M}_{n,p})\longrightarrow H_{1}(Y,\mathcal{M}_{n,p})\longrightarrow H_{1}(Y^{\mathrm{BS}},\partial Y^{\mathrm{BS}},\mathcal{M}_{n,p})\stackrel{{\scriptstyle\partial}}{{\longrightarrow}}H_{0}(\partial Y^{\mathrm{BS}},\mathcal{M}_{n,p})\longrightarrow 0.

By Lemma 4.15, the ordinary part H1ord(YBS,YBS,n,p)H_{1}^{\mathrm{ord}}(Y^{\mathrm{BS}},\partial Y^{\mathrm{BS}},\mathcal{M}_{n,p}) of the relative homology group is generated by the set {eTp[Cν]0νn}\{e_{T_{p}}[C_{\nu}]\mid 0\leq\nu\leq n\}. Then by the same argument as in Lemma 4.17 using Lemma 5.6 instead of Lemma 4.16, we find that the kernel of the boundary map :H1ord(YBS,YBS,n,p)H0ord(YBS,n,p)\partial\colon H_{1}^{\mathrm{ord}}(Y^{\mathrm{BS}},\partial Y^{\mathrm{BS}},\mathcal{M}_{n,p})\longrightarrow H_{0}^{\mathrm{ord}}(\partial Y^{\mathrm{BS}},\mathcal{M}_{n,p}) is generated by the set {eTp[Cν]1νn1}\{e_{T_{p}}[C_{\nu}]\mid 1\leq\nu\leq n-1\}. Since the homomorphism

Tp:H1ord(YBS,YBS,n,p)H1ord(YBS,YBS,n,p)T_{p}\colon H_{1}^{\mathrm{ord}}(Y^{\mathrm{BS}},\partial Y^{\mathrm{BS}},\mathcal{M}_{n,p})\longrightarrow H_{1}^{\mathrm{ord}}(Y^{\mathrm{BS}},\partial Y^{\mathrm{BS}},\mathcal{M}_{n,p})

is an isomorphism and the boundary map \partial is Hecke equivariant, it follows that the kernel of the boundary map is generated by the set {eTpTpm[Cν]1νn1}\{e_{T_{p}}T_{p}^{m}[C_{\nu}]\mid 1\leq\nu\leq n-1\}. This fact together with the above exact sequence implies this proposition. ∎

Corollary 5.8.

Let mnm\geq n be an integer. For each integer ν{1,,n1}\nu\in\{1,\ldots,n-1\}, recall the cycle [TpmCν(τ)~]H1(YBS,n,(p))[\widetilde{T_{p}^{m}C_{\nu}(\tau)}]\in H_{1}(Y^{\mathrm{BS}},\mathcal{M}_{n,(p)}) defined in Definition 3.13 (see also Lemma 3.15(2)). Then for any ordinary cohomology class cHord1(YBS,n,p)pc\in H^{1}_{\rm ord}(Y^{\mathrm{BS}},\mathcal{M}_{n,p})\otimes\mathbb{Q}_{p} satisfying c,H1(YBS,n)p\langle c,H_{1}(\partial Y^{\rm BS},\mathcal{M}_{n})\rangle\subset\mathbb{Z}_{p}, we have

δp(c)=min{δ0pδc,[TpmCν(τ)~]p for any integer 1νn1}.\delta_{p}(c)=\min\{\delta\in\mathbb{Z}_{\geq 0}\mid p^{\delta}\langle c,[\widetilde{T_{p}^{m}C_{\nu}(\tau)}]\rangle\in\mathbb{Z}_{p}\textrm{ for any integer }1\leq\nu\leq n-1\}.
Proof.

By Lemma 3.15 (3), we see that eTp[TpmCν(τ)~]H1ord(YBS,n,p)e_{T_{p}}[\widetilde{T_{p}^{m}C_{\nu}(\tau)}]\in H_{1}^{\mathrm{ord}}(Y^{\mathrm{BS}},\mathcal{M}_{n,p}) maps to eTpTpm[Cν(τ)]H1ord(YBS,YBS,n,p)e_{T_{p}}T_{p}^{m}[C_{\nu}(\tau)]\in H_{1}^{\mathrm{ord}}(Y^{\mathrm{BS}},\partial Y^{\mathrm{BS}},\mathcal{M}_{n,p}) under the homomorphism

H1ord(YBS,n,p)H1ord(YBS,YBS,n,p).H_{1}^{\mathrm{ord}}(Y^{\mathrm{BS}},\mathcal{M}_{n,p})\longrightarrow H_{1}^{\mathrm{ord}}(Y^{\mathrm{BS}},\partial Y^{\mathrm{BS}},\mathcal{M}_{n,p}).

Therefore, Proposition 5.7 shows that the p\mathbb{Z}_{p}-module H1ord(YBS,n,p)H_{1}^{\mathrm{ord}}(Y^{\mathrm{BS}},\mathcal{M}_{n,p}) is generated by the image of H1ord(YBS,n)H_{1}^{\mathrm{ord}}(\partial Y^{\rm BS},\mathcal{M}_{n}) and the set {eTp[TpmCν(τ)~]1νn1}\{e_{T_{p}}[\widetilde{T_{p}^{m}C_{\nu}(\tau)}]\mid 1\leq\nu\leq n-1\}. On the other hand, by Lemma 5.5, we have

c,eTp[TpmCν(τ)~]=c,[TpmCν(τ)~].\langle c,e_{T_{p}}[\widetilde{T_{p}^{m}C_{\nu}(\tau)}]\rangle=\langle c,[\widetilde{T_{p}^{m}C_{\nu}(\tau)}]\rangle.

Now, since c,H1(YBS,n)p\langle c,H_{1}(\partial Y^{\mathrm{BS}},\mathcal{M}_{n})\rangle\subset\mathbb{Z}_{p} by assumption, Lemma 5.5 shows that

δp(c)=min{δ0pδc,[TpmCν(τ)~]p for any integer 1νn1}.\delta_{p}(c)=\min\{\delta\in\mathbb{Z}_{\geq 0}\mid p^{\delta}\langle c,[\widetilde{T_{p}^{m}C_{\nu}(\tau)}]\rangle\in\mathbb{Z}_{p}\textrm{ for any integer }1\leq\nu\leq n-1\}.

Corollary 5.9.

For any integer mnm\geq n, we have

δp(Eisn)=min{δ0pδEisn,[TpmCν(τ)~]p for any integer 1νn1}.\delta_{p}(\mathrm{Eis}_{n})=\min\{\delta\in\mathbb{Z}_{\geq 0}\mid p^{\delta}{\langle\mathrm{Eis}_{n},[\widetilde{T_{p}^{m}C_{\nu}(\tau)}]\rangle}\in\mathbb{Z}_{p}\textrm{ for any integer }1\leq\nu\leq n-1\}.
Proof.

By Lemma 2.8, we have Eisn,H1(YBS,n)p\langle\mathrm{Eis}_{n},H_{1}(\partial Y^{\rm BS},\mathcal{M}_{n})\rangle\subset\mathbb{Z}_{p} and eTpEisn=Eisne_{T_{p}^{\prime}}\mathrm{Eis}_{n}=\mathrm{Eis}_{n}, this corollary follows from Corollary 5.8. ∎

6. Relation between the denominators of the Eisenstein classes

Recall that Δp(Eisn)\Delta_{p}(\mathrm{Eis}_{n}) denotes the pp-part of the denominator Δ(Eisn)\Delta(\mathrm{Eis}_{n}) of the Eisenstein class (see Definition 5.2). In this section, we fix a prime number p5p\geq 5 and discuss another expression for the denominator Δ(Eisn)\Delta(\mathrm{Eis}_{n}) of the Eisenstein class Eisn\mathrm{Eis}_{n}. Moreover, we study a relation of the denominators Δp(Eisn)\Delta_{p}(\mathrm{Eis}_{n}) and Δp(Eisn)\Delta_{p}(\mathrm{Eis}_{n^{\prime}}) of the Eisenstein classes when nn and nn^{\prime} are pp-adically close.

6.1. Structure of the ordinary part of cohomology groups

In this subsection, we study the structure of the ordinary part of cohomology groups. Results similar to those obtained in this subsection can be found in the paper [Hid86] of Hida. In the papers [Hid86, Hid88], Hida studied the ordinary part of cohomology groups for Γ\\Gamma^{\prime}\backslash\mathbb{H} in the case that Γ/{±1}\Gamma^{\prime}/\{\pm 1\} is torsion-free. However, in the present paper we consider the group Γ=SL2()\Gamma=\mathrm{SL}_{2}(\mathbb{Z}) which has torsion elements other than ±(11)\pm\begin{pmatrix}1&\\ &1\end{pmatrix}. Hence, for the completeness of the present paper, we give the details of the proof of all the necessary facts.

Note that since we assume that p5p\geq 5, any short exact sequence of Γ\Gamma-modules induces a long exact sequence in cohomology.

Lemma 6.1.

The inclusion map n,p\longhookrightarrown,p\mathcal{M}_{n,p}^{\flat}\longhookrightarrow\mathcal{M}_{n,p} induces an isomorphism Hord(YBS,n,p)Hord(YBS,n,p)H^{\bullet}_{\mathrm{ord}}(Y^{\mathrm{BS}},\mathcal{M}_{n,p}^{\flat})\stackrel{{\scriptstyle\sim}}{{\longrightarrow}}H^{\bullet}_{\mathrm{ord}}(Y^{\mathrm{BS}},\mathcal{M}_{n,p}).

Proof.

It suffices to prove that eTpHi(YBS,n,p/n,p)=0e_{T_{p}^{\prime}}H^{i}(Y^{\mathrm{BS}},\mathcal{M}_{n,p}/\mathcal{M}_{n,p}^{\flat})=0 for any integer i0i\geq 0. Since X1n,X2nn,pX_{1}^{n},X_{2}^{n}\in\mathcal{M}_{n,p}^{\flat}, any element in n,p/n,p\mathcal{M}_{n,p}/\mathcal{M}_{n,p}^{\flat} can be represented by a polynomial of the form X1X2f(X1,X2)X_{1}X_{2}f(X_{1},X_{2}), where fn2,pf\in\mathcal{M}_{n-2,p}. Hence the fact that

(p1)~X1X2f(X1,X2)\displaystyle\widetilde{\begin{pmatrix}p&\\ &1\end{pmatrix}}\cdot X_{1}X_{2}f(X_{1},X_{2}) =pX1X2f(pX1,X2),\displaystyle=pX_{1}X_{2}f(pX_{1},X_{2}),
(1jp)~X1X2f(X1,X2)\displaystyle\widetilde{\begin{pmatrix}1&j\\ &p\end{pmatrix}}\cdot X_{1}X_{2}f(X_{1},X_{2}) =p(X1+jX2)X2f(X1+jX2,pX2)\displaystyle=p(X_{1}+jX_{2})X_{2}f(X_{1}+jX_{2},pX_{2})

shows that eTpc=0e_{T_{p}^{\prime}}c=0 for any element cHom(S(BS),n,p/n,p)c\in\operatorname{Hom}_{\mathbb{Z}}(S_{\bullet}(\mathbb{H}^{\mathrm{BS}}),\mathcal{M}_{n,p}/\mathcal{M}_{n,p}^{\flat}). In particular, we have eTpHi(YBS,n,p/n,p)=0e_{T_{p}^{\prime}}H^{i}(Y^{\mathrm{BS}},\mathcal{M}_{n,p}/\mathcal{M}_{n,p}^{\flat})=0. ∎

Thanks to Lemma 6.1, in the following, we focus on the ordinary cohomology groups with coefficient n,p\mathcal{M}_{n,p}.

Lemma 6.2.

For any polynomial f(X1,X2)n/pnf(X_{1},X_{2})\in\mathcal{M}_{n}/p\mathcal{M}_{n}, we have f|(Tp)2𝔽pX2nf|({T_{p}^{\prime}})^{2}\in\mathbb{F}_{p}X_{2}^{n}.

Proof.

By definition, we have

(f|Tp)(X1,X2)=f(0,X2)+j=0p1f(X1+jX2,0).\displaystyle(f|{T_{p}^{\prime}})(X_{1},X_{2})=f(0,X_{2})+\sum_{j=0}^{p-1}f(X_{1}+jX_{2},0).

Hence we see that (f|Tp)(X1,0)=j=0p1f(X1,0)=0(f|{T_{p}^{\prime}})(X_{1},0)=\sum_{j=0}^{p-1}f(X_{1},0)=0, and we obtain

(f|(Tp)2)(X1,X2)=(f|Tp)(0,X2)𝔽pX2n.(f|({T_{p}^{\prime}})^{2})(X_{1},X_{2})=(f|{T_{p}^{\prime}})(0,X_{2})\in\mathbb{F}_{p}X_{2}^{n}.

The boundary YBS\partial Y^{\mathrm{BS}} is of real dimension 11, and hence H2(YBS,)H^{2}(\partial Y^{\mathrm{BS}},\mathcal{M}) vanishes for any Γ\Gamma-module \mathcal{M}. Therefore, for any integer r0r\geq 0, the short exact sequence 0n,p×prn,pn,p/prn,p00\longrightarrow\mathcal{M}_{n,p}\stackrel{{\scriptstyle\times p^{r}}}{{\longrightarrow}}\mathcal{M}_{n,p}\longrightarrow\mathcal{M}_{n,p}/p^{r}\mathcal{M}_{n,p}\longrightarrow 0 induces an isomorphism

(6.1) Hord1(YBS,n,p)p/(pr)Hord1(YBS,n,p/prn,p).\displaystyle H^{1}_{\mathrm{ord}}(\partial Y^{\mathrm{BS}},\mathcal{M}_{n,p})\otimes\mathbb{Z}_{p}/(p^{r})\stackrel{{\scriptstyle\sim}}{{\longrightarrow}}H^{1}_{\mathrm{ord}}(\partial Y^{\mathrm{BS}},\mathcal{M}_{n,p}/p^{r}\mathcal{M}_{n,p}).
Lemma 6.3.

The ordinary part Hord1(YBS,n,p)H^{1}_{\mathrm{ord}}(\partial Y^{\mathrm{BS}},\mathcal{M}_{n,p}) is torsion-free.

Proof.

By using the exact sequence of M2+()M_{2}^{+}(\mathbb{Z})-modules

0n,p×pn,pn,p/pn,p0,0\longrightarrow\mathcal{M}_{n,p}\stackrel{{\scriptstyle\times p}}{{\longrightarrow}}\mathcal{M}_{n,p}\longrightarrow\mathcal{M}_{n,p}/p\mathcal{M}_{n,p}\longrightarrow 0,

we obtain an isomorphism of Hecke modules

coker(n,pΓ/pn,pΓ(n,p/pn,p)Γ)H1(YBS,n,p)[p],\operatorname{coker}\left(\mathcal{M}_{n,p}^{\Gamma_{\infty}}/p\mathcal{M}_{n,p}^{\Gamma_{\infty}}\longrightarrow(\mathcal{M}_{n,p}/p\mathcal{M}_{n,p})^{\Gamma_{\infty}}\right)\stackrel{{\scriptstyle\sim}}{{\longrightarrow}}H^{1}(\partial Y^{\mathrm{BS}},\mathcal{M}_{n,p})[p],

where for an abelian group MM we write M[p]:=ker(M×pM)M[p]:=\ker(M\stackrel{{\scriptstyle\times p}}{{\longrightarrow}}M) for the subgroup of pp-torsion elements of MM. A direct computation shows that n,pΓ=pX2n\mathcal{M}_{n,p}^{\Gamma_{\infty}}=\mathbb{Z}_{p}X_{2}^{n}. Hence Lemma 6.2 implies that Hord1(YBS,n,p)[p]=coker(𝔽pX2neTp(n,p/pn,p)Γ)=0H^{1}_{\mathrm{ord}}(\partial Y^{\mathrm{BS}},\mathcal{M}_{n,p})[p]=\operatorname{coker}\left(\mathbb{F}_{p}X_{2}^{n}\longrightarrow e_{T_{p}^{\prime}}(\mathcal{M}_{n,p}/p\mathcal{M}_{n,p})^{\Gamma_{\infty}}\right)=0. ∎

Corollary 6.4.
  • (1)

    The ordinary part Hord1(YBS,n,p)H^{1}_{\mathrm{ord}}(\partial Y^{\mathrm{BS}},\mathcal{M}_{n,p}) is a free p\mathbb{Z}_{p}-module of rank 11.

  • (2)

    We have a canonical isomorphism Hint1(YBS,n)pHord1(YBS,n,p)H^{1}_{\mathrm{int}}(\partial Y^{\mathrm{BS}},\mathcal{M}_{n})\otimes\mathbb{Z}_{p}\stackrel{{\scriptstyle\sim}}{{\longrightarrow}}H^{1}_{\mathrm{ord}}(\partial Y^{\mathrm{BS}},\mathcal{M}_{n,p}).

  • (3)

    We have c|T=(1+n+1)cc|T_{\ell}^{\prime}=(1+\ell^{n+1})c for any element cHord1(YBS,n,p)c\in H^{1}_{\mathrm{ord}}(\partial Y^{\mathrm{BS}},\mathcal{M}_{n,p}) and prime number \ell.

Proof.

By Lemma 6.3, we have a surjective homomorphism

Hint1(YBS,n)pHord1(YBS,n,p).H^{1}_{\mathrm{int}}(\partial Y^{\mathrm{BS}},\mathcal{M}_{n})\otimes\mathbb{Z}_{p}\longrightarrow H^{1}_{\mathrm{ord}}(\partial Y^{\mathrm{BS}},\mathcal{M}_{n,p}).

Lemma 2.7 shows that Hint1(YBS,n)H^{1}_{\mathrm{int}}(\partial Y^{\mathrm{BS}},\mathcal{M}_{n})\cong\mathbb{Z} and c|T=(1+n+1)cc|T_{\ell}^{\prime}=(1+\ell^{n+1})c for any element cHint1(YBS,n)c\in H^{1}_{\mathrm{int}}(\partial Y^{\mathrm{BS}},\mathcal{M}_{n}) and prime number \ell. These facts imply this corollary. ∎

Proposition 6.5.

The ordinary part Hord1(YBS,n,p)H^{1}_{\rm ord}(Y^{\mathrm{BS}},\mathcal{M}_{n,p}) is torsion-free.

Proof.

Since H0(YBS,n)=nΓ=0H^{0}(Y^{\mathrm{BS}},\mathcal{M}_{n})=\mathcal{M}_{n}^{\Gamma}=0, the exact sequence

0n,p×pn,pn,p/pn,p00\longrightarrow\mathcal{M}_{n,p}\stackrel{{\scriptstyle\times p}}{{\longrightarrow}}\mathcal{M}_{n,p}\longrightarrow\mathcal{M}_{n,p}/p\mathcal{M}_{n,p}\longrightarrow 0

implies that

Hord0(YBS,n,p/pn,p)=Hord1(YBS,n,p)[p].H^{0}_{\mathrm{ord}}(Y^{\mathrm{BS}},\mathcal{M}_{n,p}/p\mathcal{M}_{n,p})=H^{1}_{\mathrm{ord}}(Y^{\mathrm{BS}},\mathcal{M}_{n,p})[p].

Since 𝔽pX2nH0(YBS,n,p/pn,p)=(𝔽pX2n)(n,p/pn,p)Γ=0\mathbb{F}_{p}X_{2}^{n}\cap H^{0}(Y^{\mathrm{BS}},\mathcal{M}_{n,p}/p\mathcal{M}_{n,p})=(\mathbb{F}_{p}X_{2}^{n})\cap(\mathcal{M}_{n,p}/p\mathcal{M}_{n,p})^{\Gamma}=0, Lemma 6.2 shows that the module Hord0(YBS,n,p/pn,p)H^{0}_{\mathrm{ord}}(Y^{\mathrm{BS}},\mathcal{M}_{n,p}/p\mathcal{M}_{n,p}) vanishes. ∎

Lemma 6.6.

We have Hord2(YBS,YBS,n,p)=Hord2(YBS,n,p)=0H^{2}_{\mathrm{ord}}(Y^{\mathrm{BS}},\partial Y^{\mathrm{BS}},\mathcal{M}_{n,p})=H^{2}_{\mathrm{ord}}(Y^{\mathrm{BS}},\mathcal{M}_{n,p})=0.

Proof.

Since the boundary YBS\partial Y^{\mathrm{BS}} is homeomorphic to the circle, we have H2(YBS,n,p)=0H^{2}(\partial Y^{\mathrm{BS}},\mathcal{M}_{n,p})=0. Hence the canonical homomorphism Hord2(YBS,YBS,n,p)Hord2(YBS,n,p)H^{2}_{\mathrm{ord}}(Y^{\mathrm{BS}},\partial Y^{\mathrm{BS}},\mathcal{M}_{n,p})\longrightarrow H^{2}_{\mathrm{ord}}(Y^{\mathrm{BS}},\mathcal{M}_{n,p}) is surjective. Therefore, we only need to show that Hord2(YBS,YBS,n,p)=0H^{2}_{\mathrm{ord}}(Y^{\mathrm{BS}},\partial Y^{\mathrm{BS}},\mathcal{M}_{n,p})=0.

Moreover, since YY is a two-dimensional real manifold, we have H3(YBS,YBS,n,p)=0H^{3}(Y^{\mathrm{BS}},\partial Y^{\mathrm{BS}},\mathcal{M}_{n,p})=0, and hence the short exact sequence 0n,p×pn,pn,p/pn,p00\longrightarrow\mathcal{M}_{n,p}\stackrel{{\scriptstyle\times p}}{{\longrightarrow}}\mathcal{M}_{n,p}\longrightarrow\mathcal{M}_{n,p}/p\mathcal{M}_{n,p}\longrightarrow 0 induces an isomorphism

H2(YBS,YBS,n,p)/(p)H2(YBS,YBS,n,p/pn,p).H^{2}(Y^{\mathrm{BS}},\partial Y^{\mathrm{BS}},\mathcal{M}_{n,p})\otimes\mathbb{Z}/(p)\stackrel{{\scriptstyle\sim}}{{\longrightarrow}}H^{2}(Y^{\mathrm{BS}},\partial Y^{\mathrm{BS}},\mathcal{M}_{n,p}/p\mathcal{M}_{n,p}).

Therefore, it suffices to prove that Hord2(YBS,YBS,n,p/pn,p)=0H^{2}_{\mathrm{ord}}(Y^{\mathrm{BS}},\partial Y^{\mathrm{BS}},\mathcal{M}_{n,p}/p\mathcal{M}_{n,p})=0.

For notational simplicity, set ¯n,p:=n,p/pn,p\overline{\mathcal{M}}_{n,p}:=\mathcal{M}_{n,p}/p\mathcal{M}_{n,p}. Let

S2(BS)\mathscr{F}\in S_{2}(\mathbb{H}^{\mathrm{BS}})

be a representative of a fundamental class of H2(YBS,YBS,)H_{2}(Y^{\mathrm{BS}},\partial Y^{\mathrm{BS}},\mathbb{Z})\cong\mathbb{Z}. Then it is known that the homomorphism Hom(S2(BS),¯n,p)¯n,p;ϕϕ()\operatorname{Hom}_{\mathbb{Z}}(S_{2}(\mathbb{H}^{\mathrm{BS}}),\overline{\mathcal{M}}_{n,p})\longrightarrow\overline{\mathcal{M}}_{n,p};\phi\mapsto\phi(\mathscr{F}) induces an isomorphism

(6.2) ev:H2(YBS,YBS,¯n,p)(¯n,p)Γ.\displaystyle\mathrm{ev}_{\mathscr{F}}\colon H^{2}(Y^{\mathrm{BS}},\partial Y^{\mathrm{BS}},\overline{\mathcal{M}}_{n,p})\stackrel{{\scriptstyle\sim}}{{\longrightarrow}}(\overline{\mathcal{M}}_{n,p})_{\Gamma}.

See [Shi71, Proposition 8.2] or [Hid93, Proposition 1, §6.1] for example.

We will show that for any [ϕ]H2(YBS,YBS,¯n,p)[\phi]\in H^{2}(Y^{\mathrm{BS}},\partial Y^{\mathrm{BS}},\overline{\mathcal{M}}_{n,p}), we have [ϕ]|Tp=0[\phi]|_{T_{p}^{\prime}}=0. By (6.2), it suffices to show that ev([ϕ]|Tp)0\mathrm{ev}_{\mathscr{F}}([\phi]|_{T_{p}^{\prime}})\equiv 0. Here we use \equiv to emphasize that it is an identity in (¯n,p)Γ(\overline{\mathcal{M}}_{n,p})_{\Gamma}. We then compute

ev([ϕ]|Tp)\displaystyle\mathrm{ev}_{\mathscr{F}}([\phi]|_{T_{p}^{\prime}}) ϕ|Tp()(X1,X2)\displaystyle\equiv\phi|_{T_{p}^{\prime}}(\mathscr{F})(X_{1},X_{2})
=(p001)~ϕ((p001))(X1,X2)+j=0p1(1j0p)~ϕ((1j0p))(X1,X2)\displaystyle=\widetilde{\begin{pmatrix}p&0\\ 0&1\end{pmatrix}}\phi\left(\begin{pmatrix}p&0\\ 0&1\end{pmatrix}\mathscr{F}\right)(X_{1},X_{2})+\sum_{j=0}^{p-1}\widetilde{\begin{pmatrix}1&j\\ 0&p\end{pmatrix}}\phi\left(\begin{pmatrix}1&j\\ 0&p\end{pmatrix}\mathscr{F}\right)(X_{1},X_{2})
ϕ((p001))(0,X2)+j=0p1ϕ((1j0p))(X1+jX2,0).\displaystyle\equiv\phi\left(\begin{pmatrix}p&0\\ 0&1\end{pmatrix}\mathscr{F}\right)(0,X_{2})+\sum_{j=0}^{p-1}\phi\left(\begin{pmatrix}1&j\\ 0&p\end{pmatrix}\mathscr{F}\right)(X_{1}+jX_{2},0).

Put

ap:=ϕ((p001))(0,1)𝔽p,aj:=ϕ((1j0p))(1,0)𝔽p.a_{p}:=\phi\left(\begin{pmatrix}p&0\\ 0&1\end{pmatrix}\mathscr{F}\right)(0,1)\in\mathbb{F}_{p},\quad a_{j}:=\phi\left(\begin{pmatrix}1&j\\ 0&p\end{pmatrix}\mathscr{F}\right)(1,0)\in\mathbb{F}_{p}.

Then we find that

ev([ϕ]|Tp)\displaystyle\mathrm{ev}_{\mathscr{F}}([\phi]|_{T_{p}^{\prime}}) apX2n+j=0p1aj(X1+jX2)n\displaystyle\equiv a_{p}X_{2}^{n}+\sum_{j=0}^{p-1}a_{j}(X_{1}+jX_{2})^{n}
=apX2n+j=0p1aj(1101)j(0110)X2n\displaystyle=a_{p}X_{2}^{n}+\sum_{j=0}^{p-1}a_{j}\begin{pmatrix}1&1\\ 0&1\end{pmatrix}^{-j}\begin{pmatrix}0&1\\ -1&0\end{pmatrix}X_{2}^{n}
apX2n+j=0p1ajX2n\displaystyle\equiv a_{p}X_{2}^{n}+\sum_{j=0}^{p-1}a_{j}X_{2}^{n}
=(ap+j=0p1aj)((1101)1)X1X2n1\displaystyle=\left(a_{p}+\sum_{j=0}^{p-1}a_{j}\right)\left(\begin{pmatrix}1&-1\\ 0&1\end{pmatrix}-1\right)X_{1}X_{2}^{n-1}
0.\displaystyle\equiv 0.

For any Γ\Gamma-module \mathcal{M}, we define the inner cohomology H!1(YBS,)H^{1}_{!}(Y^{\mathrm{BS}},\mathcal{M}) by

H!1(YBS,):=im(H1(YBS,YBS)H1(YBS,))H^{1}_{!}(Y^{\mathrm{BS}},\mathcal{M}):=\operatorname{im}\left(H^{1}(Y^{\mathrm{BS}},\partial Y^{\mathrm{BS}}\mathcal{M})\longrightarrow H^{1}(Y^{\mathrm{BS}},\mathcal{M})\right)

and, when \mathcal{M} is a finitely generated p\mathbb{Z}_{p}-module, we put

H!,ord1(YBS,):=eTpH!1(YBS,).H^{1}_{!,\mathrm{ord}}(Y^{\mathrm{BS}},\mathcal{M}):=e_{T_{p}^{\prime}}H^{1}_{!}(Y^{\mathrm{BS}},\mathcal{M}).

Then the following corollary follows from Lemma 6.6 and the isomorphism (6.1).

Corollary 6.7.

Let rr be a non-negative integer and {n,p,n,p/prn,p}\mathcal{M}\in\{\mathcal{M}_{n,p},\mathcal{M}_{n,p}/p^{r}\mathcal{M}_{n,p}\}. Then we have a natural exact sequence of Hecke modules:

0H!,ord1(YBS,)Hord1(YBS,)Hord1(YBS,)0.\displaystyle 0\longrightarrow H^{1}_{!,\mathrm{ord}}(Y^{\mathrm{BS}},\mathcal{M})\longrightarrow H^{1}_{\mathrm{ord}}(Y^{\mathrm{BS}},\mathcal{M})\longrightarrow H^{1}_{\mathrm{ord}}(\partial Y^{\mathrm{BS}},\mathcal{M})\longrightarrow 0.
Proof.

For notational simplicity, set n,p/pr:=n,p/prn,p\mathcal{M}_{n,p}/p^{r}:=\mathcal{M}_{n,p}/p^{r}\mathcal{M}_{n,p}. Consider the natural commutative diagram

The upper row is exact by Lemma 6.6. Moreover, by (6.1), the right vertical map is surjective, and the bottom row is also exact. ∎

Corollary 6.8.

For any integer r0r\geq 0, the canonical homomorphism n,pn,p/prn,p\mathcal{M}_{n,p}\longrightarrow\mathcal{M}_{n,p}/p^{r}\mathcal{M}_{n,p} induces isomorphisms

Hord1(YBS,n,p)p/(pr)\displaystyle H^{1}_{\mathrm{ord}}(Y^{\mathrm{BS}},\mathcal{M}_{n,p})\otimes\mathbb{Z}_{p}/(p^{r}) Hord1(YBS,n,p/prn,p),\displaystyle\stackrel{{\scriptstyle\sim}}{{\longrightarrow}}H^{1}_{\mathrm{ord}}(Y^{\mathrm{BS}},\mathcal{M}_{n,p}/p^{r}\mathcal{M}_{n,p}),
H!,ord1(YBS,n,p)p/(pr)\displaystyle H^{1}_{!,\mathrm{ord}}(Y^{\mathrm{BS}},\mathcal{M}_{n,p})\otimes\mathbb{Z}_{p}/(p^{r}) H!,ord1(YBS,n,p/prn,p).\displaystyle\stackrel{{\scriptstyle\sim}}{{\longrightarrow}}H^{1}_{!,\mathrm{ord}}(Y^{\mathrm{BS}},\mathcal{M}_{n,p}/p^{r}\mathcal{M}_{n,p}).
Proof.

The exact sequence

0n,p×prn,pn,p/prn,p00\longrightarrow\mathcal{M}_{n,p}\stackrel{{\scriptstyle\times p^{r}}}{{\longrightarrow}}\mathcal{M}_{n,p}\longrightarrow\mathcal{M}_{n,p}/p^{r}\mathcal{M}_{n,p}\longrightarrow 0

shows that we have an exact sequence

0Hord1(YBS,n,p)p/(pr)Hord1(YBS,n,p/prn,p)Hord2(YBS,n,p)[pr]0.\displaystyle 0\longrightarrow H^{1}_{\mathrm{ord}}(Y^{\mathrm{BS}},\mathcal{M}_{n,p})\otimes\mathbb{Z}_{p}/(p^{r}){\longrightarrow}H^{1}_{\mathrm{ord}}(Y^{\mathrm{BS}},\mathcal{M}_{n,p}/p^{r}\mathcal{M}_{n,p})\longrightarrow H^{2}_{\mathrm{ord}}(Y^{\mathrm{BS}},\mathcal{M}_{n,p})[p^{r}]\longrightarrow 0.

Hence by Lemma 6.6, we obtain the first isomorphism.

By Corollary 6.4(1), we see that Tor1p(Hord1(YBS,n,p),p/(pr))=0\operatorname{Tor}_{1}^{\mathbb{Z}_{p}}(H^{1}_{\mathrm{ord}}(\partial Y^{\mathrm{BS}},\mathcal{M}_{n,p}),\mathbb{Z}_{p}/(p^{r}))=0, and hence Corollary 6.7 for =n,p\mathcal{M}=\mathcal{M}_{n,p} shows that

H!,ord1(YBS,n,p)p/(pr)=ker(Hord1(YBS,n,p)p/(pr)Hord1(YBS,n,p)p/(pr)).\displaystyle H^{1}_{!,\mathrm{ord}}(Y^{\mathrm{BS}},\mathcal{M}_{n,p})\otimes\mathbb{Z}_{p}/(p^{r})=\ker\left(H^{1}_{\mathrm{ord}}(Y^{\mathrm{BS}},\mathcal{M}_{n,p})\otimes\mathbb{Z}_{p}/(p^{r})\longrightarrow H^{1}_{\mathrm{ord}}(\partial Y^{\mathrm{BS}},\mathcal{M}_{n,p})\otimes\mathbb{Z}_{p}/(p^{r})\right).

Hence the second isomorphism follows from the first isomorphism, the isomorphism (6.1), and Corollary 6.7 for =n,p/prn,p\mathcal{M}=\mathcal{M}_{n,p}/p^{r}\mathcal{M}_{n,p}. ∎

Theorem 6.9.

For any positive integers rr and nn^{\prime} with nn(mod(p1)pr1)n\equiv n^{\prime}\pmod{(p-1)p^{r-1}}, we have the following canonical isomorphism of exact sequences which is TT_{\ell}^{\prime}-equivalent for any prime number p\ell\neq p:

where n,p/pr:=n,p/prn,p\mathcal{M}_{n,p}/p^{r}:=\mathcal{M}_{n,p}/p^{r}\mathcal{M}_{n,p}.

Proof.

Theorem 6.9 follows from the results proved by Hida in [Hid86] (see also [Har11]). In the following, we briefly explain how we derive Theorem 6.9 from Hida’s results in [Hid86].

First, note that since p5p\geq 5, we have canonical isomorphisms between a sheaf cohomology on YBSY^{\mathrm{BS}} and a group cohomology of Γ\Gamma:

(6.3) H1(YBS,m,p/pr)\displaystyle H^{1}(Y^{\mathrm{BS}},\mathcal{M}_{m,p}/p^{r}) H1(Γ,m,p/pr).\displaystyle\stackrel{{\scriptstyle\sim}}{{\longrightarrow}}H^{1}(\Gamma,\mathcal{M}_{m,p}/p^{r}).

Moreover, the inner cohomology group H!1(YBS,m,p/pr)H^{1}_{!}(Y^{\mathrm{BS}},\mathcal{M}_{m,p}/p^{r}) corresponds to the parabolic subgroup HP1(Γ,m,p/pr)H^{1}_{P}(\Gamma,\mathcal{M}_{m,p}/p^{r}) of H1(Γ,m,p/pr)H^{1}(\Gamma,\mathcal{M}_{m,p}/p^{r}) under the isomorphism (6.3) (see [Hid86, (4.1a)] for the definition of the parabolic subgroup).

Let m{n,n}m\in\{n,n^{\prime}\}. Hida showed in [Hid86, Proposition 4.7] that we have isomorphisms

(6.4) eTpH1(Γ,m,p/pr)eUpH1(Γ0(pr),m,p/pr);xeUpres(x),eTpHP1(Γ,m,p/pr)eUpHP1(Γ0(pr),m,p/pr);xeUpres(x)\displaystyle\begin{split}e_{T_{p}^{\prime}}H^{1}(\Gamma,\mathcal{M}_{m,p}/p^{r})&\stackrel{{\scriptstyle\sim}}{{\longrightarrow}}e_{U_{p}^{\prime}}H^{1}(\Gamma_{0}(p^{r}),\mathcal{M}_{m,p}/p^{r});x\mapsto e_{U_{p}^{\prime}}\mathrm{res}(x),\\ e_{T_{p}^{\prime}}H^{1}_{P}(\Gamma,\mathcal{M}_{m,p}/p^{r})&\stackrel{{\scriptstyle\sim}}{{\longrightarrow}}e_{U_{p}^{\prime}}H^{1}_{P}(\Gamma_{0}(p^{r}),\mathcal{M}_{m,p}/p^{r});x\mapsto e_{U_{p}^{\prime}}\mathrm{res}(x)\end{split}

which are TT_{\ell}^{\prime}-equivariant for any prime number p\ell\neq p. Here res\mathrm{res} denotes the restriction map.

Let Lm,rL_{m,r} denote the Γ0(pr)\Gamma_{0}(p^{r})-module whose underlying abelian group is p/(pr)\mathbb{Z}_{p}/(p^{r}) and the Γ0(pr)\Gamma_{0}(p^{r})-action is given by the homomorphism Γ0(pr)(p/(pr))×;(abcd)ammodpr\Gamma_{0}(p^{r})\longrightarrow(\mathbb{Z}_{p}/(p^{r}))^{\times};\begin{pmatrix}a&b\\ c&d\end{pmatrix}\mapsto a^{m}\bmod{p^{r}}. Then Hida also showed in [Hid86, Corollary 4.5 and (6.8)] that the Γ0(pr)\Gamma_{0}(p^{r})-homomorphism ir:m,p/prLm,r;f(X1,X2)f(1,0)i_{r}\colon\mathcal{M}_{m,p}/p^{r}\longrightarrow L_{m,r};f(X_{1},X_{2})\mapsto f(1,0) induces Hecke-equivariant isomorphisms

(6.5) eUpH1(Γ0(pr),m,p/pr)eUpH1(Γ0(pr),Lm,r);xir,(x),eUpHP1(Γ0(pr),m,p/pr)eUpHP1(Γ0(pr),Lm,r);xir,(x).\displaystyle\begin{split}e_{U_{p}^{\prime}}H^{1}(\Gamma_{0}(p^{r}),\mathcal{M}_{m,p}/p^{r})&\stackrel{{\scriptstyle\sim}}{{\longrightarrow}}e_{U_{p}^{\prime}}H^{1}(\Gamma_{0}(p^{r}),L_{m,r});x\mapsto i_{r,*}(x),\\ e_{U_{p}^{\prime}}H^{1}_{P}(\Gamma_{0}(p^{r}),\mathcal{M}_{m,p}/p^{r})&\stackrel{{\scriptstyle\sim}}{{\longrightarrow}}e_{U_{p}^{\prime}}H^{1}_{P}(\Gamma_{0}(p^{r}),L_{m,r});x\mapsto i_{r,*}(x).\end{split}

Since nn(mod(p1)pr)n\equiv n^{\prime}\pmod{(p-1)p^{r}}, we have Ln,r=Ln,rL_{n,r}=L_{n^{\prime},r} as Γ0(pr)\Gamma_{0}(p^{r})-modules, by combining the isomorphisms (6.4) and (6.5) for m=n,nm=n,n^{\prime}, we obtain the following commutative diagram:

where horizontal arrows are isomorphisms and TT_{\ell}^{\prime}-equivariant for any prime number p\ell\neq p. This completes the proof. ∎

6.2. Another expression for Δp(Eisn)\Delta_{p}(\mathrm{Eis}_{n})

Let p5p\geq 5 be a prime number and take a prime number p\ell\neq p. Let

,p:=p[X]\mathcal{H}_{\ell,p}:=\mathbb{Z}_{p}[X]

be the polynomial ring over p\mathbb{Z}_{p}, and by using the Hecke operator TT_{\ell}^{\prime} at \ell, we regard cohomology groups that appear in the present paper as ,p\mathcal{H}_{\ell,p}-modules. For notational simplicity, we put

x,n:=X(1+n+1) and ,p,n:=,p/(x,n).x_{\ell,n}:=X-(1+\ell^{n+1})\,\,\,\textrm{ and }\,\,\,\mathcal{B}_{\ell,p,n}:=\mathcal{H}_{\ell,p}/(x_{\ell,n}).

Note that by Corollary 6.4, we have ,p,nHord1(YBS,n,p);1eTp[en]\mathcal{B}_{\ell,p,n}\stackrel{{\scriptstyle\sim}}{{\longrightarrow}}H^{1}_{\mathrm{ord}}(\partial Y^{\mathrm{BS}},\mathcal{M}_{n,p});1\mapsto e_{T_{p}^{\prime}}[e_{n}] as ,p\mathcal{H}_{\ell,p}-modules.

Lemma 6.10.

Let cH1(YBS,n)c\in H^{1}(Y^{\mathrm{BS}},\mathcal{M}_{n})\otimes\mathbb{C} be a cohomology class. If c|T=(1+n+1)cc|T_{\ell}^{\prime}=(1+\ell^{n+1})c, then we have c|T=(1+n+1)cc|T_{\ell^{\prime}}^{\prime}=(1+\ell^{\prime n+1})c for any prime number \ell^{\prime}, that is, the cohomology class cc is a scalar multiple of Eisn\mathrm{Eis}_{n}.

Proof.

It is well-known that one can take a TT_{\ell}^{\prime}-Hecke-eigen basis f1,,ftH1(YBS,n)f_{1},\ldots,f_{t}\in H^{1}(Y^{\mathrm{BS}},\mathcal{M}_{n})\otimes\mathbb{C} such that f1=Eisnf_{1}=\mathrm{Eis}_{n} and that the elements f2,,ftf_{2},\ldots,f_{t} correspond to either cusp forms or their complex conjugates via the Eichler–Shimura homomorphism. Then the Ramanujan conjecture proved by Deligne shows that the absolute value of the TT_{\ell}^{\prime}-eigenvalue of fif_{i} (2it2\leq i\leq t) is less than 1+n+11+\ell^{n+1}, which implies this lemma. ∎

Lemma 6.11.

We have Hord1(YBS,n,p)[x,n]=pΔp(Eisn)EisnH^{1}_{\mathrm{ord}}(Y^{\mathrm{BS}},\mathcal{M}_{n,p})[x_{\ell,n}]=\mathbb{Z}_{p}\Delta_{p}(\mathrm{Eis}_{n})\mathrm{Eis}_{n}.

Proof.

By Proposition 6.5 and Lemma 6.10, we have Hord1(YBS,n,p)[x,n]pEisnH^{1}_{\mathrm{ord}}(Y^{\mathrm{BS}},\mathcal{M}_{n,p})[x_{\ell,n}]\subset\mathbb{Q}_{p}\cdot\mathrm{Eis}_{n}. Hence this lemma follows from the definition of the denominator Δp(Eisn)\Delta_{p}(\mathrm{Eis}_{n}) of the Eisenstein class Eisn\mathrm{Eis}_{n} and Lemmas 5.5 and 6.1. ∎

Definition 6.12.

We define [,p,n]Ext,p1(,p,n,H!,ord1(YBS,n,p))[\mathcal{E}_{\ell,p,n}]\in\operatorname{Ext}^{1}_{\mathcal{H}_{\ell,p}}(\mathcal{B}_{\ell,p,n},H^{1}_{!,\mathrm{ord}}(Y^{\mathrm{BS}},\mathcal{M}_{n,p})) to be the element corresponding to the exact sequence of ,p\mathcal{H}_{\ell,p}-modules in Corollary 6.7 for =n,p\mathcal{M}=\mathcal{M}_{n,p}:

0H!,ord1(YBS,n,p)Hord1(YBS,n,p),p,n0.0\longrightarrow H^{1}_{!,\mathrm{ord}}(Y^{\mathrm{BS}},\mathcal{M}_{n,p})\longrightarrow H^{1}_{\mathrm{ord}}(Y^{\mathrm{BS}},\mathcal{M}_{n,p})\longrightarrow\mathcal{B}_{\ell,p,n}\longrightarrow 0.

The following lemma follows directly from Lemma 6.11.

Lemma 6.13.

Annp([,p,n])=Δp(Eisn)p\operatorname{Ann}_{\mathbb{Z}_{p}}([\mathcal{E}_{\ell,p,n}])=\Delta_{p}(\mathrm{Eis}_{n})\mathbb{Z}_{p}.

Lemma 6.14.

We have a natural identification

H!,ord1(YBS,n,p),p,p,n=Ext,p1(,p,n,H!,ord1(YBS,n,p)).H^{1}_{!,\mathrm{ord}}(Y^{\mathrm{BS}},\mathcal{M}_{n,p})\otimes_{\mathcal{H}_{\ell,p}}\mathcal{B}_{\ell,p,n}=\operatorname{Ext}^{1}_{\mathcal{H}_{\ell,p}}(\mathcal{B}_{\ell,p,n},H^{1}_{!,\mathrm{ord}}(Y^{\mathrm{BS}},\mathcal{M}_{n,p})).
Proof.

Since x,nx_{\ell,n} is an regular element of ,p\mathcal{H}_{\ell,p}, we have an exact sequence of ,p\mathcal{H}_{\ell,p}-modules:

0,p×x,p,p,n0.0\longrightarrow\mathcal{H}_{\ell,p}\xrightarrow{\times x_{\ell}}\mathcal{H}_{\ell,p}\longrightarrow\mathcal{B}_{\ell,p,n}\longrightarrow 0.

Applying the functor Hom,p(,H!,ord1(YBS,n,p))\operatorname{Hom}_{\mathcal{H}_{\ell,p}}(-,H^{1}_{!,\mathrm{ord}}(Y^{\mathrm{BS}},\mathcal{M}_{n,p})) to this short exact sequence, we obtain the desired identification. ∎

Lemma 6.15.

For any positive integers rr and nn^{\prime} with nn(mod(p1)pr1)n\equiv n^{\prime}\pmod{(p-1)p^{r-1}}, we have a natural isomorphism of ,p\mathcal{H}_{\ell,p}-modules:

H!,ord1(YBS,n,p),p,p,n/(pr)H!,ord1(YBS,n,p),p,p,n/(pr).H^{1}_{!,\mathrm{ord}}(Y^{\mathrm{BS}},\mathcal{M}_{n,p})\otimes_{\mathcal{H}_{\ell,p}}\mathcal{B}_{\ell,p,n}/(p^{r})\cong H^{1}_{!,\mathrm{ord}}(Y^{\mathrm{BS}},\mathcal{M}_{n^{\prime},p})\otimes_{\mathcal{H}_{\ell,p}}\mathcal{B}_{\ell,p,n^{\prime}}/(p^{r}).

Moreover, the image of [,p,n]modpr[\mathcal{E}_{\ell,p,n}]\bmod{p^{r}} is [,p,n]modpr[\mathcal{E}_{\ell,p,n^{\prime}}]\bmod{p^{r}} under this isomorphism (and the identification in Lemma 6.14).

Proof.

This lemma follows from Theorem 6.9 and Corollary 6.8. ∎

Definition 6.16.

We define a polynomial Φ,n(t)p[t]\Phi_{\ell,n}(t)\in\mathbb{Z}_{p}[t] to be the characteristic polynomial associated with T:H!,ord1(YBS,n,p)H!,ord1(YBS,n,p)T_{\ell}^{\prime}\colon H^{1}_{!,\mathrm{ord}}(Y^{\mathrm{BS}},\mathcal{M}_{n,p})\longrightarrow H^{1}_{!,\mathrm{ord}}(Y^{\mathrm{BS}},\mathcal{M}_{n,p}):

Φ,n(t):=det(tidTH!,ord1(YBS,n,p)).\Phi_{\ell,n}(t):=\det(t\cdot\mathrm{id}-T_{\ell}^{\prime}\mid H^{1}_{!,\mathrm{ord}}(Y^{\mathrm{BS}},\mathcal{M}_{n,p})).
Lemma 6.17.

The p\mathbb{Z}_{p}-module H!,ord1(YBS,n,p),p,p,nH^{1}_{!,\mathrm{ord}}(Y^{\mathrm{BS}},\mathcal{M}_{n,p})\otimes_{\mathcal{H}_{\ell,p}}\mathcal{B}_{\ell,p,n} is annihilated by Φ,n(1+n+1)\Phi_{\ell,n}(1+\ell^{n+1}).

Proof.

By the Cayley–Hamilton theorem, the Hecke module H!,ord1(YBS,n,p)H^{1}_{!,\mathrm{ord}}(Y^{\mathrm{BS}},\mathcal{M}_{n,p}) is annihilated by Φ,n(T)\Phi_{\ell,n}(T_{\ell}^{\prime}). Hence the p\mathbb{Z}_{p}-module H!,ord1(YBS,n,p),p,p,nH^{1}_{!,\mathrm{ord}}(Y^{\mathrm{BS}},\mathcal{M}_{n,p})\otimes_{\mathcal{H}_{\ell,p}}\mathcal{B}_{\ell,p,n} is annihilated by Φ,n(1+n+1)\Phi_{\ell,n}(1+\ell^{n+1}) since ,p,n=,p/(X(1+1+n))\mathcal{B}_{\ell,p,n}=\mathcal{H}_{\ell,p}/(X-(1+\ell^{1+n})). ∎

Lemma 6.18.

Let rr be a positive integer satisfying r>ordp(Φ,n(1+n+1))r>\mathrm{ord}_{p}(\Phi_{\ell,n}(1+\ell^{n+1})). Then for any even integer n2n^{\prime}\geq 2 with nn(mod(p1)pr1)n\equiv n^{\prime}\pmod{(p-1)p^{r-1}}, we have

ordp(Φ,n(1+n+1))=ordp(Φ,n(1+n+1)).\mathrm{ord}_{p}(\Phi_{\ell,n}(1+\ell^{n+1}))=\mathrm{ord}_{p}(\Phi_{\ell,n^{\prime}}(1+\ell^{n^{\prime}+1})).
Proof.

By Theorem 6.9, we have

Φ,n(t)Φ,n(t)(modpr).\Phi_{\ell,n}(t)\equiv\Phi_{\ell,n^{\prime}}(t)\pmod{p^{r}}.

The fact that nn(mod(p1)pr1)n\equiv n^{\prime}\pmod{(p-1)p^{r-1}} implies that n+1n+1(modpr)\ell^{n+1}\equiv\ell^{n^{\prime}+1}\pmod{p^{r}}, and we obtain Φ,n(1+n+1)Φ,n(1+n+1)(modpr)\Phi_{\ell,n}(1+\ell^{n+1})\equiv\Phi_{\ell,n^{\prime}}(1+\ell^{n^{\prime}+1})\pmod{p^{r}}. Hence the assumption that r>ordp(Φ,n(1+n+1))r>\mathrm{ord}_{p}(\Phi_{\ell,n}(1+\ell^{n+1})) shows that ordp(Φ,n(1+n+1))=ordp(Φ,n(1+n+1))\mathrm{ord}_{p}(\Phi_{\ell,n}(1+\ell^{n+1}))=\mathrm{ord}_{p}(\Phi_{\ell,n^{\prime}}(1+\ell^{n^{\prime}+1})). ∎

Proposition 6.19.

Let rr be a positive integer satisfying r>ordp(Φ,n(1+n+1))r>\mathrm{ord}_{p}(\Phi_{\ell,n}(1+\ell^{n+1})). Then for any even integer n2n^{\prime}\geq 2 with nn(mod(p1)pr1)n\equiv n^{\prime}\pmod{(p-1)p^{r-1}}, we have

Δp(Eisn)=Δp(Eisn).\Delta_{p}(\mathrm{Eis}_{n})=\Delta_{p}(\mathrm{Eis}_{n^{\prime}}).
Proof.

By Lemmas 6.15, 6.17, and 6.18, we have the natural isomorphism

H!,ord1(YBS,n,p),p,p,n\displaystyle H^{1}_{!,\mathrm{ord}}(Y^{\mathrm{BS}},\mathcal{M}_{n,p})\otimes_{\mathcal{H}_{\ell,p}}\mathcal{B}_{\ell,p,n} =H!,ord1(YBS,n,p),p,p,n/(pr)\displaystyle=H^{1}_{!,\mathrm{ord}}(Y^{\mathrm{BS}},\mathcal{M}_{n,p})\otimes_{\mathcal{H}_{\ell,p}}\mathcal{B}_{\ell,p,n}/(p^{r})
H!,ord1(YBS,n,p),p,p,n/(pr)\displaystyle\cong H^{1}_{!,\mathrm{ord}}(Y^{\mathrm{BS}},\mathcal{M}_{n^{\prime},p})\otimes_{\mathcal{H}_{\ell,p}}\mathcal{B}_{\ell,p,n^{\prime}}/(p^{r})
=H!,ord1(YBS,n,p),p,p,n.\displaystyle=H^{1}_{!,\mathrm{ord}}(Y^{\mathrm{BS}},\mathcal{M}_{n^{\prime},p})\otimes_{\mathcal{H}_{\ell,p}}\mathcal{B}_{\ell,p,n^{\prime}}.

Moreover, the image of [,p,n][\mathcal{E}_{\ell,p,n}] under this isomorphism is [,p,n][\mathcal{E}_{\ell,p,n^{\prime}}], and Lemma 6.13 implies that

Δp(Eisn)p=Annp([,p,n])=Annp([,p,n])=Δp(Eisn)p.\displaystyle\Delta_{p}(\mathrm{Eis}_{n})\mathbb{Z}_{p}=\operatorname{Ann}_{\mathbb{Z}_{p}}([\mathcal{E}_{\ell,p,n}])=\operatorname{Ann}_{\mathbb{Z}_{p}}([\mathcal{E}_{\ell,p,n^{\prime}}])=\Delta_{p}(\mathrm{Eis}_{n^{\prime}})\mathbb{Z}_{p}.

7. Kubota–Leopoldt pp-adic LL-function

Let pp be a prime number. In this section, we introduce the Kubota–Leopoldt pp-adic LL-functions and prove certain congruence properties that will be used in the proof of Theorem 2.13.

Let ω:Gal(¯/)p×\omega\colon\operatorname{Gal}(\overline{\mathbb{Q}}/\mathbb{Q})\longrightarrow\mathbb{Z}_{p}^{\times} denote the Teichmüller character, and let 𝟙\mathbbm{1} denote the trivial character. For any Dirichlet character χ\chi, we denote by Lp(s,χ)p[[s]]L_{p}(s,\chi)\in\mathbb{C}_{p}[[s]] the Kubota–Leopoldt pp-adic LL-function attached to χ\chi.

Proposition 7.1 ([Was97, Theorems 5.11 and 5.12, Exercises 5.11(1)] ).
  • (1)

    For any Dirichlet character χ\chi, the pp-adic LL-function Lp(s,χ)L_{p}(s,\chi) converges on p{1}\mathbb{Z}_{p}\!-\!\{1\}. Moreover, for any integer m2m\geq 2, we have

    Lp(1m,χ)=(1χωm(p)pm1)L(1m,χωm).L_{p}(1-m,\chi)=(1-\chi\omega^{-m}(p)p^{m-1})L(1-m,\chi\omega^{-m}).

    In particular, we have ordp(Lp(1m,ωm))=ordp(ζ(1m))\mathrm{ord}_{p}(L_{p}(1-m,\omega^{m}))=\mathrm{ord}_{p}(\zeta(1-m)).

  • (2)

    We have

    Lp(s,𝟙)p1p(s1)+p[[s1]]L_{p}(s,\mathbbm{1})\in\frac{p-1}{p(s-1)}+\mathbb{Z}_{p}[[s-1]]
  • (3)

    If m0(modp1)m\not\equiv 0\pmod{p-1}, then we have

    Lp(s,ωm)p+pp[[s1]].L_{p}(s,\omega^{m})\in\mathbb{Z}_{p}+p\mathbb{Z}_{p}[[s-1]].

By using the Kubota–Leopoldt pp-adic LL-functions, Theorem 4.1 can be restated as follows.

Corollary 7.2.

If we put

Dp(n,ν):=Lp(ν,ω1+ν)Lp(νn,ωnν+1)Lp(1n,ωn+2)Lp(ν,ω1+ν)Lp(νn,ωnν+1),\displaystyle D_{p}(n,\nu):=\frac{L_{p}(-\nu,\omega^{1+\nu})L_{p}(\nu-n,\omega^{n-\nu+1})}{L_{p}(-1-n,\omega^{n+2})}-L_{p}(-\nu,\omega^{1+\nu})-L_{p}(\nu-n,\omega^{n-\nu+1}),

then for any integer ν{1,,n1}\nu\in\{1,\ldots,n-1\} we have

limmEisn,Tpm!(Cν(τ))~=1pn+1(1pν)(1pnν)Dp(n,ν).\displaystyle\lim_{m\to\infty}\langle\mathrm{Eis}_{n},\widetilde{T_{p}^{m!}(C_{\nu}(\tau))}\rangle=\frac{1-p^{n+1}}{(1-p^{\nu})(1-p^{n-\nu})}D_{p}(n,\nu).

For any even integer mm, we define a positive integer NmN_{m} by

Nm:=the numerator of ζ(1m).N_{m}:=\textrm{the numerator of $\zeta(1-m)$}.
Corollary 7.3.

Let m2m\geq 2 be an even integer.

  • (1)

    If m0(modp1)m\not\equiv 0\pmod{p-1}, then we have ordp(ζ(1m))=ordp(Nm)\mathrm{ord}_{p}(\zeta(1-m))=\mathrm{ord}_{p}(N_{m}).

  • (2)

    If m0(modp1)m\equiv 0\pmod{p-1}, then we have ordp(Nm)=0\mathrm{ord}_{p}(N_{m})=0.

  • (3)

    Let rr and mm^{\prime} be positive integers with mm(mod(p1)pr1)m\equiv m^{\prime}\pmod{(p-1)p^{r-1}}. If r>ordp(Nm)r>\mathrm{ord}_{p}(N_{m}), then

    ordp(Nm)=ordp(Nm).\mathrm{ord}_{p}(N_{m})=\mathrm{ord}_{p}(N_{m^{\prime}}).
Proof.

Claims (1) and (2) follows immediately from Proposition 7.1. When m0(modp1)m\equiv 0\pmod{p-1}, claim (3) follows from claim (2). When m0(modp1)m\not\equiv 0\pmod{p-1}, claim (3) follows from Proposition 7.1. ∎

Corollary 7.4.

Let xx be an integer with x0(modp1)x\not\equiv 0\pmod{p-1}. For any integer yy, we have

Lp(1x,ωx)Lp(1y,𝟙)Lp(1xy,ωx)Lp(1y,𝟙)+pLp(1xy,ωx).\displaystyle\frac{L_{p}(1-x,\omega^{x})L_{p}(1-y,\mathbbm{1})}{L_{p}(1-x-y,\omega^{x})}\in L_{p}(1-y,\mathbbm{1})+\frac{\mathbb{Z}_{p}}{L_{p}(1-x-y,\omega^{x})}.
Proof.

Since x0(modp1)x\not\equiv 0\pmod{p-1}, by Proposition 7.1(3), we have

Lp(1x,ωx)Lp(1xy,ωx)+pyp.L_{p}(1-x,\omega^{x})\in L_{p}(1-x-y,\omega^{x})+py\mathbb{Z}_{p}.

Moreover, by Proposition 7.1(2), we have pyLp(1y,𝟙)1+pppyL_{p}(1-y,\mathbbm{1})\in 1+p\mathbb{Z}_{p}, which shows

Lp(1x,ωx)Lp(1y,𝟙)Lp(1xy,ωx)Lp(1y,𝟙)+pLp(1xy,ωx).\displaystyle\frac{L_{p}(1-x,\omega^{x})L_{p}(1-y,\mathbbm{1})}{L_{p}(1-x-y,\omega^{x})}\in L_{p}(1-y,\mathbbm{1})+\frac{\mathbb{Z}_{p}}{L_{p}(1-x-y,\omega^{x})}.

Corollary 7.5.

For any integers xx and yy, we have

Lp(1x,𝟙)Lp(1y,𝟙)Lp(1xy,𝟙)\displaystyle\frac{L_{p}(1-x,\mathbbm{1})L_{p}(1-y,\mathbbm{1})}{L_{p}(1-x-y,\mathbbm{1})} p1px+p1py(modp)\displaystyle\equiv\frac{p-1}{px}+\frac{p-1}{py}\pmod{\mathbb{Z}_{p}}
Lp(1x,𝟙)+Lp(1y,𝟙)(modp).\displaystyle\equiv L_{p}(1-x,\mathbbm{1})+L_{p}(1-y,\mathbbm{1})\pmod{\mathbb{Z}_{p}}.
Proof.

For notational simplicity, we put

R(s):=p1ps and H(s1):=Lp(s,𝟙)R(1s).R(s):=-\frac{p-1}{ps}\,\,\,\textrm{ and }\,\,\,H(s-1):=L_{p}(s,\mathbbm{1})-R(1-s).

By Proposition 7.1(2), we have H(s)p[[s]]H(s)\in\mathbb{Z}_{p}[[s]], xR(x),yR(y),(x+y)R(x+y)p1p×xR(x),yR(y),(x+y)R(x+y)\in p^{-1}\mathbb{Z}_{p}^{\times}, and R(x+y)1=R(x)1+R(y)1R(x+y)^{-1}=R(x)^{-1}+R(y)^{-1}. Since H(s)p[[s]]H(s)\in\mathbb{Z}_{p}[[s]], we have

H(x)H(x+y)+yp,H(y)H(x+y)+xp.\displaystyle H(x)\in H(x+y)+y\mathbb{Z}_{p},\quad H(y)\in H(x+y)+x\mathbb{Z}_{p}.

Put α:=1+R(x+y)1H(x+y)1+pp\alpha:=1+R(x+y)^{-1}H(x+y)\in 1+p\mathbb{Z}_{p}. Then

Lp(1x,𝟙)Lp(1y,𝟙)Lp(1xy,𝟙)(R(x)+H(x+y)+yp)(R(y)+H(x+y)+xp)R(x+y)+H(x+y)\displaystyle\frac{L_{p}(1-x,\mathbbm{1})L_{p}(1-y,\mathbbm{1})}{L_{p}(1-x-y,\mathbbm{1})}\in\frac{(R(x)+H(x+y)+y\mathbb{Z}_{p})(R(y)+H(x+y)+x\mathbb{Z}_{p})}{R(x+y)+H(x+y)}

and we have

(R(x)+H(x+y)+yp)(R(y)+H(x+y)+xp)R(x+y)+H(x+y)\displaystyle\quad\frac{(R(x)+H(x+y)+y\mathbb{Z}_{p})(R(y)+H(x+y)+x\mathbb{Z}_{p})}{R(x+y)+H(x+y)}
R(x)1+R(y)1α(R(x)R(y)+(R(x)+R(y))H(x+y)+p1p)\displaystyle\subset\frac{R(x)^{-1}+R(y)^{-1}}{\alpha}(R(x)R(y)+(R(x)+R(y))H(x+y)+p^{-1}\mathbb{Z}_{p})
=R(x)+R(y)+p.\displaystyle=R(x)+R(y)+\mathbb{Z}_{p}.

8. Proof of Theorem 2.13

Let pp be a prime number. As in Corollary 7.2, for any integer 1νn11\leq\nu\leq n-1, we define

Dp(n,ν):=Lp(ν,ω1+ν)Lp(νn,ωnν+1)Lp(1n,ωn+2)Lp(ν,ω1+ν)Lp(νn,ωnν+1)\displaystyle D_{p}(n,\nu):=\frac{L_{p}(-\nu,\omega^{1+\nu})L_{p}(\nu-n,\omega^{n-\nu+1})}{L_{p}(-1-n,\omega^{n+2})}-L_{p}(-\nu,\omega^{1+\nu})-L_{p}(\nu-n,\omega^{n-\nu+1})

and set

δp(n,ν):=max{ordp(Dp(n,ν)),0}.\displaystyle\delta_{p}(n,\nu):=\max\left\{-\mathrm{ord}_{p}(D_{p}(n,\nu)),0\right\}.
Proposition 8.1.

We have

δp(Eisn)=max1νn1δp(n,ν).\delta_{p}(\mathrm{Eis}_{n})=\max_{1\leq\nu\leq n-1}\delta_{p}(n,\nu).
Proof.

By Corollary 7.2, we have

limmEisn,Tpm!(Cν(τ))~=1pn+1(1pν)(1pnν)Dp(n,ν).\displaystyle\lim_{m\to\infty}\langle\mathrm{Eis}_{n},\widetilde{T_{p}^{m!}(C_{\nu}(\tau))}\rangle=\frac{1-p^{n+1}}{(1-p^{\nu})(1-p^{n-\nu})}D_{p}(n,\nu).

Hence, for any sufficiently large integer mm, we have

ordp(Eisn,Tpm!(Cν(τ))~)=ordp(Dp(n,ν)),\mathrm{ord}_{p}\left(\langle\mathrm{Eis}_{n},\widetilde{T_{p}^{m!}(C_{\nu}(\tau))}\rangle\right)=\mathrm{ord}_{p}(D_{p}(n,\nu)),

and this proposition follows from Corollary 5.9. ∎

Recall that NmN_{m} denotes the numerator of ζ(1m)\zeta(1-m).

Proposition 8.2.

Let pp be a prime number.

  • (1)

    δp(Eisn)ordp(Nn+2)\delta_{p}(\mathrm{Eis}_{n})\leq\mathrm{ord}_{p}(N_{n+2}).

  • (2)

    If p<np<n, then δp(Eisn)=ordp(Nn+2)\delta_{p}(\mathrm{Eis}_{n})=\mathrm{ord}_{p}(N_{n+2}).

The proof of Proposition 8.2 is given in §8.1. First, we give the proof of Theorem 2.13 assuming Proposition 8.2, that is, we show that Δ(Eisn)=Nn+2\Delta(\mathrm{Eis}_{n})=N_{n+2}.

Proof of Theorem 2.13.

Take a prime number pp. It suffices to show that δp(Eisn)=ordp(Nn+2)\delta_{p}(\mathrm{Eis}_{n})=\mathrm{ord}_{p}(N_{n+2}). When p1n+2p-1\mid n+2, by Proposition 8.2 we have 0δp(Eisn)ordp(Nn+2)=00\leq\delta_{p}(\mathrm{Eis}_{n})\leq\mathrm{ord}_{p}(N_{n+2})=0, and hence we may assume that n2(modp1)n\not\equiv-2\pmod{p-1}. Note that p5p\geq 5 in this case. Take a prime number p\ell\neq p, and positive integers rr and nn^{\prime} satisfying

  • r>ordp(Φ,n(1+n))r>\mathrm{ord}_{p}(\Phi_{\ell,n}(1+\ell^{n})),

  • p<np<n^{\prime},

  • nn(modpr1(p1))n\equiv n^{\prime}\pmod{p^{r-1}(p-1)}.

Then by Propositions 8.2(2), we have δp(Eisn)=ordp(Nn+2)\delta_{p}(\mathrm{Eis}_{n^{\prime}})=\mathrm{ord}_{p}(N_{n^{\prime}+2}), and Proposition 6.19 implies that

δp(Eisn)=δp(Eisn)=ordp(Nn+2).\delta_{p}(\mathrm{Eis}_{n})=\delta_{p}(\mathrm{Eis}_{n^{\prime}})=\mathrm{ord}_{p}(N_{n^{\prime}+2}).

Hence Corollary 7.3 shows that δp(Eisn)=ordp(Nn+2)=ordp(Nn+2)\delta_{p}(\mathrm{Eis}_{n})=\mathrm{ord}_{p}(N_{n^{\prime}+2})=\mathrm{ord}_{p}(N_{n+2}). ∎

8.1. Proof of Proposition 8.2

In this subsection, we prove Proposition 8.2. The proof is divided into the following two cases:

  • p1n+2p-1\nmid n+2,

  • p1n+2p-1\mid n+2.

8.1.1. p1n+2p-1\nmid n+2

Lemma 8.3.

If p1n+2p-1\nmid n+2, then we have δp(Eisn)ordp(Nn+2)\delta_{p}(\mathrm{Eis}_{n})\leq\mathrm{ord}_{p}(N_{n+2}).

Proof.

Take an integer ν{1,,n1}\nu\in\{1,\ldots,n-1\}. When p11+νp-1\nmid 1+\nu and p1nν+1p-1\nmid n-\nu+1, both Lp(ν,ω1+ν)L_{p}(-\nu,\omega^{1+\nu}) and Lp(νn,ωnν+1)L_{p}(\nu-n,\omega^{n-\nu+1}) are pp-adic integers, and hence we have

δp(n,ν)ordp(Lp(1n,ωn+2))=ordp(Nn+2).\delta_{p}(n,\nu)\leq\mathrm{ord}_{p}(L_{p}(-1-n,\omega^{n+2}))=\mathrm{ord}_{p}(N_{n+2}).

Suppose p11+νp-1\mid 1+\nu (resp. p1nν+1p-1\mid n-\nu+1). Then since p1n+2p-1\nmid n+2, we see that nν+1n-\nu+1 (resp. 1+ν1+\nu) is not divisible by p1p-1. Therefore, Corollary 7.4 shows that

Dp(n,ν)pLp(1n,ωn+2)+p,\displaystyle D_{p}(n,\nu)\in\frac{\mathbb{Z}_{p}}{L_{p}(-1-n,\omega^{n+2})}+\mathbb{Z}_{p},

which implies that δp(n,ν)ordp(Nn+2)\delta_{p}(n,\nu)\leq\mathrm{ord}_{p}(N_{n+2}). Hence Proposition 8.1 implies this lemma. ∎

Moreover, if p<np<n, the result of Carlitz concerning the index of irregularity of a prime shows the following lemma.

Lemma 8.4.

If p1n+2p-1\nmid n+2 and p<np<n, there is an (odd) integer ν{1,,n1}\nu\in\{1,\ldots,n-1\} such that δp(n,ν)=ordp(Nn+2)\delta_{p}(n,\nu)=\mathrm{ord}_{p}(N_{n+2}). In particular, we have δp(Eisn)=ordp(Nn+2)\delta_{p}(\mathrm{Eis}_{n})=\mathrm{ord}_{p}(N_{n+2}) in this case.

Proof.

By Lemma 8.3, for any regular prime pp and (odd) integer ν{1,,n1}\nu\in\{1,\ldots,n-1\}, we have δp(n,ν)=ordp(Nn+2)=0\delta_{p}(n,\nu)=\mathrm{ord}_{p}(N_{n+2})=0. Therefore, we may assume that pp is an irregular prime. In particular, p37p\geq 37.

We define the index d(p)d(p) of irregularity of the prime number pp by

d(p)\displaystyle d(p) :=#{1tp3t2,Btpp}\displaystyle:=\#\{1\leq t\leq p-3\mid t\in 2\mathbb{Z},B_{t}\in p\mathbb{Z}_{p}\}
=#{1tp3t2,Lp(1t,ωt)pp}.\displaystyle=\#\{1\leq t\leq p-3\mid t\in 2\mathbb{Z},L_{p}(1-t,\omega^{t})\in p\mathbb{Z}_{p}\}.

Then by using the result of Carlitz in [Car61, (21)], Skula proved in [Sku80, Theorem 2.2, Remark 2.3] that

d(p)<p+34log(2)log(p)p14.d(p)<\frac{p+3}{4}-\frac{\log(2)}{\log(p)}\frac{p-1}{4}.

Hence if p47p\geq 47, then we have

d(p)<p54.d(p)<\frac{p-5}{4}.

Moreover, since the only irregular prime smaller than 4747 is 3737 and d(37)=1<(375)/4d(37)=1<(37-5)/4, the inequality d(p)<(p5)/4d(p)<(p-5)/4 holds true.

For any integer aa, we define an integer [a]p1[a]_{p-1} by

0[a]p1p2 and [a]p1a(modp1).0\leq[a]_{p-1}\leq p-2\,\,\,\textrm{ and }\,\,\,[a]_{p-1}\equiv a\pmod{p-1}.

Since d(p)<(p5)/4d(p)<(p-5)/4, there is an even integer t{2,4,,p3}t\in\{2,4,\ldots,p-3\} with t[n+2]p1t\neq[n+2]_{p-1} such that

Lp(1t,ωt)p× and Lp(1[n+2t]p1,ω[n+2t]p1)p×.L_{p}(1-t,\omega^{t})\in\mathbb{Z}_{p}^{\times}\,\,\,\textrm{ and }\,\,\,L_{p}(1-[n+2-t]_{p-1},\omega^{[n+2-t]_{p-1}})\in\mathbb{Z}_{p}^{\times}.

Furthermore, Proposition 7.1(3) shows that

Lp(1[n+2t]p1,ω[n+2t]p1)Lp(1+tn,ωn+2t)pp,L_{p}(1-[n+2-t]_{p-1},\omega^{[n+2-t]_{p-1}})-L_{p}(-1+t-n,\omega^{n+2-t})\in p\mathbb{Z}_{p},

and hence we have Lp(1+tn,ωn+2t)p×L_{p}(-1+t-n,\omega^{n+2-t})\in\mathbb{Z}_{p}^{\times}. Therefore, we put ν:=t1\nu:=t-1 and get

δp(n,ν)=ordp(Nn+2).\delta_{p}(n,\nu)=\mathrm{ord}_{p}(N_{n+2}).

8.1.2. p1n+2p-1\mid n+2

Lemma 8.5.

If p1n+2p-1\mid n+2, we have δp(Eisn)=0=ordp(Nn+2)\delta_{p}(\mathrm{Eis}_{n})=0=\mathrm{ord}_{p}(N_{n+2}).

Proof.

The fact that ordp(Nn+2)=0\mathrm{ord}_{p}(N_{n+2})=0 follows from Corollary 7.3(2). Hence by Proposition 8.1, it suffices to show that δp(n,ν)=0\delta_{p}(n,\nu)=0 for any integer 1νn11\leq\nu\leq n-1. Since p1n+2p-1\mid n+2, we have ordp(Lp(1n,ωn+2))<0\mathrm{ord}_{p}(L_{p}(-1-n,\omega^{n+2}))<0. If p11+νp-1\nmid 1+\nu, then we also have p1nν+1p-1\nmid n-\nu+1, and hence we get δp(n,ν)=0\delta_{p}(n,\nu)=0 since Lp(ν,ω1+ν)L_{p}(-\nu,\omega^{1+\nu}) and Lp(νn,ωnν+1)L_{p}(\nu-n,\omega^{n-\nu+1}) are pp-adic integers. When p11+νp-1\mid 1+\nu, we also have p1nν+1p-1\mid n-\nu+1, and Corollary 7.5 implies that Dp(n,ν)pD_{p}(n,\nu)\in\mathbb{Z}_{p}. Hence, again, we obtain δp(n,ν)=0\delta_{p}(n,\nu)=0. ∎

This completes the proof of Proposition 8.2, and in particular of Theorem 2.13.

9. Applications

In this section, we discuss some applications of Theorem 2.13. For notational simplicity, in the following, the (co)homology groups will be denoted by H(Y,n)H^{\bullet}(Y,\mathcal{M}_{n}) (resp. H(Y,n)H_{\bullet}(Y,\mathcal{M}_{n})) rather than H(YBS,n)H^{\bullet}(Y^{\mathrm{BS}},\mathcal{M}_{n}) (resp. H(YBS,n)H_{\bullet}(Y^{\mathrm{BS}},\mathcal{M}_{n})) since they are naturally isomorphic.

First, note that we have the following corollary of Theorem 2.13.

Corollary 9.1.

Let n2n\geq 2 be an even integer and γΓ\gamma\in\Gamma a matrix. Take a polynomial P(X1,X2)nP(X_{1},X_{2})\in\mathcal{M}_{n} such that γP(X1,X2)=P(X1,X2)\gamma P(X_{1},X_{2})=P(X_{1},X_{2}). Then for any element τ\tau\in\mathbb{H}, we have

Nn+2τγτEn+2(z)P(z,1)𝑑z.\displaystyle N_{n+2}\int_{\tau}^{\gamma\tau}E_{n+2}(z)P(z,1)\,dz\in\mathbb{Z}.

Here Nn+2>0N_{n+2}>0 is the numerator of ζ(1n)\zeta(-1-n).

Proof.

Since γP=P\gamma P=P, we have ({τ,γτ}P)=0\partial(\{\tau,\gamma\tau\}\otimes P)=0, and hence {τ,γτ}P\{\tau,\gamma\tau\}\otimes P defines an element in the homology group H1(Y,n)H_{1}(Y,\mathcal{M}_{n}). Therefore, by Theorem 2.13, we obtain

Nn+2τγτEn+2(z)P(z,1)𝑑z=Nn+2Eisn,[{τ,γτ}P].\displaystyle N_{n+2}\int_{\tau}^{\gamma\tau}E_{n+2}(z)P(z,1)\,dz=\langle N_{n+2}\mathrm{Eis}_{n},[\{\tau,\gamma\tau\}\otimes P]\rangle\in\mathbb{Z}.

9.1. Duke’s conjecture

In the paper [Duk23], Duke defined a certain map called the higher Rademacher symbol

Ψk:Γ\Psi_{k}\colon\Gamma\longrightarrow\mathbb{Q}

for each integer k2k\in\mathbb{Z}_{\geq 2} which is a generalization of the classical Rademacher symbol and gave a conjecture concerning the integrality of the higher Rademacher symbol Ψk\Psi_{k}.

Conjecture 9.2 ([Duk23, Conjecture, p. 4]).

For any integer k2k\in\mathbb{Z}_{\geq 2} and matrix γΓ\gamma\in\Gamma, we have

Ψk(γ).\Psi_{k}(\gamma)\in\mathbb{Z}.

In the following, we show that Duke’s Conjecture 9.2 follows from Theorem 2.13.

Remark 9.3.

Conjecture 9.2 is recently proved also by O’Sullivan using a more direct method (see [O’S23]).

Here, instead of giving the original definition of the higher Rademacher symbols, we recall an integral representation of the higher Rademacher symbols, also given by Duke in [Duk23], which is equivalent to the original definition and more suitable for our purpose.

Proposition 9.4 ([Duk23, Definition (2.4) and Lemma 6]).

Let k2k\in\mathbb{Z}_{\geq 2} be an integer. For any matrix γ:=(abcd)Γ\gamma:=\begin{pmatrix}a&b\\ c&d\end{pmatrix}\in\Gamma, we define a binary quadratic polynomial Qγ(X1,X2)2Q_{\gamma}(X_{1},X_{2})\in\mathcal{M}_{2} associated with γ\gamma by

Qγ(X1,X2):=sgn(a+d)gcd(c,ad,b)(cX12(ad)X1X2bX22).Q_{\gamma}(X_{1},X_{2}):=-\frac{\operatorname{sgn}(a+d)}{\mathrm{gcd}(c,a-d,b)}(cX_{1}^{2}-(a-d)X_{1}X_{2}-bX_{2}^{2}).

Then for any element τ\tau\in\mathbb{H}, we have

Ψk(γ)=N2kτγτE2k(z)Qγ(z,1)k1𝑑z,\Psi_{k}(\gamma)=N_{2k}\int_{\tau}^{\gamma\tau}E_{2k}(z)Q_{\gamma}(z,1)^{k-1}\,dz,

where N2k>0N_{2k}>0 is the numerator of ζ(12k)\zeta(1-2k).

Corollary 9.5.

Duke’s Conjecture 9.2 holds true.

Proof.

By definition, the binary quadratic polynomial Qγ(X1,X2)Q_{\gamma}(X_{1},X_{2}) defined in Proposition 9.4 is γ\gamma-invariant. Hence Corollary 9.1 and Proposition 9.4 imply that Ψk(γ)\Psi_{k}(\gamma)\in\mathbb{Z}. ∎

9.2. Partial zeta functions of real quadratic fields

In this subsection, we discuss an application to the denominators of the special values of the partial zeta functions of real quadratic fields.

Let FF be a real quadratic field, and let 𝒪F\mathcal{O}\subset F be an order of FF with discriminant D𝒪D_{\mathcal{O}}. We denote by I𝒪I_{\mathcal{O}} the group of proper fractional 𝒪\mathcal{O}-ideals and P𝒪+I𝒪P_{\mathcal{O}}^{+}\subset I_{\mathcal{O}} the subgroup of totally positive principal ideals. We define the narrow ideal class group Cl𝒪+Cl_{\mathcal{O}}^{+} of 𝒪\mathcal{O} by

Cl𝒪+:=I𝒪/P𝒪+.Cl_{\mathcal{O}}^{+}:=I_{\mathcal{O}}/P_{\mathcal{O}}^{+}.

See [Cox13, §7]. We fix an embedding FF\subset\mathbb{R}, and for any element αF\alpha\in F\subset\mathbb{R}, we denote by αF\alpha^{\prime}\in F\subset\mathbb{R} its conjugate over \mathbb{Q}.

Moreover, let 𝒪+×\mathcal{O}_{+}^{\times} denote the group of totally positive units in 𝒪\mathcal{O}, and let ε0𝒪+×\varepsilon_{0}\in\mathcal{O}_{+}^{\times} denote the generator of 𝒪+×\mathcal{O}_{+}^{\times} such that ε0>1\varepsilon_{0}>1.

Definition 9.6.

We define a map

𝔷𝒪,k:Cl𝒪+H1(Y,2k2)\displaystyle\mathfrak{z}_{\mathcal{O},k}\colon Cl_{\mathcal{O}}^{+}\longrightarrow H_{1}(Y,\mathcal{M}_{2k-2})

as follows: Let 𝒜Cl𝒪+\mathcal{A}\in Cl_{\mathcal{O}}^{+}, and take a representative 𝔞I𝒪\mathfrak{a}\in I_{\mathcal{O}} of 𝒜\mathcal{A}. We also take a basis α1,α2𝔞\alpha_{1},\alpha_{2}\in\mathfrak{a} over \mathbb{Z} such that α1α2α1α2>0\alpha_{1}\alpha^{\prime}_{2}-\alpha^{\prime}_{1}\alpha_{2}>0, and let γ0Γ\gamma_{0}\in\Gamma be a matrix such that

γ0(α1α2)=(ε0α1ε0α2).\displaystyle\gamma_{0}\begin{pmatrix}\alpha_{1}\\ \alpha_{2}\end{pmatrix}=\begin{pmatrix}\varepsilon_{0}\alpha_{1}\\ \varepsilon_{0}\alpha_{2}\end{pmatrix}.

Moreover, set

Nα1,α2(X1,X2):=1N𝔞(α2X1α1X2)(α2X1α1X2).\displaystyle N_{\alpha_{1},\alpha_{2}}(X_{1},X_{2}):=-\frac{1}{N\mathfrak{a}}(\alpha_{2}X_{1}-\alpha_{1}X_{2})(\alpha^{\prime}_{2}X_{1}-\alpha^{\prime}_{1}X_{2}).

We see that Nα1,α2(X1,X2)2N_{\alpha_{1},\alpha_{2}}(X_{1},X_{2})\in\mathcal{M}_{2} and that γ0Nα1,α2(X1,X2)=Nα1,α2(X1,X2)\gamma_{0}N_{\alpha_{1},\alpha_{2}}(X_{1},X_{2})=N_{\alpha_{1},\alpha_{2}}(X_{1},X_{2}). We then define

𝔷𝒪,k(𝒜):=[{τ,γ0τ}Nα1,α2(X1,X2)k1]H1(Y,2k2),\displaystyle\mathfrak{z}_{\mathcal{O},k}(\mathcal{A}):=[\{\tau,\gamma_{0}\tau\}\otimes N_{\alpha_{1},\alpha_{2}}(X_{1},X_{2})^{k-1}]\in H_{1}(Y,\mathcal{M}_{2k-2}),

where τ\tau is an arbitrary element in \mathbb{H}.

Lemma 9.7.

The homology class 𝔷𝒪,k(𝒜)\mathfrak{z}_{\mathcal{O},k}(\mathcal{A}) does not depend on the choices we made.

Proof.

The independence of τ\tau\in\mathbb{H} is clear. Let 𝔟𝒜\mathfrak{b}\in\mathcal{A} be another representative. Then there exists a totally positive element αF×\alpha\in F^{\times} such that 𝔟=α𝔞\mathfrak{b}=\alpha\mathfrak{a}. Take a basis β1,β2𝔞\beta_{1},\beta_{2}\in\mathfrak{a} over \mathbb{Z} with β1β2β1β2>0\beta_{1}\beta^{\prime}_{2}-\beta^{\prime}_{1}\beta_{2}>0. Then we obtain a matrix γ0,𝔟Γ\gamma_{0,\mathfrak{b}}\in\Gamma and a binary quadratic polynomial Nαβ1,αβ2(X1,X2)N_{\alpha\beta_{1},\alpha\beta_{2}}(X_{1},X_{2}) from the basis αβ1,αβ2\alpha\beta_{1},\alpha\beta_{2} of 𝔟\mathfrak{b}. Note that since α\alpha is totally positive, we have (αβ1)(αβ2)(αβ1)(αβ2)>0(\alpha\beta_{1})(\alpha^{\prime}\beta^{\prime}_{2})-(\alpha^{\prime}\beta^{\prime}_{1})(\alpha\beta_{2})>0. Let γGL2()\gamma\in\mathrm{GL}_{2}(\mathbb{Z}) be a matrix satisfying

(β1β2)=γ(α1α2).\begin{pmatrix}\beta_{1}\\ \beta_{2}\end{pmatrix}=\gamma\begin{pmatrix}\alpha_{1}\\ \alpha_{2}\end{pmatrix}.

Then the facts that α1α2α1α2>0\alpha_{1}\alpha^{\prime}_{2}-\alpha^{\prime}_{1}\alpha_{2}>0 and β1β2β1β2>0\beta_{1}\beta^{\prime}_{2}-\beta^{\prime}_{1}\beta_{2}>0 imply that γΓ\gamma\in\Gamma. Since γ0,𝔟γ(α1α2)=γ(ε0α1ε0α2)\gamma_{0,\mathfrak{b}}\gamma\begin{pmatrix}\alpha_{1}\\ \alpha_{2}\end{pmatrix}=\gamma\begin{pmatrix}\varepsilon_{0}\alpha_{1}\\ \varepsilon_{0}\alpha_{2}\end{pmatrix}, we have γ1γ0,𝔟γ=γ0\gamma^{-1}\gamma_{0,\mathfrak{b}}\gamma=\gamma_{0}. Moreover, we have

(X1X2)γ~t(0110)(α1α2)=(X1X2)(0110)γ(α1α2)=(X1X2)(0110)(β1β2),\begin{pmatrix}X_{1}&X_{2}\end{pmatrix}{}^{t}\widetilde{\gamma}\begin{pmatrix}0&1\\ -1&0\end{pmatrix}\begin{pmatrix}\alpha_{1}\\ \alpha_{2}\end{pmatrix}=\begin{pmatrix}X_{1}&X_{2}\end{pmatrix}\begin{pmatrix}0&1\\ -1&0\end{pmatrix}\gamma\begin{pmatrix}\alpha_{1}\\ \alpha_{2}\end{pmatrix}=\begin{pmatrix}X_{1}&X_{2}\end{pmatrix}\begin{pmatrix}0&1\\ -1&0\end{pmatrix}\begin{pmatrix}\beta_{1}\\ \beta_{2}\end{pmatrix},

which implies that Nαβ1,αβ2(X1,X2)=γNα1,α2(X1,X2)N_{\alpha\beta_{1},\alpha\beta_{2}}(X_{1},X_{2})=\gamma N_{\alpha_{1},\alpha_{2}}(X_{1},X_{2}). Therefore, we have

[{τ,γ0,𝔟τ}Nαβ1,αβ2(X1,X2)k1]\displaystyle[\{\tau,\gamma_{0,\mathfrak{b}}\tau\}\otimes N_{\alpha\beta_{1},\alpha\beta_{2}}(X_{1},X_{2})^{k-1}] =[{τ,γγ0γ1τ}γNα1,α2(X1,X2)k1]\displaystyle=[\{\tau,\gamma\gamma_{0}\gamma^{-1}\tau\}\otimes\gamma N_{\alpha_{1},\alpha_{2}}(X_{1},X_{2})^{k-1}]
=[{γ1τ,γ0γ1τ}Nα1,α2(X1,X2)k1]\displaystyle=[\{\gamma^{-1}\tau,\gamma_{0}\gamma^{-1}\tau\}\otimes N_{\alpha_{1},\alpha_{2}}(X_{1},X_{2})^{k-1}]
=[{τ,γ0τ}Nα1,α2(X1,X2)k1]\displaystyle=[\{\tau,\gamma_{0}\tau\}\otimes N_{\alpha_{1},\alpha_{2}}(X_{1},X_{2})^{k-1}]

as elements of H1(Y,2k2)H_{1}(Y,\mathcal{M}_{2k-2}). ∎

Remark 9.8.
  • (1)

    Since the matrix γ0\gamma_{0} in Definition 9.6 is hyperbolic (i.e., |trace(γ0)|>2|\mathrm{trace}(\gamma_{0})|>2), we have dim{Q2γQ=Q}=1\dim_{\mathbb{Q}}\{Q\in\mathcal{M}_{2}\otimes\mathbb{Q}\mid\gamma Q=Q\}=1. This fact together with [Cox13, (7.6)] shows that Nα1,α2(X1,X2)=Qγ0(X1,X2)N_{\alpha_{1},\alpha_{2}}(X_{1},X_{2})=Q_{\gamma_{0}}(X_{1},X_{2}).

  • (2)

    Gauss’s theory concerning binary quadratic forms (see [Cox13, Exercise 7.21] for example) shows that for any hyperbolic element γΓ\gamma\in\Gamma, there is an order 𝒪\mathcal{O} of a real quadratic field and a narrow ideal class 𝒜Cl𝒪+\mathcal{A}\in Cl_{\mathcal{O}}^{+} such that

    [{z,γz}Qγ(X1,X2)]𝔷𝒪,2(𝒜).[\{z,\gamma z\}\otimes Q_{\gamma}(X_{1},X_{2})]\in\mathbb{Z}\mathfrak{z}_{\mathcal{O},2}(\mathcal{A}).
Definition 9.9.

For each ideal class 𝒜Cl𝒪+\mathcal{A}\in Cl_{\mathcal{O}}^{+}, the partial zeta function ζ𝒪(𝒜,s)\zeta_{\mathcal{O}}(\mathcal{A},s) associated with 𝒜\mathcal{A} is defined by

ζ𝒪(𝒜,s):=𝔞𝒪,𝔞𝒜1(N𝔞)s,(Re(s)>1),\displaystyle\zeta_{\mathcal{O}}(\mathcal{A},s):=\sum_{\mathfrak{a}\subset\mathcal{O},\mathfrak{a}\in\mathcal{A}}\frac{1}{(N\mathfrak{a})^{s}},\quad\quad(\operatorname{Re}(s)>1),

and it is well-known that ζ𝒪(𝒜,s)\zeta_{\mathcal{O}}(\mathcal{A},s) can be continued meromorphically to ss\in\mathbb{C} and has a simple pole at s=1s=1.

The following integral representation of the special values of the partial zeta function is classically known.

Proposition 9.10.

For any integer k2k\in\mathbb{Z}_{\geq 2} and ideal class 𝒜Cl𝒪+\mathcal{A}\in Cl_{\mathcal{O}}^{+}, we have

Eis2k2,𝔷𝒪,k(𝒜)=(1)kζ𝒪(𝒜1,1k)ζ(12k).\displaystyle\langle\mathrm{Eis}_{2k-2},\mathfrak{z}_{\mathcal{O},k}(\mathcal{A})\rangle=(-1)^{k}\frac{\zeta_{\mathcal{O}}(\mathcal{A}^{-1},1-k)}{\zeta(1-2k)}.

Before we give a proof of Proposition 9.10, we recall (a special case of) the so-called Feynman parametrization.

Lemma 9.11.

Let x1,x2,a,bx_{1},x_{2},a,b\in\mathbb{C} be complex numbers such that x1a+x20x_{1}a+x_{2}\neq 0 and x1b+x20x_{1}b+x_{2}\neq 0. Then for any non-negative integers k1k_{1} and k2k_{2}, we have

ab(bz)k1(za)k2(x1z+x2)2+k1+k2𝑑z=k1!k2!(k1+k2+1)!(ba)k1+k2+1(x1a+x2)k1+1(x1b+x2)k2+1.\int_{a}^{b}\frac{\left(b-z\right)^{k_{1}}\left(z-a\right)^{k_{2}}}{(x_{1}z+x_{2})^{2+k_{1}+k_{2}}}\,dz=\frac{k_{1}!k_{2}!}{(k_{1}+k_{2}+1)!}\frac{(b-a)^{k_{1}+k_{2}+1}}{(x_{1}a+x_{2})^{k_{1}+1}(x_{1}b+x_{2})^{k_{2}+1}}.
Proof.

We may assume that aba\neq b. By setting y1=x1a+x2y_{1}=x_{1}a+x_{2} and y2=x1b+x2y_{2}=x_{1}b+x_{2}, it suffices to prove that

(ba)ab(bz)k1(za)k2((bz)y1+(za)y2)2+k1+k2𝑑z=k1!k2!(k1+k2+1)!1y1k1+1y2k2+1.(b-a)\int_{a}^{b}\frac{\left(b-z\right)^{k_{1}}\left(z-a\right)^{k_{2}}}{((b-z)y_{1}+(z-a)y_{2})^{2+k_{1}+k_{2}}}\,dz=\frac{k_{1}!k_{2}!}{(k_{1}+k_{2}+1)!}\frac{1}{y_{1}^{k_{1}+1}y_{2}^{k_{2}+1}}.

The case where k1=k2=0k_{1}=k_{2}=0 is clear, i.e., we have

(ba)ab1((bz)y1+(za)y2)2𝑑z=1y1y2.(b-a)\int_{a}^{b}\frac{1}{((b-z)y_{1}+(z-a)y_{2})^{2}}\,dz=\frac{1}{y_{1}y_{2}}.

Then by viewing the both sides as holomorphic functions in (y1,y2)×××(y_{1},y_{2})\in\mathbb{C}^{\times}\times\mathbb{C}^{\times} and applying the differential operator (y1)k1(y2)k2\left(\frac{\partial}{\partial y_{1}}\right)^{k_{1}}\left(\frac{\partial}{\partial y_{2}}\right)^{k_{2}}, we obtain the desired identity. ∎

Proof of Proposition 9.10.

We use the same notations as in Definition 9.6. Since

2ζ(2k)E2k(z)=(0,0)(m,n)21(mz+n)2k2\zeta(2k)E_{2k}(z)=\sum_{(0,0)\neq(m,n)\in\mathbb{Z}^{2}}\frac{1}{(mz+n)^{2k}}

and Eis2k2=r(E2k)\mathrm{Eis}_{2k-2}=r(E_{2k}), for any element τ\tau\in\mathbb{H} we have

2ζ(2k)Eis2k2,𝔷𝒪,k(𝒜)=τγ0τ(0,0)(m,n)2Nα1,α2(z,1)k1(mz+n)2kdz.\displaystyle 2\zeta(2k)\langle\mathrm{Eis}_{2k-2},\mathfrak{z}_{\mathcal{O},k}(\mathcal{A})\rangle=\int_{\tau}^{\gamma_{0}\tau}\sum_{(0,0)\neq(m,n)\in\mathbb{Z}^{2}}\frac{N_{\alpha_{1},\alpha_{2}}(z,1)^{k-1}}{(mz+n)^{2k}}\,dz.

Since Nα1,α2(γ0z,1)=j(γ0,z)2Nα1,α2(z,1)N_{\alpha_{1},\alpha_{2}}(\gamma_{0}z,1)=j(\gamma_{0},z)^{-2}N_{\alpha_{1},\alpha_{2}}(z,1), where j((abcd),z):=cz+dj(\begin{pmatrix}a&b\\ c&d\end{pmatrix},z):=cz+d is the factor of automorphy, if we fix a complete set Sγ0S_{\gamma_{0}} of representatives of (2{(0,0)})/γ0(\mathbb{Z}^{2}\!-\!\{(0,0)\})/\gamma_{0}^{\mathbb{Z}}, then we have

τγ0τ(0,0)(m,n)2Nα1,α2(z,1)k1(mz+n)2kdz\displaystyle\int_{\tau}^{\gamma_{0}\tau}\sum_{(0,0)\neq(m,n)\in\mathbb{Z}^{2}}\frac{N_{\alpha_{1},\alpha_{2}}(z,1)^{k-1}}{(mz+n)^{2k}}\,dz =τγ0τl(m,n)Sγ0Nα1,α2(z,1)k1j(γ0l,z)2k(m(γ0lz)+n)2kdz\displaystyle=\int_{\tau}^{\gamma_{0}\tau}\sum_{l\in\mathbb{Z}}\sum_{(m,n)\in S_{\gamma_{0}}}\frac{N_{\alpha_{1},\alpha_{2}}(z,1)^{k-1}}{j(\gamma_{0}^{l},z)^{2k}(m(\gamma_{0}^{l}z)+n)^{2k}}\,dz
=lγ0lτγ0l+1τ(m,n)Sγ0Nα1,α2(γ0lz,1)k1j(γ0l,γ0lz)2k(mz+n)2kd(γ0lz)\displaystyle=\sum_{l\in\mathbb{Z}}\int_{\gamma_{0}^{l}\tau}^{\gamma_{0}^{l+1}\tau}\sum_{(m,n)\in S_{\gamma_{0}}}\frac{N_{\alpha_{1},\alpha_{2}}(\gamma_{0}^{-l}z,1)^{k-1}}{j(\gamma_{0}^{l},\gamma_{0}^{-l}z)^{2k}(mz+n)^{2k}}\,d(\gamma_{0}^{-l}z)
=lγ0lτγ0l+1τ(m,n)Sγ0Nα1,α2(z,1)k1(mz+n)2kdz.\displaystyle=\sum_{l\in\mathbb{Z}}\int_{\gamma_{0}^{l}\tau}^{\gamma_{0}^{l+1}\tau}\sum_{(m,n)\in S_{\gamma_{0}}}\frac{N_{\alpha_{1},\alpha_{2}}(z,1)^{k-1}}{(mz+n)^{2k}}\,dz.

Set α0:=α1/α2F\alpha_{0}:=\alpha_{1}/\alpha_{2}\in F\subset\mathbb{R}. Then the point α0\alpha_{0}\in\mathbb{R} (resp. α0\alpha^{\prime}_{0}) is the attractive fixed point (resp. repelling fixed point) of the hyperbolic matrix γ0Γ\gamma_{0}\in\Gamma, i.e., we have limlγ0lτ=α0\lim_{l\to\infty}\gamma_{0}^{l}\tau=\alpha_{0} and limlγ0lτ=α0\lim_{l\to\infty}\gamma_{0}^{-l}\tau=\alpha^{\prime}_{0} in 1()\mathbb{P}^{1}(\mathbb{C}) for any element τ\tau\in\mathbb{H}. Hence we obtain

lγ0lτγ0l+1τ(m,n)Sγ0Nα1,α2(z,1)k1(mz+n)2kdz\displaystyle\sum_{l\in\mathbb{Z}}\int_{\gamma_{0}^{l}\tau}^{\gamma_{0}^{l+1}\tau}\sum_{(m,n)\in S_{\gamma_{0}}}\frac{N_{\alpha_{1},\alpha_{2}}(z,1)^{k-1}}{(mz+n)^{2k}}\,dz =α0α0(m,n)Sγ0Nα1,α2(z,1)k1(mz+n)2kdz\displaystyle=\int_{\alpha^{\prime}_{0}}^{\alpha_{0}}\sum_{(m,n)\in S_{\gamma_{0}}}\frac{N_{\alpha_{1},\alpha_{2}}(z,1)^{k-1}}{(mz+n)^{2k}}\,dz
=NF/(α2)k1(N𝔞)k1(m,n)Sγ0α0α0((α0z)(zα0))k1(mz+n)2k𝑑z.\displaystyle=\frac{N_{F/\mathbb{Q}}(\alpha_{2})^{k-1}}{(N\mathfrak{a})^{k-1}}\sum_{(m,n)\in S_{\gamma_{0}}}\int_{\alpha^{\prime}_{0}}^{\alpha_{0}}\frac{((\alpha_{0}-z)(z-\alpha^{\prime}_{0}))^{k-1}}{(mz+n)^{2k}}\,dz.

By using Lemma 9.11, we find

α0α0((α0z)(zα0))k1(mz+n)2k𝑑z=((k1)!)2(2k1)!(α0α0)2k1NF/(mα0+n)k.\displaystyle\int_{\alpha^{\prime}_{0}}^{\alpha_{0}}\frac{((\alpha_{0}-z)(z-\alpha^{\prime}_{0}))^{k-1}}{(mz+n)^{2k}}\,dz=\frac{((k-1)!)^{2}}{(2k-1)!}\frac{(\alpha_{0}-\alpha^{\prime}_{0})^{2k-1}}{N_{F/\mathbb{Q}}(m\alpha_{0}+n)^{k}}.

Note that we have the identity α1α2α1α2=D𝒪N𝔞\alpha_{1}\alpha^{\prime}_{2}-\alpha^{\prime}_{1}\alpha_{2}=\sqrt{D_{\mathcal{O}}}N\mathfrak{a}, and this shows that

NF/(α2)k1(N𝔞)k1(m,n)Sγ0(α0α0)2k1NF/(mα0+n)k=D𝒪k1/2(N𝔞)kα(𝔞{0})/𝒪+×1NF/(α)k.\displaystyle\frac{N_{F/\mathbb{Q}}(\alpha_{2})^{k-1}}{(N\mathfrak{a})^{k-1}}\sum_{(m,n)\in S_{\gamma_{0}}}\frac{(\alpha_{0}-\alpha^{\prime}_{0})^{2k-1}}{N_{F/\mathbb{Q}}(m\alpha_{0}+n)^{k}}=D_{\mathcal{O}}^{k-1/2}(N\mathfrak{a})^{k}\sum_{\alpha\in(\mathfrak{a}\!-\!\{0\})/\mathcal{O}_{+}^{\times}}\frac{1}{N_{F/\mathbb{Q}}(\alpha)^{k}}.

For any subset XFX\subset F, we put

X+\displaystyle X_{+} :={αXα>0,α>0},\displaystyle:=\{\alpha\in X\mid\alpha>0,\,\alpha^{\prime}>0\},
X\displaystyle X_{-} :={αXα>0,α<0}.\displaystyle:=\{\alpha\in X\mid\alpha>0,\,\alpha^{\prime}<0\}.

Let 𝒥Cl𝒪+\mathcal{J}\in Cl_{\mathcal{O}}^{+} denote the ideal class containing the principal ideal (D𝒪)𝒪(\sqrt{D_{\mathcal{O}}})\subset\mathcal{O}. Note that 𝒥1=𝒥\mathcal{J}^{-1}=\mathcal{J} in Cl𝒪+Cl_{\mathcal{O}}^{+}. Then we further compute

(N𝔞)kα(𝔞{0})/𝒪+×1NF/(α)k\displaystyle(N\mathfrak{a})^{k}\sum_{\alpha\in(\mathfrak{a}\!-\!\{0\})/\mathcal{O}_{+}^{\times}}\frac{1}{N_{F/\mathbb{Q}}(\alpha)^{k}} =α𝔞+/𝒪+×2(N𝔞)kNF/(α)k+α𝔞/𝒪+×2(N𝔞)kNF/(α)k\displaystyle=\sum_{\alpha\in\mathfrak{a}_{+}/\mathcal{O}_{+}^{\times}}\frac{2(N\mathfrak{a})^{k}}{N_{F/\mathbb{Q}}(\alpha)^{k}}+\sum_{\alpha\in\mathfrak{a}_{-}/\mathcal{O}_{+}^{\times}}\frac{2(N\mathfrak{a})^{k}}{N_{F/\mathbb{Q}}(\alpha)^{k}}
=α𝔞+/𝒪+×2(N𝔞)kNF/(α)k+(1)kα(D𝒪𝔞)+/𝒪+×2N(D𝒪𝔞)kNF/(α)k\displaystyle=\sum_{\alpha\in\mathfrak{a}_{+}/\mathcal{O}_{+}^{\times}}\frac{2(N\mathfrak{a})^{k}}{N_{F/\mathbb{Q}}(\alpha)^{k}}+(-1)^{k}\sum_{\alpha\in(\sqrt{D_{\mathcal{O}}}\mathfrak{a})_{+}/\mathcal{O}_{+}^{\times}}\frac{2N(\sqrt{D_{\mathcal{O}}}\mathfrak{a})^{k}}{N_{F/\mathbb{Q}}(\alpha)^{k}}
=2(ζ𝒪(𝒜1,k)+(1)kζ𝒪(𝒥𝒜1,k)).\displaystyle=2\left(\zeta_{\mathcal{O}}(\mathcal{A}^{-1},k)+(-1)^{k}\zeta_{\mathcal{O}}(\mathcal{J}\mathcal{A}^{-1},k)\right).

Now, we recall the functional equations of the partial zeta functions. Set

Λ𝒪+(𝒜1,s)\displaystyle\Lambda_{\mathcal{O}}^{+}(\mathcal{A}^{-1},s) :=πsΓ(s2)2D𝒪s12(ζ𝒪(𝒜1,s)+ζ𝒪(𝒜1𝒥,s)),\displaystyle:=\pi^{-s}\Gamma\left(\frac{s}{2}\right)^{2}D_{\mathcal{O}}^{\frac{s-1}{2}}\left(\zeta_{\mathcal{O}}(\mathcal{A}^{-1},s)+\zeta_{\mathcal{O}}(\mathcal{A}^{-1}\mathcal{J},s)\right),
Λ𝒪(𝒜1,s)\displaystyle\Lambda_{\mathcal{O}}^{-}(\mathcal{A}^{-1},s) :=πsΓ(s+12)2D𝒪s2(ζ𝒪(𝒜1,s)ζ𝒪(𝒜1𝒥,s)).\displaystyle:=\pi^{-s}\Gamma\left(\frac{s+1}{2}\right)^{2}D_{\mathcal{O}}^{\frac{s}{2}}\left(\zeta_{\mathcal{O}}(\mathcal{A}^{-1},s)-\zeta_{\mathcal{O}}(\mathcal{A}^{-1}\mathcal{J},s)\right).

Then we have

Λ𝒪+(𝒜1,s)\displaystyle\Lambda_{\mathcal{O}}^{+}(\mathcal{A}^{-1},s) =Λ𝒪+(𝒜1,1s),\displaystyle=\Lambda_{\mathcal{O}}^{+}(\mathcal{A}^{-1},1-s),
Λ𝒪(𝒜1,s)\displaystyle\Lambda_{\mathcal{O}}^{-}(\mathcal{A}^{-1},s) =Λ𝒪(𝒜1,1s).\displaystyle=\Lambda_{\mathcal{O}}^{-}(\mathcal{A}^{-1},1-s).

See [DIT18, Equations (59), (60)] or [Scz93, p. 545] for example. Although [DIT18] deals only with the maximal orders, we can apply the same argument to general orders. See also [Sie80], [Duk23, (4.19)] or [VZ13, p. 42]. Using these functional equations, we find

D𝒪k1/2(ζ𝒪(𝒜1,k)+(1)kζ𝒪(𝒥𝒜1,k))=(2π)2k2((k1)!)2ζ𝒪(𝒜1,1k).D_{\mathcal{O}}^{k-1/2}\left(\zeta_{\mathcal{O}}(\mathcal{A}^{-1},k)+(-1)^{k}\zeta_{\mathcal{O}}(\mathcal{J}\mathcal{A}^{-1},k)\right)=\frac{(2\pi)^{2k}}{2((k-1)!)^{2}}\zeta_{\mathcal{O}}(\mathcal{A}^{-1},1-k).

Therefore, by also using the functional equation for the Riemann zeta function (see, for example, [Hid93, p. 29]), we obtain

Eis2k2,𝔷𝒪,k(𝒜)=(2π)2k2(2k1)!ζ(2k)ζ𝒪(𝒜1,1k)=(1)kζ𝒪(𝒜1,1k)ζ(12k).\displaystyle\langle\mathrm{Eis}_{2k-2},\mathfrak{z}_{\mathcal{O},k}(\mathcal{A})\rangle=\frac{(2\pi)^{2k}}{2(2k-1)!\zeta(2k)}\zeta_{\mathcal{O}}(\mathcal{A}^{-1},1-k)=(-1)^{k}\frac{\zeta_{\mathcal{O}}(\mathcal{A}^{-1},1-k)}{\zeta(1-2k)}.

We define the positive integer J2kJ_{2k} by

J2k:=the denominator of ζ(12k).J_{2k}:=\text{the denominator of }\zeta(1-2k).
Corollary 9.12.

Let FF be a real quadratic field, 𝒪F\mathcal{O}\subset F be an order in FF, and let 𝒜Cl𝒪+\mathcal{A}\in Cl_{\mathcal{O}}^{+} be a narrow ideal class of 𝒪\mathcal{O}. Then for any integer k2k\geq 2, we have

J2kζ𝒪(𝒜,1k).J_{2k}\zeta_{\mathcal{O}}(\mathcal{A},1-k)\in\mathbb{Z}.
Proof.

By Proposition 9.10, we have

N2kEis2k2,𝔷𝒪,k(𝒜1)=±N2kζ𝒪(𝒜,1k)ζ(12k)=±J2kζ𝒪(𝒜,1k).N_{2k}\langle\mathrm{Eis}_{2k-2},\mathfrak{z}_{\mathcal{O},k}(\mathcal{A}^{-1})\rangle=\pm N_{2k}\frac{\zeta_{\mathcal{O}}(\mathcal{A},1-k)}{\zeta(1-2k)}=\pm J_{2k}\zeta_{\mathcal{O}}(\mathcal{A},1-k).

Since N2kEis2k2,𝔷𝒪,k(𝒜1)N_{2k}\langle\mathrm{Eis}_{2k-2},\mathfrak{z}_{\mathcal{O},k}(\mathcal{A}^{-1})\rangle\in\mathbb{Z} by Theorem 2.13, we obtain J2kζ𝒪(𝒜,1k)J_{2k}\zeta_{\mathcal{O}}(\mathcal{A},1-k)\in\mathbb{Z}. ∎

Remark 9.13.

By Proposition 9.4 ([Duk23, Lemma 6]) and Proposition 9.10, we see that Duke’s conjecture 9.2 is equivalent to Corollary 9.12.

Remark 9.14.

As for the denominator of the special values of the Dedekind zeta functions of real quadratic fields, or more generally of totally real fields, the same (a slightly stronger at p=2p=2) universal upper bound was obtained by Serre in the paper [Ser73, §2, Théorème 6]. Moreover, if we fix a totally real field FF, then a more refined description for the denominators and even for the numerators of the special values of the Dedekind zeta function of FF is obtained from the classical Iwasawa main conjecture proved by Wiles in [Wil90] (see [Kol04]).

9.3. Sharpness of the universal upper bound in Corollary 9.12

Let k2k\geq 2 be an integer. We define a \mathbb{Z}-submodule kH1(Y,2k2)\mathfrak{Z}_{k}\subset H_{1}(Y,\mathcal{M}_{2k-2}) to be the \mathbb{Z}-submodule generated by homology classes of the form 𝔷𝒪,k(𝒜)\mathfrak{z}_{\mathcal{O},k}(\mathcal{A}), that is,

k:=𝔷𝒪,k(𝒜)𝒪 is an order of a real quadratic field and 𝒜Cl𝒪+ .\mathfrak{Z}_{k}:=\langle\,\mathfrak{z}_{\mathcal{O},k}(\mathcal{A})\mid\textrm{$\mathcal{O}$ is an order of a real quadratic field and $\mathcal{A}\in Cl_{\mathcal{O}}^{+}$ }\rangle_{\mathbb{Z}}.

This subsection is devoted to proving the following theorem.

Theorem 9.15.

We have Eis2k2,k=1N2k\displaystyle\left\langle\mathrm{Eis}_{2k-2},\mathfrak{Z}_{k}\right\rangle=\frac{1}{N_{2k}}\mathbb{Z}.

Theorem 9.15 has the following interesting application.

Corollary 9.16.

The universal bound in Corollary 9.12 is sharp, namely, for any prime number pp, there exist an order 𝒪\mathcal{O} of a real quadratic field and a narrow ideal class 𝒜Cl𝒪+\mathcal{A}\in Cl_{\mathcal{O}}^{+} such that

ordp(J2kζ𝒪(𝒜,1k))=0.\mathrm{ord}_{p}(J_{2k}\zeta_{\mathcal{O}}(\mathcal{A},1-k))=0.

In other words, we have

J2k=min{J>0|Jζ𝒪(𝒜,1k) for all orders 𝒪 in all real quadratic fieldsand narrow ideal classes 𝒜Cl𝒪+}.J_{2k}=\min\left\{J\in\mathbb{Z}_{>0}\,\,\middle|\,\,\begin{aligned} J\zeta_{\mathcal{O}}(\mathcal{A},1-k)\in\mathbb{Z}\text{ for all orders $\mathcal{O}$ in all real quadratic fields}\\ \text{and narrow ideal classes $\mathcal{A}\in Cl_{\mathcal{O}}^{+}$}\end{aligned}\,\right\}.
Proof.

Let pp be a prime number. By the definition of the module k\mathfrak{Z}_{k} and Theorem 9.15, one can find an order 𝒪\mathcal{O} of a real quadratic field and a narrow ideal class 𝒜Cl𝒪+\mathcal{A}\in Cl_{\mathcal{O}}^{+} such that

ordp(Eis2k2,𝔷𝒪,k(𝒜1))=ordp(N2k).\mathrm{ord}_{p}(\langle\mathrm{Eis}_{2k-2},\mathfrak{z}_{\mathcal{O},k}(\mathcal{A}^{-1})\rangle)=-\mathrm{ord}_{p}(N_{2k}).

Since ζ(12k)=±N2k/J2k\zeta(1-2k)=\pm N_{2k}/J_{2k}, Proposition 9.10 shows that

0\displaystyle 0 =ordp(N2k)+ordp(Eis2k2,𝔷𝒪,k(𝒜1))\displaystyle=\mathrm{ord}_{p}(N_{2k})+\mathrm{ord}_{p}(\langle\mathrm{Eis}_{2k-2},\mathfrak{z}_{\mathcal{O},k}(\mathcal{A}^{-1})\rangle)
=ordp(N2k)ordp(ζ(12k))+ordp(ζ𝒪(𝒜,1k))\displaystyle=\mathrm{ord}_{p}(N_{2k})-\mathrm{ord}_{p}(\zeta(1-2k))+\mathrm{ord}_{p}(\zeta_{\mathcal{O}}(\mathcal{A},1-k))
=ordp(J2kζ𝒪(𝒜,1k)).\displaystyle=\mathrm{ord}_{p}(J_{2k}\zeta_{\mathcal{O}}(\mathcal{A},1-k)).

9.3.1. Preparations for proving Theorem 9.15

Let N1N\geq 1 be an integer and define

Γ1(N):={(abcd)Γ|a1cd10(modN)}.\Gamma_{1}(N):=\left\{\begin{pmatrix}a&b\\ c&d\end{pmatrix}\in\Gamma\,\middle|\,a-1\equiv c\equiv d-1\equiv 0\pmod{N}\right\}.

We also put

Y1(N):=Γ1(N)\,Y1(N)BS:=Γ1(N)\BS,Y1(N)BS:=Y1(N)BSY1(N).Y_{1}(N):=\Gamma_{1}(N)\backslash\mathbb{H},\,\,\,\,\,\,Y_{1}(N)^{\mathrm{BS}}:=\Gamma_{1}(N)\backslash\mathbb{H}^{\mathrm{BS}},\,\,\,\,\,\,\partial Y_{1}(N)^{\mathrm{BS}}:=Y_{1}(N)^{\mathrm{BS}}\!-\!Y_{1}(N).

We note that the similar facts in §2.1 and §2.2 hold true for the congruence subgroup Γ1(N)\Gamma_{1}(N). Moreover, the Hecke operators TpT_{p} (pNp\nmid N) and UpU_{p} (pNp\mid N) act on the homology group H1(Y1(N),2k2)H_{1}(Y_{1}(N),\mathcal{M}_{2k-2}).

Let k1k\geq 1 be an integer. For any hyperbolic matrix γΓ1(N)\gamma\in\Gamma_{1}(N) (i.e., |trace(γ)|>2|\mathrm{trace}(\gamma)|>2), we set

𝔷Γ1(N),k(γ):=[{z,γz}Qγ(X1,X2)k1]H1(Y1(N),2k2).\mathfrak{z}_{\Gamma_{1}(N),k}(\gamma):=[\{z,\gamma z\}\otimes Q_{\gamma}(X_{1},X_{2})^{k-1}]\in H_{1}(Y_{1}(N),\mathcal{M}_{2k-2}).
Definition 9.17.

For any integer k1k\geq 1, we define a \mathbb{Z}-submodule

Γ1(N),kH1(Y1(N),2k2)\mathfrak{Z}_{\Gamma_{1}(N),k}\subset H_{1}(Y_{1}(N),\mathcal{M}_{2k-2})

by

Γ1(N),k:=𝔷Γ1(N),k(γ)γΓ1(N) with |trace(γ)|>2.\mathfrak{Z}_{\Gamma_{1}(N),k}:=\langle\,\mathfrak{z}_{\Gamma_{1}(N),k}(\gamma)\mid\gamma\in\Gamma_{1}(N)\textrm{ with }|\mathrm{trace}(\gamma)|>2\,\rangle_{\mathbb{Z}}.
Remark 9.18.

By Remark 9.8, we have Γ1(1),k=k\mathfrak{Z}_{\Gamma_{1}(1),k}=\mathfrak{Z}_{k}.

Lemma 9.19.

For any integer N1N\geq 1 and prime number pp, we have

[{z,γz}]γΓ1(Np)Γ(p) with |trace(γ)|>2=Γ1(Np),1=H1(Y1(Np),).\langle\,[\{z,\gamma z\}]\mid\gamma\in\Gamma_{1}(Np)\!-\!\Gamma(p)\textrm{ with }|\mathrm{trace}(\gamma)|>2\,\rangle_{\mathbb{Z}}=\mathfrak{Z}_{\Gamma_{1}(Np),1}=H_{1}(Y_{1}(Np),\mathbb{Z}).
Proof.

It suffices to show that

γγΓ1(Np)Γ(p) with |trace(γ)|>2=Γ1(Np).\langle\,\gamma\mid\gamma\in\Gamma_{1}(Np)\!-\!\Gamma(p)\textrm{ with }|\mathrm{trace}(\gamma)|>2\,\rangle=\Gamma_{1}(Np).

Here \langle~{}\rangle means the group generated by the elements inside the bracket. Moreover, since Γ1(2)=Γ1(6),Γ1(10)\Gamma_{1}(2)=\langle\Gamma_{1}(6),\Gamma_{1}(10)\rangle, we may assume that Np3Np\geq 3. Put γ:=(11Np1+Np)\gamma:=\begin{pmatrix}1&1\\ Np&1+Np\end{pmatrix}. Then the quotient group Γ1(Np)/(Γ(p)Γ1(Np))\Gamma_{1}(Np)/(\Gamma(p)\cap\Gamma_{1}(Np)) is generated by the image of γ\gamma. For any matrix γΓ(p)Γ1(Np)\gamma^{\prime}\in\Gamma(p)\cap\Gamma_{1}(Np), we have γγ1+apΓ(p)\gamma^{\prime}\gamma^{1+ap}\not\in\Gamma(p) for any integer aa. Since trace(γ)=Np+2>2\mathrm{trace}(\gamma)=Np+2>2 and det(γ)=1\det(\gamma)=1, the matrix γ\gamma is hyperbolic, and one can take a matrix QGL2()Q\in\mathrm{GL}_{2}(\mathbb{R}) such that Q1γQ=(αα1)Q^{-1}\gamma Q=\begin{pmatrix}\alpha&\\ &\alpha^{-1}\end{pmatrix}. If we put Q1γQ=:(xyzw)Q^{-1}\gamma^{\prime}Q=:\begin{pmatrix}x&y\\ z&w\end{pmatrix}, then we have trace(γγ1+ap)=xα1+ap+wα1ap\mathrm{trace}(\gamma^{\prime}\gamma^{1+ap})=x\alpha^{1+ap}+w\alpha^{-1-ap}. Since trace(γγ1+ap)2(modNp)\mathrm{trace}(\gamma^{\prime}\gamma^{1+ap})\equiv 2\pmod{Np} and Np3Np\geq 3, we have trace(γγ1+ap)0\mathrm{trace}(\gamma^{\prime}\gamma^{1+ap})\neq 0. Hence the set {trace(γγ1+ap)a}\{\mathrm{trace}(\gamma^{\prime}\gamma^{1+ap})\mid a\in\mathbb{Z}\}\subset\mathbb{Z} is infinite. Therefore we can find an integer aa such that |trace(γγ1+ap)|>2|\mathrm{trace}(\gamma^{\prime}\gamma^{1+ap})|>2. ∎

Next, we recall an important result proved by Hida (see [Hid86, Corollary 4.5] or [Hid88, Corollary 8.2]). Take an integer N1N\geq 1 and a prime number pp, and put q:=pordp(Np)q:=p^{\mathrm{ord}_{p}(Np)}. Then we have a Γ1(q)\Gamma_{1}(q)-homomorphism

j:/(q)2k2/(q);bbX22k2,j\colon\mathbb{Z}/(q)\longrightarrow\mathcal{M}_{2k-2}\otimes\mathbb{Z}/(q);b\mapsto bX_{2}^{2k-2},

which induces a Hecke-equivariant homomorphism

j:H1(Y1(Np),/(q))H1(Y1(Np),2k2,p/(q)).j_{*}\colon H_{1}(Y_{1}(Np),\mathbb{Z}/(q))\longrightarrow H_{1}(Y_{1}(Np),\mathcal{M}_{2k-2,p}\otimes\mathbb{Z}/(q)).
Proposition 9.20.

When Np4Np\geq 4, the homomorphism jj induces a Hecke-equivariant isomorphism

j:H1ord(Y1(Np),/(q))H1ord(Y1(Np),2k2,p/(q)).j_{*}\colon H_{1}^{\mathrm{ord}}(Y_{1}(Np),\mathbb{Z}/(q))\stackrel{{\scriptstyle\sim}}{{\longrightarrow}}H_{1}^{\mathrm{ord}}(Y_{1}(Np),\mathcal{M}_{2k-2,p}\otimes\mathbb{Z}/(q)).

Here we define H1ord(Y1(Np),):=eUpH1(Y1(Np),)H_{1}^{\mathrm{ord}}(Y_{1}(Np),-):=e_{U_{p}}H_{1}(Y_{1}(Np),-).

Proof.

The proof of this proposition is essentially the same as that of [Hid93, Theorem 2 in §7.2] for cohomology groups. We note that Γ1(Np)\Gamma_{1}(Np) is torsion-free since Np4Np\geq 4. Hence any short exact sequence of Γ1(Np)\Gamma_{1}(Np)-modules induces a long exact sequence in homology.

If we put

𝒞:=(2k2/X22k2)/(q),\mathcal{C}:=(\mathcal{M}_{2k-2}/\mathbb{Z}X_{2}^{2k-2})\otimes\mathbb{Z}/(q),

then the short exact sequence 0/(q)𝑗2k2/(q)𝒞00\longrightarrow\mathbb{Z}/(q)\overset{j}{\longrightarrow}\mathcal{M}_{2k-2}\otimes\mathbb{Z}/(q)\longrightarrow\mathcal{C}\longrightarrow 0 induces an exact sequence of /(q)\mathbb{Z}/(q)-modules

H2(Y1(Np),𝒞)H1(Y1(Np),/(q))jH1(Y1(N),2k2/(q))H1(Y1(Np),𝒞).\displaystyle H_{2}(Y_{1}(Np),\mathcal{C})\longrightarrow H_{1}(Y_{1}(Np),\mathbb{Z}/(q))\stackrel{{\scriptstyle j_{*}}}{{\longrightarrow}}H_{1}(Y_{1}(N),\mathcal{M}_{2k-2}\otimes\mathbb{Z}/(q))\longrightarrow H_{1}(Y_{1}(Np),\mathcal{C}).

Since the operator UpU_{p} is defined by Up=u=0p1(1u0p)U_{p}=\sum_{u=0}^{p-1}\begin{pmatrix}1&u\\ 0&p\end{pmatrix}, we have

(1u0p)P(X1,X2)\displaystyle\begin{pmatrix}1&u\\ 0&p\end{pmatrix}\cdot P(X_{1},X_{2}) P(uX2,X2)(modp)\displaystyle\equiv P(-uX_{2},X_{2})\pmod{p}
𝔽pX22k2\displaystyle\in\mathbb{F}_{p}X_{2}^{2k-2}

for any polynomial P2k2P\in\mathcal{M}_{2k-2}. In other words, we have

(1u0p)(2k2/X22k1)p(2k2/X22k1),\begin{pmatrix}1&u\\ 0&p\end{pmatrix}\left(\mathcal{M}_{2k-2}/\mathbb{Z}X_{2}^{2k-1}\right)\subset p\left(\mathcal{M}_{2k-2}/\mathbb{Z}X_{2}^{2k-1}\right),

and this shows that H1ord(Y1(N),𝒞)=H2ord(Y1(N),𝒞)=0H_{1}^{\mathrm{ord}}(Y_{1}(N),\mathcal{C})=H_{2}^{\mathrm{ord}}(Y_{1}(N),\mathcal{C})=0, which completes the proof. ∎

Lemma 9.21.

Let NN be an integer and let pp be a prime number. Set q:=pordp(Np)q:=p^{\mathrm{ord}_{p}(Np)}. Then for any matrix γΓ1(Np)Γ(p)\gamma\in\Gamma_{1}(Np)\!-\!\Gamma(p), we have

j(𝔷Γ1(Np),1(γ)modq)(/(q))×𝔷Γ1(Np),k(γ).j_{*}(\mathfrak{z}_{\Gamma_{1}(Np),1}(\gamma)\bmod{q})\in(\mathbb{Z}/(q))^{\times}\cdot\mathfrak{z}_{\Gamma_{1}(Np),k}(\gamma).
Proof.

If we put γ:=(abcd)\gamma:=\begin{pmatrix}a&b\\ c&d\end{pmatrix}, then we have

Qγ(X1,X2)=sgn(a+d)gcd(c,ad,b)(cX12(ad)X1X2bX22).\displaystyle Q_{\gamma}(X_{1},X_{2})=-\frac{\operatorname{sgn}(a+d)}{\mathrm{gcd}(c,a-d,b)}(cX_{1}^{2}-(a-d)X_{1}X_{2}-bX_{2}^{2}).

Since γΓ1(Np)Γ(p)\gamma\in\Gamma_{1}(Np)\!-\!\Gamma(p), we have cad0(modq)c\equiv a-d\equiv 0\pmod{q} and b0(modp)b\not\equiv 0\pmod{p}, which shows that

Qγ(X1,X2)±bgcd(c,ad,b)X22(modq)Q_{\gamma}(X_{1},X_{2})\equiv\pm\frac{b}{\mathrm{gcd}(c,a-d,b)}X_{2}^{2}\pmod{q}

and

j(𝔷Γ1(Np),1(γ)modq)\displaystyle j_{*}(\mathfrak{z}_{\Gamma_{1}(Np),1}(\gamma)\bmod{q}) =(±gcd(c,ad,b)b)k1𝔷Γ1(Np),k(γ)(modq)\displaystyle=\left(\pm\frac{\mathrm{gcd}(c,a-d,b)}{b}\right)^{k-1}\mathfrak{z}_{\Gamma_{1}(Np),k}(\gamma)\pmod{q}
(/(q))×𝔷Γ1(Np),k(γ).\displaystyle\in(\mathbb{Z}/(q))^{\times}\cdot\mathfrak{z}_{\Gamma_{1}(Np),k}(\gamma).

Theorem 9.22.

For any integer N1N\geq 1 and prime number pp satisfying Np4Np\geq 4, we have

eUp(Γ1(Np),kp)=H1ord(Y1(Np),2k2,p).e_{U_{p}}(\mathfrak{Z}_{\Gamma_{1}(Np),k}\otimes\mathbb{Z}_{p})=H_{1}^{\mathrm{ord}}(Y_{1}(Np),\mathcal{M}_{2k-2,p}).
Proof.

We first note that eUpH0(Y1(Np),2k2,p)=eUp((2k2,p)Γ1(Np))e_{U_{p}}H_{0}(Y_{1}(Np),\mathcal{M}_{2k-2,p})=e_{U_{p}}((\mathcal{M}_{2k-2,p})_{\Gamma_{1}(Np)}) vanishes since UpX22k2=pX22k2U_{p}\cdot X_{2}^{2k-2}=pX_{2}^{2k-2} and (1u0p)P(X1,X2)P(uX2,X2)(modp)\begin{pmatrix}1&u\\ 0&p\end{pmatrix}\cdot P(X_{1},X_{2})\equiv P(-uX_{2},X_{2})\pmod{p}. This fact implies that

H1ord(Y1(Np),2k2,p)/(q)H1ord(Y1(Np),2k2,p/(q)).H_{1}^{\mathrm{ord}}(Y_{1}(Np),\mathcal{M}_{2k-2,p})\otimes\mathbb{Z}/(q)\stackrel{{\scriptstyle\sim}}{{\longrightarrow}}H_{1}^{\mathrm{ord}}(Y_{1}(Np),\mathcal{M}_{2k-2,p}\otimes\mathbb{Z}/(q)).

Here q:=pordp(Np)q:=p^{\mathrm{ord}_{p}(Np)}. Hence this theorem follows from Proposition 9.20 and Lemmas 9.19 and 9.21. ∎

9.3.2. Proof of Theorem 9.15

Let k2k\geq 2 be an integer. For any positive integers MM and NN with MNM\mid N, we denote by

πN,M\displaystyle\pi_{*}^{N,M} :H1(Y1(N),2k2)H1(Y1(M),2k2),\displaystyle\colon H_{1}(Y_{1}(N),\mathcal{M}_{2k-2})\longrightarrow H_{1}(Y_{1}(M),\mathcal{M}_{2k-2}),
πM,N\displaystyle\pi^{*}_{M,N} :H1(Y1(M),2k2)H1(Y1(N),2k2)\displaystyle\colon H^{1}(Y_{1}(M),\mathcal{M}_{2k-2})\longrightarrow H^{1}(Y_{1}(N),\mathcal{M}_{2k-2})

the homomorphisms induced by the natural projection Y1(N)Y1(M);zzY_{1}(N)\longrightarrow Y_{1}(M);z\mapsto z.

Corollary 9.23.

For any integer N1N\geq 1 and prime number pp, we have

pEis2k2,πNp,1(eUp(Γ1(Np),kp)).\mathbb{Z}_{p}\subset\langle\mathrm{Eis}_{2k-2},\pi^{Np,1}_{*}(e_{U_{p}}(\mathfrak{Z}_{\Gamma_{1}(Np),k}\otimes\mathbb{Z}_{p}))\rangle.
Proof.

Take an element τ\tau\in\mathbb{H}. Since Up([{τ,τ+1}X22k2])=[{τp,τp+1}X22k2]=[{τ,τ+1}X22k2]U_{p}([\{\tau,\tau+1\}\otimes X_{2}^{2k-2}])=[\{\frac{\tau}{p},\frac{\tau}{p}+1\}\otimes X_{2}^{2k-2}]=[\{\tau,\tau+1\}\otimes X_{2}^{2k-2}], we have

[{τ,τ+1}X22k2]H1ord(Y1(Np)BS,2k2,p)H1ord(Y1(Np),2k2,p).[\{\tau,\tau+1\}\otimes X_{2}^{2k-2}]\in H_{1}^{\mathrm{ord}}(\partial Y_{1}(Np)^{\mathrm{BS}},\mathcal{M}_{2k-2,p})\subset H_{1}^{\mathrm{ord}}(Y_{1}(Np),\mathcal{M}_{2k-2,p}).

On the other hand, since

πNp,1([{τ,τ+1}X22k2])=[{τ,τ+1}X22k2]H1(Y,2k2),\pi^{Np,1}_{*}([\{\tau,\tau+1\}\otimes X_{2}^{2k-2}])=[\{\tau,\tau+1\}\otimes X_{2}^{2k-2}]\in H_{1}(Y,\mathcal{M}_{2k-2}),

we have

Eis2k2,πNp,1([{τ,τ+1}X22k2])=1,\langle\mathrm{Eis}_{2k-2},\pi^{Np,1}_{*}([\{\tau,\tau+1\}\otimes X_{2}^{2k-2}])\rangle=1,

Therefore, first when Np4Np\geq 4, Theorem 9.22 implies this lemma. When Np3Np\leq 3, we have N=1N=1 and p|6p|6. Then this case follows from the case N=3N=3 and p6p\mid 6 since π3p,1(eUp(Γ1(3p),kp)πp,1(eUp(Γ1(p),kp)\pi^{3p,1}_{*}(e_{U_{p}}(\mathfrak{Z}_{\Gamma_{1}(3p),k}\otimes\mathbb{Z}_{p})\subset\pi^{p,1}_{*}(e_{U_{p}}(\mathfrak{Z}_{\Gamma_{1}(p),k}\otimes\mathbb{Z}_{p}). ∎

Lemma 9.24.

Let N1N\geq 1 be an integer and let pp be a prime number. Then for any homology class xH1(Y1(Np),2k2)px\in H_{1}(Y_{1}(Np),\mathcal{M}_{2k-2})\otimes\mathbb{C}_{p} we have

eTpπNp,1(eUpx)=πNp,1(eUpx).e_{T_{p}}\pi_{*}^{Np,1}(e_{U_{p}}x)=\pi_{*}^{Np,1}(e_{U_{p}}x).
Proof.

By using the formal duality, it suffices to show that

eUpπ1,Np(eTpy)=eUpπ1,Np(y)e_{U_{p}^{\prime}}\pi_{1,Np}^{*}(e_{T_{p}^{\prime}}y)=e_{U_{p}^{\prime}}\pi_{1,Np}^{*}(y)

for any cohomology class yH1(Y,2k2)py\in H^{1}(Y,\mathcal{M}_{2k-2})\otimes\mathbb{C}_{p}. This claim is well-known (see [Gou92, Lemma 2] for example). ∎

Corollary 9.25.

For any prime number p5p\geq 5, we have

Eis2k2,πp,1(eUp(Γ1(p),kp))=1N2kp.\langle\mathrm{Eis}_{2k-2},\pi_{*}^{p,1}(e_{U_{p}}(\mathfrak{Z}_{\Gamma_{1}(p),k}\otimes\mathbb{Z}_{p}))\rangle=\frac{1}{N_{2k}}\mathbb{Z}_{p}.
Proof.

Take a prime number p5p\geq 5. Then Theorem 9.22 shows that

Eis2k2,πp,1(eUp(Γ1(p),kp))=Eis2k2,πp,1(H1ord(Y1(p),2k2,p)).\langle\mathrm{Eis}_{2k-2},\pi_{*}^{p,1}(e_{U_{p}}(\mathfrak{Z}_{\Gamma_{1}(p),k}\otimes\mathbb{Z}_{p}))\rangle=\langle\mathrm{Eis}_{2k-2},\pi_{*}^{p,1}(H_{1}^{\mathrm{ord}}(Y_{1}(p),\mathcal{M}_{2k-2,p}))\rangle.

By Lemma 9.24, we have a natural homomorphism

πp,1:H1ord(Y1(p),2k2,p)/(torsion)H1ord(Y,2k2,p)/(torsion),\pi_{*}^{p,1}\colon H_{1}^{\mathrm{ord}}(Y_{1}(p),\mathcal{M}_{2k-2,p})/(\mathrm{torsion})\longrightarrow H_{1}^{\mathrm{ord}}(Y,\mathcal{M}_{2k-2,p})/(\mathrm{torsion}),

which is the dual of the homomorphism

Hord1(Y,2k2,p)Hord1(Y1(p),2k2,p);yeUpπ1,p(y).H^{1}_{\mathrm{ord}}(Y,\mathcal{M}_{2k-2,p})\longrightarrow H^{1}_{\mathrm{ord}}(Y_{1}(p),\mathcal{M}_{2k-2,p});y\mapsto e_{U_{p}^{\prime}}\pi^{*}_{1,p}(y).

The fact that the index [Γ0(p):Γ1(p)]=p1[\Gamma_{0}(p)\colon\Gamma_{1}(p)]=p-1 is relatively prime to pp together with the isomorphism (6.4) implies that πp,1:H1ord(Y1(p),2k2,p)/(torsion)H1ord(Y,2k2,p)/(torsion)\pi_{*}^{p,1}\colon H_{1}^{\mathrm{ord}}(Y_{1}(p),\mathcal{M}_{2k-2,p})/(\mathrm{torsion})\longrightarrow H_{1}^{\mathrm{ord}}(Y,\mathcal{M}_{2k-2,p})/(\mathrm{torsion}) is surjective. Hence we have

Eis2k2,πp,1(H1ord(Y1(p),2k2,p))=Eis2k2,H1ord(Y,2k2,p).\displaystyle\langle\mathrm{Eis}_{2k-2},\pi_{*}^{p,1}(H_{1}^{\mathrm{ord}}(Y_{1}(p),\mathcal{M}_{2k-2,p}))\rangle=\langle\mathrm{Eis}_{2k-2},H_{1}^{\mathrm{ord}}(Y,\mathcal{M}_{2k-2,p})\rangle.

Therefore, this corollary follows from Theorem 2.13 and Lemma 5.5. ∎

We note that for any prime number pp, the operators Vp:=(p001)V_{p}:=\begin{pmatrix}p&0\\ 0&1\end{pmatrix} and Vp:=(p001)~V_{p}^{\prime}:=\widetilde{\begin{pmatrix}p&0\\ 0&1\end{pmatrix}} induce homomorphisms

Vp\displaystyle V_{p} :H1(Y1(Np),2k2)H1(Y1(N),2k2),\displaystyle\colon H_{1}(Y_{1}(Np),\mathcal{M}_{2k-2})\longrightarrow H_{1}(Y_{1}(N),\mathcal{M}_{2k-2}),
Vp\displaystyle V_{p}^{\prime} :H1(Y1(N),2k2)H1(Y1(Np),2k2).\displaystyle\colon H^{1}(Y_{1}(N),\mathcal{M}_{2k-2})\longrightarrow H^{1}(Y_{1}(Np),\mathcal{M}_{2k-2}).
Lemma 9.26.

For any prime number pp and integer N1N\geq 1, we have

eUp(π1,Np(Eis2k2))=11p2k1(π1,Np(Eis2k2)π1,N(Eis2k2)|Vp).e_{U_{p}^{\prime}}(\pi^{*}_{1,Np}(\mathrm{Eis}_{2k-2}))=\frac{1}{1-p^{2k-1}}\left(\pi^{*}_{1,Np}(\mathrm{Eis}_{2k-2})-\pi^{*}_{1,N}(\mathrm{Eis}_{2k-2})|V_{p}^{\prime}\right).
Proof.

We put

Eis2k2(1)\displaystyle\mathrm{Eis}_{2k-2}^{(1)} :=11p2k1(π1,Np(Eis2k2)π1,N(Eis2k2)|Vp),\displaystyle:=\frac{1}{1-p^{2k-1}}\left(\pi^{*}_{1,Np}(\mathrm{Eis}_{2k-2})-\pi^{*}_{1,N}(\mathrm{Eis}_{2k-2})|V_{p}^{\prime}\right),
Eis2k2(p2k1)\displaystyle\mathrm{Eis}_{2k-2}^{(p^{2k-1})} :=11p2k1(p2k1π1,Np(Eis2k2)+π1,N(Eis2k2)|Vp).\displaystyle:=\frac{1}{1-p^{2k-1}}\left(-p^{2k-1}\pi^{*}_{1,Np}(\mathrm{Eis}_{2k-2})+\pi^{*}_{1,N}(\mathrm{Eis}_{2k-2})|V_{p}^{\prime}\right).

Since π1,NpTp=Upπ1,Np+Vpπ1,N\pi_{1,Np}^{*}T_{p}^{\prime}=U_{p}^{\prime}\pi_{1,Np}^{*}+V_{p}^{\prime}\pi_{1,N}^{*} and UpVpπ1,N=p2k1π1,NpU_{p}^{\prime}V_{p}^{\prime}\pi_{1,N}^{*}=p^{2k-1}\pi_{1,Np}^{*}, the relation that Eis2k2|Tp=(1+p2k1)Eis2k2\mathrm{Eis}_{2k-2}|T_{p}^{\prime}=(1+p^{2k-1})\mathrm{Eis}_{2k-2} shows that

Eis2k2(1)|Up=Eis2k2(1) and Eis2k2(p2k1)|Up=p2k1Eis2k2(p2k1).\displaystyle\mathrm{Eis}_{2k-2}^{(1)}|U_{p}^{\prime}=\mathrm{Eis}_{2k-2}^{(1)}\,\,\,\textrm{ and }\,\,\,\mathrm{Eis}_{2k-2}^{(p^{2k-1})}|U_{p}^{\prime}=p^{2k-1}\mathrm{Eis}_{2k-2}^{(p^{2k-1})}.

Since π1,Np(Eis2k2)=Eis2k2(1)+Eis2k2(p2k1)\pi^{*}_{1,Np}(\mathrm{Eis}_{2k-2})=\mathrm{Eis}_{2k-2}^{(1)}+\mathrm{Eis}_{2k-2}^{(p^{2k-1})}, these facts imply this lemma. ∎

Lemma 9.27.

For any integer N1N\geq 1 and a prime number pp, we have

Eis2k2,πNp,1(eUp(Γ1(Np),kp))Eis2k2,Γ1(1),kp.\langle\mathrm{Eis}_{2k-2},\pi^{Np,1}_{*}(e_{U_{p}}(\mathfrak{Z}_{\Gamma_{1}(Np),k}\otimes\mathbb{Z}_{p}))\rangle\subset\langle\mathrm{Eis}_{2k-2},\mathfrak{Z}_{\Gamma_{1}(1),k}\rangle\mathbb{Z}_{p}.
Proof.

Lemma 9.26 implies that

eUpπNp,1(Eis2k2),Γ1(Np),k\displaystyle\langle e_{U_{p}^{\prime}}\pi^{*}_{Np,1}(\mathrm{Eis}_{2k-2}),\mathfrak{Z}_{\Gamma_{1}(Np),k}\rangle π1,Np(Eis2k2),Γ1(Np),kp+Vpπ1,N(Eis2k2),Γ1(Np),kp\displaystyle\subset\langle\pi^{*}_{1,Np}(\mathrm{Eis}_{2k-2}),\mathfrak{Z}_{\Gamma_{1}(Np),k}\rangle\mathbb{Z}_{p}+\langle V_{p}^{\prime}\pi^{*}_{1,N}(\mathrm{Eis}_{2k-2}),\mathfrak{Z}_{\Gamma_{1}(Np),k}\rangle\mathbb{Z}_{p}
Eis2k2,Γ1(1),kp+Eis2k2,πN,1(VpΓ1(Np),k)p.\displaystyle\subset\langle\mathrm{Eis}_{2k-2},\mathfrak{Z}_{\Gamma_{1}(1),k}\rangle\mathbb{Z}_{p}+\langle\mathrm{Eis}_{2k-2},\pi_{*}^{N,1}(V_{p}\mathfrak{Z}_{\Gamma_{1}(Np),k})\rangle\mathbb{Z}_{p}.

Hence we have

Eis2k2,πNp,1(eUp(Γ1(Np),kp))\displaystyle\langle\mathrm{Eis}_{2k-2},\pi^{Np,1}_{*}(e_{U_{p}}(\mathfrak{Z}_{\Gamma_{1}(Np),k}\otimes\mathbb{Z}_{p}))\rangle =eUpπ1,Np(Eis2k2),Γ1(N),kp\displaystyle=\langle e_{U_{p}^{\prime}}\pi^{*}_{1,Np}(\mathrm{Eis}_{2k-2}),\mathfrak{Z}_{\Gamma_{1}(N),k}\rangle\mathbb{Z}_{p}
Eis2k2,Γ1(1),kp+Eis2k2,πN,1(VpΓ1(Np),k)p.\displaystyle\subset\langle\mathrm{Eis}_{2k-2},\mathfrak{Z}_{\Gamma_{1}(1),k}\rangle\mathbb{Z}_{p}+\langle\mathrm{Eis}_{2k-2},\pi_{*}^{N,1}(V_{p}\mathfrak{Z}_{\Gamma_{1}(Np),k})\rangle\mathbb{Z}_{p}.

Therefore, it suffices to show that VpΓ1(Np),kΓ1(N),kV_{p}\mathfrak{Z}_{\Gamma_{1}(Np),k}\subset\mathfrak{Z}_{\Gamma_{1}(N),k}. Let γΓ1(Np)\gamma\in\Gamma_{1}(Np) be a matrix. Then we have γp:=(p001)γ(p001)1Γ1(N)\gamma_{p}:=\begin{pmatrix}p&0\\ 0&1\end{pmatrix}\gamma\begin{pmatrix}p&0\\ 0&1\end{pmatrix}^{-1}\in\Gamma_{1}(N). Moreover, by the definitions of 𝔷Γ1(N),k\mathfrak{z}_{\Gamma_{1}(N),k} and 𝔷Γ1(Np),k\mathfrak{z}_{\Gamma_{1}(Np),k}, we obtain

Vp𝔷Γ1(Np),k(γ)𝔷Γ1(N),k(γp).V_{p}\cdot\mathfrak{z}_{\Gamma_{1}(Np),k}(\gamma)\in\mathbb{Z}\mathfrak{z}_{\Gamma_{1}(N),k}(\gamma_{p}).

In particular, we have VpΓ1(Np),kΓ1(N),kV_{p}\mathfrak{Z}_{\Gamma_{1}(Np),k}\subset\mathfrak{Z}_{\Gamma_{1}(N),k}. ∎

Proof of Theorem 9.15.

By Theorem 2.13 and Remark 9.18, we only need to show that

1N2kpEis2k2,Γ1(1),kp\frac{1}{N_{2k}}\mathbb{Z}_{p}\subset\langle\mathrm{Eis}_{2k-2},\mathfrak{Z}_{\Gamma_{1}(1),k}\rangle\mathbb{Z}_{p}

for any prime number pp. When p5p\geq 5, this claim follows from Corollary 9.25 and Lemma 9.27 applied to N=1N=1. Suppose p=2p=2 or p=3p=3. Then Since the primes 22 and 33 are regular, these primes does not divide N2kN_{2k}. Hence, Corollary 9.23 and Lemma 9.27 show that

1N2kp=pEis2k2,Γ1(1),kp.\frac{1}{N_{2k}}\mathbb{Z}_{p}=\mathbb{Z}_{p}\subset\langle\mathrm{Eis}_{2k-2},\mathfrak{Z}_{\Gamma_{1}(1),k}\rangle\mathbb{Z}_{p}.

References

  • [BCG20] Nicolas Bergeron, Pierre Charollois, and Luis E. Garcia, Transgressions of the Euler class and Eisenstein cohomology of GLN(𝐙){\rm GL}_{N}({\bf Z}), Jpn. J. Math. 15 (2020), no. 2, 311–379. MR 4120422
  • [Bel21] Joël Bellaïche, The eigenbook—eigenvarieties, families of Galois representations, pp-adic LL-functions, Pathways in Mathematics, Birkhäuser/Springer, Cham, [2021] ©2021. MR 4306639
  • [Ber08] Tobias Berger, Denominators of Eisenstein cohomology classes for GL2{\rm GL}_{2} over imaginary quadratic fields, Manuscripta Math. 125 (2008), no. 4, 427–470. MR 2392880
  • [Ber09] by same author, On the Eisenstein ideal for imaginary quadratic fields, Compos. Math. 145 (2009), no. 3, 603–632. MR 2507743
  • [BHYY22] Kenichi Bannai, Kei Hagihara, Kazuki Yamada, and Shuji Yamamoto, pp-adic polylogarithms and pp-adic Hecke LL-functions for totally real fields, J. Reine Angew. Math. 791 (2022), 53–87. MR 4489625
  • [BKL18] Alexander Beilinson, Guido Kings, and Andrey Levin, Topological polylogarithms and pp-adic interpolation of LL-values of totally real fields, Math. Ann. 371 (2018), no. 3-4, 1449–1495. MR 3831278
  • [Bra23] Romain Branchereau, An upper bound on the denominator of Eisenstein classes in Bianchi manifolds, Preprint, arXiv:2305.11341, 2023.
  • [Car61] L. Carlitz, A generalization of Maillet’s determinant and a bound for the first factor of the class number, Proc. Amer. Math. Soc. 12 (1961), 256–261. MR 121354
  • [CDG15] Pierre Charollois, Samit Dasgupta, and Matthew Greenberg, Integral Eisenstein cocycles on 𝔾𝕃n\mathbb{GL}_{n}, II: Shintani’s method, Comment. Math. Helv. 90 (2015), no. 2, 435–477. MR 3351752
  • [CN79] Pierrette Cassou-Noguès, Valeurs aux entiers négatifs des fonctions zêta et fonctions zêta pp-adiques, Invent. Math. 51 (1979), no. 1, 29–59. MR 524276
  • [Cox13] David A. Cox, Primes of the form x2+ny2x^{2}+ny^{2}, second ed., Pure and Applied Mathematics (Hoboken), John Wiley & Sons, Inc., Hoboken, NJ, 2013, Fermat, class field theory, and complex multiplication. MR 3236783
  • [CS74a] J. Coates and W. Sinnott, An analogue of Stickelberger’s theorem for the higher KK-groups, Invent. Math. 24 (1974), 149–161. MR 369322
  • [CS74b] by same author, On pp-adic LL-functions over real quadratic fields, Invent. Math. 25 (1974), 253–279. MR 354615
  • [CS77] by same author, Integrality properties of the values of partial zeta functions, Proc. London Math. Soc. (3) 34 (1977), no. 2, 365–384. MR 439815
  • [DIT18] W. Duke, Ö. Imamoḡlu, and Á. Tóth, Kronecker’s first limit formula, revisited, Res. Math. Sci. 5 (2018), no. 2, Paper No. 20, 21. MR 3782448
  • [DR80] Pierre Deligne and Kenneth A. Ribet, Values of abelian LL-functions at negative integers over totally real fields, Invent. Math. 59 (1980), no. 3, 227–286. MR 579702
  • [Duk23] William Duke, Higher rademacher symbols, Experimental Mathematics 0 (2023), no. 0, 1–20.
  • [Gor05] Mark Goresky, Compactifications and cohomology of modular varieties, Harmonic analysis, the trace formula, and Shimura varieties, Clay Math. Proc., vol. 4, Amer. Math. Soc., Providence, RI, 2005, pp. 551–582. MR 2192016
  • [Gou92] Fernando Q. Gouvêa, On the ordinary Hecke algebra, J. Number Theory 41 (1992), no. 2, 178–198. MR 1164797
  • [Hab83] Klaus Haberland, Perioden von Modulformen einer Variabler and Gruppencohomologie. I, II, III, Math. Nachr. 112 (1983), 245–282, 283–295, 297–315. MR 726861
  • [Har] Günter Harder, Cohomology of Arithmetic Gruops, available on Harder’s webpage, http://www.math.uni-bonn.de/people/harder/Manuscripts/buch/.
  • [Har81] by same author, Period integrals of cohomology classes which are represented by Eisenstein series, Automorphic forms, representation theory and arithmetic (Bombay, 1979), Tata Inst. Fund. Res. Studies in Math, vol. 10, Tata Institute of Fundamental Research, Bombay, 1981, pp. pp 41–115. MR 633658
  • [Har82] by same author, Period integrals of Eisenstein cohomology classes and special values of some LL-functions, Number theory related to Fermat’s last theorem (Cambridge, Mass., 1981), Progr. Math., vol. 26, Birkhäuser Boston, Boston, MA, 1982, pp. 103–142. MR 685293
  • [Har11] by same author, Interpolating coefficient systems and pp-ordinary cohomology of arithmetic groups, Groups Geom. Dyn. 5 (2011), no. 2, 393–444. MR 2782179
  • [Har21] by same author, Can we compute denominators of Eisenstein classes?, Adv. Stud. Euro-Tbil. Math. J. Special Issue (2021), no. 9, 131–165, Special issue on Cohomology, Geometry, Explicit Number Theory.
  • [Hid86] Haruzo Hida, Galois representations into GL2(𝐙p[[X]]){\rm GL}_{2}({\bf Z}_{p}[[X]]) attached to ordinary cusp forms, Invent. Math. 85 (1986), no. 3, 545–613. MR 848685
  • [Hid88] by same author, On pp-adic Hecke algebras for GL2{\rm GL}_{2} over totally real fields, Ann. of Math. (2) 128 (1988), no. 2, 295–384. MR 960949
  • [Hid93] by same author, Elementary theory of LL-functions and Eisenstein series, London Mathematical Society Student Texts, vol. 26, Cambridge University Press, Cambridge, 1993. MR 1216135
  • [HP92] Günter Harder and Richard Pink, Modular konstruierte unverzweigte abelsche pp-Erweiterungen von 𝐐(ζp){\bf Q}(\zeta_{p}) und die Struktur ihrer Galoisgruppen, Math. Nachr. 159 (1992), 83–99. MR 1237103
  • [Kai90] Christian Kaiser, Die nenner von eisensteinklassen für gewisse kongruenzgruppen, Diploma thesis, 1990, Rheinische Friedrich-Wilhelms-Universität Bonn.
  • [Kol04] Manfred Kolster, KK-theory and arithmetic, Contemporary developments in algebraic KK-theory, ICTP Lect. Notes, vol. XV, Abdus Salam Int. Cent. Theoret. Phys., Trieste, 2004, pp. 191–258. MR 2175640
  • [Mae93] Hartmut Maennel, Nenner von Eisensteinklassen auf Hilbertschen Modulvarietäten und die pp-adische Klassenzahlformel, Bonner Mathematische Schriften [Bonn Mathematical Publications], vol. 247, Universität Bonn, Mathematisches Institut, Bonn, 1993, Dissertation, Rheinische Friedrich-Wilhelms-Universität Bonn, Bonn, 1992. MR 1290786
  • [Mah00] Joachim Mahnkopf, Eisenstein cohomology and the construction of pp-adic analytic LL-functions, Compositio Math. 124 (2000), no. 3, 253–304. MR 1809337
  • [O’S23] Cormac O’Sullivan, Integrality of the higher Rademacher symbols, Preprint, arXiv:2306.08831, 2023.
  • [Scz93] Robert Sczech, Eisenstein group cocycles for GLn{\rm GL}_{n} and values of LL-functions, Invent. Math. 113 (1993), no. 3, 581–616. MR 1231838
  • [Ser73] Jean-Pierre Serre, Congruences et formes modulaires [d’après H. P. F. Swinnerton-Dyer], Séminaire Bourbaki, 24e année (1971/1972), Exp. No. 416, Lecture Notes in Math, vol. Vol. 317, Springer, Berlin, 1973, pp. pp 319–338. MR 466020
  • [Shi71] Goro Shimura, Introduction to the arithmetic theory of automorphic functions, vol. No. 1., Iwanami Shoten Publishers, Tokyo; Princeton University Press, Princeton, N.J., 1971, Publications of the Mathematical Society of Japan, No. 11. MR 314766
  • [Shi76] Takuro Shintani, On evaluation of zeta functions of totally real algebraic number fields at non-positive integers, J. Fac. Sci. Univ. Tokyo Sect. IA Math. 23 (1976), no. 2, 393–417. MR 427231
  • [Sie80] Carl Ludwig Siegel, Advanced analytic number theory, second ed., vol. 9., Tata Institute of Fundamental Research, Bombay, 1980. MR 659851
  • [Sku80] Ladislav Skula, Index of irregularity of a prime, J. Reine Angew. Math. 315 (1980), 92–106. MR 564526
  • [VZ13] Maria Vlasenko and Don Zagier, Higher Kronecker “limit” formulas for real quadratic fields, J. Reine Angew. Math. 679 (2013), 23–64. MR 3065153
  • [Wan89] Xiang Dong Wang, Die Eisensteinklasse in H1(SL2(𝐙),Mn(𝐙))H^{1}({\rm SL}_{2}({\bf Z}),M_{n}({\bf Z})) und die Arithmetik spezieller Werte von LL-Funktionen, Bonner Mathematische Schriften [Bonn Mathematical Publications], vol. 202, Universität Bonn, Mathematisches Institut, Bonn, 1989, Dissertation, Rheinische Friedrich-Wilhelms-Universität Bonn, Bonn, 1989. MR 1015506
  • [Was97] Lawrence C. Washington, Introduction to cyclotomic fields, second ed., Graduate Texts in Mathematics, vol. 83, Springer-Verlag, New York, 1997. MR 1421575
  • [Wes88] Uwe Weselmann, Eisensteinkohomologie und Dedekindsummen für GL2{\rm GL}_{2} über imaginär-quadratischen Zahlkörpern, J. Reine Angew. Math. 389 (1988), 90–121. MR 953667
  • [Wil90] Andrew Wiles, The Iwasawa conjecture for totally real fields, Ann. of Math. (2) 131 (1990), no. 3, 493–540. MR 1053488
  • [Zag76] Don Zagier, On the values at negative integers of the zeta-function of a real quadratic field, Enseign. Math. (2) 22 (1976), no. 1-2, 55–95. MR 406957
  • [Zag77] by same author, Valeurs des fonctions zêta des corps quadratiques réels aux entiers négatifs, Journées Arithmétiques de Caen (Univ. Caen, Caen, 1976), Astérisque No. 41–42, Soc. Math. France, Paris, 1977, pp. pp 135–151. MR 441925