Harder’s denominator problem for and its applications
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 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 conjecture2020 Mathematics Subject Classification:
11F75, 11R42, 11F11, 11F67, 11R231. Introduction
1.1. The denominator of Eisenstein classes for
Let be the modular group, and let be the upper half plane, on which acts by the linear fractional transformation. We denote by the modular curve of level , by its Borel–Serre compactification, and by the Borel–Serre boundary of . For any even integer , we define a left -module to be the -th symmetric power of , and define to be its dual -module. Then the -module naturally defines a sheaf on (which we also denote by ), and we can consider the cohomology groups and . Since has a left action of , these cohomology groups carry the structure of Hecke modules.
The boundary is identified with , where . Hence it is easy to see that and we have a natural generator .
Harder considered in his book [Har] a unique Hecke-equivariant section
of the canonical homomorphism induced by the inclusion map , and he defined the Eisenstein cohomology class
to be the image of under this section. Then Harder studied the denominator of the Eisenstein cohomology class , that is, the smallest positive integer such that
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 , we have
where 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.
Remark 1.3.
For any prime numbers and with , the denominators over of Eisenstein classes for 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 ). Eisenstein classes for over totally real fields have been studied by Maennel in the dissertation [Mae93]. Eisenstein classes for 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 and Theorem 1.1 in terms of the holomorphic Eisenstein series and the Eichler–Shimura homomorphism.
In the following, let be an even integer and let denotes the space of modular forms of level and weight . Then we have the Hecke-equivariant homomorphism
called the Eichler–Shimura homomorphism which is defined by certain path integrals on (see §2.6 for the precise definition of the Eichler–Shimura homomorphism). Let
denote the holomorphic Eisenstein series of weight . 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)
We have , i.e., the class coincides with the Eisenstein class .
-
(2)
The denominator of is equal to the numerator of .
Remark 1.5.
From the -expansion of the Eisenstein series , we see that the denominator of with respect to the integral structure coming from the -expansion (the de Rham integral structure) is clearly the numerator of . An interesting and non-trivial point in Theorem 1.4 is that on the Eisenstein parts and , 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 , we have
Therefore, it suffices to prove that
for each prime number . Then the proof consists roughly of the following four parts.
- (I)
-
(II)
Next, in §4, we compute the -adic limit of the value of the pairing
where is the pairing induced by . More precisely, we will show in Theorem 4.1 and Corollary 7.2 that this -adic limit can be described in terms of the Kubota–Leopoldt -adic -functions, namely, for any integer we obtain the following interesting formula
where
and denotes the Kubota–Leopoldt -adic -function associated with the -th power of the Teichmüller character .
- (III)
- (IV)
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
for each integer 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 and matrix , we have
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 be a real quadratic field, be an order in , and be a narrow ideal class of . Then we have the associated partial zeta function
which can be continued meromorphically to , and it is known that
for any integer . We also define the positive integer by
Then in §9.2, we obtain the following as another consequence of Theorem 1.4.
Proposition 1.8 (Corollary 9.12).
Let be an integer. Then the integer gives a universal upper bound for the denominator of with respect to orders and narrow ideal classes . In other words, we have
for all orders in all real quadratic fields and narrow ideal classes .
In fact, one can construct a natural map for any integer (see Definition 9.6), and we show in Proposition 9.10 that
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
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 , 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 , 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 , we have
which implies that
(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 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 of the denominators of the values and examined its sharpness briefly. More precisely, he observed that the upper bound is not sharp and discussed how one can improve this upper bound when (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 .
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, denotes an even integer.
2.1. Definitions of Modular curve and Borel–Serre compactification
Let
denote the upper half plane, and let
be the Borel–Serre compactification of (see [Gor05] or [Har, §1,2,7]). We set
The group acts on and by the linear fractional transformation as usual. We denote by
the modular curve of level and its Borel–Serre compactification, respectively. Moreover, we denote by the boundary of . The boundary is homeomorphic to the circle and the fundamental group can be identified with .
In the following, (resp. ) means that it is either or (resp. or ). Any left -module can be regarded as (co)sheaf on in a natural way, and hence we can consider the homology groups
which fit into the long exact sequence
Similarly, we have the cohomology groups
which fit into the long exact sequence
We note that the inclusion map induces isomorphisms
Moreover, if has an action of , then these homology groups (resp. cohomology groups) carry the structure of Hecke modules, namely, for each prime number we have a Hecke operator (resp. ) 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 , the stalk at coincides with , where is a lift of and . This fact shows that a short exact sequence of left -modules does not give a short exact sequence of sheaves on in general, that is, the sheafification functor is not exact. However, a short exact sequence of left ]-modules induces a short exact sequence of sheaves on since the order of divides .
2.2. Modular symbols and (co)homology
Let and let denote the usual singular chain complex of , i.e., is the free abelian group generated by singular -simplices in and is the boundary operator.
The left action of on induces a left action of on , and is actually an -equivariant complex. Then it is known that for any left -module , which is also seen as a (co)sheaf on , we have natural isomorphisms
where denotes the -coinvariant functor. Here the left -action on is defined by
where , , and . Set
For any elements , we denote the equivalence class of a path from to in by
The boundary map induces a -homomorphism , and we have a natural isomorphism
Similarly, we also have natural isomorphisms
where denotes the -invariant functor. Here the left -action on is defined by
where , , and is the adjugate of . Since
we have a natural isomorphism
2.3. -modules and
For any matrix , we denote the adjugate of by
Note that if , then we have .
Let denote the ring of polynomials of two variables over , and we equip with a left action of by
where and . For each integer , we set
We then define submodules and of by
The -modules and are closed under the left action of on . In particular, both and are left -modules. We also define the pairing
by
where is the Kronecker delta. The pairing is perfect and -equivariant in the sense that for any polynomials and and matrix , we have
Hence the pairing induces an -equivariant isomorphism
Here the left action of on is given by
where , , and .
Remark 2.2.
The left actions of on and are slightly different from the left actions used in Harder’s book [Har, (1.57)]. However, since
they are isomorphic as left -modules. Therefore, there are no essential differences.
2.4. Hecke operators
Let and let be a left -module. In this subsection, we define the Hecke operators on and , explicitly.
For each prime number , we have the following double coset decomposition:
Hence the endomorphism
induces an endomorphism of . Similarly, the endomorphism
induces an endomorphism of .
Definition 2.3.
Let be a prime number.
-
(1)
We define the Hecke operator at on by
for any simplex and element . The operator induces operators on and , etc., also written as .
-
(2)
We define the Hecke operator at on by
for any homomorphism and simplex . The operator induces operators on and , etc., also written as .
For later use, we also define auxiliary operators and on by
so that .
Lemma 2.4.
The composite acts on , and we have as operators on .
Proof.
Since diagonal matrices act trivially on , we have
for any simplex and polynomial . Since for any integer , we obtain this lemma from this equality. ∎
2.5. Formal duality
Let . As explained in §2.2, the homology and cohomology groups can be computed as
The pairing induces a pairing
which is computed as
Note that for any matrix , we have
Therefore, we have
In particular, we obtain a Hecke-equivariant pairing
which induces an isomorphism
2.6. Eichler–Shimura homomorphism
Let denote the space of modular forms of weight and level . We define a homomorphism
by
for any modular form and . It is well-known that
and the homomorphism induces an injective homomorphism (called Eichler–Shimura homomorphism)
See [Bel21, Section 5.3] for example.
Remark 2.5.
The definition of the Eichler–Shimura homomorphism shows that, for any element and modular form , the pairing can be computed as
Remark 2.6.
For each prime number , the double coset operator acts on the space of modular forms from the right by using the weight slash operator . One can easily show that
for any matrix . Hence the Eichler–Shimura homomorphism is Hecke-equivariant, that is, for all prime numbers , we have
In other words, our Hecke operator 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
Here denotes the space of cusp forms of weight and level .
2.7. Definition of the Eisenstein class
In this subsection, we define the Eisenstein class and explain its basic properties.
We put and let denote the sum-of-positive-divisors function, namely, . Let
denote the normalized holomorphic Eisenstein series of weight .
Lemma 2.8.
-
(1)
For any element , we have
-
(2)
For any prime number , we have
Proof.
Claim (1) follows from the fact that the constant term of is , and claim (2) follows from the fact that and the Eichler–Shimura homomorphism is Hecke-equivariant. ∎
Definition 2.9.
We define the Eisenstein class by
2.8. Main theorem
Proposition 2.11.
The Eisenstein class is rational, that is, .
This proposition is proved in Corollary 4.18.
Definition 2.12.
For any -module , we define
Thanks to Proposition 2.11, we define the denominator of the Eisenstein class with respect to the integral structure by
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 of the Eisenstein class is equal to the numerator of the special value of the Riemann zeta function.
Remark 2.14.
-
(1)
Since , we have another integral structure , and one can consider another denominator of the Eisenstein class :
However, we show in Lemma 6.1 that .
-
(2)
By using the -expansion at the cusp , one can regard as a submodule of , and we obtain the de Rham rational structure of by . The rationality of the critical values of the -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 . 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 and coincide, namely,
since is a regular prime. Cf. [Har21, §1.1].
3. Construction of the cycle
Fix a prime number . In this section, we construct a special homology cycle
that is used to compute the -part of the denominator of the Eisenstein class .
For any integer and element , we set
where recall that .
3.1. Computation of
Recall also the operators
on . We have . For each integer , set
Note that is the identity map. For any (commutative) ring and cycle , we denote by the image of in .
Lemma 3.1.
Let be an integer and .
-
(1)
We have and .
-
(2)
We have
where is the greatest integer less than or equal to and
Here we assume if .
Proof.
Claim (1) follows from the fact that proved in Lemma 2.4. Let us prove claim (2). For notational simplicity, we put
Then claim (1) shows that we can write
with such that
Therefore, we have for any integer , and hence
Let us show that for any integer by induction on . When or , this claim is clear. If and , then the induction hypothesis shows that
Moreover, if is even, i.e., , then we have
∎
By definition, we have
Definition 3.2.
Take elements . For any integers , , and satisfying and , we define
Note that we have
To sum up, we obtain the following corollary.
Corollary 3.3.
We have
3.2. Computation of the boundary
Next, we compute the boundary .
Definition 3.4.
-
(1)
For any integers and with and , we denote by and the integers uniquely determined by
We also put and for any integer .
-
(2)
For any integers and , we set
Note that . We also put .
In the following, for integers and with , we often write as
(3.1) |
for simplicity.
Definition 3.5.
For any integers , , and with and , we define homogeneous polynomials and in by
Lemma 3.6.
We have
in .
Proof.
By definition, we have
The definition of shows that
Moreover, we have
Since
and , we obtain
∎
3.3. A cycle in
In this subsection, we construct a cycle in which is a lift of and is -adically integral for any sufficiently large integer .
3.3.1. Bernoulli polynomials
Since a key tool for constructing the cycle is the Bernoulli polynomials, we briefly recall the basic properties of the Bernoulli polynomials.
Let be a non-negative integer. We denote by the -th Bernoulli polynomial and by
the -th Bernoulli number. For notational simplicity, we put
In this paper, we use the following well-known facts without any notice:
The last fact is called the von Staudt–Clausen theorem. We note that the second and third facts imply that
3.3.2. and
Set
and let
be the -linear map defined by
For any integer , we have , and hence . This fact shows that
Similarly, let
be the -linear map defined by
The following lemma follows from the definitions of and .
Lemma 3.7.
For any polynomial , we have
3.3.3. Definitions of polynomials and
Definition 3.8.
For any integers , , and with and , we define polynomials and in by
Lemma 3.9.
For each , we have .
Proof.
This lemma follows from the facts that and and that the coefficient of in is divided by if . ∎
3.3.4. Definition of the cycle
Now we can define and .
Definition 3.10.
Let . For any integers , , and with and , we define an element by
We also put
Note that by Lemma 3.9.
Lemma 3.11.
We have
In particular, defines a homology class
Lemma 3.12.
The homology class is independent of the choices of and .
Proof.
Let and be another pair of points in . We will prove that
It suffices to construct an element such that
For notational simplicity, set . First, since is simply connected, there exist elements such that
Then we see that
satisfies the desired property. Indeed, we have
where
Then the same computation as in the proof of Lemma 3.6 shows that
where means that it is an equality in the -coinvariant . Hence we obtain . ∎
Definition 3.13.
For any integer and element , we define a cycle by
Definition 3.14.
For any integer , we define a homology class
to be the element represented by the cycle
where is a point such that
Lemma 3.15.
Let be an integer.
-
(1)
in .
-
(2)
If , then we have . Hence defines a -integral homology class in which is independent of the choice of .
-
(3)
If , then the image of the homology class under the homomorphism is .
Proof.
Claim (1) follows from Lemma 3.11. Let us show claim (2). By Lemma 3.9, we have . Moreover, since and , we have
for any non-negative integer . This fact shows that if . It follows from Lemma 3.12 that the homology class does not depend on the choice of . Claim (3) follows from the definition of and Lemma 3.1.
∎
4. Period
The aim of this section is to compute the value
and its -adic limit as . In this paper, we often consider the -adic limit, and hence the symbol will always mean the -adic limit. The following is the main result of this section.
Theorem 4.1.
For any integer , we have
Lemma 4.2.
Suppose that the -adic limit
exists. We then have
Proof.
For notational simplicity, we put
We then have
Take a positive integer . Then there is a positive integer such that for any integer . Hence we have
The sequence is bounded in , and hence for any sufficiently large integer , we have
This implies that
Since (note that ), we have
Since
we obtain that
which completes the proof. ∎
4.1.
We start with computing the value . In this subsection, we fix integers and with . Recall that
are taken as in Definition 3.4 and (3.1) and that
Hence we have
(4.2) |
The definition of the Eisenstein series shows that
where means that there is a constant which does not depend on and such that . Set
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 converges for , and continued to a meromorphic function on . Moreover, it has at most simple poles at and . In particular, is holomorphic at for any integer .
Proposition 4.4.
We have
Note that since by Lemma 3.11, the value does not depend on the choices of and . Therefore, in the following we take and for .
4.1.1. Computation of the first term of (4.2)
Here we compute the first term of (4.2):
This integral is transformed as follows:
Set
Then we have
Now, we see that
-
•
the first terms of and converge for all ,
-
•
the second term of is meromorphic and has at most simple pole at .
In addition, we also see that
In particular, all of these functions are meromorphically continued to and are holomorphic at . This proves Lemma 4.3, and moreover, we get the following.
Lemma 4.5.
4.1.2. Computation of the second and the third terms of (4.2)
4.1.3. Proof of Proposition 4.4
4.2. Summation over
In this subsection, we compute the sum
We keep using the abbreviation
that are actually depending on and . Recall that .
Lemma 4.6.
-
(1)
We have
-
(2)
We have
-
(3)
We have
Proof.
Recall that
Hence we have
For notational simplicity, we put
We then have
Hence the well-known relation that implies that
By setting and using the functional equation of the Riemann zeta function (see [Hid93, p.29] for example), we find
Lemma 4.7.
-
(1)
We have
-
(2)
We have
Proof.
Note that . Since , we have
Hence we have
Since , we have
First note that in the case where , since , we find
Then by using the same argument as in the proof of Lemma 4.6(3), we also obtain
This completes the proof. ∎
4.3. Summation over and the -adic limits
In this subsection, we compute the value
and its -adic limit as . This enables us to complete the proof of Theorem 4.1.
We keep the notation in the previous sections. Proposition 4.4 shows that
and hence
We set
so that we have .
Lemma 4.8.
We have
Lemma 4.9.
Let and be integers with .
-
(1)
For any positive integer , we have
-
(2)
For any integer and any with , we have
Proof.
We have
Hence claim (2) is clear. Moreover if , then all the terms with vanish as , and therefore,
∎
Lemma 4.10.
We have
Proof.
Lemma 4.11.
Let , , , and be integers with and .
-
(1)
If and , then
-
(2)
For any with , we have
Proof.
Lemma 4.12.
We have
Proof.
Proposition 4.13.
We have
4.4. Rationality of the Eisenstein class
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 , we have
Proof.
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 is generated by the set .
Proof.
Note that the relative homology group can be computed as
Let be a class represented by a -chain
where and . The condition that
implies that
(4.3) |
in for some , and . Then we can rewrite the identity (4.3) as
in . Since , this shows that for any , we have
(4.4) |
Now, let denote the point defined by
Then the identity (4.4) implies that in , we have
(4.5) |
Using the identity (4.5), in , we compute
Moreover, (as we are considering the relative homology classes) we may replace with any point in the same connected component of as , and in particular, we may replace by for some . Thus we conclude that the class is represented by a -chain of the form
for some and . Now the lemma follows from the facts that the group is generated by matrices and and that for any and . ∎
Recall that is the subgroup of generated by . Since , the boundary can be identified with . Hence we obtain the following lemma.
Lemma 4.16.
We have an identification , and hence is a -dimensional -vector space generated by , where .
Lemma 4.17.
The kernel of the boundary homomorphism
is generated by the set .
Proof.
Corollary 4.18.
We have .
5. Denominator of an ordinary cohomology class
In order to study the denominator of the Eisenstein class , in this section, we interpret the denominator in terms of the values of the pairing between the Eisenstein class and the cycles .
5.1. Definition of the ordinary part
Let be a prime number and a finitely generated -module with an endomorphism . In this subsection, we introduce the notion of the -ordinary part of .
Since is a finitely generated -module, the -adic limit
always exists, and . We define the -ordinary part of by
and we say that is (-)ordinary if , that is, . We also put . We then have .
The following lemma follows from the fact that ,
Lemma 5.1.
The functor is exact.
5.2. Denominator of a cohomology class
Recall
Definition 5.2.
For any cohomology class , we define the denominator of by
and for each prime number , we set
Lemma 5.3.
Let be a cohomology class. We have
Moreover, for any prime number , we have
Proof.
5.3. Denominator of an ordinary cohomology class
In this subsection, we fix a prime number .
Definition 5.4.
-
(1)
We put and .
-
(2)
For any finitely generated -algebra, we put
-
(3)
For any finitely generated -algebra , we put
Lemma 5.5.
Let be a cohomology class. For any homology class , we have
In particular, if is ordinary, that is, , then
and hence
Proof.
This lemma follows from the facts that the pairing is continuous and . ∎
Recall the identification by Lemma 4.16.
Lemma 5.6.
The -module is free of rank and is generated by .
Proof.
Let be an integer. Since , we have
In particular, we have , and hence . Therefore, inductively, we obtain that for any integer and that , which implies that the -module is generated by . Hence by Lemma 4.16, is a free -module of rank . ∎
Recall .
Proposition 5.7.
For any integer , the -module is generated by (the image of) and a set of lifts of .
Proof.
By definition, we have an exact sequence of Hecke modules
By Lemma 4.15, the ordinary part of the relative homology group is generated by the set . 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 is generated by the set . Since the homomorphism
is an isomorphism and the boundary map is Hecke equivariant, it follows that the kernel of the boundary map is generated by the set . This fact together with the above exact sequence implies this proposition. ∎
Corollary 5.8.
Proof.
Corollary 5.9.
For any integer , we have
6. Relation between the denominators of the Eisenstein classes
Recall that denotes the -part of the denominator of the Eisenstein class (see Definition 5.2). In this section, we fix a prime number and discuss another expression for the denominator of the Eisenstein class . Moreover, we study a relation of the denominators and of the Eisenstein classes when and are -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 in the case that is torsion-free. However, in the present paper we consider the group which has torsion elements other than . 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 , any short exact sequence of -modules induces a long exact sequence in cohomology.
Lemma 6.1.
The inclusion map induces an isomorphism .
Proof.
It suffices to prove that for any integer . Since , any element in can be represented by a polynomial of the form , where . Hence the fact that
shows that for any element . In particular, we have . ∎
Thanks to Lemma 6.1, in the following, we focus on the ordinary cohomology groups with coefficient .
Lemma 6.2.
For any polynomial , we have .
Proof.
By definition, we have
Hence we see that , and we obtain
∎
The boundary is of real dimension , and hence vanishes for any -module . Therefore, for any integer , the short exact sequence induces an isomorphism
(6.1) |
Lemma 6.3.
The ordinary part is torsion-free.
Proof.
By using the exact sequence of -modules
we obtain an isomorphism of Hecke modules
where for an abelian group we write for the subgroup of -torsion elements of . A direct computation shows that . Hence Lemma 6.2 implies that . ∎
Corollary 6.4.
-
(1)
The ordinary part is a free -module of rank .
-
(2)
We have a canonical isomorphism .
-
(3)
We have for any element and prime number .
Proof.
Proposition 6.5.
The ordinary part is torsion-free.
Proof.
Lemma 6.6.
We have .
Proof.
Since the boundary is homeomorphic to the circle, we have . Hence the canonical homomorphism is surjective. Therefore, we only need to show that .
Moreover, since is a two-dimensional real manifold, we have , and hence the short exact sequence induces an isomorphism
Therefore, it suffices to prove that .
For notational simplicity, set . Let
be a representative of a fundamental class of . Then it is known that the homomorphism induces an isomorphism
(6.2) |
See [Shi71, Proposition 8.2] or [Hid93, Proposition 1, §6.1] for example.
We will show that for any , we have . By (6.2), it suffices to show that . Here we use to emphasize that it is an identity in . We then compute
Put
Then we find that
∎
For any -module , we define the inner cohomology by
and, when is a finitely generated -module, we put
Then the following corollary follows from Lemma 6.6 and the isomorphism (6.1).
Corollary 6.7.
Let be a non-negative integer and . Then we have a natural exact sequence of Hecke modules:
Proof.
Corollary 6.8.
For any integer , the canonical homomorphism induces isomorphisms
Proof.
The exact sequence
shows that we have an exact sequence
Hence by Lemma 6.6, we obtain the first isomorphism.
Theorem 6.9.
For any positive integers and with , we have the following canonical isomorphism of exact sequences which is -equivalent for any prime number :
where .
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 , we have canonical isomorphisms between a sheaf cohomology on and a group cohomology of :
(6.3) |
Moreover, the inner cohomology group corresponds to the parabolic subgroup of under the isomorphism (6.3) (see [Hid86, (4.1a)] for the definition of the parabolic subgroup).
Let . Hida showed in [Hid86, Proposition 4.7] that we have isomorphisms
(6.4) |
which are -equivariant for any prime number . Here denotes the restriction map.
Let denote the -module whose underlying abelian group is and the -action is given by the homomorphism . Then Hida also showed in [Hid86, Corollary 4.5 and (6.8)] that the -homomorphism induces Hecke-equivariant isomorphisms
(6.5) |
Since , we have as -modules, by combining the isomorphisms (6.4) and (6.5) for , we obtain the following commutative diagram:
where horizontal arrows are isomorphisms and -equivariant for any prime number . This completes the proof. ∎
6.2. Another expression for
Let be a prime number and take a prime number . Let
be the polynomial ring over , and by using the Hecke operator at , we regard cohomology groups that appear in the present paper as -modules. For notational simplicity, we put
Note that by Corollary 6.4, we have as -modules.
Lemma 6.10.
Let be a cohomology class. If , then we have for any prime number , that is, the cohomology class is a scalar multiple of .
Proof.
It is well-known that one can take a -Hecke-eigen basis such that and that the elements 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 -eigenvalue of () is less than , which implies this lemma. ∎
Lemma 6.11.
We have .
Proof.
Definition 6.12.
We define to be the element corresponding to the exact sequence of -modules in Corollary 6.7 for :
The following lemma follows directly from Lemma 6.11.
Lemma 6.13.
.
Lemma 6.14.
We have a natural identification
Proof.
Since is an regular element of , we have an exact sequence of -modules:
Applying the functor to this short exact sequence, we obtain the desired identification. ∎
Lemma 6.15.
For any positive integers and with , we have a natural isomorphism of -modules:
Moreover, the image of is under this isomorphism (and the identification in Lemma 6.14).
Definition 6.16.
We define a polynomial to be the characteristic polynomial associated with :
Lemma 6.17.
The -module is annihilated by .
Proof.
By the Cayley–Hamilton theorem, the Hecke module is annihilated by . Hence the -module is annihilated by since . ∎
Lemma 6.18.
Let be a positive integer satisfying . Then for any even integer with , we have
Proof.
By Theorem 6.9, we have
The fact that implies that , and we obtain . Hence the assumption that shows that . ∎
Proposition 6.19.
Let be a positive integer satisfying . Then for any even integer with , we have
7. Kubota–Leopoldt -adic -function
Let be a prime number. In this section, we introduce the Kubota–Leopoldt -adic -functions and prove certain congruence properties that will be used in the proof of Theorem 2.13.
Let denote the Teichmüller character, and let denote the trivial character. For any Dirichlet character , we denote by the Kubota–Leopoldt -adic -function attached to .
Proposition 7.1 ([Was97, Theorems 5.11 and 5.12, Exercises 5.11(1)] ).
-
(1)
For any Dirichlet character , the -adic -function converges on . Moreover, for any integer , we have
In particular, we have .
-
(2)
We have
-
(3)
If , then we have
By using the Kubota–Leopoldt -adic -functions, Theorem 4.1 can be restated as follows.
Corollary 7.2.
If we put
then for any integer we have
For any even integer , we define a positive integer by
Corollary 7.3.
Let be an even integer.
-
(1)
If , then we have .
-
(2)
If , then we have .
-
(3)
Let and be positive integers with . If , then
Proof.
Corollary 7.4.
Let be an integer with . For any integer , we have
Proof.
Corollary 7.5.
For any integers and , we have
Proof.
For notational simplicity, we put
By Proposition 7.1(2), we have , , and . Since , we have
Put . Then
and we have
∎
8. Proof of Theorem 2.13
Proposition 8.1.
We have
Proof.
Recall that denotes the numerator of .
Proposition 8.2.
Let be a prime number.
-
(1)
.
-
(2)
If , then .
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 .
Proof of Theorem 2.13.
Take a prime number . It suffices to show that . When , by Proposition 8.2 we have , and hence we may assume that . Note that in this case. Take a prime number , and positive integers and satisfying
-
•
,
-
•
,
-
•
.
Then by Propositions 8.2(2), we have , and Proposition 6.19 implies that
Hence Corollary 7.3 shows that . ∎
8.1. Proof of Proposition 8.2
In this subsection, we prove Proposition 8.2. The proof is divided into the following two cases:
-
•
,
-
•
.
8.1.1.
Lemma 8.3.
If , then we have .
Proof.
Moreover, if , the result of Carlitz concerning the index of irregularity of a prime shows the following lemma.
Lemma 8.4.
If and , there is an (odd) integer such that . In particular, we have in this case.
Proof.
By Lemma 8.3, for any regular prime and (odd) integer , we have . Therefore, we may assume that is an irregular prime. In particular, .
We define the index of irregularity of the prime number by
Then by using the result of Carlitz in [Car61, (21)], Skula proved in [Sku80, Theorem 2.2, Remark 2.3] that
Hence if , then we have
Moreover, since the only irregular prime smaller than is and , the inequality holds true.
For any integer , we define an integer by
Since , there is an even integer with such that
Furthermore, Proposition 7.1(3) shows that
and hence we have . Therefore, we put and get
∎
8.1.2.
Lemma 8.5.
If , we have .
Proof.
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 (resp. ) rather than (resp. ) since they are naturally isomorphic.
First, note that we have the following corollary of Theorem 2.13.
Corollary 9.1.
Let be an even integer and a matrix. Take a polynomial such that . Then for any element , we have
Here is the numerator of .
Proof.
Since , we have , and hence defines an element in the homology group . Therefore, by Theorem 2.13, we obtain
∎
9.1. Duke’s conjecture
In the paper [Duk23], Duke defined a certain map called the higher Rademacher symbol
for each integer which is a generalization of the classical Rademacher symbol and gave a conjecture concerning the integrality of the higher Rademacher symbol .
Conjecture 9.2 ([Duk23, Conjecture, p. 4]).
For any integer and matrix , we have
Remark 9.3.
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 be an integer. For any matrix , we define a binary quadratic polynomial associated with by
Then for any element , we have
where is the numerator of .
Corollary 9.5.
Duke’s Conjecture 9.2 holds true.
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 be a real quadratic field, and let be an order of with discriminant . We denote by the group of proper fractional -ideals and the subgroup of totally positive principal ideals. We define the narrow ideal class group of by
See [Cox13, §7]. We fix an embedding , and for any element , we denote by its conjugate over .
Moreover, let denote the group of totally positive units in , and let denote the generator of such that .
Definition 9.6.
We define a map
as follows: Let , and take a representative of . We also take a basis over such that , and let be a matrix such that
Moreover, set
We see that and that . We then define
where is an arbitrary element in .
Lemma 9.7.
The homology class does not depend on the choices we made.
Proof.
The independence of is clear. Let be another representative. Then there exists a totally positive element such that . Take a basis over with . Then we obtain a matrix and a binary quadratic polynomial from the basis of . Note that since is totally positive, we have . Let be a matrix satisfying
Then the facts that and imply that . Since , we have . Moreover, we have
which implies that . Therefore, we have
as elements of . ∎
Remark 9.8.
- (1)
-
(2)
Gauss’s theory concerning binary quadratic forms (see [Cox13, Exercise 7.21] for example) shows that for any hyperbolic element , there is an order of a real quadratic field and a narrow ideal class such that
Definition 9.9.
For each ideal class , the partial zeta function associated with is defined by
and it is well-known that can be continued meromorphically to and has a simple pole at .
The following integral representation of the special values of the partial zeta function is classically known.
Proposition 9.10.
For any integer and ideal class , we have
Before we give a proof of Proposition 9.10, we recall (a special case of) the so-called Feynman parametrization.
Lemma 9.11.
Let be complex numbers such that and . Then for any non-negative integers and , we have
Proof.
We may assume that . By setting and , it suffices to prove that
The case where is clear, i.e., we have
Then by viewing the both sides as holomorphic functions in and applying the differential operator , we obtain the desired identity. ∎
Proof of Proposition 9.10.
We use the same notations as in Definition 9.6. Since
and , for any element we have
Since , where is the factor of automorphy, if we fix a complete set of representatives of , then we have
Set . Then the point (resp. ) is the attractive fixed point (resp. repelling fixed point) of the hyperbolic matrix , i.e., we have and in for any element . Hence we obtain
By using Lemma 9.11, we find
Note that we have the identity , and this shows that
For any subset , we put
Let denote the ideal class containing the principal ideal . Note that in . Then we further compute
Now, we recall the functional equations of the partial zeta functions. Set
Then we have
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
Therefore, by also using the functional equation for the Riemann zeta function (see, for example, [Hid93, p. 29]), we obtain
∎
We define the positive integer by
Corollary 9.12.
Let be a real quadratic field, be an order in , and let be a narrow ideal class of . Then for any integer , we have
Remark 9.13.
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 ) universal upper bound was obtained by Serre in the paper [Ser73, §2, Théorème 6]. Moreover, if we fix a totally real field , then a more refined description for the denominators and even for the numerators of the special values of the Dedekind zeta function of 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 be an integer. We define a -submodule to be the -submodule generated by homology classes of the form , that is,
This subsection is devoted to proving the following theorem.
Theorem 9.15.
We have .
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 , there exist an order of a real quadratic field and a narrow ideal class such that
In other words, we have
Proof.
9.3.1. Preparations for proving Theorem 9.15
Let be an integer and define
We also put
We note that the similar facts in §2.1 and §2.2 hold true for the congruence subgroup . Moreover, the Hecke operators () and () act on the homology group .
Let be an integer. For any hyperbolic matrix (i.e., ), we set
Definition 9.17.
For any integer , we define a -submodule
by
Remark 9.18.
By Remark 9.8, we have .
Lemma 9.19.
For any integer and prime number , we have
Proof.
It suffices to show that
Here means the group generated by the elements inside the bracket. Moreover, since , we may assume that . Put . Then the quotient group is generated by the image of . For any matrix , we have for any integer . Since and , the matrix is hyperbolic, and one can take a matrix such that . If we put , then we have . Since and , we have . Hence the set is infinite. Therefore we can find an integer such that . ∎
Next, we recall an important result proved by Hida (see [Hid86, Corollary 4.5] or [Hid88, Corollary 8.2]). Take an integer and a prime number , and put . Then we have a -homomorphism
which induces a Hecke-equivariant homomorphism
Proposition 9.20.
When , the homomorphism induces a Hecke-equivariant isomorphism
Here we define .
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 is torsion-free since . Hence any short exact sequence of -modules induces a long exact sequence in homology.
If we put
then the short exact sequence induces an exact sequence of -modules
Since the operator is defined by , we have
for any polynomial . In other words, we have
and this shows that , which completes the proof. ∎
Lemma 9.21.
Let be an integer and let be a prime number. Set . Then for any matrix , we have
Proof.
If we put , then we have
Since , we have and , which shows that
and
∎
Theorem 9.22.
For any integer and prime number satisfying , we have
9.3.2. Proof of Theorem 9.15
Let be an integer. For any positive integers and with , we denote by
the homomorphisms induced by the natural projection .
Corollary 9.23.
For any integer and prime number , we have
Proof.
Take an element . Since , we have
On the other hand, since
we have
Therefore, first when , Theorem 9.22 implies this lemma. When , we have and . Then this case follows from the case and since . ∎
Lemma 9.24.
Let be an integer and let be a prime number. Then for any homology class we have
Proof.
By using the formal duality, it suffices to show that
for any cohomology class . This claim is well-known (see [Gou92, Lemma 2] for example). ∎
Corollary 9.25.
For any prime number , we have
Proof.
Take a prime number . Then Theorem 9.22 shows that
By Lemma 9.24, we have a natural homomorphism
which is the dual of the homomorphism
The fact that the index is relatively prime to together with the isomorphism (6.4) implies that is surjective. Hence we have
Therefore, this corollary follows from Theorem 2.13 and Lemma 5.5. ∎
We note that for any prime number , the operators and induce homomorphisms
Lemma 9.26.
For any prime number and integer , we have
Proof.
We put
Since and , the relation that shows that
Since , these facts imply this lemma. ∎
Lemma 9.27.
For any integer and a prime number , we have
Proof.
Lemma 9.26 implies that
Hence we have
Therefore, it suffices to show that . Let be a matrix. Then we have . Moreover, by the definitions of and , we obtain
In particular, we have . ∎
Proof of Theorem 9.15.
References
- [BCG20] Nicolas Bergeron, Pierre Charollois, and Luis E. Garcia, Transgressions of the Euler class and Eisenstein cohomology of , Jpn. J. Math. 15 (2020), no. 2, 311–379. MR 4120422
- [Bel21] Joël Bellaïche, The eigenbook—eigenvarieties, families of Galois representations, -adic -functions, Pathways in Mathematics, Birkhäuser/Springer, Cham, [2021] ©2021. MR 4306639
- [Ber08] Tobias Berger, Denominators of Eisenstein cohomology classes for 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, -adic polylogarithms and -adic Hecke -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 -adic interpolation of -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 , 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 -adiques, Invent. Math. 51 (1979), no. 1, 29–59. MR 524276
- [Cox13] David A. Cox, Primes of the form , 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 -groups, Invent. Math. 24 (1974), 149–161. MR 369322
- [CS74b] by same author, On -adic -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 -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 -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 -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 attached to ordinary cusp forms, Invent. Math. 85 (1986), no. 3, 545–613. MR 848685
- [Hid88] by same author, On -adic Hecke algebras for over totally real fields, Ann. of Math. (2) 128 (1988), no. 2, 295–384. MR 960949
- [Hid93] by same author, Elementary theory of -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 -Erweiterungen von 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, -theory and arithmetic, Contemporary developments in algebraic -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 -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 -adic analytic -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 and values of -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 und die Arithmetik spezieller Werte von -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 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