1. Introduction and the Underlying Flow of Ideas
In this paper, by a zeta-function, we mean a certain Dirichlet series that is absolutely convergent in a right half-plane, continued meromorphically over the whole plane (with a possible simple pole at
, say), and satisfying a certain functional equation with gamma factors. We write
,
. Given a zeta-function
analytic at
, it is common to regard the value
as the Kronecker limit formula, i.e., the Laurent constant at the simple pole at
, say. This depends on a heuristic reason that
, whence formally
Cf. [
1], pp. 94–131, Chapter III.
Another reason comes from Stark’s intensive work [
2,
3,
4,
5] on ”
L-functions at
” in which he made clear that instead of the original Kronecker limit formula at
, the one at
is simpler (in view of a possible functional equation). In Part IV, he gave the Kronecker limit formula.
In [
6], (§2.4), we developed the Chowla–Selberg integral formula (or the Fourier–Bessel expansion, which is equivalent to the Hecke functional equation) and the Kronecker limit formula as its consequence for
. Here,
is the
Epstein-type Eisenstein series defined by [
6], (1.1.7):
where
and the prime on the summation sign means that
is excluded. Here,
is the upper half-plane
, and the far-right side of Equation (
1) is the Epstein zeta-function of a positive definite quadratic form. In what follows, we always write
(
).
Ref. [
1], p. 96, Example 7.3, gives
where
is the Dedekind eta-function, defined as follows:
Here Equation (
2) is the case of the discriminant
, the prime on the product sign has the same meaning as with the sum Equation (
1), and the zeta-function is not [
6], (1.1.7), but the non-holomorphic Eisenstein series
This is a typical example of a non-holomorphic automorphic function for which we may employ the results known of the associated Epstein zeta-function.
In this paper, we shall make use of the generating zeta-function for Equation (
4) stated in Lemma 1. For general basics on zeta-functions, we refer to [
6,
7]. As will be seen subsequently, the functional equation Equation (
5) is rather involved, save for the case
being an odd integer, and so directly proving modular relations is not simple. We shall examine the papers of [
8,
9,
10,
11] from the view point of lattice zeta-functions as an alternative. In Lewittes [
8], the Bochner modular relation is deduced by dissection (applied to the whole sum for
) and the Lipschitz summation formula (which itself is a modular relation, applied to the component of
), in Komori et al. [
9], the ramified Hecke function equation is proved, and in Kurokawa [
10], the Kronecker limit formula without absolute value is deduced as the value at
, as per Stark (with implicit influence of the functional equation).
Lemma 1. Consider the product of two Riemann zeta-functions,where the series is absolutely convergent for andis the sum-of-divisors function. This satisfies the asymmetric functional equation In the case of α being an odd integer, it reduces to the Hecke functional equation Equation (6) and the Bochner modular relation in symmetric formare equivalent, where for , the residual function is given byand for odd, Equation (7) amounts toor In the literature, the zeta-values are often expressed by the
Bernoulli number, cf. Equation (
12) below. E.g., on [
12], pp. 97–98, we find
We remark that the common source of Equations (
5) and (
10) is the asymmetric functional equation for the Riemann zeta-function:
The
nth Bernoulli polynomial
is defined by the Taylor expansion
and the
nth Bernoulli number
is the value
. We agree to use both expressions in Equation (
10) interchangeably hereafter.
The transition between the Lambert series and the rapidly convergent series is carried out using
Liouville’s formula
where
with
d running through all positive divisors of
n, and the left-side expression is called the
Lambert series. It seems that this was first explicitly stated by Koshlyakov [
13,
14,
15], I, p. 138, (10.15).
Equation (
7) reads (
)
which is Ramanujan’s formula in the Bochner modular relation form, where
cf. Equation (
8). Through the change of variable
the original Ramanujan’s formula (Equation (
24)) and its Bochner modular relation form (Equation (
14)) transform into each other, which we describe as part of the following theorem.
Theorem 1. The Bochner modular relation (Equation (7)), which holds for and every odd integer α, entails at one end of the spectrum , Ramanujan’s formula (Equation (14)), and at the other end, , the automorphy of the general Eisenstein series in Equation (17), where is an even integer:At the threshold point , it entails automorphy (Equation (32)) of the modular form of the half-integral weight. Definition 1. We denote the Fourier series as follows:which is continued analytically over the whole plane. We define the generalized Eisenstein series as follows: Summarizing [
16], pp. 82–83 or [
17], pp. 117–137, Chapter 8, we have the following:
Proposition 1. Each -lattice may be viewed as with . Each -lattice up to homothety may be identified with an element of , where . We may identify modular functions of weight with some lattice functions of weight :It follows that the automorphic property is intrinsic to lattice functions (and its equivalent, with the ratio τ of the bases) as well as modular functions. Equation (
1) is a typical example of the lattice function. Another is the classical Eisenstein series [
16], p. 83, (12), given as follows:
Lemma 2. The Eisenstein series is defined bywhich is a modular form of weight . The Laurent expansion (or q-expansion) reads as follows:Also, is the constant multiple of this: Here, Equation (
19) follows from Equation (
11).
We shall show that plenty of the subsequent derived results are inheritance of the
q-expansion (Equation (
18)) (rather than the Bochner modular relation). By equating the partial fraction expansion for the cotangent function and the polylogarithm function of order 0, we obtain the following (cf. Equation (
53)):
By differentiating this
times, we deduce that
where
is defined in Equation (
40). The
q-expansion (Equation (
18)) follows by substituting Equation (
21) with
replaced by
in the following:
Definition 2. Lewittes [8] introduced the generalized Eisenstein series,absolutely convergent for , where the summation sign indicates that range over all integers except . Combining Equations (
41) and (
43), we deduce the following:
This indicates that
accommodates both
(as a special case) and
(as an accommodated element) and suggests that one can derive many results on those Eisenstein series as consequences.
2. Main Results
In view of the raison-d’être for automorphy in Proposition 1, we understand that lattice functions are the other main ingredients. From this point of view, we are to examine the papers of Chapman [
18], Komori et al. [
9], Kurokawa [
10], and Lewittes [
8] which are written in complete independence of each other but are all concerned with lattice functions and use some dissection methods. Ref. [
18] is concerned with the pseudo-automorphy of
for odd integer
k in Theorem 2. Save for the Chapman Dedekind sum part, others can be accommodated in [
8], and we dwell on those three papers. Lewittes uses the Barnes double zeta-function implicitly (proving the results on the Barnes double zeta-function anew) and gives a generalization of the Bochner modular relation (Theorem 2), while Komori et al. use the Barnes multiple zeta-function and prove the original version (Equation (
24)). It is stated for
satisfying the relation
and for any positive integer
k, as
As is stated in
Section 1, the key point is to extend this as in Lemma 1.
It seems that Komori et al. do not use a dissection, but in their proof of the functional equation, the Atkinson dissection was used. Since the Barnes multiple zeta-functions are semi-lattice functions, they may be accommodated in this category. The Barnes–Hurwitz zeta-function (or multiple zeta-function) is defined in Equation (
25) for a basis
and the associated semi-lattice:
suppose
lie in the same side of a line going through the origin. Then, it is defined as follows:
absolutely convergent for
. It is continued meromorphically over the whole plane and satisfies the functional equation of Hecke type, cf.
Section 4. The common procedure is to reduce the lattice zeta-functions to Barnes zeta-functions and apply the known results. This seems to have been done effectively by Hardy and Littlewood and much later by Shintani. The shift in
is rather delicate. For Lewittes’ case, it is 0, but then in the final regularized product results, the required case
is not included. Komori et al. incorporated the parameter
, (Equation (
61)) in Equation (
25). This is convenient to make the case of
resp.
accommodated, which presents the case of summation variables being
resp.
, which in turn amounts to the case of ordinary zeta-functions and the hyperbolic sine case. But again, the required case
is not included. And, it is included only in Kurokawa’s zeta-function. This suggests the following:
We are in a position to state the main results (including those from references [
8,
9,
10,
18]), in somewhat revised and organized form.
Theorem 2 (Ramanujan formula as per Lewittes)
. The general Ramanujan formula holds true for all :where is the Fourier series (Equation (16)) and is the Barnes double zeta-function in Equation (45). The special case of being an odd integer of Equation (26) is equivalent to the functional equation given in Equation (6) as the Bochner modular relation, where is defined by Equation (3). Both Equation (
47) and Equation (
26) hold for all
s values, and each term is analytically continued over the whole plane. Therefore, it follows that
also has an analytic continuation over the whole plane.
Theorem 2 implies the following:
Corollary 1. which entailsfor every . Proof. For a proof of Equation (
28), we note the following. The series for
in Equation (
45) converges uniformly for
for each
. Hence, it follows that for
. □
The following corollary provides a further refinement of the elaboration of [
18], Theorem 1, given in [
19].
Corollary 2 ([
18])
. For odd integer and for every ,wherewhich satisfies the transformation formulaFurther, the closed formula holds true ():whereis Chapman’s Dedekind sum. Proof. We denote by
the right-hand side of Equation (
28) with
replaced by
:
Then,
since
based on Equation (
46). Hence,
which entails Equation (
29).
Proof of Equation (
30) amounts to establishing
where
is the Lerch zeta-function. More computation is needed. □
Corollary 3. Lewittes’ generalized Ramanujan formula (Equation (26)) in the case of being an integer amounts towhere on account of Equation (60), we havewhere C is a Hankel contour in Lemma 3. The special case of Equation (31) with leads to the transformation formula for the Dedekind eta-function. For Equation (
32), cf. e.g., [
20,
21].
In view of Equation (
45), Equation (
31) is a restatement of Equation (
26). Then, Lewittes evaluates
directly, but we appeal to Equation (
63), which gives the evaluation for
(
), and so Corollary 3 leads to Ramanujan’s formula Equation (
79). The negative even integer case is also included as stated in Equation (
9).
Now, we turn to Kronecker limit formulas without absolute value. We use the
q-notation:
Theorem 3. We have the limit formulas.
(i)
([22], Proposition 2.3) For the function in Equation (40), we have the following:(ii)
We havewhereis the arithmetic function studied by Erdös and Zaremba [23] andor As a corollary to Theorem 7, we have the zeta-regularization formula corresponding to Equation (
2):
Theorem 4 ([
24])
. Let , . Then, the path integral (with imaginary time) has the expression as the product of determinants (if they exist):We have the following explicit formula: Proof. The eigenvalues
of
resp.
are
resp.
, where
. Then, consider the following zeta-function:
where
is defined by Equation (
71), and we interpret the determinant of the operators
as the zeta-regularized determinant:
Hence, Theorem 7 is applicable to give the following:
and similarly for
. Substituting these into Equation (
37), we deduce Equation (
38). □
As a consequence of Theorem 6 and results from the Barnes double zeta-function, we have the following:
Theorem 5. The following functional equation holds:This may be viewed as a generalization of Ramanujan’s formula (Equation (79)), which is the case of with . For the derivative, we have the following:which leads to a variant of Ramanujan’s formula and the limit formula. 3. Lewittes’ Modular Relation for the Generalized Eisenstein Series
We provide proofs of some of the results stated in
Section 2 by slightly modifying the results of [
8,
18]. It turns out that most of Chapman’s results are consequences of Lewittes’. We mainly state Lewittes’ results and add remarks in square brackets on the corresponding results of Chapman’s.
Proof of Theorem 2. Proof of equivalence follows from Lemma 1, and Equation (
26) follows from Equation (
48) in view of
. Hence, it suffices to prove Equation (
48).
Lewittes [
8] restricts the argument not as in Equation (
72) but as follows:
The zeta-function is defined as follows:
which is absolutely and uniformly convergent for
(
).
Lewittes extracts the cases of from and and divides into three cases: , , and . Similarly, the remaining two cases, and , are unified as .
This proves Lewttes’ formula:
corresponding to Equation (
22), where
is defined by Equation (
23).
Based on the upper half-plane version of the Lipschitz summation formula,
we have
Hence, based on the upper half-plane version of Liouville’s formula, we obtain the following:
[This corresponds to [
18], (3) in the following form:
]
Hence, substituting Equation (
42) into Equation (
41) leads to [
8], (3)
where
is defined by Equation (
16). The series for
is absolutely and uniformly convergent for
in any compact subset of
, hence
, and a fortiori,
, have analytic continuation over the whole
s-plane. However,
is a meromorphic function with a simple at
with residue 1.
[In [
18],
for odd integer
. However, in the original definition given by Kurokawa [
11], there is the correction term
, which happens to vanish for
for an odd integer
. Hence, we are led to define the generalized Eisenstein series by Equation (
17). Cf. Corollary 1.]
If we can find the Laurent constant of Equation (
23) in closed form, then it is the Kronecker limit formula. Here, we stick to the zeta-regularization and prove Equation (
36).
To proceed further, Lewittes introduces another dissection that is almost the same as the Kurokawa dissectionn except for the excludion of two cases of and . Here, e.g., means Lewittes’ dissection no. (1), and similarly for for Kurokawa below.
,
,
,
,
,
,
, and
, where the term in (n’) is
times of (n). We use the zeta-functions
as well as
where the far-right member indicates the Barnes double zeta-function Equation (
25) with
,
.
Note that
gives
,
gives
,
gives
, and
gives
. Substituting these into
values in Equation (
44), we obtain the following:
[[
18], (2),
is
in Equation (
45).]
Therefore, we could apply the method of
Section 4, but we partly reproduce Lewittes’ argement combined with the theory of Barnes zeta-functions. Moreover, since
, we have
, and we have the following transformation formula:
Substituting Equation (
44) in the above and solving in
, we conclude the following:
which is to be proved. □
Remark 1. In [25], Kim introduced “two-sided” L-series, called the H-series (), which corresponds to the Laurent expansion. The following relation was proved:whereFrom Equation (49), which corresponds to Equation (44), he deduced the integral representation of as the Mellin transform of the Weierstrass zeta-function. This will be touched on elsewhere. Proof of Corollary 1. Another proof of Corollary 1 can be given using Equation (
42) and Equation (
17) and transforming the difference
Equation (
50) amounts to the following:
Extracting the special case with
resp.
, which gives
resp.
, we have the following:
This proves Equation (
27).
[In the case of Chapman, is an odd integer, which gives .] □
Finally, the
case again leads to the eta-transformation formula (Equation (
32)) based on the following:
Proof of Theorem 3. A proof is given by Kurokawa, but Lewittes could have proved it based on the following dissection:
Hence,
Substituting the Lerch formula
and
, we deduce that
based on the reciprocity relation. Hence,
which leads to the first formula of Equation (
33). In the case of
, the dissection (Equation (
52)) is changed into the following:
and the argument proceeds similarly.
Recall Equation (
44) as
We observe the ease of computation at
. For as long as
survives, the terms are 0. Hence, the following is needed:
because of
which is part of Equation (
20) and
. Hence, the first equality of Equation (
35) follows, and the second equality follows from Equation (
51). □
Note that
for
is the
kth derivative of the cotangent function. Also note that Equation (
36) is very plausible since it contains the Dedekind eta-function and is an analogue of Equation (
33). One can see how it will be messy to work at
, in which case we need to evaluate many terms, and use is made of Equation (
34). Then connecting two results (based on the functional equation) would produce some interesting results.
4. Generalized Ramanujan Formula as per Komori et al. as the Modular Relation for the Barnes Multiple Zeta-Function
We assemble results on the Barnes multiple and double zeta-functions, which are basic in Lewittes’ and Kurokawa’s results, and we derive Theorem 5 from Theorem 6 using the data on the Barnes double zeta-function.
Suppose at least three of are linearly independent and that all s and lie on the same side of some straight line L going through the origin. Let denote the semi-lattice consisting of linear forms with .
Then, the Barnes multiple zeta-function (or
r-ple Hurwitz zeta-function)
of the complex variable
s with parameter
and basis
is defined as the Dirichlet series (Equation (
25)). Cf. [
26,
27].
From [
26], we have the following:
Lemma 3. where the contour C can be taken as the one from to λ along the real axis, going along the circle around 0 of radius λ counterclockwise to λ, and then going back to , where . This expression gives the analytic continuation of to the whole complex plane with simple poles at ; in particular, is holomorphic at . Also, the integrals along the real axis cancel each other for for , so that the value is given by the residue of the integrand at , which is as follows:whereunder the convention that the jth power of is . We define the Stirling modular form
by
and the
r-ple gamma function
by
We state the special case of Lemma 3.
Lemma 4. We have the formulafor , , and the formula for the double zeta-function: (54) amounts toor equivalently, Proof. We have
upon summing the geometric series. Hence, multiplying the integrand by
, we deduce Equation (
58).
In the case of the double zeta-function, we have the following:
We obtain Equation (
59) in the same way.
It turns out that the only difference is whether one starts summing the geometric series from 1 or 0, which gives the difference to the effect that there appears the additional factor (in the general case ) or not. After this, it is a well-known Hankel contour method with the specified contour
, cf. e.g., [
28], §12.22. □
Let
be the open half-plane whose normal vector is
([
26], p. 388), and let
We write
.
Theorem 6 ([
9], Theorem 2.1)
. The functional equation holds true.where and the right-hand side converges absolutely and uniformly in the whole plane if and in the region if . As is noted in [
29], p. 3 and [
9], pp. 55–56, the proof of Theorem 6 follows the proof in [
30]. Matsumoto [
29] established a proof of the functional equation for the double zeta-function
for
. In [
31], an asymptotic formula for the special case
is established.
For each
, let
, and similarly, we define
. For
, we define
Theorem 6 for the case of the generalized Eisenstein series ((
25) with
; cf. the special case Equation (
45)), reads as follows:
Corollary 4 ([
9], Corollary 2.2)
. We havewhere the series on the right-hand side converge absolutely and uniformly in and where To find special values, we appeal to Equation (
55). Therefore, we recall the
nth Bernoulli polynomial
and denote its generating function by
.
Lemma 5. As , we haveFrom this, we have for or if In particular, Lemma 6. We have the closed-form evaluation:for an odd integer. This gives the closed-form evaluation of in Equation (45) in view of Equation (45). Proof. We apply Corollary 4 with
,
. Then,
and
. Hence, we have the following
Table 1:
Incorporating these data in the equalities in Corollary 4, we deduce that
We immediately observe that if
is an even integer, Equation (
65) leads to the well-known relation between the zeta-values and Bernoulli numbers, while if
is an odd integer, Equation (
65) amounts to the following:
where the right-hand side is the value at
, being an odd integer including the case
. Indeed, to include the case, we have to consider the limit case as
of the last term
, say, of Equation (
66).
We find that as
where
, so that
Cf. [
17], §69. We understand Equation (
67) to mean the following:
under Equation (
68).
Incorporating Equations (
69) and (
63), we transform Equation (
66) into Equation (
64).
□
Proof of Theorem 5. Proof of Theorem 5 follows from Lemma 6 as follows. The
case of Equation (
79) leads to Equation (
32) as in Corollary 3.
It suffices to derive Equation (
70) from Equation (
62).
Using
we find from Equation (
62) that
The coefficient
may be expressed as
and also as
. Hence, (
39) follows as a special case of Equation (
70). □
6. Appendix: Some Notes on Ramanujan’s Formula as the Modular Relation
In [
33], pp. 275–276, Entry 21 (i), Chapter 14, it is claimed that Entry 21 (i) yields the
case in [
33], p. 261, Entry 13, so that the generalized Ramanujan formula for all integers but 0 is in order upon modifying the residual function
, which is
of Equation (
24). But, this is not convenient, and we stated explicit formulas above for the residual function in the form of the Bochner modular relation (or its upper half-plane verison—the Hecke correspondence).
By changing the variable
we have the correspondence: the right half-plane
↔ the upper half-plane
.
Based on Equation (
77), Equation (
13) becomes the upper half-plane version
where the right-hand side is more often recognized as a Fourier series (or a
q-expansion, or a Laurent expansion, etc.).
Problem: Is it possible to draw some information from the functional equation when is not necessarily an integer?
Even in the case of
being an even integer, (
5) leads to the following:
where the expansion holds for
.
Bellman’s work [
34] applies in place of Equation (
78)
Both are rather involved. Cf. [
35].
After changing into the rapidly convergent series, it is customary to use Equation (
15) and express Equation (
24) as the relation in Equation (
14) between the correspondence
in
.
The
upper half-plane version of Equation (
14) ([
36], (2.4), (2.5)) reads as follows, with
and any integer
:
where
. In this form, it is an example of the
Hecke modular relation and may be referred to as a
q-expansion. In general, for a modular form
f, the Fourier series
is often called the
q-expansion.
Note that we changed the notation
in Equation (
14) to
to include the case of
. This inclusion is important, since the
case implies the eta-transformation formula (Equation (
32)).
Generalization of Equation (
24) (
) is performed in [
37], pp. 429–432, Entry 20 and [
33], pp. 253–254, Entry 8. Let
with
and
and let
be an entire function. Then, with
,
and
C the rectangle with vertices at
, the following lemma corresponds to Entry 20, where
and
m is a positive integer.
Lemma 7. The following generalization of the argument principle holds:if has finitely many poles , with multiplicity resp. finitely many zeros , with multiplicity , and is meromorphic with poles at , (assuming that poles of f and do not coincide for simplicity) in a domain D with boundary C. In view of the global expressions
in Lemma 7 can be chosen to be the cotangent function
, then Entry 20 is a consequence of Lemma 7. In view of the fact that the partial fraction expansion for the cotangent function is equivalent to the functional equation for the Riemann zeta-function and that the cotangent function is essentially the polylogarithm of order 0, it is the most natural way to derive Ramanujan’s formula and the eta transformation formula, cf. [
38,
39,
40].
In the above, we were restricted to the case of the Hecke modular relation or the Hecke correspondence, which is equivalent to the functional equation with the single gamma factor
. More generally, this is called the RHB (Riemann–Hecke–Bochner) correspondence. We shall mention generalities of the Hecke correspondence, cf. [
6,
41].
In relation to the
q-expansion (Equation (
80)), we introduce the modular-type functions corresponding to the Dirichlet series (Equation (
83)):
which are absolutely convergent and satisfy the (modular) transformation formula:
Definition 3. Letbe increasing sequences of real numbers. For complex sequences form the Dirichlet serieswhich we assume are absolutely convergent for and , respectively. Then, and are said to satisfy Hecke’s functional equation (HFE)where is a constant if there exists a regular function outside of a compact set such thatandand such that is convex, in the sense thatuniformly in . We refer to the Dirichet series (Equation (83)) as Hecke L-functions (HLF) which satisfy (HFE) (Equation (84)). Following Bochner [42], the residual function is defined as follows:where C encircles all the singularities of in . Then, Equation (
84) needs to become Equation (
85), being modified by replacing
with
.
It is known that the partial fraction expansion (Equation (
81)) is equivalent to the functional equation for the Riemann zeta-function, and in the long run, it is equivalent to the above modular relations.
Lemma 8 (Hecke)
. The Dirichlet series (Equation (83)) satisfies the following functional equation:and is BEV (bounded in every vertical strip) and equivalent to the modular transformation (Equation (82)). Lemma 8 is a useful statement of Hecke’s epoch-making discovery [
43,
44].