ON RAMANUJAN’S FORMULA FOR ζ(1/2) AND ζ(2m + 1)
arXiv:2106.04797v1 [math.NT] 9 Jun 2021
ANUSHREE GUPTA AND BIBEKANANDA MAJI
Abstract. Page 332 of Ramanujan’s Lost Notebook contains a compelling identity for
ζ(1/2), which has been studied by many mathematicians over the years. On the same page,
Ramanujan also recorded the series,
1r
2r
3r
+
+
+··· ,
s
s
exp(1 x) − 1
exp(2 x) − 1
exp(3s x) − 1
where s is a positive integer and r −s is any even integer. Unfortunately, Ramanujan doesn’t
give any formula for it. This series was rediscovered by Kanemitsu, Tanigawa, and Yoshimoto, although they studied it only when r − s is a negative even integer. Recently, Dixit
and the second author generalized the work of Kanemitsu et al. and obtained a transformation formula for the aforementioned series with r − s is any even integer. While extending
the work of Kanemitsu et al., Dixit and the second author obtained a beautiful generalization of Ramanujan’s formula for odd zeta values. In the current paper, we investigate
transformation formulas for an infinite series, and interestingly, we derive Ramanujan’s formula for ζ(1/2), Wigert’s formula for ζ(1/k) as well as Ramanujan’s formula for ζ(2m + 1).
Furthermore, we obtain a new identity for ζ(−1/2) in the spirit of Ramanujan.
1. Introduction
The study of understanding the algebraic nature of the special values of the Riemann zeta
function ζ(s) has been a constant source of motivation for mathematicians to produce many
beautiful results. In 1735, Euler established the following elegant formula for ζ(2m). For
every positive integer m,
(2π)2m B2m
,
(1.1)
ζ(2m) = (−1)m+1
2(2m)!
where B2m denotes the 2mth Bernoulli number. These are certain rational numbers, defined
by the generating function:
∞
X
z
zn
=
,
B
n
ez − 1 n=0
n!
|z| < 2π.
Euler’s formula along with the transcendence of π, proved by F. Lindemann, immediately
tells us that all even zeta values are transcendental. However, the arithmetic nature of the
odd zeta values is yet to be determined. Surprisingly in 1979, Roger Apéry [2, 3] proved
the irrationality of ζ(3). But till today we do not know about the algebraic nature of ζ(3),
that is, whether ζ(3) is algebraic or transcendental. In 2001, Rivoal [28] Ball and Rivoal [5]
2010 Mathematics Subject Classification. Primary 11M06; Secondary 11J81 .
Keywords and phrases. Riemann zeta function, Odd zeta values, Ramanujan’s formula.
1
2
ANUSHREE GUPTA AND BIBEKANANDA MAJI
made a breakthrough by proving that there exist infinitely many odd zeta values which are
irrational, but their result does not tells us anything about the algebraic nature of a specific
odd zeta value. Around the same time, Zudilin [32] proved that at least one of the members
of ζ(5), ζ(7), ζ(9) or ζ(11) is irrational, which is the current best result in this direction.
Before going to England, Ramanujan, in his second notebook [26, p. 173, Ch. 14, Entry
21(i)], noted down the following remarkable formula for odd zeta values ζ(2m + 1).
Let α and β be two positive numbers such that αβ = π 2 . For any non-zero integer m,
1
(4α)m
∞
X n−2m−1
1
ζ(2m + 1) +
2
(e2nα − 1)
n=1
!
(−1)m+1
+
(4β)m
=
m+1
X
∞
X n−2m−1
1
ζ(2m + 1) +
2
(e2nβ − 1)
n=1
(−1)k−1
k=0
!
B2k
B2m+2−2k
αm+1−k β k .
(2k)! (2m + 2 − 2k)!
(1.2)
This formula can also be found in Ramanujan’s Lost Notebook [27, pp. 319-320, formula
(28)]. In sharp contrast to Euler’s formula, it falls short of providing an explicit formula for
odd zeta values, but it is indeed another wonderful discovery of Ramanujan that has attracted
the attention of several mathematicians. The first published proof was given by Malurkar
[21] in 1925, who had no idea about its presence in Ramanujan’s Notebook. Lerch, in 1901,
proved a particular case of the formula (1.2), namely, corresponding to α = β = π and m an
odd positive integer. Replacing m by 2m + 1 in (1.2), it reduces to
ζ(4m + 3) + 2
2m+2
∞
X (−1)j+1 B2j B4m+4−2j
X
n−4m−3
4m+3 4m+2
=
π
2
.
(e2πn − 1)
(2j)!(4m + 2 − 2j)!
n=1
(1.3)
j=0
Note that the infinite series present in the left side of (1.3) is a rapidly convergent series,
which indicates that ζ(4m + 3) is almost a rational multiple of π 4m+3 . Moreover, one can
P
n−4m−3
say that atleast one of the expressions ζ(4m + 3) or ∞
n=1 (e2πn −1) must be transcendental
since the right side expression is transcendental. A folklore conjecture is that π and odd zeta
values do not satisfy any non-zero polynomial with rational coefficients, which implies that
all odd zeta values must be transcendental.
Ramanujan’s formula (1.2) has a deep connection with the theory of Eisenstien series
over SL2 (Z) and their Eichler integrals. To know more about this connection, one can
see [12], [18]. In 1977, Bruce Berndt [7] obtained a modular transformation formula for a
generalized Eisenstien series from which he was able to derive Euler’s formula for ζ(2m) as
well as Ramanujan’s formula for ζ(2m + 1). Over the years, many mathematicians found
generalizations of (1.2) in various directions. Readers are encouraged to see the paper of
Berndt and Straub [13] for more information on the history of Ramanujan’s formula for odd
zeta values.
ON RAMANUJAN’S FORMULA FOR ζ(1/2) AND ζ(2m + 1)
3
In 2001, Kanemitsu, Tanigawa and Yoshimoto [20] studied a generalized Lambert series
defined by
∞
X
nN −2h
,
exp(nN x) − 1
n=1
(1.4)
where h, N ∈ N such that 1 ≤ h ≤ N/2. They established an important modular transformational relation for this generalized Lambert series which enabled them to find a formula
for the Riemann zeta function at rational arguments. Recently, Dixit and the second author
[15] pointed out that the above generalized Lambert series is in fact present on page 332 of
Ramanujan’s Lost Notebook [27] with more general conditions on the parameters. At the
end of page 332, Ramanujan wrote
2r
3r
1r
+
+
+ ··· ,
exp(1s x) − 1 exp(2s x) − 1 exp(3s x) − 1
(1.5)
where s is a positive integer and r − s is any even integer. Unfortunately, Ramanujan does
not give any formula for the series (1.5). Comparing (1.4) and (1.5), one can observe that
Kanemitsu et al. studied (1.5) with the restriction that r−s is a negative even integer that lies
in a restricted domain. This motivated Dixit and the second author to generalize the main
result of Kanemitsu et al. [20, Theorem 1.1] and find a formula for (1.5). While extending
the main result of Kanemitsu et al., Dixit and the second author [15, Theorem 1.2] obtained
the following beautiful generalization of the Ramanujan’s formula for ζ(2m + 1).
Let N ≥ 1 be an odd positive integer and α, β be two positive real numbers such that
αβ N = π N +1 . Then for any integer non-zero integer m 6= 0, we have
!
∞
−2N m−1
X
n
1
− 2Nm
ζ(2N m + 1) +
α N+1
2
exp ((2n)N α) − 1
n=1
N−1
−m 22m(N −1)
2N
= −β N+1
N
2N m
m+ N+3
2
2
+ (−1)
N+3
1
ζ(2m + 1) + (−1) 2
2
+m
⌊ N+1
2N
X ⌋
j=0
2
X
j= −(N−1)
2
(−1)j
∞
X
n=1
n−2m−1
iπj
1
exp (2n) N βe N − 1
2N 2 (m−j)
2j
(−1)j B2j BN +1+2N (m−j)
α N+1 β N + N+1 .
(2j)!(N + 1 + 2N (m − j))!
!
(1.6)
This formula is of interest as it establishes a relation between two distinct odd zeta values,
namely ζ(2m + 1) and ζ(2N m + 1) when N > 1. To derive the above generalization of
Ramanujan’s formula for ζ(2m + 1), Dixit and the second author studied the following line
integration representation of (1.4):
∞
X
n=1
nN −2h
1
=
N
exp(n x) − 1
2πi
where c0 > max(1, (N − 2h + 1)/N ).
Z
c0 +i∞
c0 −i∞
Γ(s)ζ(s)ζ(N s − (N − 2h))x−s ds,
4
ANUSHREE GUPTA AND BIBEKANANDA MAJI
Inspired by the work of Dixit and the second author, in the current paper, we would like
to investigate the following line integration:
Z c+i∞
1
Γ(s)ζ(ks)ζ(s − r)x−s ds,
2πi c−i∞
where k ∈ N, r ∈ Z and c > max (1/k, 1 + r). Surprisingly, when k − r is an even integer,
we obtain nice transformation formulas which allow us to derive many well-known formulas
in the literature, for example, Ramanujan’s formula for ζ(1/2), Wigert’s formula for ζ(1/k)
for k ≥ 2 even, and Ramanujan’s formula for odd zeta values. Furthermore, we also obtain
a new identity for ζ(−1/2) analogous to Ramanujan’s formula for ζ(1/2).
2. Preliminaries
The study of the well-known divisor function d(n) plays an important role in analytic number theory. This function has been generalized in many directions, and one of the important
generalizations is defined by
X
σk (m) :=
dk ,
d|m
where k ∈ Z. In this paper, we encountered two new divisor functions that are defined by
X n r
X n −r
Dk,r (n) :=
,
and
S
(n)
:=
dk−1 ,
(2.1)
k,r
k
k
d
d
k
k
d |n
d |n
where k ∈ N and r ∈ Z. At this point, it may look artificial to define the above two divisor
functions, but in due time we will see that these divisor functions naturally arise with the
theory. One can easily observe that
D1,r (n) = σr (n),
and S1,r = σ−r (n).
(2.2)
P
σk (n)
It is well-known that ζ(s)ζ(s−k) = ∞
n=1 ns , absolutely convergent for Re(s) > max{1, k+
1}. We now look for the Dirichlet series associated to Dk,r (n) and Sk,r (n) in terms of the
Riemann zeta function. Using the definition (2.1), we see that
r
∞
∞
∞
∞
X
X
X
X
Dk,r (n)
1 X 1
1
n
= ζ(ks)ζ(s − r),
(2.3)
=
=
ns
ns
nk1
nks
ns−r
k
n=1
n=1
n =1 1 n =1 2
n1 |n
1
2
valid for Re(s) > max k1 , 1 + r . Similarly, with the help of the definition (2.1), one can
show that, for Re(s) > max(1, 1 − r),
∞
X
Sk,r (n)
n=1
ns
= ζ(1 + ks − k)ζ(s + r).
(2.4)
P∞
−mx associated to the divisor function
We observe that the Lambert series
m=1 σk (m)e
σk (m) is in fact present in Ramanujan’s formula (1.2) as it can be clearly seen from the
ON RAMANUJAN’S FORMULA FOR ζ(1/2) AND ζ(2m + 1)
5
relation that
∞
X
m=1
σk (m)e−mx =
∞
X
m=1
X
d|m
dk e−mx =
∞ X
∞
X
dk e−mdx =
m=1 d=1
∞
X
nk
.
enx − 1
n=1
P
−nx associated
Let us try to find a similar expression for the Lambert series ∞
n=1 Dk,r (n)e
P∞
k
to the new divisor function Dk,r (n). We know that, for x > 0, m=1 e−n mx = nk1x . Now
e
−1
differentiating both sides of this series r times, and then taking sum over n, we obtain
∞
X
dr
1
r k
r −nk mx
(−1) (n m) e
=
dxr enk x − 1
m=1
∞
∞ X
∞
X
X
1
(−1)r dr
k
mr e−n mx =
⇒
nkr dxr enk x − 1
n=1
n=1 m=1
∞
∞
X
X
(−1)r dr
1
−nx
, for r ≥ 0.
(2.5)
⇒
Dk,r (n)e
=
nkr dxr enk x − 1
n=1
n=1
Next, we state another important special function, the Meijer G-function, which is the
generalization of many well-known special functions in the literature.
Let 0 ≤ m ≤ q, 0 ≤ n ≤ p. Let a1 , · · · , ap and b1 , · · · , bq complex numbers such that
ai − bj 6∈ N for 1 ≤ i ≤ n and 1 ≤ j ≤ m. Then the Meijer G-function [24, p. 415, Definition
16.17] is defined by the line integral:
!
Qn
Z Qm
−s
1
j=1 Γ(bj + s)
j=1 Γ(1 − aj − s)z
m,n a1 , · · · , ap
Qq
Qp
Gp,q
z =
ds,
(2.6)
2πi L j=m+1 Γ(1 − bj − s) j=n+1 Γ(aj + s)
b1 , · · · , b q
where the vertical line of integration L, going from −i∞ to +i∞, separates the poles of
the factors Γ(1 − aj − s) from those of the factors Γ(bj + s). This integral converges if
p + q < 2(m + n) and | arg(z)| < (m + n − p+q
2 )π.
Now we are ready to state the main results.
3. Main results
Theorem 3.1. Let k ≥ 2 and r be an even integers. Let Dk,r (n) and Sk,r (n) be defined as
in (2.1). Then for any x > 0, we have
k+r−2
k+1−2r
∞
X
(−1) 2 (2π) 2
1
1
1
1
−nx
− k1
Dk,r (n)e
= − ζ(−r) + Γ
− r x + R1+r +
ζ
2k−1
2
k
k
k
xk 2
n=1
!
(k−1)
∞
X ′′ X
{}
k,0
Sk,r (n) G0,k
×
(k−1) X(j) ,
1
r,
−
,
·
·
·
,
−
k
k
j=−(k−1) n=1
where
′′
means summation runs over j = −(k − 1), −(k − 3), · · · , (k − 3), (k − 1), and
X(j) := Xx,n,k (j) :=
e−
iπj
2
(2π)k+1 n
,
kk x
(3.1)
6
ANUSHREE GUPTA AND BIBEKANANDA MAJI
and
R1+r =
r!ζ(k(1 + r))x−(1+r) ,
(−1)1+r kζ ′ (k(1
(−(1+r))!
+ r))x−(1+r) ,
if r ≥ 0,
if r < 0.
(3.2)
As an immediate consequence of the above identity, we obtain Ramanujan’s formula for
ζ( 12 ).
Corollary 3.2. Let α and β be two positive numbers such that αβ = 4π 3 .Then
√
√ √ X
√
√
∞
1
1 π2
β
β
cos( mβ) − sin( mβ) − e− mβ
√
√
.
= +
+
ζ
+
√
n2 α − 1
4
6α
4π
2
4π
m(cosh(
mβ)
−
cos(
mβ))
e
n=1
m=1
∞
X
1
(3.3)
Ramanujan1 recorded this formula on the page 332 of his Lost Notebook [27], where Dixit
and the second author encountered the series (1.5). On the same page, Ramanujan mentioned
another form of this identity, which can also be found in Ramanujan’s second notebook [26,
Entry 8, Chapter 15] and in [8, p. 314]. The formula (3.3) has been generalized by many
mathematicians, for more information readers can see [4, pp. 191-193], [11, p. 859, Theorem
10.1].
More generally, substituting r = 0 in Theorem 3.1, we obtain the following Wigert’s formula
[31] for ζ( k1 ), for k ≥ 2 even.
Corollary 3.3. For any x > 0 and k ≥ 2 even, we have
1
1
k
ζ(k) 1
1
1
1 k 1 (−1) 2 −1 2π k
=
+ +
+ Γ
ζ
x
k
k
k
x
4
k
x
enk x − 1
n=1
k
!
−1 (
1
2
X
iπ(2j+1)(k−1)
2π k − iπ(2j+1))
2k
2k
e
×
e
L̄k 2π
x
j=0
!)
1
k
iπ(2j+1)
iπ(2j+1)(k−1)
2π
2k
,
L̄k 2π
e 2k
+ e−
x
∞
X
1
(3.4)
where
L̄k (x) :=
∞
X
n=1
1
n k −1
1
exp(xn k ) − 1
.
This formula has been generalized by Dixit and the second author [15, Theorem 1.5]
and further a two-variable generalization obtained by Dixit, Gupta, Kumar, and Maji [17,
Theorem 2.12].
1Ramanujan [27, p. 332] missed the factor 1/√m on the right side series expression of (3.3).
ON RAMANUJAN’S FORMULA FOR ζ(1/2) AND ζ(2m + 1)
7
Theorem 3.4. Let k ≥ 2 be an even integer and r 6= −1 be an odd integer. Let Dk,r (n) and
Sk,r (n) be defined as in (2.1). Then for any x > 0, we have
2k+r−1
k+1−2r
∞
X
1
1
1
1
(−1) 2 (2π) 2
− k1
−nx
− r x + R1+r +
ζ
Dk,r (n)e
= − ζ(−r) + Γ
2k−1
2
k
k
k
xk 2
n=1
!
(k−1)
∞
X ′′ X
{}
k,0
×
X(j) ,
Sk,r (n)G0,k
ij
r, − k1 , · · · , − (k−1)
k
n=1
j=−(k−1)
where X(j) and R1+r are defined as in (3.1) and (3.2) respectively.
Substituting k = 2 and r = 1 in the above result, we derive an interesting identity for
ζ − 12 .
Corollary 3.5. For α, β > 0 such that αβ = 4π 3 ,
√
∞
X
1
1
1 d
1
π4
β
+
ζ
−
=
+
2
n2 dα 1 − en α
24
4π
2
90α2
n=1
(
!
√
√
√
∞
cos( mβ) + sin( mβ) − e− mβ
π X 1
√
√
√
−
4α
m
cosh(
mβ
mβ)
−
cos(
mβ)
m=1
)
√
√
2 sin( mβ) sinh( mβ)
√
√
√
.
+
(cos(mβ) cosh( mβ) − 1)2 + (sin( mβ) sinh( mβ))2
Note that, Theorem 3.4 does not hold for r = −1, so we present the next result corresponding to r = −1.
Theorem 3.6. Let k ≥ 2 be an even integer. Then for x > 0, we have
∞
X
1
x
1
1
1
1
Dk,−1 (n)e−nx = log
Γ
+
ζ
1
+
x− k
k
2
(2π)
k
k
k
n=1
(k−1)
k
+ (−1)
X
j=−(k−1)
′′
iπj
e
∞
X
m=1
− iπj
2k
log 1 − exp −e
1+ k1
(2π)
m 1
k
x
.
Till now, we have stated identities when k ≥ 2 is even. From the next result onwards we
dealt with k ≥ 1 odd. It turns out that, Ramanujan’s formula for ζ(2m + 1) can be derived
as a special case of our result when k is an odd integer.
Theorem 3.7. Let k ≥ 1 be an odd integer and r 6= −1 be an odd integer. Let Dk,r (n) and
Sk,r (n) be defined as in (2.1). Then for x > 0, we have
2k+r−1
k+1−2r
∞
X
1
1
1
1
(−1) 2 (2π) 2
1
− r x− k + R1+r + R +
ζ
Dk,r (n)e−nx = − ζ(−r) + Γ
2k−1
2
k
k
k
xk 2
n=1
!
(k−1)
∞
X ′′ X
{}
k,0
X(−j) ,
(3.5)
×
ij
Sk,r (n)G0,k
r, − k1 , · · · , − (k−1)
k
n=1
j=−(k−1)
8
ANUSHREE GUPTA AND BIBEKANANDA MAJI
where X(j) and R1+r are defined as in (3.1) and (3.2) respectively, and
0,
if r ≥ 1,
1+r
R=
−(1+r)
i+1
−r
P
(−1)
Bk(2i+1)+1 B−2i−1−r (2π)
(−1) 2
x 2i+1
2
, if r ≤ −3.
i=0
2
2π
(2i+1)!(k(2i+1)+1) (−(2i+1+r))!
The next result provides a relation between two odd zeta values, namely, ζ(2m + 1) and
ζ(2km + 1), and a zeta value at rational argument ζ k1 + 2m + 1 .
Corollary 3.8. Let k ≥ 1 be an odd integer and m be a positive integer. For any x > 0, we
have
∞
X
1
1
1
1
1
−nx
Dk,−(2m+1) (n)e
= − ζ(2m + 1) + Γ
+ 2m + 1 x− k
ζ
2
k
k
k
n=1
2m
x
(−1)m k(2km)!ζ(2km + 1)
+
2
(2m)!
(2π)k
m
x 2i+1
(−1)i+1 Bk(2i+1)+1 B2m−2i
(−1)m (2π)2m+1 X
+
2
(2i + 1)!(k(2i + 1) + 1)(2m − 2i)! 2π
i=0
+
(−1)m (2π)
k+1
+2m+1
2
(k−1)
1
x kk− 2
X
j=−(k−1)
k,0
× G0,k
′′
j
i
∞
X
Sk,−(2m+1) (n)
n=1
!
{}
X(−j) .
−(2m + 1), − k1 , · · · , − (k−1)
k
Substituting k = 1 in Corollary 3.8, and simplifying, we derive Ramanujan’s formula for
odd zeta values.
Corollary 3.9. Let m be a positive integer. For any x > 0,
!
∞
∞
x 2m ζ(2m + 1) X
X
2
1
−nx
− 4πx n
m
σ−(2m+1) (n)e
+ ζ(2m + 1) = (−1)
σ−(2m+1) (n)e
+
2
2π
2
n=1
n=1
m+1
1 X (−1)i+1 B2i B2m+2−2i x2m+1 2π 2i
.
(3.6)
+
2
(2i)!(2m + 2 − 2i)!
x
i=0
One can easily show that (3.6) is equivalent to (1.2) under the substitution x = 2α and
αβ = π 2 . Again, letting k = 1 and r = 2m + 1, for m ≥ 1, in Theorem 3.7, one can obtain
the following interesting identity.
Corollary 3.10. Let m be a positive integer. For α, β > 0 with αβ = π 2 , we have
αm+1
∞
∞
X
X
B2m+2
n2m+1
n2m+1
m+1
−
(−β)
= (αm+1 − (−β)m+1 )
.
2nα
2nβ
e
−1
e
−1
4m + 4
n=1
(3.7)
n=1
This formula is a special case of (1.2) and it can be found in [26, Vol. 1, p. 259].
Note that Theorem 3.7 is not valid for r = −1. Corresponding to r = −1 and k ≥ 1 odd,
we obtain the next result.
ON RAMANUJAN’S FORMULA FOR ζ(1/2) AND ζ(2m + 1)
9
Theorem 3.11. Let k ≥ 1 be an odd integer. Then for x > 0, we have
∞
X
1
x
1
1
1
ζ(−k)x
1
−nx
+ Γ
ζ 1+
x− k +
Dk,−1 (n)e
= log
k
2
(2π)
k
k
k
2
n=1
(k−1)
−
X
′′
j=−(k−1)
∞
X
m=1
log 1 − exp −e
iπj
2k
1+ k1
(2π)
m 1
k
x
.
(3.8)
Plugging k = 1 in the above identity, we obtain an interesting identity.
Corollary 3.12. For α, β > 0 and αβ = π 2 ,
∞
X
n=1
1
n(e2nα − 1)
−
∞
X
n=1
1
n(e2nβ
β−α 1
=
− log
− 1)
12
4
α
.
β
(3.9)
Remark 1. This identity is popularly known as the transformation formula for the logarithm
πiz Q
2nπiz ), which is an important
of the Dedekind eta function η(z), defined as e 12 ∞
n=1 (1 − e
example of a half-integral weight modular form. Ramanujan recorded this identity twice in
his second notebook [26, Ch. 14, Sec. 8, Cor. (ii) and Ch. 16, Entry 27(iii)]. For more
information, one can see [8, p. 256], [9, p. 43], [27, p. 320, Formula (29)].
In the next chapter we mention a few important results which will be useful throughout
the manuscript. We mention the proof of those results which are not well-known.
4. Well-known results
An asymmetric form of the functional equation of ζ(s) can be written as
πs
.
ζ(s) = 2s π s−1 ζ(1 − s)Γ(1 − s) sin
2
(4.1)
We know that ζ(s) has a simple pole at s = 1 with residue 1 and the Laurent series expansion
is
ζ(s) =
∞
X
1
(−1)n γn (s − 1)n
+γ+
,
s−1
n!
n=1
where γn are called the Stieltjes constants and defined by
i
n+1
n
X
(log
i)
(log
j)
,
−
γn = lim
i→∞
j
n+1
(4.2)
j=1
and γ is the well-known Euler–Mascheroni constant. The Laurent series expansion of Γ(s) at
s = 0 is
1
1
π2
1
π2
2
3
+ 2ζ(3) s2 + · · · .
(4.3)
Γ(s) = − γ +
γ +
s−
γ +γ
s
2
6
6
2
More generally, Γ(s) has Laurent series expansion at every negative integer.
10
ANUSHREE GUPTA AND BIBEKANANDA MAJI
Lemma 4.1. For r ≤ −2, we have
a−1
+ a0 + a1 (s − (1 + r)) + · · · ,
Γ(s) =
s − (1 + r)
P−(1+r)
(−1)r+1
(−1)r+1
−γ + k=1 k1 .
and a0 = (−(1+r))!
where a−1 = (−(1+r))!
Euler proved that Γ(s) satisfy the following beautiful reflection formula:
π
Γ(s)Γ(1 − s) =
, ∀s ∈ C − Z.
sin(πs)
(4.4)
The multiplication formula for the gamma function stated as
Lemma 4.2. For any positive integer k,
kks
Γ(ks) = √
k−1
k(2π) 2
In particular, substituting k = 2, we have
k−1
Y
l=1
l
Γ(s)Γ s +
k
.
(4.5)
1
22s−1
Γ(s)Γ s +
.
Γ(2s) =
π
2
This is known as duplication formula.
The next result gives an important information about the asymptotic expansion of the
gamma function, popularly known as Stirling’s formula.
Lemma 4.3. For s = σ + iT in a vertical strip α ≤ σ ≤ β,
√
1
σ−1/2 − 12 π|T |
,
1+O
e
|Γ(σ + iT )| = 2π|T |
|T |
(4.6)
as |T | → ∞.
Proof. The proof of this result can be found in [19, p. 151].
Lemma 4.4. For σ ≥ σ0 , there exist a constant C(σ0 ), such that
|ζ(σ + iT )| ≪ |T |C(σ0 )
as |T | → ∞.
Proof. One can find the proof in [30, p. 95].
Next, we shall mention a few important special cases of the Meijer G-function. An immediate special case is the following result. Putting m = q = 1 and n = p = 0 in (2.6), one can
see that
!
Z c+i∞
1
1,0 {}
G0,1
z =
Γ(b + s)z −s ds = e−z z b ,
(4.7)
2πi c−i∞
b
provided Re(s) = c > − Re(b) and Re(z) > 0. Again, plugging m = q = 2 and n = p = 0 in
the definition (2.6) of the Meijer G-function, we will have the following result.
ON RAMANUJAN’S FORMULA FOR ζ(1/2) AND ζ(2m + 1)
11
Lemma 4.5. Assuming the real part of the line of integration is bigger than max(−Re(b1 ), − Re(b2 )).
For | arg(z)| < π, we have
!
√
1
{}
2,0
G0,2
z = 2z 2 (b1 +b2 ) Kb1 −b2 (2 z),
b1 , b2
where Kν (z) is the modified Bessel function.
Proof. This can be verified from [23, p. 115, Equn (11.1)].
Lemma 4.6. Let k ≥ 2 be a positive integer. Then
k,0
G0,k
provided | arg(z)| <
{}
2
1
b, b + k , b + k , · · · , b +
(k−1)
k
z
!
k−1
(2π) 2 b −kz 1/k
= √
z e
,
k
kπ
2 .
Proof. With the help of the definition (2.6) of the Meijer-G function, we write
!
Z Y
k
{}
1
l
k,0
z =
G0,k
z −s ds,
Γ(s + b)Γ s + b +
2πi
k
b, b + k1 , · · · , b + (k−1)
L
k
l=1
provided the poles of each of these gamma factor lie on the left of the line of integration.
Replacing s by s + b in the multiplication formula (4.5) for Γ(s), we have
k−1
kk(s+b) Y
l
.
Γ(k(s + b)) = √
Γ(s + b)Γ s + b +
k−1
k
k(2π) 2 l=1
Plugging this expression, one can see that
!
Z
√
k−1 1
{}
Γ(k(s + b)) −s
k,0
2
G0,k
k(2π)
z
=
z ds
(k−1)
1
2πi L kk(s+b)
b, b + k , · · · , b + k
Z
√
s1
k−1 1
ds1
2
Γ(s1 )z − k +b s1 +1
= k(2π)
2πi L′
k
k−1
Z
1
(2π) 2 1
Γ(s1 )(kz k )−s1 z b ds1
= √
k 2πi L′
k−1
(2π) 2 z b −kz k1
√
,
=
e
k
the last step is possible only when | arg(z)| <
kπ
2 .
Dixit and Maji [15, Lemma 3.1] used the below lemma to prove (1.6).
Lemma 4.7. Let a, u, v be three real numbers. Then
cos(a sin(u) + uv) − e−a cos(u) cos(uv)
eiuv
.
=
2 Re
exp(ae−iu ) − 1
cosh(a cos(u)) − cos(a sin(u))
In a similar vein, we have the following lemma.
12
ANUSHREE GUPTA AND BIBEKANANDA MAJI
Lemma 4.8. Let a, u, v be three real numbers. Then
ieiuv
− sin(uv + a sin(u)) + e−a cos(u) sin(uv)
.
2 Re
=
exp(ae−iu ) − 1
cosh(a cos(u)) − cos(a sin(u))
Proof. The left hand side expression can be written as
− sin(uv) + i cos(uv)
ieiuv
= 2 Re a cos(u)
2 Re
,
exp(ae−iu ) − 1
e
(cos(a sin(u)) − i sin(a sin(u))) − 1
now multiplying the numerator and the denominator by the conjugate of the denominator
reduces to
!
(− sin(uv) + i cos(uv))(ea cos(u) cos(a sin(u)) + iea cos(u) sin(a sin(u))) − 1)
2 Re
e2a cos(u) − 2ea cos(u) (cos(a sin(u))) + 1
!
−ea cos(u) (cos(a sin(u)) sin(uv)) + sin(uv) − ea cos(u) (sin(a sin(u)) cos(uv))
=2
e2a cos(u) − 2ea cos(u) (cos(a sin(u))) + 1
!
−ea cos(u) (sin(uv + a sin(u)) + sin(uv)
,
=2
e2a cos(u) − 2ea cos(u) (cos(a sin(u))) + 1
=
− sin(uv + a sin(u)) + e−a cos(u) sin(uv)
.
cosh(a cos(u)) − cos(a sin(u))
In the final step we divided by ea cos(u) on the numerator as well as on the denominator.
The next lemma will play a crucial role to obtain our main results.
Lemma 4.9. Let s ∈ C and k ∈ N. Then
sin(ks)
=
sin(s)
(k−1)
X
′′
exp(ijs),
j=−(k−1)
where ′′ means summation runs over j = −(k − 1), −(k − 3), · · · , (k − 3), (k − 1). Thus, for
k even,
(k−1)
X ′′
k
sin(ks)
2
= (−1)
ij exp(ijs),
(4.8)
cos(s)
j=−(k−1)
and for k odd,
k−1
cos(ks)
= (−1) 2
cos(s)
(k−1)
X
′′
ij exp(−ijs).
(4.9)
j=−(k−1)
5. Proof of main results
Proof of Theorem 3.1. It is well-known that e−x is the inverse Mellin transform of the
gamma function Γ(s), that is, for any x > 0,
Z c+i∞
1
−x
Γ(s)x−s ds,
(5.1)
e =
2πi c−i∞
ON RAMANUJAN’S FORMULA FOR ζ(1/2) AND ζ(2m + 1)
13
valid for any c > 0. Replacing x by nx in (5.1) and using (2.3), one can easily show that
∞
X
Dk,r (n)e−nx =
n=1
1
2πi
Z
c+i∞
c−i∞
Γ(s)ζ(ks)ζ(s − r)x−s ds,
(5.2)
provided with Re(s) = c > max( k1 , 1 + r). In order to evaluate this line integral we choose
a contour C in a way so that all the poles of the integrand function lie inside the contour C,
determined by the line segments [c − iT, c + iT ], [c + iT, −λ + iT ], [−λ + iT, −λ − iT ], and
[−λ − iT, c − iT ], where we wisely take λ > max(0, −r) and T is some large positive real
number. The essence of this choice of λ is justified at the later stage in our proof. Now
applying Cauchy’s residue theorem, we have
Z c+iT Z −λ+iT Z −λ−iT Z c−iT
1
Γ(s)ζ(ks)ζ(s − r)x−s ds
+
+
+
2πi c−iT
−λ−iT
−λ+iT
c+iT
X
=
Res Γ(s)ζ(ks)ζ(s − r)x−s ,
(5.3)
ρ
s=ρ
where the sum over ρ runs through all the poles of Γ(s)ζ(ks)ζ(s − r)x−s inside the contour
C.
First, let us analyse the poles of the integrand function. We know that Γ(s) has simple
poles at s = 0 and negative integers. The poles at negative integers are getting cancelled by
the trivial zeros of ζ(ks) as we are dealing with k even. Thus, the only simple poles of the
integrand function Γ(s)ζ(ks)ζ(s − r)x−s are at s = 0, k1 and 1 + r, due to the simple pole of
ζ(ks) at s = 1/k and the pole of ζ(s − r) at s = 1 + r. Here we would like to mention that,
as we have considered r is any even integer, so it may happen that 1 + r is a negative integer,
and in that case, it seems that 1 + r is a pole of order 2 due to the factor Γ(s)ζ(s − r), but
ζ(ks) will have a zero at 1 + r since k is even. This shows that 1 + r remains a simple pole
of the integrand function as we are dealing with k ≥ 2 is even and r is even.
One can easily evaluate residues at s = 0 and s = k1 . Let Rρ denotes the residue term at
ρ. Using residue calculation methods, we have
1
R0 = lim sΓ(s)ζ(ks)ζ(s − r)x−s = ζ(0)ζ(−r) = − ζ(−r).
s→0
2
(5.4)
Similarly, we obtain
R 1 = lim
k
s→ k1
s−
1
k
Γ(s)ζ(ks)ζ(s − r)x−s =
1
Γ
k
1
1
1
− r x− k .
ζ
k
k
(5.5)
Here we note that while calculating the residue at s = 1 + r, we have to make two different
cases. First, if r ≥ 0, then
R1+r = lim (s − (1 + r))Γ(s)ζ(ks)ζ(s − r)x−s = Γ(1 + r)ζ(k(1 + r))x−(1+r)
s→1+r
= r!ζ(k(1 + r))x−(1+r) .
(5.6)
14
ANUSHREE GUPTA AND BIBEKANANDA MAJI
Second, if r < 0 is a negative even integer, then 1 + r is also a negative integer. Thus, Γ(s)
satisfy the Laurent series expansion (4.1) at 1 + r. Therefore, we have
lim (s − 1 − r)Γ(s) =
s→1+r
(−1)1+r
.
(−(1 + r))!
(5.7)
Again, ζ(ks) has a trivial zero at s = 1 + r as k is even, so we have
ζ(ks)
= kζ ′ (k(1 + r)).
s→1+r (s − 1 − r)
lim
(5.8)
Combining (5.7) and (5.8) and together with the fact that 1 + r is a simple pole of ζ(s − r),
we obtain
R1+r =
lim (s − (1 + r))Γ(s)ζ(ks)ζ(s − r)x−s =
s→(1+r)
(−1)1+r
kζ ′ (k(1 + r))x−(1+r) . (5.9)
(−(1 + r))!
We now proceed to show that the horizontal integrals
Z −λ+iT
1
H1 (T, x) :=
Γ(s)ζ(ks)ζ(s − r)x−s ds
2πi c+iT
and
H2 (T, x) =
1
2πi
Z
c−iT
−λ−iT
Γ(s)ζ(ks)ζ(s − r)x−s ds
vanish as T → ∞. Replace s by σ + iT in H1 (T, x) to see
Z d
1
|H1 (T, x)| =
Γ(σ + iT )ζ(kσ + ikT )ζ(σ + iT − r)x−σ+iT dσ
2πi c
Z d
|Γ(σ + iT )ζ(kσ + ikT )ζ(σ + iT − r)|x−σ dσ
≪
c
π
≪ |T |A exp − |T | ,
2
for some constant A. In the final step, we used Stirling’s formula for Γ(s) i.e., Lemma 4.3 and
bound for ζ(s) i.e., Lemma 4.4. This immediately implies that H1 (T, x) vanishes as T → ∞.
Similarly, one can show that H2 (T, x) also vanishes as T → ∞. Now letting T → ∞ in (5.3)
and collecting all the residual terms, we have
Z
1
Γ(s)ζ(ks)ζ(s − r)x−s ds = R0 + R 1 + R1+r
k
2πi (c)
Z
1
Γ(s)ζ(ks)ζ(s − r)x−s ds.
(5.10)
+
2πi (−λ)
R
R c+i∞
Here and throughout the paper we denote (c) by c−i∞ . Now one of our main objectives is
to evaluate the following vertical integral
Z
Γ(s)ζ(ks)ζ(s − r)x−s ds.
(5.11)
Vk,r (x) :=
(−λ)
ON RAMANUJAN’S FORMULA FOR ζ(1/2) AND ζ(2m + 1)
Utilize the asymmetric form (4.1) of the functional equation of ζ(s) to write
πks
ks ks−1
ζ(ks) = 2 π
ζ(1 − ks)Γ(1 − ks) sin
,
2
π(s − r)
s−r s−r−1
.
ζ(s − r) = 2 π
ζ(1 − s + r)Γ(1 − s + r) sin
2
15
(5.12)
(5.13)
Substituting (5.12) and (5.13) in (5.11), the vertical integral becomes
Z
1
1
Γ(s)Γ(1 − ks)Γ(1 − s + r)ζ(1 − ks)ζ(1 − s + r)
Vk,r (x) =
(2π)r π 2 2πi (−λ)
s
π(s − r)
(2π)k+1
πks
ds.
sin
× sin
2
2
x
Now to shift the line of integration, we replace s by 1 − s. Then we have
Z
(2π)k+1−r 1
Γ(1 − s)Γ(1 − k + ks)Γ(s + r)ζ(1 − k + ks)ζ(s + r)
Vk,r (x) =
π2 x
2πi (1+λ)
−s
π(1 − s − r)
(2π)k+1
πk(1 − s)
ds.
(5.14)
sin
× sin
2
2
x
Since k and r are both even, the following identities hold
k
ksπ
πk(1 − s)
−1
sin
= (−1) 2 sin
,
2
2
πs
r
π(1 − s − r)
.
= (−1) 2 cos
sin
2
2
(5.15)
(5.16)
Using the above two expressions in (5.14), we see
k+r
Z
(−1) 2 −1 (2π)k+1−r 1
Vk,r (x) =
Γ(1 − s)Γ(1 − k + ks)Γ(s + r)ζ(1 − k + ks)
π2 x
2πi (1+λ)
πs (2π)k+1 −s
πks
×ζ(s + r) sin
cos
ds
2
2
x
k+r
Z
(−1) 2 −1 (2π)k+1−r 1
Γ(1 − s)Γ(1 − k + ks)Γ(s + r)ζ(1 − k + ks)
=
2π 2 x
2πi (1+λ)
−s
sin πks
(2π)k+1
2
sin(πs)
×ζ(s + r)
ds
x
sin πs
2
k+r
Z
Γ(1 − k + ks)Γ(s + r)
(−1) 2 −1 (2π)k+1−r 1
ζ(1 − k + ks)
=
2πx
2πi (1+λ)
Γ(s)
−s
sin πks
(2π)k+1
2
×ζ(s + r)
ds.
(5.17)
x
sin πs
2
Here in the ultimate step, we have used Euler’s reflection formula (4.4) for Γ(s) and in the
in the denominator as well as in the numerator.
penultimate step we multiplied by sin πs
2
Now we notice that Re(s + r) = 1 + λ + r > 1 and Re(1 − k + ks) = 1 + kλ > 1 as
we have considered λ > max(0, −r). This explains our choice of λ. Thus, we can express
16
ANUSHREE GUPTA AND BIBEKANANDA MAJI
ζ(1−k+ks)ζ(s+r) as an infinite series, mainly, use (2.4) and then interchange the summation
and integration in (5.17) to deduce that
Vk,r (x) =
(−1)
k+r
−1
2
x
∞
(2π)k−r X
Sk,r (n)Ik,r (n, x),
(5.18)
n=1
where
1
Ik,r (n, x) :=
2πi
Z
(1+λ)
Γ(1 − k + ks)Γ(s + r) sin
Γ(s)
sin
πks
2
πs
2
(2π)k+1 n
x
−s
ds.
(5.19)
Now our main goal is to simplify this integral. Invoking Lemma 4.9 in (5.19) we see that
!−s
iπj
(k−1)
X ′′ 1 Z
Γ(1 − k + ks)Γ(s + r) e− 2 (2π)k+1 n
ds.
(5.20)
Ik,r (n, x) =
2πi (1+λ)
Γ(s)
x
j=−(k−1)
Employing multiplication formula (4.5) for Γ(s), one can derive that
Γ(1 − k + ks)
=
Γ(s)
kks
1
kk− 2 (2π)
k−1
2
k−1
Y
l=1
l
Γ s−
k
.
(5.21)
Substituting (5.21) in (5.20) yields that
Ik,r (n, x) =
(k−1)
1
1
kk− 2 (2π)
k−1
2
X
j=−(k−1)
′′
1
2πi
Z
Γ(s + r)
(1+λ)
k−1
Y
l=1
l
Γ s−
k
X(j)−s ds,
(5.22)
where
X(j) =
e−
iπj
2
(2π)k+1 n
kk x
(5.23)
is exactly same as defined in (3.1). Now we shall try to express, the line integral in (5.22),
in terms of the Meijer G-function and for that we have to verify all the necessary conditions
for the convergence of the integral. First, one can check that all the poles of the integrand
function lie on the left of the line of integration Re(s) = 1 + λ since λ > max(0, −r). Now
using the definition (2.6) of the Meijer G-function, with m = q = k and n = p = 0 and
b1 = r, b2 = − k1 , b3 = − k2 , · · · , bk = − k−1
k , we find that
!
Z
k−1
Y
{}
1
l
k,0
−s
(5.24)
Γ(s + r)
Γ s−
X(j) ds = G0,k
(k−1) X(j) ,
1
2πi (1+λ)
k
,
·
·
·
,
−
r,
−
k
k
l=1
p+q
πk
convergent since p + q < 2(m + n) and | arg(X(j))| = | πj
2 | < (m + n − 2 ) = 2 as |j| ≤ k − 1.
Substituting (5.24) in (5.22), we arrive at
!
(k−1)
X ′′ k,0
{}
1
G0,k
X(j) .
(5.25)
Ik,r (n, x) =
k−1
1
r, − k1 , · · · , − (k−1)
kk− 2 (2π) 2 j=−(k−1)
k
ON RAMANUJAN’S FORMULA FOR ζ(1/2) AND ζ(2m + 1)
17
Now plugging this final expression (5.25) of Ik,r (n, x) in (5.18), the left vertical integral
becomes
!
k+1
k+r
(k−1)
∞
{}
(−1) 2 −1 (2π) 2 −r X ′′ X
k,0
X(j) .
Vk,r (x) =
Sk,r (n)G0,k
k− 21
r, − k1 , · · · , − (k−1)
k
x
k
j=−(k−1) n=1
(5.26)
At this moment, we must show that the infinite series
∞
X
k,0
Sk,r (n)G0,k
n=1
!
{}
X(j)
r, − k1 , · · · , − (k−1)
k
(5.27)
is convergent for any fixed k ≥ 1, r ∈ Z, |j| ≤ k − 1. To show the convergence of this series,
we shall use the integral representation (5.24) of the Meijer G-function. Employing Stirling’s
formula (4.6) on each gamma factor that is present in (5.24), letting s = 1 + λ + iT , we have
π
1
|Γ(s + r)| = |Γ(1 + λ + r + iT ) = O |T |λ+r+ 2 e− 2 |T | ,
1
π
l
l
l
Γ s−
= Γ 1 + λ − + iT = O |T |λ− k + 2 e− 2 |T | , for 1 ≤ l ≤ k − 1,
k
k
as T → ∞. Using the definition (5.23) of X(j), we note that
−s
|X(j)
|=
e−
iπj
2
(2π)k+1 n
xkk
!−1−λ−iT
= Ok,x
1
n1+λ
e
π|j||T |
2
.
Utilizing the above bounds for gamma functions and the bound for X(j), and upon simplification, we derive that
!
Z ∞
{}
1
k,0
kλ+r+ 12 − π2 (k−|j|)|T |
|T |
e
dT
G0,k
X(j) ≪ Ok,x
n1+λ −∞
r, − k1 , · · · , − (k−1)
k
1
≪ Ok,r,x
.
(5.28)
n1+λ
The integral present inside the above big-oh bound is convergent since |j| ≤ k − 1 and
kλ + r + 32 is a positive quantity as we have λ > max(0, −r). Plugging (5.28) in (5.27), we can
P
−(1+λ)
see that the infinite series (5.27) is convergent since the Dirichlet series ∞
n=1 Sk,r (n)n
is convergent.
Finally, combining (5.2), (5.10), (5.11), and (5.26), and together with all the residual terms
(5.4), (5.5) (5.6), and (5.9), we complete the proof of Theorem 3.1.
5.1. Recovering Ramanujan’s formula for ζ(1/2).
18
ANUSHREE GUPTA AND BIBEKANANDA MAJI
Proof of Corollary 3.2. Substituting k = 2 and r = 0 in Theorem 3.1, one can see that
√
∞
X
π
1
1
π2
−nx
D2,0 (n)e
= + √ ζ
+
4 2 x
2
6x
n=1
!
!#
"
iπ
iπ
∞
{} e 2 2nπ 3
{} e− 2 2nπ 3
π 3/2 X
2,0
2,0
+ G0,2
.
S2,0 (n) G0,2
+
x
x
x
0, − 12
0, − 12
n=1
Observe that the Meijer G-functions that are present in the above equation are conjugate to
each other. Thus, letting x = α and αβ = 4π 3 , and using (2.5), we obtain
√
∞
X
1
1
1
π2
π
√
=
+
+
ζ
2
4 2 α
2
6α
en α − 1
n=1
!#
"
∞
− iπ
2 nβ
e
2π 3/2 X
{}
2,0
+
.
(5.29)
S2,0 (n) Re G0,2
α
2
0, − 21
n=1
Recall Lemma 4.6 with k = 2 and b = − 12 , to produce
! r
p
− iπ
2 nβ
2π iπ
e
{}
2,0
− iπ
4 exp −
4
.
2nβe
G0,2
e
=
2
nβ
0, − 21
(5.30)
Now utilizing (5.30) in (5.29), and using the definition (2.1) of S2,0 (n), we see that
√
∞
X
π
1
1
1
π2
√
=
+
ζ
+
4 2 α
2
6α
en2 α − 1
n=1
r
∞
i
h iπ
p
iπ
2π X X d
√ Re e 4 exp − 2nβe− 4
+
.
α
n
n=1 2
(5.31)
d |n
Next aim is to simplify the infinite series
∞ X
i
p
h iπ
X
iπ
d
√ Re e 4 exp − 2nβe− 4
.
n
n=1 2
d |n
Writing n = d2 m, one can simplify the infinite sum in a following way
∞ X
i
h iπ
p
X
iπ
d
√ Re e 4 exp − 2nβe− 4
n
n=1 2
d |n
∞ X
∞
X
h iπ
i
p
iπ
1
√ Re e 4 exp − 2mβde− 4
m
m=1 d=1
iπ
∞
X
1
e4
√
√ Re
=
− iπ
m
exp
2mβe 4 − 1
m=1
=
∞
X
1
√
=
2 2m
m=1
!
√
√
√
cos( mβ) − sin( mβ) − e− mβ
√
√
.
(cosh( mβ) − cos( mβ))
In the final step, we have used Lemma 4.7 with u = π/4, v = 1, and a =
substituting (5.32) in (5.31), one can finish the proof of (3.3).
(5.32)
√
2mβ. Now
ON RAMANUJAN’S FORMULA FOR ζ(1/2) AND ζ(2m + 1)
19
5.2. Wigert’s formula for ζ(1/k) for k ≥ 2 even.
Proof. First, substituting r = 0 in Theorem 3.1, we find that
∞
X
1
1
1 1
1
ζ(k)
−nx
Dk,0 (n)e
= + Γ
ζ
x− k +
4 k
k
k
x
n=1
+
(−1)
k−2
2
k
iπj
− 2
(2π)
2k−1
2
(k−1)
k+1
2
x
′′
X
∞
X
!
{}
X(j) ,
0, − k1 , · · · , − (k−1)
k
k,0
Sk,0 (n) G0,k
n=1
j=−(k−1)
(5.33)
k+1
(2π)
n
where X(j) = e
. Now we shall try to simplify the last term and to do that we
kk x
make use of Lemma 4.6, with b = − k−1
k , which gives rise to
!
k−2
k+1
{}
(−1) 2 (2π) 2
k,0
G0,k
X(j)
2k−1
0, − k1 , · · · , − (k−1)
k 2 x
k
1
k
1
1
(−1) 2 −1 2π k iπj(k−1) 1 −1
− iπj
+1 n k
2k
k
2k
k
n
(2π)
exp −e
e
.
=
k
x
x
Implement this simplification of the Meijer G-function in the last term of (5.33), and use
(2.5) and the definition (2.1) of Sk,0 (n), then (5.33) becomes
1
k
∞
X
1
1
ζ(k) (−1) 2 −1 2π k
1 1
1
− k1
+
Γ
+
ζ
x
+
=
4 k
k
k
x
k
x
enk x − 1
n=1
(k−1)
×
X
′′
∞ X
X
k−1
d
n
1
−1
k
e
iπj(k−1)
2k
n=1 dk |n
j=−(k−1)
− iπj
2k
exp −e
(2π)
1
+1
k
n1
k
x
.
(5.34)
Now writing n = dk m and upon simplification, the last term reduces to
1 (k−1)
k
∞
∞
1
1
(−1) 2 −1 2π k X ′′ X 1 −1 iπj(k−1) X
+1 m k
− iπj
k
2k
2k
k
d
e
(2π)
exp −e
m
k
x
x
m=1
d=1
j=−(k−1)
k
(−1) 2 −1
=
k
k
(−1) 2 −1
=
k
(k−1)
2π
x
1
2π
x
k
−1 "
1 X
2
k
k
′′
X
e
iπj(k−1)
2k
m=1
j=−(k−1)
e
∞
X
iπ(2j+1)(k−1)
2k
∞
X
m=1
j=0
−
+e
iπ(2j+1)(k−1)
2k
1
m k −1
iπj
1
exp e− 2k (2π) k +1
m
x
1
m k −1
1
iπ(2j+1)
1
exp e− 2k (2π) k +1
∞
X
1
m k −1
iπ(2j+1)
1
exp
e 2k (2π) k +1
m=1
k
m
x
−1
1
m
x
k
−1
1
k
−1
#
.
(5.35)
Ultimately, plugging (5.35) in (5.34), we complete the proof of Wigert’s formula (3.4) for
ζ( k1 ), k ≥ 2 even.
20
ANUSHREE GUPTA AND BIBEKANANDA MAJI
Proof of Theorem 3.4. The proof goes in accordance with the proof of Theorem 3.1, so we
mention those steps where it differs from Theorem 3.1. Note that, in this case, we are dealing
with k ≥ 2 even and r 6= −1 is any odd integer. First, we point out that, the analysis of
poles will be exactly same as in Theorem 3.1, so won’t repeat it here. The only changes will
be in the calculation of the vertical integral of Vk,r (x). One can recall that, while calculating
Vk,r (x), the equation (5.14) depends on k and r. Since k is even, the equation (5.15) will
remain same. For clarity, we mention it once again. Mainly, the equation (5.15) is
k
ksπ
πk(1 − s)
−1
2
sin
sin
= (−1)
,
(5.36)
2
2
whereas the equation (5.16) changes to
πs
r+1
π(1 − s − r)
,
sin
= (−1) 2 sin
2
2
(5.37)
as r is odd. In view of (5.36) and (5.37), the vertical integral (5.14) reduces to
k+r+1
Z
(−1) 2 −1 (2π)k+1−r 1
Γ(1 − s)Γ(1 − k + ks)Γ(s + r)ζ(1 − k + ks)
Vk,r (x) =
π2 x
2πi (1+λ)
πs (2π)k+1 −s
πks
ds.
sin
×ζ(s + r) sin
2
2
x
(5.38)
Here we remark that the only changes happened in the power of −1 and cos πs
got replaced
2
πs
πs
by sin 2 . At this point, to simplify further, we multiply and divide by cos 2 in (5.38).
Then using Euler’s reflection formula and simplifying, we arrive at
k+r+1
Z
(−1) 2 −1 (2π)k−r 1
Γ(1 − k + ks)Γ(s + r)
Vk,r (x) =
ζ(1 − k + ks)
x
2πi (1+λ)
Γ(s)
−s
sin πks
(2π)k+1
2
ds.
×ζ(s + r)
x
cos πs
2
Now we shall make use of (4.8), that is,
sin πks
k
2
= (−1) 2
πs
cos 2
(k−1)
X
j=−(k−1)
′′
j
i exp
iπjs
2
.
This is one of the crucial changes in this proof. Substituting this expression, one can show
that
2k+r+1
Z
(k−1)
Γ(1 − k + ks)Γ(s + r)
(−1) 2 −1 (2π)k−r X ′′ j 1
i
ζ(1 − k + ks)
Vk,r (x) =
x
2πi (1+λ)
Γ(s)
j=−(k−1)
×ζ(s + r)
e−
iπj
2
(2π)k+1
x
!−s
ds.
ON RAMANUJAN’S FORMULA FOR ζ(1/2) AND ζ(2m + 1)
21
From here onwards, the analysis of simplifying Vk,r (x) is exactly same as in Theorem 3.1, so
we avoid reproducing the calculations. If we continue the calculations along the same spirit
of Theorem 3.1, then the vertical integral takes the shape of
!
k+1−2r
2k+r−1
(k−1)
∞
X ′′ X
{}
(−1) 2 (2π) 2
k,0
X(j) .
Sk,r (n)G0,k
ij
Vk,r (x) =
2k−1
r, − k1 , · · · , − (k−1)
xk 2
k
n=1
j=−(k−1)
(5.39)
Now (5.39) and along with all the residual terms in Theorem 3.1, one can complete the proof
of Theorem 3.4.
5.3. A new identity for ζ(−1/2). First, let us state a lemma which will be useful in proving
the next corollary.
Lemma 5.1. Let b and u be two real numbers. Then
"
#
i exp(beiu )
1
sin(b sin u) sinh(b cos u)
Re
.
=
2
2 (cos(b sin u) cosh(b cos u) − 1)2 + (sin(b sin u) sinh(b cos u))2
(exp(beiu ) − 1)
The proof of this lemma goes along the same vein as the proof of Lemma 4.8, so we left to
readers to verify.
4
1
Proof of Corollary 3.5. It is well-known that ζ(−1) = − 12
and ζ(4) = π90 . Letting k = 2
and r = 1, Theorem 3.4 yields that
√
∞
X
1
π4
π
1
−nx
√
+
ζ −
+
D2,1 (n)e
=
24 2 x
2
90x2
n=1
"
!
!#
√ X
∞
{}
π
{}
2,0
2,0
+
S2,1 (n) iG0,2
X(−1)
. (5.40)
1 X(1) − iG0,2
2x
1,
−
1, − 12
2
n=1
Recall Lemma 4.5, with b1 = 1 and b2 = − 21 , and use [1, p. 444, Equation (10.2.17)] the fact
that
√
1
−2 z √π 1 + √
e
√
2 z
K 3 (2 z) =
,
1
2
2z 4
to deduce that
2,0
G0,2
!#
√ −2√X(j)
{}
X(j)
=
πe
1, −1
2
where
p
2 X(j) = 2π
r
1
1+ p
2 X(j)
2πn − iπj
e 4 .
x
!
,
(5.41)
22
ANUSHREE GUPTA AND BIBEKANANDA MAJI
From (2.5), we see
∞
X
−nx
D2,1 (n)e
n=1
∞
X
1 d
1
=
.
n2 dx 1 − en2 x
n=1
(5.42)
Simplifying the definition (2.1) of S2,1 (n), one can write
X d
.
S2,1 (n) =
m
2
(5.43)
n=d m
Now plugging (5.41), (5.42), and (5.43) in
∞
X
1
1
1 d
+
=
2x
2
n
n dx 1 − e
24
n=1
(5.40), we derive that
√
1
π4
π
√ ζ −
+
2 x
2
90x2
"
#
∞
∞
√
π X 1 X
1
−θd m
+
,
1+ √
Re ide
x m=1 m
θd m
(5.44)
d=1
where θ = 2π
Let us write
q
− iπ
4
2π
x e
. Next aim is to decipher the sum:
#
" ∞
X
√
√
i
−θd m
−θd m
+ √ e
Re i
de
θ m
d=1
"
#
√
i eθ m
i
.
√
= Re
√
2 + √
θ m eθ m − 1
eθ m − 1
A(θ) =
i eθ
eθ
√
√
m
i
√
.
and B(θ) = √
θ m eθ m − 1
,
m−1 2
√
− iπ
4 , where
To extract
q real part of A(θ), we need to use Lemma 5.1. We write θ m = be
π
b = 2π 2πm
x . Then, employing Lemma 5.1, with u = − 4 , we get
"
#
− iπ
i ebe 4
Re[A(θ)] = Re
2
− iπ
ebe 4 − 1
√b
√b
sin
sinh
1
2
2
=−
(5.45)
2
2 .
2
b
b
cos √2 cosh √2 − 1 + sin √b2 sinh √b2
Again, to find the real part of B(θ), we shall use Lemma 4.8, with u =
we have
#
"
iπ
ie 4
1
Re[B(θ)] = Re
− iπ
b
ebe 4 − 1
− √b
2
√b
√b
+
cos
−
e
sin
1
2 2 .
=−
4 √b cosh √b − cos √b
2
2
2
π
4
and v = 1. Thus,
(5.46)
ON RAMANUJAN’S FORMULA FOR ζ(1/2) AND ζ(2m + 1)
23
Now in view of (5.45) and (5.46), the equation (5.44) becomes
√
∞
X
1
1
1 d
π4
π
1
√
+
=
+
ζ
−
n2 dx 1 − en2 x
24 2 x
2
90x2
n=1
" b
− √b
∞
√
√b
sin
+
cos
−e 2
X
1
π
2
2
−
4x
m √b cosh √b − cos √b
m=1
2
2
2
#
2 sin √b2 sinh √b2
+
2
2
cos √b2 cosh √b2 − 1 + sin √b2 sinh √b2
3
Finally, replace x by α and β = 4πα , then one check that
variables, we complete the proof of Corollary (3.5).
√b
2
=
√
mβ. Plugging all these
Proof of Theorem 3.6. In this theorem, we are concerned with k ≥ 2 even and r = −1, so
the integrand function changes to
F (s) = Γ(s)ζ(ks)ζ(s + 1)x−s .
One can easily show that the only poles of this integrand function are at s = k1 and s = 0.
We shall note that s = 0 is a pole of order 2, where as k1 remains a simple pole. Thus, the
residue at k1 will be
1
1
1
1
+ 1 x− k .
R1/k = Γ
ζ
k
k
k
And the residue at s = 0 will be
d 2
s Γ(s)ζ(ks)ζ(s + 1)x−s .
s→0 ds
R0 = Res Γ(s)ζ(ks)ζ(s + 1)x−s = lim
s=0
From (4.2) and (4.3), we write the Laurent series expansion at s = 0 of each factor of the
integrand function:
1
π2
1
2
γ +
s + ··· ,
Γ(s) = − γ +
s
2
6
1
ζ(ks) = − + kζ ′ (0)s + · · · ,
2
1
ζ(s + 1) = + γ − γ1 s + · · · ,
s
x−s = 1 − log(x)s + · · · .
Multiplying these expansions, one can find that the coefficient of s in s2 Γ(s)ζ(ks)ζ(s + 1)x−s
is kζ ′ (0) + log(x)
2 . Thus, the residue R0 equals to
x
1
,
(5.47)
R0 = log
2
(2π)k
24
ANUSHREE GUPTA AND BIBEKANANDA MAJI
since ζ ′ (0) = − 12 log(2π). The remaining part of the proof is exactly same as in Theorem 3.4,
as we are dealing a particular case of when k is even and r is odd. Hence, considering the
above residual terms, Theorem 3.4, with r = −1, yields that
k+3
∞
X
x
1
1
1
(−1)k−1 (2π) 2
1
− k1
−nx
+1 x +
+ Γ
ζ
Dk,−1 (n)e
= log
2k−1
2
(2π)k
k
k
k
xk 2
n=1
!
(k−1)
∞
X ′′ X
{}
k,0
×
X(j) .
(5.48)
Sk,−1 (n)G0,k
ij
−1, − k1 , · · · , − (k−1)
k
n=1
j=−(k−1)
Here we invoke Lemma 4.6 with b = −1 to simplify that
{}
k,0
G0,k
−1, − k1 , · · ·
, − (k−1)
k
!
X(j)
1
exp −kX(j) k
k−1
2
(2π)
√
k
=
X(j)
.
(5.49)
Substituting (5.49) in (5.48), the right side infinite sum reduces to
1
(k−1)
∞
k
exp
−kX(j)
k−1
k+1
X
X
′′ j
(−1) (2π)
Sk,−1 (n)
i
k
xk
X(j)
n=1
j=−(k−1)
(k−1)
k−1
= (−1)
′′
X
j
i e
iπj
2
(k−1)
= (−1)
′′
X
iπj
e
(k−1)
X
j=−(k−1)
′′
∞ X
∞
X
1
m=1 d=1
j=−(k−1)
= (−1)k
n
n=1
j=−(k−1)
k−1
∞
X
Sk,−1 (n)
eiπj
d
− iπj
2k
exp −e
− iπj
2k
exp −e
∞
X
(2π)
1+ k1
(2π)
1+ k1
d
n1
k
x
m1
k
x
m 1
iπj
1
k
.
log 1 − exp −e− 2k (2π)1+ k
x
m=1
(5.50)
In the penultimate step, we have used the definition of Sk,−1 (n), and in the final step we
P
ym
used the identity − log(1 − y) = ∞
m=1 m when |y| < 1. Finally, in view (5.48) and (5.50),
we complete the proof of Theorem 3.6.
Proof of Theorem 3.7. Note that here we are dealing with k ≥ 1 and r 6= −1 are both as
an odd integer. Let us recall that the integrand function is
F (s) = Γ(s)ζ(ks)ζ(s − r)x−s .
We shall divide the analysis of poles in two cases.
Case I: First, we consider r as an odd positive integer. We know that Γ(s) has simple
poles at 0, −1, −2, −3, · · · . The poles at even negative integers are getting cancelled by the
trivial zeros of ζ(ks), whereas the poles at odd negative integers will be cancelled by ζ(s − r).
Therefore, in this case, the only poles of the integrand function are at s = 0, k1 , and 1 + r.
These are all simple poles and their corresponding residues have been already calculated in
Theorem 3.1, so we won’t repeat it here.
ON RAMANUJAN’S FORMULA FOR ζ(1/2) AND ζ(2m + 1)
25
Case II: We consider r 6= −1 as an odd negative integer. One can easily check that 0 and
are simple poles in this case too. As r ≤ −3 is odd, so 1 + r is an even negative integer.
Thus, 1 + r is a pole of order 2 of the factor Γ(s)ζ(s − r). And we also note that 1 + r is a
trivial zero of ζ(ks). Therefore, 1 + r remains as a simple pole of the integrand function.
1
k
An important point is to note that, the poles of Γ(s) at s = −1, −3, · · · , r will not get
neutralized by ζ(s−r), but the poles at the negative odd integers beyond r, say r−2, r−4, · · ·
will be neutralized by ζ(s − r) since they are trivial zeros of ζ(s − r). Therefore, in this case,
we have to take into account the residual terms coming from the contribution of the poles at
s = −1, −3, · · · , r.
To calculate the residual term corresponding to s = 1 + r, we have to use the Laurent
series expansion of Γ(s) i.e., Lemma 4.1, and the Laurent series expansion (4.2) of ζ(s) with
s replace by s − r. Thus, we will have
R1+r = lim (s − (1 + r))Γ(s)ζ(ks)ζ(s − r)x−s
s→1+r
= lim (s − (1 + r))Γ(s)
s→1+r
=
ζ(ks)
(s − (1 + r))ζ(s − r)x−s
(s − (1 + r))
(−1)1+r
kζ ′ (k(1 + r))x−(1+r) .
(−(1 + r))!
(5.51)
Let R be the sum of the residual terms corresponding to the poles at s = −1, −3, · · · , r.
Then, we have
− 1+r
2
R=
X
i=0
− 1+r
2
=
X
i=0
(−1)
=
2
Res
s=−(2i+1)
Γ(s)ζ(ks)ζ(s − r)x−s
−1
ζ(−k(2i + 1))ζ(−2i − 1 − r)x2i+1
(2i + 1)!
1+r
2
− 1+r
2
X
i=0
(−1)i+1 Bk(2i+1)+1 B−2i−1−r (2π)−r x 2i+1
.
(2i + 1)!(k(2i + 1) + 1) (−(2i + 1 + r))! 2π
(5.52)
The remaining part of the proof is in the same direction of the proof of Theorem 3.1, so we
briefly mention all the steps. Mainly, we will concentrate on the calculation of the vertical
integral Vk,r (x). From (5.14), we have
Vk,r (x) =
(2π)k+1−r 1
π2 x
2πi
Z
(1+λ)
× sin
Γ(1 − s)Γ(1 − k + ks)Γ(s + r)ζ(1 − k + ks)ζ(s + r)
πk(1 − s)
2
sin
π(1 − s − r)
2
(2π)k+1
x
−s
ds.
26
ANUSHREE GUPTA AND BIBEKANANDA MAJI
Since k and r both odd, so we must use
k−1
πk(1 − s)
kπs
2
sin
= (−1)
,
cos
2
2
πs
r+1
π(1 − s − r)
.
= (−1) 2 sin
sin
2
2
Use these two trigonometric identities to see
k+r
Z
(−1) 2 (2π)k+1−r 1
Vk,r (x) =
Γ(1 − s)Γ(1 − k + ks)Γ(s + r)ζ(1 − k + ks)ζ(s + r)
π2x
2πi (1+λ)
πs (2π)k+1 −s
kπs
ds.
sin
× cos
2
2
x
Here we multiply and divide by cos πs
and then use Euler’s reflection formula to deduce
2
that
k+r
Z
Γ(1 − k + ks)Γ(s + r)
(−1) 2 (2π)k−r 1
ζ(1 − k + ks)ζ(s + r)
Vk,r (x) =
x
2πi (1+λ)
Γ(s)
−s
cos kπs
(2π)k+1
2
×
ds.
x
cos πs
2
Next, we shall use (4.9), that is,
cos kπs
k−1
2
= (−1) 2
πs
cos 2
(k−1)
X
j=−(k−1)
′′
iπjs
i exp −
2
j
.
This is one of the important changes in this proof. Plugging this expression in the above
integral, we obtain
2k+r−1
Z
(k−1)
(−1) 2 (2π)k−r X ′′ j 1
Γ(1 − k + ks)Γ(s + r)
Vk,r (x) =
i
ζ(1 − k + ks)
x
2πi (1+λ)
Γ(s)
j=−(k−1)
×ζ(s + r)
e
iπj
2
(2π)k+1
x
!−s
ds.
Henceforth the simplification of the vertical integral Vk,r (x) is exactly same as in Theorem
3.1. If we continue the calculations in the same spirit, then the final expression of Vk,r (x)
will be
!
k+1−2r
2k+r−1
(k−1)
∞
X ′′
X
{}
(−1) 2 (2π) 2
k,0
j
Sk,r (n) G0,k
i
X(−j) .
Vk,r (x) =
2k−1
r, − k1 , · · · , − (k−1)
xk 2
k
n=1
j=−(k−1)
(5.53)
An important point is to note that the argument of the Meijer G-function is X(−j). Finally,
collecting all the residual terms (5.51), (5.52), and together with (5.53), one can complete
the proof of (3.5).
ON RAMANUJAN’S FORMULA FOR ζ(1/2) AND ζ(2m + 1)
27
Proof of Corollary 3.7. Setting r = −(2m + 1), with m ≥ 1, in Theorem 3.7, and using the
identity
ζ ′ (−2km) =
(−1)km (2km)!ζ(2km + 1)
,
2
(2π)2km
we complete the proof.
5.4. Derivation of Ramanujan’s formula for odd zeta values.
Proof of Corollary 3.9. Plugging k = 1 in Corollary 3.8, we deduce that
∞
X
n=1
x 2m
1
ζ (2m + 2) (−1)m
D1,−(2m+1) (n)e−nx = − ζ(2m + 1) +
+
ζ(2m + 1)
2
x
2
2π
m
(−1)m (2π)2m+1 X (−1)i+1 B2i+2 B2m−2i x 2i+1
+
2
(2i + 2)!(2m − 2i)! 2π
+
i=0
∞
m
2m+2
X
(−1) (2π)
x
1,0
S1,−(2m+1) (n)G0,1
n=1
!
(2π)2 n
{}
.
x
−(2m + 1)
(5.54)
From (2.2), we know
D1,−(2m+1) (n) = σ−(2m+1) (n),
and
S1,−(2m+1) (n) = σ2m+1 (n).
(5.55)
The equation (4.7) yields that
1,0
G0,1
(2π)2 n
{}
x
−(2m + 1)
!
=
x 2m+1 4π2 n
e− x .
4π 2 n
(5.56)
Now substituting (5.55) and (5.56) in (5.54), one can derive that
∞
X
x 2m
1
ζ (2m + 2) (−1)m
σ−(2m+1) (n)e−nx + ζ(2m + 1) =
+
ζ(2m + 1)
2
x
2
2π
n=1
+
m
(−1)m (2π)2m+1 X (−1)i+1 B2i+2 B2m−2i x 2i+1
2
(2i + 2)!(2m − 2i)! 2π
i=0
∞
x 2m X
σ2m+1 (n) − 4π2 n
m
+ (−1)
e x .
2π
n2m+1
n=1
Finally, invoking Euler’s formula (1.1) for ζ(2m + 2) and using the fact that
σ−(2m+1) (n) and upon simplification, one can derive (3.6).
σ2m+1 (n)
n2m+1
=
28
ANUSHREE GUPTA AND BIBEKANANDA MAJI
Proof of Corollary 3.10. Letting k = 1 and r = 2m + 1 with m ≥ 1 in Theorem 3.7, and
2m+2
with the fact that ζ(−(2m + 2)) = 0 and ζ(−(2m + 1)) = − B2m+2
, we find that
∞
X
D1,2m+1 (n)e−nx =
n=1
ζ(2m + 2)
B2m+2
+ (2m + 1)!
4m + 4
x2m+2
∞
x −(2m+1) 4π2 n
X
1
S
(n)
e− x
1,2m+1
x(2π)2m n=1
4π 2 n
2m+2
∞
X
2π
B2m+2
B2m+2
+ (−1)m
σ2m+1 (n)e−nx =
⇒
4m
+
4
x
4m
+4
n=1
2m+2 X
∞
4π 2 n
m+1 2π
+ (−1)
(5.57)
σ2m+1 (n)e− x .
x
n=1
+ (−1)m+1
Here we have used (2.2) to write D1,2m+1 (n) and S1,2m+1 (n)n2m+1 in terms of σ2m+1 (n).
Finally, to obtain (3.7), replace x = 2α and αβ = π 2 in (5.57).
Substituting k = r = 1 in Theorem 3.7 and setting x = 2α and αβ = π 2 , one can obtain
the following identity of Schlömilch [29], and rediscovered by Ramanujan [26, Ch. 14, Sec.
8, Cor. (i)], [27, p. 318].
Corollary 5.2. For α, β > 0 with αβ = π 2 ,
∞
X
∞
X
n
n
α+β
1
α
+β
=
−
2nα − 1
2nβ − 1
e
e
24
4α
n=1
n=1
In particular, if we consider α = β = π, then
∞
X
n=1
n
1
1
=
−
.
−1
24 8π
e2nπ
Proof of Theorem 3.11. The proof of this theorem travels in the same direction as of Theorem 3.7 since we are dealing k odd and r = −1. In this case, the integrand function is
F (s) = Γ(s)ζ(ks)ζ(s + 1)x−s .
One can easily check that the integrand function has poles are at s = 0, k1 and −1. At s = 0,
we have a pole order 2, but the other two points are simple pole. In view of (5.47), we know
x
1
R0 = log
.
2
(2π)k
One can find that
R−1 =
ζ(−k)x
.
2
ON RAMANUJAN’S FORMULA FOR ζ(1/2) AND ζ(2m + 1)
29
Now along with the residue at k1 , Theorem 3.7, with r = −1, leads to
k+3
∞
X
x
1
ζ(−k) (2π) 2
1
1
1
− k1
−nx
+1 x +
+
Dk,−1 (n)e
= log
+ Γ
ζ
2k−1
2
(2π)k
k
k
k
2
xk 2
n=1
!
(k−1)
∞
X ′′ X
{}
k,0
×
Sk,−1 (n)G0,k
ij
(5.58)
(k−1) X(−j) .
1
−1,
−
,
·
·
·
,
−
k
k
n=1
j=−(k−1)
Calculations of the above right hand side sum goes along the same line of (5.48). For completeness, we mention it briefly. First, employ Lemma 4.6 with b = −1 to see
1
!
k−1
k
exp
−kX(−j)
{}
(2π) 2
k,0
√
G0,k
.
(k−1) X(−j) =
1
X(−j)
−1, − k , · · · , − k
k
Substituting it in (5.58), together with definition (3.1) of X(j), the right side infinite sum
takes the shape of
(k−1)
∞
n1
X
X ′′
iπj
1
Sk,−1 (n)
k
1+
j − iπj
exp −e 2k (2π) k
i e 2
n
x
n=1
j=−(k−1)
(k−1)
′′
X
=
j=−(k−1)
∞ X
∞
X
1
m=1 d=1
(k−1)
=−
X
j=−(k−1)
′′
∞
X
m=1
d
exp −e
iπj
2k
1+ k1
(2π)
log 1 − exp −e
iπj
2k
d
m1
k
x
1+ k1
(2π)
m 1
k
x
.
Now substituting (5.59) in (5.58), one can finish the proof of (3.8).
(5.59)
Proof of Corollary 3.12. Plugging k = 1 in (3.8), we arrive at
∞
∞
x π2
X
X
4π 2 m
x
1
−nx
+
log 1 − exp −
−
−
D1,−1 (n)e
= log
2
2π
6x 24
x
m=1
n=1
⇒
∞ X
X
n=1 d|n
∞
X
1 −nx
e
d
∞ X
∞
x π2
X
1
x
1 − 4π2 md
x
+
= log
−
+
e
2
2π
6x 24
d
m=1 d=1
∞
x π2
X
1
1
1
x
4π2 n
.
⇒
+
=
log
−
+
nx
n(e
−
1)
2
2π
6x
24
x
n
e
−
1
n=1
n=1
Finally, replacing x = 2α and β = π 2 /α and simplifying we obtain (3.9).
6. Concluding Remarks
On page 332, Ramanujan recorded an intriguing formula for ζ(1/2) and also mentioned
the following interesting infinite series
X
nr
,
N x) − 1
exp(n
n=1
30
ANUSHREE GUPTA AND BIBEKANANDA MAJI
where N ∈ N, r ∈ Z and N − r is any even integer. This infinite series can be written as
Z c+i∞
1
Γ(s)ζ(s)ζ(N s − r)x−s ds,
2πi c−i∞
for c > max 1, 1+r
N . Recently, Dixit and the second author obtained a beautiful generalization of Ramanujan’s formula for odd zeta values while studying the above integral.
In the current paper, we have studied a variant of the above integral,
Z c+i∞
∞
X
1
−nx
Γ(s)ζ(ks)ζ(s − r)x−s ds,
Dk,r (n)e
=
2πi c−i∞
n=1
where for k ∈ N, r ∈ Z, and c > max (1/k, 1 + r). The analysis of this integral can be divided
into the following four cases.
Case 1: When k and r are both are even integers, Theorem 3.1 gives us a transformation
formula, which allows us to derive Ramanujan’s formula for ζ(1/2) and Wigert’s formula for
ζ(1/k).
Case 2: When k is even and r is odd, i.e., k − r is an odd integer. In this case, Theorem
3.4 allows us to obtain a new identity for ζ(−1/2) analogous to Ramanujan’s formula for
ζ(1/2).
Case 3: When k and r are both odd, i.e., k − r is an even integer. This is the most
interesting case. Theorem 3.7 allows us to derive Ramanujan’s formula for odd zeta values.
Corollary 3.8 can be conceived as a one-variable generalization of Ramanujan’s formula for
odd zeta values, which connects two odd zeta values ζ(2m + 1), and ζ(2km + 1) and a zeta
value at a rational argument ζ k1 + 2m + 1 .
Case 4: When k is odd and r is even, i.e., k − r is an odd integer. In this case, the
integrand function will have infinitely many poles, namely, at all odd negative integers. In
the current paper, we have not discussed this case. It would be interesting to study this case
too.
Ramanujan’s formula (1.2) for ζ(2m + 1) has been generalized by many mathematicians
in various directions. Ramanujan himself gave a huge generalization [10, Entry 20, p. 430].
Berndt [7] derived Euler’s formula for ζ(2m) as well as Ramanujan’s formula for ζ(2m + 1)
from a single transformation formula for a generalized Eisenstein series. The Dirichlet Lfunction L(s, χ) and the Hurwitz zeta function ζ(s, a) are natural generalizations of the
Riemann zeta function. A character analogue of Ramanujan’s formula was obtained by
Merrill [22] and more generally, for any periodic function was obtained by Bradley [14].
Dixit, Kumar, Gupta and the second author [17] found a generalization for Hurwitz zeta
function, which allowed them to connect many odd zeta values from a single formula. Dixit
and Gupta [16] established an interesting analogue of Ramanujan’s formula which connects
square of odd zeta values. Very recently, Banerjee and Kumar [6] found an analogue of
Ramanujan’s formula (1.2) over an imaginary quadratic fields. For more references on the
generalizations of Ramanujan’s formula, readers are encouraged to see [13].
ON RAMANUJAN’S FORMULA FOR ζ(1/2) AND ζ(2m + 1)
31
Acknowledgements. The authors want to thank IIT Indore for the conductive research
environment. The second author is grateful to SERB for the Start-Up Research Grant
SRG/2020/000144.
References
[1] M. Abramowitz and I. Stegun, Handbook of Mathematical Functions with Formulas, Graphs, and Mathematical Tables. Number 55 in National Bureau of Standards Applied Mathematics Series. U.S. Government Printing Office, Washington, D.C. 1965.
[2] R. Apéry, Irrationalité de ζ(2) et ζ(3), Astéerisque 61 (1979), 11–13.
[3] R. Apéry, Interpolation de fractions continues et irrationalité de certaines constantes, Bull. Section des
Sci., Tome III, Bibliothéque Nationale, Paris, 1981, 37—63.
[4] G.E. Andrews and B.C. Berndt. Ramanujan’s lost notebook, Part IV (New York: Springer- Verlag, 2013).
[5] K. Ball and T. Rivoal, Irrationalité d’une infinité de valeurs de la fonction zêta aux entiers impairs
(French), Invent. Math. 146 no. 1 (2001), 193–207.
[6] S. Banerjee, R. Kumar, Explicit identities on zeta values over imaginary quadratic field, submitted for
publication, arXiv:2105.04141, 2021.
[7] B. C. Berndt, M odular transformations and generalizations of several formulae of Ramanujan, Rocky
Mountain J. Math. 7 (1977), 147–189.
[8] B. C. Berndt, Ramanujan’s Notebooks, Part II, Springer-Verlag, New York, 1989.
[9] B. C. Berndt, Ramanujan’s Notebooks, Part III, Springer-Verlag, New York, 1991.
[10] B. C. Berndt, Ramanujan’s Notebooks, Part IV, Springer-Verlag, New York, 1994.
[11] B. C. Berndt, A. Dixit, A. Roy, and A. Zaharescu, New pathways and connections in number theory and
analysis motivated by two incorrect claims of Ramanujan, Adv. Math. 304 (2017), 807–929.
[12] B. C. Berndt and A. Straub, On a secant Dirichlet series and Eichler integrals of Eisenstein series, Math.
Z. 284 (2016), 827–852.
[13] B. C. Berndt and A. Straub, Ramanujan’s formula for ζ(2n + 1), Exploring the Riemann zeta function,
Eds. H. Montgomery, A. Nikeghbali, and M. Rassias, pp. 13–34, Springer, 2017.
[14] D. M. Bradley, S eries acceleration formulas for Dirichlet series with periodic coefficients, Ramanujan J.
6 (2002), 331–346.
[15] A. Dixit and B. Maji, Generalized Lambert series and arithmetic nature of odd zeta values, Proc. Roy.
Soc. Edinburg Sect. A: Mathematics 150 (2020), 741–769.
[16] A. Dixit and R. Gupta, On squares of odd zeta values and analogues of Eisenstein series, Adv. Appl.
Math. 110 (2019), 86–119.
[17] A. Dixit, R. Gupta, R. Kumar, and B. Maji, Generalized Lambert series, Raabe’s cosine transform and a
generalization of Ramanujan’s formula for ζ(2m + 1), Nagoya Math. J. 239 (2020), 232–293.
[18] S. Gun, M. R. Murty, and P. Rath, Transcendental values of certain Eichler integrals, Bull. London Math.
Soc. 43 (2011), 939–952.
[19] H. Iwaniec, E. Kowalski, Analytic number theory, American Mathematical Society Colloquium Publications, vol. 53, 2004.
[20] S. Kanemitsu, Y. Tanigawa, and M. Yoshimoto, On the values of the Riemann zeta-function at rational
arguments. Hardy-Ramanujan J. 24 (2001), 11–19.
[21] S. L. Malurkar, On the application of Herr Mellin’s integrals to some series, J. Indian Math. Soc. 16
(1925/26), 130–138.
[22] K. J. Merrill, Ramanujan’s formula for the Riemann zeta function extended to L-functions, 2005.
[23] F. Oberhettinger, Tables of Mellin Transforms, Springer-Verlag, New York-Heidelberg, 1974.
32
ANUSHREE GUPTA AND BIBEKANANDA MAJI
[24] F. W. J. Olver, D. W. Lozier, R. F. Boisvert, C. W. Clark, eds., NIST Handbook of Mathematical
Functions, Cambridge University Press, Cambridge, 2010.
[25] B. Riemann, U eber die Anzahl der Primzahlen unter einer gegebenen Grösse, Monatsberichte der Berliner
Akademie, 1859.
[26] S. Ramanujan, Notebooks (2 volumes), Tata Institute of Fundamental Research, Bombay, 1957; second
ed., 2012.
[27] S. Ramanujan, The Lost Notebook and Other Unpublished Papers, Narosa, New Delhi, 1988.
[28] T. Rivoal, La fonction zêta de Riemann prend une infinité de valeurs irrationnelles aux entiers impairs,
C. R. Acad. Sci. Paris Sér. I Math. 331 no. 4 (2000), 267–270.
[29] O. Schlömilch, Ueber einige unendliche Reihen, Berichte über die Verh. d. Könige Sächsischen Gesell.
Wiss. zu Leipzig 29 (1877), 101–105.
[30] E. C. Titchmarsh, The Theory of the Riemann Zeta-function, Clarendon Press, Oxford, 1986.
[31] S. Wigert, Sur une extension de la série de Lambert, Arkiv Mat. Astron. Fys. 19 (1925), 13 pp.
[32] W. W. Zudilin, One of the numbers ζ(5), ζ(7), ζ(9) and ζ(11) is irrational (Russian), Uspekhi Mat.
Nauk 56 No. 4 (2001), 149–150; translation in Russian Math. Surveys 56 No. 4 (2001), 774–776.
Anushree Gupta, Department of Mathematics, Indian Institute of Technology Indore, Simrol, Indore, Madhya Pradesh 453552, India.
Email address:
[email protected],
[email protected]
Bibekananda Maji, Department of Mathematics, Indian Institute of Technology Indore, Simrol, Indore, Madhya Pradesh 453552, India.
Email address:
[email protected],
[email protected]