You are on page 1of 6

Volume 00, Number 0, Pages 000–000

S (XX)0000-0

FOURIER ANALYSIS AND THE ZETA FUNCTION AT EVEN


POSITIVE INTEGERS

JIBRAN IQBAL SHAH

Abstract. Using a summation identity obtained for the Fourier coefficients


of x2k for k ∈ Z+ , this paper derives the closed form expression for the Zeta
function at even positive integers by deriving a recurrence relation for the
Bernoulli Numbers and comparing it to the recurrence relation derived for the
Zeta function at even positive integers with the help of Fourier Series. This
paper uses the idea of Fourier Series in a way similar to one in an existing paper
by Aladdi and Defant, but avoids the use Parseval’s Theorem as it considers
the 2π periodic Fourier Series representation of x2k in the interval [−π, π]
whereas the paper by Aladdi and Defant uses the 2π periodic Fourier Series
representation of xk in the interval [−π, π]. This simplifies the derivation
immensely and hence provides a shorter, more readable, and more elegant
proof.

1. Introduction
The Basel problem, more specifically the problem of evaluating the sum

X 1
n=1
n2
was a very famous problem back in the day. The problem was popularized by the
Bernoullis who lived in Basel, Switzerland, so it became known as the Basel Prob-
lem. Leonhard Euler solved it in 1735 and along with solving the Basel Problem,
he also found a closed-form evaluation of

X 1
ζ(2k) =
n=1
n2k
for all even integers 2k ≥ 2. The value of ζ(2k) is a rational multiple of π 2k . Euler
showed that
(−1)k+1 B2k (2π)2k
ζ(2k) =
2(2k)!
Where Bk are the Bernoulli Numbers, note that these are defined as [7]
  ∞ − m
t t t X Bm t
t
= coth − 1 =
e −1 2 2 m=0
m!
  ∞ + m
t t t X Bm t
−t
= coth + 1 =
1−e 2 2 m=0
m!

2010 Mathematics Subject Classification. 11R42 (Primary), 11B68 (Secondary).


Key words and phrases. Zeta Function, Number Theory, Analytic Number Theory, Bernoulli
Numbers.
1
2 JIBRAN IQBAL SHAH

+
The only difference between Bm and Bm −
being that B1+ = 21 and B1− = − 12 . There
are many ways to prove Euler’s closed form for ζ(2k), some simpler than others but
they appear in [3] and [4].

In this paper, we will, using the 2π periodic Fourier Series representation of


x2k in the interval [−π, π], derive a recurrence relation for ζ(2k) and obtain the
closed form expression Euler obtained with the help of Bernoulli numbers and a
new identity with them. This proof, although uses the same core ideas, takes a
different route than that of Alladi and Defant[1] as it avoids the use of Parseval’s
theorem, and hence ends up being shorter, easier to understand and more elegant.
The recurrence for ζ(2k) in this paper relies on ζ(2j) for j < k, but a paper by
Kuo[5] provides a recurrence for ζ(2k) for j ≤ n2 . The novelty in this paper is the
use of the Fourier Coefficients of x2k rather than xk and a new Bernoulli Identity
which we prove and use to derive our final result. An example of the case for k = 1
is given below.
Consider the Fourier Series representation of f (x) = x2 for x ∈ [−π, π]

X
f (x) = a0 + an cos(nx) + bn sin(nx)
n=1
π
π2 1 π 2 4(−1)n
Z Z
1 2
a0 = x dx = , an = x cos(nx) dx = , bn = 0
2π −π 3 π −π n2

π 2 X 4(−1)n
f (x) = + cos(nx)
3 n=1
n2
At x = π, we obtain

π2 X 1 π2
π2 = +4 ⇒ ζ(2) =
3 n=1
n2 6

2. The proof of the generalized sum ζ(2k)


Lemma 2.1.
∞ π
kπ 2k+1
X Z
(−1)n I(n, k) = where I(n, k) = x2k cos(nx) dx
n=1
2k + 1 0

Proof. Consider the Fourier series representation of f (x) = x2k for k ∈ N in the
interval [−π, π] and calculate a0 , an and bn
Z π
π 2k 1 π 2k
Z
1 2k 2
a0 = x = , an = x cos(nx) dx = I(n, k)
2π −π 2k + 1 π −π π
as x2k cos(nx) is an even function.
1 π 2k
Z
bn = x sin(nx) dx = 0
π −π
as x2k sin(nx) is an odd function. Substituting x = π yields
∞ ∞
π 2k 2X X kπ 2k+1
π 2k = + (−1)n I(n, k) ⇒ (−1)n I(n, k) = .
2k + 1 π n=1 n=1
2k + 1
FOURIER ANALYSIS AND THE ZETA FUNCTION AT EVEN POSITIVE INTEGERS 3


Lemma 2.2.
k π
(−1)n (−1)i−1 (2k)! 2k−(2i−1)
X Z
I(n, k) = π where I(n, k) = x2k cos(nx) dx
i=1
n2i [2k − (2i − 1)]! 0

Proof. Z π
I(n, k) = x2k cos(nx) dx
0
Carrying out integration by parts repeatedly, one obtains

(2k)!x2k−1 cos(nx) (2k)!x2k−3 cos(nx) (2k)!x cos(nx)
I(n, k) = 2
− 4
+ · · · (−1)k−1
(2k − 1)!n (2k − 3)!n n2k 0
Substituting the limits of integration, we obtain
k
X (−1)n (−1)i−1 (2k)! 2k−(2i−1)
I(n, k) = π .
i=1
n2i [2k − (2i − 1)]!

Lemma 2.3.
k
X (−1)i−1 (2k)! kπ 2k
ζ(2i)π 2k−2i =
i=1
[2k − (2i − 1)]! 2k + 1

Proof. Substituting the expression obtained for I(n, k) in Lemma 2.2.into Lemma
2.1. (and noting the (−1)n in Lemma 2.1. canceling out the (−1)n in Lemma 2)
we obtain
∞ X k
X (2k)!π 2k−(2i−1) (−1)i−1 kπ 2k+1
2i
=
n=1 i=1
n [2k − (2i − 1)]! 2k + 1
Taking n2i out of the first summation as it is independent of k, we obtain
k
X (−1)i−1 (2k)!π 2k−2i+1 kπ 2k+1
ζ(2i) =
i=1
[2k − (2i − 1)]! 2k + 1
This is the summation identity mentioned in the abstract. It can be rearranged to
provide a recurrence relation for ζ(2k) in terms of ζ(2j) where j < n.

Dividing both sides of the equation by π proves the lemma


k
X (−1)i−1 (2k)! kπ 2k
ζ(2i)π 2k−2i = .
i=1
[2k − (2i − 1)]! 2k + 1

Notice that the expression on the right hand side is a single fraction containing
a power of π. This means the terms in the summation all must be multiples of the
same power of π. If that is the case, then each term in the summation must contain
π 2k and hence ζ(2i) = Ci π 2i where Ci are coefficients to be determined.
Lemma 2.4.
k  
X
2i 2k 1 2k
2 B2i =
i=1
2i 2k − 2i + 1 2k + 1
4 JIBRAN IQBAL SHAH

Proof. First we prove a similar identity, specifically


m  
i + m 1 2m + 1
X
(2.1) 2 Bi =
i=1
i m−i+1 m+1
Using the property of Bernoulli Polynomials, which are defined as
n  
X n
Bn (x) = Bk− xn−k
k
k=0
More on the definition of the Bernoulli Polynomials and its generalization is present
in [6]. We then have
m m  Z 1 Z 21
2i
 
− m − m
X X 2
m+1 m−i m+1
Bi =2 Bi x dx = 2 Bm (x)dx
i=0
k m−i+1 i=0
i 0 0

using the substitution,


y =1−x dy = −dx
1
Z 2
Z 1 Z 1 Z 1
Bm (x)dx = Bm (1 − x)dx = (−1)m Bm (x)dx = Bm (x)dx
1 1 1
0 2 2 2

(−1)m = 1 as m is even and hence using a well known property of Bernoulli


Polynomials [2]
Z 21
1 1
Z
Bm (x)dx = Bm (x)dx = 0
0 2 0
The only difference between B1+ and B1− is that B1+ = 12 and B1− = − 21 Plugging
in i = 0, subtracting the term with i = 1 for Bi− and adding the term for Bi+ , and
rearranging, we get
m  
X m 1 2m + 1
2i Bi+ =
i=1
i m − i + 1 m+1
when m is a positive even integer. This can be rewritten as
2k  
X 2k + 1
(2.2) 2i Bi+ = 4k + 1
i=1
i
If m is an even positive integer, then m = 2k for some k ∈ N, this can be
rewritten as
2k  
X 2k 1 2k
2i Bi+ = +1
i=1
i 2k − i + 1 2k + 1
4k+1 2k
Notice how 2k+1 = 2k+1 + 1.

If Eq. (1) is true then as Bk = 0 when k is odd, we have


2k   k  
i + 2k 1 2k 1
X X
2i
2 Bi −1= 2 B2i
i=1
i 2k − i + 1 i=1
2i 2k − 2i + 1
The −1 is present because at i = 1, the expression evaluates to 1 (B1 = 12 ). This
completes the proof of the identity
k  
X
2i 2k 1 2k
2 B2i = .
i=1
2i 2k − 2i + 1 2k +1
FOURIER ANALYSIS AND THE ZETA FUNCTION AT EVEN POSITIVE INTEGERS 5

Lemma 2.5.
k
X (−1)i−1 (2k)! k (−1)i−1 22i B2i
If Ci = then Ci =
i=1
(2k − 2i + 1)! 2k + 1 2(2i)!

Proof. Begin with Lemma 2.4. which states


k  
X 2k 1 2k
22i B2i =
i=1
2i 2k − 2i + 1 2k + 1

Substituting in the identity,


 
2k 1 (2k)!
=
2i 2k − 2i + 1 (2i)!(2k − 2i + 1)!
We obtain
k
X (2k)! k
22i−1 B2i =
i=1
(2i)!(2k − 2i + 1)! 2k + 1
k
As both expressions sum up to 2k+1 , we can equate them to give
k k
X (2k)! X (−1)i−1 (2k)!
22i−1 B2i = Ci
i=1
(2i)!(2k − 2i + 1)! i=1
(2k − 2i + 1)!

Comparing both sides, we obtain


(−1)i+1 22i B2i
Ci = .
2(2i)!


Theorem 2.6. For any even positive integer i


(2π)2i B2i
ζ(2i) = (−1)i+1
2(2i)!
Proof. Substituting ζ(2i) = Ci π 2i into Lemma 2.3. we get
k
X (−1)i−1 (2k)! kπ 2k
Ci π 2k =
i=1
[2k − (2i − 1)]! 2k + 1

Dividing both sides by π 2k we obtain :


k
X (−1)i−1 (2k)! k
Ci =
i=1
(2k − 2i + 1)! 2k + 1

As ζ(2i) = Ci π 2i , according to Lemma 2.5., we obtain


(−1)i+1 (2π)2i B2i
ζ(2i) = .
2(2i)!

6 JIBRAN IQBAL SHAH

3. Remarks
In this paper we obtained and used the new identity
k  
X
2i 2k 1 2k
2 B2i =
i=1
2i 2k − 2i + 1 2k + 1
In the proof of Lemma 2.4., we have Eq. (2) which stated that
2k  
i + 2k + 1
X
2 Bi = 4k + 1
i=1
i
Readers can attempt to generalize the sum by considering finding a closed form for
2kx  
X 2kx + 1
xi Bi .
i=1
i

4. Acknowledgements
The author would like to thank Colin Defant, Sai Teja Somu, Juan Luis Varona
and Fabio M. S. Lima for spotting typographical errors and providing him with
their insightful comments and suggestions on the paper. The author is still a high
school student and needed guidance.

References
[1] K. Alladi anc C. Defant. (2017). Revisiting The Riemann Zeta Function at Positive Even
Integers. Int. J. Number Theory 14 (2018), 1849-1856. Link
[2] Abramowitz M. & Stegun I. A. (Eds.). ”Bernoulli and Euler Polynomials and the Euler-
Maclaurin Formula.” §23.1 in Handbook of Mathematical Functions with Formulas, Graphs, and
Mathematical Tables, 9th printing. New York: Dover, pp. 804-806, 1972.
[3] Ó. Ciaurri, L. M. Navas, F. J. Ruiz and J. L. Varona, (2012). A simple computation of ζ(2k),
Amer. Math. Monthly 122 (2015), 444-451.
[4] E. de Amo, M. Dı́az Carrillo and J. Fernández-Sánchez (2012). Another proof of Euler’s formula
for ζ(2k). Proc. Am. Math. 139. 1441-1444.
[5] H. Kuo, A recurrence formula for ζ(2n), Bull. Amer. Math. Soc. 55 (1949) 573–574.
[6] P. Natalini and A. Bernardini (2003). A generalization of the Bernoulli polynomials. J Appl
Math. 3. Link
[7] Weisstein, Eric W. (4 January 2016), ”Bernoulli Number”, MathWorld, Wolfram. Link

Student, Arrowad International School, Riyadh, Saudi Arabia


Email address: jibraniqbal2003@gmail.com

You might also like