Skip to main content

A silly little derivation of \( \zeta(2) \)

(This is a cleaned-up and somewhat expanded version of this Twitter thread.)

What follows is a silly little proof that

\[ \zeta(2) = \sum_{n=1}^{\infty} \frac{1}{n^2} = \frac{\pi^2}{6} \]

where \( \zeta \) is the Riemann zeta function.

Consider the integral

\[ I := \int_0^1 \frac{\log(1 - x + x^2)}{x(x - 1)} \, dx. \]

We have, by using partial fractions and performing some other algebraic manipulations,

\[ \begin{align*} I &=  -\int_0^1 \! \frac{\log(1 - x + x^2)}{x} \, dx - \int_0^1 \! \frac{\log(1 - x + x^2)}{1 - x} \, dx  \\
&= -2\int_0^1 \! \frac{\log(1 - x + x^2)}{x} & (x \mapsto 1 - x ) \\
&= 2\left( \int_0^1 \! \frac{\log(1 + x)}{x} \, dx - \int_0^1 \! \frac{\log(1 + x^3)}{x} \, dx \right) \\
&= \frac{4}{3}\int_0^1 \frac{\log(1 + x)}{x} \, dx & (x \mapsto x^{1/3}). \end{align*} \]

To evaluate this integral, we take the Maclaurin series:

\[ \int_0^1 \! \frac{\log(1 + x)}{x} \, dx = \int_0^1 \! \sum_{n=1}^{\infty} \frac{(-1)^nx^{n-1}}{n} \, dx \]

Since for all positive integers \( N \),

\[ \left|\sum_{n=1}^N \frac{(-1)^nx^{n-1}}{n}\right| \le \sum_{n=1}^{N} \frac{x^{n-1}}{n} \le \sum_{n=1}^{\infty} \frac{x^{n-1}}{n} = -\frac{\log(1 - x)}{x} \]

on \( [0, 1) \) and

\[ -\int_0^1 \! \frac{\log(1 - x)}{x} \, dx < -\int_0^{1/2} \! \frac{\log(1 - x)}{x} \, dx - 2\int_{1/2}^1 \! \log(1-x) \, dx < \infty, \]

we can invoke the dominated convergence theorem to switch summation and limit processes. We then have

\[ \begin{align*} \int_0^1 \! \frac{\log(1 + x)}{x} \, dx &= \sum_{n=1}^{\infty} \int_0^1 \frac{(-x)^{n-1}}{n} \, dx \\
&= \sum_{n=1}^{\infty} \frac{(-1)^{n-1}}{n^2} = \eta(2) \end{align*} \]

where \( \eta(s) \) is the Dirichlet eta function.

Now, we have \( \eta(s) = 2^{1-s}\zeta(s) \); for \(s > 1 \), this can be seen by separating and rearranging all even terms in the summation. Thus, we have

\[ I = \frac{4}{3}\eta(2) = \frac{2}{3}\zeta(2). \]

We look at another way to evaluate \( I \). Noticing that \( 1 - x + x^2 = 1 - x(1 - x) \), we can write the integrand as a power series in \( x (1 - x) \):

\begin{align*} I &= -\int_0^1 \! \frac{\log(1 - x(1 - x))}{x(1 - x)} \, dx \\
&= \int_0^1 \sum_{n=1}^{\infty} \frac{x^{n-1}(1 - x)^{n-1}}{n} \, dx \\
&= \sum_{n=1}^{\infty} \int_0^1 \! \frac{x^{n-1}(1 - x)^{n-1}}{n}
\end{align*}

where we justify the switch of integral and sum by the monotone convergence theorem (since the summand is nonnegative on \([0, 1]\)). Recall the Euler Beta function given by

\[ \mathrm{B}(m,n) = \int_0^1 \! x^{m-1}(1 - x)^{n-1} \, dx = \frac{\Gamma(m)\Gamma(n)}{\Gamma(m+n)}. \]

In the case that \( m \) and \( n \) are integers, we get

\[ \mathrm{B}(m,n) = \frac{(m - 1)!(n - 1)!}{(m + n - 1)!} \]

and in particular

\[ \int_0^1 \! \frac{x^{n-1}(1 - x)^{n-1}}{n} \, dx = \frac{\mathrm{B}(n,n)}{n} = \frac{((n - 1)!)^2}{n(2n - 1)!} = \frac{2}{n^2\binom{2n}{n}}. \]

Thus,

\[ I = 2\sum_{n=1}^{\infty} \frac{1}{n^2\binom{2n}{n}}. \]

The next part may strike you as something I pulled out of nowhere. We invoke the identity

\[ \arcsin^2{x} = \frac{1}{2}\sum_{n=0}^{\infty} \frac{(2x)^{2n}}{n^2\binom{2n}{n}}, \qquad x \in [-1, 1] , \]

a nice proof of which can be found by following the links starting here. (Okay, I'll admit it's not a very commonly-taught series, and the only reason I recognized it is that I used to spend too much time on AoPS. As such, I feel bad about blackboxing it like this. But it's cute!) We deduce that

\[ I = 4\arcsin^2\left(\frac{1}{2}\right) = \frac{\pi^2}{9}. \]

Thus, \( (2/3)\zeta(2) = \pi^2/9 \), and so \( \zeta(2) = \pi^2/6 \). QED.

Comments

Popular posts from this blog

On infinite decimal expansions, missing numbers, and generating functions

(This post is a cleaned up and expanded version of this thread .) A cool fact I've seen shared around the internet   a few times : The decimal expansion of \( 1/998001 \) starts with \[ \frac{1}{998001} = 0.000001002003\dots 996997999\dots \] That is, it begins with three-digit strings from \( 000 \) to \( 999 \), in order, except that it skips \(998\) for some reason. The first thing to observe is that \( 998001 = 999^2 \). Recall the formula for the infinite geometric series: \[ \sum_{n=0}^{\infty} r^n = \frac{1}{1 - r}. \] If we differentiate both sides with respect to \( r\), we get \[ \sum_{n=1}^{\infty} nr^{n-1} = \frac{1}{(1 - r)^2}, \] and multiplying by \( r \) gives \[ \sum_{n=1}^{\infty} nr^n = \frac{r}{(1 - r)^2}. \] (This can also be obtained by some series manipulations.) Now, take \( r = 0.001 \). We have \[ \frac{0.001}{(1 - 0.001)^2} = \frac{1000}{998001} = 0.001 + 0.000002 + 0.000000003 + \dots \] From here, the appearance of the numbers from \( 001 \) to \( 997

On My Favorite Number, 76923 (A Brief Survey of Cyclic Numbers)

(This is a cleaned-up, somewhat revised/expanded version of my Twitter thread here .) Among math enthusiasts, the number \( 142857 \) is pretty cool. Move its leftmost digit to the right, and you get \( 428571 \), which is three times the original: \( 428571 = 142857 \times 3 \). Do this again, and you get \( 285714 \), which is two times the original: \( 285714 = 142857 \times 2 \). We can keep doing this until we return to \( 142857 \), as follows: \[ \begin{align*} 142857 &= 142857 \times 1 & 142857 \times 1 &= \color{red} 142857 \\ 428571 &= 142857 \times 3 & 142857 \times 2 &= 2857\color{red}14 \\ 285714 &= 142857 \times 2 & 142857 \times 3 &= 42857\color{red}1 \\ 857142 &= 142857 \times 6 & 142857 \times 4 &= 57\color{red}1428 \\ 571428 &= 142857 \times 4 & 142857 \times 5 &= 7\color{red}14285 \\ 714285 &= 142857 \times 5 & 142857 \times 6 &= 857\color{red}142 \end{align*} \] Numbers that give you consecutive