In the paper of Riemann and the book of Edwards I encountered the following representation for $\zeta$:
$$
\int_0^\infty\frac{x^{s-1}}{e^x-1}\,dx=\Pi(s-1)\zeta(s)
$$
which due to the fast growth of $e^x$ converges at infinity, for $x$ tending to $0$ the integrand grows like $x^{s-2}$, and therefore due to this pole the integral converges for $Re(s)>1$. Then he considers the following integral,taken over the Hankel contour from infinity to along the real axis anti-clockwise around the pole and back to infinity.
\begin{align*}
& \int_\infty^\infty\frac{(-x)^{s-1}}{e^x-1} \, dx \\[8pt]
= {} & (e^{i\pi s}-e^{-i\pi s}) \int_0^\infty\frac{x^{s-1}}{e^x-1}\,dx \\[8pt]
= {} & 2i\sin(\pi s) \Pi(s-1)\zeta(s)
\end{align*}
It is claimed that the left integral converges for all $s.$ However, I was unable to see this convergence. I suspect by going around $z=0$ we omitted the pole, and therefore we only have to look at the convergence at infinity. But around $z=0$, a limiting process still takes place.
How can we rigorously prove the convergence of the left integral?
- 68
-
What do you mean with 'around 0' ? – reuns Feb 20 '21 at 23:36
-
2Maybe you mean when we shrink the contour to $(+\infty,0]\cup [0,e^{-2i\pi} \infty)$ ? We do this only for $\Re(s)>1$, for $\Re(s)\le 1$ we don't shrink the contour. – reuns Feb 20 '21 at 23:57
-
This works for any $\Re(s)>1$ so you can represent the integral as $\Gamma(s)\zeta(s)$ – Henry Lee Feb 21 '21 at 19:52
-
You might also be interested in this: https://math.stackexchange.com/questions/2226247/a-integral-involving-riemann-zeta-function-and-gamma-function-int-0-infty – Henry Lee Feb 21 '21 at 19:54
-
@reuns by around 0 I meant the piece of the contour of $|z|=\delta$ – Libertas Feb 27 '21 at 19:20
1 Answers
I wasn't clear exactly where you were stuck in the proof, so I posted a (more-or-less) full proof.
We assume one easily proved lemma:
lemma 1: For $|z| < 1/2$, $\left|e^z -1\right| \geq |z|/2$.
Next, we define a version of the Hankel contour. For small positive $\delta$ select two points in the complex plane: \begin{align*} p_1&=\delta e^{i\delta} = \delta_x + i\delta_y\quad\text{where}\quad \delta_x = \delta\cos(\delta), \delta_y = \delta\sin(\delta)\\ p_2&=\delta e^{- i\delta} = \overline{\delta e^{i\delta}} = \delta_x - i\delta_y \end{align*} Our contour $C$ consists of three piecewise smooth curves $C_1+C_2+C_3$ defined as follows:
$C_1$ follows a path parallel to (and just above) the real axis from $\infty$ to $p_1$.
$C_2$ follows a counterclockwise circular path from $p_1$ to $p_2$, tracing the boundary of a circle that is centered at 0 with radius $\delta$. (Making just barely less than a full circle).
$C_3$ follows a path parallel to (and just below) the real axis from $p_2$ to $\infty$.
Lemma 2 Let $ z\in\mathbb{C} \backslash [0,\infty)$ and $Re(s) > 1$. Then $\displaystyle \lim_{\delta \to 0^+}\int_{|z|=\delta}\frac{-z^{s-1}}{e^z -1}\, dz =0$.
Proof - Lemma 2 Fix $s$. Let $A = | Re(s) - 1|$ and let $B = Im(s)$. We use the ML Inequality to obtain an upper bound for our integral. We also assume that $\delta < 1/2$ and apply lemma 1: \begin{equation*} \int_{|z|=\delta}\frac{-z^{s-1}}{e^z -1}\, dz \leq 2\pi\delta \cdot \max_{|z| = \delta} \left| \frac{-z^{s-1}}{e^z -1} \right| \leq 2\pi\delta \cdot \max_{|z| = \delta} \left| \frac{-z^{s-1}}{\delta /2} \right| = 4\pi \cdot \max_{|z| = \delta} \left|-z^{s-1}\right|. \end{equation*} So it only remains to show that \begin{equation*} \lim_{\delta \to 0^+} \max_{|z| = \delta} \left|-z^{s-1}\right| = 0. \end{equation*} Using the principal value of the complex logarithm (for argument $\theta$: $-\pi < \theta \leq\pi$), we rewrite $-z = \delta e^{i\theta} = e^{\log(\delta) +i\theta}$, so that \begin{align*} \left|-z^{s-1}\right| = \left| e^{ \left( \log(\delta) +i\theta \right) [ A + i B] } \right| = \left| e^{A\log(\delta) - B\theta} e^{ i \left( B\log(\delta) + A\theta \right) } \right| = \left| e^{A \log(\delta) - B\theta } \right| = \delta^A e^{- B\theta} \leq \delta^A e^{\pi \left| B \right| }. \end{align*} But $e^{\pi \left| B \right| }$ is fixed. Because $A > 0$, we have $\lim_{\delta \to 0^+} \delta^A = 0$, completing the proof.
End of Proof - Lemma 2
Main Theorem Let $s \in \mathbb{C}$. We recall the traditional definition of $\zeta(s)$: \begin{align*} \zeta(s) = \sum_{n=1}^{\infty}\frac{1}{n^s} = \prod_{p}\frac{1}{ \left( 1-p^{-s} \right)} \quad\text{(where $p$ ranges over an ordered list of all the primes)}, \end{align*} convergent for $Re(s) > 1$. For the contour of integration $C$ (a Hankel contour oriented in the positive direction defined above), we have that: \begin{align} \zeta(s)=\frac{\Gamma(1-s)}{ -2 \pi i} \int_C \frac{-z^{s-1}}{e^z - 1} \, dz \end{align} extends $\zeta(s)$ to a meromorphic function on $\mathbb{C}$, holomorphic except for a simple pole at $s = 1$.
Proof - Main Theorem.
Unless otherwise stated, we assume $Re(s) > 1$. We start with the Gamma function:
$$ \Gamma(s) = \int_0^\infty e^{-x} x^{(s-1)} \, dx. $$
Next, we replace $x$ by $nt$ in the integral (so that $dx=ndt$):
\begin{align*}
\Gamma(s) &= \int_0^\infty e^{-nt} (nt)^{(s-1)} n\, dt = \int_0^\infty e^{-nt}n^s t^{(s-1)} \, dt = n^s \int_0^\infty e^{-nt} t^{(s-1)} \, dt \\
n^{-s}\Gamma(s) &=\int_0^\infty e^{-nt} t^{(s-1)} \, dt \\
\frac{1}{n^s} &= \frac{1}{\Gamma(s)}\int_0^\infty e^{-nt} t^{(s-1)} \, dt.
\end{align*}
Now sum both sides over $n$:
\begin{align}
\sum_{n=1}^\infty \frac{1}{n^s} &= \frac{1}{\Gamma(s)}\sum_{n=1}^\infty \int_0^\infty e^{-nt} t^{(s-1)} \, dt \\
\zeta(s) &= \frac{1}{\Gamma(s)} \int_0^\infty \sum_{n=1}^\infty e^{-nt} t^{(s-1)} \, dt \\
\zeta(s) &= \frac{1}{\Gamma(s)} \int_0^\infty t^{(s-1)}\sum_{n=1}^\infty e^{-nt} \, dt.\tag{1}
\end{align}
Because we are assuming $Re(s) > 1$, the interchange of the sum and integral is justified by the uniform convergence of the sum. Notice that the sum over $n$ in the last integral is just a geometric series with the $n=0$ term missing. (We are justified in using a geometric series, because $0 < e^{-t} < 1$). So, we have: $$ \sum_{n=1}^\infty e^{-nt} = \sum_{n=0}^\infty e^{-nt} - e^{-0t} = \sum_{n=0}^\infty (e^{-t})^n - 1= \frac{1}{1-e^t} -1 = \frac{e^t}{1-e^t}=\frac{1}{e^t-1}. $$
But that means we can restate equation (1) (and while we are at it replace $dt$ with the slightly more standard $dx$), and have: $$ \zeta(s) = \frac{1}{\Gamma(s)} \int_0^\infty x^{(s-1)}\sum_{n=1}^\infty e^{-nx} \, dx = \frac{1}{\Gamma(s)} \int_0^\infty \frac{x^{s-1}}{e^x - 1} \, dx. $$
If we set $\displaystyle I(s)= \int_0^\infty \frac{x^{s-1}}{e^x - 1} \, dx$, then we have: $\displaystyle \zeta(s) = \frac{1}{\Gamma(s)} I(s) $.
Our next goal is to evaluate $I(s)$ in the complex plane and see where that takes us. Riemann starts with a related but not identical integral, which we call $I_0(s)$: $$ I_0(s) = \int_C \frac{-z^{s-1}}{e^z - 1} \, dz, $$ using the contour of integration $C$ described above. This contour stays well away from the singularities of the integrand at $2\pi i \mathbb{Z}$. We have: $$ I_0(s) = \int_C \frac{-z^{s-1}}{e^z - 1} \, dz = \int_\infty^{p_1} \frac{-z^{s-1}}{e^z - 1} \, dz + \int_{|z|=\delta} \frac{-z^{s-1}}{e^z - 1} \, dz + \int_{p_2}^\infty \frac{-z^{s-1}}{e^z - 1} \, dz. $$ We now study $I_0(s)$ as $\delta\rightarrow 0$. By Lemma 2, the second integral can be disregarded so the contour includes only the horizontal lines $\infty\rightarrow p_1$ and $p_2\rightarrow\infty$.
Note that $(-z)^{(s-1)}=e^{(s-1)log(-z)}$ and that $log(-z)=log|z|+iarg(-z)$. We will use $Arg(z)$, the principal value of $arg(z)$, defined as the value $\theta$ satisfying $-\pi < \theta \leq \pi$. Recall that $z \mapsto Arg(z)$ is discontinuous at each point on the nonpositive real axis. Let $z=x_0+iy$ for some fixed $x_0 < 0$. If $y\downarrow 0$ then $Arg(z) \rightarrow \pi$, whereas, if $y \uparrow 0$ then $Arg(z) \rightarrow -\pi$.
In the first integral, $-z=-\delta_x - i\delta_y$ approaches the nonpositive real axis from below, so (as $\delta\rightarrow 0$) we have
$Arg(-z) \rightarrow -\pi$. In the third integral, $-z=-\delta_x + i\delta_y$ approaches the nonpositive real axis from above, so we have $Arg(-z) \rightarrow \pi$.
\begin{align*}
I_0(s) &= %\int_C \frac{-z^{s-1}}{e^z - 1} \, dz =
\int_{\infty}^{p_1} \frac{e^{(s-1)(\log |z| -\pi i)}}{e^z - 1} \, dz
+ \int_{p_2}^{\infty} \frac{e^{(s-1)(\log |z| +\pi i)}}{e^z - 1} \, dz \\
&= -\int_{p_1}^{\infty} \frac{ e^{(s-1) \log |z| } e^{(s-1)(-\pi i)} }{e^z - 1} \, dz
+ \int_{p_2}^{\infty} \frac{ e^{(s-1) \log |z| } e^{(s-1)(\pi i)} }{e^z - 1} \, dz \\
&= -e^{(s-1)(-\pi i)} \int_{p_1}^{\infty} \frac{ e^{(s-1) \log |z| } }{e^z - 1} \, dz
+ e^{(s-1)(\pi i)} \int_{p_2}^{\infty} \frac{ e^{(s-1) \log |z| } }{e^z - 1} \, dz.
\end{align*}
Taken to the limit, our integrals are over a horizontal line with $dy = 0$, so we can replace $dz$ with $dx$ (and $z$ with $x$), replace $p_1$ and $p_2$ with $0$,
and consider $e^{(s-1) \log |z|} = x^{(s-1)}$:
\begin{align*}
\lim_{\delta \to 0+} I_0(s)
&=-e^{(s-1)(-\pi i)} \int_{0}^{\infty} \frac{ e^{(s-1) \log |z| } }{e^z - 1} \, dz
+ e^{(s-1)(\pi i)} \int_{0}^{\infty} \frac{ e^{(s-1) \log |z| } }{e^z - 1} \, dz \\
&= \left[ e^{i\pi(s-1)} - e^{-i\pi(s-1)} \right] \int_0^\infty \frac{x^{(s-1)}}{e^x - 1} \, dx.
\end{align*}
But we have: $e^{iw}-e^{-iw} = \cos(w) + i \sin(w) - \left[ \cos(w)-i\sin(w) \right] = 2i\sin(w)$.
So, the value in brackets is: $2i\sin(\pi (s-1)) = 2i\sin(\pi s - \pi) = -2i\sin(\pi s)$.
This gives our value for $I_0(s)$:
\begin{equation*}
\lim_{\delta \to 0+} I_0(s) =
-2i\sin(\pi s)\int_0^\infty \frac{x^{(s-1)}}{e^x - 1} \, dx = -2i\sin(\pi s) I(s).
\end{equation*}
Reviewing, we have:
\begin{equation*}
\zeta(s)= \frac{1}{\Gamma(s)} I(s)\quad\text{and}\quad I_0(s) = -2i\sin(\pi s) I(s)
\quad\text{so that}\quad \zeta(s)=\frac{1}{\Gamma(s)}\cdot\frac{1}{ -2i\sin(\pi s) } I_0(s).
\end{equation*}
Now use a rearrangement of Euler's Reflection Formula: $\displaystyle \frac{1}{\Gamma(s)} = \frac{\Gamma(1-s)\sin(\pi s)}{\pi}$, giving
\begin{equation*} \zeta(s)= \left[ \frac{1}{\Gamma(s)} \right] \frac{1}{ -2i\sin(\pi s) } I_0(s) = \left[ \frac{\Gamma(1-s)\sin(\pi s)}{\pi} \right] \frac{1}{ -2i\sin(\pi s) } I_0(s) = \frac{\Gamma(1-s)}{ -2 \pi i} I_0(s). \end{equation*} We therefore have: \begin{align} \zeta(s)=\frac{\Gamma(1-s)}{ -2 \pi i} \int_C \frac{-z^{s-1}}{e^z - 1} \, dz. \end{align}
Up to this point, we have assumed $Re(s) > 1$. But the integral is an entire function (it is uniformly convergent in any compact subset of $\mathbb{C}$ because $e^z$ grows faster than any power of $z$). That means that $\zeta(s)$ is analytic, except possibly at the positive integers where $\Gamma(1-s)$ has simple poles. But we already know that $\zeta(s)$ is analytic for $Re(s) > 1$, so the possible poles at $2,3,4...$ must be removable and cancel against zeros of the integral. Thus, the only pole of $\zeta(s)$ is at $s= 1$.
End of Proof - Main Theorem.
- 194
-
First of all thank you for the beautiful proof. In my question, however, I was trying to ask how to deal with the pole at $z=0$ of the second integral (which occurs for Re(s)<2 and bothered me for $Re(s)\leq1$). Therefore your statement that the second integrand is an entire function is also not true I believe. – Libertas Feb 27 '21 at 19:14
-
@Libertas First, your correction: yes, I should say “integral” not “integrand”. Next, I want to make sure I understand your point. By the “second integral”, I assume you are referring to the very last integral in my answer. Right? Assuming that is true, can we agree that the Hankel contour never assumes the value $z=0$? If we agree so far, are you saying you are concerned with what happens as $\delta$ gets very small? Or are you simply saying that you don’t think the second integral converges for $Re(s)\leq 1$? How I think of it...(continued) – TMurphy Feb 28 '21 at 22:39
-
@Libertas ... Fix any $s\in\mathbb{C}$ and fix any $\delta>0$. Choose large $R>0$ and make a disk $D(R)$ of radius $R$ centered at the origin. Inside the disk, the integrand (along the contour) is bounded. Outside the disk, the integrand (along the contour) quickly approaches $0$, and (soon) the “zeroness” of the integrand swamps the length of the contour. The ML inequality outside the disk quickly approaches zero for large $R$. With the inner contour bounded, the ML inequality for the entire contour is bounded. – TMurphy Feb 28 '21 at 22:39
-
1@Liberta I should add an important note. We use the $\delta$ approach because it makes it easy to obtain the last integral. However, once we have $\zeta(s)$ equal to the last integral, we can then use Cauchy's theorem and modify the contour without changing the value of the integral. For example, we can change the contour so that at $1 + i\delta$ the contour goes to $1+i\pi$ to $-1 + i\pi$ to $-1-i\pi$ to $1 -i\pi$ to $1-i\delta$. – TMurphy Feb 28 '21 at 23:02
-
Thank you very much. This really helped me. I am so gratefull for your time and effort! – Libertas Mar 05 '21 at 11:17