The Prime Number Theorem states that the number of prime numbers less
than \(x\) is asymptotic to \(x/\log x\) as
\(x\) becomes large.
Its proof was a crowning achievement of 19th century mathematics, and led to the development of a
vast amount of mathematics ever since. So understanding the classical proof
is more than just of historical interest. This article details
a minimal path to that first classical proof.
Introduction
The first complete proof of the Prime Number Theorem was given (independently)
by Hadamard and de la Vallé Poussin in 1896 [1,4]. It
was the culmination of work by many of the leading mathematicians of the 18th and 19th centuries
such as Euler, Chebyshev, Riemann and von Mangoldt.
Many simpler and ''elementary'' proofs of the theorem have been produced in the following 128 years [3,8,5,9].
However, the ideas involved in the
classical proof have been responsible for the development of a large amount of 20th and 21st century mathematics.
In particular, Riemann's short and legendary 1859 paper [7] can be rightly considered to be the birth of
Analytic Number Theory.
This article describes the first complete proof of the theorem.
H. Edwards's masterly historical account [2] traces (in the first four chapters)
the development of the proof from Riemann's 1859
paper until the final steps by Hadamard and de la Vallé Poussin in 1896, and it provides the main reference work for
this article. While Edwards's account
is focussed on the theory of the Riemann zeta function, the present article aims at providing a minimal path to the
proof of the Prime Number Theorem.
The proof consists of the following 13 items, containing
six propositions and five theorems.
The five theorems describe the crucial ideas behind the proof.
The proofs of the propositions or theorems are
accessible by clicking on the collapsible boxes
in display mode, labeled Proof, just after the statement.
There are also several references to background material
and results needed in the proofs,
that can be accessed by clicking on inline collapsible boxes.
Two of the theorems (Theorem 3 on the product expansion for
\(\xi(s)\) and Theorem 4 on the analytic
formula for \(\psi(x)\))
depend on some lemmas, that are numbered Lemma 1,
Lemma 2, etc.The proofs of these lemmas are also
accessible by clicking on collapsible boxes.
A few of these lemmas in turn depend on auxiliary
results, also called lemmas, but labeled Lemma A,
Lemma B, etc. whose proof is accessible by collapsible boxes.
Let \(\pi(x)\) be the prime counting function
\[\pi(x) =|\{p: p \mbox{ is prime and } p\leq x\}|.\]
For positive functions \(f(x)\) and \(g(x)\), the notation
\(f(x)\sim g(x)\) means that \(f(x)/g(x) \rightarrow 1 \) as \(x\rightarrow \infty\).
The Prime Number Theorem states that \(\pi(x)\sim \mbox{Li}(x)\),
where
\[\mbox{Li}(x)=\lim_{\epsilon \rightarrow 0^+} \left(\int_0^{1-\epsilon} \frac{dt}{\log t}+\int_{1+\epsilon}^x\frac{dt}{\log t}\right).\]
It is easy to see that \(\mbox{Li}(x) \sim x/\log x\), so an equivalent statement of the Prime Number Theorem is \(\pi(x)\sim x/\log x\).
.
Chebyshev's function
\(\pi(x)\)
arises from the point measure \(d\pi(t)\) that assigns weight 1 to
prime numbers and 0 to every
other number.
Let \(d\theta(t) =(\log t) d\pi(t)\). The corresponding function
\[\theta(x) = \sum_{p\lt x} \log p\]
(where the sum is over the prime numbers)
is Chebyshev's function. We assume by convention that the values of step functions such as
this at the points of
discontinuity are defined to be the average of their left and right limits.
The next proposition shows that in order to prove the PNT,
it is enough to derive the asymptotic behavior of \(\theta(x)\).
Proposition
\[ \hspace{2ex}\theta(x)\sim x \mbox{ implies }
\pi(x)\sim \mbox{Li}(x).\]
Let \(d\theta\) be the point measure that assigns weight \(\log p\) to prime \(p\). Then
\[\int_0^x\frac{d\theta (t)}{\log t} = \sum_{p\lt x}\frac{\log p}{\log p} = \pi(x).\]
So for \(x\lt y\),
\begin{equation} \label{pi}
\pi(y)-\pi(x) =\int_x^y\frac{d\theta(t)}{\log t} = \frac{\theta(y)}{\log y} - \frac{\theta(x)}{\log x} +\int_x^y\theta(t)\left(-\frac{1}{\log t}\right)'dt.\end{equation}
Suppose that \(\theta(x)\sim x\) for large \(x\). Let
\(\epsilon \gt 0\) be given, and take \(x\) large enough so that
\[(1-\epsilon) x \lt \theta(x) \lt (1+\epsilon) x.\]
Then from (\ref{pi})
we find
\begin{eqnarray*}
\pi(y)-\pi(x) &\leq & 2\epsilon \frac{x}{\log x}+
(1+\epsilon)\frac{y}{\log y} -(1+\epsilon)\frac{x}{\log x} +
(1+\epsilon)\int_x^y t\left(-\frac{1}{\log t}\right)' dt\\[2ex]
&=&2\epsilon \frac{x}{\log x}+(1+\epsilon) \left.\frac{t}{\log t}\right|_x^y -
(1+\epsilon)\left.\frac{t}{\log t}\right|_x^y+(1+\epsilon) \int_x^y\frac{dt}{\log t}
=2\epsilon \frac{x}{\log x}+(1+\epsilon)(\mbox{Li}(y)-\mbox{Li}(x)).
\end{eqnarray*}
So we find
\[\frac{\pi(y)}{\mbox{Li}(y)}\leq \frac{\pi(x)}{\mbox{Li}(y)}+\frac{2\epsilon x}{\mbox{Li}(y) \log x}-(1+\epsilon)\frac{\mbox{Li}(x)}{\mbox{Li}(y)} + 1 + \epsilon.\]
For a fixed \(x\), when \(y\) is large enough, we get
\[\frac{\pi(y)}{\mbox{Li}(y)}\leq 1+2\epsilon.\]
In a similar way,
\[\frac{\pi(y)}{\mbox{Li}(y)}\geq 1-2\epsilon .\]
So we conclude that
\[\lim_{x\rightarrow \infty} \frac{\pi(x)}{\mbox{Li}(x)}=1.\]
.
von Mangoldt's function
The function \(\theta(x)\) is constructed
using prime numbers only. If we also include prime powers, we get von Mangoldt's function
\[\psi(x)=\sum_{p^n\lt x} \log p =\sum_{n\lt x} \Lambda(n),\]
where the first sum is over all prime powers \(p^n\), and
\[\Lambda(n)=\begin{cases} \log p & \mbox{if \(n\) is a prime power }p^m\\
0 & \mbox{if \(n\) is not a prime power}
\end{cases}. \]
von Mangoldt's function
has a very useful connection
with the infinite sum
\begin{equation} \label{zeta0}
\zeta(s)=\sum_{n=1}^\infty\frac{1}{n^s}, \mbox{ Re}(s)\gt 1.
\end{equation}
The connection is via
Let \(p_1,p_2,p_3,\ldots\) be the prime numbers, and \(P_m=\{p_i: 1\leq i \leq m\}\).
Let \(s\) be a complex number with \(\sigma =\mbox{Re}(s)\gt 1\). Then
\[\sum_{j=1}^\infty \frac{1}{p_i^{sj}}=\left(1-\frac{1}{p_i^s}\right)^{-1}, \]
and so
\[\prod_{j=i}^m \left(1-\frac{1}{p_j^s}\right)^{-1}=\sum_{f\in {{\mathbb Z}_+}^m}\frac{1}{p_1^{sf(1)}p_2^{sf(2)} \cdots p_m^{sf(m)}} \]
and by the fundamental theorem of arithmetic, this last sum is
\(\displaystyle \sum_{n\in P_m}\frac{1}{n^s}\),
that is absolutely convergent because dominated by the series \( \displaystyle \sum_{n=1}^\infty \frac{1}{n^\sigma}\). So we can write
\[\zeta(s)=\sum_{n=1}^\infty \frac{1}{n^s} = \prod_{j=i}^m \left(1-\frac{1}{p_j^s}\right)^{-1}
+\sum_{n\not \in P_m} \frac{1}{n^s}.\]
If \(n\in P_m\), then \(n\gt m\), and so
\[\left|\sum_{n\not \in P_m} \frac{1}{n^s}\right| \leq \sum_{n=m}^\infty \frac{1}{n^\sigma} \rightarrow 0 \mbox{ as } m\rightarrow \infty .\]
So letting \(m\rightarrow \infty\) in
\[\prod_{j=i}^m \left(1-\frac{1}{p_j^s}\right)^{-1} = \zeta(s) - \sum_{n\not \in P_m} \frac{1}{n^s}\]
we conclude that
\[\prod_{j=i}^\infty \left(1-\frac{1}{p_j^s}\right)^{-1} = \zeta(s)\]
and the convergence is uniform on \(\mbox{Re}(s)\geq a\) for any \(a\gt 1\).
\begin{equation} \label{EulProdForm} \zeta(s) = \prod_p \left(1-\frac{1}{p^s}\right)^{-1}, \hspace{2ex}\mbox{Re}(s)\gt 1,\end{equation}
that is the heart of the Prime Number Theorem.
Let \(d\psi(t)\) be the point measure corresponding to \(\psi(t)\),
that assigns weight \(\log p\) to prime powers \(p^n\),
and zero to all other numbers. By taking the logarithmic derivative of (\ref{EulProdForm}), we obtain
the next result.
By definition,
\[\int_0^\infty \frac{d\psi(x)}{x^s}=\sum_{n=2}^\infty \frac{1}{n^s}\Lambda(n).\]
The zeta function
\[\zeta(s) =\sum_{n=1}^\infty \frac{1}{n^s}, \hspace{2ex}\mbox{Re}(s)\gt 1 \]
converges locally uniformly for \(\mbox{Re}(s)\gt 1\), and using Euler's product formula
we obtain by logarithmic differentiation
\[-\frac{\zeta'(s)}{\zeta(s)} = \sum_p \frac{\log p}{p^s-1} =\sum_p \log p \sum_{n=1}^\infty p^{-sn}=\sum_{n=2}\frac{1}{n^s}\Lambda(n)=\int_0^\infty \frac{1}{x^s} d\psi , \hspace{2ex}\mbox{Re}(s)\gt 1 .\]
Equation (\ref{int}) can be considered a re-statement in Stieltjes integral form of
Euler's product formula.
.
The PNT from \(\psi(x)\sim x\).
The heart of the proof of the Prime Number Theorem consists in exploiting equation (\ref{int})
to derive the asymptotic estimates \[\psi(x)\sim x.\] Once that is done,
the Prime Number Theorem is a consequence of the following simple result.
Proposition
.
\(\psi(x) \sim x \mbox{ implies } \theta(x)\sim x.\)
Suppose that \(\psi(x)\sim x\) as \(x\rightarrow \infty\).
Note that
\[\psi(x)=\sum_{n=1}^\infty \theta\left(x^{1/n}\right),\]
and the sum is finite, because \(\theta\left(x^{1/n}\right)\) is zero
as soon as \( 2^n\gt x \).
Since \(\theta(x)=0\) for \(n\gt \log x /\log 2\), and
\(\theta(x)\geq \theta\left(x^{1/n}\right)\) for \(n\geq 2\), we find
\[\psi(x)\leq \theta(x) +\frac{\log x}{\log 2} \theta\left(x^{1/2}\right).\]
So
\begin{equation} \label{theta}
\psi(x)-\frac{ \theta\left(x^{1/2}\right) \log x}{\log 2} \lt \theta(x) \lt \psi(x).
\end{equation}
Note that since \(\psi(x)\sim x\), if \(\epsilon \gt 0\),
\[\frac{\theta(x)}{x^{1+\epsilon}} \leq \frac{\psi(x)}{x^{1+\epsilon}} \rightarrow 0 \mbox{ as } x\rightarrow \infty.\]
So for small \(\epsilon \gt 0\),
\[\frac{ \theta\left(x^{1/2}\right)\log x}{x}= \frac{\theta\left(x^{1/2}\right)}{\left(x^{1/2}\right)^{1+\epsilon}} \frac{\log x}{\left(x^{1/2}\right)^{1-\epsilon}}\rightarrow 0 \mbox{ as }x\rightarrow \infty .\]
Dividing (\ref{theta}) by \(x\), we conclude that
\[\lim_{x\rightarrow \infty}\frac{\theta(x)}{x} =1.\]
.
Extension of \(\zeta(s)\)
So we have now reduced the proof of the PNT to the derivation of the asymptotic behavior
of \(\psi(x)\). This is where complex analysis comes in.
The function \(\zeta(s)\) in (\ref{zeta0})
is Riemann's zeta function, that had been used by Euler
mostly for integer values of \(s\), and later by Dirichlet for real values \(s\gt 1\).
The crucial connection (\ref{int}) between the prime numbers and \(\zeta(s)\) can now be exploited using Riemann's new and
far reaching idea that \(\zeta(s)\) extends
analytically to a meromorphic function with just a simple pole at \(s=1\), and then using
complex function theory to derive results on the prime numbers. Riemann's work in this area
can be considered the birth of Analytic Number Theory.
In order to state the next theorem, we introduce the
Let \(z\in {\mathbb C}\), and \(z\) not a negative integer. Define, for \(n\geq 0\), \(u_0=1\) and
\[u_n=\frac{n!(n+1)^z}{(z+1)(z+2)\cdots (z+n)}, n\geq 1.\]
Then
\[\frac{u_n}{u_{n-1}}=\frac{1}{1+\frac{z}{n}}\left(1+\frac{1}{n}\right)^z=\left(1-\frac{z}{n}+\frac{1}{n^2}a_n\right)\left(1+\frac{z}{n}+\frac{1}{n^2}b_n\right)
=1+\frac{c_n}{n^2},\]
where \(a_n,b_n,c_n\) are bounded as \(n\rightarrow \infty\).
So by standard results about
Suppose \(z_k, k\geq 1\) are complex numbers, and let \(\displaystyle p_n=z_1z_2\cdots z_n=\prod_{k=1}^nz_k\).
If \( \displaystyle \lim_{n\rightarrow \infty}p_n\) exists and is non-zero, we denote it by \(\displaystyle \prod_{k=1}^\infty z_k\), we call it
the infinite product of
the sequence \(z_k\), and say that the infinite product converges. In this case, clearly we must have \(z_k\neq 0\) for all \(k\), and
if we let \( \displaystyle p= \prod_{k=1}^\infty z_k\), we have
\[z_n=\frac{p_n}{p_{n-1}}\rightarrow \frac{p}{p} = 1 \mbox{ as } n\rightarrow \infty.\]
So a necessary condition for the infinite product \( \displaystyle \prod_{k=1}^\infty z_k \) to converge is that
\(z_k\rightarrow 1\) as \(k\rightarrow \infty\).
If \(z_k\neq 0\) for all \(k\) but \(\displaystyle \lim_{n\rightarrow \infty} p_n=0\), we say that the infinite product diverges to zero. An example is the case
\(z_k=a\) for all \(k\), where \(|a|\lt 1\).
If \(z_k=0\) for some \(k\), but there is some \(m\) such that \(z_k\neq 0\) for \(k\geq m\), and the infinite product \(z_m z_{m+1} z_{m+2}\cdots\)
converges, we say that the infinite product converges to zero. Note that in this terminology (used for example by Whittaker & Watson in A course of modern analysis)
an infinite product that converges to zero does not converge. Other sources will say that the infinite product diverges as soon as
the \(\displaystyle \lim_{n\rightarrow \infty}p_n=0\).
Since
\[\log\prod_{k=1}^n(1+a_k) =\sum_{k=1}^n \log(1+a_k),\]
we see from the definitions that
\[\prod_{n=1}^\infty (1+a_n)\mbox{ converges } \Longleftrightarrow \sum_{n=1}^\infty \log(1+a_n) \mbox{ converges }.\]
Note that if \(\displaystyle \sum_{n=1}^\infty \log(1+a_n)\) converges, then
\[\prod_{n=1}^\infty (1+a_n) =\exp\left(\sum_{n=1}^\infty \log(1+a_n)\right) \neq 0.\]
Let \(a_n\) be a sequence of complex numbers, with \(a_n\neq -1\) for all \(n\). So \(\log(1+a_n)\) is defined for all \(n\). Suppose \(a_n\rightarrow 0\) as \(n\rightarrow \infty\).
Find \(m\) such that \(|a_k|\leq 1/2\) for \(k\geq m\). Then we find, for \(k\geq m\),
\[\log(1+a_k)=a_k\left(1+\sum_{n=1}^\infty \frac{(-1)^n}{n+1}a_k^n\right).\]
Since
\[ \left|\sum_{n=1}^\infty \frac{(-1)^n}{n+1}a_k^n\right|\leq \sum_{n=1}^\infty \frac{1}{n+1}|a_k|^n\leq
\frac{1}{2} \sum_{n=1}^\infty \frac{1}{2^n}=\frac{1}{2},\]
we find
\[\frac{1}{2}|a_k|\leq |\log(1+a_k)|\leq\frac{3}{2} |a_k|.\]
So by the comparison test we see that \(\displaystyle \sum_{n=1}^\infty a_n\) converges absolutely if and only if \(\displaystyle \sum_{n=1}^\infty \log(1+a_n)\) converges absolutely.
We say that the infinite product \(\displaystyle \prod_{n=1}^\infty (1+a_n)\) converges absolutely if the infinite series \(\displaystyle \sum_{n=1}^\infty \log(1+a_n)\) converges
absolutely. If \(I_+=\{i:a_i\geq 0\}\) and \(I_-=\{i:a_i\lt 0\}\), absolute convergence of \(\displaystyle \prod_{n=1}^\infty (1+a_n)\) means that both
\(\displaystyle \prod_{i\in I_+}(1+a_i) \) and \( \displaystyle \prod_{i\in I_-}(1+a_i)\) converge. By the previous item, \(\displaystyle \prod_{n=1}^\infty (1+a_n)\) converges absolutely if and only if
\(\displaystyle \sum_{n=1}^\infty a_n\) converges absolutely.
,
\(\displaystyle \prod_{n=1}^\infty \frac{u_n}{u_{n-1}}\) converges absolutely. This means that
\[\lim_{n\rightarrow \infty}\prod_{k=1}^n \frac{u_k}{u_{k-1}} = \lim_{n\rightarrow \infty} u_n \]
exists for all \(z\in {\mathbb C}\), \(z\) not a negative integer, and we define it to be \(\Pi(z)\). So
\[\Pi(z) = \lim_{n\rightarrow \infty} \frac{n!(n+1)^z}{(z+1)(z+2)\cdots (z+n)}, z\neq -1, -2, -3, \ldots.\]
If \(z=m\) is a positive integer, then
\[u_n=\frac{n!(n+1)^m}{(m+1)\cdots (m+n)}=m!\frac{n!(n+1)^m}{(m+n)!}=m!\frac{(n+1)(n+1) \cdots (n+1)}{(n+1) (n+2)\cdots (n+m)}\rightarrow m! \mbox{ as } n\rightarrow \infty.\]
So \(\Pi(m)=m!\) and \(\Pi(z)\) is an extension of the factorial.
\begin{equation} \label{prodform} \Pi(s)=\prod_{k=1}^\infty \left(1+\frac{s}{k}\right)^{-1}\left(1+\frac{1}{k}\right)^s
\end{equation}
that was already known to Euler.
\(\Pi(s)\) is defined and analytic for all \(s\) except the negative integers, and
\(\Pi(n)=n!\) if \(n\) is a non-negative integer. A more common (but less natural)
notation for it
is \(\Gamma(s) = \Pi(s-1)\).
Theorem
For each \(s\neq 1\), define
\begin{equation} \label{zeta}
\zeta(s)=\frac{\Pi(-s)}{2\pi i}\int_{C} \frac{(-z)^s}{e^z-1}\frac{dz}{z},
\end{equation}
where \(C=C(\delta, \varepsilon)\) is the Hankel path that starts at a small distance \(\varepsilon\)
above the positive real axis at infinity, circles counterclockwise around the origin with radius
\(\delta\gt \varepsilon\), and returns to infinity traveling at distance \(\varepsilon\) below the positive real axis (see picture below).
\(\img{Fig1.svg}{}{25em}{}\)
The Hankel path \(C(\delta,\varepsilon)\)
Then \(\zeta(s)\) is a meromorphic function with a simple
pole at \(s=1\), and it coincides with the previous definition for \(\mbox{Re}(s)\gt 1\).
Moreover, \(\zeta(-2n)=0\) for all positive integers \(n\).
The starting point is the extension of the factorial (\ref{prodform}).
The integral representation
\begin{equation} \label{gamma}
\Pi(s)=\int_0^\infty e^{-s}x^sdx.
\end{equation}
valid for \(\mbox{Re}(s)\gt -1\), was also known to Euler. Using
(\ref{gamma})
with \(s\) replaced by \(s-1\), making the substitution \(x\mapsto nx\), and summing over \(n\geq 1\), we obtain the
integral representation
\begin{equation} \label{IntRep}
\int_0^\infty \frac{x^{s-1}}{e^x-1}dx = \Pi(s-1)\sum_{n=1}^\infty \frac{1}{n^s},
\end{equation}
valid for \(s\gt 1\).
In view of the integral representation (\ref{IntRep}),
Riemann considers the contour integral
\[
\int_{C} \frac{(-z)^s}{e^z-1}\frac{dz}{z},\]
where the integral is along the Hankel path \(C=C(\delta,\varepsilon)\) that starts at distance \(\varepsilon\)
above the positive real axis at infinity, circles counterclockwise around the origin with radius
\(\delta\gt \varepsilon\), and returns to infinity traveling at distance \(\varepsilon\) below the positive real axis (see figure below).
\(\img{Fig1.svg}{}{35em}{}\)
The Hankel path \(C(\delta,\varepsilon)\)
It is not hard to see that the integral does not depend on the values of \(\varepsilon\) and \(\delta\) (as long as they are small enough so that no singularities are enclosed).
This can be proved by truncating the path at \(x=R\) for some large \(R\), then joining it
via vertical segments to a similar path \(C(\delta_1,\varepsilon_1)\), where
\(\delta_1\lt \delta\), \(\varepsilon_1\lt \varepsilon\),
in order to form a closed contour without singularities inside, so that by Cauchy's theorem the integral will be zero, then let \(R\rightarrow \infty\) and notice that the integral over the
vertical segments will go to zero. This will prove that the integral over \(C(\delta,\varepsilon)\)
is the same as the integral over \(C(\delta_1,\varepsilon_1)\).
On \(L_+\) we have \(-z=-x-i\varepsilon\) and \(\log(-x-i\varepsilon)\rightarrow \log|z| -i\pi\) as \(\varepsilon \rightarrow 0\), while on \(L_-\), \(-z=-x+i\varepsilon\), and
\(\log(-x+i\varepsilon)\rightarrow \log|z| +i\pi\) as \(\varepsilon \rightarrow 0\). So the integral over \(L_+\) approaches \(\displaystyle e^{-i\pi s} \int_0^\infty \frac{x^s}{e^x-1}\frac{dx}{x}\)
and the integral over \(L_-\) approaches \(\displaystyle e^{i\pi s} \int_0^\infty \frac{x^s}{e^x-1}\frac{dx}{x}\)
as \(\varepsilon, \delta \rightarrow 0\). Also, if we write \(z=\delta e^{i\theta}\) on \(C_0\), the integrand is bounded by \(\displaystyle \frac{\delta^s}{|e^{\delta e^{i\theta}}-1|}=\frac{\delta^{s-1}}{|e^{i\theta} +O(\delta)|}\) and so the integral over \(C_0\) will approach zero as long as \(\mbox{Re}(s)\gt 1\).
So we conclude that, for \(\mbox{Re}(s) \gt 1\),
\[\int_C \frac{(-z)^s}{e^z-1}\frac{dz}{z} =2i\sin(\pi s)\int_0^\infty \frac{x^s}{e^x-1}\frac{dx}{x}=2i\sin(\pi s) \zeta(s) \Pi(s-1).\]
But the left side of this equation is an entire function of \(s\). Using standard identities for the factorial function, this leads to the definition of the Riemann zeta function as in the statement
of the theorem:
\[
\zeta(s)=\frac{\Pi(-s)}{2\pi i}\int_{C} \frac{(-z)^s}{e^z-1}\frac{dz}{z}.
\]
The integral on the right side
defines an entire function. So the only singularities can come from the simple poles of \(\Pi(-s)\) at the positive integers. But for \(s=2,3,4,\ldots\)
the function coincides with \(\sum_{n\geq 1} 1/n^s\), so the simple poles of \(\Pi(-s)\) must be cancelled by zeros of the integral, while at \(s=1\) we know that \(\sum_{n\geq 1} 1/n^s\) diverges,
so there must be a simple pole (with residue 1) coming from \(\Pi(-s)\).
So formula (\ref{zeta}) defines a function for all complex values of \(s\) except \(s=1\), where it has a simple pole, and it coincides with the previous
definition for \(\mbox{Re}(s)\gt 1\).
The Maclaurin expansion
\[\frac{x}{e^x-1}=\sum_{n=0}^\infty \frac{B_n}{n!}x^n\]
introduces the Bernoulli numbers \(B_n\) in the evaluation of \(\zeta(s)\). For \(s=-n\), \(n=0,1,2,3,\ldots\), the integrals over \(L_+\) and \(L_-\) will
cancel each other, and the integral over \(L_0\) as \(\delta, \varepsilon \rightarrow 0\) can easily be evaluated to obtain
\[
\zeta(-n)=(-1)^n\frac{B_{n+1}}{n+1}.
\]
In particular, since \(B_n=0\) for \(n\) odd and greater than 1, we find \(\zeta(-2n)=0\)
for all positive integers \(n\).
The first few values of \(\zeta(s)\) are:
\(n\)
\(0\)
\(-1\)
\(-2\)
\(-3\)
\(-4\)
\(-5\)
\(\zeta(n)\)
\[-\frac{1}{2}\]
\[-\frac{1}{12}\]
\(0\)
\[\frac{1}{120}\]
\(0\)
\[-\frac{1}{252}\]
The zeros at \(s=-2n\) are called the trivial zeros. But \(\zeta(s)\) has many
other zeros, which we
will denote by \(\rho\), that are crucial in understanding the distribution
of prime numbers. The Riemann hypothesis, proposed in his 1859 paper, is that all
the non-trivial zeros \(\rho\) of \(\zeta(s)\) have \(\mbox{Re}(\rho)=1/2\).
.
The function \(\xi(s)\)
So the zeta function contains information about the prime numbers.
But in order to extract it in the most efficient way, we now strip \(\zeta(s)\)
of some inessential information.
First of all we multiply it by \(s-1\) in order to make it an entire
function. Then we multiply it by \(\Pi(s/2)\) in order to eliminate the trivial zeros
at the negative even integers. At this point we would be left with an entire function without
real zeros (because as the next theorem shows \(\mbox{Im}(\rho)\neq 0\) for all \(\rho\)).
An additional simple factor of \(\pi^{-s/2}\) will turn it into the function
that plays a central role in the proof of the PNT, and that is
described in the next theorem.
Theorem
Let
\[ \xi(s)=(s-1)\Pi\left(\frac{s}{2}\right) \pi^{-s/2} \zeta(s).
\]
Then:
\(\xi(s)\) is entire.
\(\xi(s)\) satisfies the functional equation
\begin{equation} \label{funceq}
\xi(1-s)=\xi(s).
\end{equation}
\(\xi(s)\) has no zeros
outside the strip \(0\leq \mbox{Re}(s) \leq 1\).
\(\xi(s)\) is an even function of \(s-1/2\), and the
coefficients of the Taylor expansion of \(\xi(s)\) at \(s=1/2\) are positive real numbers.
In particular, \(\xi(s)\) has no zeros on the real axis.
Formula
(\ref{IntRep})
in the proof of Theorem 1
that led to the analytic continuation of \(\zeta(s)\) from \(s\gt 1\)
to the whole complex plane (except \(s=1\)) was obtained by substituting \(nx\) for \(x\) in the
integral representation (\ref{gamma})
for \(\Pi(s-1)\), and then summing over \(n\). If we instead make the substitution \(x\mapsto n^2\pi x\)
in the integral representation for \(\Pi(s/2-1)\) and sum, we find
\begin{equation} \label{intpsi}
\Pi\left(\frac{s}{2}-1\right)\pi^{-s/2} \zeta(s) = \int_0^\infty \phi(x)x^{s/2}\frac{dx}{x},
\end{equation}
where
\[\phi(x)= \sum_{n=1}^\infty e^{-\pi n^2x}.\]
The function \(\phi(x)\) is related to the Jocobi theta function
\( \displaystyle G(u)=\vartheta(0;iu^2)=\sum_{n\in {\mathbb Z}} e^{-\pi n^2 u^2}\) by
\[2\phi(u^2) +1 =G(u)\] and so the
Jacobi theta function \(\vartheta(z;\tau)\) is defined for \(z\in {\mathbb C}\) and \(\mbox{Im} \tau \gt 0\) to be
\[\vartheta(z;\tau)=\sum_{n\in {\mathbb Z}} e^{\pi i n^2\tau +2\pi i nz}.\]
Consider the special case
\[G(u)=\vartheta(0;iu^2)=\sum_{n\in {\mathbb Z}} e^{-\pi n^2 u^2}, u\gt 0\]
We prove that \(G\) satisfies the functional equation
\begin{equation} \label{functional}G(u)=\frac{1}{u}G\left(\frac{1}{u}\right).\end{equation}
Let \(u\gt 0\) be fixed, and consider the function
\[f(x)=e^{-\pi x^2/u^2}.\]
Let
\[\hat{f}(s)=\int_{-\infty}^\infty f(x) e^{2\pi i xs}dx\]
be its Fourier transform.
Then by
Let \(f: {\mathbb R} \rightarrow {\mathbb C}\) be a function that goes to zero sufficiently fast at \(\pm \infty\). The Fourier transform of \(f\) is
\[\hat{f}(u) =\int_{-\infty}^\infty f(x) e^{2\pi i x u} dx.\]
We can "periodify" \(f\) by defining
\begin{equation}\label{periodify} F(x) =\sum_{n\in {\mathbb Z}} f(x+n).\end{equation}
Then by the theory of Fourier series [see Rudin, Real and Complex Analysis],
\begin{equation} \label{Fourier} F(x) = \sum_{n\in {\mathbb Z}} a_n e^{2\pi i nx},\end{equation}
where
\[a_n= \int_0^1F(x) e^{-2\pi i nx}dx.\]
We then find
\[a_n=\int_0^1 \sum_{m\in{\mathbb Z}} f(x+m)e^{-2\pi i nx}dx =\sum_{m\in{\mathbb Z}}\int_0^1 f(x+m) e^{-2\pi i nx} dx \\
=\sum_{m\in{\mathbb Z}} \int_m^{m+1} f(t) e^{-2\pi i n(t-m)}dt = \sum_{m\in{\mathbb Z}} \int_m^{m+1} f(t) e^{-2\pi i nt}dt = \int_{-\infty}^\infty f(t) e^{-2\pi i nt} dt = \hat{f}(-n).\]
Setting \(x=0\) in (\ref{periodify}) and (\ref{Fourier}), we find
\[ \sum_{n\in{\mathbb Z}} f(n) =F(0) = \sum_{n\in{\mathbb Z}} a_n=\sum_{n\in{\mathbb Z}}\hat{f}(-n)=\sum_{n\in{\mathbb Z}}\hat{f}(n).\]
This proves Parseval's Theorem:
\[\sum_{n\in{\mathbb Z}}f(n) = \sum_{n\in{\mathbb Z}} \hat{f}(n).\]
\begin{equation} \label{Parseval}\sum_{n\in {\mathbb Z}} f(n) =\sum_{n\in{\mathbb Z}} \hat{f}(n).\end{equation}
We find
\[\sum_{n\in {\mathbb Z}} f(n) =\sum_{n\in {\mathbb Z}}e^{-\pi n^2/u^2}=G\left(\frac{1}{u}\right).\]
Using the integral formula
\[\int_{-\infty}^\infty e^{-\pi x^2}dx=1\]
and Cauchy's theorem to change the path of integration from \((-\infty +iu, \infty+iu)\) to \((-\infty,\infty)\), we also find
\[\hat{f}(n) =\int_{-\infty}^\infty e^{-\pi x^2/u^2}e^{2\pi i xn}dx= u\int_{-\infty}^\infty e^{-\pi t^2}e^{2\pi i u n t}dt= ue^{-\pi u^2 n^2}\int_{-\infty}^{\infty} e^{-\pi(t-iu)^2} dt\\
=ue^{-\pi u^2n^2} \int_{-\infty +iu}^{\infty +iu}e^{-\pi z^2} dz=ue^{-\pi u^2n^2}\int_{-\infty}^\infty e^{-\pi x^2}dx=ue^{-\pi u^2n^2}.\]
So we conclude that
\[\sum_{n\in {\mathbb Z}} \hat{f}(n) =u G(u),\]
and so by Parseval's theorem (\ref{Parseval}) the functional equation (\ref{functional}) follows.
\(uG(u)=G(1/u)\) for \(G(u)\) gives
\begin{equation} \label{functpsi}
\phi(x) =\frac{1}{\sqrt{x}}\phi\left(\frac{1}{x}\right) +\frac{1}{\sqrt{x}}-\frac{1}{2}.
\end{equation}
Differentiation of this identity gives
\begin{equation} \label{psi}
\phi(1)+4\phi'(1)+\frac{1}{2}=0.
\end{equation}
By splitting the integral
(\ref{intpsi})
from \(0\) to \(1\) and \(1\) to \(\infty\), and then using the functional equation
(\ref{functpsi}),
some routine calculations give
\begin{equation} \label{entint}
\Pi\left(\frac{s}{2}-1\right) \pi^{-s/2} \zeta(s) = \int_1^\infty \phi(x) \left(x^{s/2}+x^{(1-s)/2}\right) \frac{dx}{x} -\left(\frac{1}{s}+\frac{1}{1-s}\right).
\end{equation}
The integral on the right side of
(\ref{entint})
is an entire function of \(s\). So multiplying the equation by \(s(s-1)/2\) we get an entire function, denoted by \(\xi(s)\):
\begin{equation} \label{xi}
\xi(s)=(s-1)\Pi\left(\frac{s}{2}\right)\pi^{-s/2}\zeta(s) =\frac{1}{2} -\frac{s(1-s)}{2}\int_1^\infty \phi(x)\left(x^{s/2}+x^{(1-s)/2}\right)\frac{dx}{x}.
\end{equation}
The right side of
(\ref{xi})
is invariant under \(s\rightarrow 1-s\), and so we
obtain the functional equation
\[\xi(s)=\xi(1-s).\]
We show that \(\zeta(s)\) (and hence \(\xi(s)\)) has no zeros for \(\mbox{Re}(s)\gt 1\).
Let \(s\in {\mathbb C}\) with \(\sigma=\mbox{Re}(s)\gt 1\), and let \((p_n:n\geq 1)\) be the prime numbers.
Let \(a_n=(p_n^s-1)^{-1}\), so that \(1+a_n=(1-p_n^{-s})^{-1}\). Since
\[\sum_{n=1}^\infty |a_n|=\sum_{n=1}^\infty \frac{1}{|p_n^s-1|}\leq \sum_{n=2}^\infty \frac{1}{n^\sigma-1}\lt \infty ,\]
we conclude from basic results on
Suppose \(z_k, k\geq 1\) are complex numbers, and let \(\displaystyle p_n=z_1z_2\cdots z_n=\prod_{k=1}^nz_k\).
If \( \displaystyle \lim_{n\rightarrow \infty}p_n\) exists and is non-zero, we denote it by \(\displaystyle \prod_{k=1}^\infty z_k\), we call it
the infinite product of
the sequence \(z_k\), and say that the infinite product converges. In this case, clearly we must have \(z_k\neq 0\) for all \(k\), and
if we let \( \displaystyle p= \prod_{k=1}^\infty z_k\), we have
\[z_n=\frac{p_n}{p_{n-1}}\rightarrow \frac{p}{p} = 1 \mbox{ as } n\rightarrow \infty.\]
So a necessary condition for the infinite product \( \displaystyle \prod_{k=1}^\infty z_k \) to converge is that
\(z_k\rightarrow 1\) as \(k\rightarrow \infty\).
If \(z_k\neq 0\) for all \(k\) but \(\displaystyle \lim_{n\rightarrow \infty} p_n=0\), we say that the infinite product diverges to zero. An example is the case
\(z_k=a\) for all \(k\), where \(|a|\lt 1\).
If \(z_k=0\) for some \(k\), but there is some \(m\) such that \(z_k\neq 0\) for \(k\geq m\), and the infinite product \(z_m z_{m+1} z_{m+2}\cdots\)
converges, we say that the infinite product converges to zero. Note that in this terminology (used for example by Whittaker & Watson in A course of modern analysis)
an infinite product that converges to zero does not converge. Other sources will say that the infinite product diverges as soon as
the \(\displaystyle \lim_{n\rightarrow \infty}p_n=0\).
Since
\[\log\prod_{k=1}^n(1+a_k) =\sum_{k=1}^n \log(1+a_k),\]
we see from the definitions that
\[\prod_{n=1}^\infty (1+a_n)\mbox{ converges } \Longleftrightarrow \sum_{n=1}^\infty \log(1+a_n) \mbox{ converges }.\]
Note that if \(\displaystyle \sum_{n=1}^\infty \log(1+a_n)\) converges, then
\[\prod_{n=1}^\infty (1+a_n) =\exp\left(\sum_{n=1}^\infty \log(1+a_n)\right) \neq 0.\]
Let \(a_n\) be a sequence of complex numbers, with \(a_n\neq -1\) for all \(n\). So \(\log(1+a_n)\) is defined for all \(n\). Suppose \(a_n\rightarrow 0\) as \(n\rightarrow \infty\).
Find \(m\) such that \(|a_k|\leq 1/2\) for \(k\geq m\). Then we find, for \(k\geq m\),
\[\log(1+a_k)=a_k\left(1+\sum_{n=1}^\infty \frac{(-1)^n}{n+1}a_k^n\right).\]
Since
\[ \left|\sum_{n=1}^\infty \frac{(-1)^n}{n+1}a_k^n\right|\leq \sum_{n=1}^\infty \frac{1}{n+1}|a_k|^n\leq
\frac{1}{2} \sum_{n=1}^\infty \frac{1}{2^n}=\frac{1}{2},\]
we find
\[\frac{1}{2}|a_k|\leq |\log(1+a_k)|\leq\frac{3}{2} |a_k|.\]
So by the comparison test we see that \(\displaystyle \sum_{n=1}^\infty a_n\) converges absolutely if and only if \(\displaystyle \sum_{n=1}^\infty \log(1+a_n)\) converges absolutely.
We say that the infinite product \(\displaystyle \prod_{n=1}^\infty (1+a_n)\) converges absolutely if the infinite series \(\displaystyle \sum_{n=1}^\infty \log(1+a_n)\) converges
absolutely. If \(I_+=\{i:a_i\geq 0\}\) and \(I_-=\{i:a_i\lt 0\}\), absolute convergence of \(\displaystyle \prod_{n=1}^\infty (1+a_n)\) means that both
\(\displaystyle \prod_{i\in I_+}(1+a_i) \) and \( \displaystyle \prod_{i\in I_-}(1+a_i)\) converge. By the previous item, \(\displaystyle \prod_{n=1}^\infty (1+a_n)\) converges absolutely if and only if
\(\displaystyle \sum_{n=1}^\infty a_n\) converges absolutely.
that
\[\sum_{n=1}^\infty \log(1+a_n) = \log \prod_{n=1}^\infty \left(1-\frac{1}{p_n^s}\right)^{-1} =L \in {\mathbb C},\]
and so
\[\zeta(s)=\prod_{n=1}^\infty \left(1-\frac{1}{p_n^s}\right)^{-1}=e^L\neq 0.\]
Using the functional equation \(\xi(s)=\xi(1-s)\) it follows that \(\xi(s)\) has no zeros
outside \(0\leq \mbox{Re}(s) \leq 1\).
Using the estimates
\[\phi(x)=\sum_{n=1}^\infty e^{-\pi n^2x}\leq \sum_{n=1}^\infty e^{-\pi nx}=\frac{1}{e^{\pi x}-1}\]
and
\[|\phi'(x)|=\pi \sum_{n=1}^\infty n^2 e^{-\pi n^2x}\leq \pi \sum_{n=1}^\infty n^2 e^{-\pi nx}= \pi e^{-\pi x}\frac{1+e^{-\pi x}}{(1-e^{-\pi x})^3},\]
we see that for every \(s\),
\[
|x^s \phi(x)|\rightarrow 0 \mbox{ and } |x^s\phi'(x)|\rightarrow 0 \mbox{ as } x\rightarrow \infty .
\]
Integrating
(\ref{xi})
by parts twice and using
(\ref{psi})
we then find
\[\xi(s) = 2 \int_1^\infty \left(x^{s/2-1/2} +x^{-s/2}\right) \left(x^{3/2} \phi'(x)\right)'dx\\
=4\int_1^\infty x^{-1/4}\cosh\left(\frac{1}{2}\left(s-\frac{1}{2}\right)\log x \right) \left(x^{3/2} \phi'(x)\right)'dx.\]
Using the Maclaurin expansion for \(\cosh\), this gives
\[
\xi(s)=\sum_{n=0}^\infty c_n \left(s-\frac{1}{2}\right)^{2n}, \mbox{ or } \xi\left(\frac{1}{2}+it\right) =\sum_{n=0}^\infty (-1)^n c_n t^{2n},
\]
where
\[c_n=\frac{4}{(2n)!}\int_1^\infty x^{-1/4}\left(\frac{\log x}{2}\right)^{2n} \left(x^{3/2}\phi'(x)\right)' dx.\]
We find
\[\frac{d}{dx}\left(x^{3/2}\phi'(x)\right)= \pi x^{1/2} \sum_{n=1}^\infty n^2\left(\pi xn^2-\frac{3}{2}\right) e^{-\pi n^2 x}\]
and so all coefficients \(c_n\) are positive.
Note that as a consequence of the functional
equation (\ref{funceq}), the zeros of \(\xi(s)\) will occur in pairs \((\rho, 1-\rho)\).
.
The product expansion for \(\xi(s)\)
Riemann had the insight that the function \(\xi(s)\) admits an infinite product expansion
whose factors are \(1-s/\rho\), where \(\rho \) are the zeros of \(\xi(s)\), and that this
infinite product expansion is the key to the derivation of the
information about the prime numbers.
This product expansion (that essentially says
that \(\xi(s)\) behaves like an infinite degree polynomial) is similar
to Euler's expansion for \(\sin(\pi s)\) in terms of factors \(1-s^2/n^2\).
But Riemann did not prove his statement in his short 1859 paper.
He simply stated that the Taylor series for \(\xi(s)\)
around \(s=1/2\) converges ''very rapidly''. A rigorous proof of the product
expansion for \(\xi(s)\) (described in the next theorem)
was given by Hadamard about 30 years later, and it was the first major
progress towards a rigorous proof of the PNT following Riemann's 1859 paper.
The proof of the product expansion is not easy. The essential ingredients are an upper bound on the number of zeros
of \(\xi(s)\) inside a circle of radius \(R\) (for which part (e) of Theorem 2 is used) and a general complex analysis
result that is an analogue of Liouville's theorem when an entire function is also even.
Theorem
The function \(\xi(s)\) has a product expansion
\[
\xi(s) =\xi(0)\prod_\rho \left(1-\frac{s}{\rho}\right),
\]
where the product is over all the zeros \(\rho\) of \(\xi(s)\), and the zeros \(\rho, 1-\rho\)
are assumed to be paired, so that the product is
\[\xi(s) =\xi(0)\prod_{\rm Im(\rho)\gt 0} \left(1-\frac{s(1-s)}{\rho(1-\rho)}\right).\]
It is reasonable to expect that in order to prove that a product
\[\prod_{\rho} \left(1-\frac{s}{\rho}\right)\]
converges we need to have an estimate of how many zeros \(\rho\) there are.
The first step in the proof will be to put a bound on the growth of \(\xi(s)\), as described
in the next result.
Lemma
For all large enough \(R\) and \(|s-1/2|\leq R\), we have
\(|\xi(s)|\leq R^R.\)
This will be used to provide the estimate on the number of zeros.
According to Theorem 2 d., the coefficients \(c_n\) for the power series of \(\xi\)
as an even function of \(s-1/2\) are positive real numbers.
Hence if \(|s-1/2|\leq R\), we find
\[|\xi(s)|=\left|\sum_{n=0}^\infty c_n \left(s-\frac{1}{2}\right)^{2n}\right|\leq \sum_{n=0}^\infty c_n \left|s-\frac{1}{2}\right|^{2n}\leq
\sum_{n=0}^\infty c_n R^{2n}=\xi(R+1/2). \] For a fixed \(R\gt 0\), let \(N\in {\mathbb N}\) be such that
\[\frac{1}{4} +\frac{R}{2} \leq N \lt \frac{1}{4}+\frac{R}{2}+1.\]
Recall that
\[\xi(s) =(s-1) \Pi\left(\frac{s}{2}\right) \pi^{-s/2} \zeta(s).\]
Since \(\displaystyle \xi(t) = \sum_{n=0}^\infty c_n (t-1/2)^{2n}\) is increasing for \(t\) real and \(t\geq 1/2\), we have
\[\xi\left(\frac{1}{2}+R\right) \leq \xi(2N)=(2N-1) \Pi(N) \pi^{-N}\zeta(2N)\leq (2N) N! \zeta(2) \\
\leq 2\zeta(2) N^{N+1}\leq 2\zeta(2) \left(\frac{1}{4}+\frac{R}{2} +1\right)^{1/4+R/2+2}
\leq R^R\]
for \(R\) large enough. \(\blacksquare\)
Lemma
For \(R\gt 0\), let \(n(R)\) be the number of zeros of \(\xi(s)\) such that \(|s-1/2|\leq R\). Then \(n(R)\leq 3R\log R\) for all large enough \(R\).
If \(R\gt 0\) and \(f\) is a function defined and analytic for \(|z|\leq R\) and with \(f(0)\neq 0\), let
\[Z(f;R)=\{z\in {\mathbb C}: f(z)=0 \mbox{ and } |z|\lt R\}\]
\[Z_0(f;R)=\{z\in{\mathbb C}: f(z)=0 \mbox{ and } |z|\leq R\}.\]
Then \(Z_0(f;R)\subset Z(f;2R)\). If \(z\in Z(f;2R)\), then \(2R/|z|\gt 1\), so \(\log(|2R/z|)\gt 0\), and we find
\[\sum_{z\in Z(f;2R)}\log\left|\frac{2R}{z}\right| \geq \sum_{z\in Z_0(f;R)}\log\left|\frac{2R}{z}\right|\geq \log(2) |Z_0(f;R)|,\]
because \(2R/|z|\geq 2\) for \(z\in Z_0(f;R)\). If \(f\) has no zeros on the circle \(|z|=2R\),
Suppose that \(f\) is analytic on an open set containing \(\overline{B}(0,R)\), and \(f(z)\neq 0\) in \(B(0,R)\). Integrating the Maclaurin series for \(f\) we find
\[f(0)=\frac{1}{2\pi}\int_0^{2\pi}f\left(Re^{i\theta}\right)d\theta.\]
Note that for any \(w\neq 0\) we have \(\mbox{Re}(\log w)=\log|w|\). So applying the above formula to \(\log(f(z))\) and taking the real part, we find
\[\log|f(0)|=\frac{1}{2\pi}\int_0^{2\pi}\log\left|f(R e^{i\theta})\right|d\theta.\]
Now suppose \(f\) has a simple zero at \(z_0\in B(0,R)\), and \(z_0\neq 0\).
If we can find some \(\phi(z)\) with a simple pole at \(z_0\), without zeros for \(|z|\leq R\), and such that \(|\phi(z)|=1\) on \(|z|=R\), then the function \(f(z)\phi(z)\) has no zeros
in \(B(0,R)\), and since \(|f(z)\phi(z)|=|f(z)|\) for \(|z|=R\), applying the above formula to \(f(z)\phi(z)\) we have
\[\log(|f(0)\phi(0)|) =\frac{1}{2\pi} \int_0^{2\pi} \log(|f(Re^{i\theta})| d\theta.\]
Such a factor is
\[\phi(z) =\frac{R^2-\overline{z_0}z}{R(z-z_0)}, \]
as we can see by
noting that \(R^2-\overline{z_0}z\neq 0\) in \(B(0,R)\), and for \(|z|=R\),
\[\left|\frac{R^2-\overline{z_0}z}{R(z-z_0)}\right|=\left|\frac{(R^2-\overline{z_0}z)\overline{z}}{R(z-z_0)\overline{z}}\right|=
\left|\frac{R^2(\overline{z}-\overline{z_0})}{R^2(z-z_0)}\right|=1.\]
Since \(\phi(0)=R/|z_0|\), we find
\[\log\left(|f(0)|\frac{R}{|z_0|}\right)=\frac{1}{2\pi}\int_0^{2\pi} \log\left|f(R e^{i\theta})\right| d\theta.\]
In the same way, if \(f(0)\neq 0\) and \(f\) has zeros \(z_1,\ldots, z_k\) in \(B(0,R)\) (repeated according to multiplicity), consider
\(\displaystyle f(z)\frac{R^2-\overline{z_1}z}{R(z-z_1)}\cdots \frac{R^2-\overline{z_k}z}{R(z-z_k)}\). We then find Jensen's formula:
\[ \log\left(|f(0)|\frac{R}{|z_1|}\cdots \frac{R}{|z_k|}\right)=\frac{1}{2\pi}\int_0^{2\pi} \log\left|f(R e^{i\theta})\right| d\theta.\]
applies and we find
\begin{equation} \label{Jen}
\log|f(0)|+\log(2)|Z_0(f;R)|\leq \log|f(0)|+\sum_{z\in Z(f;2R)} \log\left|\frac{2R}{z}\right|\\
\hspace{25ex} =\displaystyle \log\left(|f(0)|\prod_{z\in Z(f;2R)}\frac{2R}{|z|}\right)=\frac{1}{2\pi}\int_0^{2\pi} \log\left|f(2R e^{i\theta})\right|d\theta.
\end{equation}
Now let \(f(z)=\xi(z+1/2)\). Then \(n(R)=|Z_0(f;R)|\), and by Lemma 1, we find
\[\left|f\left(2Re^{i\theta}\right)\right| =\left|\xi\left(\frac{1}{2}+2Re^{i\theta}\right)\right|\leq \left(2R\right)^{2R}.\]
If \(\xi\) has no zeros on the circle \(|s-1/2|= 2R\), we find from (\ref{Jen})
\[\log\left|\xi\left(\frac{1}{2}\right)\right|+\log(2) n(R) \leq 2R\log(2R) =2R\log R +2R\log(2),\]
or
\[n(R) \leq \frac{2}{\log 2} R\log R +2R -\frac{\log|\xi(1/2)|}{\log 2}\]
Let \(R_0\) be such that the right side of the above inequality is \(\leq 3 R \log R\) for \(R\geq R_0\).
So \(n(R)\leq 3R\log R\) if \(R\geq R_0\) and there are no zeros on the circle \(|s-1/2|=2R\).
If \(R\geq R_0\) and there are some zeros on the circle \(|s-1/2|=2R\), we can find some \(\epsilon_R \gt 0\) such that \(f\) has no
zeros on \(|s-1/2|=2(R+\epsilon)\) for all \(0\lt \epsilon \leq \epsilon_R\), and then
\[n(R)\leq n(R+\epsilon)\leq 3(R+\epsilon)\log(R+\epsilon) \]
for all \(0\lt \epsilon \leq \epsilon_R\). Letting \(\epsilon\rightarrow 0\), we find
\(n(R)\leq 3R\log R\). So the inequality holds for all \(R\geq R_0\). \(\blacksquare\)
The above estimate of the number of zeros of \(\xi(s)\) allows us to prove the next result.
Lemma
Let \(\rho_k: k=1,2,3,\ldots\) be the zeros of \(\xi(s)\) ordered so that
\(|\rho_k -1/2|\leq |\rho_{k+1} -1/2|\). Let \(\varepsilon \gt 0\) be given. Then
the series
Define \(R_n\) implicitly by \(4R_n \log R_n=n\). Then if \(n(R)\) is as
defined in the previous lemma, we find \(n(R_n)\leq 3R_n\log R_n =3n/4\lt n\).
This means that the \(n\)-th zero is outside the circle of radius \(R_n\) centered at \(1/2\), so \(|\rho_n-1/2|\gt R_n\), and
\[\frac{1}{|\rho_n-1/2|^{1+\epsilon}}\lt \frac{1}{R_n^{1+\epsilon}}=\frac{(4\log R_n)^{1+\epsilon}}{n^{1+\epsilon}}=
\frac{(4\log R_n)^{1+\epsilon}}{n^{1+\epsilon/2}}\frac{1}{n^{1+\epsilon/2}}.\]
Since \(\log R_n \lt \log 4+\log R_n +\log \log R_n = \log(4R_n \log R_n)=\log n\), we find
\[\frac{(4\log R_n)^{1+\epsilon}}{n^{1+\epsilon/2}} \lt \frac{(4\log n)^{1+\epsilon}}{n^{1+\epsilon/2}}\lt 1\] for large \(n\),
and so
\[\frac{1}{|\rho_n-1/2|^{1+\epsilon}}\lt \frac{1}{n^{1+\epsilon/2}}\]
holds for all large enough \(n\), and the series \(\displaystyle \sum_{n=1}^\infty \frac{1}{|\rho_n-1/2|^{1+\epsilon}}\) converges.\(\blacksquare\)
The infinite product \(P(s)\) has exactly the same zeros as \(\xi(s)\).
We will show that in fact it is the same as \(\xi(s)\), up to a multiplicative constant.
Our main tool is the next result from complex analysis,
that can be thought of as an analogue of Liouville's theorem
when an entire function is also even.
Lemma
Suppose that \(f(s)\) is entire and even and for each \(\varepsilon \gt 0\),
\(\mbox{Re} (f(s))\leq \varepsilon |s|^2\) for all large enough \(|s|\). Then \(f(s)\) is a
constant function.
The proof of Lemma depends on the following result,
that provides a bound on \(|f(s)|\) on a disk \(|s|\leq r\) if we are
given a bound for \(\mbox{Re}f(s)\) on a larger disk \(|s|\leq R\).
Lemma A Suppose \(f(s)\) is analytic on \(|s|\leq R\), \(f(0)=0\), and \(r\lt R\). If \(\mbox{Re}f(s)\leq M \) for \(|s|\leq R\), then
\[
|f(s)|\leq \frac{2Mr}{R-r} \mbox{ for } |s|\leq r.
\]
Write \(f(s)=u(s)+iv(s)\). Since \(u(s)\) is harmonic, its maximum on \(|s|\leq R\) is achieved at the boundary. Let
\[\phi(s)=\frac{f(s)}{s(2M-f(s))}\]
Note that
\[u-2M \leq u \leq M \leq 2M-u,\]
and so
\[|u(s)|\leq 2M-u(s).\]
Then we find
\[|2M-f(s)|^2=(2M-u(s))^2+v(s)^2\geq u(s)^2+v(s)^2=|f(s)|^2.\]
Hence
\[|\phi(s)|\leq \frac{1}{|s|}=\frac{1}{R} \mbox{ on } |s|\leq R.\]
Then
\[|f(s)|=\left|\frac{2Ms\phi(s)}{1+s\phi(s)}\right|\leq \frac{2Mr/R}{1-r/R}=\frac{2Mr}{R-r} \mbox{ on } |s|\leq r. \blacksquare\]
To prove Lemma ,
suppose that \(f(s)\) is entire and even, and for every \(\epsilon \gt 0\), \(\mbox{Re}f(s)\leq \epsilon |s|^2\) for large enough |s|. Assume first \(f(0)=0\). Write
\[f(s)=\sum_{n=2}a_ns^n.\]
Then given \(r\gt 0\) and using Lemma A with \(R=2r\), we find
\[|a_n|=\left|\frac{1}{2\pi i}\int_{|s|=r}\frac{f(re^{i\theta}}{s^{n+1}} ds\right| \leq \frac{1}{2\pi} \int_0^{2\pi} \frac{|f(re^{i\theta})|}{r^n}d\theta
\leq \frac{1}{2\pi} \frac{\epsilon 4r^2}{r^n}=\frac{2\epsilon}{\pi r^{n-2}}.\]
Since \(n\geq 2\) and \(\epsilon \) is arbitrary, we must have \(a_n=0\).
If \(f(0)\neq 0\), we can consider \(g(s)=f(s)-f(0)\), that satisfies the same growth condition, and conclude that \(f(s)=f(0)\). \(\blacksquare\)
The next result provides the asymptotic estimate needed to make use of Lemma
.
Its proof depends on the estimate given by Lemma
.
Lemma
Let \(\varepsilon \gt 0 \) be given. Then for all large enough \(|s-1/2|\),
Lemma B If \(|z|\leq 1/2\), then \(-\mbox{Re}\log (1-z) \leq 2|z|\).
\[-\log(1-z)=\int_0^z\frac{dw}{1-w}.\]
So
\[-\mbox{Re}\log(1-z)=\mbox{Re}\int_0^z\frac{dw}{1-w}\leq \left|\int_0^z\frac{dw}{1-w}\right|\leq |z|\max\left\{\frac{1}{|1-w|}: |w|\leq \frac{1}{2}\right\} \leq 2|z| \hspace{2ex}\blacksquare\] Lemma C
Suppose \(a_n\geq 0\), and \(\displaystyle \sum_{n=1}^\infty a_n\) converges. For \(t\gt 0\), let \(A(t)=\{n\in {\mathbb N}: a_n\leq t\}\).
Then \[\lim_{t\rightarrow 0^+} \sum_{n\in A(t)} a_n=0.\]
Let \(\epsilon \gt 0\). Find \(m\) such that \(\displaystyle \sum_{n=m}^\infty a_n\lt \epsilon\). If \(a_n=0\) for \(1\leq n \leq m\), then
\(\displaystyle \sum_{n\in A(t)}a_n \leq \sum_{n=m}^\infty a_n\lt \epsilon\). Otherwise, let \(\delta = \min\{a_n: a_n\gt 0, 1\leq n \leq m\}\).
Suppose \(t\lt \delta\). If \(n\leq m\), then either \(a_n=0\), or \(a_n\geq \delta \gt t\), and then \(n\not \in A(t)\), so
\[\sum_{n\in A(t)}a_n \leq \sum_{n=m}^\infty a_n \lt \epsilon \hspace{2ex} \blacksquare\]
Lemma D
Suppose \(z_n\) is such that for every \(\epsilon \gt 0\), \(\displaystyle \sum_{n=1}^\infty \frac{1}{|z_n|^{1+\epsilon}}\) converges. For \(R\gt 0\) and \(s\in {\mathbb C}\),
let
\[v_R(s)=\mbox{Re}\log \frac{1}{\displaystyle \prod_{|z_n|\geq 2R}\left(1-\frac{s^2}{z_n^2}\right)}.\]
Then given \(\epsilon \gt 0\), if \(|s|=R\) and \(R\) is large enough, \(v_R(s) \leq R^{1+\epsilon}\).
Let \(\epsilon\gt 0\), and suppose \(|s|=R\). Then
\begin{eqnarray*}
v_R(s)&=&-\mbox{Re} \sum_{|z_n|\geq 2R}\log\left(1-\frac{s^2}{z_n^2}\right)\\
&\leq &2 \sum_{|z_n|\geq 2R}\frac{R^2}{|z_n|^2} & \mbox{ from Lemma B}\\
&=& 2\sum_{|z_n|\geq 2R} \left(\frac{R}{|z_n|}\right)^{1-\epsilon}\left(\frac{R}{|z_n|}\right)^{1+\epsilon}\\
&\leq & 2\left(\frac{1}{2}\right)^{1-\epsilon} \sum_{|z_n|\geq 2R} \left(\frac{R}{|z_n|}\right)^{1+\epsilon}\\
&=&2^\epsilon R^{1+\epsilon} \sum_{|z_n|\geq 2R} \left(\frac{1}{|z_n|}\right)^{1+\epsilon}.
\end{eqnarray*}
From Lemma C, the sum will go to zero as \(R\rightarrow \infty\), so \(v_R(s) \leq R^{1+\epsilon}\) when \(|s|=R\) and \(R\) is large enough.\( \blacksquare\)
To prove Lemma ,
consider the function
\begin{equation} \label{F}
F(s)=\frac{\xi(s)}{P(s)}.
\end{equation}
For each zero \(\rho\) of \(\xi(s)\), the factor \(1-(s-1/2)/(\rho-1/2)=2(s-\rho)/(1-2\rho)\) in the product \(P(s)\)
will cancel the zero of \(\xi(s)\) at \(\rho\), and so \(F(s) \) is entire and without zeros.
Fix \(R\gt 0\), and write
\[A_R(s)=\prod_{|\rho-1/2|\leq 2R}\left(1-\left(\frac{s-1/2}{\rho-1/2}\right)^2\right),\]
\[B_R(s)=\prod_{|\rho-1/2|\gt 2R}\left(1-\left(\frac{s-1/2}{\rho-1/2}\right)^2\right),\]
so that \(P(s)=A_R(s)B_R(s)\). Then on the disk \(|s-1/2|\leq R\), \(\xi(s)/A_R(s)\) is analytic and has no zeros, because the zeros of \(\xi(s)\) are cancelled by those
of \(A_R(s)\) , and \(1/B_R(s)\) is analytic, because all the zeros of \(B_R\) are in \(|s-1/2|\gt 2R\). So we can consider
\[
u_R(s)=\mbox{Re} \log \frac{\xi(s)}{A_R(s)}=\log \frac{|\xi(s)|}{\displaystyle \prod_{|\rho-1/2|\leq 2R}\left|1-\frac{(s-1/2)^2}{(\rho-1/2)^2}\right|}.
\]
\[
v_R(s)=\mbox{Re} \log \frac{1}{B_R(s)}=\log \frac{1}{\displaystyle \prod_{|\rho-1/2|\gt 2R}\left|1-\frac{(s-1/2)^2}{(\rho-1/2)^2}\right|},
\]
both defined and harmonic on \(|s-1/2|\leq R\), and
\[u_R(s)+v_R(s)=\mbox{Re} \log F(s).\]
Now consider \(u_R(s)\) on \(|s-1/2|\leq 4R\). Then \(u_R\) is harmonic there except at the points \(s=\rho\) with \(2R\lt |\rho-1/2|\leq 4R\). But at those points \(u_R(s)\rightarrow
-\infty\). If we draw closed disks of small
radius \(\epsilon\)
around each point of the finite set of zeros of \(\xi(s)\) in the disk \(|s-1/2|\leq 4R\), then \(u_R\) is harmonic on
the same disk with those small disks removed, and so it will achieve
its maximum on the boundary. By taking the radius of the disks small enough, since
\(u\rightarrow -\infty\) on the boundary of each small disk, we conclude that
the maximum of \(u\) is achieved on the circle \(|s-1/2|= 4R\). But if
\(|s-1/2|=4R\), then for \(|\rho-1/2|\leq 2R\), we have
\[\left|1-\frac{s-1/2}{\rho-1/2}\right|\geq \left|\frac{s-1/2}{\rho-1/2}\right|-1
=\frac{4R}{|\rho-1/2|}-1\geq \frac{4R}{2R}-1=1.\]
So each factor in the denominator has absolute value at least \(1\), and
We now prove the same inequality for \(v_R(s)\).
By taking \(z_n=\rho-1/2\) in Lemma D, we find that for all \(\epsilon \gt 0\),
\begin{equation} \label{vR}
v_R(s)=\mbox{Re} \log \frac{1}{B_R(s)}=\log \frac{1}{\displaystyle \prod_{|\rho-1/2|\gt 2R}\left|1-\frac{(s-1/2)^2}{(\rho-1/2)^2}\right|}\leq R^{1+\epsilon}
\end{equation}
for \(|s-1/2|=R\) and all large \(R\).
Let now \(\epsilon \gt 0\) be given, and let \(F(s) =\xi(s)/P(s)\) as given in
(\ref{F}).
From (\ref{uR}) and
(\ref{vR}), we find \(\mbox{Re}\log F(s) =u_R(s)+v_R(s) \leq 2|s-1/2|^{1+\epsilon/2}\leq |s-1/2|^{1+\epsilon}\) for large
enough \(|s-1/2|\).
So we conclude that
\[
\mbox{Re} \log F(s) \leq \left|s-\frac{1}{2}\right|^{1+\epsilon}
\]
for all large enough \(|s-1/2|\). \(\blacksquare\)
We can now derive the product expansion for \(\xi(s)\) by applying Lemma
to the function \(F(s)=\xi(s)/P(s)\). This completes the proof of Theorem 3.
.
The integral \(I(b)\)
The next important step was the derivation of an analytic formula for \(\psi(x)\). This was
achieved by von Mangoldt in 1895. His strategy was to evaluate the integral
\[\frac{1}{2\pi i}\int_{a-i\infty}^{a+i\infty} \left(-\frac{\zeta'(s)}{\zeta(s)}\right)x^s\frac{ds}{s}\]
(where \(a\gt 1\))
in two different ways. The first is to substitute the expression for
\(-\zeta'(s)/\zeta(s)\) given by (\ref{int})
and obtain \(\psi(x)\). The second is to use the definition of \(\xi(s)\) and its product expansion to write
\begin{equation} \label{xiprod0} \xi(0)\prod_\rho \left(1-\frac{s}{\rho}\right)=\Pi\left(\frac{s}{2}\right)\pi^{-s/2} (s-1) \zeta(s),
\end{equation}
take the logarithmic derivative, and substitute the resulting expression for
\(-\zeta'(s)/\zeta(s)\) (given in the next proposition)
in the integral to obtain an analytic formula.
However,
von Mangoldt's derivation of the asymptotic behavior of \(\psi(x)\) for large \(x\) depends on
first deriving the asymptotic behavior of the antiderivative
\[\int_0^x\psi(t) dt.\]
To derive a formula for the antiderivative we will evaluate the integral
\[\frac{1}{2\pi i}\int_{a-i\infty}^{a+i\infty} \left(-\frac{\zeta'(s)}{\zeta(s)}\right)x^s\frac{ds}{s+1}\]
in a similar way.
In order to combine the two very similar derivations, we consider the integral
\begin{equation} \label{Ib} I(b)=\frac{1}{2\pi i}\int_{a-i\infty}^{a+i\infty} \left(-\frac{\zeta'(s)}{\zeta(s)}\right)x^s\frac{ds}{s+b},\end{equation}
where \(b\) is a non-negative real number.
.
The formula for \(-\zeta'(s)/\zeta(s)\)
Taking the logarithmic derivative of (\ref{xiprod0}) and using the product representation
(\ref{prodform})
for \(\Pi(s)\), we derive the formula for \(\zeta'(s)/\zeta(s)\) given in the next proposition.
Beside routine calculations, the proof only requires checking that the infinite series involved
can be differentiated termwise.
Proposition
Let \(b\) be a non-negative real number. Then
\begin{equation} \label{logderb} -\frac{\zeta'(s)}{\zeta(s)}=\sum_{n=1}^\infty \frac{s+b}{(2n-b)(2n+s)} +\frac{s+b}{(s-1)(1+b)} +\sum_\rho \frac{s+b}{(\rho+b)(\rho-s)} -\frac{\zeta'(-b)}{\zeta(-b)}.\end{equation}
Taking the logarithmic derivative of (\ref{xiprod0}), and using the product formula
(\ref{prodform}) for \(\Pi(s)\),
we find
\begin{equation} \label{logder}
-\frac{\zeta'(s)}{\zeta(s)}=\sum_{n=1}^\infty \left[\frac{1}{2}\log\left(1+\frac{1}{n}\right)-\frac{1}{2n+s}\right]-\frac{1}{2}
\log \pi+\frac{1}{s-1}+\sum_\rho\frac{1}{\rho -s}.
\end{equation}
Subtracting from
(\ref{logder})
the same equation evaluated at \(s=-b\), we find
(\ref{logderb}).
To justify the pointwise differentiation of infinite series,
it is enough to show that the two
infinite series in
(\ref{logder})
converge locally uniformly.
We find
\[\frac{1}{2}\log\left(1+\frac{1}{n}\right) -\frac{1}{2n+s} =\frac{1}{2n^2}\sum_{k=0}^\infty \frac{(-1)^k}{n^k}\left(\frac{s^{k+1}}{2^{k+1}}-\frac{1}{k+2}\right)\]
and so
\[\left|\frac{1}{2}\log\left(1+\frac{1}{n}\right) -\frac{1}{2n+s} \right| \leq \frac{c}{n^2}\]
for some constant \(c\), all \(s\) in a compact set, and all large enough \(n\). This proves that the pointwise differentiation
of the first series is justified.
For the second series, we need to pair terms with \(\rho\) and \((1-\rho)\). Then we find
\[\frac{1}{s-\rho}+\frac{1}{s-(1-\rho)} = \frac{2(s-1/2)}{(s-1/2)^2-(\rho-1/2)^2}\]
and so
\[\left|\frac{1}{s-\rho}+\frac{1}{s-(1-\rho)}\right|\leq \frac{K}{|\rho-1/2|^2}\]
for a constant \(K\), \(s\) in a compact set,
and \(|\rho -1/2|\) large enough. Then the series converges by
Lemma
in the proof of Theorem 3.
.
The formulas for \(\psi(x)\) and \(\int_0^x\psi(t)dt\)
The next theorem gives analytic formulas for \(\psi(x)\) and \(\int_0^x \psi(t)dt\).
If we ignore the fact that taking a limit inside an integral or an
infinite sum must be justified, the proof amounts to routine calculations obtained by
substituting (\ref{int}) and (\ref{logderb}) in (\ref{Ib}). In this respect, a substantial
part of the classical proof of the PNT could be described as the task of proving
that some limits can be switched.
Probably the most difficult part of the proof of this theorem is
proving that when substituting (\ref{logderb}) into (\ref{Ib}), the term of (\ref{logderb}) involving the roots \(\rho\) of \(\xi(s)\)
results in the limit
\begin{equation} \label{difflimit} \lim_{h\rightarrow \infty}\sum_{|{\rm Im}(\rho)|\leq h}\frac{x^\rho}{\rho+b}.
\end{equation}
This was done by von Mangoldt, and he was only able to prove
that the infinite series converges when the terms are arranged in increasing order of
\(|\mbox{Im}(\rho)|\).
We will often need to consider the integral
\[F_h(x,\beta)=\frac{1}{2\pi i}\int_{a-ih}^{a+ih}\frac{x^{s-\beta}}{s-\beta}ds\]
where \(a\gt 0\), \(x\gt 0\), \(h\gt 0\) and \(\beta\) is a complex number with
\(a\gt \mbox{Re}(\beta)\).
The next lemma describes the rate of convergence of \(F_h(x,0)\) as \(h\rightarrow \infty\),
and it will be used to justify taking limits inside integrals or infinite series.
These estimates are derived by evaluating the integral along
suitable large rectangles and making use of the residue theorem.
\[F_h(1,0)=
\frac{1}{2\pi i}\int_{a-ih}^{a+ih} \frac{ds}{s} = \frac{1}{2\pi}\int_{-h}^h \frac{a-it}{a^2+t^2}dt
= \frac{a}{2\pi}\int_{-h}^h \frac{dt}{a^2+t^2}=\frac{1}{\pi} \tan^{-1}\left(\frac{h}{a}\right)
.\]
If \(0\lt x \lt 1\), consider the integral over the rectangle shown:
\[\img{Ch3fig1.png}{}{}{16em}.\]
The integral over the closed path is zero, because the integrand is analytic inside. The integral over the top or bottom sides of the rectangle are easily seen to be bounded by
\[\frac{1}{2\pi}\frac{x^K-x^a}{h|\log x|},\]
and the integral on the right side is bounded by
\[\frac{1}{2\pi} \frac{x^K}{K}(2h).\]
Letting \(K\rightarrow \infty\), we find
\[
|F_h(x,0)|=\left|\frac{1}{2\pi i} \int_{a-ih}^{a+ih} \frac{x^s}{s} ds\right| \leq \frac{x^a}{\pi h |\log x|}.\]
If \(x\gt 1\), consider the integral over the rectangle:
\[\img{Ch3fig2.png}{}{}{16em}\]
The integrand has a simple pole at \(s=0\), and the integral over the closed path is \(1\). The integral over the top or bottom is bounded by
\[\frac{1}{2\pi}\frac{x^a-x^{-K}}{h\log x}\]
and the integral over the left side is bounded by
\[\frac{1}{2\pi}\frac{2h}{K x^K}.\]
Letting \(K\rightarrow \infty\), we find the bound
\[
\left|F_h(x,0)-1\right|=\left|\frac{1}{2\pi i}\int_{a-ih}^{a+ih} \frac{x^s}{s}ds -1 \right| \leq \frac{x^a}{\pi h \log x}. \blacksquare \]
The next two lemmas are simple consequences of the above estimates. They will be needed to
justify taking some limits inside integrals.
Lemma
\[
\mbox{If \(x \gt 1\) and \(a\gt \mbox{Re}(\beta)\), then } \hspace{2ex} \lim_{h\rightarrow \infty} F_h(x,\beta)=1.
\]
Consider the integral
\[\frac{1}{2\pi i}\int_{a-i\infty}^{a+i\infty} x^s\frac{ds}{s-\beta},\]
where \(x\gt 1 \), \(\beta=\sigma+i\tau\) and \(a\gt \sigma\).
Let \(t=s-\beta\). Then
\[\int_{a-ih}^{a+ih}x^s\frac{ds}{s-\beta} = x^\beta \int_{a-\sigma-i(h+\tau)}^{a-\sigma+i(h-\tau)}x^t\frac{dt}{t}
=x^\beta \int_{a-\sigma-i(h+\tau)}^{a-\sigma+i(h+\tau)}x^t\frac{dt}{t} +
x^\beta \int_{a-\sigma+i(h+\tau)}^{a-\sigma+i(h-\tau)}x^t\frac{dt}{t}.\]
The limit of the first term as \(h\rightarrow \infty\) is \(2\pi i x^\beta\),
by (\(*\)). For the second term, we find
\[\left|x^\beta \int_{a-\sigma+i(h+\tau)}^{a-\sigma+i(h-\tau)}x^t\frac{dt}{t}\right|\leq x^{a-\sigma}\max\left\{\frac{1}{\sqrt{(a-\sigma)^2+y^2}}: h-\tau\leq y \leq h+\tau \right\}2\tau \leq \frac{2\tau x^{a-\sigma}}{h-\tau} \rightarrow 0 \mbox{ as } h\rightarrow
\infty.\]
So we conclude that
\[\frac{1}{2\pi i}\int_{a-i\infty}^{a+i\infty}
x^s\frac{ds}{s-\beta} = x^\beta, \hspace{2ex} x\gt 1, \hspace{2ex} a\gt \mbox{Re}(\beta).
\blacksquare\]
Lemma
\[
\mbox{If \(x \gt 1, d\gt c \geq 0\), then } \hspace{5ex} \left|\frac{1}{2\pi i} \int_{a+ic}^{a+id}\frac{x^s}{s}ds\right| \leq K \frac{x^a}{(a+c)\log x},
\mbox{ where } K=\sqrt{2}\left(\frac{1}{\pi}+\frac{1}{4}\right).
\]
Integrating by parts we find
\begin{equation} \label{parts}
\int_{a+ic}^{a+id} \frac{x^s}{s}ds =\frac{x^{a+id}}{(a+id)\log x} - \frac{x^{a+ic}}{(a+ic)\log x} + \frac{ix^a}{\log x} \int_c^d\frac{x^{it}}{(a+it)^2}dt.
\end{equation}
Using the inequality
\[\sqrt{a^2+c^2}\geq \frac{a+c}{\sqrt{2}},\]
we find
\[|a+id|\geq |a+ic|\geq \frac{a+c}{\sqrt{2}}\]
and so the first two terms on the right side of
(\ref{parts}) are bounded by
\[\frac{x^a\sqrt{2}}{(a+c)\log x}.\]
The integral on the right side of (\ref{parts})
is bounded by
\[\left|\int_c^d\frac{x^{it}}{(a+it)^2}dt\right| \leq \int_0^\infty \frac{du}{a^2+(c+u)^2}\leq \int_0^\infty \frac{du}{a^2+c^2+u^2}=\frac{1}{\sqrt{a^2+c^2}}\frac{\pi}{2}
\leq \frac{\sqrt{2}}{a+c} \frac{\pi}{2}.\]
So we conclude that
\[ \left|\frac{1}{2\pi i} \int_{a+ic}^{a+id}\frac{x^s}{s}ds\right| \leq K \frac{x^a}{(a+c)\log x},\]
where
\[K=\sqrt{2}\left(\frac{1}{\pi}+\frac{1}{4}\right). \blacksquare\]
The next lemma is obtained by carrying out the first substitution of
\(-\zeta'(s)/\zeta(s)\) in \(I(b)\), as mentioned in item
,
using equation (\ref{int}).
Let \(x\gt 2\) be fixed, and \(m=\lceil x \rceil\), \(a\gt 1\). Then for each \(h\gt 0\),
\begin{array}{l}
\displaystyle \frac{1}{2\pi i}\int_{a-ih}^{a+ih}\left(-\frac{\zeta'(s)}{\zeta(s)}\right)x^s\frac{ds}{s+b}=\frac{1}{2\pi i}\int_{a-ih}^{a+ih}\sum_{n=2}^\infty \Lambda(n)
\frac{x^s}{n^s}\frac{ds}{s+b}\\
\displaystyle = \frac{1}{2\pi i}\int_{a-ih}^{a+ih}\sum_{n=2}^{m-1}\Lambda(n)\left(\frac{x}{n}\right)^s\frac{ds}{s+b}+\Lambda(m)\frac{1}{2\pi i}\int_{a-ih}^{a+ih}\left(\frac{x}{m}\right)^s\frac{ds}{s+b}
+\frac{1}{2\pi i}\int_{a-ih}^{a+ih}\sum_{n=m+1}^\infty \Lambda(n) \left(\frac{x}{n}\right)^s\frac{ds}{s+b}.
\end{array}
The first term involves a finite sum and its limit as \(h\rightarrow \infty\) is
\[\sum_{n=2}^{m-1}\Lambda(n) \frac{1}{2\pi i}\int_{a-i\infty}^{a+i\infty}\left(\frac{x}{n}\right)^s\frac{ds}{s+b} =
\sum_{n=2}^{m-1}\Lambda(n) \frac{n^b}{2\pi i x^b}\int_{a+b-i\infty}^{a+b+i\infty}\left(\frac{x}{n}\right)^t\frac{dt}{t}=\frac{1}{x^b}\sum_{n=2}^{m-1}n^b\Lambda(n) =
\frac{1}{x^b}\sum_{n\lt x}n^b\Lambda(n),\]
where we have used (\ref{main}).
In a similar way, the limit of the second term is
\[\Lambda(m) \frac{1}{2}[x=m].\]
So if \(b=0\), the sum of the first two terms gives
\[
\frac{1}{2}\sum_{n\lt x}\Lambda(n) + \frac{1}{2}\sum_{n\lt x}\Lambda(n) +\frac{1}{2}\Lambda(n)[n=x]=\frac{1}{2}\left(\sum_{n\lt x}\Lambda(n)+\sum_{n\leq x}\Lambda(n)\right) =\psi(x),
\]
while if \(b=1\) we get
\[\frac{1}{2x}\left(\sum_{n\lt x}n\Lambda(n)+\sum_{n\leq x}n\Lambda(n)\right)
\]
For the third term, since the series converges uniformly on any halfplane \(\mbox{Re}(s)\geq a \gt 1\), we can integrate termwise on finite paths. Using \(\Lambda(n)\leq \log(n)\)
and the estimate (\ref{xlt1})
we find
\[\left|\int_{a-ih}^{a+ih} \sum_{n=m+1}^\infty \Lambda(n) \left(\frac{x}{n}\right)^s\frac{ds}{s+b}\right|\leq \sum_{n=m+1}^\infty\Lambda(n) \frac{n^b}{x^b}\left|\int_{a+b-ih}^{a+b+ih}\left(\frac{x}{n}\right)^t\frac{dt}{t}\right|\leq \sum_{n=m+1}^\infty \Lambda (n)
\frac{x^a}{n^a \pi h \log(1+1/x)}\leq \frac{1}{h}\sum_{n=m+1}^\infty \frac{c}{n^a}\]
for some constant \(c\). So the limit of the third term is zero as \(h\rightarrow \infty\).
This proves that
\[ I(0)=\frac{1}{2\pi i}\int_{a-i\infty}^{a+i\infty} \left(-\frac{\zeta'(s)}{\zeta(s)}\right)x^s\frac{ds}{s}=\psi(x) = \sum_{n\lt x}\Lambda(n)\]
and
\[
I(1)=\frac{1}{2\pi i}\int_{a-i\infty}^{a+i\infty} \left(-\frac{\zeta'(s)}{\zeta(s)}\right)x^s\frac{ds}{s+1}= \frac{1}{x}\left(\sum_{n\lt x}n \Lambda(n)\right). \blacksquare
\]
The analytic formulas for \(\psi(x)\) and its antiderivative will now be derived by substituting formula (\ref{logderb}) into \(I(b)\). There are four terms in (\ref{logderb}).
The calculations, and the switching of infinite sums and
integral, will be relatively straightforward for all terms except the one involving the
sum of the roots \(\rho\). That term will produce the limit
\[\lim_{h\rightarrow \infty} \sum_\rho \frac{x^\rho}{\rho+b} F_h(x,\rho),\]
because the infinite series (that is locally uniformly convergent) can be integrated termwise
on finite paths. So we know that the limit exists, since all other terms in the formula
obtained by substituting (\ref{logderb}) into \(I(b)\) exist. But the difficult
part (as mentioned earlier)
will be in proving that the same limit is equal to
\[ \lim_{h\rightarrow \infty}\sum_{|{\rm Im}(\rho)\leq h}\frac{x^\rho}{\rho+b}.\]
One ingredient of the proof is the estimate of the
vertical density of the roots of \(\xi(s)\) given in the next lemma.
Note that this formula can be considered a
refinement of the formula for \(n(R)\) given in Lemma
(in the proof of Theorem 3),
because it is easy
to derive that formula if we have the vertical density estimate. But the proof of the
vertical density given in the next Lemma depends on the product formula for \(\xi(s)\), and
the formula for \(n(R)\) was needed in order to prove the product formula.
We will also need Stirling's
formula for \(\Pi(s)\), and in fact it will be necessary to use not just the dominant term, but also the fact that the relative error is of order \(1/|s|\) for large \(|s|\).
Lemma
Let \(D(T)\) be the number of roots of \(\xi(s)\)
with imaginary part between \(T\) and \(T+1\).
Then
\begin{equation} \label{DT}
D(T)\leq 2 \log T \hspace{5ex}\mbox{ for all large }T.
\end{equation}
We first prove a lemma that will be needed in order to put an upper bound on the number of roots \(\rho\) in a horizontal
strip of height 1.
Lemma E Let \(\rho=\sigma +i\tau\) be a root of \(\zeta(s)\) with \(0\leq \sigma \leq 1\), and let \(T\geq 0\). Then
If \(T\leq \tau \leq T+1\), then
\( \displaystyle \mbox{Im} \int_{2+iT}^{2+i(T+1)}\frac{ds}{s-\rho} \geq \tan^{-1}\left(\frac{1}{2}\right)\)
From a geometric point view, part a. follows from the fact that the left side is the angle \(\theta_2-\theta_1\) in
the following picture,
\[\img{Ch3p1.png}{}{}{12em}\]
and part b. from the fact that in case \(T\leq \tau \leq T+1\), the same angle is at least
the angle \(\varphi\) in the following picture.
\[\img{Ch3p2.png}{}{}{12em}\]
Let \(s=\rho+re^{i\theta}\). Then (using Cauchy's theorem to replace the segment \([2+iT,2+i(T+1)]\) with an
arc centered at \(\rho\)), we find
\[\mbox{Im} \int_{2+iT}^{2+i(T+1)}\frac{ds}{s-\rho} = \theta_2 -\theta_1,\]
where \[\tan \theta_1 =\frac{T-\tau}{2-\sigma} \lt \frac{T+1-\tau}{2-\sigma} = \tan \theta_2.\]
So \(\theta_1 \lt \theta_2\).
Let \(s=2+i(u+T)\), \(\rho=\sigma +i(b+T)\). Then \(0\leq \sigma \leq 1\), \(0\leq b \leq 1\), and
\begin{eqnarray*}
\mbox{Im} \int_{2+iT}^{2+i(T+1)}\frac{ds}{s-\rho}& =&\mbox{Im} \int_0^1\frac{idu}{2-\sigma+i(u-b)}\\[2ex]
&=&\mbox{Im } i \int_0^1\frac{2-\sigma -i(u-b)idu}{(2-\sigma)^2+(u-b)^2}\\[2ex]
&=&\int_0^1 \frac{2-\sigma}{(2-\sigma)^2+(u-b)^2} \\[2ex]
&=& \tan^{-1}\frac{b}{2-\sigma} -\tan^{-1}\frac{b-1}{2-\sigma}\\[2ex]
&=&\tan^{-1} \frac{2-\sigma}{(2-\sigma)^2+b(b-1)}
\end{eqnarray*}
Now the lemma follows from the fact that the inequality \(\sigma (2-\sigma) +b(1-b)\geq 0\) is equivalent to
\(\displaystyle \frac{2-\sigma}{(2-\sigma)^2+b(b-1)}\geq \frac{1}{2} \hspace{1ex}
\blacksquare \)
Since the Hadamard product
\[\xi(s)=\xi(0) \prod_\rho \left(1-\frac{s}{\rho}\right)\]
was shown to converge locally uniformly, we can take the logarithmic derivative and then integrate termwise
over finite paths. So using Lemma E, we find
\begin{equation} \label{Im}
\mbox{Im} \int_{2+iT}^{2+i(T+1)} \frac{\xi'(s)}{\xi(s)} ds = \mbox{Im} \sum_\rho \int_{2+iT}^{2+i(T+1)}\frac{ds}{s-\rho}
\geq \sum_{T\leq \textrm{Im}(\rho) \leq T+1}\int_{2+iT}^{2+i(T+1)}\frac{ds}{s-\rho} \geq D(T) \tan^{-1} \left(\frac{1}{2}\right).
\end{equation}
So we can get an upper bound for \(D(T)\) from an upper bound for \(\displaystyle \mbox{Im} \int_{2+iT}^{2+i(T+1)} \frac{\xi'(s)}{\xi(s)} ds\). This is done in the next lemma.
We denote by \(O(s^0)\) any function \(f(s)\) that remains bounded as \(|s|\rightarrow \infty\).
Lemma F For \(T\geq 0\), we have
\[\int_{2+iT}^{2+i(T+1)} \frac{\xi'(s)}{\xi(s)} ds = \frac{i}{2}\log T +O(T^0).\]
We will use the Stirling's approximation
\[\Pi(s) = s^se^{-s} \sqrt{2\pi s} \left(1+\frac{O(s^0)}{s}\right).\]
From the definition of \(\xi(s)\), we find
\[
\xi(s) = \prod_{k=0}^5f_k(s)
\]
where
\begin{eqnarray*}
f_0(s)&=&\left(\frac{s}{2}\right)^{s/2}\\[1ex]
f_1(s)&=& (e\pi)^{-s/2}\\[1ex]
f_2(s)&=& \sqrt{\pi s}\\[1ex]
f_3(s)&=& (s-1)\\[1ex]
f_4(s)&=& \zeta(s)\\[1ex]
f_5(s)&=& \left(1+\frac{O(s^0)}{s}\right)
\end{eqnarray*}
For any non-zero differentiable function \(f(s)\), let \(\displaystyle L(f)(s)=\frac{f'(s)}{f(s)}\). Then
\(L(fg)=L(f)+L(g)\). So we find
\[L(\xi)(s))=\sum_{k=0}^5L(f_k)(s),\]
and we need to show that
\[\sum_{k=0}^5\int_{2+iT}^{2+i(T+1)} L(f_k)(s)ds = \frac{i}{2}\log T +O(T^0).\]
We will show that all terms in the sum with \(k\gt 0\) are bounded as \(T\rightarrow \infty\), while the term with \(k=0\) is
\((i/2) \log T +O(T^0)\).
For \(k=1\), we have
\[L\left((e\pi)^{-s/2}\right)=-\frac{1}{2}\log(e\pi)\]
and so
\[\int_{2+iT}^{2+i(T+1)} L(f_1)(s)ds=-\frac{i}{2}\log(e\pi)\]
For \(k=2\),
\[L(\sqrt{\pi s}) = \frac{1}{2s},\]
and
\[\int_{2+iT}^{2+i(T+1)} L(f_2)(s)ds=\frac{1}{2}\log\left(1+\frac{i}{2+iT}\right) = O(T^0).\]
For \(k=3\),
\[L(s-1)=\frac{1}{s-1},\]
and
\[\int_{2+iT}^{2+i(T+1)} L(f_3)(s)ds=\log\left(1+\frac{i}{1+iT}\right) = O(T^0).\]
For \(k=4\),
\[L(\zeta(s))=\frac{d}{ds}\log \zeta(s)\]
and for \(\mbox{Re}(s)=2\),
\[|\log\zeta(2+it)|=\left|\sum_p\sum_{n=1}^\infty\frac{1}{n}\frac{1}{p^{2n+nit}}\right|
\leq \sum_p\sum_{n=1}^\infty \frac{1}{n}\frac{1}{p^{2n}}=\log \zeta(2),\]
so
\[\int_{2+iT}^{2+i(T+1)} L(f_4)(s)ds = O(T^0).\]
For \(k=5\),
\[\int_{2+iT}^{2+i(T+1)} L(f_5)(s)ds =\left.\log\left(1+\frac{O(s^0)}{s}\right)\right|_{2+iT}^{2+i(T+1)}=O(T^0).\]
For \(k=0\),
\[\int L\left(\frac{s}{2}\right)^{s/2}=\frac{1}{2}s \log\left(\frac{s}{2}\right)=\frac{1}{2}s\log s-\frac{1}{2}s \log 2,\]
and
\[\left.\frac{s \log 2}{2}\right|_{2+iT}^{2+i(T+1)}=O(T^0).\]
So it remains to show that
\[\left.\frac{1}{2}s\log s\right|_{2+iT}^{2+i(T+1)} = \frac{1}{2} i\log T + O(T^0).\]
This is easily seen from the computation
\begin{eqnarray*}
(2+i(T+1))\log(2+i(T+1))-(2+iT)\log(2+iT)&=& (2+iT)\log(2+i(T+1))+i \log(2+i(T+1))-(2+iT)\log(2+iT)\\[1ex]
&=&(2+iT)\log\left(1+\frac{i}{2+iT}\right)+i\log(2+i(T+1)) \\[1ex]
&=& O(T^0) + i\log(2+i(T+1))\\[1ex]
&=& i\log T +O(T^0) \hspace{1ex} \blacksquare
\end{eqnarray*}
Combining Lemma F and (\ref{Im}),
we find that for large \(T\),
\[D(T)\leq \frac{\log T}{2 \tan^{-1}(1/2)} +c\leq 1.1 \log T +c\]
for some constant \(c\), and so
\[
D(T)\leq 2 \log T \mbox{ for all large } T \hspace{1ex} \blacksquare
\]
The last step in the proof of Theorem 4 is the derivation of the term that results from the
sum involving the roots of \(\xi(s)\) when substituting (\ref{logderb}) into \(I(b)\).
The main idea in von Mangoldt's proof is to split the difference
\[ \sum_\rho \frac{x^\rho}{\rho+b} F_h(x,\rho)- \sum_{|{\rm Im}(\rho)|\leq h}\frac{x^\rho}{\rho+b}\]
as
\begin{equation} \label{split}
\left(\sum_\rho \frac{x^\rho}{\rho+b} F_h(x,\rho)-\sum_{|{\rm Im}(\rho)|\leq h} \frac{x^\rho}{\rho+b}
F_h(x,\rho)\right) + \left(\sum_{|{\rm Im}(\rho)|\leq h} \frac{x^\rho}{\rho+b}
F_h(x,\rho)- \sum_{|{\rm Im}(\rho)|\leq h}\frac{x^\rho}{\rho+b}\right)
\end{equation}
and then use the estimate for \(D(T)\) to prove that the differences in both parentheses
go to zero as \(h\rightarrow \infty\). The term that is added and subtracted is what Edwards
calls a "diagonal" term.
Note that \(|F_h(\rho)|=|F_h(\bar{\rho})|\). Write \(\rho=\beta+i\gamma\).
We will consider the first two and the last two terms of
(\ref{split})
separately.
For the first two terms, we find
\begin{eqnarray*}
&& \left|\sum_\rho \frac{x^\rho}{\rho+b} F_h(\rho)-\sum_{{\rm |Im(\rho)|} \leq h} \frac{x^\rho}{\rho+b} F_h(\rho)\right|=\left|\sum_{{\rm |Im(\rho)|} \gt h} \frac{x^\rho}{\rho+b} F_h(\rho)\right|\\
&\leq &\sum_{\gamma \gt h} \frac{x^\beta}{\gamma}\left|F_h(\beta+i\gamma)\right|
+\sum_{\gamma \lt -h} \frac{x^\beta}{|\gamma|}\left|F_h(\beta+i\gamma)\right|\\
&=&\sum_{\gamma \gt h} \frac{x^\beta}{\gamma}\left|F_h(\beta+i\gamma)\right|+
\sum_{\gamma \gt h} \frac{x^\beta}{\gamma}\left|F_h(\beta-i\gamma)\right|\\
&=&2\sum_{\gamma \gt h} \frac{x^\beta}{\gamma}\left|F_h(\beta-i\gamma)\right|
\end{eqnarray*}
Since \(0\leq \mbox{Re}(\rho) \leq 1\), and \(a\gt 1\), we have \(a-\beta\geq a-1 \gt 0\).
Using Lemma ,
we have, for some constant \(K\),
\[|F_h(\beta-i\gamma)|=\left|\frac{1}{2\pi}\int_{a-ih}^{a+ih}\frac{x^{s-\beta+i\gamma}}{s-\beta+i\gamma}ds\right|
=\left|\frac{1}{2\pi}\int_{a-\beta+i(\gamma-h)}^{a-\beta+i(\gamma+h)}\frac{x^t}{t}dt\right|
\leq \frac{K x^{a-\beta}}{(a-\beta+\gamma -h)\log x}\leq \frac{K x^{a-\beta}}{(a-1+\gamma-h)\log x}\]
and so
\[\sum_{\gamma\gt h}\frac{x^\beta}{\gamma} |F_h(\beta-i\gamma)|\leq
\frac{Kx^a}{\log x}\sum_{\gamma\gt h}\frac{1}{\gamma (a-1+\gamma-h)}.\]
Let \(T\gt 0\) be large enough so that (according to Lemma
) \(D(h)\leq 2\log h\) for \(h\geq T\).
By increasing \(T\) if necessary, we can also assume that \(\log(T+j)\leq (T+j)^{1/2}\) for all \(j\geq 0\). Then
if \(h\geq T\), we have
\begin{eqnarray*}
&&\sum_{\gamma\gt h}\frac{1}{\gamma (a-1+\gamma-h)}=\sum_{j=0}^\infty\sum_{h+j\lt \gamma \leq h+j+1}
\frac{1}{\gamma (a-1+\gamma-h)}\leq \sum_{j=0}^\infty \frac{2\log(h+j)}{(h+j)(a-1+j)}\\
&&\leq 2\sum_{j=0}^\infty\frac{(h+j)^{1/2}}{(h+j)(j+a-1)}\leq 2\sum_{j=0}^\infty\frac{1}{(h+j)^{1/4} (h+j)^{1/4}(j+a-1)}
\leq \frac{2}{h^{1/4}}\sum_{j=0}^\infty\frac{1}{(h+j)^{1/4}(j+a-1)}.
\end{eqnarray*}
The last infinite series converges, and so we conclude that
\[\lim_{h\rightarrow \infty}
\left|\sum_\rho \frac{x^\rho}{\rho+b} F_h(\rho)-\sum_{{\rm |Im(\rho)|} \leq h} \frac{x^\rho}{\rho+b} F_h(\rho)\right|=0.\]
Before considering the last two terms of
(\ref{split}),
we prove a lemma. We will use the fact that the harmonic numbers
\(\displaystyle H_n=\sum_{k=1}^n 1/k\) are of order \(\log n\) as \(n\rightarrow \infty\), so that we can write
\(H_n=\log n (1+\epsilon_n)\) where \(\epsilon_n\rightarrow 0\) as \(n\rightarrow \infty\).
Lemma G Denote by \(\gamma\) the imaginary part of a root of \(\xi(s)\). Then
\begin{equation} \label{pta}
\lim_{h\rightarrow \infty} \frac{1}{h}\sum_{0\lt \gamma \leq h}\frac{1}{\gamma} =0
\end{equation}
and if \(c\gt 0\),
\begin{equation} \label{ptb}
\lim_{h\rightarrow \infty} \frac{1}{h}\sum_{0\lt \gamma \leq h}\frac{1}{c+h-\gamma} =0
\end{equation}
Let \(N\) be an integer large enough so that (according to Lemma
)
the number of roots in \([N+j,N+j+1]\) is at most
\(2\log(N+j)\), for all \(j\geq 0\). We then find
\[\frac{1}{h}\sum_{0\lt \gamma \leq h} \frac{1}{\gamma} = \frac{1}{h}\sum_{0\lt \gamma \leq N} \frac{1}{\gamma}
+\frac{1}{h}\sum_{N\lt \gamma \leq h} \frac{1}{\gamma}\]
and the limit of the first (finite) sum is \(0\) as \(h\rightarrow \infty\). For the second sum, we find
\begin{eqnarray*}
&&\frac{1}{h}\sum_{N\lt \gamma \leq h} \frac{1}{\gamma}= \frac{1}{h}\sum_{j=0}^{\lfloor h-N\rfloor} \sum_{N+j\lt \gamma
\leq N+j+1} \frac{1}{\gamma} \leq \frac{1}{h}\sum_{j=0}^{\lfloor h-N\rfloor}\frac{2\log(N+j)}{N+j}\\
&&\leq \frac{2\log h}{h} \sum_{j=0}^{\lfloor h \rfloor} \frac{1}{N+j}\leq \frac{2\log h}{h} H_{N+\lfloor h \rfloor}
=\frac{2\log h}{h}\log(N+ h)(1+\epsilon_h).
\end{eqnarray*}
Since \((\log h)^2/h \rightarrow 0\) as \(h\rightarrow \infty\), this proves
(\ref{pta}).
To prove (\ref{ptb}),
it is enough to consider (as for (\ref{pta})) the sum for \(N\lt \gamma \leq h\). We find
\[\sum_{N\lt \gamma \leq h}\frac{1}{c+h-\gamma}=\sum_{j=0}^{\lfloor h-N\rfloor-1}
\sum_{N+j\lt \gamma \leq N+j+1}\frac{1}{c+h-\gamma}+\sum_{N+\lfloor h-N\rfloor -1\lt \gamma \leq h}\frac{1}{c+h-\gamma}.\]
The second sum is at most \((2\log h)/c\). The first sum is at most
\[\sum_{j=0}^{\lfloor h-N\rfloor-1} \frac{2\log h}{c+h-N-j-1}\leq 2 \log h\left(\frac{1}{c} +H_{\lfloor h-N\rfloor}\right)
\leq 2\log h\left( \frac{1}{c} +\log h (1+\epsilon_h)\right).\]
This proves (\ref{ptb})
\(\blacksquare\)
We now consider the last two terms of (\ref{split}). We have
\begin{eqnarray*}
F_h(\rho)-1 &=&\frac{1}{2\pi i}\int_{a-ih}^{a+ih}\frac{x^{s-\rho}}{s-\rho}ds -1 =
\frac{1}{2\pi i} \int_{a-\beta -i \gamma -ih}^{a-\beta -i\gamma +ih}\frac{x^t}{t}dt -1\\[2ex]
&=&\frac{1}{2\pi i} \int_{a-\beta -i (\gamma +h)}^{a-\beta +i(\gamma +h)}\frac{x^t}{t}dt -1-
\frac{1}{2\pi i} \int_{a-\beta +i(h-\gamma)}^{a-\beta +i(h+\gamma)}\frac{x^t}{t}dt.
\end{eqnarray*}
Using the estimate
(\ref{xgt1})
and
Lemma ,
we find
\begin{eqnarray*}
&&\frac{x^\beta}{\gamma}|F_h(\rho)-1|\leq \frac{x^a}{\pi\log x}\frac{1}{\gamma(\gamma +h)} +
\frac{Kx^a}{\log x}\frac{1}{\gamma(c+h-\gamma)}\\[2ex]
&&=\frac{x^a}{\pi h\log x}\left(\frac{1}{\gamma}-\frac{1}{\gamma +h}\right)+\frac{Kx^a}{(c+h)\log x}\left(\frac{1}{\gamma}+
\frac{1}{c+h-\gamma}\right)\\[2ex]
&&\leq \left(\frac{x^a}{\pi \log x}+\frac{Kx^a}{\log x}\right) \frac{1}{h\gamma} +\frac{Kx^a}{\log x}\frac{1}{h(c+h-\gamma)}.
\end{eqnarray*}
So using
(\ref{pta})
and
(\ref{ptb}) we find
\[\lim_{h\rightarrow \infty} \sum_{|{\rm Im}(\gamma)|\leq h} \frac{x^\rho}{\rho+b}(F_h(\rho)-1) =0.\]
This completes the proof of Lemma \(\blacksquare\).
We are now ready to substitute (\ref{logderb}) into (\ref{Ib}).
Since the series in (\ref{logderb}) converge locally uniformly, we
can integrate them termwise over finite paths and
we find
\begin{eqnarray*}
&&\frac{1}{2\pi i}\lim_{h\rightarrow \infty}\int_{a-ih}^{a+ih} \left(-\frac{\zeta'(s)}{\zeta(s)}\right)x^s\frac{ds}{s+b}\\
&=& \frac{1}{2\pi i}\lim_{h\rightarrow \infty}\int_{a-ih}^{a+ih}
\left(\frac{s+b}{(s-1)(1+b)} -\frac{\zeta'(-b)}{\zeta(-b)}+\sum_{n=1}^\infty \frac{s+b}{(2n-b)(2n+s)} +\sum_\rho \frac{s+b}{(\rho+b)(\rho-s)} \right)x^s\frac{ds}{s+b}\\
&=&\frac{1}{2\pi i(1+b)}\int_{a-i\infty}^{a+i\infty}\frac{x^s}{s-1}ds+\frac{1}{2\pi i}\int_{a-i\infty}^{a+i\infty}\left(- \frac{\zeta'(-b)}{\zeta(-b)}\right)x^s \frac{ds}{s+b}\\
&+&\lim_{h\rightarrow \infty}\sum_{n=1}^\infty \frac{1}{(2n-b)}\frac{1}{2\pi i}\int_{a-ih}^{a+ih}\frac{x^s}{s+2n}ds
+\lim_{h\rightarrow \infty}\sum_\rho\frac{1}{(\rho+b)}\frac{1}{2\pi i}\int_{a-ih}^{a+ih} \frac{x^s}{\rho-s} ds\\
&=& \frac{x}{1+b}- x^{-b}\frac{\zeta'(-b)}{\zeta(-b)}+\lim_{h\rightarrow \infty}\sum_{n=1}^\infty \frac{1}{(2n-b)}\frac{1}{2\pi i}\int_{a-ih}^{a+ih}\frac{x^s}{s+2n}ds+
\lim_{h\rightarrow \infty}\sum_\rho\frac{1}{(\rho+b)}\frac{1}{2\pi i}\int_{a-ih}^{a+ih} \frac{x^s}{\rho-s} ds
\end{eqnarray*}
where we have used Lemma to evaluate the first two terms. We now need to show that we can switch the limit with the two infinite series.
First consider the series in \(n\). We will show that the it converges uniformly in \(h\).
We find
\[\int_{a-ih}^{a+ih}\frac{x^{s+2n}}{s+2n}ds=\int_0^h\left(\frac{x^{a+2n+iy}}{a+2n+iy}+
\frac{x^{a+2n-iy}}{a+2n-iy}\right)dy=2\mbox{Re}\int_0^h\frac{x^{a+2n+iy}}{a+2n+iy}dy
=2\mbox{Re}\int_{a+2n}^{a+2n+ih}\frac{x^t}{t}dt\]
and so using Lemma ,
\[\left|\int_{a-ih}^{a+ih}\frac{x^{s+2n}}{s+2n}ds\right|\leq 2 \left|\int_{a+2n}^{a+2n+ih}
\frac{x^t}{t}dt\right|\leq \frac{K x^{a+2n}}{(a+2n)\log x}\]
for a constant \(K\). It follows that the \(n\)-th term of the series in \(n\) is bounded by \(c/n^2\) for a constant \(c\) independent of \(h\). So the series converges uniformly in \(h\), and it follows from
Lemma that
\[\lim_{h\rightarrow \infty}\sum_{n=1}^\infty \frac{1}{2n-b}\frac{1}{2\pi i}\int_{a-ih}^{a+ih}\frac{x^s}{s+2n}ds=\sum_{n=1}^\infty \frac{1}{2n-b}\frac{1}{2\pi i}\int_{a-i\infty}^{a+i\infty}\frac{x^s}{s+2n}ds=\sum_{n=1}^\infty \frac{x^{-2n}}{2n-b}.\]
So we have now shown that
\begin{equation} \label{exist}
I(b)=\frac{x}{1+b}- x^{-b}\frac{\zeta'(-b)}{\zeta(-b)}+\sum_{n=1}^\infty \frac{x^{-2n}}{2n-b}-
\lim_{h\rightarrow \infty}\sum_\rho\frac{x^\rho}{\rho+b}F_h(x,\rho)
\end{equation}
and using Lemma
we obtain the formula for \(I(b)\)
\[I(b)=\frac{x}{1+b}- x^{-b}\frac{\zeta'(-b)}{\zeta(-b)}+\sum_{n=1}^\infty \frac{x^{-2n}}{2n-b}-
\lim_{h\rightarrow \infty}\sum_{|{\rm Im}(\rho)|\leq h}\frac{x^\rho}{\rho+b}. \]
Setting \(b=0\) we find, according to
(\ref{I0}), van Mangoldt's formula for \(\psi(x)\)
\[
I(0)=\psi(x)=x-\frac{\zeta'(0)}{\zeta(0)}+\sum_{n=1}^\infty \frac{x^{-2n}}{2n}
-\lim_{h\rightarrow \infty}\sum_{|{\rm Im}(\rho)\leq h}\frac{x^\rho}{\rho}
=x-\frac{\zeta'(0)}{\zeta(0)}+\frac{1}{2}\log\left(\frac{x^2}{x^2-1}\right)
-\lim_{h\rightarrow \infty}\sum_{|{\rm Im}(\rho)|\leq h}\frac{x^\rho}{\rho}
\]
while setting \(b=1\) we find, according to
(\ref{I1}),
\[
I(1)=\frac{1}{2x}\left(\sum_{n\lt x}n \Lambda(n)+\sum_{n\leq x}
n \Lambda(n)\right)=
\frac{x}{2}- \frac{\zeta'(-1)}{x\zeta(-1)} +
\sum_{n=1}^\infty \frac{x^{-2n}}{2n-1}-
\lim_{h\rightarrow \infty}\sum_{|{\rm Im}(\rho)|\leq h}\frac{x^\rho}{\rho+1}
\]
and the formula for the antiderivative of \(\psi(x)\) follows by computing \(xI(0)-xI(1)\). This completes the proof of Theorem 4.
.
\(\mbox{Re}(\rho) \lt 1\)
The last significant step in the derivation of the asymptotic behavior of \(\psi(x)\)
(and so the proof of the PNT) is the
proof that \(\zeta(s)\) has no zeros on the line \(\mbox{Re}(s)=1\). This was proved by both Hadamard and de la Vallée Poussin independently in 1896.
Theorem
\begin{equation}\label{rholt1}
\mbox{If \(\rho \) is a zero of \(\zeta(s)\), then \(\mbox{Re}(\rho)\lt 1\)}
\end{equation}
Let \(\displaystyle F(s) = \sum_p \frac{1}{p^s}\), for \(\mbox{Re}(s)\gt 1\). We claim that
\(F(\sigma) +\log (\sigma -1)\) is bounded as \(\sigma \rightarrow 1^+\).
Using the representation
\[\Pi(s) =\lim_{n\rightarrow \infty}\frac{n! (n+1)^s}{(s+1)\cdots (s+n)}\]
we find
\[\lim_{s\rightarrow 1} (s-1) \Pi(-s)=-1.\]
Since
\[\frac{1}{2\pi i}\int_C\frac{-z}{e^z-1}\frac{dz}{z} =-1\]
(where \(C\) is the Hankel path from Chapter 1), we find from the integral representation
\[\zeta(s)=\frac{\Pi(-s)}{2\pi i}\int_C\frac{(-z)^s}{e^z-1}\frac{dx}{z} \]
that
\[\lim_{s\rightarrow 1} (s-1)\zeta(s) =1.\]
As a
Let \(p_1,p_2,p_3,\ldots\)
be the prime numbers, and \(P_m=\{p_i: 1\leq i \leq m\}\). Then
\[\log\sum_{k\not \in P_n}\frac{1}{k^s}=\sum_{i=1}^n\log\left(1-\frac{1}{p_i^s}\right)^{-1}, s\gt 1.\]
The sum on the right is
\[\sum_{i=1}^n\sum_{m=1}^\infty \frac{1}{m}\frac{1}{p_i^{ms}}=\sum_{i=1}^n\frac{1}{p_i^s}+\sum_{i=1}^n\sum_{m=2}^\infty \frac{1}{m}\frac{1}{p_i^{ms}}\]
and the second sum on the right is bounded by
\[\sum_{r=2}^{p_n}\sum_{m=2}^\infty \frac{1}{m}\frac{1}{r^{ms}}\leq \frac{1}{2}\sum_{r=2}^{p_n}\sum_{m=2}^\infty \left(\frac{1}{r^s}\right)^m=\frac{1}{2}\sum_{r=2}^{p_n}\frac{1}{r^{2s}}\frac{1}{1-r^{-s}}
=\frac{1}{2}\sum_{r=2}^\infty\frac{1}{r^s(r^s-1)}\leq \frac{1}{2}.\]
So letting \(n\rightarrow \infty\) we see that
\[\log\zeta(s)=\sum_{i=1}^\infty\frac{1}{p_i^s}+B(s) \mbox{ for } s\gt 1,\]
where \(|B(s)|\leq 1/2\) for \(s\geq 1\). Hence the sum \(\displaystyle \sum_{i=1}^\infty \frac{1}{p_i^s}\) diverges as \(s\rightarrow 1^+\) like \(\displaystyle \log\zeta(s)\).
we find
\[\log \zeta(s) =F(s) + B(s), |B(s)|\leq \frac{1}{2} \mbox{ for Re}(s)\gt 1.\]
So
\[\log((\sigma -1)\zeta(\sigma))=\log(\sigma -1) +F(\sigma) +B(\sigma) \rightarrow 0 \mbox{ as } \sigma\rightarrow 1^+\]
and the claim follows.
Suppose now \(t\neq 0\) is such that \(\zeta(1+it)=0\). Let \(\delta\gt 0\) be
such that \(1-\cos\delta \gt 0\) and \(\cos(2\delta)\gt 0\). Consider the set
\[A=\{p \mbox{ prime}: |(2n+1)\pi -t \log p|\lt \delta \mbox{ for some } n\in {\mathbb Z}\}\]
Let
\[F_1(\sigma) = \sum_{p\in A}\frac{1}{p^\sigma}, \hspace{2ex} F_2(\sigma) = \sum_{p\not \in A}\frac{1}{p^\sigma}, \hspace{2ex} \sigma \gt 1.\]
Since \(\zeta(s)/(s-1-it)\) is analytic, using \(s=\sigma +it\) we find that
\(\zeta(\sigma +it)/(\sigma -1)\) is bounded as \(\sigma \rightarrow 1^+\), and so
\(\mbox{Re}\log \zeta(\sigma +it) -\log(\sigma -1) \) is bounded. This implies
\[\mbox{Re} F(\sigma +it) -\log(\sigma -1)=\sum_p\frac{\cos(t \log p)}{p^\sigma} -\log(\sigma -1) \mbox{ is bounded as } \sigma\rightarrow 1^+.\]
If \(p \not \in A\), then \(\cos(t \log p)\geq 1-\cos(\delta)\), and so
\[\sum_p\frac{\cos(t \log p)}{p^\sigma}\geq -\sum_{p\in A}\frac{1}{p^\sigma}-\cos(\delta)
\sum_{p\not \in A}\frac{1}{p^\sigma} = -F_1(\sigma) -\cos(\delta) F_2(\sigma).\]
Hence
\begin{equation} \label{1} -F_1(\sigma) -\cos(\delta) F_2(\sigma)-\log(\sigma -1) \leq \sum_p\frac{\cos(t \log p)}{p^\sigma} -\log(\sigma -1) \mbox{ is bounded as }\sigma\rightarrow 1^+.\end{equation}
According to the claim,
\begin{equation} \label{2}
F(\sigma)+\log(\sigma-1) =F_1(\sigma) + F_2(\sigma) +\log(\sigma -1) \mbox{ is bounded as }
\sigma\rightarrow 1^+ .
\end{equation}
Adding (\ref{1}) and (\ref{2}), we find
\[(1-\cos(\delta))F_2(\sigma) \mbox{ is bounded as } \sigma\rightarrow1^+.\]
So we conclude that \(F_2(\sigma)\) is bounded as \(\sigma \rightarrow 1^+\).
But \(\zeta(s)\) is analytic at \(s=1+2it\), and so \(\mbox{Re}F(\sigma +2it)\) is bounded
as \(\sigma \rightarrow 1^+\). If \(p\in A\), then \(|2(2n+1)\pi -2t\log p|\lt 2\delta\)
for some \(n\), and so \(\cos(2t\log p)\geq \cos(2\delta)\). So
\[\mbox{Re} F(\sigma +2it) \geq \cos(2\delta) F_1(\sigma) -F_2(\sigma)\]
implies that \(F_1(\sigma)\) is also bounded as \(\sigma\rightarrow 1^+\). But this contradicts
the fact that \(F(\sigma) = F_1(\sigma) + F_2(\sigma)\) is unbounded as \(\sigma \rightarrow 1^+\blacksquare \).
.
The asymptotic estimate for \(\int_0^x \psi(t)dt\)
From (\ref{antider}) and (\ref{rholt1}), it is easy to derive the asymptotic estimate for the antiderivative of
\(\psi(x)\).
From (\ref{antider}), we find
\[\frac{1}{x^2}\left(\int_0^x\psi(t)dt-\frac{x^2}{2}\right)=
-\frac{\zeta'(0)}{x \zeta(0)} +\frac{\zeta'(-1)}{x^2\zeta(-1)}
-\sum_{n=1}^\infty \frac{x^{-2n-1}}{2n(2n-1)}
-\lim_{h\rightarrow \infty}
\sum_{|{\rm Im}(\rho)\leq h}\frac{x^{\rho-1}}{\rho(\rho+1)}
\]
Since this last series in \(\rho\) now
converges absolutely and uniformly in \(x\), we can evaluate
the limit as \(x\rightarrow \infty\) termwise and since \(\mbox{Re}(\rho)-1 \lt 0\),
we find
\[\lim_{x\rightarrow \infty} \frac{1}{x^2}\left(\int_0^x\psi(t)dt-\frac{x^2}{2}\right)=0.\blacksquare\]
.
The asymptotic estimate for \(\psi(x)\)
The asymptotic behavior of \(\psi(x)\) (and hence the Prime Number Theorem) now follows
from the following general result on the antiderivative of a positive, increasing function.
Proposition
Suppose that \(f\) is positive and increasing, and
\[\int_0^xf(t)dt \sim \frac{x^2}{2} \hspace{2ex} \mbox{as } x\rightarrow \infty.\]
Then
\[f(x)\sim x \hspace{2ex} \mbox{as } x\rightarrow \infty.\]
Write \(\displaystyle F(x)=\int_0^x f(t) dt\).
Let \(\epsilon \gt 0\) be given. Then for large enough \(x\),
\[(1-\epsilon)\frac{x^2}{2} \leq F(x) \leq (1+\epsilon) \frac{x^2}{2}.\]
If \(x\) is large enough and \(y=\beta x\), where \(\beta \gt 1\), we have
\[(1-\epsilon)\frac{y^2}{2} \leq f(y)(y-x)+(1+\epsilon)\frac{x^2}{2}=(y-x)f(y)+(1-\epsilon)
\frac{x^2}{2} + \epsilon x^2,\]
and then we find
\[\frac{1}{2}(1-\epsilon)(y^2-x^2)\leq (y-x)f(y) +\epsilon x^2\]
and using \(y=\beta x\),
\[\frac{f(y)}{y}\geq \frac{1}{2}(1-\epsilon)\frac{\beta +1}{\beta}-\frac{\epsilon}{\beta(\beta-1)}.\]
Hence
\[\liminf_{x\rightarrow \infty} \frac{f(x)}{x} \geq \frac{1}{2}(1-\epsilon)\frac{\beta +1}{\beta}-\frac{\epsilon}{\beta(\beta-1)}.\]
Letting \(\epsilon \) go to zero, we find
\[\liminf_{x\rightarrow \infty} \frac{f(x)}{x} \geq\frac{\beta+1}{2 \beta},\]
and letting \(\beta \) approach 1,
\[\liminf_{x\rightarrow \infty} \frac{f(x)}{x}\geq 1.\]
In a similar way, we find
\[\frac{f(x)}{x} \leq \frac{1}{2} (1-\epsilon)(\beta+1) +\epsilon \frac{\beta^2}{\beta-1}\]
and so
\[\limsup_{x\rightarrow \infty} \frac{f(x)}{x}
\leq \frac{1}{2} (1-\epsilon)(\beta+1) +\epsilon \frac{\beta^2}{\beta-1}.\]
Letting \(\epsilon \) go to zero, we find
\[\limsup_{x\rightarrow \infty}\frac{f(x)}{x} \leq \frac{\beta+1}{2},\]
and then letting \(\beta \) approach 1,
\[\limsup_{x\rightarrow \infty}\frac{f(x)}{x} \leq 1\leq \liminf_{x\rightarrow \infty} \frac{f(x)}{x}.\]
So we conclude that
\[\lim_{x\rightarrow \infty}\frac{f(x)}{x} =1.\blacksquare \]
References
[1] C. de la Vallée Poussin, Recherches analytiques sur la théorie des nombres
premiers, Annales de la Société scientifique de Bruxelles,
Imprimeur de l'Académie Royale de Belgique, (1896) 20 B, 21 B: 183--256, 281--352, 363--397, 351--368.
[3] P. Erdös, On a new method in elementary number theory which leads to an
elementary proof of the prime number theorem, Proceedings of the National Academy of Sciences,
National Academy of Sciences, (1949), 35 (7): 374--384.
[4] J. Hadamard, Sur la distribution des zèros de la fonction \(\zeta(s)\) et ses
conséquences arithmétiques. Bulletin de la Société Mathématique de France, Société
Mathématique de France, (1896) 24: 199–220.
[5] D. J. Newman, Simple analytic proof of the prime number theorem,
American Mathematical Monthly, (1980), 87 (9): 693--696.
[6] F. Richter, A new elementary proof of the Prime Number Theorem, Bulletin of the London Mathematical Society, Volume 53 (2021), Issue 5, pp. 1365-1375.
[7] B. Riemann, Ueber die Anzahl der Primzahlen unter einer gegebenen Grösse,
Monatsberichte der Berliner Akademie, November 1859. (An English translation appears as an Appendix in [2]).
[8] A. Selberg, An Elementary Proof of the Prime-Number Theorem,
Annals of Mathematics, (1949), 50 (2): 305--313,
[9] D. Zagier, Newman's short proof of the prime number theorem,
American Mathematical Monthly, (1997), 104 (8): 705--708.