Skip to main content

Notes on complex analysis

Section 3.3 The prime number theorem

Subsection 3.3.1 Preliminaries

Our first proof of the prime number theorem will run through a complex analytic argument due to Newman. The particular paper I'm referencing here is a very nice exposition of Newman's paper by D. Zagier. There is a wonderful stroll through the Zagier paper in Helgason's notes for a complex analysis class on MIT's OCW. Tao's notes follow a similar line, but works through a different set of functions.
We aim to prove the following description of the asymptotic density of the primes in the integers. Let \(\pi(x)\) denote the number of primes less than or equal to \(x\text{.}\) We will use the notation \(f \sim g\) (that is, \(f\) and \(g\) are asymptotically equal) to mean \(\lim_{x \to \infty} f(x)/g(x) = 1\text{.}\) We will also use the order notation \(g = \mathrm{O}(f)\) as \(x \to \infty\) to mean that there exists some \(C>0\) so that \(\abs{g(x)} \leq Cf(x)\) for all sufficiently large \(x\text{.}\) (Every order argument we make in this section will be at infinity, so we may sometimes just write \(g = \mathrm{O}(x)\text{.}\))
Let us introduce our dramatis personae. Our primary character is, of course, the zeta function
\begin{equation*} \zeta(s) = \sum_{n=1}^\infty \frac{1}{n^s}. \end{equation*}
From the logarithmic derivative (\(f\dd/f\)) of \(\zeta\text{,}\) we'll consider the function
\begin{equation*} \Phi(s) = \sum_p \frac{\log p}{p^s}. \end{equation*}
Finally, our asymptotic prime checking will be done by the function
\begin{equation*} \vtheta(s) = \sum_{p \leq x} \log p. \end{equation*}
Essentially, the argument goes by showing that there can be no zeros of \(\zeta\) on the line \(\RE(s) = 1\text{,}\) and then using that fact to force an asymptotic estimate on \(\vtheta\text{.}\) The steps to the proof are as follows:
  1. Product formula.
    Prove the product formula for \(\zeta\) for \(\RE s > 1\text{.}\) That is,
    \begin{equation*} \zeta(s) = \prod_{p} (1 - p^{-s})\inv. \end{equation*}
  2. Continuation across the critical strip.
    Show that the function
    \begin{equation*} \zeta(s) - \frac{1}{s -1} \end{equation*}
    has an analytic continuation to \(\RE s > 0\text{.}\)
  3. Order asymptotic for \(\vtheta\).
    Show that
    \begin{equation*} \vtheta(x) = \mathrm{O}(x). \end{equation*}
  4. No zeros on \(\RE s = 1\).
    Show that
    \begin{equation*} \zeta(s) \neq 0 \text{ on } \RE s \geq 1 \end{equation*}
    and that
    \begin{equation*} \Phi(s) - \frac{1}{s-1} \text{ is holomorphic on } \RE s \geq 1. \end{equation*}
  5. Integral estimate.
    Show that
    \begin{equation*} \int_1^\infty \frac{\vtheta(x) - x}{x^2} \, dx \text{ converges.} \end{equation*}
  6. Exact asymptotic for \(\vtheta\).
    Show that
    \begin{equation*} \vtheta(x) \sim x. \end{equation*}
  7. Prime number theorem.
    Conclude that
    \begin{equation*} \pi(x) \sim \frac{x}{\log x}. \end{equation*}
We will prove these as a series of theorems
This theorem was proved in the earlier section.
The idea is to write the function as a power series and show that it converges absolutely by comparison. Both functions are analytic on \(\RE s > 1\text{,}\) so we'll use that as a basis for our series. In order to leverage bounding in integral, note that we can write \(1/(s-1)\) in a sort of continuous \(p\)-series form via
\begin{equation*} \frac{1}{s-1} = \int_1^\infty \frac{1}{x^s}\, dx. \end{equation*}
Then,
\begin{align*} \zeta(s) - \frac{1}{s-1} \amp= \sum_{n=1}^\infty \frac{1}{n^s} - \int_1^\infty \frac{1}{x^s}\, dx\\ \amp= \sum_{n=1}^\infty \int_n^{n+1} (\frac{1}{n^s} - \frac{1}{x^s})\, dx. \end{align*}
The terms \(1/n^s - 1/x^s\) themselves look like the output of a definite integral - indeed,
\begin{equation*} \int_n^x \frac{s}{u^{s+1}}\, dx = \frac{1}{n^s} - \frac{1}{x^s}. \end{equation*}
Then, invoking the ML inequality,
\begin{align*} \abs{ \int_n^{n+1} (\frac{1}{n^s} - \frac{1}{x^s})\, dx} \amp= \abs{s \int_n^{n+1} \int_n^x \frac{1}{u^{s+1}}\, du dx}\\ \amp\leq \max_{n \leq u \leq n+1} \abs{\frac{s}{u^{s+1}}}\\ \amp= \frac{\abs{s}}{n^{\RE s + 1}}. \end{align*}
Thus, while we defined the series for \(\RE s > 1\text{,}\) the terms of the series will converge absolutely for \(\RE s > 0\text{,}\) (for example by successive application of the \(M\)-test on an appropriate family of half-disks).
The argument is combinatorial-esque. For \(n \in \mathbb{N}\text{,}\) we have
\begin{align*} 2^{2n} = (1+1)^{2n} \amp= \binom{2n}{0} + \ldots \binom{2n}{2n} \\ \amp\geq \binom{2n}{n}\\ \amp\geq \prod_{n \leq p \leq 2n} p\\ \amp= \exp(\vtheta(2n) - \vtheta(n)). \end{align*}
This implies that
\begin{equation*} \vtheta(x) - \vtheta(\frac{x}{2}) \leq C x \end{equation*}
for any \(C > \log 2\) for all \(x > x_0\) for some \(x_0\) whose choice depends on \(C\text{.}\) Now, given some large value of \(x\text{,}\) consider the sequence
\begin{equation*} x, \frac{x}{2}, \frac{x}{4}, \ldots, \frac{x}{2^r} \geq x_0 \geq \frac{x}{2^{r+1}}.\text{.} \end{equation*}
The corresponding inequalities are
\begin{align*} \vtheta(x) - \vtheta(\frac{x}{2}) \amp\leq Cx\\ \vtheta(\frac{x}{2}) - \vtheta(\frac{x}{4}) \amp\leq C\frac{x}{2}\\ \vdots\\ \vtheta(\frac{x}{2^{r}})- \vtheta(\frac{x}{2^r+1}) \amp\leq C\frac{x}{2^{r}} \end{align*}
Adding these results in the inequality
\begin{equation*} \vtheta(x) - \vtheta(\frac{x}{2^{r+1}}) \leq (1 + \frac{1}{2} + \ldots + \frac{1}{2^{r}})Cx \leq 2Cx, \end{equation*}
and so, noting that \(\vtheta(\frac{x}{2^{r+1}}) \leq \vtheta(x_0)\text{,}\) we get
\begin{equation*} \vtheta(x) \leq 2Cx + \vtheta(x_0), \end{equation*}
or in order notation
\begin{equation*} \vtheta(x) = \mathrm{O}(x) \text{ as } x\to \infty \end{equation*}
as \(\vtheta(x_0) = \mathrm{O}(1)\text{.}\)
We now arrive at the crucial step, which the assertion that \(\zeta\) has no zeros on the line \(\RE s = 1\text{.}\)
Start with the assumption that \(\RE s > 1\text{.}\) Now compute the logarithmic derivative of \(\zeta\text{:}\)
\begin{align*} \frac{\zeta\dd}{\zeta} \amp= \frac{d}{ds} \log \zeta(s) = \frac{d}{ds} \log \prod_p (1 - p^{-s})\inv\\ \amp= \frac{d}{ds} \sum_p \log (1-p^{-s})\inv\\ \amp= -\sum_p \frac{d}{ds} \log (1-p^{-s})\\ \amp= -\sum_p \frac{\log p} {p^s(1-p^{-s})}\\ \amp= -(\sum_p \frac{\log p}{p^s} + \sum_p \frac{\log p}{p^s(p^s - 1)}).\\ \amp= -(\Phi(s) + \sum_p \frac{\log p}{p^s(p^s - 1)}) \end{align*}
This gives the identity
\begin{equation*} -\frac{\zeta\dd}{\zeta} = \Phi(s) + \sum_p\frac{\log p}{p^s(p^s - 1)}. \end{equation*}
Notice that the series on the right hand side converges for \(\RE s > \frac{1}{2}\text{.}\) From Step II above, we know that \(\zeta - \frac{1}{s-1}\) is holomorphic on \(\RE s > 0\text{,}\) and so
\begin{equation*} \zeta = (\zeta -\frac{1}{s-1}) + \frac{1}{s-1} \end{equation*}
is meromorphic on \(\RE s \geq 0\) with a simple pole at \(s = 1\text{.}\) Likewise, \(\zeta\dd\) will be meromorphic with a pole at \(s = 1\) of order 2. Then \(\zeta\dd/\zeta\) is also a meromorphic function on \(\RE s > 0\) with poles at \(s = 1\) and at the zeros of \(\zeta\text{,}\) which implies for \(\Phi\text{,}\) which is
\begin{equation} \Phi = -\frac{\zeta\dd}{\zeta} - \sum_p\frac{\log p}{p^s(p^s - 1)},\tag{3.3.1} \end{equation}
that we have a meromorphic extension of \(\Phi\) to \(\RE s > \frac{1}{2}\) with poles at \(s = 1\) and the zeros of \(\zeta\text{.}\)
Now let's analyze the zeros. First, because \(\zeta(s) = \cc{\zeta(\cc s)}\text{,}\) we see that if \(s\) is zero of \(\zeta\) then so is \(\cc{s}\text{.}\) Now let \(a\in \R\text{.}\) If \(s_0 = 1 + ia\) is a zero of \(\zeta\) of order \(\mu \geq 0\text{,}\) then \(\zeta = (s - s_0)^\mu h(s)\) for some holomorphic function \(h\) that is non-zero near \(s_0\text{.}\) Computing the logarithmic derivative gets us
\begin{equation*} -\frac{\zeta\dd}{\zeta} = -\frac{\mu}{s - s_0} + H(s) \end{equation*}
for some function \(H = \log h\) holomorphic near \(s_0\text{.}\) The function \(\Phi(s_0 + \eps)\) then has a simple pole at \(\eps = 0\text{,}\) and by the residue theorem
\begin{equation*} \lim_{\eps \to 0} \eps \Phi(1 + \eps + ia) = -\mu. \end{equation*}
Now, consider
\begin{equation*} \Phi(1+\eps) = \sum_p \frac{\log p}{p^{1 + \eps}}, \end{equation*}
and note that the terms are positive for \(\eps > 0\text{.}\) Then
\begin{equation*} \sum_p \frac{\log p}{p^{1 + \eps}}(p^{\frac{ia}{2}} + p^{\frac{-ia}{2}})^2 \geq 0. \end{equation*}
Multiplying this out gives
\begin{equation*} \Phi(1 + \eps -ia) + 2\Phi(1 + \eps) + \Phi(1 + \eps + ia) \geq 0. \end{equation*}
From Part II, as noted above we have that \(s = 1\) is a simple pole of \(-\zeta\dd/\zeta\) with residue 1, and so
\begin{equation*} \lim_{\eps \to 0} \eps \Phi(1_+ \eps) = 1. \end{equation*}
Then
\begin{align*} \amp\lim_{\eps \to 0} \eps(\Phi(1 + \eps -ia) + 2\Phi(1 + \eps) + \Phi(1 + \eps + ia))\\ \amp= -2\mu + 2 \geq 0 \end{align*}
and so \(0 \leq \mu \leq 1\text{.}\) As we intend to show that \(\mu = 0\text{,}\) we need more. So assume as well that \(\zeta\) has a zero at \(s_1 = 1 \pm i2a\) with order \(\nu \geq 0\text{.}\) (That is, we are admitting the possibility that there is no zero at those points.) Again, note that
\begin{equation*} \lim_{\eps \to 0} \eps \Phi(1 + \eps \pm i2a) = -\nu. \end{equation*}
Then the same idea applied to
\begin{equation*} \sum_p \frac{\log p}{p^{1 + \eps}}(p^{\frac{ia}{2}} + p^{\frac{-ia}{2}})^4 \geq 0 \end{equation*}
produces the equation
\begin{equation*} 6 - 8 \mu - 2 \nu \geq 0. \end{equation*}
This requires that \(\mu = 0\) since \(\mu, \nu \geq 0\text{.}\) We have shown that if \(\zeta\) has a zero of the form \(s_0 = 1 + ia\text{,}\) then it must have order 0, and hence not be a zero. That is, \(\zeta\) has no zeros on the line \(\RE s > 1\text{.}\)
Finally, recall that the poles of \(\Phi\) are \(s = 1\) and the zeros of \(\zeta\text{,}\) none of which can lie on \(\RE s = 1\text{.}\) If we trace through the form of \(\Phi\) from (3.3.1) and the implications of Step II, we conclude that \(\Phi - \frac{1}{s-1}\) is holomorphic on \(\RE s \geq 1\text{.}\)
In the next step, we hone our understanding of the behavior of \(\vtheta\text{.}\)
We'll use Stieltjes integrals and their properties. Assume that \(\RE s > 1\text{.}\) First, note that \(\vtheta\) is a step function increasing with a step at each new prime number \(p\text{.}\) Then we can take the view that
\begin{equation*} \Phi(s) = \sum_{p} \frac{\log p}{p^s} \, ds = \int_1^\infty \frac{d\vtheta(s)}{x^s}. \end{equation*}
Now, applying the integration by parts formula for Stieltjes integrals, we get
\begin{equation*} \int_1^\infty \frac{d\vtheta(s)}{x^s} = \int_1^\infty \vtheta(x) \,d(x^{-s}) = s\int_1^\infty \frac{\vtheta(x)}{x^{s+1}}\, dx \end{equation*}
and finally under the change of variables \(x = e^t\text{,}\) we get
\begin{equation} \Phi(s) = s\int_0^\infty \vtheta(e^t) e^{-st}\, dt.\tag{3.3.2} \end{equation}
We're setting up to invoke a sort of dominated integral theorem from Newman's original paper, so we now introduce functions \(f\) and \(g\text{.}\) Define the functions
\begin{align*} f(t) \amp= \vtheta(e^t)e^{-t} - 1\\ g(s) \amp= \frac{\Phi(s+1)}{s+1} - \frac{1}{s} \end{align*}
By Step III, \(\vtheta\) is \(O(x)\text{,}\) and so \(f\) is bounded. Proceeding formally for a moment, (that is, ignoring issues of convergence, which will be pinned down shortly) the same change of variables from above will give
\begin{equation*} \int_1^\infty \frac{\vtheta(x) - x}{x^2} \, dx = \int_0^\infty f(t) \, dt. \end{equation*}
The function \(\Phi(s) - 1/(s-1)\) is holomorphic on \(\RE s \geq 1\) by Step IV, and so
\begin{equation*} \Phi(s + 1) = h(s) + \frac{1}{s} \end{equation*}
for some holomorphic function \(h\) on \(\RE z \geq 0\text{.}\) As a consequence,
\begin{equation*} g(s) = \frac{\Phi(s+1)}{s+1} - \frac{1}{s} = \frac{h(s) -1}{s+1} \end{equation*}
is holomorphic on \(\RE s \geq 0\text{.}\) On the other hand, for \(\RE s > 0\text{,}\) we have by (3.3.2)
\begin{align*} g(s) \amp= \int_0^\infty e^{-st}(f(t) + 1)\, dt - \int_0^\infty e^{-st}\, dt\\ \amp= \int_0^\infty e^{-st}f(t)\, dt. \end{align*}
All of the pieces are set up. We now need Newman's Analytic Theorem (which we defer the proof of until later). Notice that \(f\) and \(g\) as we have defined them satisfy the hypotheses of the lemma, and so
\begin{equation*} \int_1^\infty \frac{\vtheta(x) - x}{x^2} \, dx = \int_0^\infty f(t) \, dt \text{ converges}, \end{equation*}
which establishes the theorem.
We'll proceed by cases on the assumption that \(\vtheta(x) \nsim x\text{.}\) First, assume that there exists \(\la > 1\) so that \(\vtheta(x) > \la x\) for sufficiently large \(x\text{.}\) Given such an \(x\text{,}\) as \(\vtheta\) is an increasing function,
\begin{equation*} \int_x^{\la x} \frac{\vtheta(t) - t}{t^2} \, dt \geq \int_x^{\la x} \frac{\la x - t}{t^2}\, dt = \int_1^\la \frac{\la - s}{s^2} \, ds > 0, \end{equation*}
where this constant does not depend on \(x\text{.}\) However, the convergence in Step V means that for all \(\eps > 0\text{,}\) there exists some \(M\) for which \(M_1, M_2 \geq M\) we have
\begin{equation*} \int_{M_1}^{M_2} \frac{\vtheta(x) - x}{x^2} \, dx \lt \eps, \end{equation*}
and so such a \(\la\) cannot exist.
Likewise, if there exists \(\la \lt 1\) with \(\vtheta(x) \leq \la x\) for sufficiently large \(x\text{,}\) then for \(t \leq x,\)
\begin{equation*} \vtheta(t) \leq \vtheta(x) \leq \la x, \end{equation*}
and so
\begin{equation*} \int_{\la x}^{x} \frac{\vtheta(t) - t}{t^2} \, dt \leq \int_{\la x}^{ x} \frac{\la x - t}{t^2}\, dt = \int_\la^1 \frac{\la - s}{s^2} \, ds \lt 0, \end{equation*}
which leads to the same contradiction. Thus,
\begin{equation*} \limsup_{x \to \infty} \frac{\vtheta(x)}{x} \leq 1 \leq \liminf_{x\to \infty} \frac{\vtheta(x)}{x}, \end{equation*}
which means that \(\vtheta(x) \sim x\text{.}\)
First,
\begin{equation*} \vtheta(x) = \sum_{p \leq x} \log p \leq \sum_{p \leq x} \log x = \pi(x) \log x. \end{equation*}
Thus,
\begin{equation*} \liminf_{x \to \infty} \frac{\pi(x) \log x}{x} \geq \liminf_{x \to \infty} \frac{\vtheta(x)}{x} = 1. \end{equation*}
In the other direction, for \(0 \lt \eps \lt 1\text{,}\)
\begin{align*} \vtheta(x) \amp\geq \sum_{x^{1 - \eps} \leq p \leq x} \log p\\ \amp\geq (1 - \eps) \sum_{x^{1 - \eps} \leq p \leq x} \log x\\ \amp= (1-\eps) \log x (\pi(x) + O(x^{1-\eps})). \end{align*}
Then for each \(\eps\text{,}\)
\begin{equation*} \limsup_{x \to \infty} \frac{\pi(x) \log x}{x} \leq \frac{1}{1 - \eps} \limsup_{x \to \infty} \frac{\vtheta(x)}{x}. \end{equation*}
We conclude that
\begin{equation*} \lim_{x \to \infty} \frac{\pi(x) \log x}{x} = 1. \end{equation*}

Subsection 3.3.2 The analytic theorem

As is typical in mathematics, the technical lemma that drove our proof above, Newman's so-called Analytic Theorem, is the key to the entire enterprise. Indeed, according to J. Korevaar's On Newman's quick way to the prime number theorem 1 , the simplicity of Newman's approach boils down to his proof of that theorem. Between the 1930s and the 1980s, the standard proof of the prime number theorem went by way of Fourier analysis, and a method due to Weiner called Tauberian integrals 2 . In fact, the Analytic Theorem appears in a paper by Ingham from the 1930s with a different proof and more general statement. We restate Newman's result here.
Korevaar points out that this is related to a very useful theorem called the Ikehara-Weiner theorem 3 . (The structure of this should remind you of the functions we used in the proof of the prime number theorem.)
Korevaar notes that Newman's Analytic Theorem has as a consequence the following “poor-man's Ikehara-Weiner Theorem”.
Notice that the content of the above theorem is essentially Steps IV-V of Newman's argument. The ideas that we're tossing around here are really ideas in analytic number theory, where complex analysis is used to analyze objects like Dirichlet series 4  via the theory of generating functions. Whereas the Ikehara-Weiner theorem is quite difficult, the proof of the Analytic Theorem requires little more than Cauchy's theorem and careful choice of contour.
I'm going to follow Korevaar's exposition here, which is quite nice.
Figure 3.3.13. Illustration of main contour
By assumption, the Laplace transform of \(f\) given by
\begin{equation*} g(s) = \int_0^\infty e^{-st} f(t)\, dt \end{equation*}
extends holomorphically to \(\RE s \geq 0\text{.}\) Choose some radius \(R\) and then choose a small \(\delta\) so that \(g\) is analytic on and inside the contour \(C_1 \cup C_2\text{.}\) Set the function
\begin{equation*} g_T(z) = \int_0^T e^{-st} f(t)\, dt, \end{equation*}
which is holomorphic in \(\C\) (say by Morera's theorem). We will show that
\begin{equation*} \lim_{T \to \infty} g_T(0) = \int_0^\infty f(t)\, dt = g(0). \end{equation*}
As \(g, g_T\) are analytic on and across \(C\text{,}\) the Cauchy Integral Formula gives us
\begin{equation*} g(0) - g_T(0) = \frac{1}{2\pi i}\int_C \frac{g(z) - g_T(z)}{z} \, dz. \end{equation*}