You are currently browsing the category archive for the ‘246B – complex analysis’ category.
Previous set of notes: Notes 3. Next set of notes: 246C Notes 1.
One of the great classical triumphs of complex analysis was in providing the first complete proof (by Hadamard and de la Vallée Poussin in 1896) of arguably the most important theorem in analytic number theory, the prime number theorem:
Theorem 1 (Prime number theorem) Letdenote the number of primes less than a given real number
. Then
(or in asymptotic notation,
as
).
(Actually, it turns out to be slightly more natural to replace the approximation in the prime number theorem by the logarithmic integral
, which happens to be a more precise approximation, but we will not stress this point here.)
The complex-analytic proof of this theorem hinges on the study of a key meromorphic function related to the prime numbers, the Riemann zeta function . Initially, it is only defined on the half-plane
:
Definition 2 (Riemann zeta function, preliminary definition) Letbe such that
. Then we define
Note that the series is locally uniformly convergent in the half-plane , so in particular
is holomorphic on this region. In previous notes we have already evaluated some special values of this function:
The Riemann zeta function has several remarkable properties, some of which we summarise here:
Theorem 3 (Basic properties of the Riemann zeta function)
- (i) (Euler product formula) For any
with
, we have
where the product is absolutely convergent (and locally uniform in
) and is over the prime numbers
.
- (ii) (Trivial zero-free region)
has no zeroes in the region
.
- (iii) (Meromorphic continuation)
has a unique meromorphic continuation to the complex plane (which by abuse of notation we also call
), with a simple pole at
and no other poles. Furthermore, the Riemann xi function
is an entire function of order
(after removing all singularities). The function
is an entire function of order one after removing the singularity at
.
- (iv) (Functional equation) After applying the meromorphic continuation from (iii), we have
for all
(excluding poles). Equivalently, we have
for all
. (The equivalence between the (5) and (6) is a routine consequence of the Euler reflection formula and the Legendre duplication formula, see Exercises 26 and 31 of Notes 1.)
Proof: We just prove (i) and (ii) for now, leaving (iii) and (iv) for later sections.
The claim (i) is an encoding of the fundamental theorem of arithmetic, which asserts that every natural number is uniquely representable as a product
over primes, where the
are natural numbers, all but finitely many of which are zero. Writing this representation as
, we see that
The claim (ii) is immediate from (i) since the Euler product is absolutely convergent and all terms are non-zero.
We remark that by sending to
in Theorem 3(i) we conclude that
The meromorphic continuation (iii) of the zeta function is initially surprising, but can be interpreted either as a manifestation of the extremely regular spacing of the natural numbers occurring in the sum (1), or as a consequence of various integral representations of
(or slight modifications thereof). We will focus in this set of notes on a particular representation of
as essentially the Mellin transform of the theta function
that briefly appeared in previous notes, and the functional equation (iv) can then be viewed as a consequence of the modularity of that theta function. This in turn was established using the Poisson summation formula, so one can view the functional equation as ultimately being a manifestation of Poisson summation. (For a direct proof of the functional equation via Poisson summation, see these notes.)
Henceforth we work with the meromorphic continuation of . The functional equation (iv), when combined with special values of
such as (2), gives some additional values of
outside of its initial domain
, most famously
From Theorem 3 and the non-vanishing nature of , we see that
has simple zeroes (known as trivial zeroes) at the negative even integers
, and all other zeroes (the non-trivial zeroes) inside the critical strip
. (The non-trivial zeroes are conjectured to all be simple, but this is hopelessly far from being proven at present.) As we shall see shortly, these latter zeroes turn out to be closely related to the distribution of the primes. The functional equation tells us that if
is a non-trivial zero then so is
; also, we have the identity
Conjecture 4 (Riemann hypothesis) All the non-trivial zeroes oflie on the critical line
.
This conjecture would have many implications in analytic number theory, particularly with regard to the distribution of the primes. Of course, it is far from proven at present, but the partial results we have towards this conjecture are still sufficient to establish results such as the prime number theorem.
Return now to the original region where . To take more advantage of the Euler product formula (3), we take complex logarithms to conclude that
The series and
that show up in the above formulae are examples of Dirichlet series, which are a convenient device to transform various sequences of arithmetic interest into holomorphic or meromorphic functions. Here are some more examples:
Exercise 5 (Standard Dirichlet series) Letbe a complex number with
.
- (i) Show that
.
- (ii) Show that
, where
is the divisor function of
(the number of divisors of
).
- (iii) Show that
, where
is the Möbius function, defined to equal
when
is the product of
distinct primes for some
, and
otherwise.
- (iv) Show that
, where
is the Liouville function, defined to equal
when
is the product of
(not necessarily distinct) primes for some
.
- (v) Show that
, where
is the holomorphic branch of the logarithm that is real for
, and with the convention that
vanishes for
.
- (vi) Use the fundamental theorem of arithmetic to show that the von Mangoldt function is the unique function
such that
for every positive integer
. Use this and (i) to provide an alternate proof of the identity (8). Thus we see that (8) is really just another encoding of the fundamental theorem of arithmetic.
Given the appearance of the von Mangoldt function , it is natural to reformulate the prime number theorem in terms of this function:
Theorem 6 (Prime number theorem, von Mangoldt form) One has(or in asymptotic notation,
as
).
Let us see how Theorem 6 implies Theorem 1. Firstly, for any , we can write
Exercise 7 Show that Theorem 1 conversely implies Theorem 6.
The alternate form (8) of the Euler product identity connects the primes (represented here via proxy by the von Mangoldt function) with the logarithmic derivative of the zeta function, and can be used as a starting point for describing further relationships between and the primes. Most famously, we shall see later in these notes that it leads to the remarkably precise Riemann-von Mangoldt explicit formula:
Theorem 8 (Riemann-von Mangoldt explicit formula) For any non-integer, we have
where
ranges over the non-trivial zeroes of
with imaginary part in
. Furthermore, the convergence of the limit is locally uniform in
.
Actually, it turns out that this formula is in some sense too precise; in applications it is often more convenient to work with smoothed variants of this formula in which the sum on the left-hand side is smoothed out, but the contribution of zeroes with large imaginary part is damped; see Exercise 22. Nevertheless, this formula clearly illustrates how the non-trivial zeroes of the zeta function influence the primes. Indeed, if one formally differentiates the above formula in
, one is led to the (quite nonrigorous) approximation
Comparing Theorem 8 with Theorem 6, it is natural to suspect that the key step in the proof of the latter is to establish the following slight but important extension of Theorem 3(ii), which can be viewed as a very small step towards the Riemann hypothesis:
Theorem 9 (Slight enlargement of zero-free region) There are no zeroes ofon the line
.
It is not quite immediate to see how Theorem 6 follows from Theorem 8 and Theorem 9, but we will demonstrate it below the fold.
Although Theorem 9 only seems like a slight improvement of Theorem 3(ii), proving it is surprisingly non-trivial. The basic idea is the following: if there was a zero at , then there would also be a different zero at
(note
cannot vanish due to the pole at
), and then the approximation (9) becomes
In fact, Theorem 9 is basically equivalent to the prime number theorem:
Exercise 10 For the purposes of this exercise, assume Theorem 6, but do not assume Theorem 9. For any non-zero real, show that
as
, where
denotes a quantity that goes to zero as
after being multiplied by
. Use this to derive Theorem 9.
This equivalence can help explain why the prime number theorem is remarkably non-trivial to prove, and why the Riemann zeta function has to be either explicitly or implicitly involved in the proof.
This post is only intended as the briefest of introduction to complex-analytic methods in analytic number theory; also, we have not chosen the shortest route to the prime number theorem, electing instead to travel in directions that particularly showcase the complex-analytic results introduced in this course. For some further discussion see this previous set of lecture notes, particularly Notes 2 and Supplement 3 (with much of the material in this post drawn from the latter).
Previous set of notes: Notes 2. Next set of notes: Notes 4.
On the real line, the quintessential examples of a periodic function are the (normalised) sine and cosine functions ,
, which are
-periodic in the sense that
What about periodic functions on the complex plane? We can start with singly periodic functions which obey a periodicity relationship
for all
in the domain and some period
; such functions can also be viewed as functions on the “additive cylinder”
(or equivalently
). We can rescale
as before. For holomorphic functions, we have the following characterisations:
Proposition 1 (Description of singly periodic holomorphic functions)In both cases, the coefficients
- (i) Every
-periodic entire function
has an absolutely convergent expansion
where
is the nome
, and the
are complex coefficients such that
Conversely, every doubly infinite sequence
of coefficients obeying (2) gives rise to a
-periodic entire function
via the formula (1).
- (ii) Every bounded
-periodic holomorphic function
on the upper half-plane
has an expansion
where the
are complex coefficients such that
Conversely, every infinite sequence
obeying (4) gives rise to a
-periodic holomorphic function
which is bounded away from the real axis (i.e., bounded on
for every
).
can be recovered from
by the Fourier inversion formula
for any
in
(in case (i)) or
(in case (ii)).
Proof: If is
-periodic, then it can be expressed as
for some function
on the “multiplicative cylinder”
, since the fibres of the map
are cosets of the integers
, on which
is constant by hypothesis. As the map
is a covering map from
to
, we see that
will be holomorphic if and only if
is. Thus
must have a Laurent series expansion
with coefficients
obeying (2), which gives (1), and the inversion formula (5) follows from the usual contour integration formula for Laurent series coefficients. The converse direction to (i) also follows by reversing the above arguments.
For part (ii), we observe that the map is also a covering map from
to the punctured disk
, so we can argue as before except that now
is a bounded holomorphic function on the punctured disk. By the Riemann singularity removal theorem (Exercise 35 of 246A Notes 3)
extends to be holomorphic on all of
, and thus has a Taylor expansion
for some coefficients
obeying (4). The argument now proceeds as with part (i).
The additive cylinder and the multiplicative cylinder
can both be identified (on the level of smooth manifolds, at least) with the geometric cylinder
, but we will not use this identification here.
Now let us turn attention to doubly periodic functions of a complex variable , that is to say functions
that obey two periodicity relations
Within the world of holomorphic functions, the collection of doubly periodic functions is boring:
Proposition 2 Letbe an entire doubly periodic function (with periods
linearly independent over
). Then
is constant.
In the language of Riemann surfaces, this proposition asserts that the torus is a non-hyperbolic Riemann surface; it cannot be holomorphically mapped non-trivially into a bounded subset of the complex plane.
Proof: The fundamental domain (up to boundary) enclosed by is compact, hence
is bounded on this domain, hence bounded on all of
by double periodicity. The claim now follows from Liouville’s theorem. (One could alternatively have argued here using the compactness of the torus
.
To obtain more interesting examples of doubly periodic functions, one must therefore turn to the world of meromorphic functions – or equivalently, holomorphic functions into the Riemann sphere . As it turns out, a particularly fundamental example of such a function is the Weierstrass elliptic function
Previous set of notes: Notes 1. Next set of notes: Notes 3.
In Exercise 5 (and Lemma 1) of 246A Notes 4 we already observed some links between complex analysis on the disk (or annulus) and Fourier series on the unit circle:
- (i) Functions
that are holomorphic on a disk
are expressed by a convergent Fourier series (and also Taylor series)
for
(so in particular
), where
conversely, every infinite sequenceof coefficients obeying (1) arises from such a function
.
- (ii) Functions
that are holomorphic on an annulus
are expressed by a convergent Fourier series (and also Laurent series)
, where
conversely, every doubly infinite sequenceof coefficients obeying (2) arises from such a function
.
- (iii) In the situation of (ii), there is a unique decomposition
where
extends holomorphically to
, and
extends holomorphically to
and goes to zero at infinity, and are given by the formulae
whereis any anticlockwise contour in
enclosing
, and and
whereis any anticlockwise contour in
enclosing
but not
.
This connection lets us interpret various facts about Fourier series through the lens of complex analysis, at least for some special classes of Fourier series. For instance, the Fourier inversion formula becomes the Cauchy-type formula for the Laurent or Taylor coefficients of
, in the event that the coefficients are doubly infinite and obey (2) for some
, or singly infinite and obey (1) for some
.
It turns out that there are similar links between complex analysis on a half-plane (or strip) and Fourier integrals on the real line, which we will explore in these notes.
We first fix a normalisation for the Fourier transform. If is an absolutely integrable function on the real line, we define its Fourier transform
by the formula
Exercise 1 (Fourier transform of Gaussian) Ifis a complex number with
and
is the Gaussian function
, show that the Fourier transform
is given by the Gaussian
, where we use the standard branch for
.
The Fourier transform has many remarkable properties. On the one hand, as long as the function is sufficiently “reasonable”, the Fourier transform enjoys a number of very useful identities, such as the Fourier inversion formula
Exercise 2 (Decay ofimplies regularity of
) Let
be an absolutely integrable function.
Hint: to establish holomorphicity in each of these cases, use Morera’s theorem and the Fubini-Tonelli theorem. For uniqueness, use analytic continuation, or (for part (iv)) the Schwartz reflection principle.
- (i) If
has super-exponential decay in the sense that
for all
and
(that is to say one has
for some finite quantity
depending only on
), then
extends uniquely to an entire function
. Furthermore, this function continues to be defined by (3).
- (ii) If
is supported on a compact interval
then the entire function
from (i) obeys the bounds
for
. In particular, if
is supported in
then
.
- (iii) If
obeys the bound
for all
and some
, then
extends uniquely to a holomorphic function
on the horizontal strip
, and obeys the bound
in this strip. Furthermore, this function continues to be defined by (3).
- (iv) If
is supported on
(resp.
), then there is a unique continuous extension of
to the lower half-plane
(resp. the upper half-plane
) which is holomorphic in the interior of this half-plane, and such that
uniformly as
(resp.
). Furthermore, this function continues to be defined by (3).
Later in these notes we will give a partial converse to part (ii) of this exercise, known as the Paley-Wiener theorem; there are also partial converses to the other parts of this exercise.
From (3) we observe the following intertwining property between multiplication by an exponential and complex translation: if is a complex number and
is an absolutely integrable function such that the modulated function
is also absolutely integrable, then we have the identity
The material in these notes is loosely adapted from Chapter 4 of Stein-Shakarchi’s “Complex Analysis”.
Previous set of notes: 246A Notes 5. Next set of notes: Notes 2.
— 1. Jensen’s formula —
Suppose is a non-zero rational function
, then by the fundamental theorem of algebra one can write
Exercise 1 Letbe a complex polynomial of degree
.
- (i) (Gauss-Lucas theorem) Show that the complex roots of
are contained in the closed convex hull of the complex roots of
.
- (ii) (Laguerre separation theorem) If all the complex roots of
are contained in a disk
, and
, then all the complex roots of
are also contained in
. (Hint: apply a suitable Möbius transformation to move
to infinity, and then apply part (i) to a polynomial that emerges after applying this transformation.)
There are a number of useful ways to extend these formulae to more general meromorphic functions than rational functions. Firstly there is a very handy “local” variant of (1) known as Jensen’s formula:
Theorem 2 (Jensen’s formula) Letbe a meromorphic function on an open neighbourhood of a disk
, with all removable singularities removed. Then, if
is neither a zero nor a pole of
, we have
where
and
range over the zeroes and poles of
respectively (counting multiplicity) in the disk
.
One can view (3) as a truncated (or localised) variant of (1). Note also that the summands are always non-positive.
Proof: By perturbing slightly if necessary, we may assume that none of the zeroes or poles of
(which form a discrete set) lie on the boundary circle
. By translating and rescaling, we may then normalise
and
, thus our task is now to show that
An important special case of Jensen’s formula arises when is holomorphic in a neighborhood of
, in which case there are no contributions from poles and one simply has
Exercise 3 Use (6) to give another proof of Liouville’s theorem: a bounded holomorphic functionon the entire complex plane is necessarily constant.
Exercise 4 Use Jensen’s formula to prove the fundamental theorem of algebra: a complex polynomialof degree
has exactly
complex zeroes (counting multiplicity), and can thus be factored as
for some complex numbers
with
. (Note that the fundamental theorem was invoked previously in this section, but only for motivational purposes, so the proof here is non-circular.)
Exercise 5 (Shifted Jensen’s formula) Letbe a meromorphic function on an open neighbourhood of a disk
, with all removable singularities removed. Show that
for all
in the open disk
that are not zeroes or poles of
, where
and
. (The function
appearing in the integrand is sometimes known as the Poisson kernel, particularly if one normalises so that
and
.)
Exercise 6 (Bounded type)
- (i) If
is a holomorphic function on
that is not identically zero, show that
.
- (ii) If
is a meromorphic function on
that is the ratio of two bounded holomorphic functions that are not identically zero, show that
. (Functions
of this form are said to be of bounded type and lie in the Nevanlinna class for the unit disk
.)
Exercise 7 (Smoothed out Jensen formula) Letbe a meromorphic function on an open set
, and let
be a smooth compactly supported function. Show that
where
range over the zeroes and poles of
(respectively) in the support of
. Informally argue why this identity is consistent with Jensen’s formula. (Note: as many of the functions involved here are not holomorphic, complex analysis tools are of limited use. Try using real variable tools such as Stokes theorem, Greens theorem, or integration by parts.)
When applied to entire functions , Jensen’s formula relates the order of growth of
near infinity with the density of zeroes of
. Here is a typical result:
Proposition 8 Letbe an entire function, not identically zero, that obeys a growth bound
for some
and all
. Then there exists a constant
such that
has at most
zeroes (counting multiplicity) for any
.
Entire functions that obey a growth bound of the form for every
and
(where
depends on
) are said to be of order at most
. The above theorem shows that for such functions that are not identically zero, the number of zeroes in a disk of radius
does not grow much faster than
. This is often a useful preliminary upper bound on the zeroes of entire functions, as the order of an entire function tends to be relatively easy to compute in practice.
Proof: First suppose that is non-zero. From (6) applied with
and
one has
Just as (3) and (7) give truncated variants of (1), we can create truncated versions of (2). The following crude truncation is adequate for many applications:
Theorem 9 (Truncated formula for log-derivative) Letbe a holomorphic function on an open neighbourhood of a disk
that is not identically zero on this disk. Suppose that one has a bound of the form
for some
and all
on the circle
. Let
be constants. Then one has the approximate formula
for all
in the disk
other than zeroes of
. Furthermore, the number of zeroes
in the above sum is
.
Proof: To abbreviate notation, we allow all implied constants in this proof to depend on .
We mimic the proof of Jensen’s formula. Firstly, we may translate and rescale so that and
, so we have
when
, and our main task is to show that
Suppose has a zero
with
. If we factor
, where
is the Blaschke product (5), then
Similarly, given a zero with
, we have
, so using Blaschke products to remove all of these zeroes also only affects the left-hand side of (8) by
(since the number of zeroes here is
), with
also modified by at most
. Thus we may assume in fact that
has no zeroes whatsoever within the unit disk. We may then also normalise
, then
for all
. By Jensen’s formula again, we have
Exercise 10
- (i) (Borel-Carathéodory theorem) If
is analytic on an open neighborhood of a disk
and
, show that
(Hint: one can normalise
,
,
, and
. Now
maps the unit disk to the half-plane
. Use a Möbius transformation to map the half-plane to the unit disk and then use the Schwarz lemma.)
- (ii) Use (i) to give an alternate way to conclude the proof of Theorem 9.
A variant of the above argument allows one to make precise the heuristic that holomorphic functions locally look like polynomials:
Exercise 11 (Local Weierstrass factorisation) Let the notation and hypotheses be as in Theorem 9. Then show thatfor all
in the disk
, where
is a polynomial whose zeroes are precisely the zeroes of
in
(counting multiplicity) and
is a holomorphic function on
of magnitude
and first derivative
on this disk. Furthermore, show that the degree of
is
.
Exercise 12 (Preliminary Beurling factorisation) Letdenote the space of bounded analytic functions
on the unit disk; this is a normed vector space with norm
- (i) If
is not identically zero, and
denote the zeroes of
in
counting multiplicity, show that
and
- (ii) Let the notation be as in (i). If we define the Blaschke product
where
is the order of vanishing of
at zero, show that this product converges absolutely to a holomorphic function on
, and that
for all
. (It may be easier to work with finite Blaschke products first to obtain this bound.)
- (iii) Continuing the notation from (i), establish a factorisation
for some holomorphic function
with
for all
.
- (iv) (Theorem of F. and M. Riesz, special case) If
extends continuously to the boundary
, show that the set
has zero measure.
Remark 13 The factorisation (iii) can be refined further, withbeing the Poisson integral of some finite measure on the unit circle. Using the Lebesgue decomposition of this finite measure into absolutely continuous parts one ends up factorising
functions into “outer functions” and “inner functions”, giving the Beurling factorisation of
. There are also extensions to larger spaces
than
(which are to
as
is to
), known as Hardy spaces. We will not discuss this topic further here, but see for instance this text of Garnett for a treatment.
Exercise 14 (Littlewood’s lemma) Letbe holomorphic on an open neighbourhood of a rectangle
for some
and
, with
non-vanishing on the boundary of the rectangle. Show that
where
ranges over the zeroes of
inside
(counting multiplicity) and one uses a branch of
which is continuous on the upper, lower, and right edges of
. (This lemma is a popular tool to explore the zeroes of Dirichlet series such as the Riemann zeta function.)
Just a short announcement that next quarter I will be continuing the recently concluded 246A complex analysis class as 246B. Topics I plan to cover:
- Schwartz-Christoffel transformations and the uniformisation theorem (using the remainder of the 246A notes);
- Jensen’s formula and factorisation theorems (particularly Weierstrass and Hadamard); the Gamma function;
- Connections with the Fourier transform on the real line;
- Elliptic functions and their relatives;
- (if time permits) the Riemann zeta function and the prime number theorem.
Notes for the later material will appear on this blog in due course.
Recent Comments