UNSW MATH2621 Course Notes
UNSW MATH2621 Course Notes
Assumed Knowledge
This is a review of basic facts about complex numbers that ought to be familiar:
Students who do not feel confident about this material need to do lots of exercises
about these, such as those in the MATH1141 notes.
1. Complex numbers
Definition 0.1. A complex number is an expression of the form x + iy, where
x and y are real numbers. The real part of x + iy is x and the imaginary part of
x + iy is y. We denote this by Re(x + iy) = x and Im(x + iy) = y. The set of all
complex numbers is denoted C.
We often write w = u + iv and z = x + iy, and work with w and z rather than
u + iv and x + iy. In this case, we write, for instance, Re(z) = x and Im(w) = v.
We abbreviate x + i0 and 0 + iy to x and iy, and 0 + i1 to i.
w + z = (u + x) + i(v + y)
−z = −x + i(−y)
wz = (ux − vy) + i(uy + vx)
z −1 = (x2 + y 2 )−1 (x − iy)
w/z = wz −1 = (x2 + y 2 )−1 [(ux + vy) + i(vx − uy)].
Inequality (g) is called the triangle inequality. If |z| = 1, then z −1 = z, from (j).
Proof. We only prove the triangle inequality, because this is hardest.
First, here is an algebraic proof. Recall that 2ab ≤ a2 + b2 for real numbers a
and b. Taking a and b to be x1 y2 and x2 y1 , we deduce that
2x1 x2 y1 y2 ≤ x21 y22 + x22 y12 ,
and hence
x21 x22 + y12 y22 + 2x1 x2 y1 y2 ≤ x21 x22 + y12 y22 + x21 y22 + x22 y12 = (x21 + y12 )(x22 + y22 ),
so, taking square roots,
x1 x2 + y1 y2 ≤ |z1 | |z2 | .
Finally, z1 + z2 = (x1 + x2 ) + i(y1 + y2 ), and so
|z1 + z2 |2 = (x1 + x2 )2 + (y1 + y2 )2
= x21 + x22 + y12 + y22 + 2(x1 x2 + y1 y2 )
≤ |z1 |2 + |z2 |2 + 2|z1 | |z2 |
2
= |z1 | + |z2 | ;
the triangle inequality follows by taking square roots.
2. EULER’S FORMULA 3
2. Euler’s formula
The usual exponential function has a Taylor series:
x x2 x3 x4 x5
ex = 1 + + + + + + ··· ,
1! 2! 3! 4! 5!
so at least formally, for a real number θ,
θ θ2 θ3 θ4 θ5
eiθ = 1 + i − −i + + i + ···
1! 2! 3! 4! 5!
θ 2
θ 4 θ θ3 θ5
= 1− + + ··· + i − + + ···
2! 4! 1! 3! 5!
= cos θ + i sin θ,
from the Taylor series for the cosine and sine functions.
Later we will make this rigorous and use power series very effectively. This
observation leads us to make the following definition.
Definition 0.6. We define eiθ to be cos θ + i sin θ, for any real number θ.
Suppose that z = x + iy 6= 0. Write r instead of |z|. Then (x/r, y/r) lies on the
unit circle in the Cartesian plane, so (x/r, y/r) = (cos θ, sin θ) for an appropriate
choice of θ, and hence
x y
z=r +i = r cos θ + i sin θ = reiθ .
r r
Definition 0.7. The Cartesian form of a complex number z is its representation
in the form x + iy, where x and y are real. The polar form of a complex number z is
its representation in the form reiθ , where r ≥ 0 and θ ∈ R. The number θ is called
the argument of z, and is written arg(z).
Lemma 0.8. Suppose that r, s ∈ R+ and θ, ϕ ∈ R. Then
r(cos θ + i sin θ) = s(cos ϕ + i sin ϕ)
if and only if r = s and θ − ϕ = 2kπ for some k ∈ Z.
This lemma follows immediately from trigonometry. It tells us that the argument
of a nonzero complex number z is ambiguous. The next definition is to avoid this
ambiguity.
Definition 0.9. The principal value of the argument of a nonzero complex
number z, written Arg(z), is the unique number θ such that z = |z| eiθ and −π <
θ ≤ π.
4. De Moivre’s formula
Theorem 0.10. If θ, ϕ ∈ R, then
(cos θ + i sin θ)(cos ϕ + i sin ϕ) = cos(θ + ϕ) + i sin(θ + ϕ).f
Proof. For all θ, ϕ ∈ R,
(cos θ + i sin θ)(cos ϕ + i sin ϕ)
= (cos θ cos ϕ − sin θ sin ϕ) + i(cos θ sin ϕ + sin θ cos ϕ)
= cos(θ + ϕ) + i sin(θ + ϕ),
as required. □
Corollary 0.11 (de Moivre’s formula). If n ∈ Z and θ ∈ R, then
(cos θ + i sin θ)n = cos(nθ) + i sin(nθ). (0.1)
Proof. This is obviously true if n = 0 or 1. The result may be proved for
n ∈ Z+ by induction. Suppose that k ∈ Z+ and formula (0.1) holds when n = k,
i.e.,
(cos θ + i sin θ)k = cos kθ + i sin kθ.
Then by Theorem 0.10,
(cos θ + i sin θ)k+1 = (cos θ + i sin θ)(cos θ + i sin θ)k
= (cos θ + i sin θ)(cos kθ + i sin kθ)
= cos(θ + kθ) + i sin(θ + kθ)
= cos(k + 1)θ + i sin(k + 1)θ,
Im
w v w = u + iv = seiϕ
b b
s
ϕ Re
b
u O
so the result holds when n = k + 1. By induction, the result holds for all n ∈ Z+ .
To prove the result when n ∈ Z− , we use the fact that if |z| = 1, then zz = 1, so
z = z −1 . That is,
(cos θ + i sin θ)−1 = cos θ − i sin θ = cos(−θ) + i sin(−θ),
as required. □
From Lemma 0.8, r = s1/n and nθ = ϕ + 2kπ for some k ∈ Z. Therefore z = s1/n eiθ ,
where
ϕ 2kπ
θ= +
n n
for some k ∈ Z. If k = nl + r, where l ∈ Z and r = 0, 1, 2, . . . , n − 1, then
e(ϕ/n+2kπ/n) = e(ϕ/n+2rπ/n+2lπ) = e(ϕ/n+2rπ/n) ,
so we get the same value of z by taking k = nl + r as by taking k = r. Thus, to
get all n possible values of z, it suffices to take the first n values of k, or any n
consecutive values, or indeed any n values of k which give the n possible different
remainders when divided by n.
The nth roots of any nonzero complex number w are uniformly spaced around
the circle with centre 0 and radius |w|1/n . For example, Figure 0.2 shows the seventh
roots of unity. A symmetry argument shows that the sum of all these numbers is 0.
Im
e4πi/7 b
b
e2πi/7
e6πi/7 b
b
1 Re
b
e−6πi/7 b
b
b
e−2πi/7
e−4πi/7
6. History†
Leonhard Euler found the formula that bears his name in 1748; he was one of
the most prolific mathematicians ever. Some mathematicians did not believe that
the geometric representation afforded by the Argand diagram was legitimate. The
story of de Moivre’s death is very curious. For more information, see
http://www-groups.dcs.st-and.ac.uk/~history/Biographies/
and then find Euler, Argand, and de Moivre.
LECTURE 1
2. Properties of sets
Definition 1.5. The open ball with centre z0 and radius ε, written B(z0 , ε), is
the set {z ∈ C : |z − z0 | < ε}.
The punctured open ball with centre z0 and radius ε, written B◦ (z0 , ε), is the set
{z ∈ C : 0 < |z − z0 | < ε}.
Sometimes these sets are called discs rather than balls.
Definition 1.6. Suppose that S ⊆ C. For any point z0 in C, there are three
mutually exclusive and exhaustive possibilities:
(1) When the positive real number ε is sufficiently small, B(z0 , ε) is a subset of
S, that is, B(z0 , ε) ∩ S = B(z0 , ε). In this case, z0 is an interior point of S.
(2) When the positive real number ε is sufficiently small, B(z0 , ε) does not meet
S, that is, B(z0 , ε) ∩ S = ∅. In this case, z0 is an exterior point of S.
(3) No matter how small the positive real number ε is, neither of the above
holds, that is, ∅ ⊂ B(z0 , ε) ∩ S ⊂ B(z0 , ε). In this case, z0 is a boundary
point of S.
These definitions are illustrated in Figure 1.1. We consider points z1 , z2 and z3 .
If the radius of the ball centred at z1 is small enough, then the ball lies inside
the set S, and B(z1 , ε) ∩ S = B(z1 , ε). Thus z1 is an interior point.
If the radius of the ball centred at z2 is small enough, then the ball lies outside
the set S, and B(z2 , ε) ∩ S is empty. Thus z2 is an exterior point.
Im
z1 b
B(z0 , ε)
z0 ε
b z2 b
z3 Re
b
Figure 1.1. The ball with centre z0 and radius ε, and interior, ex-
terior and boundary points z1 , z2 and z3 of the set S
3. DESCRIBING SETS IN THE COMPLEX PLANE 9
No matter how small the radius of the ball centred at z3 is, part of the ball lies
inside S, and part lies outside S, and B(z3 , ε) ∩ S is neither empty nor all of B(z3 , ε).
Thus z3 is a boundary point.
Definition 1.7. Suppose that S ⊆ C.
(1) The set S is open if all its points are interior points.
(2) The set S is closed if it contains all of its boundary points, or equivalently,
if its complement C \ S is open.
(3) The closure of S, written S̄, is the set consisting of all the points of S
together with all its boundary points.
(4) The set S is bounded if S ⊆ B(0, R) for some positive real number R.
(5) The set S is compact if it is both closed and bounded.
(6) The set S is a region if it is an open set together with none, some, or all of
its boundary points.
For example, the dashed boundary lines of the set S in Figure 1.1 indicate that
it does not contain any boundary points. Consequently, this set is open.
Note that open and closed are not exclusive nor exhaustive. There are sets that
are open and closed, such as the whole plane, and sets that are neither open nor
closed, such as {z ∈ C : Re(z) ≥ 0, Im(z) > 0}. In complex analysis, we focus on
open sets. We often write Ω for an open set.
We now present connectedness.
Definition 1.8. A polygonal path is a finite sequence of finite line segments,
where the end point of one line segment is the initial point of the next one. A simple
closed polygonal path is a polygonal path that does not cross itself, but the final point
of the last segment is the initial point of the first segment. The complement of a
simple closed polygonal path is made up of two pieces: one, the interior of the path,
is bounded, and the other, the exterior, is not.
Definition 1.9. Let X ⊆ C be a subset of the complex plane.
(1) The set X is polygonally path-connected if any two points of X can be joined
by a polygonal arc lying inside X.
(2) The set X is simply polygonally connected if it is polygonally path-connected
and if the interior of every simple closed polygonal arc in X lies in X, that
is, if “X has no holes”.
(3) The set X is a domain if it is open and polygonally path-connected.
The set Ω in Figure 1.2 is polygonally path-connected, because any two points in
Ω (such as z1 and z4 ) can be joined by a polygonal path. However, Ω is not simply
polygonally connected, because part of the interior of the closed path shown going
through z5 is not in Ω.
Answer. If |az + b| = |cz + d|, then |az + b|2 = |cz + d|2 , so (az + b)(az + b) =
(cz + d)(cz + d) , whence
¯ + (āb − c̄d)z̄ + (cc̄ − dd)
(aā − cc̄)z z̄ + (ab̄ − cd)z ¯ = 0.
We may rewrite this in the form e|z|2 + f z + f¯z̄ + g = 0, where e and g are real,
while f may be complex. Now we pass to Cartesian coordinates:
ex2 + ey 2 + (f + f¯)x + i(f − f¯)y + g = 0;
all the coefficients are now real since f + f¯ = 2 Re f while i(f − f¯) = −2 Im f .
The desired result follows from coordinate geometry. For instance, if e 6= 0 and
eg − |f |2 < 0, then we have a circle with centre −f¯/e, while if e = 0 and f 6= 0,
then we have a straight line.
The other types of solutions arise when many of the coefficients are 0. If a = c = 0
and |b| = |d|, then the equation holds in the whole plane; if a = c = 0 while |b| 6= |d|,
then there are no solutions at all. If a = b = 0 and c 6= 0, then |z − d/c| = 0, and
the radius of the circle is 0, which gives a point solution. The situation is similar if
c = d = 0. 4
Exercise 1.11. Sketch the set {z ∈ C : |z − 3 − 2i| < 4, Re(z) > 0} in the
complex plane. Is it open, closed, bounded, compact, polygonally path-connected,
simply polygonally connected, a region, or a domain?
Im
z1 b
Re
z5 b
z4
Ω
Im Im
4 4
b b 3
3 + 2i Re 3 + 2i Re
Answer. The set includes points inside or on the circle with centre 3 + 2i and
radius 5, but outside or on the circle with centre 3 + 2i and radius 4. It includes the
points on the two concentric circles. The sketch is on the right in Figure 1.3.
The set {z ∈ C : 4 < |z − 3 − 2i| < 5} is open, because all points are interior
points. The set we are considering is this set together will all its boundary points,
so it is closed, and also a region. It is not open, because some points are boundary
points, and hence not a domain. It is closed and bounded, hence also compact. It
is connected, but not simply connected, because there are points in the complement
of the set inside a polygonal path around the annulus. 4
|z + i| + |z − i| = 4
|z + i| = 4 − |z − i|
|z + i| = 16 − 8|z − i| + |z − i|2
2
x2 y 2
+ = 1.
3 4
This is the circumference of an ellipse, as shown in Figure 1.4. The ellipse is the set
of points for which the sum of the distances from i and −i is exactly 4. It is closed
but not open, and bounded.
12 1. INEQUALITIES AND SETS OF COMPLEX NUMBERS
Im
b
2i
b
√
3 b
Re
The sets
{z ∈ C : |z + i| + |z − i| > 4} and {z ∈ C : |z + i| + |z − i| < 4}
are the exterior and interior of the ellipse. 4
(1) Note that connected is not defined for closed sets, but there are questions
about closed sets being connected.
(2) Note that simply connected does not imply connected, or vice versa.
(3) Point out that it is best to use Cartesian coordinates late in a calculation.
(4) It is not important to do the last example, but if it is done, then it is nice
to make a comment that the sign of |z + i| + |z − i| − 4 has to be constant
on connected components of the complement of the set where it is 0.
LECTURE 2
1. Functions
In Mathematics, we often think of functions as machines: you give the machine
a number, x say, press a button, and out comes f (x). We sometimes write x 7→ f (x)
to indicate that x is the input and f (x) is the output.
• The domain of a function f , written Domain(f ), is the set of all the numbers
you are allowed to put in. Sometimes this is restricted in some way. If there
is no explicit restriction, you should consider the natural domain, that is,
the largest domain possible.
• A codomain is a set of numbers that includes all the numbers that you can
get out, and perhaps more.
• The range or image of a function f , written Range(f ), is the set of the
numbers that you can get out, and no others.
• The image of a subset S of the domain of a function f , sometimes written
f (S), is the set of all possible values f (s) as s varies over S.
• The preimage of a subset T of the codomain of a function f , sometimes
written f −1 (T ), is the set of all x in Domain(f ) such that f (x) ∈ T .
Definition 2.1. A complex function is one whose domain, or whose range, or
both, is a subset of the complex plane C that is not a subset of the real line R.
To emphasize that the domain is complex, not real, the expression function of a
complex variable may be used. To emphasize that the range is complex, not real,
the expression complex-valued function may be used.
Remark 2.2. The word ”domain” is then used in two different ways in this
course. We say that a subset of C is a domain if it is open and connected which has
nothing to do with a function. We just saw that the domain of a function f is the
set of points where f is defined. Be careful to not confuse them and note that the
domain of a function is not necessarily a domain in the sense of open and connected:
for instance consider f : {0, 1} → C, 0 7→ 0, 1 7→ 1 and not that its domain is equal
to two points which forms a nonconnected and nonopen subset of C.
2. Examples of functions
Examples of functions of a complex variable include the real part function Re,
the imaginary part function Im, the modulus function z 7→ |z|, and the principal
value of the argument Arg; these are all real-valued. Complex conjugation z 7→ z̄ is
an example of a complex-valued function of a complex variable.
13
14 2. FUNCTIONS OF A COMPLEX VARIABLE
In this course, we are going to learn about a number of useful complex functions.
Shortly we will define complex polynomials and then rational functions. In future
lectures, we will define log z, sin z, and cosh z for a complex number z, and there
are many other functions in the menagerie of complex functions.
Exercise 2.3. Suppose that f (z) = 1/z for all z ∈ C \ {0}, and that g(z) = z
for all z ∈ C. Show that f ◦ f (z) = z for all z ∈ C \ {0}. Is f ◦ f = g?
Answer. By definition,
f ◦ f (z) = f (1/z) = 1/(1/z) = z = g(z).
However, the domain of f ◦ f is C \ {0} and the domain of g is C, so these functions
are different. 4
where the αj are all distinct, and the multiplicities mj add to give the degree of p.
Theorem 2.6 (Polynomial division and partial fractions). Suppose that p and q
are polynomials of degrees m and n. Then the rational function p/q may be written
as a sum
p(z) r(z)
= s(z) + ,
q(z) q(z)
where r and s are polynomials, and the degree of r is strictly less than the degree of
q. Further, if
Y e
q(z) = c (z − βj )mj ,
j=1
At this stage, we do not prove these results, which should be familiar, though
perhaps not in this generality; we will give proofs later.
The natural domain of any complex polynomial is C. Sometimes we cannot
determine the range of a real polynomial exactly, because we cannot find maxima
or minima exactly. However, for complex polynomials, things are easier.
Exercise 2.7. Suppose that p is a nonconstant complex polynomial. Show that
the range of p is C.
Answer. Take a nonconstant complex polynomial p, and a complex number w.
We need to show that there is z ∈ C such that p(z) = w. Define q(z) = p(z) − w.
Then q is also a nonconstant complex polynomial, so has a root by the fundamental
theorem of algebra. That is, there exists z ∈ C such that q(z) = 0. It follows that
p(z) = w, as required. 4
Exercise 2.9. Suppose that f (z) = z 3 + z − 2. Write the real and imaginary
parts of this function as functions u and v of (x, y), where z = x + iy.
Answer. Observe that
f (x + iy) = (x + iy)3 + x + iy − 2
= x3 + 3ix2 y − 3xy 2 − iy 3 + x − iy − 2
= (x3 − 3xy 2 + x − 2) + i(3x2 y − y 3 − y).
Thus u(x, y) = x3 − 3xy 2 + x − 2 and v(x, y) = 3x2 y − y 3 − y. 4
Exercise 2.10. Suppose that f (z) = 1/z. Write the real and imaginary parts
of this function as functions of x and y, where z = x + iy.
Answer. Note that
1 x − iy x −y
f (x + iy) = = 2 = 2 +i .
x + iy x + y2 x + y2 x2 + y 2
x −y
Thus Re f (x + iy) = 2 and Im f (x + iy) = . 4
x + y2 x2 + y 2
16 2. FUNCTIONS OF A COMPLEX VARIABLE
Exercise 2.11. Write ez in the form u(x, y) + iv(x, y), where z = x + iy.
Answer. Observe that
ez = ex (cos(y) + i sin(y)) = ex cos(y) + iex sin(y) = u(x, y) + iv(x, y),
where u(x, y) = ex cos(y) and v(x, y) = ex sin(y). 4
Sometimes we view the complex number z in polar coordinates, that is, we write
z = reiθ . In this case, we consider
f (z) = u(r, θ) + iv(r, θ).
Exercise 2.12. Write ez in the form u(r, θ) + iv(r, θ), where z = reiθ .
Answer. Observe that
ez = er cos θ+ir sin θ
= er cos θ (cos(r sin θ) + i sin(r sin θ))
= er cos θ cos(r sin θ) + i er cos θ sin(r sin θ).
Thus u(r, θ) = er cos θ cos(r sin θ) and v(r, θ) = er cos θ sin(r sin θ). 4
the points where f (z) varies will include ∞, and this means that f (z) must vary
on a line. Conversely, if the points where z varies do not include −d/c, then f (z)
will stay bounded, and this means that f (z) must vary on a circle. Once we know
whether f (z) varies on a line or on a circle, we may find the equation of the line or
the circle quite easily by finding a few values of f (z).
Example 2.14. Let f (z) = 1/z. As z varies on the line x = 1, its image f (z)
varies on a circle, because z stays away from 0 and so 1/z stays away from ∞.
This circle passes through the points 1 and 0 (since f (z) → 0 as z → ∞), and is
symmetric about the real axis, since 1/(1 − it) = (1/(1 + it)) . This must be the
circle that we found above.
Exercise 2.16. Suppose that p(z) = 10z 4 − 3z 3 + z − 10. Show that when |z|
is large enough, |p(z)| ≤ 11|z|4 .
Answer. Write p(z) = z 4 (10 − 3z −1 + z −3 − 10z −4 ). If |z| > 10, then, by the
triangle inequality applied several times,
3 1 10
|10 − 3z −1 + z −3 − 10z −4 | ≤ 10 + + + ≤ 11,
10 1000 10000
and so |p(z)| ≤ 11|z|4 when |z| > 10. 4
LECTURE 3
We often use graphical methods to gain useful intuition about complex functions,
and we spend some time investigating these. It is hard to represent complex func-
tions, because there are up to four variables involved. Just as we often write y = f (x)
for a real function, it is common to consider a function in the form w = f (z), and
to use real variables x and y to describe the domain and u and v to describe the
range. Typically, we draw “elementary” curves in the z plane, such as lines parallel
to the axes, or concentric circles around and rays exiting from the origin, and then
examine their images in the w plane, or we draw similar elementary curves in the w
plane and then examine their preimages.
Im Im
Re Re
Im Im
Re Re
The inverse of an affine mapping is an affine mapping. It follows that the pre-
images of lines are lines and preimages of circles are circles; the preimage of a grid
parallel to the axes is a rectangular grid, but not necessarily parallel to the axes.
In Figures 3.1 and 3.2, we illustrate the images of lines parallel to the axes and
circles around the origin under affine mappings.
This is a good way to show how the function behaves, although a lot of space is
used and care is needed to choose the points that are moved in a way that is not
ambiguous. Indeed, it might be better to draw a very asymmetrical figure in the z
plane and then its image in the w plane.
Sometimes we just draw the right hand figure of the two drawn above, labelling
the curves with the corresponding curve in the domain of the function. In Figure 3.1,
these are the circles r = 1, r = 2, and r = 3; in Figure 3.2, they are the horizontal
lines x = 0, x = ±1, x = ±2, and the vertical lines y = 0, y = ±1, y = ±2.
On the other hand, we may look for curves in the xy plane whose images in the
uv plane are the lines u = c and v = d. So we are finding the level curves of the
3. THE FUNCTION w = 1/z 21
real and imaginary parts of the function. People who are used to map reading can
build a picture in their mind of terrain, just knowing the contours that represent
different heights. They imagine a surface above the page at the height indicated by
the contour, and then fill in the gaps.
2. Quadratic functions
Now we consider the function z 7→ z 2 . Notice that this function is two-to-one in
B◦ (0, ∞).
On the one hand, we may represent the images in the uv plane of the curves in
the xy plane given by x = a and y = b, or by r = a and θ = b. For instance, if x = a
and y = t, where a is fixed and t varies, then
w = z 2 = (a + it)2 = a2 − t2 + 2iat.
That is, u = a2 − t2 and v = 2at. We eliminate t to show that
v2
u = a2 − .
4a2
Alternatively, if y = b and x = t, where b is fixed and t varies, then
w = z 2 = (t + ib)2 = t2 − b2 + 2ibt.
That is, u = t2 − b2 and v = 2bt. We eliminate t to show that
v2
u= − b2 .
4b2
See, for example, Figure 3.3.
Exercise† 3.1. Find the focus and the directrix of the parabola u = v 2 /4b2 − b2
in the uv plane.
On the other hand, we may look for the values of x and y so that Re(z 2 ) or
Im(z 2 ) takes a fixed value. For instance, if Re(z 2 ) = a, then
x2 − y 2 = a,
and this is a hyperbola opening to the left and right, or up and down, depending on
the sign of a. Similarly, if Im(z 2 ) = b, then
2xy = b,
and this is a right hyperbola in the first and third quadrants, or in the second and
fourth quadrants, depending on the sign of b. See, for example, Figure 3.3.
Exercise 3.2. How would you “sketch the graph” of w = (z − 1)2 − 1?
Im Im
Re Re
Im Im
Re Re
Im Im
Re Re
Similarly, as θ varies, the point reiθ moves around a circle centred at the origin.
Now w = (1/r)e−iθ , and this point moves around the same circle, but in the opposite
direction.
5. MORE ON GRAPHICAL REPRESENTATIONS OF COMPLEX FUNCTIONS 23
Im Im
Re Re
Im Im
Re Re
25
26 4. FRACTIONAL LINEAR TRANSFORMATIONS
Our discussion of the domain and range had several cases; we can simplify the
statements by enlarging the set of complex numbers by adding ∞. Indeed, we define
az + b az + b a
f (−d/c) = lim = ∞ and f (∞) = lim =
z→−d/c cz + d z→∞ cz + d c
(we will define limits formally in the next lecture) and now we can just write
TM : C ∪ {∞} → C ∪ {∞}.
We often call C ∪ {∞} the Riemann sphere, and write it S: we imagine a unit
sphere in three dimensions with centre at 0. We can define a function σ : C → S
geometrically by joining a point p in the plane to the north pole n of the sphere by a
straight line. The line will cut the sphere at n and at one other point, which we call
σ(p). Then we may think of n as being σ(∞). The function σ is called stereographic
projection.
In this lecture, we outline the key ideas and facts about limits and continuity,
as a preliminary to defining differentiability.
1. Limits
We define limits for complex functions much as for real functions.
Recall that, given a set S, we define its closure S̄ or S to be the set consisting
of all points of S together with all boundary points.
Definition 5.1. Suppose that f is a complex function, ℓ ∈ C, and z0 is in
Domain(f ) . We say that f (z) tends to ℓ as z tends to z0 , or that ℓ is the limit of
f (z) as z tends to z0 , and we write f (z) → ℓ as z → z0 , or
lim f (z) = ℓ,
z→z0
if, for every ε ∈ R+ , there exists δ ∈ R+ such that |f (z) − ℓ| < ε provided that z is
in Domain(f ) and 0 < |z − z0 | < δ.
Suppose also S is a subset of Domain(f ) and that z0 ∈ S̄. We say that f (z)
tends to ℓ as z tends to z0 in S, or that ℓ is the limit of f (z) as z tends to z0 in S,
and we write f (z) → ℓ as z → z0 in S, or
lim f (z) = ℓ,
z→z0
z∈S
if, for every ε ∈ R+ , there exists δ ∈ R+ such that |f (z) − ℓ| < ε provided that z ∈ S
and 0 < |z − z0 | < δ.
Informally, f (z) tends to ℓ if we can make f (z) arbitrarily close to ℓ by taking z
close to, but not equal to, z0 .
Most of what follows about limits of the form limz→z0 f (z) also applies to re-
stricted limits, that is, limits of the form limz→z0 f (z).
z∈S
We may rewrite the conditions 0 < |z − z0 | < δ and |f (z)−ℓ| < ε as z ∈ B◦ (z0 , δ)
and f (z) ∈ B(ℓ, ε). We define limits involving infinity in a similar way by defining
balls centred at infinity, and extending our previous definition slightly.
Definition 5.2. Suppose that ε > 0. We define both B(∞, ε) and B◦ (∞, ε) to
be the set {z ∈ C : |z| > 1/ε}.
Definition 5.3. Suppose that f is a complex function, that ℓ ∈ C ∪ {∞}, and that
either z0 ∈ Domain(f ) or Domain(f ) is unbounded and z0 = ∞. We say that f (z)
tends to ℓ as z tends to z0 , or that ℓ is the limit of f (z) as z tends to z0 , and we
write f (z) → ℓ as z → z0 , or
lim f (z) = ℓ,
z→z0
31
32 5. LIMITS AND CONTINUITY
if for all ε ∈ R+ , there exists δ ∈ R+ such that f (z) ∈ B(ℓ, ε) provided that
z ∈ Domain(f ) ∩ B◦ (z0 , δ).
With this definition, the following lemma holds; we omit the proof.
Lemma 5.4 (Standard limits). Suppose that α, c ∈ C. Then
lim c = c lim c = c
z→α z→∞
lim z − c = α − c lim z − α = ∞
z→α z→∞
1 1
lim =∞ lim = 0.
z→α z − α z→∞ z − α
Im
δ π − sin−1 (δ/a) Re
z b
−a
in the sense that if the right hand side exists, then so does the left hand side, and
they are equal. In particular, f (z) tends to ℓ as z tends to z0 if and only if Re(f (z))
tends to Re(ℓ) and Im(f (z)) tends to Im(ℓ) as z tends to z0 .
Proof. The proof of part (1) uses that fact that
f (z) − ℓ̄ = |f (z) − ℓ| .
The rest follows from the first part and the first two parts of Theorem 5.7. □
2. Examples of limits
Exercise 5.9. Show from first principles that limz→z0 z = z0 .
Answer. Given ε ∈ R+ , take δ = ε. Then if 0 < |z − z0 | < δ, it follows
immediately that |z − z0 | < ε, and the result is proved. 4
Thus Arg is not continuous at any point of (−∞, 0]. The function Arg is one of
many important discontinuous complex functions.
Exercise 5.12. Suppose that f (z) = z̄/z and g(z) = z 2 /z̄. Does limz→0 f (z) or
limz→0 g(z) exist? If so, find the limit; otherwise, explain why it does not exist.
4. Continuity
Definition 5.13. Suppose that f is a complex function. We say that f is continuous
at a point z0 if f (z0 ) is defined and limz→z0 f (z) = f (z0 ). We say that f is continuous
in a set S if it is continuous at all points of S. We say that f is continuous if it is
continuous at all points of its domain.
Exercise 5.20. Show that Log(z) is continuous in C \ (−∞, 0], and is not
continuous at any point on (−∞, 0].
Answer. We know that z 7→ |z| is continuous, and from calculus, ln is also
continuous. Hence z 7→ ln |z| is continuous.
We claim that Log is continuous where Arg is continuous. Indeed, if Arg is
continuous, then so is Log, as it is the sum of continuous functions; conversely, if
Log is continuous, then so is z 7→ Log(z) − ln |z|, for the same reason, that is, Arg
is continuous. 4
Complex Differentiability
1. Definition
Definition 6.1. Suppose that S ⊆ C and that f : S → C is a complex function.
Then we say that f is differentiable at the point z0 in S if
f (z) − f (z0 )
lim (6.1)
z→z0 z − z0
exists and is finite. If it does, it is called the derivative of f at z0 , and is written
df (z0 )
or f ′ (z0 ).
dz
We say that f is differentiable in a set S if it is differentiable at all points of S,
and that f is differentiable if it is differentiable at all points of its domain.
2. Examples
Exercise 6.3. Suppose that f1 (z) = z 2 + iz + 2. Is f1 differentiable at z0 in C?
If so, find f1′ (z0 )?
Answer. We use formula (6.1): if z 6= z0 , then
f1 (z) − f1 (z0 ) z 2 + iz + 2 − z02 − iz0 − 2
=
z − z0 z − z0
(z − z0 )(z + z0 ) + i(z − z0 )
=
z − z0
= z + z0 + i.
From the algebra of limits and standard limits,
f1 (z) − f1 (z0 )
lim = lim z + z0 + i = 2z0 + i.
z→z0 z − z0 z→z0
′
Hence f1 is differentiable at z0 and f1 (z0 ) = 2z0 + i. This holds for any point z0 , so
f1 is differentiable in C, and f1′ (z) = 2z + i. 4
37
38 6. COMPLEX DIFFERENTIABILITY
The computation above is almost identical to that to find the derivative of the
real function x2 + x + 2. Indeed, many formulae from the real case also hold in the
complex case when x is replaced by z. So do many theorems.
which depends on θ. By the uniqueness of limits, limw→0 w/w does not exist, and
f2 (z0 + w) − f2 (z0 )
lim
w→0 w
fails to exist. As z0 is arbitrary, f2 is not differentiable anywhere in C. 4
Remark 6.7. The pair of equations (6.3), which relate the partial derivatives of
u and v, are known as the Cauchy–Riemann equations.
Proof. If f ′ (z0 ) exists, then
f (z0 + w) − f (z0 )
f ′ (z0 ) = lim
w→0
w∈R
w
u(x0 + w, y0 ) + iv(x0 + w, y0 ) − u(x0 , y0 ) − iv(x0 , y0 )
= lim
w→0
w∈R
w
u(x0 + w, y0 ) − u(x0 , y0 ) v(x0 + w, y0 ) − v(x0 , y0 )
= lim + i lim
w→0
w∈R
w w→0
w∈R
w
∂u ∂v
= (x0 , y0 ) + i (x0 , y0 ),
∂x ∂x
because a limit exists if and only if its real and imaginary parts do. Thus
∂u ∂v
(x0 , y0 ) = Re(f ′ (z0 )) and (x0 , y0 ) = Im(f ′ (z0 )).
∂x ∂x
′ ′ ′
Since f (z0 ) = Re(f (z0 )) + i Im(f (z0 )), part of (6.4) follows.
Similarly, if f ′ (z0 ) exists, then
f (z0 + w) − f (z0 )
f ′ (z0 ) = lim
w→0
w∈iR
w
u(x0 , y0 + h) + iv(x0 , y0 + h) − u(x0 , y0 ) − iv(x0 , y0 )
= lim
h→0 ih
h∈R
v(x0 , y0 + h) − v(x0 , y0 ) u(x0 , y0 + h) − u(x0 , y0 )
= lim − i lim
h→0 h h→0 h
h∈R h∈R
∂v ∂u ∂u ∂v
= (x0 , y0 ) − i (x0 , y0 ) = −i (x0 , y0 ) + i (x0 , y0 ) .
∂y ∂y ∂y ∂y
Thus
∂v ∂u
(x0 , y0 ) = Re(f ′ (z0 )) and (x0 , y0 ) = − Im(f ′ (z0 )).
∂y ∂y
40 6. COMPLEX DIFFERENTIABILITY
The Cauchy–Riemann equations follow by equating the two expressions for the real
part of f ′ (z0 ) and the two expressions for the imaginary part of f ′ (z0 ), and the
remaining part of (6.4) also follows.. □
One consequence of the previous theorem is that if f is differentiable at every
point of an open set Ω in C, then the Cauchy–Riemann equations hold at every
point of Ω. Later on, we will see that in addition the four partial derivatives ∂u/∂x,
∂v/∂x, ∂u/∂y and ∂v/∂y are all continuous. For open sets Ω, the converse is true.
Theorem 6.8. If the four partial derivatives ∂u/∂x, ∂v/∂x, ∂u/∂y and ∂v/∂y are
all continuous in an open set Ω, then f is complex differentiable at z0 ∈ Ω if and
only if the Cauchy–Riemann equations hold at z0 , and if so, then
∂u ∂v
f ′ (x0 + iy0 ) = (x0 , y0 ) + i (x0 , y0 ).
∂x ∂x
We will justify this result later. When the partial derivatives are continuous in
a set that is not open, the function might be differentiable, or it might not be.
5. Examples
We revisit our previous examples, and add another, using the Cauchy–Riemann
equations.
Examples 6.9. (1) Suppose that f1 (z) = z 2 + iz + 2. Then
u(x, y) = x2 − y 2 − y + 2 and v(x, y) = 2xy + x,
so
∂u ∂u ∂v ∂v
= 2x, = −2y − 1, = 2y + 1 and = 2x,
∂x ∂y ∂x ∂y
and hence the Cauchy–Riemann equations hold for all (x, y) in R2 . Since the partial
derivatives are continuous and C is open, f1 is differentiable in C, and
f1′ (z) = 2x + i(2y + 1) = 2z + i.
(2) Suppose that f2 (z) = z. Then
u(x, y) = x and v(x, y) = −y,
so
∂u ∂u ∂v ∂v
= 1, = 0, = 0 and = −1,
∂x ∂y ∂x ∂y
and hence the Cauchy–Riemann equations do not hold for any (x, y) in R2 .
Hence f2 is not differentiable at any point in C.
(3) Suppose that f3 (z) = |z|2 . Then
u(x, y) = x2 + y 2 and v(x, y) = 0,
so
∂u ∂u ∂v ∂v
= 2x, = 2y, = 0 and = 0,
∂x ∂y ∂x ∂y
and hence the Cauchy–Riemann equations hold if and only if x = y = 0. The partial
derivatives are continuous in C, which is open, and hence f3 is differentiable at 0.
Finally, f is not differentiable at any other point than 0, since the Cauchy–Riemann
equations do not hold at any other point.
6. PROPERTIES OF THE DERIVATIVE 41
(2) We use d, ∂ and δ in this course and it is important to write these clearly
and correctly. Marks will be deducted for incorrectly written letters!
(3) It is common to use ux instead of ∂∂xu ; this is fine.
LECTURE 7
if, for every ε ∈ R+ , there exists δ ∈ R+ such that |u(x, y) − ℓ| < ε provided that
(x, y) ∈ Domain(f ) and 0 < |(x, y) − (x0 , y0 )| < δ. The same definition applies to a
vector-valued function F , provided that we interpret |F (x, y) − ℓ| as a vector length.
Given a complex function f , we associate real-valued functions u and v and an
R2 -valued function F of two real variables by the formulae
f (x + iy) = u(x, y) + iv(x, y)
F (x, y) = (u(x, y), v(x, y)).
The definitions of limit imply the following link between the real and complex
functions.
Connection 1. Let functions f , u, v and F be related as above. Then the
following are equivalent:
(1) f (z) → ℓ as z → z0 ;
(2) u(x, y) → Re ℓ and v(x, y) → Im ℓ as (x, y) → (x0 , y0 );
(3) F (x, y) → (Re ℓ, Im ℓ) as (x, y) → (x0 , y0 ).
This means that to most theorems about limits of real functions such as u and v,
there is a corresponding theorem about complex-valued functions f , and vice versa.
For example, a theorem about vector-valued functions states that a vector-valued
function tends to a limit ℓ if and only if each component of the function tends to the
corresponding component of ℓ. This is the analogue of the theorem that a complex-
valued function tends to a complex limit if and only if the real and imaginary parts
of the function tend to the real and imaginary parts of the limit.
Continuity for functions of two real variables is defined much as for functions of
a complex variable.
43
44 7. CONNECTIONS WITH CALCULUS IN THE PLANE† (NOT EXAMINABLE)
2. Differentiability
For functions of two real variables, we cannot define the derivative as for functions
of a real variable, because this would involve dividing by a vector. But we note that
an equivalent definition of the derivative of a function of one (real or complex)
variable is that f is differentiable at z0 and f ′ (z0 ) = D if
|f (z0 + h) − f (z0 ) − Dh|
lim = 0.
h→0 |h|
To see this, note that
|f (z0 + h) − f (z0 ) − Dh| f (z0 + h) − f (z0 ) − Dh
lim = lim
h→0 |h| h→0 h
f (z0 + h) − f (z0 )
= lim −D .
h→0 h
We define the derivative by extending this modified definition.
Definition 7.3. A real- or vector-valued function u of two real variables is
differentiable at (x0 , y0 ), and its derivative is the linear transformation D, if
|u((x0 , y0 ) + (h, k)) − u(x0 , y0 ) − D(h, k)|
lim = 0,
(h,k)→(0,0) |(h, k)|
or equivalently, if
|u(x, y) − u(x0 , y0 ) − D(x − x0 , y − y0 )|
lim = 0.
(x,y)→(x0 ,y0 ) |(x − x0 , y − y0 )|
The linear transformation D sends R2 to R if u is real-valued, and sends R2 to R2
if u is R2 -valued.
where the second line is a dot product of vectors, and the third is a product of two
matrices. The plane z = u(x0 , y0 ) + D(x − x0 , y − y0 ) in R3 is the tangent plane to
the surface z = u(x, y) in R3 at the point (x0 , y0 , u(x0 , y0 )).
To avoid having to deal with limits all the time, most treatments of multivariable
calculus prove the next result as soon as possible. For ease of notation, we state it
for a real-valued function, but it also holds for vector-valued functions.
Theorem 7.4. Suppose that u is a real-valued function of two real variables, and
that the partial derivatives ∂u/∂x and ∂u/∂y exist and are continuous in an open
set Ω. Then u is differentiable in Ω.
∂u ∂u ε ∂u ∂u ε
(x1 , y1 ) − (x0 , y0 ) < and (x1 , y1 ) − (x0 , y0 ) <
∂x ∂x 2 ∂y ∂y 2
whenever (x1 , y1 ) ∈ B((x0 , y0 ), δ). This is possible because Ω is open and because
both the partial derivatives are continuous at (x0 , y0 ). If (x, y) ∈ B((x0 , y0 ), δ),
then the line segments joining (x0 , y0 ) to (x, y0 ) and joining (x, y0 ) to (x, y) lie in
B((x0 , y0 , δ), by the geometry of balls in R2 .
The fundamental theorem of calculus and the mean value theorem for integrals
implies that there exist y1 between y and y0 and x1 between x and x0 such that
u(x, y) − u(x0 , y0 ) = u(x, y) − u(x, y0 ) + u(x, y0 ) − u(x0 , y0 )
Z y Z x
∂u ∂u
= (x, t) dt + (s, y0 ) ds
y0 ∂y x0 ∂x
∂u ∂u
= (x, y1 )(y − y0 ) + (x1 , y0 )(x − x0 ),
∂y ∂x
46 7. CONNECTIONS WITH CALCULUS IN THE PLANE† (NOT EXAMINABLE)
and so ∂p/∂y = ∂q/∂x. However there is no function on R2 \{(0, 0)} whose gradient
is F . When we integrate, we get the argument function (plus a constant), and this
cannot be defined continuously on R2 \ {(0, 0)}.
1. Examples
The Cauchy Riemann equations enable us to define new functions and show that
they are complex differentiable. For instance, recall the definition of the hyperbolic
functions:
ex + e−x ex − e−x
cosh(x) = and sinh(x) = ∀x ∈ R.
2 2
Recall also that the derivative of cosh is sinh and the derivative of sinh is cosh, and
that
cosh(x + y) = cosh(x) cosh(y) + sinh(x) sinh(y)
and
sinh(x + y) = sinh(x) cosh(y) + cosh(x) sinh(y).
We define two new functions of a complex variable as follows:
ch(x + iy) = cosh(x) cos(y) + i sinh(x) sin(y)
and
sh(x + iy) = sinh(x) cos(y) + i cosh(x) sin(y).
Exercise 8.1. Is ch differentiable? What is its derivative? What about sh?
Answer. As usual, we may write Log(x + iy) = u(x, y) + iv(x, y), where
u(x, y) = ln(x2 + y 2 )1/2 and v(x, y) = Arg(x + iy). Note that Log cannot be differ-
entiable at 0 since it is not defined there, nor on the negative real axis (−∞, 0), as
it is not continuous there.
In C \ (−∞, 0], which is open, we may apply the Cauchy–Riemann equations.
Observe that
∂u(x, y) 1 1 x
= 2 2x = ,
∂x (x + y 2 )1/2 2(x2 + y 2 )1/2 x2 + y 2
∂u(x, y) 1 1 y
= 2 2y = .
∂y (x + y 2 )1/2 2(x2 + y 2 )1/2 x2 + y 2
In the right half-plane where x > 0, we may write Arg(x + iy) = tan−1 (y/x),
and so
∂v(x, y) 1 −y −y
= 2 2
= 2 ,
∂x 1 + (y/x) x x + y2
∂v(x, y) 1 1 x
= = .
∂y 1 + (y/x)2 x x2 + y 2
In the upper half-plane where y > 0, we may write Arg(x+iy) = − tan−1 (x/y)+π/2;
in the lower half plane where y < 0, we may write Arg(x + iy) = − tan(x/y) − π/2.
Two very similar calculations show that the same formulae hold for the partial
derivatives in these cases too.
Hence the Cauchy–Riemann equations hold in the open set C \ (−∞, 0], and so
Log is differentiable in this set.
The derivative of Log is given by
∂u(x, y) ∂v(x, y) x − iy z̄ 1
Log′ (x + iy) = +i = 2 2
= =
∂x ∂x x +y z z̄ z
in C \ (−∞, 0]. 4
Im Im
0
Re Re
4. Polar coordinates
The Cauchy–Riemann equations are a very powerful tool. Consequently it is
worth stating them in polar coordinates as well.
Theorem 8.8. Suppose that the complex function f is differentiable at a point
z0 in C \ {0}, and that z0 = r0 eiθ0 . Then
∂u ∂v ∂v ∂u
(r0 , θ0 ) = −r0 (r0 , θ0 ) and (r0 , θ0 ) = r0 (r0 , θ0 ).
∂θ ∂r ∂θ ∂r
Further,
′ −iθ0 ∂u ∂v −ie−iθ0 ∂u ∂v
f (z0 ) = e (r0 , θ0 ) + i (r0 , θ0 ) = (r0 , θ0 ) + i (r0 , θ0 ) .
∂r ∂r r ∂θ ∂θ
Proof. Write z in polar coordinates as reiθ . Much as argued to prove the
Cauchy–Riemann equations, since f is differentiable at z0 ,
∂f f ((r0 + s)eiθ0 ) − f (r0 eiθ0 )
(r0 eiθ0 ) = lim
∂r s→0 s
f (r0 e + seiθ0 ) − f (r0 eiθ0 )
iθ0
= eiθ0 lim
s→0 seiθ0
f (z0 + seiθ0 ) − f (z0 )
= eiθ0 lim
s→0 seiθ0
iθ0 ′
= e f (z0 ),
and
∂f f (r0 ei(θ0 +φ) ) − f (r0 eiθ0 )
(r0 eiθ0 ) = lim
∂θ φ→0 φ
f (r0 ei(θ0 +φ) ) − f (r0 eiθ0 ) r0 ei(θ0 +φ) − r0 eiθ0
= lim
φ→0 r0 ei(θ0 +φ) − r0 eiθ0 φ
f (r0 e i(θ0 +φ)
) − f (r0 e )
iθ0
r0 e i(θ0 +φ)
− r0 eiθ0
= lim lim
φ→0 r0 ei(θ0 +φ) − r0 eiθ0 φ→0 φ
iθ0 ′
= ir0 e f (z0 ),
where we used trigonometric limits to evaluate the last limit.
We now have two formulae, both involving eiθ0 f ′ (z0 ); we equate these to get
∂f ∂f
(r0 eiθ0 ) = ir0 (r0 eiθ0 ).
∂θ ∂r
Now we write f in u + iv form, and see that
∂u ∂v ∂u ∂v
(r0 , θ0 ) + i (r0 , θ0 ) = ir0 (r0 , θ0 ) + i (r0 , θ0 ) .
∂θ ∂θ ∂r ∂r
Equating the real and imaginary parts gives the Cauchy–Riemann equations.
The formula for f ′ (z0 ) is found similarly. □
Remark 8.9. It is easiest to remember this version of the Cauchy–Riemann
equations in the form
∂f (reiθ ) ∂f (reiθ )
= ir .
∂θ ∂r
Corollary 8.10. The function Log is differentiable in C \ (−∞, 0].
54 8. PROPERTIES OF DIFFERENTIABLE FUNCTIONS
Proof. Write Log(reiθ ) in the form u(r, θ) + iv(r, θ). Then u(r, θ) = ln(r) and
v(r, θ) = θ; v is discontinuous when θ = π. Hence
∂u 1 ∂u
(r0 , θ0 ) = (r0 , θ0 ) = 0
∂r r0 ∂θ
∂v ∂v
(r0 , θ0 ) = 0 (r0 , θ0 ) = 1,
∂r ∂θ
unless θ = π. Hence the Cauchy–Riemann equations in polar form are satisfied
in C \ (−∞, 0]. Further, the partial derivatives are continuous in the open set
C \ (−∞, 0], so Log is differentiable in this set. □
5. Inverse functions
Suppose that Ω and Υ are open subsets of C, and that f is one-to-one from Ω
onto Υ. Then f has an inverse function, usually written f −1 , from Υ to Ω: we define
f −1 (w) = z if f (z) = w.
Theorem 8.11. Suppose that Ω and Υ are open subsets of C, that f : Ω → Υ is
one-to-one, and that f (z0 ) = w0 . If f is differentiable at z0 and f −1 is differentiable
at w0 , then (f −1 )′ (w0 ) = 1/f ′ (z0 ).
Proof. By definition, z = f −1 (f (z)); the chain rule implies that
df −1
1= (w0 ) f ′ (z0 ). □
dw
Later, we will investigate whether the inverse function is differentiable.
6. Definition
The examples above show that functions that are complex differentiable in an
open set have special properties; we are going to study them in much greater detail.
This justifies giving them a name.
Definition 8.12. Suppose that Ω is an open subset of C and f : Ω → C is a
function. If f is differentiable in Ω, that is, if it is differentiable at every point of Ω,
then we say that f is holomorphic or complex analytic or analytic or regular in Ω,
and we write f ∈ H(Ω).
If Ω = C and f is differentiable in Ω, then we say that f is entire.
LECTURE 9
Harmonic functions
1. Harmonic functions
We begin with a definition.
Definition 9.1. Suppose that u : Ω → R is a function, where Ω is an open
subset of R2 , and that u is twice continuously differentiable, that is, all the partial
derivatives ∂u/∂x, ∂u/∂y, ∂ 2 u/∂x2 , ∂ 2 u/∂x ∂y, ∂ 2 u/∂y ∂x and ∂ 2 u/∂y 2 exist and
are continuous. Then we say that u is harmonic in Ω if it satisfies Laplace’s equation:
∂ 2u ∂ 2u
+ = 0.
∂x2 ∂y 2
Remark 9.2. In three dimensions, where the three real variables are labelled
x, y and z, we say that a real-valued twice continuously differentiable function u is
harmonic if
∂ 2u ∂ 2u ∂ 2u
+ + = 0.
∂x2 ∂y 2 ∂z 2
If a function u on R3 is independent of the variable z, then we may consider it as
a function of x and y only, and then it is harmonic in the two-dimensional sense
above.
How do we find harmonic functions? The next theorem tells us that it is enough
to find holomorphic functions.
Theorem 9.3. Suppose that f ∈ H(Ω), where Ω is an open subset of C, that f
is twice continuously differentiable, and that
f (x + iy) = u(x, y) + iv(x, y)
for all x + iy in Ω, where u and v are real-valued. Then u and v are harmonic
functions.
Remark 9.4. Later we will see that f ∈ H(Ω) implies that f is actually in-
finitely differentiable, so the “twice continuously differentiable” hypothesis is not
really necessary.
Proof. Suppose that f ∈ H(Ω), where Ω is an open subset of C, and that
f (x + iy) = u(x, y) + iv(x, y)
55
56 9. HARMONIC FUNCTIONS
2. Examples
Exercise 9.5. Suppose that u(x, y) = x3 − 3xy 2 . Show that u is harmonic in C,
and find a function v such that the function f , given by f (x+iy) = u(x, y)+iv(x, y),
is holomorphic in C.
Answer. The partial derivatives of u are given by:
∂u ∂u
= 3x2 − 3y 2 = − 6xy
∂x ∂y
∂ 2u ∂ 2u
= 6x = − 6x.
∂x2 ∂y 2
Hence
∂ 2u ∂ 2u
+ = 6x − 6x = 0,
∂x2 ∂y 2
and u is harmonic.
If v exists such that the function f given by
f (x + iy) = u(x, y) + iv(x, y)
is holomorphic, then the Cauchy–Riemann equations must hold, so we must have
∂v
= 3x2 − 3y 2 (9.1)
∂y
∂v
= 6xy. (9.2)
∂x
Fix y ∈ R. From (9.2), ∂v(x, y)/∂x = 6xy, and integrating with respect to x
shows that
v(x, y) = 3x2 y + C1 ,
2. EXAMPLES 57
where C1 does not depend on x. However, C1 may depend on y. Thus the general
solution to equation (9.2), when we allow y to vary, is
v(x, y) = 3x2 y + c(y),
where c is an unknown function.
The function v must also satisfy (9.1), and this happens as long as
∂v
3x2 − 3y 2 = = 3x2 + c′ (y), (9.3)
∂y
i.e., c′ (y) = −3y 2 , so c(y) = −y 3 + C, where C is a constant, and
v(x, y) = 3x2 y − y 3 + C.
Set f (x + iy) = u(x, y) + iv(x, y). Then
f (z) = f (x + iy) = x3 − 3xy 2 + i(3x2 y − y 3 ) + iC = z 3 + iC,
which is holomorphic. 4
Note that, in the solution above, if we had just used the constant C1 rather than
the function c(y), then we would not have been able to satisfy (9.3).
Exercise 9.6. Suppose that u(x, y) = x2 − y 2 . Show that u is harmonic in C.
Find a function v such that the function f , given by f (x + iy) = u(x, y) + iv(x, y),
is holomorphic in C.
Answer. The partial derivatives of u are given by:
∂u ∂u
= 2x = − 2y
∂x ∂y
∂ 2u ∂ 2u
= 2 = − 2.
∂x2 ∂y 2
Hence
∂ 2u ∂ 2u
+ = 2 − 2 = 0,
∂x2 ∂y 2
and u is harmonic.
If v exists such that the function f given by
f (x + iy) = u(x, y) + iv(x, y)
is holomorphic, then the Cauchy–Riemann equations must hold, so we must have
∂v ∂u
= = 2x (9.4)
∂y ∂x
∂v ∂u
=− = 2y. (9.5)
∂x ∂y
The general solution to equation (9.4) is
v(x, y) = 2xy + c(x),
where c is an unknown function.
For (9.5) to hold as well, we require that
∂v
2y = = 2y + c′ (x)
∂x
58 9. HARMONIC FUNCTIONS
for all x. This tells us that c′ (x) = 0 for all x and c has to be a constant, C say.
Set f (x + iy) = u(x, y) + iv(x, y). Then
which is holomorphic. 4
in fact, often when we write out ux (x, y) − iuy (x, y), we can express this in terms of
z and hence find f .
Exercise 9.9. Suppose that u(x, y) = ln(x2 + y 2 )1/2 in R2 \ {(0, 0)}. Show that
u is harmonic. Can you find a function v such that the function f in C \ {0}, defined
by f (x + iy) = u(x, y) + iv(x, y), is holomorphic.
The trouble with this method is that it may be hard to express ux − iuy as a
function of z, or hard to find the integral of this.
4. PROOF OF THE EXISTENCE THEOREM† (NOT EXAMINABLE) 59
u is harmonic here
u is known here
Now
Z y Z x
∂v ∂ ∂u ∂u ∂u
(x, y) = (x0 , t) dt − (s, y) ds + C =− (x, y)
∂x ∂x y0 ∂x x0 ∂y ∂x
60 9. HARMONIC FUNCTIONS
Power series
In this lecture, we define and study complex power series. The proofs of many
theorems about complex power series are almost identical to the proof of the corre-
sponding theorem for real power series, and so we omit most proofs. Power series
are important for several reasons.
(a) There are formulae for manipulating them, so they may be used for calculations,
for instance, in MAPLE.
(b) Holomorphic functions may be expressed in power series, and, as we will see
later, vice versa.
61
62 10. POWER SERIES
2. Examples
X
∞
Exercise 10.4. Find the centre and radius of convergence of 3n−1 (z + 1)n .
n=0
Answer. The centre is −1. Take an = 3n−1 for all n ∈ N. Now |an |/|an+1 | =
1/3, so limn→∞ |an |/|an+1 | = 1/3, and ρ = 1/3 by the ratio test. 4
−1
This is a geometric series whose sum is 3(1 − 3(z + 1)) .
X
∞
(z − 2)n
Exercise 10.5. Find the centre and radius of convergence of .
n=0
n!
α(α − 1) . . . (α − n + 1)
an = .
n!
Answer. The centre is 1. By the ratio test,
|an | |α(α − 1) . . . (α − n + 1)| (n + 1)!
=
|an+1 | |α(α − 1) . . . (α − n + 1)(α − n)| n!
n+1 n+1 1
= = →1
|α − n| n |α/n − 1|
as n → ∞. Thus ρ = 1, and the series converges in B(1, 1). 4
This is a binomial series, which converges to (1 + (z − 1))α , that is, z α .
2. EXAMPLES 63
Exercise 10.7. Find the centre and radius of convergence of the series
X
∞
k(k + 1)(k 2 + 2)
(z + 3)k .
k=1
2k
Exercise 10.8. Find the centre and radius of convergence of the series
X
∞
(−j)j
(z − 5)j .
j=1
j!
Exercise 10.9. Find the centre and radius of convergence of the series
X zm
.
m∈N
2m
m even
This is as a geometric series with ratio z 2 /4, whose sum is 1/(1 − z 2 /4). The series
converges when |z 2 /4| < 1, that is, |z| < 2, so the radius of convergence is 2. 4
n=0 bn (z − z0 ) both converge in B(z0 , ρ), and that c ∈ C. Then the following
n
X
∞ X
∞ X
∞
(b) (an + bn )(z − z0 )n , and its sum is an (z − z0 )n + bn (z − z0 )n ;
n=0 n=0 n=0
X
∞ X
n
(c) cn (z − z0 ) , where c0 = a0 b0 , c1 = a0 b1 + a1 b0 , …, and cn =
n
aj bn−j ,
n=0 j=0
X
∞ X
∞
and its sum is an (z − z0 ) ×
n
bn (z − z0 )n .
n=0 n=0
in B(z0 , ρ).
Proof. We omit the proof. □
This theorem allows us to differentiate power series term by term.
P
Corollary 10.12. Suppose that f (z) = ∞ n=0 an (z − z0 ) in B(z0 , ρ). Then f
n
P∞
Corollary
P∞ 10.14. Suppose that f (z) = n=0 an (z − z0 )n and moreover that
g(z) = n=0 bn (z − z0 )n in B(z0 , ρ). If f (z0 + t) = g(z0 + t) for all t ∈ (−ε, ε), then
f (z) = g(z) for all z ∈ B(z0 , ρ).
Proof. We apply the previous corollary to f − g. □
There is a stronger version of Corollary 10.13 that says that if f is holomorphic
in a domain Ω, and f (zn ) = 0 for distinct points zn ∈ Ω such that zn → z0 ∈ Ω as
n → ∞, then f = 0. This leads to a stronger version of the last corollary.
These last corollaries will later lead us to the concept of analytic continuation:
if a function is defined in a domain Ω, then it is determined by its values in a small
set. In particular, if f is an entire function, then it is determined by its values on R.
This explains why, in finding an analytic function with certain properties, it suffices
to find it on R; this fact is useful in dealing with harmonic functions.
LECTURE 11
This is the only possible power series extension of the real exponential into the
whole P
complex plane. Indeed, we saw in the previous lecture that a complex power
series ∞ n
n=0 an z is determined by its values on any interval (−ε, ε).
y
y
cosh x
x
4. Graph sketching
Example 11.10. Find the images of the lines Re(z) = c and Im(z) = d under
the function cosh, and sketch these for various values of c and d.
axis between −∞ and −1, or between 1 and +∞; if cos(d) = 0, then w = ±i sinh(x),
and so w varies along the imaginary axis.
Some of these curves are sketched in Figure 11.1. In this figure, the blue lines
on the left are the lines Im(z) = d, where d = 0.5, 1.0, 1.5, 2.0, 2.5, and 3, and
the half-hyperbolae on the right are the images of these lines; as the value of d
increases from 0.5 to 3.0, the half-hyperbola moves to the left. The images of the
lines Im(z) = d where d = 0 or π would be the half-lines [1, +∞) and (−∞, −1]; it
seems evident, and it may be proved, that the images of the lines Im(z) = d where
0 < d < π fill the plane apart from the two half-lines mentioned above. The red line
segments on the left are the parts of the lines Re(z) = c, where c = 0.5, 1.0, 1.5 and
2.0, where the imaginary part varies from 0 t π. Note that we could complete the
red ellipses in the figure in two different ways: we could either include segments of
the lines Re(z) = c, where c is negative and the imaginary part varies from 0 to π,
or we could allow the imaginary part of the line segments already present to range
from −π to π. 4
Exercise 11.11. Find the images of the lines Re(z) = c and Im(z) = d under
the functions sinh, sin and cos, and sketch these for various values of c and d.
5. Harmonic functions
Exercise 11.12. Suppose that u(x, y) = cos(x) sinh(y) for all x, y ∈ R. Show
that u is harmonic and find its harmonic conjugate.
Answer. First,
∂ 2 u(x, y) ∂ 2 u(x, y) ∂ 2 cos(x) ∂ 2 sinh(y)
+ = sinh(y) + cos(x)
∂x2 ∂y 2 ∂x2 ∂y 2
= − cos(x) sinh(y) + cos(x) sinh(y)
= 0.
Thus u is harmonic.
The harmonic conjugate v of u must satisfy:
∂v(x, y) ∂u(x, y)
=− = − cos(x) cosh(y)
∂x ∂y
and
∂v(x, y) ∂u(x, y)
= = − sin(x) sinh(y).
∂y ∂x
It follows that
v(x, y) = − sin(x) cosh(y) + c(y),
where c(y) depends only on y, and then
∂v(x, y)
sin(x) sinh(y) + c′ (y) = = − sin(x) sin(y),
∂y
so c′ (y) = 0, that is, c(y) = C, a constant. Then v(x, y) = − sin(x) cosh(y) + C.
72 11. EXPONENTIAL, HYPERBOLIC AND TRIGONOMETRIC FUNCTIONS
1. Some algebra
Many complex functions are not one-to-one. This means that inverse functions
are not well-defined. One solution is to deal with “multi-valued functions”, and
another is to restrict the domain of the original function. The archetypical example
is the argument function, and we usually write arg(z) to indicate the multi-valued
function and Arg(z) to indicate the particular choice where the values lie in the
range (−π, π].
√
Example 12.1. Suppose that w = z 2 . Then we may write z = w or w1/2 ; the
question is what this means. Unfortunately, different writers mean different things.
Some writers mean that z may be any of the two possible values; others mean that
a particular choice has been made. We will try to use the expressions above for the
multi-valued function, and add the words “principal value” (symbolically, PV) or
“principal branch” to indicate that a particular choice is being made. In particular,
we define
(
1/2 |w|1/2 ei Arg(w)/2 if w 6= 0
PV w =
0 if w = 0.
In words, we might say “the principal branch of the square root”.
We may think of the graph of the multi-valued function w1/2 as being like two
copies of the plane, slit along the branch cut, and then joined together cleverly. It
is not possible to do this in three dimensions, but it is possible in four dimensions.
Example 12.2. Suppose that w = ez and z = x + iy. Then w = ex eiy , so
|w| = ex and Arg w = Arg eiy .
Then x = ln |w|, and x is single-valued, but y = Arg(w) + 2πk, where k ∈ Z; and
y is multiple-valued. When w 6= 0, we write z = log(w) to indicate that z may be
any one of the infinitely many complex numbers such that ez = w and we write
z = Log(w) to indicate the choice that z = ln |w| + i Arg(w).
Im Im
z 7→ z 6
Ω
Re Υ Re
z 7→ z 1/6
Im Im
z 7→ z 6
Re Υ Re
Ω
z 7→ z 1/6
Note that where this function is differentiable and where it is continuous are
nearly the same; the only difference is that PV z 1/n is continuous at 0 but not
differentiable.
All the possible inverse functions are constant multiples of each other on con-
nected sets where both are defined, where the multiplying factors are nth roots of
unity. The different possible inverse functions are called branches of the nth root
function.
3. SQUARE ROOTS OF POLYNOMIALS 75
Figure 12.1 illustrates some possibilities in the case where n = 6. In the upper
figure, we usually choose the inverse function on the negative real axis to map to
the upper side of the cone. It is given by f −1 (w) = |w|1/6 ei Arg(w)/6 .
In the lower figure, we usually choose the inverse function to map the positive
real axis to the lower side of the cone. It is given by f −1 (w) = |w|1/6 ei Arg (w)/6 ,
0
where Arg0 denotes the choice of argument in the range [0, 2π).
The first inverse function is the “standard” choice; it is called the principal value
of the 6th root; the notation PV z 1/6 should be used.
The regions Ω that we chose are relatively simple, but we could have chosen
more complicated versions, such as a region between two curves coming out of the
origin. This would have led to a curved branch cut.
In general, when we restrict a function f to a smaller domain in order to define
an inverse function for f , we try to make the smaller domain Ω as large as possible,
so that the domain Υ of the inverse function in as large as possible. The boundary
of Υ is called the branch cut (or cuts, as it may have a number of connected pieces).
By choosing Ω and Υ carefully, we may usually avoid having an inverse function
with discontinuities where we want to work. Although we may vary these sets by
varying the branch cut, there are some points, called branch points, which must
appear in any branch cut. For the case of the function f (z) = z n , the only branch
point is 0.
(z − 1) = PV(z − 1) =
2 1/2 2 1/2
(12.1)
0 when z = ±1.
Im
Arg(z 2 − 1) → −π Arg(z 2 − 1) → π
bc bc
Re
Arg(z 2 − 1) → π Arg(z 2 − 1) → −π
Let us now determine whether these possible definitions coincide and where they
are continuous and where they are differentiable.
First, we consider |z 2 − 1|1/2 ei Arg(z −1)/2 . Since |z 2 − 1|1/2 is continuous, the
2
so the function is continuous at ±1. There are many possibilities to consider, best
represented in a diagram—see Figure 12.2. We deduce that this function is contin-
uous in C \ (iR ∪ (−1, 1)).
We may check that this function is differentiable in C \ (iR ∪ [−1, 1]): indeed, it
is the composition of differentiable functions there.
Next, we consider |z 2 −1|1/2 ei Arg(z−1)/2+i Arg(z+1)/2 . Since |z 2 −1|1/2 is continuous,
the question is linked to where ei Arg(z−1)/2+i Arg(z+1)/2 is continuous. The possible
discontinuities are when Arg(z − 1) or Arg(z + 1) are discontinuous, that is, when
z is real and z ≤ 1. Clearly
lim |z 2 − 1|1/2 ei Arg(z−1)/2+i Arg(z+1)/2 = 0,
z→±1
so the function is continuous at ±1. Observe that, on the one hand, if −1 < x0 < 1,
then
lim Arg(z − 1)/2 + Arg(z + 1)/2 = π/2
z→x 0
Im(z)>0
The next step up in complication is the square root of a cubic polynomial. For
instance, we might take the multi-valued function w = (z(z 2 − 1))1/2 . “Graphs” of
functions like this are called elliptic curves. They are important in number theory—
they are central to Andrew Wiles’ proof of Fermat’s last theorem—and they are
important in cryptography—they give rise to “better” codes.
Exercise 12.5. Show that there is a choice of definition of (z(z 2 − 1))1/2 that
is continuous except on the intervals (−∞, −1) and (0, 1).
The graph of the multi-valued function is like two copies of the plane, with slits
along the branch cuts, joined together appropriately.
78 12. LOGARITHMS AND ROOTS
4. More examples
The logarithm log is multi-valued because the argument arg is multi-valued.
There are two common choices of argument, one between −π and π, and the other
between 0 and 2π. Both of these give logarithms with the property that log(i) = π/2,
but the first has its branch cut along the negative real axis and the second has its
branch cut along the positive real axis. If we want to deal with a logarithm that is
continuous around 1, the first choice is better, but if we want to arrange continuity
around −1, then the second is better.
Soon we will need to look at expressions such as log(w + (w2 − 1)1/2 ). We try to
choose a branch of log and a branch of the square root to make the function con-
tinuous in as large a domain as possible. Usually this function will be differentiable
where it is continuous, with some possible exceptions.
(1) While the arg–Arg notation seems fairly√standard, there does not seem to
be any standard about whether z 1/n or n z is uniquely defined. We use the
word “multi-valued” wherever appropriate to try to avoid any ambiguity.
LECTURE 13
In this lecture, we discuss the inverse functions for the exponential, hyperbolic
and trigonometric functions introduced in the last lecture.
1. Inverse functions
The following exercise is a warm-up.
Exercise 13.1. Fix w in C. Find all z in C such that
(a) exp(z) = w, (b) cosh(z) = w, (c) sinh(z) = w.
Im Im
i(θ + 2π) b
Sθ
Re Re
b
Rθ
iθ
exp
are both defined, The point 0, which is common to all the branch cuts, is called a
branch point.
Lemma 13.3. For any branch logθ of the complex logarithm,
1
log′θ (w) =
w
for all w ∈ C \ Rθ .
The notation logθ is not standard, and we will not use it any more. Rather, we
use the expression “the branch of the logarithm with imaginary part in (θ, θ + 2π)”.
A different sort of “inverse function” of the exponential function is the “multi-
function” log from C \ {0} to C given by
log(z) = ln |z| + i arg(z).
This is a “multifunction” in the sense that it takes multiple values, because arg(z)
takes multiple values.
3. Complex powers
We define complex powers of complex numbers using exponentials and loga-
rithms.
Definition 13.4. Given z ∈ C \ {0} and α ∈ C, we define
z α = exp(α log(z)).
The principal branch of z α is found by using Log, the principal branch of the loga-
rithm. That is, PV z α = exp(α Log(z)).
Often we write ez rather than exp(z). Note that this is ambiguous, since complex
powers are multi-valued! So arguably exp(z) is better notation.
82 13. INVERSES OF EXPONENTIAL AND RELATED FUNCTIONS
Both the logarithm and the square root are possible causes of discontinuity. The
function PV(w2 + 1)1/2 is continuous as long as w2 + 1 is not in the interval (−∞, 0],
and the logarithm is continuous as long as w + PV(w2 + 1)1/2 is not in the interval
(−∞, 0].
On the one hand, if w2 + 1 is not in (−∞, 0], then w2 is not in (−∞, −1]. So
one possible discontinuity is when w = iv, where |v| ≥ 1.
4. THE INVERSE HYPERBOLIC SINE 83
On the other hand, we may try to solve the equation w + PV(w2 + 1)1/2 = −t
for t ∈ [0, ∞); we get
PV(w2 + 1)1/2 = −t − w
w2 + 1 = t2 + w2 + 2tw
1 − t2
w= ,
2t
and so w is real. But if w is real, then
w + PV(w2 + 1)1/2 = w + (w2 + 1)1/2 > 0,
and so w + PV(w2 + 1)1/2 is not in the interval (−∞, 0]. Thus the only possible
discontinuities are when w = iv, where v is real and |v| ≥ 1.
Lemma 13.9. The principal branch of the inverse hyperbolic sine function is
differentiable in C \ ([i, +i∞) ∪ (−i∞, −i]). Further,
′ 1
PV sinh−1 (w) = √ .
PV w2 + 1
Proof. We compute the derivative:
In this lecture, we define paths and path integrals, and see a key theorem about
these. This material should be familiar to students who have studied multi-variable
calculus.
The main question underlying this lecture is the relation between integration
and differentiation. As before, we represent points in R2 by row vectors rather than
by column vectors.
1. Curves
Definition 14.1. A curve γ in R2 is a continuous function from an interval [a, b]
of real numbers into R2 . We may write γ(t) = (γ1 (t), γ2 (t)), where γ1 and γ2 are
real-valued; then the continuity of γ is equivalent to the continuity of both γ1 and
γ2 .
The initial point of the curve is γ(a) and the final point of the curve is γ(b). The
range of the curve is the set of points {γ(t) : t ∈ [a, b]}.
A curve γ : [a, b] → R2 is said to be closed if γ(a) = γ(b), and simple if
γ(s) 6= γ(t) when s < t, except perhaps if s = a and t = b.
Examples 14.2. In Figure 14.1, the curve α moves from p to q, the curve β
moves up the parabolic arc, the curve γ moves from the right to the left of the
line segment and then back again, the curve δ moves once around the circle in the
anticlockwise direction, starting at the right-most point, the curve ε moves twice
around the “infinity-shaped” figure, starting at the right-most point and moving
upwards, and the curve ζ moves along the “alpha-shaped” figure.
The curves α and β are simple but not closed; γ and ε are closed but not simple;
δ is both simple and closed; and ζ is neither closed not simple. The curve γ in the
figure “goes back on itself”, while the curves ε and ζ intersect themselves. The
curve ε repeats itself.
Definition 14.3. Suppose that α : [a, b] → R2 and β : [c, d] → R2 are curves such
that the final point of [a, b] is the initial point of [c, d] and the final point of α is the
initial point of β, that is, α(b) = β(c). The join α t β of α and β is the curve
(
α(t) when a ≤ t ≤ b
α t β(t) =
β(t) when c ≤ t ≤ d.
85
86 14. PATHS AND PATH INTEGRALS† (NOT EXAMINABLE)
y
α(t) = p + (q − p)t b q
= p(1 − t) + qt ∀t ∈ [0, 1]
α
β(t) = ( 23 + t, t ) ∀t ∈ [0, 1]
2
p b
ε
b
β δ
γ(t) = (1 + 2t , −3) ∀t ∈ [−1, 1]
2
x
b b b
δ(t) = (1 + 22
10
cos(t), 22
10
sin(t)) ∀t ∈ [0, 2π]
ε(θ) = (cos(θ), sin(θ) cos(θ)) ∀θ ∈ [0, 4π]
ζ
ζ(θ) = (cos(θ) − 2, sin(θ) cos(θ) − 3) b
γ
∀θ ∈ [π/4, 7π/4]
b
Example 14.4. In Figure 14.2, the join αtβ tγ tδ moves anticlockwise around
the perimeter of the square with vertices (0, 0), (1, 0), (1, 1) and (0, 1), starting and
ending at (0, 0), as t varies from 0 to 4. The curve α∗ moves from 1 to 0 as t varies
between 0 and 1. What does (α t β t γ t δ)∗ do?
Examples 14.7. In Figure 14.3, both curves α and β are continuously differ-
entiable. Only α is smooth.
2. Orientation
If γ is a simple curve that is not closed, then it has an initial point and an
endpoint and a “direction of motion”, and even if we reparametrise it, these do not
change. Similarly, simple closed curves have a “direction of motion”. We call this
orientation. For simple closed curves, the next theorem lets us find an orientation
that does not depend on the parametrisation.
Theorem 14.9 (The Jordan curve theorem). If γ : [a, b] → R2 is a simple closed
curve, then the complement of the range of γ is the union of two disjoint domains.
One of these is bounded and the other is not. The bounded component is called the
interior of γ and written Int(γ), and the unbounded component is called the exterior
of γ and written Ext(γ).
Proof. We do not prove this, but remark that it relies on approximating a
simple curve by a simple polygonal curve. □
This theorem seems obvious but is actually quite difficult, particularly when
curves such as snowflakes are concerned. We will consider only curves for which it
is easy to identify the interior and exterior.
If γ is a simple closed curve, then as we travel around γ, the interior of γ will
always lie on our left, or always on our right. The standard orientation of γ is the
direction of motion such that the interior is always on our left. With the standard
orientation, we move around the perimeter of Int(γ) in an anti-clockwise direction.
88 14. PATHS AND PATH INTEGRALS† (NOT EXAMINABLE)
Exercise 14.10. What is the standard orientation of each of the simple closed
curves described above? Can you define orientation for closed curves that are not
simple?
We will use the expression oriented range of a curve to describe the image of a
simple curve in R2 and a “direction of motion” along the curve.
Again, γ ′ (t) may not be defined for finitely many points, where different con-
tinuously differentiable curves are joined. We take γ ′ (t) to be 0 wherever it is not
defined, and then the integral exists as a Riemann integral.
We now consider the approximation of piecewise smooth curves by polygonal
curves. If γ : [a, b] → R2 is a piecewise smooth curve, we may define “approximating
curves” γ N (where N ∈ Z+ ) as follows. Fix N , and subdivide the interval [a, b] into
N equal subintervals of equal length, [an−1 , an ] say, where n = 1, . . . , N (we do
this by choosing an = (N − n)a/N + nb/N ). Now let γ N be the polygonal curve
composed of the N line segments from γ(an−1 ) to γ(an ), in order.
Lemma 14.12. Suppose also V is a continuous vector field in a domain Ω in R2 .
Suppose that γ is a piecewise smooth curve in Ω, and that the polygonal curves γ N
defined above also lie in Ω. Then the curves γ N approximate γ, in the sense that
γ N (t) → γ(t) for all t ∈ [a, b] and Length(γ N ) → Length(γ) as N → ∞; further,
Z Z
V (s) ds → V (s) ds.
γN γ
Proof. Omitted. □
The point of this lemma is that we may prove many results for integration along
general curves by proving easier results for integration along polygonal curves.
The line integral has the usual properties of integration.
Theorem 14.13. Suppose that λ, µ ∈ R, that γ : [a, b] → R2 is a piecewise
smooth curve, and that V and W are vector fields defined on Range(γ). Then the
following hold.
(a) The integral is linear:
Z Z Z
λV (s) + µW (s) ds = λ V (s) ds + µ W (s) ds.
γ γ γ
3. VECTOR FIELDS AND LINE INTEGRALS 89
(e) The size of the integral is controlled by the size of the vector field V and the
length of the curve γ:
Z
V (s) ds ≤ M L,
γ
where L is the length of γ and M is a number such that |V (s)| ≤ M for all
s ∈ Range(γ).
We may extend the notation for joins and reverse curves to paths.
Note that for paths that are not closed, different parametrisations differ by a
change of variable, and part (d) of the theorem applies. But for closed paths,
parametrisations might have different initial points and endpoints. To show that
the integral around a closed path does not depend on where we start, we break up
the closed path Γ into two parts: Γ = A t B, and observe that
Z Z Z
V (s) ds = V (s) ds + V (s) ds
A⊔B
Z A
Z B
Z
= V (s) ds + V (s) ds = V (s) ds.
B A B⊔A
This implies that the integral around Γ starting at the initial point of A is the same
as the integral around Γ starting at the initial point of B.
90 14. PATHS AND PATH INTEGRALS† (NOT EXAMINABLE)
R Proof. We omit this proof, but mention that the previous theorem implies that
γ
V (s) ds depends only on the initial and final points of the curve γ, which allows
us to define Z
u(q) = V (s) ds
γ
for any curve γ from a fixed point p to q. □
This theorem has variants; in particular, it is equivalent to the statement that a
line integral depends only on the initial point and the end point.
(1) We have seen that not all vector fields can be integrated “nicely”. However,
the integral of a derivative is the function that we started with, up to a
constant of integration.
(2) We will prove similar results in the context of complex analysis, where line
integrals become contour integrals, vector fields become complex-valued
functions, and the condition for a vector field to be closed becomes the
Cauchy–Riemann equations.
(3) The material in this lecture is not examinable, except that we re-use without
further mention definitions such as closed and simple (for curves) in future
lectures.
LECTURE 15
Contour integrals
δ(t) = (t − 1) − it ∀t ∈ [0, 1]
ε(t) = t + i(t − 1) ∀t ∈ [0, 1]
provided that the two real integrals on the right hand side exist. In other words,
Z b Z b Z b Z b
Re f (t) dt = Re (f (t)) dt and Im f (t) dt = Im (f (t)) dt.
a a a a
We prove only the last part: if the left hand side is 0, there is nothing to prove,
otherwise we take θ to be the argument of the left hand side. Then
Z b Z b Z b
−iθ
f (t) dt = e f (t) dt = e−iθ f (t) dt.
a a a
3. CONTOUR INTEGRALS 93
3. Contour integrals
Definition 15.6. Given a piecewise smooth curve γ : [a, b] → C and a contin-
uous (not necessarily differentiable)
R function f defined on the range of γ, we define
the complex line integral γ f (z) dz by
Z Z b
f (z) dz = f (γ(t)) γ ′ (t) dt,
γ a
(e) The size of the integral is controlled by the size of the function f and the length
of the curve γ: Z
f (z) dz ≤ M L,
γ
where L is the length of γ and M is a maximiser for |f | on the curve, that is, a
number such that |f (z)| ≤ M for all z ∈ Range(γ).
Part (e) of the theorem is often called the M L Lemma. Note that M is not
necessarily the maximum of |f | on the curve.
Recall that a contour Γ is the oriented range of a piecewise smooth curve γ. The
theorem above implies that the complex line integral depends only on Γ, and not
on the parametrisation γ.
Definition 15.8. We define
Z Z
f (z) dz = f (z) dz,
Γ γ
Next
Z Z 1 Z 1
′
z dz = γ(t) γ (t) dt = [p + t(q − p)] (q − p) dt
γ 0 0
q 2 − p2
= p(q − p) + 12 (q − p)2 = .
2
Third,
Z Z 1 Z 1
′
z dz = γ(t) γ (t) dt = p + t(q − p) (q − p) dt
γ 0 0
(q + p)(q − p)
= p(q − p) + 12 (q − p)(q − p) = .
2
Finally,
Z Z 1 Z 1
′
z
e dz = e γ(t)
γ (t) dt = ep et(q−p) (q − p) dt
γ 0 0
h it=1
= ep et(q−p) = ep e(q−p) − 1 = eq − ep .
t=0
Hence
Z
λ1 + λ2 z + λ3 z + λ4 ez dz
γ
q 2 − p2 (q + p)(q − p)
= λ1 (q − p) + λ2 + λ3 + λ4 (eq − ep ). 4
2 2
Alternatively, if λ3 = 0, then we may write
Z Z 1
z
λ1 + λ2 z + λ4 e dz = λ1 + λ2 γ(t) + λ4 eγ(t) γ ′ (t) dt
γ 0
Z 1
dγ(t) λ2 dγ 2 (t) deγ(t)
= + λ1 + λ4 dt
0 dt 2 dt dt
h λ2 it=1
= λ1 γ(t) + γ 2 (t) + λ4 eγ(t)
2 t=0
q −p
2 2
= λ1 (q − p) + λ2 + λ4 (eq − ep ).
2
96 15. CONTOUR INTEGRALS
We would get the same answer if the parameter t varied from −kπ to kπ.
Alternatively, we may write, when n 6= 0,
Z Z 2π Z 2π h i2π
n n ′ 1 dγ n+1 (t) 1−1
z dz = γ (t) γ (t) dt = dt = γ n+1 (t) = = 0.
γ 0 0 n+1 dt 0 n+1
When n = −1,
Z Z 2π ′ Z 2π h i2π
1 γ (t) d log γ(t)
dz = dt = dt = log γ(t) = 2πi,
γ z 0 γ(t) 0 dt 0
In this lecture, we begin with an exercise, and then state and discuss one of the
key theorems of complex analysis.
1. An exercise
Proof. We prove a more general version of this result later for the case in which
Γ is simple. To extend from the simple case to the general case, we argue as we did
for path integrals in Lecture 13. □
Now we take p ∈ Ω, and show that F ′ (p) = f (p). To do so, we need to make
F (q) − F (p)
− f (p)
q−p
small by taking q sufficiently close to p.
Take q ∈ Ω close to p, and let ∆ be the line segment from p to q. On the one
hand, Z
F (q) − F (p) 1
= f (z) dz;
q−p q−p ∆
3. MULTIPLY CONNECTED DOMAINS 99
this is correct
R provided that ∆ is contained in Ω. On the other hand, by direct
calculation, ∆ dz = q − p, whence
Z
1
f (p) = f (p) dz.
q−p ∆
Thus
Z Z
F (q) − F (p) 1 1
− f (p) = f (z) dz − f (p) dz
q−p q−p ∆ q−p ∆
Z
1
= f (z) − f (p) dz .
q−p ∆
We can make this small by making ∆ small enough that ∆ is contained in Ω and f
does not vary much on ∆(f is differentiable and hence continuous), and then using
the M L Lemma.
More precisely, take any small positive ε. Since Ω is open and f is continuous at
p, there exists δ such that B(p, δ) ⊂ Ω and |f (z) − f (p)| < ε when z ∈ B(p, δ). Take
q ∈ B ◦ (p, δ) and let ∆ be the straight line segment from p to q. Then ∆ ⊂ B(p, δ)
and so |f (z) − f (p)| < ε for all z ∈ ∆. Thus
Z
F (q) − F (p) 1
− f (p) = f (z) − f (p) dz
q−p q−p ∆
1
≤ max{|f (z) − f (p)| : z ∈ ∆} |q − p|
|q − p|
= max{|f (z) − f (p)| : z ∈ ∆} < ε,
by the M L lemma and the fact that max{|f (z) − f (p)| : z ∈ ∆} = f (z ∗ ) for some
z ∗ ∈ ∆ (the maximum is attained). It follows that F is differentiable at p, with
derivative f (p), as required.
If F1 is another function such that F1′ = f , then (F1 − F )′ = 0, so F1 − F is a
constant, C say. This means that
F1 (q) − F1 (p) = (F (q) + C) − (F (p) + C) = F (q) − F (p),
R
so that F1 can also be used to compute Γ f (z) dz. □
Im
Γ2
Re
Γ0
Γ1
functions which include the hypotheses that f is holomorphic and that f ′ is con-
tinuous, and it is useful to know that the continuity hypothesis is automatically
true. At least in principle, we should check that the hypotheses of a theorem are
satisfied before we apply the theorem, and so it is good to make these hypotheses
unnecessary.
Augustin Cauchy was one of the finest French mathematicians of the first half of
the 1800s, and he developed much of what is in a course on complex analysis today, as
well as making precise the idea of limit that had been worrying mathematicians and
philosophers of mathematics since the time of Newton and Leibnitz. The Cauchy–
Goursat theorem is Goursat’s 1884 improvement of a theorem of Cauchy from 1829—
it says something of Goursat’s ability that he could improve the work of Cauchy.
Some of Cauchy’s ideas were being developed simultaneously by George Green,
an “uneducated miller” from Nottingham, who gave us Green’s theorem in 1828.
Proof. The proof of the theorem involves three steps. First, we prove it in the
case where Ω is a triangle. Second, we consider the case where Ω is a domain whose
boundary is made up of finitely many closed polygonal contours. Third, we treat
the general case.
Step one. Suppose that Ω is a triangle in the complex plane. We write T0 for Ω
and ∂T0 for its boundary. Suppose that f ∈ H(Υ), where T0 ⊂ Υ, and let
Z
f (z) dz = I.
∂T0
T′′′′
′′
T
T′
T′′′
Now
Z
I= f (z) dz
Z ∂T0
Z Z Z
= f (z) dz + f (z) dz + f (z) dz + f (z) dz.
∂T′ ∂T′′ ∂T′′′ ∂T′′′′
At least one of the triangles T′ , T′′ , T′′′ and T′′′′ , which we call T1 , must satisfy
Z Z
1 |I|
f (z) dz ≥ f (z) dz = .
∂T1 4 ∂T0 4
We now subdivide T1 into 4 congruent triangles, and argue in the same way that
there must be one of these, T2 say, with the property that
Z Z
1 |I|
f (z) dz ≥ f (z) dz ≥ .
∂T2 4 ∂T1 16
Continuing inductively in this way, we find a sequence (Tn )n∈N of nested trian-
gles, such that
Z
|I|
f (z) dz ≥ n . (17.1)
∂Tn 4
Write Length(∂Tn ) for the perimeter of Tn . Then
Length(∂Tn ) = 2−n Length(∂T0 ).
The first two integrals on the right hand side are 0, by calculation, and hence by
(17.1), the last equation, the M L Lemma, properties of maxima, and (17.2),
Z
|I| ≤ 4 n
f (z) dz
Z ∂Tn
n
=4 E(z) dz
∂Tn
≤ 4 max{|E(z)| : z ∈ ∂Tn } Length(∂Tn )
n
|E(z)|
n
= 4 max |z − z0 | : z ∈ ∂Tn Length(∂Tn )
|z − z0 |
|E(z)|
≤ 4 max
n
: z ∈ ∂Tn max {|z − z0 | : z ∈ ∂Tn } Length(∂Tn )
|z − z0 |
|I| Length(∂Tn )2
≤ 4n
Length(∂T0 )2 2
|I|
= ,
2
which is absurd. Hence I = 0.
Step 2. The next step is to deal with a domain Ω with a polygonal boundary.
Any such domain may be subdivided into triangles Tn (although this may seem
obvious, it is hard to prove), in such a way that
Z XZ
f (z) dz = f (z) dz;
∂Ω n ∂Tn
Step 3. Finally, we have to deal with a domain whose boundary is the union of
finitely many disjoint closed contours. This can be done by approximating unions of
closed contours by unions of polygonal contours; the integral is 0 for all the unions of
approximating polygonal contours, and so the integral around the union of general
contours that we want is also 0. □
Proof. Let Γε be the circle with centre w and radius ε, traversed clockwise,
and take ε small enough that Γε ⊂ Int(Γ). We consider the domain Υ consisting
of Int(Γ) ∩ Ext(Γε ), the domain between Γ and Γε , whose boundary consists of
Γ, traversed anti-clockwise, and Γε , traversed clockwise. The quotient function
z 7→ f (z)/(z − w) is holomorphic in Ω \ {w}, a domain that contains Υ ∪ ∂Υ.
106 17. CAUCHY’S INTEGRAL FORMULA
that is,
Z Z
f (z) f (z)
dz = dz.
Γ z−w Γ∗ε z−w
The left hand side of this equality does not depend on ε, so the limit as ε tends to
0 of the right hand side exists.
To compute the limit, we parametrise Γ∗ε . Define γε∗ (θ) = w + εeiθ , where 0 ≤
θ ≤ 2π, and observe that
Z Z
f (z) f (z)
dz = lim dz
Γ∗ε z − w ε→0 Γ∗ z − w
Z 2π
ε
f (w + εeiθ )
= lim iεeiθ dθ
ε→0 0 εeiθ
Z 2π
= lim f (w + εeiθ ) i dθ
ε→0 0
Z 2π
=i lim f (w + εeiθ ) dθ
0 ε→0
Z 2π
=i f (w) dθ
0
= 2πif (w).
We can move the limit inside the integral because limε→0 f (w + εeiθ ) = f (w) uni-
formly in θ, since limz→w f (z) = f (w). Formula (17.3) follows. □
Proof. By Cauchy’s integral formula, both sides are equal to 2πif (w). □
Z
f (z)
This means that if we need to compute the integral dz, we may change
Γ z −w
the contour to make the calculation easier.
Corollary 17.3 (mean value formula). Suppose that Ω is a simply connected
domain in C, that f ∈ H(Ω), and that w ∈ Ω. If B(w, r) ⊂ Ω, then
Z 2π
1
f (w) = f (w + reiθ ) dθ. (17.4)
2π 0
4. A COMPUTATION 107
Proof. This formula is virtually proved in the course of the proof of the Cauchy
integral formula; let γ(θ) = w + reiθ , where 0 ≤ θ ≤ 2π. Then
Z
1 f (z)
f (w) = dz
2πi γ z − w
Z 2π
1 f (w + reiθ )
= ireiθ dθ
2πi 0 reiθ
Z 2π
1
= f (w + reiθ ) dθ,
2π 0
as required. □
The Cauchy integral formula may be considered as a way to write f (w) as a
weighted average of the values of f (z) around any contour surrounding w.
4. A computation
Cauchy’s integral formula enables us to compute some integrals without inte-
grating!
Z
sin z
Exercise 17.4. Compute dz, where Γ is the circle with centre 0 and
Γ z
radius R.
Answer. Take f (z) = sin z and z0 = 0, and apply Cauchy’s integral formula:
Z Z
sin z f (z)
dz = dz = 2πif (z0 ) = 0. 4
Γ z Γ z − z0
Corollary 18.1. Suppose that f ∈ H(B(z0 , R)), and that Γ is a simple closed
contour in B(z0 , R) such that z0 ∈ Int(Γ). Then
X
∞
f (w) = cn (w − z0 )n ∀w ∈ B(z0 , R),
n=0
where
Z
1 f (z)
cn = dz.
2πi Γ (z − z0 )n+1
The radius of convergence of the power series is at least R.
Remark 18.2. This corollary, combined with the fact that f (n) (z0 ) = n! cn ,
implies that
Z
(n) n! f (z)
f (z0 ) = dz.
2πi Γ (z − z0 )n+1
This is often called Cauchy’s generalised integral formula.
Notice that we just assumed that f is differentiable once; Corollary 18.1 implies
that f is actually infinitely differentiable.
Proof. Write Γr for the circle with centre z0 and radius r, where r < R. By
independence of contour,
Z Z
f (z) f (z)
dz = dz,
Γ (z − z0 ) Γr (z − z0 )
n+1 n+1
109
110 18. CAUCHY’S GENERALISED INTEGRAL FORMULA
where Z Z
1 f (z) 1 f (z)
cn = dz = dz,
2πi Γr (z − z0 )n+1 2πi Γ (z − z0 )n+1
by independence of contour. (Here we have exchanged the order of summation and
integration.) Once we know that f has this power series representation, it follows
(n)
that cn = f n!(z0 ) from results on power series in Lecture 9.
We chose r and w such that |w − z0 | < r < R. If we take any w ∈ B(z0 , R),
then we may choose r such that these inequalities hold, so the series converges at
w. Since w is an arbitrary element of B(z0 , R), the series converges in B(z0 , R), and
the radius of convergence is at least R. □
where
Z
(n) n! f (z)
f (0) = dz.
2πi ΓR z n+1
The power series (18.2) converges inside ΓR , so its radius of convergence ρ is at least
R; as R may be made arbitrarily large, ρ = ∞.
Further, when |z| = R,
f (z) C
n+1
≤ n+1 ,
z R
and so, by the M L Lemma,
Z
n! f (z) n! C n! C
f (n)
(0) = n+1
dz ≤ n+1
2πR = n .
2π ΓR z 2π R R
If n ≥ 1, then the left hand side of the formula above is 0, since R may be made
arbitrarily large, and so f (n) (0) = 0. Thus the power series (18.2) simplifies to show
that f (z) = f (0). □
4. Examples
Z 2π
4
Exercise 18.8. Compute dθ.
0 5 + 3 cos(θ)
Answer. First we “reverse engineer” this into an integral that arises from inte-
grating around a parametrised circle: take γ(θ) = eiθ , where 0 ≤ θ ≤ 2π.
Z 2π Z 2π
4 8
dθ = dθ
0 5 + 3 cos(θ) 0 10 + 3e + 3e−iθ
iθ
Z 2π
8eiθ
= dθ
0 3(eiθ )2 + 10eiθ + 3
Z
1 2π 8
= ieiθ dθ
i 0 3(e ) + 10eiθ + 3
iθ 2
Z
1 2π 8
= 2
γ ′ (θ) dθ
i 0 3(γ(θ)) + 10γ(θ) + 3
Z
1 8
= 2
dz.
i γ 3z + 10z + 3
Now we use partial fractions. Clearly 3z 2 + 10z + 3 = (3z + 1)(z + 3), and it is easy
to check that
8 3 1
2
= − .
3z + 10z + 3 3z + 1 z + 3
Thus
Z 2π Z
4 1 3 1
dθ = − dz
0 5 + 3 cos(θ) i γ 3z + 1 z + 3
Z Z
1 1 1 1
= dz − dz
i γ z + 1/3 i γ z+3
Z
1 1
= dz,
i γ z + 1/3
because z 7→ 1/(z +3) is holomorphic in B(0, 3) and Int(γ) ⊂ B(0, 3), so the second
integral is zero by the Cauchy–Goursat theorem. Finally, we apply Cauchy’s integral
formula, where f (z) is identically 1 and w = −1/3, which lies inside the curve γ. It
114 18. CAUCHY’S GENERALISED INTEGRAL FORMULA
follows that
Z 2π
4 2πi
dθ = f (−1/3)
0 5 + 3 cos(θ) i
= 2π. 4
This integral may also be computed using the substitution t = tan(θ/2).
LECTURE 19
1. Morera’s theorem
Theorem 19.1 (Morera’s theorem). Suppose that Ω is a domain, that the function
f : Ω → C is continuous, and that
Z
f (z) dz = 0,
Γ
whenever Γ is a closed contour in Ω. Then f is holomorphic in Ω.
in B(z0 , r). We may differentiate a power series term by term in any ball in which
it converges, so
X
∞ X
∞
f (z) = F ′ (z) = cn n(z − z0 )n−1 = cm+1 (m + 1)(z − z0 )m
n=0 m=0
break such an integral into a sum of integrals over closed contours in Ω \ Λ together
with an error term that may be made arbitrarily small. □
(a) f is holomorphic in Ω
Z
(b) f (z) dz = 0 for every closed Γ
Γ
Z
′
(b ) f (z) dz depends only on the start and end of Γ
Γ
Z
′′
(b ) there is a function F in Ω such that f (z) dz = F (q) − F (p), where Γ goes
Γ
from p to q, and F ′ = f
Z
1 f (z)
(c) f (w) = dz for any simple closed Γ such that w ∈ Int(Γ)
2πi Γ z − w
X
∞
(d) f (z) = cn (z − z0 )n in any open ball B(z0 , r) contained in Ω.
n=0
We have seen that:
(a) =⇒ (b) (the Cauchy–Goursat theorem)
′
(b) =⇒ (b ) (independence of contour)
(b′ ) =⇒ (b′′ ) (existence of primitives)
(a) =⇒ (c) (Cauchy’s integral formula)
(c) =⇒ (d) (corollary to Cauchy’s integral formula)
(d) =⇒ (a) (power series)
′
(b ) =⇒ (a) (Morera’s theorem)
′′
(b ) =⇒ (a) (proof of Morera’s theorem)
3. Analytic continuation
We have already seen that the values of a holomorphic function f on an interval
determine f in an open ball. The key to this is showing that if f is 0 on an interval,
then f is 0 on the ball. Now we see that the values of a holomorphic function in an
open subset of a domain determine the values in the whole domain.
Lemma 19.3. Suppose that B(z1 , r1 ) ⊂ B(w, R), that f ∈ H(B(w, R)), and that
f (z) = 0 for all z ∈ B(z1 , r1 ). Then f (z) = 0 for all z ∈ B(w, R).
and the centre of B(zj+1 , rj+1 ) is contained in B(zj , rj ). We will show that f (z) = 0
for all z ∈ B(zJ , rJ ) by induction.
By hypothesis, f (z) = 0 for all z ∈ B(z1 , r1 ).
Suppose that f (z) = 0 for all z ∈ B(zj , rj ). Then f (z) = 0 for all z near the
centre of B(zj+1 , rj+1 ). Hence f (n) (zj+1 ) = 0 for all n ∈ N. We may expand f in a
power series in B(zj+1 , rj+1 ), and the coefficients in the power series are multiples
of the derivatives f (n) (zj+1 ). It follows that f (z) = 0 for all z ∈ B(zj+1 , rj+1 ).
By induction, f (z) = 0 for all z ∈ B(zJ , rJ ). □
Answer. Then
Z 2π Z 2π
cos θ 2 cos θ
dθ = dθ
0 5 + 3 cos θ 0 10 + 6 cos θ
Z 2π
eiθ + e−iθ
= dθ
0 10 + 3(eiθ + e−iθ )
Z 2π
e2iθ + 1
= dθ
0 3e2iθ + 10eiθ + 3
Z
1 2π e2iθ + 1
= ieiθ dθ.
i 0 eiθ (3e2iθ + 10eiθ + 3)
Let γ(θ) = eiθ , where 0 ≤ θ ≤ 2π. Then
Z 2π Z
cos θ 1 z2 + 1
dθ = dz
0 5 + 3 cos θ i γ z(3z 2 + 10z + 3)
Z
1 1 5 5
= − + dz
i γ 3z 4(3z + 1) 12(z + 3)
Z Z Z
1 1 5 1 5 1
= dz − dz + dz
3i γ z 12i γ z + 1/3 12i γ z + 3
2πi 10πi
= − +0
3i 12i
−π
= ,
6
by the Cauchy integral formula for the first two integrals and the Cauchy–Goursat
theorem for the last integral. 4
Z 2π
sin θ
Exercise 19.7. Compute dθ.
0 cos θ
Answer. This is a trick question: the integral is not defined. 4
In all the following exercises, Γ is the circle of radius 3 and centre 0, traversed
in the usual anticlockwise direction, and f ∈ H(B(0, π)).
Z
f (z)
Exercise 19.8. Compute dz.
Γ z −1
2
Z
f (z)
Exercise 19.10. Compute dz.
Γ (z 2 − 1)2
Answer. We may use partial fractions again:
1 A B C D
= + + + ,
(z − 1)2 (z + 1)2 (z − 1)2 z − 1 (z + 1)2 z + 1
from which it follows that
A(z + 1)2 + B(z − 1)(z + 1)2 + C(z − 1)2 + D(z − 1)2 (z + 1) = 1.
Looking at the coefficients of z 3 tells us that B + D = 0. Putting z = 1 and then
putting z = −1 tells us first that 4A = 1 and then that 4C = 1. Putting z = 0 tells
us that A − B + C + D = 1. It follows that
1
A = −B = C = D = .
4
We conclude that
Z Z Z Z Z
f (z) 1h f (z) f (z) f (z) f (z) i
dz = dz − dz + dz + dz
Γ (z − 1) 4 Γ (z − 1)2 Γ z −1
2 2 2
Γ (z + 1) Γ z +1
2πi ′
= f (1) − f (1) + f ′ (−1) + f (−1)
4
πi ′
= f (1) − f (1) + f ′ (−1) + f (−1) ,
2
by Cauchy’s (generalised) integral formula. 4
5. Remarks
The exercises that we have done suggest that we will be able to compute “arbi-
trary” integrals around closed contours of functions of the form f (z)/p(z), where p
is a polynomial. To do this, we will need to be able to factorise p, and then expand
1/p into partial fractions, that is, a sum of terms of the form 1/(z − α)k . We will
need to be able to compute with partial fractions!
LECTURE 20
Taylor series
We recall some facts about power series, Cauchy’s integral formula, and Taylor
series. We discuss computation with and manipulation of Taylor series.
where the coefficients an , the centre z0 , and the variable z are complex. A Taylor
series (with centre z0 ) for a function f is a series of the form
X∞
f (n) (z0 )
(z − z0 )n .
n=0
n!
A Maclaurin series for a function f is a Taylor series for f with centre 0, that is, a
series of the form
X∞
f (n) (0) n
z .
n=0
n!
Maclaurin series and Taylor series are particular examples of power series.
Recall from Lecture 10 that a power series has a radius of convergence, ρ, which
can often be found using the ratio test or the root test. The power series converges
inside B(z0 , ρ) and fails to converge outside B(z0 , ρ); if a power series converges in
B(z0 , r), then r ≤ ρ, but it is possible that r < ρ.
We say that (or write)
X∞
f (z) = an (z − z0 )n in B(z0 , r)
n=0
to mean that the domain of f includes B(z0 , r), the power series converges in B(z0 , r)
and that f (z) is the sum of the power series for each z ∈ B(z0 , r). If this holds, then
f is holomorphic in B(z0 , r), and moreover
f (n) (z0 )
an = , (20.1)
n!
that is, this power series is the Taylor series for f with centre z0 . Conversely, from
the lecture on Cauchy’s generalised integral formula, if a function
P∞ f is holomorphic
in B(z0 , r), then it can be represented as a power series n=0 an (z − z0 )n in the
same ball.
When we ask how large a ball with centre z0 in which a function f is represented
by a power series can be, we often find that the maximum value of the radius r
121
122 20. TAYLOR SERIES
of the ball is equal to the distance of z0 from the set of points where f fails to be
holomorphic.
Exercise 20.2. Show that the function f defined by f (z) = 1/z can be rep-
resented as a power series in a ball B(z0 , r), where z0 6= 0. Find the radius of
convergence of this power series.
Answer.
1 1 1
f (z) = = =
z z0 + z − z0 z0 (1 + (z − z0 )/z0 )
1 X
∞
(z − z0 )n
= (−1)n
z0 n=0 z0n
X
∞
(−1)n
= (z − z0 )n .
n=0
z0n+1
By the ratio test, the radius of convergence of this series is |z0 |. 4
Then
X
∞
(a) c f (z) = c an (z − z0 )n in B(z0 , r);
n=0
X
∞
(b) (f + g)(z) = (an + bn )(z − z0 )n in B(z0 , r);
n=0
X
∞
Pn
(c) (f g)(z) = cn (z − z0 )n in B(z0 , r), where cn = j=0 aj bn−j .
n=0
and
n
X
(n) n
(f g) (z0 ) = f (k) (z0 ) g (n−k) (z0 ),
k=0
k
and the theorem follows from this and (20.1). □
Further, let
(
c when m = 0
bm = am−1
when m ≥ 1.
m
X
∞
Then bn (z − z0 )n converges in B(z0 , r), to a function F say, such that F ′ = f
n=0
and F (z0 ) = c.
3. Examples
Exercise 20.5. Consider the function Log. Determine its Taylor series with
centre i − 1. What is the radius of convergence ρ of this series. Does the series
represent Log in the ball B(i − 1, ρ)?
ln 2 3 X∞
(−1)n−1
= + iπ + (z + 1 − i)n .
2 4 n=1
n(i − 1) n
√
in the ball B(i − 1, 2).
We look for a function F such that F ′ (z) = 1/z in this same ball. According to
the previous theorem, if
X
∞
(z + 1 − i)n+1 X∞
m−1 (z + 1 − i)
m
n
F (z) = (−1) = (−1) ,
n=0
(n + 1)(i − 1) n+1
m=1
m(i − 1) m
√
then F ′ (z) = 1/z. In the connected set {z ∈ C : |z + 1 − i| < 2, Im(z) > 0}, it is
also true that Log′ (z) = 1/z, and hence (Log −F )′ (z) = 0 and Log(z) − F (z) is a
constant. By considering this expression at i − 1, we see that
√ 3
Log(z) − F (z) = Log(i − 1) − 0 = ln( 2) + iπ,
4
and hence
X∞
(z + 1 − i)m ln 2 3 X∞
Log(z) = (−1)m+1 + + iπ = am (z + 1 − i)m ,
m=1
m(i − 1) m 2 4 m=0
ln 2 3 (z + 1 − i)m
where a0 = + iπ and am = (−1)m+1 when m ≥ 1. From the ratio
2 4 m(i − 1)m √
test, we see that the radius of convergence of the series is 2. The Taylor series
represents Log above the real axis, but does not represent Log on the other side of
the branch cut along the negative real axis. However, it does represent the branch
of log where the imaginary parts of the values lie in [0, 2π). 4
sin(z)
Exercise 20.6. Define the function f by f (z) = if z 6= 0 and f (0) = 1.
z
Does this function have a Maclaurin series? If so, then what is its radius of conver-
gence?
3. EXAMPLES 125
z3 z5 X (−1)n ∞
sin(z) = z − + − ··· = z 2n+1 .
3! 5! n=0
(2n + 1)!
sin(z) z2 z4 X (−1)n ∞
=1− + − ··· = z 2n .
z 3! 5! n=0
(2n + 1)!
When z = 0, the power series on the right hand side of this last expression is equal
to 1. It follows that
z2 z4 X (−1)n
∞
f (z) = 1 − + − ··· = z 2n ∀z ∈ C.
3! 5! n=0
(2n + 1)!
This power series converges for all z in C. Thus the radius of convergence of the
Maclaurin series for f (z) is infinite. 4
We note that it follows that f (2n) (0) = (−1)n /(2n + 1) and f (2n+1) (0) = 0 for
all n ∈ N. It would also be possible to compute the derivatives of the function
z 7→ sin(z)/z “by hand”, and to find a formula for these using induction, but this
would be much longer.
When we discuss singularities in more detail, we will say that “the function
z 7→ sin(z)/z has a removable singularity at 0”.
z
Exercise 20.7. Define the function g : C → C by g(z) = when z 6= 0
sin(z)
and g(0) = 1. Does this function have a Maclaurin series? If so, then what is its
radius of convergence?
Laurent series
1. Laurent’s theorem
Theorem 21.1. Suppose that A is the annulus {z ∈ C : R1 < |z − z0 | < R2 } and
R1 < r < R2 . If f ∈ H(A), then
X
∞
f (w) = cn (w − z0 )n ∀w ∈ A,
n=−∞
where Z
1 f (z)
cn = dz.
2πi ∂B(z0 ,r) (z − z0 )n+1
Proof. First, take r1 and r2 such that R1 < r1 < r2 < R2 , and let Ω1 be the
domain B(z0 , r2 ) \ B(z0 , r1 ), as shown in Figure 21.1.
R
Suppose that g ∈ H(A). By the Cauchy–Goursat theorem, ∂Ω1 g(z) dz = 0. It
follows that Z Z
g(z) dz = g(z) dz.
∂B(z0 ,r1 ) ∂B(z0 ,r2 )
R2 r2
R1 b z0 r1
A Ω1
b
w
Ω2
X∞ Z
f (z)
= (w − z0 ) n
dz.
∂B(z0 ,r2 ) (z − z0 )
n+1
n=0
2. FINDING LAURENT SERIES 129
X∞ Z
1
= f (z) (z − z0 )m dz
m=0
(w − z 0 ) m+1
∂B(z0 ,r1 )
X−1 Z
f (z)
= (w − z0 )n dz
∂B(z0 ,r1 ) (z − z0 )
n+1
n=−∞
X
∞
f (w) = cn (w − z0 )n ∀w ∈ A,
n=−∞
where
Z
1 f (z)
if n ≥ 0
dz
2πi Z∂B(z0 ,r2 ) (z − z0 )n+1
cn =
1 f (z)
dz if n ≤ −1.
2πi ∂B(z0 ,r1 ) (z − z0 )n+1
To conclude, we recall that the integrals defining cn around ∂B(z0 , r) do not depend
on r in the range (R1 , R2 ). □
1 1 −1 1/2 1/2
f (z) = = = + + .
z(z 2 − 1) z(z − 1)(z + 1) z z−1 z+1
Find Laurent series for f in the largest annuli with centre 0 in which f is holomor-
phic.
Answer. The function f has singularities at 0, ±1. The largest annuli with
centre 0 in which f is holomorphic are {z ∈ C : 0 < |z| < 1} and {z ∈ C : 1 < |z|}.
130 21. LAURENT SERIES
In the annulus {z ∈ C : 0 < |z| < 1}, also known as the punctured ball B ◦ (0, 1),
we may write
−1 1/2 1/2
f (z) = − +
z 1−z 1+z
−1 1 1
= − 1 + z + z2 + z3 + z4 + . . . + 1 − z + z2 − z3 + z4 − . . .
z 2 2
−1
= − z + z3 + z5 + . . .
z
= (−1) z −1 + z + z 3 + z 5 + . . .
Answer. The function f has singularities at 0, ±1. The three largest annuli
with centre 1 in which f is holomorphic are {z ∈ C : 0 < |z − 1| < 1}, {z ∈ C : 1 <
|z − 1| < 2}, and {z ∈ C : 1 < |z − 1|}.
2. FINDING LAURENT SERIES 131
X∞
(−1)m−1 (2m−2 − 1)
= . 4
m=3
(z − 1)m
Remarks. These series involve both odd and even powers, reflecting the fact
that f is neither even nor odd about 1.
None of these series converges in any larger annulus, since otherwise f would be
holomorphic in a larger domain.
Note also that the last Laurent series starts with z −3 . This corresponds to
the fact that f (z) behaves like z −3 at infinity, and that (z − 1)−3 and z −3 behave
similarly there. Thus we should have expected cancellations in the terms (z − 1)−1
and (z − 1)−2 .
LECTURE 22
Singularities
by setting f (0) = 0, and then f becomes an entire function, and the singularity has
been removed. This extended function has a zero of order three at 0.
1
3. Suppose that f (z) = 3 . The natural domain for f is C \ {0, ±1}, because
z −z
f (0), f (1) and f (−1) are not defined. As we will see, limz→0 f (z) = ∞, so we will
leave f (0) undefined. For this function,
X
∞
f (z) = (−1)n+1 z 2n−1
n=0
◦
in B (0, 1), and there is a pole of order 1, that is, a simple pole, at 0.
4. Suppose that f (z) = e1/z . The natural domain for f is C \ {0}, because f (0)
is not defined. As we will see, limz→0 f (z) does not exist, and f has an essential
singularity at 0.
5. Suppose that f (z) = Log(z). The domain of f is C \ {0}, and we say that f has
a singularity at 0. However, f is not differentiable at all points on the negative real
axis, due to the jump in Arg there, and so the singularity at 0 is not isolated.
1
6. Suppose that f (z) = . Then f is defined in C\({0}∪{1/n : n ∈ Z\{0}}).
sin(π/z)
In this example, f has isolated singularities at the points 1/n, where n ∈ Z \ {0},
and a nonisolated singularity at 0.
Theorem 22.3. Suppose that f ∈ H(B ◦ (z0 , R)), and that
X
∞
f (z) = cn (z − z0 )n ∀z ∈ B ◦ (z0 , R).
n=−∞
Suppose that (iv) holds. We use an argument like that to prove Liouville’s
theorem. According to Laurent’s theorem, if 0 < r < R, then
Z
1 f (z)
cn = dz .
2πi ∂B(z0 ,r) (z − z0 )n+1
Clearly when z lies on ∂B(z0 , r),
f (z) |f (z)| C
= n+1 ≤ n+1 .
(z − z0 ) n+1 r r
The length of ∂B(z0 , r) is 2πr, and so, by the M L Lemma,
Z
1 f (z) 1 C
|cn | ≤ dz ≤ 2πr = 2πCr−n .
2π ∂B(z0 ,r) (z − z0 ) n+1 2π rn+1
If n < 0, then −n > 0, and r−n → 0 as r → 0. In this case, we can make r−n
arbitrarily small, so cn = 0, i.e., (i) holds. □
The reason for the terminology “removable singularity” should now be clear:
by defining, or perhaps redefining, f (z0 ) appropriately, we may extend f to be
holomorphic in B(z0 , R), that is, we remove the singularity of f at z0 .
It is possible to generalize the proof of the theorem above to prove the following
result.
Theorem 22.4. Suppose that f ∈ H(B ◦ (z0 , R)), and that
X
∞
f (z) = cn (z − z0 )n ∀z ∈ B ◦ (z0 , R).
n=−∞
We omit the details of the proof, but observe that it involves defining the function
F by F (z) = (z − z0 )−M f (z), applying the techniques of the proof of the previous
theorem to F , and being careful about when limits are 0.
We write f (z) ∼ c(z − z0 )N as z → z0 , where c ∈ C \ {0} and N ∈ Z, to mean
that
f (z)
lim = 1.
z→z0 c(z − z0 )N
2. Examples
Exercise 22.7. Suppose that f (z) = (1 − cos(z))2 /z. What kind of singularity
does f have at 0?
Answer. Clearly
(1 − cos(z))2 2(1 − cos(z)) sin(z)
lim f (z) = lim = lim =0
z→0 z→0 z z→0 1
by l’Hôpital’s rule. So f has a zero of some currently unknown order at 0.
3. MORE ABOUT SINGULARITIES† (NOT EXAMINABLE) 137
Exercise 22.8. Suppose that f (z) = e1/z . How does f behave near 0?
Answer. From Taylor series,
w2 w3
ew = 1 + w + + + ....
2! 3!
Hence
z −2 z −3
e1/z = 1 + z −1 + + + ....
2! 3!
Since there are infinitely many nonzero negative powers of z, the singularity at 0 is
essential. 4
1. Residues
Definition 23.1. Suppose that the function f has an isolated singularity at z0 .
Then
P∞ f is holomorphic in some punctured ball B ◦ (z0 , r), and has a Laurent series
n=−∞ cn (z − z0 ) there. The residue of f at z0 , written Res(f, z0 ) or Res(f (z), z =
n
We prove this theorem later. This theorem enables us to calculate integrals over
closed curves as long as we can calculate residues.
Exercise 23.3. Suppose that
α1 β1 β2
f (z) = + + .
z − a z − b (z − b)2
R
Find the residues of f at a and b, and hence find Γ f (z) dz, where Γ is a simple
closed contour surrounding a and b.
Answer. Both the singularities a and b lie inside Γ, and both may contribute
to the sum of residues.
To find Res(f, a), we observe that the second and third terms are holomorphic
inside a very small ball B(a, ε) centred at a, so only contribute terms in (z − a)n
where n ≥ 0 to the Laurent series for f in B ◦ (a, ε), and the residue is α1 .
139
140 23. RESIDUES AND THE RESIDUE THEOREM
To find Res(f, b), we observe that the first term is holomorphic inside a very
small ball B(b, ε) centred at b, so only contributes terms in (z − b)n where n ≥ 0 to
the Laurent series for f in B ◦ (b, ε), and the residue is β1 .
Consequently,
Z
f (z) dz = 2πi Res(f, a) + Res(f, b) = 2πi α1 + β1 . 4
Γ
This example shows that, for rational functions, residues are coefficients of cer-
tain terms in the partial fraction expansions, and that not all the coefficients are
needed to compute contour integrals (in the example above, β2 does not matter at
all). If we can compute residues efficiently, then we will not need to find partial
fraction expansions.
2. Computing residues
Suppose that f has an isolated singularity at z0 . There are three ways to compute
Res(f, z0 ), according to the type of singularity,
First, if the singularity is removable, that is, if f is bounded near z0 , the residue
is 0.
Second, if f has a pole of order N at z0 , then
f (z) = c−N (z − z0 )−N + c1−N (z − z0 )1−N + . . .
+ c−1 (z − z0 )−1 + c0 (z − z0 )0 + c1 (z − z0 )1 + . . .
so
(z − z0 )N f (z) = c−N (z − z0 )0 + c1−N (z − z0 )1 + . . .
+ c−1 (z − z0 )N −1 + c0 (z − z0 )N + c1 (z − z0 )N +1 + . . .
and
dN −1
(z − z0 ) f (z)
N
dz N −1
dN −1
= N −1 c−N (z − z0 )0 + c1−N (z − z0 )1 + . . .
dz
+ c−1 (z − z0 )N −1 + c0 (z − z0 )N + c1 (z − z0 )N +1 + . . .
(N + 1)!
= (N − 1)! c−1 (z − z0 )0 + N ! c0 (z − z0 )1 + c1 (z − z0 )2 + . . . ,
2!
whence
dN −1
lim (z − z0 )N f (z) = (N − 1)! c−1 .
z→z0 dz N −1
Thus
1 dN −1
Res(f, z0 ) = lim N −1 (z − z0 )N f (z). (23.2)
(N − 1)! z→z0 dz
Note that this formula is not the definition of a residue; that definition is at the
start of this lecture. To use this formula, we have to know the order of the pole,
and it does not apply at all singularities.
Third, if f has an essential singularity at z0 then we need a different approach,
usually involving integration or series.
2. COMPUTING RESIDUES 141
for increasing values of n, starting at 0, until such time as we find a finite number;
this is then the order of the pole, N . Once we know N , we use formula (23.2). This
would be very inefficient if the order of the pole is large, and would not give a result
at all for an essential singularity! It is therefore important to know how to find the
order of the pole. The key information about the orders of zeros and poles is in the
previous lecture.
z − π/2
Exercise 23.4. Suppose that f (z) = . Find Res(f, π/2).
1 − sin(z)
We could have worked out that the pole was simple by working out that the zero
of the numerator is of order 1 and that of the denominator is of order 2, and then
doing the second calculation only.
The following result is often useful.
Proposition 23.5 (The p/q ′ formula). Suppose that f (z) = p(z)/q(z) in Ω, and
that p(z0 ) 6= 0 while q(z0 ) = 0. If z0 is a simple zero of q, that is, a zero of order 1,
then
p(z0 )
Res(f, z0 ) = ′ .
q (z0 )
Of course, we could also have done this calculation directly, without using the
formula.
Exercise 23.7. Suppose that
f (z) = 2z sin(z −2 ).
Find the residue of f at 0.
Answer. By Taylor series,
w3 w5
sin(w) = w − + − ...,
3! 5!
and so, replacing w by z −2 , we see that
z −6 z −10
2z sin(z −2 ) = 2z z −2 − + − ...
3! 5!
z −5
z −9
−1
=2 z − + − ... .
3! 5!
Hence Res(f, 0) = 2. 4
The expression for the residue in the last example does not simplify but may be
computed numerically to any desired degree of accuracy. This behaviour is common
with essential singularities.
Essential singularities usually appear because there is a “transcendental func-
tion” (such as exp or sinh or tan) which can be written as a power series, with
something like 1/z in its argument. Residues involving transcendental functions
3. PROOF OF CAUCHY’S RESIDUE THEOREM 143
Im
× z3 × z2 Γ
× z1
× z4 × z5
Re
can often be expressed as series which are derived from the corresponding power
series.
Proof. We take balls B(zk , ε) centred at the singularities zk , where the ε are
chosen small enough that the closed balls B̄(zi , ε) and B̄(zj , ε) are disjoint if i 6= j,
and such that each closed ball B̄(zk , ε) is contained in Int(Γ). This ensures that
the contours around the perimeters of the balls B(zk , ε) are well defined and do not
meet each other or Γ. Define
[ K
Υ = Int(Γ) \ B(zk , ε) .
k=1
See Figure 23.1. Then f is holomorphic on the open set Ω which contains Ῡ, and ∂Υ
is made up of Γ, traversed anti-clockwise, together with the boundaries ∂B(zk , ε)
traversed clockwise.
The Cauchy–Goursat theorem implies that
Z Z XK Z
0= f (z) dz = f (z) dz + f (z) dz,
∂Υ Γ k=1 ∂B(zk ,ε)∗
that is,
Z K Z
X
f (z) dz = f (z) dz.
Γ k=1 ∂B(zk ,ε)
144 23. RESIDUES AND THE RESIDUE THEOREM
From (23.1), Z
1
Res(f, zk ) = f (z) dz,
2πi ∂B(zk ,ε)
so
Z
f (z) dz = 2πi Res(f, zk ),
∂B(zk ,ε)
and hence
Z K Z
X X
K
f (z) dz = f (z) dz = 2πi Res(f, zk );
Γ k=1 ∂B(zk ,ε) k=1
the theorem is proved. □
LECTURE 24
Computing integrals. 1
In this lecture, we give three examples of the use of Cauchy’s residue theorem
to calculate integrals.
1. Trigonometric integrals
Z π
cos(θ)
Exercise 24.1. Evaluate dθ.
−π 5 − 4 cos(θ)
Answer. The first step is to convert this to a contour integral. We take γ(θ) =
eiθ , where −π ≤ θ ≤ π. Then γ ′ (θ) = ieiθ = iγ(θ) and
eiθ + e−iθ γ(θ) + γ(θ)−1
cos(θ) = = .
2 2
As θ varies over its domain, γ(θ) travels anticlockwise around the unit circle in the
complex plane. Thus
Z π Z
cos(θ) (z + z −1 )/2 1
dθ = dz
−π 5 − 4 cos(θ) γ 5 − 2(z + z ) iz
−1
Z
1 z2 + 1 1
= dz
2i γ (−2z + 5z − 2) z
2
Z
i z2 + 1
= dz
2 γ z(2z 2 − 5z + 2)
Z
i z2 + 1
= dz
2 γ z(2z − 1)(z − 2)
Z
i
= f (z) dz,
2 γ
where
z2 + 1
f (z) = dz.
z(2z − 1)(z − 2)
The second step is to evaluate the contour integral by evaluating the residues of f .
Clearly, f has singularities at 0, 21 and 2. Only 0 and 12 lie inside γ. Because each
of the factors in the denominator is of degree 1, and the numerator does not vanish
at these points, each of the singularities is a simple pole. Further
z2 + 1 1
Res(f, 0) = lim (z − 0)f (z) = lim zf (z) = lim =
z→0 z→0 z→0 (2z − 1)(z − 2) 2
and
z2 + 1 5/4 5
Res(f, 1/2) = lim (z − 1/2)f (z) = lim = =− .
z→1/2 z→1/2 2z(z − 2) −3/2 6
145
146 24. COMPUTING INTEGRALS. 1
Remarks on this calculation. If the integrand had contained a sin term, then
we could have used the fact that
eiθ − e−iθ z − z −1
sin(θ) = = .
2i 2i
Similarly, expressions such as sin(2θ) and tan(θ) may be expressed in terms of eiθ
and hence of z. In summary, any integral that can be put in the form
Z π
f (sin(θ), cos(θ)) dθ
−π
where f is a rational function of two variables, can be tackled in this way, and
becomes an integral of a different rational function around a closed contour, which
can be evaluated, at least in principle, and often in practice.
There are some other integrals that may be converted to integrals of this form.
For instance, since cos is an even function, so is θ 7→ cos(θ)/(5 − 4 cos(θ)), and
Z π Z 0
cos(θ) cos(θ)
dθ = dθ
0 5 − 4 cos(θ) −π 5 − 4 cos(θ)
and so Z Z
π
cos(θ) 1 π cos(θ)
dθ = dθ.
0 5 − 4 cos(θ) 2 −π 5 − 4 cos(θ)
If the integrand involves expressions like cos(θ/2), this method does not work: square
roots appear, and these mess up the holomorphy.
Answer. The real line is not a simple closed contour, and it is not obvious how
the residue theorem can be used.
Define f by
1 1
f (z) = 4 = ,
z +1 (z − ω1 )(z − ω2 )(z − ω3 )(z − ω4 )
where ωk = exp( 12 πik − 14 πi).
R The function f has four singularities,
R at the points
ωk . We will integrate f , as f (z) dz should be related to f (x) dx.
We will take Ω to be the semicircular region Ω above the interval [−R, R], and
integrate f around the boundary ∂Ω. We suppose that R > 1, so that all the
singularities of the integrand in the upper half plane lie in Ω.
2. A RATIONAL FUNCTION ON THE REAL LINE 147
Im
Ω
ω2 × × ω1 Re
−R ω3 × × ω4 R
We defined f by
1 1
f (z) = = ,
z4 + 1 (z − ω1 )(z − ω2 )(z − ω3 )(z − ω4 )
where ωk = exp( 12 πik − 14 πi). The only singularities of f that lie in Ω are ω1 and ω2 .
Since the power of (z − ωk ) in the denominator is 1, these are simple poles. Further,
ωk4 = −1, and so ωk3 = −1/ωk . We deduce that
z − ωk 1 1 1
Res(f, ωk ) = lim (z − ωk )f (z) = lim 4
= lim 3
= 3
= − ωk .
z→ωk z→ωk z +1 z→ωk 4z 4ωk 4
Now 2πi multiplied by the sum of the residues is given by
2πi 1 + i −1 + i
2πi(Res(f, ω1 ) + Res(f, ω2 )) = − √ + √
4 2 2
√
πi 2i π 2
=− √ = .
2 2 2
The boundary contour ∂Ω has two parts: the circular arc and the diameter.
Parametrise the diameter by the function λ : [−R, R] → C, defined by λ(x) = x.
Then one component of the integral around the contour is
Z Z R Z R
′
f (z) dz = f (λ(x)) λ (x) dx = f (x) dx
λ −R −R
Z R Z ∞ (24.1)
1 1
= 4
dx → 4
dx
−R x + 1 −∞ x + 1
Im
iR
Γ2
Γ1
Γ3
ir Re
× ×
−ir 1
Γ4
−iR
as R → ∞, that is,
Z
f (z) dz → 0. (24.2)
γ
where Γ is the join of the contours Γ1 , Γ2 , Γ3 and Γ4 shown in Figure 24.2, and z 1/2
denotes the principal branch of the square root. What happens to the component
integrals when r → 0 and R → ∞? Find
Z +∞ 1/2
x
dx.
0 x2 + 1
Answer. We suppose that r < 1 < R. First of all, this is a simple closed
contour, and the integrand is holomorphic on and inside the contour, except for a
3. AN INTEGRAL INVOLVING A ROOT 149
simple pole at 1, so
Z z 1/2
z 1/2
dz = 2πi Res , z = 1
Γ z2 − 1 z2 − 1
z 1/2 (z − 1)
= 2πi lim
z→1 z2 − 1
= πi.
Now, by the M L lemma,
Z n z 1/2 o
z 1/2
dz ≤ max : |z| = r, Re(z) ≥ 0 πr
Γ3 z − 1 z2 − 1
2
r1/2
≤ πr → 0 as r → 0.
1 − r2
Similarly, by the M L lemma,
Z n z 1/2 o
z 1/2
dz ≤ max : |z| = R, Re(z) ≥ 0 πR
Γ1 z − 1 z2 − 1
2
R1/2
≤ πR → 0 as R → ∞.
R2 − 1
Parametrise Γ2 by γ2 (y) = iy, where y goes from R to r, and Γ4 by γ4 (y) = −iy,
where r ≤ y ≤ R. Then,
Z Z R Z R 1/2
z 1/2 (iy)1/2 y
dz = − i dy = i(+i) 1/2
dy,
Γ2 z − 1 r −y − 1
2 2 2
r y +1
and Z Z Z
R R
z 1/2 (−iy)1/2 y 1/2
dz = (−i) dy = i(−i)1/2 dy.
Γ4 z2 − 1 r −y 2 − 1 r y2 + 1
As r → 0 and R → ∞,
Z R Z ∞
y 1/2 y 1/2
dy → dy.
r y2 + 1 0 y2 + 1
Now
Z Z Z Z
z 1/2 z 1/2 z 1/2 z 1/2
dz + dz + dz + dz
Γ1 z2 − 1 Γ2 z2 − 1 Γ3 z − 1
2
Γ4 z − 1
2
Z
z 1/2
= dz = πi ,
Γ z −1
2
whence
Z z 1/2
Z
z 1/2
Z
z 1/2
Z
z 1/2
lim dz + dz + dz + dz
r→0
R→∞ Γ1 z2 − 1 Γ2 z2 − 1 Γ3 z2 − 1 Γ4 z2 − 1
= πi,
so
Z ∞
y 1/2
1/2 1/2
i(i ) + i(−i) dy = πi,
0 y2 + 1
150 24. COMPUTING INTEGRALS. 1
and hence Z ∞
y 1/2 π
2
dy = √ . 4
0 y +1 2
We can compute integrals involving functions with branches as long as the branch
cut is outside the contour. The process of avoiding a singularity, such as 0 here, by
adding a small circular arc to the contour, is called indenting.
LECTURE 25
Computing Integrals. 2
Answer. Define f by
eiξz eiξz
f (z) = = .
z2 + 1 (z − i)(z + i)
The function f has two singularities, at the points ±i. We integrate around the
boundary of the semicircular region Ω above the interval [−R, R] shown in Figure
25.1. We suppose that R > 1, so that all the singularities of the integrand in the
upper half plane lie inside the boundary contour ∂Ω.
Since the power of (z − i) in the denominator of the fraction defining f is 1, the
pole at i is simple. From l’Hôpital’s rule,
z−i 1 1
Res(f, i) = lim(z − i)f (z) = lim eiξz 2
= e−ξ lim = e−ξ .
z→i z→i z +1 z→i 2z 2i
The contour ∂Ω has two components: the circular arc and the diameter.
Im
Ω
×i
Re
−R × −i R
eiξz 1 1
≤ 2 ≤ 2 = M,
|z + 1|
2 |z + 1| R −1
say.
From the M L lemma,
Z
1 πR
f (z) dz ≤ M length(γ) = πR = 2 →0
γ R2 −1 R −1
as R → ∞, that is,
Z
f (z) dz → 0. (25.2)
γ
that is,
Z Z
−ξ
f (z) dz = πe − f (z) dz. (25.3)
λ γ
Remarks on this calculation. There are many integrals on the whole line R
that may be evaluated in this way. Integrals between 0 and ∞ may also be tackled
if the integrand is even. When the integrand involves cos(ξx) or sin(ξx), then we
write the trigonometric function in terms of exponentials, and if necessary change
variables to get an integral involving eiξx , where ξ ≥ 0. Then |eiξz | ≤ 1 when
z ∈ Range(γ). Sometimes this inequality is not enough and we have to use Jordan’s
lemma (see below).
For example, by the changes of variable x′ = −x and then x = x′ ,
Z ∞ Z Z ∞ −iξx
cos(ξx) 1 ∞ eiξx e
2
dx = dx + dx
−∞ x + 1 2 −∞ x2 + 1 2
−∞ x + 1
Z Z ∞
1 ∞ eiξx
′
eiξx ′
= dx + ′ 2
dx
2 −∞ x2 + 1 −∞ (x ) + 1
Z ∞ iξx
e
= 2
dx,
−∞ x + 1
which we have just computed. Since cos(x) and x2 are even functions of x,
Z ∞ Z
cos(ξx) 1 ∞ cos(ξx) πe−ξ
dx = dx = .
0 x2 + 1 2 −∞ x2 + 1 2
2. Jordan’s lemma
Lemma 25.2. Suppose that f is continuous in {z ∈ C : Im(z) ≥ 0, |z| ≥ S}, for
some positive S, that ΓR is the upper half of the circle with centre 0 and radius R,
and that |f (z)| ≤ MR for all z ∈ ΓR where limR→∞ MR = 0.
Then for any ξ > 0, Z
lim eiξz f (z) dz = 0.
R→∞ ΓR
|eiξz | ≤ 1,
and further, by the triangle inequality, |z 2 + 1| ≥ |z|2 − 1 = R2 − 1; thus
z eiξz |z| R
≤ ≤ .
|z 2 + 1| |z 2 + 1| R2 − 1
From the M L lemma,
Z
R πR2
f (z) dz ≤ length(γ) = 2 →π
γ R2 − 1 R −1
as R → ∞. This is not helpful. However, since M → 0 as R → ∞, Jordan’s lemma
still tells us that Z
f (z) dz → 0. (25.5)
γ
The result implies that the first integral is 0, as it should be since the integrand is
odd, while the second integral is πe−ξ .
4. PROOF OF JORDAN’S LEMMA† 155
y
y = 2θ/π
1 b
y = sin(θ)
θ
π/2 π
In this lecture, we count the number of times a closed curve winds around a
point, and use this to count the number of zeros and poles of a function. This leads
to some surprising facts about holomorphic functions.
We start with several definitions that could have come earlier.
Definition 26.1. A point z of a subset S of a set Ω is isolated if there exists
ε ∈ R+ such that B(z, ε) ∩ S = {z}. We say that S is discrete if every point of S is
isolated.
Definition 26.2. A function is said to be meromorphic in the open set Ω if it
is holomorphic in Ω \ ∆, where ∆ is a discrete subset of Ω, and the singularity at
each point of ∆ is a pole.
157
158 26. THE THEORY OF FUNCTIONS
Proof. From Cauchy’s residue theorem, the left hand side of the equality above
is equal to the sum of the residues of f ′ /f at its singularities inside γ. The function
f ′ /f has a singularity at w if and only if w is a zero or a pole of f , so it suffices to
compute the residues at these points.
Suppose that f has a zero or a pole at w. Then, from the lecture on singularities,
we know that there exist ε ∈ R+ and m ∈ Z such that f (z) = (z − w)m g(z), where
g, 1/g ∈ H(B(w, ε)). If w is a zero, then the order of the zero is m, while if w is a
pole, then the order of the pole is −m. Hence
f ′ (z) m(z − w)m−1 g(z) + (z − w)m g ′ (z) m g(z) + (z − w) g ′ (z)
= = .
f (z) (z − w)m g(z) (z − w) g(z)
Then f ′ /f has a simple pole at w, and
m g(z) + (z − w) g ′ (z)
Res(f ′ /f, w) = lim (z − w)
z→w (z − w) g(z)
m g(z) + (z − w) g ′ (z)
= lim
z→w g(z)
= m. □
This result has been used to provide experimental confirmation of the Riemann
hypothesis, which states that the zeros of the Riemann zeta function lie on the line
{z ∈ C : Re(z) = 21 }. We will discuss this later.
Theorem 26.4 may also be used to show that functions have zeros in certain
regions.
Corollary 26.5. Suppose that Ω is a domain, that f ∈ H(Ω), that γ : [a, b] → Ω
is a simple closed curve, and that f (γ(t)) 6= 0 for all t ∈ [a, b]. Then the number of
zeros of f in Int(γ) (counting multiplicities) is equal to the number of times f ◦ γ
winds around 0.
Proof. Let δ = f ◦ γ. Then δ : [a, b] → C is also a closed curve, though not
necessarily simple. Further, δ ′ (t) = f ′ (γ(t)) γ ′ (t), and so
Z ′ Z b ′ Z b ′ Z
f (z) f (γ(t)) ′ δ (t) 1
dz = γ (t) dt = dt = dz. □
γ f (z) a f (γ(t)) a δ(t) δ z
2. Rouché’s theorem
The next result gives us a useful tool for counting zeros.
Theorem 26.6 (Rouché’s theorem). Suppose that γ : [a, b] → Ω is a closed curve
in a simply connected domain Ω, that f, g ∈ H(Ω), and that |f (z)| < |g(z)| for all
z ∈ Range(γ). Then the number of zeros of f + g inside γ is equal to the number of
zeros of g inside γ.
We need to show that the winding number of (f +g)◦γ is the same as the winding
number of g ◦ γ. We choose the argument arg(g(γ(t)) so that it varies continuously
with t. Since |f (γ(t))| < |g(γ(t))|, we may choose arg(f (γ(t)) + g(γ(t))) in such a
way that
π
|arg((f + g)(γ(t))) − arg(g(γ(t)))| < ;
2
when we do this, arg((f + g)(γ(t))) varies continuously with t, and
|arg((f + g)(γ(b))) − arg((f + g)(γ(a))) − arg(g(γ(b))) + arg(g(γ(a)))| < π .
Since the left-hand side is equal to 2πk for some integer k, it must be 0, and so
arg((f + g)(γ(b))) − arg((f + g)(γ(a))) = arg(g(γ(b))) − arg(g(γ(a))),
which means that the winding numbers are equal. □
Here is a corollary.
P P
Corollary 26.7. Suppose that p(z) = nj=0 cj z j , and that |cm | > j̸=m |cj |.
Then p has m zeros (counting multiplicities) inside the unit circle.
P
Proof. Take γ(t) = eit , where 0 ≤ t ≤ 2π, and define f (z) = j̸=m cj z j and
g(z) = cm z m . The hypothesis on the coefficients of p means that |f (eit )| < |g(eit )|,
and by Rouché’s theorem, (f + g) has the same number of zeros inside γ as g, that
is, m zeros. □
Exercise 26.8. Suppose that n ≥ 3, and let p(z) = 2z n + z 2 − 3z − 1. Using
Rouché’s theorem, show that all the zeros of p lie inside the circle of radius 2.
Answer. Take g(z) = 2z n , and f (z) = z 2 − 3z − 1; then f + g = p. If |z| = 2,
then |g(z)| = 2n+1 ≥ 16, while |f (z)| ≤ 4 + 6 + 1, by the triangle inequality. By
Rouché’s theorem, the number of zeros of p inside the circle of radius 2 is equal to
the number of zeros of g inside the circle of radius 2, that is, n.
Since p has n zeros, all the zeros of p are inside the circle of radius 2. 4
Let M be the smallest positive integer for which am 6= 0. Then there exist δ, ε ∈ R+
such that for all w ∈ B ◦ (a0 , ε), there are exactly M points z in B ◦ (z0 , δ) such that
f (z) = w (counting multiplicities).
The proof of this result may be found at the end of this lecture.
P
Remark 26.10. Suppose that f (z) = ∞ n=M an (z − z0 ) in B(z0 , r), where aM 6=
n
Another way in which this assertion is true is that both f and the leading term of
the Taylor series are both M -to-1 near z0 .
In the case where a1 6= 0, Theorem 26.9 implies that f is one-to-one near z0 .
This is the hard part of the inverse function theorem.
Corollary 26.11 (Inverse function theorem). Suppose that f is a holomorphic
function in Ω, that f (z0 ) = w0 , and that f ′ (z0 ) 6= 0. Then there is an open subset
Υ of Ω, containing z0 , and an open ball B containing w0 such that the restriction of
f to Υ is a bijection from Υ onto B, and the inverse function F −1 from B to Υ is
holomorphic. Further,
1
(f −1 )′ (w0 ) = ′ .
f (z0 )
Proof. By Theorem 26.9 above, there exist δ, ε ∈ R+ such that for every point
w ∈ B(w0 , ε), there exists exactly one point z in B(z0 , δ) such that f (z) = w. We
define f −1 (w) = z.
It may be shown (but we do not do so here) that a continuous bijection on
a compact set has a continuous inverse. From this result, it follows that f −1 is
continuous, and so f −1 (w) → f −1 (w0 ) as w → w0 .
Now, taking w = f (z) and w0 = f (z0 ), we see that
−1
f −1 (w) − f −1 (w0 ) w − w0
lim = −1 lim−1
w→w0 w − w0 f (w)→f (w0 ) f −1 (w) − f −1 (w0 )
−1
f (z) − f (z0 )
= lim = f ′ (z0 )−1 ,
z→z0 z − z0
which shows both that the derivative exists and finds it. □
Corollary 26.12. Suppose that Ω and Υ are domains, and f is a holomorphic
and bijective function from Ω to Υ. Then f ′ (z) 6= 0 for all z ∈ Ω, and hence the
inverse function f −1 is also holomorphic.
Proof. If f ′ (z0 ) = 0 for some z0 ∈ Ω, then f is not one-to-one near z0 , by
Theorem 26.9 above. This is a contradiction, and hence f ′ never vanishes in Ω.
Now the inverse function f −1 is holomorphic by the inverse function theorem. □
Theorem 26.13 (The open mapping theorem). Suppose that Ω is a domain in
C and f ∈ H(Ω) is nonconstant. Then Range(f ) is open in C.
Proof. Take w0 ∈ Range(f ); then w0 = f (z0 ) for some z0 ∈ Ω.
From the theorem above, there exists ε ∈ R+ such that every point in B(w0 , ε)
also lies in Range(f ). Hence Range(f ) is open. □
Exercise 26.14. Find an example of a holomorphic function defined on an open
set Ω that is not constant, but whose range is not open.
Answer. In light of the open mapping theorem, Ω cannot be connected. We
may take Ω to be B(0, 1) ∪ B(3, 1), and define
(
z when z ∈ B(0, 1)
f (z) =
3 when z ∈ B(3, 1).
Then the range of f is B(0, 1) ∪ {3}, which is not open. 4
4. A PROOF† 161
4. A proof†
In this section, we prove Theorem 26.9, whose statement we first recall.
Theorem. Suppose that h is a nonconstant holomorphic function on B(z0 , R),
where R > 0, and that
X
∞
h(z) = an (z − z0 )n in B(z0 , R).
n=0
Let M be the smallest positive integer for which am 6= 0. Then there exist δ, ε ∈ R+
such that for all w ∈ B ◦ (a0 , ε), there are exactly M points z in B ◦ (z0 , δ) such that
f (z) = w (counting multiplicities).
Proof† . Since h is not constant, it is not possible that an = 0 for all positive
n. Thus M exists. Note that a0 need not be 0.
We may write
X∞ X
∞
′
h(z) − a0 = an (z − z0 ) and h (z) =
n
nan (z − z0 )n−1
n=M n=M
and there exists δ0 ∈ R+ such that h(z) − a0 6= 0 for all z ∈ B(z0 , δ0 ) \ {z0 }. Take
δ = min{δ0 , δ1 }.
The function |h(z) − a0 | is continuous and nonvanishing on ∂B(z0 , δ). Define
ε = min{|h(z) − a0 | : z ∈ ∂B(z0 , δ)}; then ε > 0, by compactness.
Suppose that w ∈ B ◦ (a0 , ε). Take f (z) = a0 − w, and g(z) = h(z) − a0 ; then
|f (z)| < ε ≤ |g(z)| for all z ∈ ∂B(z0 , δ). Rouché’s Theorem implies that f + g and g
have the same number of zeros in B(z0 , δ), counting multiplicities. But z0 is a zero
of multiplicity M for g, and so f + g has M zeros. The derivative of f + g does not
vanish in B ◦ (z0 , δ), and so these zeros all have order one. Since (f +g)(z) = h(z)−w,
there are M distinct points in B(z0 , δ) such that h(z) = w. □
LECTURE 27
In this lecture, we show how residue calculus can be used to evaluate sums. We
also introduce the Riemann ζ function, a focus of current mathematical research.
1. Computing sums
We begin with an exercise.
Exercise 27.1. Denote by ΓN the perimeter of the square with vertices at the
points ±(N + 1/2) ± i(N + 1/2), where N ∈ N, as in Figure 27.1, and define
cot(πz)
f (z) = .
z2
(a) Find the singularities of f inside ΓN , classify them, and compute the
residues of f at the singularities.
(b) Show that
| cot(±(N + 1/2)π + iy)| ≤ 1 ∀y ∈ R,
and
| cot(x ± i(N + 1/2)π)| ≤ coth(π/2) ∀x ∈ R.
(c) Deduce that if Λ is a vertical side and ∆ is a horizontal side of ΓN , then
Z Z
4 4 coth(π/2)
f (z) dz ≤ and f (z) dz ≤ .
Λ 2N + 1 ∆ 2N + 1
X∞
1
(d) Hence find .
n=1
n2
Im
ΓN (N + 1/2) + i(N + 1/2)
Re
+
+
+
+
+
+
+
+
+
+
+
Answer. (a) The function f has singularities when z = 0 and when sin(πz) =
0, that is, when z ∈ Z. By L’Hôpital’s rule,
(z − n) cos(πz) 1
lim (z − n) cot(πz) = lim = 6= 0
z→n z→n sin(πz) π
for every integer n, and we deduce that cot(πz) has a simple pole at every integer
n ∈ Z. Moreover z 7→ z 2 has a zero of order 2 at 0, and it follows that the
singularities of f are at the points n, where n ∈ Z; all the singularities are poles;
the poles are simple except at 0, where the pole is of order 3.
We compute the residues using the appropriate formulae: if n ∈ Z \ {0}, then
cot(πz) 1
Res(f, n) = lim (z − n)f (z) = lim (z − n) 2
= .
z→n z→n z πn2
Further,
1 d2 3 1 d2
Res(f, 0) = lim z f (z) = lim z cot(πz)
z→0 2! dz 2 z→0 2 dz 2
1 d d2
= lim 2 cot(πz) + z 2 cot(πz) .
z→0 2 dz dz
Now
d
cot(πz) = −π cosec2 (πz) ,
dz
and
d2
cot(πz) = 2π 2 cot(πz) cosec(πz).
dz 2
By l’Hôpital’s rule,
d z d2
Res(f, 0) = lim cot(πz) + cot(πz)
z→0 dz 2 dz 2
−π π 2 cos(πz)
= lim + z
z→0 sin2 (πz) sin3 (πz)
−π sin(πz) + π 2 z cos(πz)
= lim
z→0 sin3 (πz)
−π sin(πz) + π 2 z cos(πz) π 3 z 3
= lim
z→0 π3z3 sin3 (πz)
−π sin(πz) + π 2 z cos(πz)
= lim
z→0 π3z3
−π cos(πz) + π 2 cos(πz) − π 3 z sin(πz)
2
= lim
z→0 3π 3 z 2
sin(πz) π
= lim − =− .
z→0 3z 3
(b) By the addition formula for the sine and cosine,
cos(x + iy)
cot(x + iy) =
sin(x + iy)
cos(x) cosh(y) − i sin(x) sinh(y)
=
sin(x) cosh(y) + i cos(x) sinh(y)
1. COMPUTING SUMS 165
By the estimates that we have proved for the integrals along the sides of ΓN ,
π 2 X Z
1
N
16 coth(π/2)
2πi − + = f (z) dz ≤ .
3 π n=1 n2 ΓN 2N + 1
As N → ∞, the right hand side tends to 0, so the left hand side does too. Thus
π 2 X 1
N
− + → 0,
3 π n=1 n2
or in other words,
X∞
1 XN
1 π2
= lim = .
n=1
n2 N →∞ n=1 n2 6
4
X∞
1
Exercise 27.2. Compute 2k
, where k = 2, 3, 4, . . . .
n=1
n
Replacing p by σ gives the desired result for the first sum. The second one is treated
similarly. 4
This sum converges absolutely when Re(s) > 1, and continues meromorphically into
C.
2. THE RIEMANN ZETA FUNCTION 167
It may be shown that ζ is holomorphic in the half plane {s ∈ C : Re s > 1}, and
X
∞
ln n
′
ζ (s) = − .
n=1
ns
The argument is similar to that needed to show that we may differentiate inside an
integral, which we will see shortly.
2.1. The Euler product formula. The zeta function is connected to prime
numbers via the Euler product formula:
Y
ζ(s) = (1 − p−s )−1 .
p
when Re s > 0 (and s 6= 1). To do this, we observe first that when Re(s) > 2, then
X∞
n n−s
s
−
n=1
(n + 1) ns
X∞
n+1 X
∞
1 X
∞
n X s
∞
= − − +
n=1
(n + 1)s n=1 (n + 1)s n=1 ns n=1 ns
= ζ(s − 1) − 1 − ζ(s) + 1 − ζ(s − 1) + s ζ(s)
= (s − 1)ζ(s).
However, it may be seen that the sum on the right hand side of (27.1) converges
when Re(s) > 0, and then analytic continuation (or more precisely meromorphic
continuation) gives the desired result.
We may find other ways to rewrite (s − 1)ζ(s) that converge when Re(s) > −1,
or when Re(s) > −2, and so on.
2.3. More on the zeta function. It is possible to express the zeta function
as an integral:
Z ∞ s−1
1 x
ζ(s) = dx.
Γ(s) 0 ex − 1
168 27. THE RIEMANN ZETA FUNCTION
X∞ Z ∞ X∞
s−1 −(n+1)x 1
= x e dx = Γ(s) s
.
n=0 0 n=0
(n + 1)
Taking the infinite summation out of the integral requires exchanging the order of
a limit and an integral.
Another interesting formula concerning the zeta function is the following func-
tional equation πs
s s−1
ζ(s) = 2 π sin Γ(1 − s) ζ(1 − s).
2
2.4. The Riemann hypothesis. Riemann hypothesised that if ζ(z) = 0, then
either Re(z) = 1/2 or z = −1, −3, −5, . . . . The so-called Riemann hypothesis is
one of the key problems in pure mathematics today. The first person to prove—or
disprove—it will earn a prize of USD 1 million from the Clay Mathematics Institute.
For more information, see
http://www.claymath.org/millennium/.
It is known that, apart from the so-called trivial zeros at −1, −3, −5, …, all
the zeros lie inside the strip {z ∈ C : 0 < Re(z) < 1}, and that the zeta function
has infinitely many zeros on the line {z ∈ C : Re(z) = 1/2}. It can be shown by
numerical computation that all the nontrivial zeros z such that | Im(z)| ≤ 10N lie
on the line for some large N (in fact, finding zeros numerically is one of the test
problems used by computer scientists to demonstrate their virtuosity).
To do this, one takes a rectangular contour which surrounds a segment of the
vertical axis, above the horizontal axis; the vertical sides are where Re(z) = 0 or
Re(z) = 1. Then there are no poles inside the contour; there are formulae which
give the number of zeros on the axis. It is possible to compute the integral on the
left hand side numerically, because the value must be an integer, and if the error is
strictly less than 21 , then the real value must be the integer closest to the numerical
value. If the number of zeros inside the contour is equal to the predicted number of
zeros on the line {z ∈ C : Re(z) = 12 }, then there cannot be any zeros not on the
line inside the contour.
In some sense, the Riemann hypothesis is equivalent to “randomness” of prime
numbers. For instance it tells us that the number of primes in a large interval of
integers can be given by a formula which amounts to saying that the probability of
a large number n being prime is about 1/ log(n). If we could compute enough prime
numbers, then we could test this computationally. However, it would require us to
compute more prime numbers than there are atoms in the universe, so we would
never be able to store all the necessary data.
The Riemann hypothesis is also of interest in theoretical physics, as the distri-
bution of the zeros of a random matrix are related to those of the Riemann zeta
function.
LECTURE 28
Interlude: Integrals†
and
Z Z −R Z ∞
|f (x)| dx rather than |f (x)| dx + |f (x)| dx.
R\[−R,R] −∞ R
1. An example
Just to be sure that there is a real issue to discuss, we consider an example.
Find limt→0+ F (s, t) and limt→+∞ F (s, t), and decide whether or not (28.1) holds
when t0 = 0 and when t = ∞.
3 s2 u1/2 3 u1/2
= lim = lim
u→+∞ 2 s2 exp(us2 ) u→+∞ 2 exp(us2 )
3 u−1/2
= lim = 0.
u→+∞ 4 s2 exp(us2 )
and Z ∞ Z ∞
lim F (s, t) ds = 0 ds = 0.
−∞ t→+∞ −∞
On the other hand, for any fixed t ∈ R+ , by setting u = t−1/2 s, we see that
Z ∞ Z ∞ Z ∞ √
−3/2 2 2 1/2 2 2 π
F (s, t) ds = t tu exp(−u )t du = u exp(−u ) du = ,
−∞ −∞ −∞ 2
which is independent of t, and so
Z ∞ √
π
lim F (s, t) ds =
t→0+ −∞ 2
and Z √
∞
π
lim . F (s, t) ds =
t→+∞ −∞ 2
Thus (28.1) does not hold when t0 = 0 and does not hold when t0 = +∞. 4
It is a good idea to draw the graphs of the function s 7→ F (s, t) for different
values of t to try to understand what is going on in the exercise above.
2. A theorem on limits
Now we prove a theorem giving sufficient conditions for the validity of exchanging
the orders of limits and integrals. In this theorem P may be R or C.
Theorem 28.2. Suppose that F : R × P → R and t0 ∈ P . Suppose also that
there exists µ ∈ R+ such that
(a) F is continuous on R × B(t0 , µ), and
(b) there is a nonnegative function M ∈ L1 (R) such that
|F (s, t)| ≤ M (s) ∀s ∈ R ∀t ∈ B(t0 , µ).
Then Z Z Z
lim F (s, t) ds = lim F (s, t) ds = F (s, t0 ) ds.
t→t0 R R t→t0 R
2. A THEOREM ON LIMITS 171
We make the first integral (which actually has two parts) small by making R big
and using condition (b), and we make the second integral small by using condition
(a).
To make the first integral small, recall that |F (s, t)| ≤ M (s) when t ∈ B(t0 , µ).
Hence, for such t,
Z Z
|F (s, t) − F (s, t0 )| ds ≤ |F (s, t)| + |F (s, t0 )| ds
R\[−R,R] R\[−R,R]
Z
≤2 M (s) ds.
R\[−R,R]
R R
Since the improper integral R M (s) ds converges, we may make R\[−R,R]
M (s) ds
small by taking R large enough. Take R such that
Z
ε
M (s) ds < .
R\[−R,R] 4
Then
Z
ε
|F (s, t) − F (s, t0 )| ds < . (28.3)
R\[−R,R] 2
Further, from (b), F is uniformly continuous on compact subsets of R × B(t0 , µ).
Hence there exists δ ∈ R+ such that
ε
|F (s, t) − F (s, t0 )| < ∀s ∈ [−R, R]
4R
provided that t ∈ B(t0 , δ). For such t,
Z Z
ε ε
|F (s, t) − F (s, t0 )| ds ≤ ds = . (28.4)
[−R,R] [−R,R] 4R 2
Set η = min{δ, µ}. Putting everything together, (28.3) and (28.4) hold when
t ∈ B(t0 , η). Combining with (28.2), we see that
Z Z
ε ε
F (s, t) ds − F (s, t0 ) ds < + = ε
R R 2 2
when t ∈ B(t0 , η), as required. □
172 28. INTERLUDE: INTEGRALS†
3. Remarks
Though we state our results for real-valued integrands F , they also hold for
complex-valued integrands, by an identical argument.
The arguments also apply when the parameter t is constrained to lie in a subset
of R or C; we just have to add this constraint to the statement of the theorem, and
to each step of the proof. By replacing t by 1/t, we can also establish a version of
the theorem for limits as t tends to ∞.
Our arguments also apply to integrals where one or both the limits of integration
is finite, that is, to integrals such as
Z
H(s, t) ds;
[a,b]
in this case, the proof of Theorem 28.2 simplifies, as we replace [−R, R] by [a, b] and
do not need to consider the other component of the integral.
Finally, the result also extends to integration in Rn ; we replace the integrals over
[−R, R] and over R \ [−R, R] by integrals over B(0, R) and over R \ B(0, R).
Finally, we state a theorem giving sufficient conditions for the validity of ex-
changing the order of differentiation and integration.
Theorem 28.4. Suppose that G : R × P → R and t0 ∈ P , where P ⊆ R, and
that the function s 7→ G(s, t0 ) ∈ L1 (R). Suppose also that there exists µ ∈ R+ such
that
(a) (s, t) 7→ ∂t
∂
G(s, t) is continuous on R × B(t0 , µ), and
(b) there is a nonnegative function M ∈ L1 (R) such that
∂
G(s, t) ≤ M (s) ∀s ∈ R ∀t ∈ B(t0 , µ).
∂t
5. APPLICATIONS† 173
Then Z Z
∞ ∞
∂ ∂
G(s, t) ds t=t0
= G(s, t) t=t0
ds. (28.6)
∂t −∞ −∞ ∂t
Proof. The idea of the proof is to write
Z ∞ Z ∞
∂
G(s, t) ds = lim F (s, h) ds,
∂t −∞ t=t0 h→0 −∞
where Z t0 +h
1 ∂
G(s, t) dt if h 6= 0
F (s, h) = h t 0
∂t
∂ G(s, t) if h = 0.
∂t t=t0
Then we show that F satisfies the hypotheses of the previous theorem (where we
suppose that h → 0). It follows that
Z ∞ Z ∞
∂
G(s, t) ds = lim F (s, h) ds
∂t −∞ t=t0 h→0 −∞
Z ∞
= lim F (s, h) ds
−∞ h→0
Z ∞
∂
= G(s, t0 ) ds,
−∞ ∂t
as required. □
Remark 28.5. This result also applies when t is a complex variable, and differ-
entiation with respect to t is complex differentiation.
Differentiation with respect to a parameter may be used to compute integrals.
For instance, to compute
Z 1 2
x −1
dx,
0 ln x
we may find
Z Z 1 t
d 1 xt − 1 x −1
dx and dx t=0 .
dt 0 ln x 0 ln x
5. Applications†
The gamma function Γ is defined by an integral.
Definition 28.6. The function Γ is given by
Z ∞
Γ(z) = sz−1 e−s ds.
0
This integral converges absolutely when Re z > 0, and continues meromorphically
into C.
It follows from Theorem 28.4 that the gamma function is holomorphic, and
Z ∞
′
Γ (z) = sz−1 ln(s) e−s ds.
0
174 28. INTERLUDE: INTEGRALS†
It is easy to check that that Γ(z) = (z−1)Γ(z−1), by integration by parts, and hence
Γ(n) = (n − 1)! when n ∈ Z+ . This formula allows to extend Γ meromorphically
into C.
Now we consider some integrals that may be evaluated in terms of this function.
Exercise 28.7. Find changes of variables or other transformations such that
the following integrals may be converted into an integral of the form
Z π/2
cosa (θ) sinb (θ) dθ.
0
Z 1 Z 1
(a) (1 − x ) x dx;
2 α β
(b) (1 − y)α y β dy;
Z 0 π/2 0
tanc (φ)
(c) dφ;
Z0 ∞ secd (φ) Z ∞
tγ sε
(d) dt; (e) ds.
0 (1 + t2 )δ 0 (1 + s)ζ
Answer. Omitted 4
Observe that
Z ∞ Z ∞
a −x2 ′
x e dx = 1
2
(x′ )(a−1)/2 e−x dx′ = 12 Γ( a+1
2
);
0 0
similarly, Z ∞
y b e−y dy = 12 Γ( b+1
2
2
),
0
and Z ∞
ra+b+1 e−r dr = 12 Γ( a+b+2
2
2
).
0
By using polar coordinates, we see that
Z ∞ Z ∞
a −x2
y b e−y dy
2
1 a+1 1 b+1
2
Γ( 2 ) 2 Γ( 2 ) = x e dx
Z0 ∞ Z ∞ 0
xa y b e−x −y dx dy
2 2
=
0 0
Z π/2 Z ∞
ra cosa (θ) rb sinb (θ) e−r r dr dθ
2
=
0 0
Z ∞ Z π/2
a+b+1 −r2
= r e dr cosa (θ) sinb (θ) dθ
0 0
Z π/2
= 12 Γ( a+b+2
2
) cosa (θ) sinb (θ) dθ.
0
Hence Z π/2
Γ( a+1 )Γ( b+1 )
cosa (θ) sinb (θ) dθ = 2
a+1
2
b+1
.
0 2Γ( 2 + 2 )
LECTURE 29
R 2π
x ∈ R. Then φ ∈ L1 (R) and R φ(x) dx = 1.
The convergence of the improper integral follows from that fact that | e−ixξ | = 1
for all x, ξ ∈ R, and properties of the Riemann integral.
3. Examples
Exercise 29.3. Define f : R → R by f (x) = e−|x| . Show that
2
fb(ξ) = ∀ξ ∈ R.
1 + ξ2
Answer. First of all, it is easy to check that f ∈ L1 (R). By definition,
Z
fb(ξ) = lim e−|x| e−ixξ dx.
R→∞ [−R,R]
175
176 29. THE FOURIER TRANSFORMATION
Now, Z Z Z
−|x| −ixξ x −ixξ
e e dx = e e dx + e−x e−ixξ dx
[−R,R] [−R,0] [0,R]
Z 0 Z R
= x(1−iξ)
e dx + e−x(1+iξ) dx
−R 0
−R(1−iξ)
1 e e−R(1+iξ) 1
− = − + ,
1 − iξ 1 − iξ 1 + iξ 1 + iξ
and since | e | = 1, it follows that
iRξ
Hence Z Z
∞ ∞
e−iξx eiξx
dx = dx
−∞ x2 + 1 −∞ x2 + 1
Z ∞
ei|ξ|x
= dx
−∞ x2 + 1
−|ξ|
= πe ,
as required. 4
1
Exercise 29.5. The Gaussian φ is defined by φ(x) = √ e−x /2 . Show that
2
2π
b = e−ξ
2 /2
φ(ξ) .
Answer. First, by Example 29.1,
Z ∞
φ(x) dx = 1. (29.1)
−∞
plied to the rectangular contour with vertices at −R, R, R + iξ and −R + iξ, implies
that
Z R Z R Z R+iξ Z −R+iξ
−(x+iξ)2 /2 −x2 /2 −z 2 /2
e−z /2 dz. (29.3)
2
e dx = e dx + e dz −
−R −R R −R
tend to 0 when R tends to infinity. We show how to handle the first term; the second
term may be handled similarly. We parametrise the line segment between R and
R + iξ by writing z = R + iy, where 0 ≤ y ≤ ξ. For such z,
e−z = e−(R+iy) = e−(R = e−(R
2 /2 2 /2 2 +2iRy−y 2 )/2 2 −y 2 )/2 2 −R2 2 −R2
= ey ≤ eξ .
By the M L Lemma,
Z R+iξ
e−z
2 /2 2 −R2
dz ≤ eξ |ξ| → 0 as R → ∞.
R
b =
Combining these two observations with (29.2) and (29.3), we deduce that φ(ξ)
−ξ 2 /2
e for all ξ ∈ R, as required. 4
Exercise 29.6. Let f (x) = 1 if x ∈ [−1, 1] and f (x) = 0 otherwise. Show that
fb ∈
/ L1 (R).
and so f is continuous.
178 29. THE FOURIER TRANSFORMATION
To justify this last step, write F (x, ξ) = f (x) e−ixξ for all z, ξ ∈ R. Observe that
|F (x, ξ)| = |f (x)| ∈ L1 (R),
by hypothesis, and so condition (b) of Theorem 28.2 holds. Further, if f is contin-
uous, then F is continuous, and condition (a) of Theorem 28.2 holds.
We will not consider the vanishing at infinity. □
Theorem 29.9 (The inversion formula for the Fourier transform). If f ∈ M(R)
and fb ∈ M(R), then Z
1
f (x) = fb(ξ) eixξ dξ.
2π R
as required. □
The change in the order of integration in the double integral is justified by the
fact that Z Z
|g(x)| |ψ(ξ)| dx dξ < ∞.
R R
Now we claim that we may exchange limits and integrals to show that
√ Z ∞ −y 2 /2
√ Z ∞
lim g(δy) e−y /2 dy
2
lim 2π g(δy) e dy = 2π
δ→0 −∞ −∞ δ→0
√ Z ∞
g(0) e−y /2 dy = 2π g(0),
2
= 2π
−∞
To justify the first of these, write F (y, δ) = g(δy) e−y /2 . Since g is continuous,
2
Thus we may exchange the limit and the integral. The second formula may be
justified similarly.
Hence Z ∞
1
g(0) = gb(ξ) dξ. (29.4)
2π −∞
180 29. THE FOURIER TRANSFORMATION
Now suppose that x ∈ R, f ∈ M(R) and g(y) = f (y + x). Then, by the change
of variable u = y + x, we see that
Z Z ∞
−iyξ
gb(ξ) = g(y) e dy = f (y + x) e−iyξ dy
ZR∞ −∞
Z ∞
= f (u) e −i(u−x)ξ
du = e ixξ
f (u) e−iuξ du = eixξ fb(ξ)
−∞ −∞
for all ξ ∈ R, and by applying (29.4) to this g, we obtain
Z ∞ Z ∞
1 1
f (x) = g(0) = gb(ξ) dξ = fb(ξ) eixξ dξ,
2π −∞ 2π −∞
completing the proof of Theorem 29.9. □
LECTURE 30
Answer. For part (a), we observe that ebt = eRe(bt) = eRe(b)t for all t ∈
[0, +∞); we then take C = |a| in the definition of exponential type.
For (b), observe first that if ε > 0, then
p(t)
lim p(t) e−εt = lim = 0,
t→+∞ t→+∞ eεt
from l’Hôpital’s rule. This means that |p(t) e−εt | < 1 if t is big enough, say t > R.
The function t 7→ p(t) e−εt is continuous in the compact interval [0, R], and so it
is bounded there, that is, there exists B ∈ R+ such that |p(t) e−εt | ≤ B for all
t ∈ [0, R]. We deduce that |p(t) e−εt | ≤ B + 1 for all t ∈ [0, R) and all t ∈ [R, +∞),
whence
|p(t)| ≤ C eεt ∀t ∈ [0, +∞),
where C = B + 1, and finally
p(t) ebt ≤ C eεt ebt = C e(Re(b)+ε)t ∀t ∈ [0, +∞).
Finally, for part (c), suppose with a view to a contradiction that there is a real
number A such that
2
et ≤ C eAt ∀t ∈ [0, +∞).
Then taking logarithms, we see that
t2 ≤ ln(C) + At ∀t ∈ [0, +∞),
181
182 30. THE LAPLACE TRANSFORM
We prove something very close to (c) in Theorem 30.7, and omit this.
To prove (d), we integrate by parts:
Z ∞ Z ∞
−zs
Lg(z) = g(s) e ds = f ′ (s) e−zs ds
0
Z ∞ 0
= [ f (s) e−zs ]∞
0 − f (s) (−z) e−zs ds = −f (0) + zLf (z).
0
This last argument is not really complete—in the integration by parts, we really
ought to deal with integrals over [0, R] and then let R tend to +∞. □
Remembering these properties can save time; for example, by part (a), since the
Laplace transform of t3 is 6/z 4 , the Laplace transform of t3 e−2t is just 6/(z + 2)4 .
184 30. THE LAPLACE TRANSFORM
where gσ (t) = f (t) e−σt if t ≥ 0 and gσ (t) = 0 if t < 0. By the inversion formula for
the Fourier transform, if t > 0 then
Z ∞ Z σ+iR
−σt 1 −σt 1
f (t) e = Lf (σ + iy) e dy = e
iyt
lim Lf (z) ezt dz.
2π −∞ R→∞ 2πi σ−iR
Strictly speaking, this is not a proof, because we proved the Fourier inversion
formula for a more limited class of functions. However, with a little more work the
Fourier inversion formula may be proved for more general functions, justifying the
above argument.
The key fact is that the Laplace transformation is invertible. Thus to find the
inverse Laplace transform of g, it suffices to find a continuous function f of expo-
nential type such that Lf = g; then f is the desired inverse transform.
5. A PROOF† 185
4. Examples
Exercise 30.10. Find the continuous function f : [0, ∞) → C of exponential
1
type 1+ such that Lf (z) = .
(z − 1)2
1
Answer. We have seen that the Laplace transform of t 7→ t is z 7→ 2 . Now
z
1
the Laplace transform of t 7→ t e is z 7→
t
, by Proposition 30.8. 4
(z − 1)2
5. A proof†
For convenience, we restate the result that we are going to prove.
Theorem. If f : [0, ∞) → C is locally integrable and of exponential type A+,
then Lf is holomorphic on HA . Further,
d
Lf (z) = Lg(z) ∀z ∈ HA ,
dz
where g : [0, +∞) → C is given by g(t) = −t f (t).
Proof. We do this by applying Theorem 26.4, which states that we may ex-
change the order of differentiation with respect to a parameter and integration when
three conditions are satisfied. By definition,
Z ∞
Lf (z) = f (s) e−zs ds.
0
and similarly
dG
(s, z) ≤ C e−εs/2 ,
dz
and the integrability conditions of Theorem 26.4 are also satisfied. Thus we may
differentiate under the integral and obtain the desired result. □
LECTURE 31
We may use the Laplace transform to solve differential and integral equations
on [0, ∞).
1. Differential equations
The Laplace transform may be used to solve differential equations on [0, +∞).
Exercise 31.1. Use the Laplace transform to solve the ordinary differential
equation
d2 u du
2
(t) − 2 (t) + u(t) = t ∀t ∈ [0, +∞),
dt dt
du
subject to the initial conditions u(0) = (0) = 0.
dt
Answer. Recall from Proposition 30.8 that L(f ′ )(z) = zLf (z) − f (0). We take
the Laplace transform of both sides of the equation and use this result and the initial
conditions to obtain
1
z 2 Lu(z) − 2zLu(z) + Lu(z) = 2 .
z
Thus
1 1
Lu(z) = 2 2 = 2 .
z (z − 2z + 1) z (z − 1)2
Now it is just a matter of inverting the Laplace transform. First, by partial fractions,
1 2 2 1 1
Lu(z) = = − + 2+ ;
z 2 (z − 1) 2 z z−1 z (z − 1)2
n!
Now the Laplace transform of tn is n+1 for all n ∈ N, and the Laplace transform
z
−at n n!
of e t is for all a ∈ C by Proposition 30.8, and so
(z + a)n+1
u(t) = 2 − 2 et + t + t et ,
and we are done. 4
Notice that this solution could have been found by more elementary methods,
and the use of the Laplace transform here is using a sledgehammer to crack a nut.
The same cannot be said so easily of the next examples.
Example 31.2. Suppose that someone wiggles one end of a very long straight
piece of string up and down for a certain time. How does the string move?
187
188 31. APPLICATIONS TO DIFFERENTIAL AND INTEGRAL EQUATIONS
Let us suppose that the vertical displacement of the string at position x and
time t is given by u(x, t). The string satisfies the wave equation
∂ 2u ∂ 2u
= 2,
∂x2 ∂t
and we may take the initial conditions u(x, 0) = ∂∂tu (x, 0) = 0 when x > 0 and
u(0, t) = f (t), where f (t) = 0 unless a < t < b, where b > a > 0. In addition, we
shall make other assumptions along the way to get a solution.
We write Lu(x, z) for the Laplace transform of u in the t variable; this means
that x 6= Re(z). Then (assuming that there are no problems with exchanging the
order of differentiation and integration)
∂ 2 Lu ∂ 2u ∂ 2u
(x, z) = L (x, z) = L (x, z).
∂x2 ∂x2 ∂t2
Further, from Proposition 30.8,
∂ 2u ∂u ∂u ∂u
L 2
(x, z) = zL (x, z) − (x, 0) = z 2 Lu(x, z) − u(x, 0) − (x, 0).
∂t ∂t ∂t ∂t
The initial conditions imply that u(x, 0) = ∂∂tu (x, 0) = 0, and so
∂ 2 Lu
(x, z) = z 2 Lu(x, z).
∂x2
The solution to this equation is
Lu(x, z) = A(z)ezx + B(z)e−zx .
It is physically sensible to suppose that u(x, t) → 0 as x → +∞, so it seems
reasonable to suppose that A(z) = 0. Thus, taking x = 0, we see that
B(z) = Lu(0, z) = Lf (z).
Finally,
Lu(x, z) = e−zx Lf (z).
To unravel the Laplace transform, observe that if g(t) = f (t − c), then g(t) = 0 if
t < c, and so
Z ∞ Z ∞
−tz
Lg(z) = g(t)e dt = g(t)e−tz dt
Z0 ∞ c
Z ∞
−tz
= f (t − c)e dt = f (t)e−(t+c)z dt = e−cz Lf (z).
c 0
Hence
u(x, t) = f (t − x) ∀t ∈ R+ .
Example 31.3. Use the Laplace transform to find the solution u : [0, ∞) ×
[0, ∞) → R of the partial differential equation
∂u ∂ 2u
(s, t) = 2 (s, t) ∀(s, t) ∈ R+ × R+ ,
∂t ∂s
subject to the initial conditions u(s, 0) = 0 for all s ∈ R+ and u(0, t) = 1 for all
t ∈ R+ .
We omit the answer to this problem, which is known as the heat equation on the
half-line [0, ∞).
2. CONVOLUTION ON [0, ∞) 189
2. Convolution on [0, ∞)
Throughout this section, f and g will denote locally integrable functions on
[0, ∞).
Definition 31.4 (Convolution on [0, ∞)). The convolution of f and g is the
function f ∗ g : [0, ∞) → C given by
Z t
f ∗ g(t) = f (s) g(t − s) ds.
0
Item (c) is a consequence of the fact that the exponential function is multiplica-
tive (that is, ea+b = ea eb ). Again, by the change of variables u = t − s, we see
that
Z +∞ Z t
L(f ∗ g)(z) = f (s) g(t − s) ds e−zt dt
0 0
Z +∞ Z t
= f (s) e−zs g(t − s) e−z(t−s) ds dt
0 0
Z +∞ Z +∞
= f (s) e−zs g(u) e−zu ds du
0 0
= Lf (z) Lg(z)
for all z ∈ HA+ . □
190 31. APPLICATIONS TO DIFFERENTIAL AND INTEGRAL EQUATIONS
3. Integral equations
Exercise 31.6. Use the Laplace transform to find a function u satisfying the
integral equation
Z
2 t
u(t) = t −
2
(t − s)3 u(s) ds.
3 0
This is an example of a Volterra integral equation; these arise in many areas of
application, including finance and biology.
Answer. We first note that this is a “convolution equation”, in the sense that
it may be expressed as
u(t) = t2 − h ∗ u(t),
where h(t) = 23 t3 .
Taking the Laplace transform of this equation and using Proposition 31.5, we
obtain
2 4Lu(z)
Lu(z) = 3 − ,
z z4
which may be rearranged to give
2z
Lu(z) = .
z4+4
Again, u may be found by applying the inversion formula for the Laplace transform
or by the method of partial fractions and inspection. This time we use the inversion
formula.
First observe that Lu has four simple poles, at the fourth roots of −4. These
roots are (1 + i)ik , where k = 1, 2, 3, 4. If w is any one of these roots, then, by
l’Hôpital’s rule,
2z etz z−w
Res(L u(z) etz , z = w) = lim (z − w) = 2w e tw
lim
z→w z4 + 4 z→w z 4 + 4
1 1
= 2w etw lim 3 = 2w etw 3
z→w 4z 4w
1
= etw 2 .
2w
k k
e(1+i)i t e(1+i)i t
k
Thus the residue at (1 + i)i is if k is odd and − if k is even.
4i 4i
Using the inversion formula for the Laplace transform, we obtain
e(1+i)t e(−1+i)t e(−1−i)t e(1−i)t
u(t) = − + −
4i 4i 4i 4i
−t
1 e −e
t
1 e − e−t
t
= eit − e−it
2i 2 2i 2
t −t
it −it
e −e e −e
=
2 2i
= sinh t sin t,
and we are done. 4
3. INTEGRAL EQUATIONS 191
Exercise 31.7. Use the Laplace transform to find a function m satisfying the
renewal equation (31.1), when f (t) = e−t for all t ∈ R+ .
Answer. We note that the renewal equation is a convolution equation, and
apply the Laplace transform, to get
Lm(z) = LF (z) + Lf (z) Lm(z).
Then
LF (z)
Lm(z) = .
1 − Lf (z)
Now we compute the Laplace transforms of f and F . Clearly,
1
Lf (z) = ∀z ∈ H−1 ,
z+1
z
and hence 1 − Lf (z) = . Further,
z+1
Z +∞ Z t Z +∞ Z t
−zt
LF (z) = f (s) ds e dt = e−s e−zt ds dt
0 0 0 0
Z +∞ Z ∞ Z +∞ h −zt i∞
−s −zt e
= e e dt ds = e−s ds
0 s 0 −z s
Z +∞ −zs Z +∞ −(z+1)s
e e
= e−s ds = ds
0 z 0 z
h e−(z+1)s i∞ 1
= = .
−z(z + 1) 0 z(z + 1)
Thus
1
LF (z) z(z + 1) 1
Lm(z) = = z = 2,
1 − Lf (z) z
z+1
and so m(t) = t. 4
LECTURE 32
In this lecture, we introduce the Dirichlet problem. We show that the solution to
this problem in a domain is unique, if it exists, using the so-called maximum prin-
ciple; later we consider the questions of existence and of finding explicit solutions.
Given a subset S of C, we write C(S) for the collection of all continuous functions
on S. At the risk of some confusion, we will write Ω both for an open subset of C
and for the corresponding subset of R2 .
Recall that a real-valued function h of two real variables u and v is said to be
harmonic in Υ its partial derivatives of order one and two are continuous in Υ, and
∆h(u, v) = 0 for all (u, v) ∈ Υ, where
2
∂ ∂2
∆h(u, v) = + h(u, v).
∂u2 ∂v 2
The differential operator ∆ is known as the Laplacian or Laplace operator.
Problem 32.1 (Dirichlet problem). Suppose that Ω is a bounded domain and
that ∂Ω is a finite union of contours. Given a function b ∈ C(∂Ω), find a harmonic
function u on Ω such that
∆u(x, y) = 0 for all (x, y) ∈ Ω
lim u(x, y) = b(x0 , y0 ) for all (x0 , y0 ) ∈ ∂Ω.
(x,y)→(x0 ,y0 )
Proof. If f were not constant, then Range(f ) would contain an open ball
around f (z0 ), by the open mapping theorem, and |f (z0 )| could not be the maxi-
mum value. □
Exercise 32.3. Find an example of a nonconstant holomorphic function defined
on a bounded open set Ω for which there is a point z0 ∈ Ω such that
|f (z0 )| = max{|f (z)| : z ∈ Ω}.
193
194 32. THE DIRICHLET PROBLEM
Exercise 32.4. Find the maximum value of |sin(z)| as z varies over the closed
unit disc B(0, 1).
Answer. By the maximum modulus principle, the maximum value of |sin(z)|
is attained on the boundary of the ball, where x2 + y 2 = 1. Now
|sin(x + iy)| = |sin(x) cosh(y) + i cos(x) sinh(y)|
1/2
= sin2 (x) cosh2 (y) + cos2 (x) sinh2 (y)
1/2
= sin2 (x) + sin2 (x) sinh2 (y) + cos2 (x) sinh2 (y)
1/2
= sin2 (x) + sinh2 (y)
1/2
= sin2 (x) − x2 + sinh2 (y) − y 2 + 1 .
Changing the sign of x or y does not change the value, so it will suffice to look
for a point where x ≥ 0 and y ≥ 0.
It is easy to check that, for such x and y,
d
sin2 (x) − x2 = 2 sin(x) cos(x) − 2x ≤ 2x cos(x) − 2x
dx
= −2x(1 − cos(x)) ≤ 0
and similarly
d
sinh2 (y) − y 2 = 2 sinh(y) cosh(y) − 2y ≥ 2y cosh(y) − 2y
dy
= 2y(cosh(y) − 1) ≥ 0.
It follows that we can make |sin(x + iy)| as large as possible by making y as big
and x as small as possible. Hence the maximum value is sinh(1), attained when
z = ±i. 4
Proof. We will first prove this for a set Ω1 that is simply connected. Suppose
that there is a point (x0 , y0 ) ∈ Ω1 such that |h(x0 , y0 )| = max{|h(x, y)| : (x, y) ∈ Ω1 }.
Since Ω1 is simply connected, we may find a harmonic conjugate of h, and hence
4. THE DIRICHLET PROBLEM AND HOLOMORPHIC MAPPINGS 195
write h(x, y) = Re(f (x + iy)), for some f ∈ H(Ω1 ). Consider the holomorphic
function exp ◦f . Evidently
|exp ◦f (x + iy)| = exp(Re(f (x + iy))) = exp(h(x, y)),
and this attains its maximum at x0 + iy0 , which lies in the interior of Ω1 . By the
maximum modulus principle, exp ◦f is constant in Ω1 , that is, f is constant in Ω1 ,
and so h is constant in Ω1 .
To deal with the general case of a domain Ω that need not be simply connected,
we take a arbitrary point (x1 , y1 ) of Ω, and show that h(x1 , y1 ) = h(x0 , y0 ), and
conclude that h is constant. Since Ω is connected, there is a polygonal path joining
(x1 , y1 ) and (x0 , y0 ); by omitting any loops, we may assume that this is simple. Now
by “fattening up” this path a little, we may find a connected and simply connected
open subset Ω1 of Ω that contains both points. Observe that
h(x0 , y0 ) ≤ max{h(x, y) : (x, y) ∈ Ω1 } ≤ max{h(x, y) : (x, y) ∈ Ω} ≤ h(x0 , y0 ),
since (x0 , y0 ) is just one of the points considered in the first maximum, since Ω1 ⊆ Ω,
and by hypothesis. In particular, h(x0 , y0 ) = max{h(x, y) : (x, y) ∈ Ω1 }, and
so we may apply the result for the simply connected set Ω1 , and conclude that
h(x0 , y0 ) = h(x1 , y1 ); hence h is constant. □
Then u1 = u2 .
are real-valued, and let F (x, y) = (u(x, y), v(x, y)) be the corresponding vector-valued
function. Then h ◦ F is harmonic in Ω.
then (x, y) 7→ h(u(x, y), v(x, y)) solves the Dirichlet problem
∆u(x, y) = 0 for all (x, y) ∈ Ω
lim u(x, y) = b(x0 , y0 ) for all (x0 , y0 ) ∈ B.
(x,y)→(x0 ,y0 )
5. Examples
In the exercises below, we set H = {(x, y) ∈ R2 : y > 0}, that is, H is the upper
half plane. Boundary values should be interpreted as continuous limits at points of
continuity on the boundary.
Exercise 32.10. Find bounded harmonic functions h in H such that
(a) h(x, 0) = 1 for all x ∈ R;
(b) h(x, 0) = 0 if x > 0 and h(x, 0) = π if x < 0;
(c) h(x, 0) = sgn(x) for all x ∈ R;
(d) h(x, 0) = 1 if 0 < x < 1 and h(x, 0) = 0 if x < 0 or x > 1.
Answer. These are all combinations of constants and argument functions. 4
y
= .
(x − s)2 + y 2
Theorem 32.11. Suppose that g is a bounded continuous function on R, and
define h on the upper half-plane by the formula
Z
1 ∞ y
h(x + iy) = g(s) ds ∀x ∈ R ∀y ∈ R+ .
π −∞ (x − s)2 + y 2
198 32. THE DIRICHLET PROBLEM
Then the integral defining h converges, and h has the following properties:
(a) h is harmonic
(b) h is bounded
(c) limx+iy→x0 h(x + iy) = g(x0 ) for all x0 ∈ R.
Further, h is the only function with these properties.
Proof. See next lecture. □
In this lecture, we first show that our proposed solution to the Dirichlet problem
in the upper half plane has the desired properties. Next, we examine conformal
mappings: mappings that preserve “form”, at least locally. Finally we solve the
Dirichlet problem in some other regions.
Then the integral defining h converges, and h has the following properties:
(a) h is harmonic
(b) h is bounded
(c) limx+iy→x0 h(x + iy) = g(x0 ) for all x0 ∈ R.
Further, h is the only function with these properties.
Proof. To show that h is harmonic, we observe that
Arg(z − s) = Re(−i Log(z − s)),
and so
∂ ∂ ∂ 1
Arg(z − s) = Re(−i Log(z − s)) = Re (−i Log(z − s)) = Re i ,
∂s ∂s ∂s z−s
which is harmonic, so if we can differentiate under the integral sign, the harmonicity
will follow. It is possible to verify that it is legitimate to exchange the order of
differentiation and integration; we omit the details.
To show that h is bounded, we observe that
Z Z
1 y 1 y
g(s) ds ≤ | g(s)| ds
R π (x − s) + y R π (x − s) + y
2 2 2 2
Z
1 y
≤ sup{|g(s)| : s ∈ R} ds
R π (x − s) + y
2 2
Z
1 1
= sup{|g(s)| : s ∈ R} 2
dt
R πt +1
= sup{|g(s)| : s ∈ R},
by the change of variables t = (x − s)/y.
199
200 33. CONFORMAL MAPPINGS AND HARMONIC FUNCTIONS
Yet another “limit inside the integral” argument shows that this goes to 0 as x + iy
goes to x0 , that is, as x → x0 and y → 0+. □
at least when z is close to z0 . What does this tell us about the geometric properties
of f ? Because we are only dealing with an approximation, with errors that increase
the further z is from z0 , it seems unlikely that f will preserve ratios of lengths.
However, if might still be true that f preserves angles. But since f (T ) might have
curved sides, even if T has straight sides, we need to be able to measure angles
between curves.
3. DIFFERENTIABILITY AND ANGLES BETWEEN TWO CURVES 201
and choose the coefficients an to get the desired boundary value functions.
Note that a + b ln(| · |) is a radial harmonic function in the annulus, which may
be given arbitrary boundary values that are constant on the inner and the outer
circles. This kind of problem arises in electrostatics.
LECTURE 34
Loose ends
205