an invitation to complex analysis - warwick · 2019. 12. 13. · the geometric intuitions sketched...

30
An Invitation to Complex Analysis Ryan Acosta Babb Univeristy of Warwick

Upload: others

Post on 31-Mar-2021

1 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: An Invitation to Complex Analysis - Warwick · 2019. 12. 13. · The geometric intuitions sketched above will be familiar from analysis in R. In one dimension. we use the absolute

An Invitation to Complex Analysis

Ryan Acosta BabbUniveristy of Warwick

Page 2: An Invitation to Complex Analysis - Warwick · 2019. 12. 13. · The geometric intuitions sketched above will be familiar from analysis in R. In one dimension. we use the absolute

Contents

1 Real Analysis in R2 21.1 The Fundamental Theorem of Calculus for Gradient Fields . . 21.2 Continuity in R2 . . . . . . . . . . . . . . . . . . . . . . . . . 4

2 Topology in the Plane 52.1 Open and Closed Sets . . . . . . . . . . . . . . . . . . . . . . 62.2 Continuity in C . . . . . . . . . . . . . . . . . . . . . . . . . . 82.3 Connectedness and Compactness . . . . . . . . . . . . . . . . 9

3 Differentiability in C 133.1 Differentiability . . . . . . . . . . . . . . . . . . . . . . . . . . 133.2 The Cauchy-Riemann Equations . . . . . . . . . . . . . . . . . 15

4 Complex Integration 164.1 Line and Arc-Length Integrals . . . . . . . . . . . . . . . . . . 164.2 Cauchy’s Integral Theorem . . . . . . . . . . . . . . . . . . . . 184.3 Cauchy’s Integral Formula . . . . . . . . . . . . . . . . . . . . 224.4 Complex Analytic Functions . . . . . . . . . . . . . . . . . . . 24

5 Applications 255.1 The Fundamental Theorem of Algebra . . . . . . . . . . . . . 255.2 Contour Integration . . . . . . . . . . . . . . . . . . . . . . . . 27

6 Bibliography 29

1

Page 3: An Invitation to Complex Analysis - Warwick · 2019. 12. 13. · The geometric intuitions sketched above will be familiar from analysis in R. In one dimension. we use the absolute

1 Real Analysis in R2

The complex plane C can be identified easily with the real plane R2 via thecorrespondence:

C ∼= R2

x+ iy ←→ (x, y).

Familiar concepts from multivariable calculus, such as partial derivativesand line integrals can thus be carried over almost effortlessly into the complexsetting. We begin by recalling some facts about analysis in R2 which willthen receive a surprising new life in the complex plane.

1.1 The Fundamental Theorem of Calculus for Gradi-ent Fields

Given a function F : R2 → R2 and a differentiable curve γ : [a, b] → R2, wedefine the line integral of F along γ to be∫

γ

F (x) dx :=

∫ b

a

F (γ(t)) · γ′(t) dt.

(The · in the right-hand side denotes the vector dot product in R2). Inparticular, if γ is a closed curve, meaning γ(a) = γ(b), then we write∮

γ

F (x) dx

to emphasise that the curve is closed.If f : R2 → R, has differentiable projections f(·, y) and f(x, ·), the gra-

dient of f is the vector valued function

∇f :=

(∂f

∂x,∂f

∂y

)If further g : R → R2, g(t) = (gx(t), gy(t)), is differentiable, then f g :

R→ R is differentiable and the chain rule applies

(f g)′(t) =∂f

∂x(g(t))g′x(t) +

∂f

∂y(g(t))g′y(t) = ∇f(g(t)) · g′(t)

Just as in one-dimensional calculus we can “integrate a derivative” byevaluating it at the end points of an interval, we can “integrate a gradient”along a curve γ by evaluating the function at the end-points of the curve.

2

Page 4: An Invitation to Complex Analysis - Warwick · 2019. 12. 13. · The geometric intuitions sketched above will be familiar from analysis in R. In one dimension. we use the absolute

Theorem 1.1 (Fundamental Theorem of Calculus for Gradient Fields). Letf : R2 → R have partial derivatives on [a, b], and γ : [a, b] → R2 a differen-tiable curve. Then ∫

γ

∇f(x) dx = f(γ(b))− f(γ(a)).

Proof. The theorem follows from the chain rule and the Fundamental Theo-rem of Calculus in R:∫γ

∇f(x) dx =

∫ b

a

∇f(γ(t)) · γ′(t) dt =

∫ b

a

(f γ)′(t) dt = f(γ(b))− f(γ(a)).

This implies that line integrals of gradients are path-independent. In fact,the converse is also true, which we prove with an additional result which wewill “revisit” in §4.2.

Theorem 1.2. For a continuous vector field F : R2 → R2, the following areequivalent:

(i) F is conservative, i.e. ∮γ

F (x) dx = 0.

for all closed curves γ.

(ii) All line integrals of F are path independent, i.e. if γ1 and γ2 are curveswith the same end-points, then∫

γ1

F (x) dx =

∫γ2

F (x) dx.

(iii) F is a gradient field, i.e. F = ∇f for some f : R2 → R.

Proof. The equivalence between (i) and (ii) is straightforward. If all closedpaths integrate to zero, then given any two paths γ1 and γ2 with the sameend point, we can form the path Γ = γ1−γ2 understood as “traverse γ1, thentraverse γ2 in the opposite direction”. Clearly Γ is closed, so we have

0 =

∮Γ

F (x) dx =

∫γ1

F (x) dx−∫γ2

F (x) dx.

3

Page 5: An Invitation to Complex Analysis - Warwick · 2019. 12. 13. · The geometric intuitions sketched above will be familiar from analysis in R. In one dimension. we use the absolute

Conversely, if the line integrals of F are path independent, then we cancalculate any closed path γ by picking a single point p ∈ γ and consideringthe constant path t 7→ p. Hence,∮

γ

F (x) dx =

∫ 0

0

F (p) · 0 dt = 0.

The implication (iii) =⇒ (ii) follows from Theorem 1.1. For the converse,pick any point p ∈ R2 and consider the path t 7→ (1 − t)p + tx. We writethe line integral, suggestively, as

f(x) =

∫ x

p

F (y) dy.

It can then be shown that ∇f = F . (The proof is similar to that of Theorem4.7 below.)

1.2 Continuity in R2

We now discuss continuity in the plane. Recall that, in R, continuity can beexpressed in terms of limits. Thus, a function f : (a, b)→ R is continuous atc ∈ (a, b) if, and only if,

f(c) = limx→c

f(x).

This condition is equivalent to the existence of side limits, which must beequal, i.e.

f(c) = limx→c+

f(x) = limx→c−

f(x).

Thus, continuity in one dimension is a very simple matter: it suffices to checklimits from the left and from the right.

In the plane however, one can approach a point p in all sorts of ways:parallel to the x-axis or the y-axis; along arbitrary straight lines; or evenmore “exotic” curves. For the limit to exist in the plane as x→ p, all suchlimits along all possible paths must exist and be equal.

Example 1.3. Consider the function f : R2 → R2 give by

f(x, y) :=

x2yx4+y2

, if (x, y) 6= (0, 0);

0, if (x, y) = (0, 0).

Along the x-axis, γ(t) = (t, 0), we have

f(γ(t)) = 0 for all t ∈ R.

4

Page 6: An Invitation to Complex Analysis - Warwick · 2019. 12. 13. · The geometric intuitions sketched above will be familiar from analysis in R. In one dimension. we use the absolute

Similarly, f(x, y) → (0, 0) as (x, y) → (0, 0) along the y-axis γ(t) = (0, t).In fact, for any “direction” (u, v) ∈ R2, the limit along the straight lineγ(t) = (tu, tv) is also zero, as

f(γ(t)) = f(tu, tv) =(t2u2)(tv)

t4v4 + t2v2=

tu2v

t2u4 + v2→ 0.

However, if we approach the origin along the parabola γ(t) = (t, t2), we have

f(γ(t)) = f(t, t2) =t2t2

t4 + t4=

t4

2t4→ 1

2.

Therefore, the limit of f(x, y) as (x, y)→ (0, 0) does not exist.

2 Topology in the Plane

In order to study the behaviour of functions on the complex plane, we needto become acquainted with its “geography”: in order to take limits we needto known what points can be considered “close” to one another, or if takinglimits in a set keeps the limit “inside” the same set, or whether it may escapeto the “boundary”. We want to known what sets hang together of a piece,and which are “disconnected” into separate pieces, since functions which areisolated to a single “piece” may have different behaviours on different pieces.This mathematical “cartography” is called topology. We refer the interestedreader to [3].

The geometric intuitions sketched above will be familiar from analysisin R. In one dimension. we use the absolute value function to define a“distance” between points on the line:

d(x, y) := |x− y|.

This takes care of “closeness” in terms of open intervals, (a− ε, a+ ε), whichlocalise the points that are within ε of a. A further useful property of openintervals is that they allow for “wiggle” room: given any point x ∈ (a, b), wecan always find a small enough δ > 0 such that (x− δ, x + δ) ⊂ (a, b). Thisproperty is very useful when studying limits and continuous functions, sincewe don’t need to worry about “falling out” of the set where the function isdefined.

Notice that closed intervals, [a, b], fail to allow for wiggle room at theend points, where moving to the left of a or the right of b will lead us outof the interval we want to be in. In this sense, the points a and b form the“boundary” of the interval, delimiting the “inside” from the “outside” of theset.

5

Page 7: An Invitation to Complex Analysis - Warwick · 2019. 12. 13. · The geometric intuitions sketched above will be familiar from analysis in R. In one dimension. we use the absolute

Figure 1: An open set in C. Every point has a small “wiggle ball” aroundit. The dashed line indicates that the boundary is not included.

It should be recalled that the presence of the end points makes closedintervals suitable for taking limits: if (xn)n is a convergent sequence of pointsin [a, b], then lim xn will also be a point of [a, b]. Open intervals, however,fail to have this property precisely because limits may “escape to the endpoints”, as is the case of ( 1

n+1)n ⊂ (0, 1) converging to 0 /∈ (0, 1).

The goal of this section is to formalise and generalise these familiar con-cepts so that we can translate them to the complex plane. Subsequent sec-tions will rely on many of these notions, especially when it comes to complexintegration.

2.1 Open and Closed Sets

Just as in R, the complex modulus function allows us to define a distancebetween points:

d(z, w) := |z − w|.

However, in two dimensions the equation |z − w| = r does not define aninterval around z of length 2r (r on either side); instead, we have a disk orball. The concept of a ball is fundamental, so we introduce the followingdefinitions:

Definition 2.1. The open ball of radius r > 0 centred at a ∈ C is definedto be the set

B(a, r) := z ∈ C : |z − a| < r.

The closed ball of radius r > 0 centred at a ∈ C is

B(a, r) := z ∈ C : |z − a| 6 r.

Balls provide a natural way to consider “wiggle room”. (See Figure 1.)

Definition 2.2. Let G ⊂ C. A point a ∈ G is called interior to G if thereexists an ε > 0 such that B(a, ε) ⊂ G. The set G is open if every point of Gis an interior point.

6

Page 8: An Invitation to Complex Analysis - Warwick · 2019. 12. 13. · The geometric intuitions sketched above will be familiar from analysis in R. In one dimension. we use the absolute

z

Figure 2: Any ball around the point z on the boundary intersects both the“interior” and the “exterior” of the ball (shaded red).

Notice that open balls are indeed open sets. Closed balls, however, arenot open: picking a point on the boundary circle |z − a| = r, any open ballaround z will intersect the complement of the ball. (See Figure 2.)

Balls also give a topological characterisation of limits: a sequence (zn)nconverges to z if for each ε > 0 there is an N such that zn ∈ B(z, ε) for alln > N . In other words, any ball around z contains infinitely many terms ofthe sequence.

Definition 2.3. Let F ⊂ C. A point z ∈ C is a limit point of F if everyball around z contains points of F . The set F is closed if it contains all ofits limit points.

One again, note that closed balls are closed sets according to this defini-tion, but open balls are not closed : any ball around a point on the boundarycircle intersects the inside of the circle, making these points limit points. (Seeagain Figure 2.)

We have used the term “boundary” informally, but it has a simple formaldefinition which clearly captures the geometric meaning we are used to:

Definition 2.4. The boundary of a set U ⊂ C is the collection of pointsz ∈ C such that all open balls about z meet U and its complement. Wedenote this set by ∂U .

The reader may be tempted to think that: “a set is open if and only if it isnot closed”, but this is certainly not true given the definitions above. Indeed,even in R, an open may be neither open no closed, as in the case of (a, b].We invite the reader to construct a similar example in in C. Perhaps morestrikingly, the set C is both open and closed according to the definitions.1

The “true” relationship between open and closed sets is the following:

Proposition 2.5. A set F ⊂ C is closed if, and only if, its complement C\Fis open.

1For the pedantically inclined, the empty set is also both open and closed.

7

Page 9: An Invitation to Complex Analysis - Warwick · 2019. 12. 13. · The geometric intuitions sketched above will be familiar from analysis in R. In one dimension. we use the absolute

Proof. Let F be a closed. We want to show that any point z ∈ C \ F isinterior to C \ F . Suppose it was not, i.e. there is no ε > 0 such thatB(z, ε) ⊂ C \ F . Then all balls around z meet F and, therefore, z is a limitpoint of F . But then z ∈ F , as F is closed, contradicting z ∈ C \ F .

Conversely, suppose C \ F is open. We want to show that every limitpoint of F lies in F . So suppose z is a limit point, i.e. B(z, ε) ∩ F 6= ∅ forall ε > 0. Since C \F is open and there is not ball around z lying entirely inC \ F , it follows that z ∈ F .

2.2 Continuity in CFunctions f : C→ C can also be continuous, and the definition is identical tothe real definition, the “only” change being the use of the complex modulusinstead of the absolute value: a function f : U → C on an open set U ⊂ C iscontinuous at a ∈ U if for all ε > 0 there exists a δ > 0 such that

|z − a| < δ =⇒ |f(z)− f(w)| < ε.

We chose the domain U to be open to avoid worrying about whether somez with |z− a| < δ may not be in U , where f is defined. The condition aboveis easily rephrased using open balls, and we adopt this formulation as ourdefinition.

Definition 2.6. A function f : U → C on an open set U ⊂ C is continuousat a ∈ U if for all ε > 0 there exists a δ > 0 such that f(B(a, δ)) ⊂ B(f(a), ε).

That is, all points in the ball B(a, δ) around a are mapped into the ballB(f(a), ε) around f(a). Thus, continuous functions “keep points together”.

The reader is reminded that continuity in the plane is more subtle than onthe line. We refer back the the discussion in §1.2, and in particular Example1.3.

For many purposes, the following, more topological, characterisation ofcontinuity is useful.

Proposition 2.7. A function f : U → C is continuous if, and only if, thepreimage of every open set under f is open in U .

The reader may wonder “why the pre-image?” A naive guess might bethat continuous functions map open sets to open sets, but this is certainlynot the case. All constant maps take open sets to a single point, which isclearly not open. The reason behind the formulation with pre-images shouldbecome clearer during the proof.

8

Page 10: An Invitation to Complex Analysis - Warwick · 2019. 12. 13. · The geometric intuitions sketched above will be familiar from analysis in R. In one dimension. we use the absolute

Proof. We want to show that for every open set G ⊂ C the set f−1(G) isopen in U .

Suppose f is continuous and let G ⊂ C be open, and pick any pointx ∈ f−1(G). We want to show that there is a ball B(x, δ) ⊂ f−1(G). Indeed,as G is open and f(x) ∈ G, there is a ball B(f(x), ε) ⊂ G for some ε > 0.By continuity of f , there is a δ > 0 such that |f(x) − f(y)| < ε whenever|x − y| < δ. In other words, f(y) ∈ B(f(x), ε) for all y ∈ B(x, δ), soB(x, δ) ⊂ f−1(G). This shows that f−1(G) is open.

For the converse, pick x ∈ U and ε > 0. Then B(f(x), ε) is an open set,so

f−1(B(f(x), ε)) = y ∈ U : |f(x)− f(y)| < ε

is open in U . Hence, there is a δ > 0 such that B(x, δ) ⊂ f−1(B(f(x), ε)),since x ∈ f−1(B(f(x), ε)). Thus,

|x− y| < δ =⇒ y ∈ B(x, δ) =⇒ f(y) ∈ B(f(x), ε) =⇒ |f(x)− f(y)| < ε,

which establishes continuity of f .

As an example of the power of the topological characterisation, we provethe following:

Proposition 2.8. The composition of two continuous functions is a contin-uous function.

Proof. Let f, g : C→ C be continuous functions, so g f : C→ C. Pick anyopen set G ⊂ C. Since g is continuous, g−1(G) is open. Hence, using that fis continuous, f−1(g−1(G)) = (g f)−1(G) is open.

2.3 Connectedness and Compactness

Continuous functions in R satisfy two important properties:

Intermediate Value Theorem If f : (a, b) → R is continuous, then itattains all values between f(a) and f(b).

Extreme Value Theorem If f : [a, b]→ R is continuous, then it is boundedand attains its bounds.

The crucial assumption behind the IVT is that the domain of f consistsof a single “piece”.

9

Page 11: An Invitation to Complex Analysis - Warwick · 2019. 12. 13. · The geometric intuitions sketched above will be familiar from analysis in R. In one dimension. we use the absolute

Figure 3: A disconnected subset of the plane (green), covered by disjointopen sets (red).

Example 2.9. Let f be defined on (−2,−1) ∪ (1, 2) as follows:

f(x) :=

0, if x ∈ (−2,−1);

1, if x ∈ (1, 2).

This function is continuous according to the ε− δ definition, but it does notsatisfy the intermediate value property, since f never achieves 1/2.

The property of a set being “a single piece” or “not breaking up” is calledconnectedness.

Definition 2.10. A set R ⊂ C is connected if there are no non-empty,disjoint open sets U, V ⊂ C such that R ⊂ U ∪ V .

In other words, a set is disconnected if we can cover it by disjoint opensets which “isolate its pieces”. (See Figure 3.)

It should come as no surprise that continuous functions preserve connect-edness, as they “keep points together”.

Proposition 2.11. The continuous image of a connected set is connected.

Proof. Let f : U → C be a continuous map and suppose f(U) is not con-nected. Hence, there exist open sets A,B ⊂ C such that A ∩ B = ∅ andf(U) = A ∪ B. But then U = f−1(A) ∪ f−1(B) with f−1(A) and f−1(B)disjoint and open (by continuity of f). Therefore, U is not connected.

It can be shown that the only connected subsets of R are intervals, i.e.sets I ⊂ R such that whenever x, y ∈ I and x 6 z 6 y, then z ∈ I. Hence,the special case of the above proposition in R reads: continuous functionsmap intervals to intervals, which is precisely the content of the IVT.

There is another “intuitive” notion of what it means for a set to be con-nected: any two points can be “joined together” by a path without leavingthe set.

10

Page 12: An Invitation to Complex Analysis - Warwick · 2019. 12. 13. · The geometric intuitions sketched above will be familiar from analysis in R. In one dimension. we use the absolute

Definition 2.12. A path is a continuous map γ : [0, 1] → C. A set U ⊂ Cis path connected if any two points a, b ∈ U can be joined by a path in U .That is, there exists a path γ : [0, 1] → C such that γ(0) = a, γ(1) = b andγ([0, 1]) ⊂ U .

A special case of path connected sets are convex sets. These are the setswhich contain all straight lines between their points. The “official definition”is

Definition 2.13. A set C ⊂ C is convex if for all a, b ∈ C and any t ∈ (0, 1),we have (1− t)a+ tb ∈ C.

The paths in this case are the straight lines t 7→ (1− t)a+ tb. Clearly

Proposition 2.14. Balls (open or closed) are convex, and therefore pathconnected.

Since we will often consider functions defined in open connected sets, weintroduce the following

Definition 2.15. An open and connected subset of C is called a domain.

It can be shown that an open subset of C (or indeed Rn) is connected if,and only if, it is also path connected, although we shall not prove this here.Suffice it to show that both definitions are convenient for different purposes,and that there is no risk of confusion by using them interchangeably fordomains.

We now turn to the EVT. The reader will recall from real analysis thatthe conditions that the interval be bounded and closed are essential to thetheorem. A continuous function on an unbounded interval may of course beunbounded (a trivial example is f(x) = x on (0,∞)).

Example 2.16. The function f : (0, 1)→ R given by f(x) := 1x

is continuouson the bounded interval (0, 1), but is clearly unbounded as x→ 0+.

The generalisation to C involves closed and bounded subsets of the plane.But in topology it is useful to work with the even more general notion ofcompactness.

Definition 2.17. A set K ⊂ C is called compact if whenever Gjj∈J areopen sets such that K ⊂ ∪jGj, there exist finitely many Gj1 , . . . Gjn suchthat K ⊂ Gj1 ∪ · · · ∪ Gjn . We call a collection Gjj∈J as above an opencover of K.

11

Page 13: An Invitation to Complex Analysis - Warwick · 2019. 12. 13. · The geometric intuitions sketched above will be familiar from analysis in R. In one dimension. we use the absolute

It is fairly easy to see that compact sets are bounded: cover K by theballs B(x, 1), x ∈ K. Then K is contained in a union of finitely many ballsof radius 1, and therefore is bounded. We invite the reader to show thatcompact sets are also closed. (Proposition 2.5 might be helpful.)

In C (and more generally, Rn), the compact sets are precisely the closedand bounded sets, a fact known as the Heine-Borel Theorem.

Theorem 2.18 (Heine-Borel). A subset of C is compact if, and only if, it isclosed and bounded.

Will we not prove this theorem here.2 But we do remark that this Theo-rem only holds in Rn, and not in more general metric or topological spaces.In particular, we have that closed and bounded intervals [a, b] are compactin R, and closed balls B(z, r) are compact in C.

Using Proposition 2.7, it is easy to show that

Proposition 2.19. If f : U → C is continuous and K ⊂ U is compact, thenf(K) is compact

Proof. Pick any cover Gjj∈J of f(K). Since the Gj are open and f is con-tinuous, Proposition 2.7 implies that the f−1(Gj) are all open. Furthermore,

f(K) ⊂⋃j∈J

Gj ⇐⇒ K ⊂⋃j∈J

f−1(Gj),

so the compactness of K means that we can pick finitely many j1, . . . , jn suchthat

K ⊂n⋃k=1

f−1(Gjk).

Therefore, applying the same reasoning as above,

f(K) ⊂n⋃k=1

Gjk .

From the Heine Borel Theorem, we can deduce that the continuous imageof a closed bounded set in the plane is again closed and bounded. This factwill be useful in §5.1.

Example 2.20. The function defined by f(z) = 1z

is continuous on the

bounded set B(0, 1) \ 0, but it is certainly not bounded! This “failure” isdue to the fact that the punctured disk is not compact, since it is not closed:the origin 0 ∈ C is clearly a limit point of B(0, 1) \ 0. (Compare withExample 2.16.)

2A proof may be found in any text on general topology such as [3] , or in [1].

12

Page 14: An Invitation to Complex Analysis - Warwick · 2019. 12. 13. · The geometric intuitions sketched above will be familiar from analysis in R. In one dimension. we use the absolute

We note that the same argument works for functions f : K → R. Inparticular, f(K) is a closed and bounded subset of R, and hence containedin a bounded interval. We can thus give a simple proof of the Extreme ValueTheorem:

Theorem 2.21 (Extreme Value Theorem). Let f : K → C be a continuousfunction on a compact set K ⊂ C. Then |f | is bounded and attains itsbounds.

Proof. By Proposition 2.8, |f | : K → R is continuous as f and | · | are contin-uous. By Proposition 2.19, |f |(K) is compact, hence closed and bounded. ByProposition 2.11, |f |(K) is connected. Therefore, |f |(K) is a closed boundedinterval [a, b] ⊂ R. In particular, a = min |f | and b = max |f | are achievedat some points in K, for otherwise a, b /∈ |f |(K).

3 Differentiability in CIn §2.2, we saw how the definition of continuity carried over seamlessly to C.We will see how, formally, the definition of a derivative of a complex functionis identical. Yet looks are deceiving, for a complex derivative is a far strangerobject than a real one, as we will have ample opportunity to explore in theremainder of these notes.

3.1 Differentiability

Limits in the complex plane are defined using the complex modulus. Theusual shorthand may be used to convey the definition

limz→a

f(z) = L ⇐⇒ ∀ε > 0∃δ > 0∀z ∈ U(|z− a| < δ =⇒ |f(z)− f(a)| < ε),

where f : U → C is a complex function defined on a domain U . (SeeDefinition 2.15.) We also remind the reader that limits in the complex planeare required to exists and be equal in all directions of approach, as in Example1.3.

With these preliminaries out of the way, it is routine to define a complexderivative.

Definition 3.1. A complex function f : U → C on a domain U is differen-tiable at a ∈ U if the following (complex) limit exists and is finite:

limz→a

f(z)− f(a)

z − a

13

Page 15: An Invitation to Complex Analysis - Warwick · 2019. 12. 13. · The geometric intuitions sketched above will be familiar from analysis in R. In one dimension. we use the absolute

In that case we denote the limit by f ′(a), the derivative of f at a. If f : U →C is differentiable at every point of U (open), we shall call f a holomorphicfunction on U . In particular, a holomorphic function on C is called entire.

It is routine to check the complex differentiability implies continuity, andthat the usual algebraic properties of derivatives carry over to complex deriva-tives. (The proofs, being identical to the real case, are omitted.)

Theorem 3.2. Let f, g : C be differentiable functions and α, β ∈ C. Then,

(i) f and g are continuous;

(ii) αf+βg, fg and (where g(z) 6= 0) f/g are differentiable with derivatives

(αf + βg)′ = αf ′ + βg′, (fg)′ = f ′g + fg′,

(f

g

)′=f ′g − fg′

g2;

(iii) Chain Rule: f g is differentiable with (f g)′ = (f ′ g)g′.

Example 3.3. The function f : C→ C defined as f(z) = z2 is differentiableat every z ∈ C with f ′(z) = 2z:

(z + h)2 − z2

h=

2zh+ h2

h= 2z + h→ 2z as h→ 0.

Example 3.4. The function f : C → C defined as f(z) = z is nowheredifferentiable on C. Consider h ∈ R. Then, along the real axis, we have

(z + h)− zh

=h

h≡ 1.

But along the imaginary axis, with ih, we have

(z + ih)− zih

=−ihih≡ −1.

Since the “partial limits” do not agree, the limit does not exists.

This simple example gives us some indication of how complex differen-tiability differs from real differentiability. In real analysis, it is very compli-cated to construct a function which is everywhere continuous but nowheredifferentiable. In the complex case, we have a very simple example of thisphenomenon in complex conjugation.

Some authors call holomorphic functions “analytic”. We reserve this termfor a more specific use:

14

Page 16: An Invitation to Complex Analysis - Warwick · 2019. 12. 13. · The geometric intuitions sketched above will be familiar from analysis in R. In one dimension. we use the absolute

Definition 3.5. A holomorphic f : U → C is analytic if it has complexderivatives of all orders and its Taylor series converges to f on U .

Note that, in the real setting, having derivatives of all orders is not enoughfor analyticity.

Example 3.6. Let f : R→ R be defined as follows:

f(x) :=

e−1/x2 if x 6= 0;

0 if x = 0.

Then f is infinitely differentiable, but its Taylor series does not converge tof(0) = 0 at z = 0. So f is not analytic.

So far we only known that a holomorphic function is continuous, butknow nothing about its derivative, which may not even be continuous. Wewill examine the properties of holomorphic functions and their derivatives ingreater detail with the help of complex integrals in §4.

3.2 The Cauchy-Riemann Equations

Before progressing to complex integration, we derive an important charac-terisation of holomorphic functions.

Suppose f : U → C is holomorphic on U and pick a ∈ U . Recall that wecan write f(x+ iy) = u(x, y) + iv(x, y), where u, v : R2 → R.

We now turn to the derivative of f at a, which it is convenient to writein the form

f ′(a) = limh→0

f(a+ h)− f(a)

h.

We known that the limit must be the same in all directions. If we take h→ 0along the x-axis:

f ′(a) = limh→0

f(a+ h)− f(a)

h

= limh→0

u(<a+ h,=a)− u(<a,=a)

h+ i lim

h→0

v(<a+ h,=a)− v(<a,=a)

h

=∂u

∂x(<a,=a) + i

∂v

∂x(<a,=a).

A similar calculation taking ih→ 0 (h ∈ R) along the y-axis shows

f ′(a) =∂v

∂y(<a,=a)− i∂u

∂y(<a,=a).

15

Page 17: An Invitation to Complex Analysis - Warwick · 2019. 12. 13. · The geometric intuitions sketched above will be familiar from analysis in R. In one dimension. we use the absolute

Equating real and imaginary terms in both calculations yields the Cauchy-Riemann Equations :

∂u

∂x=∂v

∂yand

∂u

∂y= −∂v

∂x.

We have shown that holomorphic functions satisfy the Cauchy-RiemannEquations. The converse, however, is not true.

Example 3.7. Let f(z) := e−1/z4 if z 6= 0 and f(0) := 0. It is easy to showthat f satisfies the Cauchy-Riemann Equations at 0. But f is not continuousat 0, hence not differentiable.

The precise theorem is as follows.

Theorem 3.8. Let f : U → C be a function given by f(x + iy) ≡ u(x, y) +iv(x, y) on a domain U ⊂ C. Then, f is complex differentiable at a ∈ U if,and only if, the function (x, y) 7→ (u(x, y), v(x, y)) is real differentiable at aand satisfies the Cauchy-Riemann Equations.

We omit the proof, as we have not developed real-differentiability in R2

in sufficient detail; it can be found in [1].

4 Complex Integration

The complex theory of integration plays a central rle in complex analy-sis. Many of the more subtle and interesting results from the discipline aremost easily proved using integration. In this section we explore two key re-sults: Cauchy’s Integral Theorem (§4.2) and one of its richest consequences,Cauchy’s Integral Formula (§4.3). Both results will shed light on the rich-ness of complex differentiation, setting it well apart from real differentiation(§4.4). We will also see how they can be applied to more specific problemsin §5.

4.1 Line and Arc-Length Integrals

Complex line integrals can be defined in much the same way as real lineintegrals. (C.f. §1.)

Let γ : [a, b] → C be a differentiable path in the plane and f : U → C acomplex function such that γ([a, b]) ⊂ U . The line integral of f along γ isdefined to be ∫

γ

f(z) dz :=

∫ b

a

f(γ(t))γ′(t) dt.

16

Page 18: An Invitation to Complex Analysis - Warwick · 2019. 12. 13. · The geometric intuitions sketched above will be familiar from analysis in R. In one dimension. we use the absolute

We will often integrate along the boundary of a bounded domain Ω inthe plane, or around a circle |z − a| = r, in which cases the more explicitnotation ∮

∂Ω

f(z) dz or

∮|z−a|=r

f(z) dz

will be used. It is understood that all curves are traversed anticlockwise,and it is easy to show, by a change of variable, that a change of orientationimplies a change of sign. Thus, if −γ denotes the curve γ traversed in theopposite orientation, we have∫

−γf(z) dz = −

∫γ

f(z) dz.

Circles |z − a| = r are most easily parametrised by t → a + reit with0 6 t 6 2π.

Example 4.1. Let f(z) := z on C. Then∮|r−a|=r

f(z) dz =

∫ 2π

0

(a+ reit)(ireit) dt = 0.

This result should come as no surprise, since z is the complex derivativeof z2

2. (Compare with Theorem 1.1.)

Example 4.2. Let f(z) = 1z−a . Then∮

|z−a|=r

1

z − adz =

∫ 2π

0

ireit

reitdt = 2πi.

Another useful integral is the arc-length integral:∫γ

f(z) |dz| :=∫ b

a

f(γ(t))|γ′(t)| dt.

This integral should be familiar from multivaribale calculus. In particular,we have ∫

γ

|dz| = length(γ).

We will often use this integral in conjunction with the following lemma.

17

Page 19: An Invitation to Complex Analysis - Warwick · 2019. 12. 13. · The geometric intuitions sketched above will be familiar from analysis in R. In one dimension. we use the absolute

Lemma 4.3 (Estimation Lemma). Let f : U → C be a complex functionand γ a differentiable path. Then∣∣∣∣∫

γ

f(z) dz

∣∣∣∣ 6 ∫γ

|f(z)| |dz|.

Proof. We first establish the estimate∣∣∣∣∫ b

a

ϕ(t) dt

∣∣∣∣ 6 ∫ b

a

|ϕ(t)| dt. (∗)

Note that, by linearity of the integral, we have∫ b

a

e−iθϕ(t) dt = e−iθ∫ b

a

ϕ(t) dt.

for all θ ∈ [0, 2π). Hence

<[e−iθ

∫ b

a

ϕ(t) dt

]=

∫ b

a

<[e−iθϕ(t)] dt 6∫ b

a

|ϕ(t)| dt.

Setting θ = arg[∫ b

aϕ(t) dt

],

<[e−iθ

∫ b

a

ϕ(t) dt

]=

∣∣∣∣∫ b

a

ϕ(t) dt

∣∣∣∣which proves (∗). The lemma follows by taking ϕ(t) = f(γ(t))γ′(t).

4.2 Cauchy’s Integral Theorem

This section introduces a fundamental result of complex analysis, with far-reaching consequences:

Theorem 4.4 (Cauchy’s Integral Theorem). Let f : U → C be holomorphicon a domain U ⊂ C. Then, for any closed curve γ in U that can be “shrunkto a point in U”, we have: ∮

γ

f(z) dz = 0.

The assumption that γ may be “shrunk to a point” can be given a rigorousmeaning. Geometrically, it is quite clear (see Figure 4), and the main appli-cation of this theorem in §4.3 will also be obvious. It is a crucial condition,as the following example shows.

18

Page 20: An Invitation to Complex Analysis - Warwick · 2019. 12. 13. · The geometric intuitions sketched above will be familiar from analysis in R. In one dimension. we use the absolute

· γ ζ

Figure 4: The curve γ (right) can be shrunk to a point in the green set. Thecurve ζ (right) cannot be shrunk to a point in the red set as it wraps aroundthe “hole”.

Example 4.5. The function f(z) = 1z

is holomorphic on B(0, 1) \ 0, but∮|z|=1

1

zdz = 2πi.

Cauchy’s Theorem does not apply to the boundary circle |z| = 1 since itcannot be “shrunk to a point” in B(0, 1) \ 0 as it encircles the “puncture”at the origin. (Compare this to the curve ζ in Figure 4.)

Cauchy’s Integral Theorem is analogous to Theorem 1.2. Just as in R2,if f happens to be the derivative of a function, then the result easily followsfrom the Fundamental Theorem of Calculus (or Theorem 1.1). But we areonly assuming that f has a derivative, without additional assumptions suchas that f has an antiderivative.

We will prove Cauchy’s Theorem for the special case of a ball. The moregeneral proof is a little more complex, and we will omit it; it can be foundin Chapter 4 of [1].3 Note that it is this more general version which we shallneed to prove Theorem 4.8 below.

We begin with an even simpler case: a triangle. This proof is due toGoursat.

Theorem 4.6 (Goursat’s Theorem). Let f : U → C be holomorphic on adomain U . Then, for every triangle ∆ ⊂ U (with ∂∆ ⊂ U too), we have∮

∂∆

f(z) dz = 0.

Proof. We Divide the triangle δ into four smaller triangles ∆′1, . . .∆′4 as per

Figure 5. Then ∮∂∆

f(z) dz =4∑j=1

∮∂∆′j

f(z) dz,

3If the reader is familiar with Green’s Theorem, we invite them to prove the generalversion of Cauchy’s Integral Theorem using the Cauchy Riemann Equations from §3.2.

19

Page 21: An Invitation to Complex Analysis - Warwick · 2019. 12. 13. · The geometric intuitions sketched above will be familiar from analysis in R. In one dimension. we use the absolute

∆∆′1

∆′3 ∆′4

∆′2

Figure 5: Triangle decomposition used in the proof of Goursat’s Theorem.

as the integrals over the edges inside of ∆ cancel out. By the triangle in-equality. ∣∣∣∣∮

∂∆

f(z) dz

∣∣∣∣ 6 4∑j=1

∣∣∣∣∣∮∂∆′j

f(z) dz

∣∣∣∣∣ .Let ∆1 be the ∆′j with largest integral in the sum above, so that∣∣∣∣∮

∂∆

f(z) dz

∣∣∣∣ 6 4

∣∣∣∣∮∂∆1

f(z) dz

∣∣∣∣ .We proceed by induction, obtaining a sequence (∆n)n of triangles satisfying

(i) ∆ ⊃ ∆1 ⊃ · · · ⊃ ∆n · · · ;

(ii) diag(∆n) 6 2−n diag(∆);

(iii) length(∂∆n) 6 2−n length(∂∆);

for which we have the inequalities∣∣∣∣∮∂∆

f(z) dz

∣∣∣∣ 6 4n∣∣∣∣∮∂∆n

f(z) dz

∣∣∣∣ . (∗)

Picking arbitrary points zn ∈ ∆n, the sequence (zn)n is Cauchy, henceconvergent to some z∗ ∈ ∆∪ ∂∆. In particular, z∗ ∈ U , so f is differentiableat z∗. Hence, for any ε > 0 there is a δ > 0 small enough that B(z∗, δ) ⊂ ∆and ∣∣∣∣f(z)− f(z∗)

z − z∗− f ′(z∗)

∣∣∣∣ < ε

for all 0 < |z − z∗| < δ. We can re-write this as

|f(z)− f(z∗)− (z − z∗)f ′(z∗)| < ε|z − z∗|.

20

Page 22: An Invitation to Complex Analysis - Warwick · 2019. 12. 13. · The geometric intuitions sketched above will be familiar from analysis in R. In one dimension. we use the absolute

By Theorem 1.1, we have that∮∂∆n

dz = 0 and

∮∂∆n

z dz = 0

since z 7→ 1 and z 7→ z have complex antiderivatives everywhere. Therefore,∮∂∆n

f(z) dz =

∮∂∆n

[f(z)− f(z∗)− (z − z∗)f ′(z∗)] dz,

and the Estimation Lemma implies∣∣∣∣∮∂∆n

f(z) dz

∣∣∣∣ 6 ε

∮∂∆n

|z − z∗| |dz|

6 ε diag(∆n) length(∂∆n)

6 ε(2−n diag(∆))(2−n length(∂∆))

= ε4−n diag(∆) length(∂∆).

Thus, (∗) implies ∣∣∣∣∮∂∆

f(z) dz

∣∣∣∣ 6 ε diag(∆) length(∂∆).

Since ε > 0 is arbitrary and diag(∆) length(∂∆) > 0 is a constant indepen-dent of ε, it follows that

∮∂∆f(z) dz = 0.

Goursat’s Theorem can be used to prove a more general statement forclosed curves in a disk. The proof fill’s in the gap left in Theorem 1.2, wherewe used an integral along a path from a fixed point p to x to define anantiderivative of F .

Theorem 4.7 (Cacuchy’s Integral Theorem for Balls). Let f : U → C beholomorphic on a domain U ⊂ C and B(a, r) ⊂ U . Then, for any closedcurve γ in B(a, r), we have ∮

γ

f(z) dz = 0.

Proof. We show that f has an antiderivative in B(a, r). Then, the resultfollows from Theorem 1.1.

Since balls are convex, let λz denote the straight line from a to z ∈ B(a, r).Define the function F : B(a, r)→ C as follows:

F (z) :=

∫λz

f(ξ) dξ.

21

Page 23: An Invitation to Complex Analysis - Warwick · 2019. 12. 13. · The geometric intuitions sketched above will be familiar from analysis in R. In one dimension. we use the absolute

a

z0

z

λz0

βλz

a

z0

z

Figure 6: The triangle 〈a, z0, z〉, oriented anticlockwise

We claim that F ′ = f .To see this, fix z0 ∈ B(a, r). Since open balls are open, there is a ρ > 0

such that B(z0, ρ) ⊂ B(a, r). Let β denote the straight line from z0 to z.Then, the three lines λz0 , λz and β form a triangle in B(a, r), since the ballis convex. (Figure 6.) Applying Goursat’s Theorem, we have∫

λz0

f(ξ) dξ +

∫β

f(ξ) dξ −∫λz

f(ξ) dξ = 0.

Hence,

F (z)− F (z0)

z − z0

=1

z − z0

(∫λz

f(ξ) dξ −∫λz0

f(ξ) dξ

)=

1

z − z0

∫β

f(ξ) dξ.

Writing down β(t) = (1− t)z0 + tz explicitly, we have

1

z − z0

∫β

f(ξ) dξ =1

z − z0

∫ 1

0

f((1−t)z0+tz)(z−z0) dt =

∫ 1

0

f((1−t)z0+tz) dt.

As z → z0, we can use the continuity of to f exchange the limit and theintegral to conclude that the right hand side is f(z0).

4.3 Cauchy’s Integral Formula

We are now ready to prove one of the most remarkable and beautiful resultsof complex analysis. Its power we become evident in the consequences wewill derive from it.

Theorem 4.8 (Cauchy’s Integral Formula). Let f : U → C be holomorphicon a bounded domain U ⊂ C and B(a, r) ⊂ U . Then

f(a) =1

2πi

∮∂B(a,r)

f(z)

z − adz.

22

Page 24: An Invitation to Complex Analysis - Warwick · 2019. 12. 13. · The geometric intuitions sketched above will be familiar from analysis in R. In one dimension. we use the absolute

∂B(a, r)

∂B(a, δ)

La

γ1

a=

γ2

Figure 7: The contour used in the proof of Theorem 4.8.

Proof. Consider the function F : B(a, r) \ a → C defined by

F (z) =f(z)− f(a)

z − a.

Note that F is holomorphic. To apply Cauchy’s integral theorem to F , wewill “cut off” a smaller ball B(a, δ) ⊂ B(a, r) at a and “integrate around it”,as shown in Figure 7.

Consider the paths γ1 and γ2 shown in Figure 7. Then∮|z−a|=r

F (z) dz −∮

|z−a|=δ

F (z) dz =

∮γ1

F (z) dz +

∮γ2

F (z) dz

where we integrate anticlockwise around |z − a| = r and clockwise around|z−a| = δ. Hence, the integrals along the line segments L cancel out as theyare traversed once in each direction. Since the function F is holomorphicinside the domains bounded by γ1 and γ2, and these curves can be shrunkto a point in B(a, r) \ a, it follows from Cauchy’s Integral Theorem that∮γjF (z) dz = 0. Therefore,∮

|z−a|=r

F (z) dz =

∮|z−a|=δ

F (z) dz.

We now use the continuity of f to estimate the integral on the right handside. Fix ε > 0. Then there exists a δ > 0 such that

|z − a| < δ =⇒ |f(z)− f(a)| < ε/2π.

23

Page 25: An Invitation to Complex Analysis - Warwick · 2019. 12. 13. · The geometric intuitions sketched above will be familiar from analysis in R. In one dimension. we use the absolute

(We require this δ to be small enough so that B(a, δ) ⊂ B(a, r), where f isholomorphic!) Then, using the Estimation Lemma,∣∣∣∣∮∂B(a,δ)

F (z) dz

∣∣∣∣ 6 ∮|z−a|=δ

|f(z)− f(a)||z − a|

|dz| 6∮

|z−a|=δ

ε/2π

δ|dz| = ε

2πδ(2πδ) = ε.

We have therefore proved that∫|z−a|=r

F (z) dz = limδ→0+

∫|z−a|=δ

F (z) dz = 0.

Expanding the integral of F (z) we see that∮∂B(a,r)

f(z)− f(a)

z − adz =

∮∂B(a,r)

f(z)

z − adz − f(a)

∫∂B(a,r)

1

z − adz

=

∮∂B(a,r)

f(z)

z − adz − 2πif(a),

and the result follows from noting that the left hand side is zero.

In fact, the integral formula can be applied to general curves encirclinga, not just circles. We will use this more general version in §5.2.

This theorem is a remarkable departure from real analysis: we can de-termine every value of a complex-differentiable function on a domain byconsidering the integral only along its boundary. But it has a much morepowerful consequence.

4.4 Complex Analytic Functions

Let us rename our variables and write

f(z) =1

2πi

∫∂Ω

f(ξ)

ξ − zdξ.

Since z 6= ξ along ∂Ω, we can take the derivative of the integrand withrespect to z for each fixed z ∈ Ω. Hence, differentiating under the integralsign:

f ′(z) =1

2πi

∫∂Ω

f(ξ)

(ξ − z)2dξ.

This shows that if f is holomorphic on Ω, then f ′ is also continuous, as theintegrand on the right hand side is continuous. In fact, it is also differentiable!Iterating differentiation by induction yields

f (n)(z) =n!

2πi

∫∂Ω

f(ξ)

(ξ − z)n+1dξ.

24

Page 26: An Invitation to Complex Analysis - Warwick · 2019. 12. 13. · The geometric intuitions sketched above will be familiar from analysis in R. In one dimension. we use the absolute

for each n ∈ N. Hence, holomorphic functions are infinitely differentiable.This is most certainly not true of real differentiable functions: having one

derivative, even in an open and connected set, in no way guarantees even thecontinuity of the derivative, let alone being smooth! But we can show more.

Fix a point z0 ∈ Ω and write

1

ξ − z=

1

ξ − z0

ξ − z0

ξ − z=

1

ξ − z0

1ξ−zξ−z0

=1

ξ − z0

1

1− z−ξξ−z0

.

For |z − z0| < |ξ − z0|, the second fraction on he right hand side can beexpanded as a geometric series:

1

1− z−ξz0−ξ

=∞∑k=0

(z − z0

ξ − z0

)k

.

Plugging this development back into Cauchy’s Integral Formula yields

f(z) =1

2πi

∫∂Ω

f(ξ)

ξ − zdz

=1

2πi

∫∂Ω

f(ξ)

ξ − z0

∞∑k=0

(z − z0

ξ − z0

)kdz

=∞∑k=0

[1

2πi

∫∂Ω

f(ξ)

(ξ − z0)k+1dz

](z − z0)k

=∞∑k=0

f (k)

k!(z − z0)k,

where the exchange of the integral and the series can be justified by theuniform convergence in ξ of the series. Therefore, we have “proved”4

Theorem 4.9. A holomorphic function f : U → C on a domain U ⊂ C isanalytic on U .

Having proved this result, we will hereafter refer to holomorphic functionsas “analytic”, following common practise in the literature.

5 Applications

5.1 The Fundamental Theorem of Algebra

A remarkable consequence of Theorem 4.9 is

4We still have to justify the differentiation under the integral sign! This is done, forexample, in Lemma 3 at [1, p. 121].

25

Page 27: An Invitation to Complex Analysis - Warwick · 2019. 12. 13. · The geometric intuitions sketched above will be familiar from analysis in R. In one dimension. we use the absolute

Theorem 5.1 (Liouville’s Theorem). An entire bounded function is constant.

Proof. Recall that an entire function is a function which is analytic (holo-morphic) on all of C. Suppose that there is an M > 0 such that |f(z)| 6Mfor all z ∈ C. Using the derivative forms of Cauchy’s Integral Formula wehave

|f (n)(z)| 6 n!

∫|ξ−z|=r

|f(ξ)||ξ − z|n+1

|dξ| 6 n!M

rn.

For n > 1, the right hand side can be made arbitrarily small as r → ∞,and f remains analytic since it is entire. Hence, all coefficients of the Taylorseries of f vanish except for the the constant term.5

Once again, this result violates all of our expectations from real analysis,where even bounded analytic functions can be injective (consider tanh).

Example 5.2. The real restriction of sin is bounded by 1, as is well known;yet this function is analytic. Indeed, we defined it to be a power serieswhen we extended it to the complex plane. Since sin is entire and clearlynot constant, Liouville’s theorem implies that it must be unbounded. As acomplex function, we have

sin(z) =eiz − e−iz

2i

Choosing zn = −in, we have eizn = en →∞ as n→∞ while e−izn = e−n 6 1,so sin(zn) is unbounded.

Liouville’s Theorem also yields a simple and elegant proof of the Funda-mental Theorem of Algebra.

Theorem 5.3 (Fundamental Theorem of Algebra). Any non-constant poly-nomial with complex coefficients has at least one complex root.

Proof Sketch. Let p(z) = anzn + · · · + a0 be a polynomial of degree n > 1

(so an 6= 0) with complex coefficients ak ∈ C. Suppose that p(z) 6= 0 for allz ∈ C. Then, the function

g(z) =1

p(z)

is entire. Since |p(z)| → ∞ as |z| → ∞, there exists an R > 0 such that|p(z)| > |a0| for all |z| > R, and therefore |g(z)| 6 1/|a0| whenever |z| >

5Alternatively, notice that for n = 1 we have that |f ′(z)| vanishes for all z, so f ′ ≡ 0.Hence, as in the real case, f is constant.

26

Page 28: An Invitation to Complex Analysis - Warwick · 2019. 12. 13. · The geometric intuitions sketched above will be familiar from analysis in R. In one dimension. we use the absolute

R. On the other hand, |z| 6 R defines the closed ball B(0, R). Since gis continuous, Proposition 2.19 implies that g is also bounded on B(0, R).Hence, g is bounded on the entire plane. By Liouville’s theorem, g must beconstant. Hence p must be constant. But this contradicts the assumptionthat p is a polynomial of degree n > 1. Therefore, p must have at least oneroot in the complex plane.

The more usual form of the Fundamental Theorem of Algebra, whichgrantees the factorisation of p into n linear factors

∏nk=1(z − zk) where

z1, . . . , zn are the (possibly repeated) roots of p, follows by the remaindertheorem and induction on n.

5.2 Contour Integration

In this final section we will apply the results from §4 to compute a “hard”real integral.

We wish to calculate∞∫

−∞

cos(x)

x2 + 1dx.

Using Euler’s identity, we can re-write the integral as

∞∫−∞

cos(x)

x2 + 1dx = <

∞∫−∞

eiz

z2 + 1dz.

For convenience, we will set

f(z) :=eiz

z2 + 1=

eiz

(z − i)(z + i), g(z) :=

eiz

z + i.

Note that f is analytic on C \ ±i and g is analytic on C \ −i.We now consider the contour Γ as described in Figure 8. The point −i

lies outside of the region bounded by Γ and, for R > 1, i is inside this region.We can therefore apply Cauchy’s Integral Formula to the analytic functiong inside Γ: ∫

Γ

f(z) dz =

∫Γ

g(z)

z − idz = 2πig(i) =

2πie−1

2i=π

e.

We can now recover our initial integral by calculating the line integrals∫CR

f(z) dz and

∫ R

−Rf(z) dz,

27

Page 29: An Invitation to Complex Analysis - Warwick · 2019. 12. 13. · The geometric intuitions sketched above will be familiar from analysis in R. In one dimension. we use the absolute

CR

−R R

i

Figure 8: The (anticlockwise) contour Γ consists of two pieces: the straightline [−R,R] and the upper semicircle CR.

since∞∫

−∞

eiz

z2 + 1dz = lim

R→∞

∫ R

−Rf(z) dz = lim

R→∞

[∫Γ

f(z) dz −∫CR

f(z) dz

].

Ideally, we want to show that the integral along CR vanishes as R→∞,so that we are left with the finite integral over Γ, which we already computedto be π

e. Indeed, this is so, by an application of the Estimation Lemma.

First note that|eiz| = |eix||e−y| = |e−y| 6 1

since 0 6 =z 6 R on the upper semicircle CR. Furthermore, using the reversetriangle inequality6, we see that

|z2 + 1| > ||z|2 − 1| = R2 − 1

as |z| = R > 1 along CR. Putting these estimates together, we have∣∣∣∣∫CR

f(z) dz

∣∣∣∣ 6 ∫CR

|eiz||z2 + 1|

|dz| 6 πR

R2 − 1→ 0 as R→∞.

Therefore,∞∫

−∞

eiz

z2 + 1dz = lim

R→∞

[∫Γ

f(z) dz −∫CR

f(z) dz

]=π

e.

Since this integral is real, our initial integral evaluates to∞∫

−∞

cos(x)

x2 + 1dx =

π

e.

6I.e. |x− y| > ||x|− |y||. This is an easy consequence of the regular triangle inequality.

28

Page 30: An Invitation to Complex Analysis - Warwick · 2019. 12. 13. · The geometric intuitions sketched above will be familiar from analysis in R. In one dimension. we use the absolute

6 Bibliography

[1] Lars V. Ahlfors: Complex Analysis (Third Edition). McGraw-Hill. (1979)

[2] T. M. Apostol: Mathematical Analysis (Second Edition). Addison-WesleyPublishing Co. (1974)

[3] J. R. Munkres: Topology. Prentice-Hall. (2000)

[4] David Preiss: Lecture notes for the course “Metric Spaces” at the Uni-versity of Warwick. (2015)

[5] T. J. Sullivan: Lecture notes for the course “Differentiation” at the Uni-versity of Warwick.

[6] Hendrik Weber: Lecture notes for the course “Complex Analysis” at theUniversity of Warwick. (2018)

29