Visual Complex Analysis
Visual Complex Analysis
VISUAL
CO MPL EX
ANALYSIS
T R I S TA N N E E D H A M
UNIVERSIT Y OF SAN FRANCISCO
Foreword by
ROGER PENROSE
Great Clarendon Street, Oxford, OX2 6DP,
United Kingdom
Oxford University Press is a department of the University of Oxford.
It furthers the University’s objective of excellence in research, scholarship,
and education by publishing worldwide. Oxford is a registered trade mark of
Oxford University Press in the UK and in certain other countries
© Tristan Needham 2023
The moral rights of the author have been asserted
First edition published in 1997
25th Anniversary Edition 2023
All rights reserved. No part of this publication may be reproduced, stored in
a retrieval system, or transmitted, in any form or by any means, without the
prior permission in writing of Oxford University Press, or as expressly permitted
by law, by licence or under terms agreed with the appropriate reprographics
rights organization. Enquiries concerning reproduction outside the scope of the
above should be sent to the Rights Department, Oxford University Press, at the
address above
You must not circulate this work in any other form
and you must impose this same condition on any acquirer
Published in the United States of America by Oxford University Press
198 Madison Avenue, New York, NY 10016, United States of America
British Library Cataloguing in Publication Data
Data available
Library of Congress Control Number: 2022948019
ISBN 978–0–19–286891–6 (hbk.)
ISBN 978–0–19–286892–3 (pbk.)
DOI: 10.1093/oso/9780192868916.001.0001
Printed and bound by
CPI Group (UK) Ltd, Croydon, CR0 4YY
Links to third party websites are provided by Oxford in good faith and
for information only. Oxford disclaims any responsibility for the materials
contained in any third party website referenced in this work.
For
Roger Penrose
and
George Burnett-Stuart
FOREWORD
Roger Penrose
Complex analysis is the theory of functions of complex numbers or, more specif-
ically, holomorphic functions of such numbers. This theory is both profoundly
beautiful and vastly influential, both in pure mathematics and in many areas of
application, particularly in physics, indeed being central to the underlying formal-
ism of quantum mechanics. However, the very concept of a complex number is an
essentially abstract one, depending upon the seemingly absurd notion of a square
root of −1, the square of any ordinary real number being, unlike −1, necessarily
non-negative.
Yet, it should be borne in mind that even the notion of a so-called “real” num-
ber is also an abstraction, and we must move far beyond the immediate notion of
“counting numbers” 0, 1, 2, 3, 4, . . ., and beyond even the fractions 21 , 53 , − 47 , etc., if
we are to express even the square root of 2. But here we are helped by a visual image,
and can perceive a straight line extended indefinitely in both directions to give us
a good intuitive impression of the full array of real numbers. The slightly mislead-
ing term “real” for this imagined array is excusable, as we can indeed imagine a
ruler, or a line of ink drawn on a piece of paper, as providing us with some sort of
conceptual image of this array. This greatly helps our understandings of what the
mathematician’s precise notion of a “real number” is intended to idealize. We are
not concerned with whatever might be the nature of the physics of the particles or
fields that might compose our ruler or ink-line; nor, indeed, do we require any con-
cept of the cosmology that may be relevant to the extension out to infinity of our
imagined ruler or ink-line. Our abstract mathematical notion of a “real number”
remains aloof from any such realities of the actual world. Yet, in a curious reversal
of roles, it is this very mathematical idealization that underlies most of our theories
of the actual world.
So, what can be the driving force behind a need to go beyond these seemingly
ubiquitous “real numbers”? What purpose might there be for the introduction of
a “square root of −1”? Such a number fits nowhere within the span of the real
numbers, and it would appear that we have no reason to demand that the equation
x2 + 1 = 0 have any kind of “solution”. The answer to this desire for such entities
lies in the magic that lies hidden within them, but it is not a magic that immediately
reveals itself. In fact, when the first hint of this magic was actually perceived, in
viii Foreword
the mid 16th century, by Girolamo Cardano, and then more completely by Rafael
Bombelli, these strange numbers were dismissed, even by them, as being as useless
as they were mysterious.
It is of some interest to note that it was not in the equation x2 + 1 = 0 that this
hint of magic was first perceived, but in cubic equations like x3 = 15x + 4, which
has the perfectly sensible solution x = 4. Yet, as Bombelli had noted, Cardano’s
general expression for the solution of such cubic equations necessarily involves a
detour into a mysterious world of numbers of a sort where the equation x2 + 1 = 0
is deemed to have the two solutions, now referred to as the imaginary units x = i
and x = −i, of an algebra—now called complex-number algebra—that had appeared
to be consistent, but not what had been regarded as “real”.
This dismissive attitude did not change much until the mid to late 18th cen-
tury, with Leonhard Euler’s remarkable formula eiθ = cos θ + i sin θ and, even
more importantly, the geometrical representation of the entire family of complex
numbers as points in a Euclidean plane, as initially proposed by Caspar Wessel,
where the algebraic operations on complex numbers are readily understood in
geometrical terms. This provided a kind of 2-dimensional “visual reality” to the
array of complex numbers that could be combined with topological notions, such
as employed initially by Carl Friedrich Gauss, and soon followed by others. The
early to mid 19th century saw many important advances, many of these being due
to Augustin-Louis Cauchy, especially with the beauty and the power of contour
integration, and, perhaps most profoundly, with ideas due to Bernhard Riemann.
The very notion of “complex smoothness” of complex functions was expressed
by use of the Cauchy–Riemann equations, and this provided the powerful con-
cept of a holomorphic function that implies that a power-series expansion always
locally exists, this leading to a vast and powerful theory with numerous magical
properties.
The two revolutions of early 20th century physics both owe a profound debt
to complex-number mathematics. This is most manifest with quantum mechanics,
since the basic formalism of that theory depends fundamentally on complex num-
bers and holomorphic functions. We see a remarkable interplay between quantum
spin and the geometry of complex numbers. The basic equations of Schrödinger
and Dirac are both complex equations. In relativity theory, the transformations
relating the visual field of two observers passing close by each other at differ-
ent relativistic speeds is most easily understood in terms of simple holomorphic
functions. Moreover, many solutions of Einstein’s equations for general relativity
benefit greatly from properties of holomorphic functions, as does the description
of gravitational waves.
In view of the undoubted importance of complex analysis in so much of math-
ematics and physics, it is clearly important that there are basic accounts of these
topics available to those unfamiliar (or only partly familiar) with the basic ideas
Foreword ix
of complex analysis. In this foreword I have very much stressed how the visual
or geometric viewpoint has been of vital importance, not only to the historical
development of complex analysis, but also to the proper understanding of the
subject. Tristan Needham’s Visual Complex Analysis as originally published in 1997
was, to my knowledge, unique in the extent to which it was able to cover these
fundamental ideas with such thoroughness, visual elegance, and clarity.
With this 25th Anniversary Edition there have been some significant improve-
ments, most particularly in the incorporation of captions to the diagrams. This
makes it easier for the reader to dip into the arguments, as illustrated so elegantly in
such wonderfully expressive pictures, without necessarily having to look through
to find the relevant portion of the text. In any case, I am sure that readers, over a
broad range of relevant knowledge—from those with no prior experience of com-
plex analysis to those already experts—will gain greatly from the charm, distinct
originality, and visual clarity of the arguments presented here.
PREFACE TO THE 25TH ANNIVERSARY EDITION
Introduction
Mathematical reality exists outside the confines of space and time, but books about
mathematics do not. A quarter century after its publication, I am grateful that an
entirely new generation of mathematicians and scientists has continued to embrace
VCA’s unorthodox, intuitive, and, above all, geometrical approach to Complex
Analysis. To mark the occasion, Oxford University Press has graciously permit-
ted me to revisit the work, resulting in the creation of this significantly improved
25th Anniversary Edition.
• The most obvious change is in the physical dimensions of the book: it has
expanded from 6 ′′ ×9 ′′ to 7 ′′ ×10 ′′ . There are two significant advantages to
this change: (1) the book is more comfortable to hold and read; (2) the 5032
1
Needham (2021).
2
The original edition only contained 501 figures: The figure in this Preface and [6.41] are both new.
xii Preface to the 25th Anniversary Edition
figures—the beating heart of the work!—are now 36% larger, and, one may dare
to hope, 36% clearer!
• Another obvious change is the introduction of the standard numbering sys-
tem for sections, subsections, equations, and figures. In hindsight, my deliberate
avoidance of the standard scheme was perhaps a rather childish, tantrum-like
expression of my disgust with the prevailing, life-sucking reduction of wildly
exciting mathematical ideas to the arid structure enshrined in traditional trea-
tises: “Lemma 12.7.2 implies Theorem 14.3.8”. Let me take this opportunity to
apologize—better 25 years late than never?—to every professor who was ever
brave enough to adopt VCA as the text for their course, only to discover that
they had to struggle to refer their students to any given figure or result!
• The Bibliography has been updated. Not only have previously cited works been
updated to their latest editions, but I have also added a number of new works,
which were not cited in the first edition, for the simple reason that 25 years ago
they did not exist!
• The Index has been improved significantly. Many new entries have been added,
and, perhaps more importantly, wherever a single main entry formerly listed
a long, frustratingly unhelpful string of undifferentiated page numbers, I have
now split it into many individual and helpfully specific subentries. By way of proof,
consider the entry for “Jacobian matrix”, then and now!
• Giving credit to Eugenio Beltrami. In the original edition I pointed out that in 1868
Beltrami discovered3 (and published) the conformal models of the hyperbolic
plane, which Poincaré then rediscovered 14 years later, in 1882. Now, following
the example I set myself in VDGF, I have gone one step further, attempting to put
the record straight by renaming these models as the Beltrami–Poincaré disc and
half-plane. Beltrami also discovered the projective model, and I have renamed
that, too, as the Beltrami–Klein model. Correspondingly, my fancifully named
inhabitants of these hyperbolic worlds have been renamed from “Poincarites”
to Beltrami–Poincarites!
• Twenty five years ago, VCA was on the bleeding edge of what was typograph-
ically possible. Indeed, my editor told me that VCA was the first mathematics
book published by Oxford University Press to be composed in LATEX and yet not
to be typeset using the standard Computer Modern fonts designed by Donald
Knuth, the creator of TEX. I was able to achieve this feat by virtue of Michael
Spivak’s4 then newly created MathTime typeface, and by virtue of the wonderful
3
See Milnor (1982), Stillwell (1996), and Stillwell (2010).
4
Yes, the famous differential geometer!
Preface to the 25th Anniversary Edition xiii
Y&Y TEX System for Windows.5 I was thereby able to typeset the book in Times
text and (mainly) MathTime mathematics—a vast aesthetic improvement over
Computer Modern, in my humble but strong opinion!
Needless to say, the TEX world has moved on considerably in the past 25 years.
But in order to take advantage of these advances, I was forced to grapple with
the task of updating my original TEX files (including my countless macros) from
the ancient (dead?) language of LATEX2.09 to the modern LATEX 2ε . This effort was
rewarded with the ability to typeset this new edition of VCA using the same type-
faces that I very carefully selected for VDGF, all three of which sprang from the
genius mind of Hermann Zapf (1918–2015): Optima for the headings, Palatino for
the text, and, crucially, the remarkable Euler fonts for the mathematics.
Let me pause for a moment to pay tribute to Zapf’s Euler mathematical fonts.
They would appear to me to be at home onboard the starship of an advanced alien
civilization, yet they also evoke the time-worn stone engravings within an ancient
Greek ruin—their beauty transcends space and time: Picture an inscription on the
Guardian of Forever!6
But my fatal attraction to these fonts brought about conflict between the Euler
mathematics in the text of the new edition and the MathTime mathematics in all
the figures of the original edition. Having been born cursed (and I suppose blessed)
with a compulsion to strive for perfection, I had no choice but to undertake the
self-inflicted, Herculean task of hand-editing (within CorelDRAW) all 501 of my
original, hand-drawn figures! I then output new versions in which each MathTime
symbol is here replaced with its matching Euler counterpart, thereby bringing the
figures into perfect alignment with the new text.
I am the impossibly proud father of remarkable twin daughters, Faith and Hope.
I am also the father of VCA and of VDGF. It is therefore a source of deep, resonating
joy that both sets of twins now look like twins!
• Now let me explain the most fundamental change, the one that took me the
greatest time and effort to accomplish, and the one that I believe transforms this
25th Anniversary Edition into a truly new edition, one that may even be of value
to owners of the original edition.
The 503 figures are the mathematical soul of the work. They crystallize all the
geometrical insights I was able to glean from my many years spent struggling to
understand Complex Analysis. Yet, in the entire original edition, you will not find a
single caption. Why?!
Sadly, the answer is simple: cowardice. As a newly minted DPhil student of Pen-
rose, with no track record or reputation, I feared that the mathematical community
5
Sadly, this system (designed by Professor Berthold Horn of MIT) ceased to exist in 2004.
6
The space–time portal featured in the Star Trek episode, “The City on the Edge of Forever”.
xiv Preface to the 25th Anniversary Edition
This approach8 is directly inspired by the works of my teacher, friend, and mentor,
Sir Roger Penrose—to whom this book is dedicated, and to whom I now offer sin-
cere thanks for his generous Foreword!—most remarkably in his Road to Reality,9
where a single figure’s caption can take up a quarter of a page!
This innovation now makes it possible to read VCA in an entirely new way—as
a highbrow comic book! Much of the geometrical reasoning of the work can now be
grasped simply by studying a figure and its accompanying caption, only turning
to the main text for the complete explanation as needed. Furthermore, instead of
undertaking a systematic, linear reading of the work, you are now invited to skip
and hop about, lighting upon whichever figure happens to catch your eye.
7
In my defence, such fears were not entirely groundless: When Princeton University Press sent out
draft chapters of VDGF for review, one of the three anonymous reviewers bluntly declared, “This is
not even mathematics!”
8
This is also my approach in VDGF.
9
Penrose (2005).
10
Needham (1993), Needham (2014).
Preface to the 25th Anniversary Edition xv
11
See Needham (1993), the original 1997 Preface to VCA, and Needham (2014).
12
Sadly, this myth originated with Newton himself, in the heat of his bitter priority battle with
Leibniz over the discovery of the calculus. See Arnol’d (1990), Bloye and Huggett (2011), de Gandt
(1995), Guicciardini (1999), Newton (1687, p.123), and Westfall (1980).
xvi Preface to the 25th Anniversary Edition
the form in which he had originally discovered the calculus in his youth—which
is different again from the Leibnizian form we all learn in college today—and had
instead embraced purely geometrical methods.
Thus it came to pass that by the 1680s Newton’s algebraic infatuation with power
series gave way to a new form of calculus—what he called the “synthetic method
of fluxions”13 —in which the geometry of the Ancients was transmogrified and
reanimated by its application to shrinking geometrical figures in their moment of
vanishing. This is the potent but non-algorithmic form of calculus that we find in
full flower in his great Principia of 1687.
Let me now immediately spell out Newton’s approach, and in significantly
greater detail than I did in the first edition of VCA, in the vain hope that this
new edition may bolster my efforts in VDGF to inspire more mathematicians and
physicists to adopt Newton’s intuitive (yet rigorous14 ) methods than did the first
edition.
If two quantities A and B depend on a small quantity ϵ, and their ratio
approaches unity as ϵ approaches zero, then I shall avoid the more cumbersome
language of limits by following Newton’s precedent in the Principia, saying simply
that, “A is ultimately equal to B”. Also, as I did in earlier works [(Needham, 1993),
(Needham, 2014)], I shall employ the symbol ≍ to denote this concept of ultimate
equality.15 In short,
A
“A is ultimately equal to B” ⇐⇒ A≍B ⇐⇒ lim
= 1.
B ϵÑ0
It follows [exercise] from the theorems on limits that ultimate equality is an equiv-
alence relation, and that it also inherits additional properties of ordinary equality,
e.g.,
13
See Guicciardini (2009, Ch. 9).
14
Fine print to follow!
15
This notation was subsequently adopted (with attribution) by the Nobel physicist, Subrahmanyan
Chandrasekhar (see Chandrasekhar, 1995, p. 44).
Preface to the 25th Anniversary Edition xvii
time I will use the “≍”–notation to present the argument rigorously, whereas in the
first edition I did not. Indeed, this example may be viewed as a recipe for trans-
forming many of VCA’s “explanations” into “proofs”,16 merely by sprinkling on
the requisite ≍’s. With the addition of figure captions, I am now able to do some of
this sprinkling myself, but some must still be left to the reader.
Before looking at this example, I suggest you first read the original presentation
of the argument in the original Preface (immediately following this one) and then
return to this point.
16
I was already using the ≍ notation (both privately and in print) at the time of writing VCA, and,
in hindsight, it was a mistake that I did not employ it throughout the original edition of VCA.
17
See Guicciardini (2009, p. 231)
xviii Preface to the 25th Anniversary Edition
Newton himself did not employ any symbol to represent his concept of “ultimate
equality”. Instead, his devotion to the geometrical method of the Ancients spilled
over into emulating their mode of expression, causing him to write out the words
“ultimately have the ratio of equality” every single time the concept was invoked
in a proof. As Newton (1687, p. 124) explained, the Principia is “written in words at
length in the manner of the Ancients”. Even when Newton claimed that two ratios
were ultimately equal, he insisted on expressing each ratio in words. As a result, I
myself was quite unable to follow Newton’s reasoning without first transcribing
and summarizing each of his paragraphs into “modern” form (which was in fact
already quite common in 1687). Indeed, back in 1982, this was the catalyst for my
private introduction and use of the symbol ≍.
It is my view that Newton’s choice not to introduce a symbol for “ultimate
equality” was a tragically consequential error for the development of mathemat-
ics. As Leibniz’s symbolic calculus swept the world, Newton’s more penetrating
geometrical method fell by the wayside. In the intervening centuries only a hand-
ful of people ever sought to repair this damage and revive Newton’s approach,
the most notable and distinguished recent champion having been V. I. Arnol’d18
[1937–2010].
Had Newton shed the trappings of this ancient mode of exposition and instead
employed some symbol (any symbol!) in place of the words “ultimately have the
ratio of equality”, his dense, paragraph-length proofs in the Principia might have
been reduced to a few succinct lines, and his mode of thought might still be widely
employed today. Both VCA and VDGF are attempts to demonstrate, very con-
cretely, the continuing relevance and vitality of Newton’s geometrical approach,
in areas of mathematics whose discovery lay a century in the future at the time of
his death in 1727.
Allow me to insert some fine print concerning my use of the words “rigour”
and “proof”. Yes, my occasional explicit use of Newtonian ultimate equalities in
this new edition represents a quantum jump in rigour, as compared to my original
exposition in VCA, but there will be some mathematicians who will object (with
justification!) that even this increase in rigour is insufficient, and that none of the
“proofs” in this work are worthy of that title, including the one just given: I did
not actually prove that the side of the triangle is ultimately equal to the arc of the
circle.
I can offer no logical defence, but will merely repeat the words I wrote in the
original Preface to VCA, 25 years ago: “... suppose one believes, as I do, that our
mathematical theories are attempting to capture aspects of a robust Platonic world
that is not of our making. I would then contend that an initial lack of rigour is a
small price to pay if it allows the reader to see into this world more directly and
18
See, for example, Arnol’d (1990).
Preface to the 25th Anniversary Edition xix
Yet, as Poincaré was the first to recognize, this strange geometry arises naturally
across many parts of mathematics and physics. For example, the final figure of this
book reveals how hyperbolic geometry unifies all the methods of solving the two-
dimensional Dirichlet Problem. Furthermore, the visionary insights of Thurston21
19
Upon reading these words, a strongly supportive member of the Editorial Board of Princeton
University Press suggested to my editor that in place of “Q.E.D.”, I conclude each of my proofs in
VDGF with the letters, “P.B.R.D.”!
20
I freely admit that I have not undertaken a systematic study of all the Complex Analysis textbooks
that have been published during the last 25 years, and I am aware that there exist excellent exceptions to
the following generalization, my favourite ones being Shaw (2006) and Stewart and Tall (2018), which
I highly recommend precisely because of their very unusual contents. Incidentally, Shaw just happens
to have been a fellow student of Penrose!
21
For details of Thurston’s Geometrization Conjecture—now Theorem!—see Thurston (1997).
xx Preface to the 25th Anniversary Edition
The reason I lavished such extravagant attention upon these deceptively sim-
ple transformations is that they are possessed of magical powers, manifesting
themselves in multifaceted guises across mathematics and physics. Again, it was
Poincaré who was the first to recognize this. They are the isometries of both the two-
dimensional hyperbolic plane (introduced above) and of three-dimensional hyper-
bolic space; they are the famous Lorentz transformations (isometries) of Minkowski
and Einstein’s four-dimensional spacetime; and when the constants are all integers,
and ad − bc = 1, they form the modular group, describing the symmetries of the
modular functions so important in modern number theory; and the list goes on . . .
The second pair of innovations with respect to the contents of VCA is centred
on the use of vector fields as an alternative means of visualizing complex mappings.
Instead of picturing a point z in C as being mapped to another point w = f(z)
in another copy of C—which is the paradigm in force throughout the first nine
chapters—we instead picture f(z) as a vector emanating from z.
• The Topology of Vector Fields is the subject of Chapter 10. Here we shed
new light on the vital topological concept of the winding number (the subject
of Chapter 7) by instead viewing it through the prism of the index of a singular
point of a vector field. The climax of the topological analysis is a proof of the
glorious Poincaré–Hopf Theorem, (10.4), which relates the indices of a flow on a
closed surface to the surface’s topological genus, which counts how many holes
the surface has.
Preface to the 25th Anniversary Edition xxi
So far as I can tell, my final22 major innovative topic in VCA has still not had the
impact I believe it deserves, despite my best efforts, and despite the even earlier,
independent efforts of its first principal champion, Professor Bart Braden, who
should also be credited23 with having named the concept in honour of its first
proponent:
Here, the imaginary flux component is positive if P flows across C from our left to our
right as we face forward in the direction of travel along C.
Next, it follows immediately from the Cauchy–Riemann equations, (4.7), that
there is a wonderfully vivid physical interpretation of the existence of the complex
derivative (what I call the amplitwist), namely,
If and only if the amplitwist f ′ (z) exists, then the Pólya vector
field P(z) ≡ f(z) is divergence-free and curl-free.
This in turn immediately provides a marvellously physical explanation of one the
central results of Complex Analysis, namely, Cauchy’s Theorem, (8.20), the simplest
version of which states that
If the amplitwist f ′ exists everywhere on and inside a simple
H
closed loop L, then f(z)dz = 0.
L
But if L contains a singularity of f(z), then the integral need not vanish. An example
of central importance is f(z) = (1/z), for which the Pólya vector field is the radial
outward flow from a source at the origin. Choosing L to be an origin-centred circle
22
In fact there are many other innovations scattered across the work, most notably my introduction
of a concept I christened the complex curvature, which I apply to central force fields in Chapter 5, and
to the geometry of harmonic functions in Chapter 12; there’s also the Topological Argument Principle in
Chapter 7, and more besides, but I cannot attempt to catalogue all these ideas and discoveries here.
Some of the observations in Chapter 12 were previously published in Needham (1994), which won the
MAA’s Carl B. Allendoerfer Award in 1995.
23
See Braden (1985), Braden (1987)—which won the MAA’s Carl B. Allendoerfer Award in 1988—and
Braden (1991).
24
Rest assured that the following ideas will all be explained ab initio in the main text; this section
of the Preface is addressed to experienced readers who are already familiar with the fundamentals of
standard Complex Analysis.
xxii Preface to the 25th Anniversary Edition
For clearly there is no flow along K, and since P(z) = (1/z) flows orthogonally
across K (from left to right) with speed (1/r), its flux across K is (2πr)(1/r) = 2π.
Furthermore, this physical interpretation explains why the value of the integral will
remain 2πi if K is continuously deformed into a general loop encircling the source
at the origin, so long as K does not cross that point as it is deformed.
Lastly, the ability of modern computers to quickly and easily draw the Pólya
vector field of any explicit formula for f(z) makes the concept all the more powerful
as a means of visualizing Complex Analysis.
While the evangelical work of VCA (and of Braden’s earlier papers) failed to
have its desired effect in the 20th century, my fervent hope now is that by explic-
itly singling out this concept for praise in this new Preface, more people may take
notice of it, and the gospel of the Pólya vector field may thereby be spread amongst
multitudes of new believers in the 21st century!
Cheers!
I raise my glass to the next 25 years!
T. N.
Mill Valley, California
June, 2022
PREFACE
Theories of the known, which are described by different physical ideas, may be equivalent
in all their predictions and hence scientifically indistinguishable. However, they are not
psychologically identical when trying to move from that base into the unknown. For differ-
ent views suggest different kinds of modifications which might be made and hence are not
equivalent in the hypotheses one generates from them in one’s attempt to understand what
is not yet understood.
Feynman (1966)
A Parable
Imagine a society in which the citizens are encouraged, indeed compelled up to
a certain age, to read (and sometimes write) musical scores. All quite admirable.
However, this society also has a very curious—few remember how it all started—
and disturbing law: Music must never be listened to or performed!
Though its importance is universally acknowledged, for some reason music is
not widely appreciated in this society. To be sure, professors still excitedly pore
over the great works of Bach, Wagner, and the rest, and they do their utmost to
communicate to their students the beautiful meaning of what they find there, but
they still become tongue-tied when brashly asked the question, “What’s the point
of all this?!”
In this parable, it was patently unfair and irrational to have a law forbid-
ding would-be music students from experiencing and understanding the subject
directly through “sonic intuition.” But in our society of mathematicians we have
such a law. It is not a written law, and those who flout it may yet prosper, but it
says, Mathematics must not be visualized!
More likely than not, when one opens a random modern mathematics text on a
random subject, one is confronted by abstract symbolic reasoning that is divorced
from one’s sensory experience of the world, despite the fact that the very phenom-
ena one is studying were often discovered by appealing to geometric (and perhaps
physical) intuition.
This reflects the fact that steadily over the last hundred years the honour of
visual reasoning in mathematics has been besmirched. Although the great mathe-
maticians have always been oblivious to such fashions, it is only recently that the
“mathematician in the street” has picked up the gauntlet on behalf of geometry.
xxiv Preface
The present book openly challenges the current dominance of purely symbolic
logical reasoning by using new, visually accessible arguments to explain the truths
of elementary complex analysis.
Computers
In part, the resurgence of interest in geometry can be traced to the mass-availability
of computers to draw mathematical objects, and perhaps also to the related, some-
what breathless, popular interest in chaos theory and in fractals. This book instead
advocates the more sober use of computers as an aid to geometric reasoning.
I have tried to encourage the reader to think of the computer as a physicist would
his laboratory—it may be used to check existing ideas about the construction of
the world, or as a tool for discovering new phenomena which then demand new
ideas for their explanation. Throughout the text I have suggested such uses of the
computer, but I have deliberately avoided giving detailed instructions. The reason
is simple: whereas a mathematical idea is a timeless thing, few things are more
ephemeral than computer hardware and software.
Having said this, the program “f(z)” is currently25 the best tool for visually
exploring the ideas in this book; a free demonstration version can be downloaded
directly from Lascaux Graphics. On occasion it would also be helpful if one had
access to an all-purpose mathematical engine such as Maple or Mathematica. How-
ever, I would like to stress that none of the above software is essential: the entire
book can be fully understood without any use of a computer.
Finally, some readers may be interested in knowing how computers were used
to produce this book. Perhaps five of the 501 diagrams were drawn using output
from Mathematica; the remainder I drew by hand (or rather “by mouse”) using
CorelDRAW, occasionally guided by output from “f(z)”. I typeset the book in LATEX
using the wonderful Y&Y TEX System for Windows26 , the figures being included as
EPS files. The text is Times27 , with Helvetica heads, and the mathematics is princi-
pally MathTime, though nine other mathematical fonts make cameo appearances.
All of these Adobe Type 1 fonts were obtained from Y&Y, Inc., with the exception
of Adobe’s MathematicalPi-Six font, which I used to represent quaternions. Finally,
OUP printed the book directly from my PostScript file.
25
Sadly, this no longer exists.
26
Sadly, this no longer exists.
27
In this new edition, the headings are Optima, the text is Palatino, and the mathematics is Euler.
Preface xxv
dT L dT
= =⇒ = L2 = 1 + T 2 .
L dθ 1 dθ
Only gradually did I come to realize how naturally this mode of thought could be
applied—almost exactly 300 years later!—to the geometry of the complex plane.
and stretch slightly to pass from one to the next. Such places are marked “[exer-
cise]”; they often require nothing more than a simple calculation or a moment of
reflection.
This brings me to the exercises proper, which may be found at the end of each
chapter. In the belief that the essential prerequisite for finding the answer to a
question is the desire to find it, I have made every effort to provide exercises that
provoke curiosity. They are considerably more wide-ranging than is common, and
they often establish important facts which are then used freely in the text itself.
While problems whose be all and end all is routine calculation are thereby avoided,
I believe that readers will automatically develop considerable computational skill
in the process of seeking solutions to these problems. On the other hand, my inten-
tion in a large number of the exercises is to illustrate how geometric thinking can
often replace lengthy calculation.
Any part of the book marked with a star (“*”) may be omitted on a first read-
ing. If you do elect to read a starred section, you may in turn choose to omit any
starred subsections. Please note, however, that a part of the book that is starred is
not necessarily any more difficult, nor any less interesting or important, than any
other part of the book.
• Traditional Course. Chapters 1 to 9, omitting all starred material (e.g., the whole
of Chapter 6).
• Vector Field Course. In order to take advantage of the Pólya vector field
approach to visualizing complex integrals, one could follow the “Tradi-
tional Course” above, omitting Chapter 9, and adding the unstarred parts of
Chapters 10 and 11.
• Non-Euclidean Course. At the expense of teaching any integration, one could
give a course focused on Möbius transformations and non-Euclidean geometry.
These two related parts of complex analysis are probably the most important
ones for contemporary mathematics and physics, and yet they are also the ones
that are almost entirely neglected in undergraduate-level texts. On the other
hand, graduate-level works tend to assume that you have already encountered
the main ideas as an undergraduate: Catch 22!
the starred subsections; all of Chapter 4; all of Chapter 6, including the starred
sections but (possibly) omitting the starred subsections.
Prof. Gerald Alexanderson of Santa Clara University has my sincere thanks for the
encouragement he offered me upon reading some of the earliest chapters, as well
as for his many subsequent acts of kindness.
I will always be grateful for the education I received at Merton College, Oxford.
It is therefore especially pleasing and fitting to have this book published by OUP,
and I would particularly like to thank Dr. Martin Gilchrist, the former Senior Math-
ematics Editor, for his enthusiastic encouragement when I first approached him
with the idea of the book.
When I first arrived at USF from England in 1989 I had barely seen a com-
puter. The fact that OUP printed this book directly from my Internet-transmitted
PostScript files is an indication of how far I have come since then. I owe all this
to James Kabage. A mere graduate student at the time we met, Jim quickly rose
through the ranks to become Director of Network Services. Despite this fact, he
never hesitated to spend hours with me in my office resolving my latest hardware or
software crisis. He always took the extra time to clearly explain to me the reasoning
leading to his solution, and in this way I became his student.
I also thank Dr. Benjamin Baab, the Executive Director of Information Technol-
ogy Services at USF. Despite his lofty position, he too was always willing to roll up
his sleeves in order to help me resolve my latest Microsoft conundrum.
Eric Scheide (our multitalented Webmaster) has my sincere gratitude for writing
an extremely nifty Perl program that greatly speeded my creation of the index.
I thank Prof. Berthold Horn of MIT for creating the magnificent Y&Y TEX System
for Windows28 , for his generous help with assorted TEXnical problems, and for his
willingness to adopt my few suggestions for improving what I consider to be the
Mercedes-Benz of the TEX world.
Similarly, I thank Martin Lapidus of Lascaux Graphics for incorporating many
of my suggestions into his “f(z)”29 program, thereby making it into an even better
tool for doing “visual complex analysis”.
This new printing of the book incorporates a great many corrections. Most of
these were reported by readers, and I very much appreciate their efforts. While I
cannot thank each one of these readers by name, I must acknowledge Dr. R. von
Randow for single-handedly having reported more than 30 errors.
As a student of Roger Penrose I had the privilege of watching him think out loud
by means of his beautiful blackboard drawings. In the process, I became convinced
that if only one tried hard enough—or were clever enough!—every mathemati-
cal mystery could be resolved through geometric reasoning. George Burnett-Stuart
and I became firm friends while students of Penrose. In the course of our endless
discussions of music, physics, and mathematics, George helped me to refine both
28
While Y & Y no longer exists, my gratitude lives on.
29
While “f(z)” no longer exists, my gratitude lives on.
Acknowledgements xxxi
Bibliography 653
Index 661
CHAPTER 1
1.1 Introduction
Visual Complex Analysis. 25th Anniversary Edition. Tristan Needham, Oxford University Press.
© Tristan Needham (2023). DOI: 10.1093/oso/9780192868916.003.0001
2 Geometry and Complex Arithmetic
[1.1] The complex plane enables us to visualize the abstract complex number a + ib as a
simple point in the plane, with Cartesian coordinates (a, b).
numbers were called “impossible” or “imaginary”, the latter term having (unfortu-
nately) lingered to the present day1 . Even in 1770 the situation was still sufficiently
confused that it was possible for so great a mathematician as Euler to mistakenly
√ √ √
argue that −2 −3 = 6.
The root cause of all this trouble seems to have been a psychological or philo-
sophical block. How could one investigate these matters with enthusiasm or
confidence when nobody felt they knew the answer to the question, “What is a
complex number?”
A satisfactory answer to this question was only found at the end of the eigh-
teenth century2 . Independently, and in rapid succession, Wessel, Argand, and
Gauss all recognized that complex numbers could be given a simple, concrete, geo-
metric interpretation as points (or vectors) in the plane: The mystical quantity a + ib
should be viewed simply as the point in the xy-plane having Cartesian coordi-
nates (a, b), or equivalently as the vector connecting the origin to that point. See
[1.1]. When thought of in this way, the plane is denoted C and is called the complex
plane3 .
The operations of adding or multiplying two complex numbers could
now be given equally definite meanings as geometric operations on the two
1
However, an “imaginary number” now refers to a real multiple of i, rather than to a general com-
plex number. Incidentally, the term “real number” was introduced precisely to distinguish such a
number from an “imaginary number”.
2
Wallis almost hit on the answer in 1673; see Stillwell (2010, §14.4) for a detailed account of this
interesting near miss.
3
Also known as the “Gauss plane” or the “Argand plane”.
Introduction 3
[1.2] [a] Geometrically, the sum of the complex numbers A and B is the diagonal of the
parallelogram with sides A and B. [b] Geometrically, the length of the product of A and
B is the product of their separate lengths, and its angle is the sum of their separate angles.
corresponding points (or vectors) in the plane. The rule for addition is illustrated in
[1.2a]:
The sum A + B of two complex numbers is given by the parallelogram
rule of ordinary vector addition. (1.1)
Note that this is consistent with [1.1], in the sense that 4+3i (for example) is indeed
the sum of 4 and 3i.
Figure [1.2b] illustrates the much less obvious rule for multiplication:
The length of AB is the product of the lengths of A and B, and the angle
of AB is the sum of the angles of A and B. (1.2)
This rule is not forced on us in any obvious way by [1.1], but note that it is at least
consistent with it, in the sense that 3i (for example) is indeed the product of 3 and i.
Check this for yourself. As a more exciting example, consider the product of i with
itself. Since i has unit length and angle (π/2), i2 has unit length and angle π. Thus
i2 = −1.
The publication of the geometric interpretation by Wessel and by Argand went
all but unnoticed, but the reputation of Gauss (as great then as it is now) ensured
wide dissemination and acceptance of complex numbers as points in the plane. Per-
haps less important than the details of this new interpretation (at least initially) was
the mere fact that there now existed some way of making sense of these numbers—
that they were now legitimate objects of investigation. In any event, the floodgates
of invention were about to open.
It had taken more than two and a half centuries to come to terms with complex
numbers, but the development of a beautiful new theory of how to do calculus with
such numbers (what we now call complex analysis) was astonishingly rapid. Most of
the fundamental results were obtained (by Cauchy, Riemann, and others) between
1814 and 1851—a span of less than forty years!
4 Geometry and Complex Arithmetic
Other views of the history of the subject are certainly possible. For example,
Stewart and Tall (2018) suggest that the geometric interpretation4 was somewhat
incidental to the explosive development of complex analysis. However, it should
be noted that Riemann’s ideas, in particular, would simply not have been possible
without prior knowledge of the geometry of the complex plane.
x2 = mx + c. (1.3)
Two thousand years BCE, it was already known that such equations could be solved
using a method that is equivalent to the modern formula,
h p i
x = 21 m ± m2 + 4c .
But what if m2 + 4c is negative? This was the very problem that led Cardano to
consider square roots of negative numbers. Thus far the textbook is being histor-
ically accurate, but next we read that the need for (1.3) to always have a solution
forces us to take complex numbers seriously. This argument carries almost as little
weight now as it did in the sixteenth century. Indeed, we have already pointed out
that Cardano did not hesitate to discard such “solutions” as useless.
It was not that Cardano lacked the imagination to pursue the matter further,
rather he had a fairly compelling reason not to. For the ancient Greeks mathe-
matics was synonymous with geometry, and this conception still held sway in
the sixteenth century. Thus an algebraic relation such as (1.3) was not so much
thought of as a problem in its own right, but rather as a mere vehicle for solving a
genuine problem of geometry. For example, (1.3) may be considered to represent
the problem of finding the intersection points of the parabola y = x2 and the line
y = mx + c. See [1.3a].
In the case of L1 the problem has a solution; algebraically, (m2 + 4c) > 0 and the
two intersection points are given by the formula above. In the case of L2 the problem
clearly does not have a solution; algebraically, (m2 + 4c) < 0 and the absence of
4
We must protest one piece of their evidence: Although Wallis did attempt a geometric interpreta-
tion of complex numbers in 1673, he did not hit upon the “correct” picture; see footnote 2.
Introduction 5
[1.3] [a] While some lines intersect a parabola, some do not, so the quadratic equation
x2 = mx + c may or may not have solutions. [b] But x3 = mx + c must always have a
solution.
[Ex. 1 shows that a general cubic can always be reduced to this form.] This equation
represents the problem of finding the intersection points of the cubic curve y = x3
and the line y = 3px+2q. See [1.3b]. Building on the work of del Ferro and Tartaglia,
Cardano’s Ars Magna showed that this equation could be solved by means of a
remarkable formula [see Ex. 2]:
q p q p
3 3
x = q + q2 − p3 + q − q2 − p3 . (1.4)
Try it yourself on x3 = 6x + 6.
Some thirty years after this formula appeared, Bombelli recognized that there
was something strange and paradoxical about it. First note that if the line
y = 3px + 2q is such that p3 > q2 then the formula involves complex numbers.
For example, Bombelli considered x3 = 15x + 4, which yields
√
3
√3
x = 2 + 11i + 2 − 11i.
In the previous case of [1.3a] this merely signalled that the geometric problem had
no solution, but in [1.3b] it is clear that the line will always hit the curve! In fact
inspection of Bombelli’s example yields the solution x = 4.
As he struggled to resolve this paradox, Bombelli had what he called a “wild
thought”: perhaps the solution x = 4 could be recovered from the above expression
√ √
if 3 2 + 11i = 2 + ni and 3 2 − 11i = 2 − ni. Of course for this to work he would
6 Geometry and Complex Arithmetic
A + B = (a + i a e = (a + b) + i (e
e) + (b + i b) e
a + b). (1.5)
√
Next, to see if there was indeed a value of n for which 3 2 + 11i = 2+in, he needed
to calculate (2 + in)3 . To do so he assumed that he could multiply out brackets as
in ordinary algebra, so that
e = ab + i (a b
e) (b + i b)
(a + i a e+a
e b) + i2 a e
e b.
Using i2 = −1, he concluded that the product of two complex numbers would be
given by
AB = (a + i a e = (ab − a
e) (b + i b) e + i (a b
e b) e+a
e b). (1.6)
This rule vindicated his “wild thought”, for he was now able to show that (2±i)3 =
2 ± 11i. Check this for yourself.
While complex numbers themselves remained mysterious, Bombelli’s work on
cubic equations thus established that perfectly real problems required complex
arithmetic for their solution.
Just as with its birth, the subsequent development of the theory of complex num-
bers was inextricably bound up with progress in other areas of mathematics (and
also physics). Sadly, we can only touch on these matters in this book; for a full and
fascinating account of these interconnections, the reader is instead referred to Still-
well (2010). Repeating what was said in the Preface, we cannot overstate the value
of reading Stillwell’s book alongside this one.
[1.4] Visual summary of the terminology and notation used to describe complex numbers.
It is valuable to grasp from the outset that (according to the geometric view) a
complex number is a single, indivisible entity—a point in the plane. Only when
we choose to describe such a point with numerical coordinates does a complex
number appear to be compound or “complex”. More precisely, C is said to be two
dimensional, meaning that two real numbers (coordinates) are needed to label a point
within it, but exactly how the labelling is done is entirely up to us.
One way to label the points is with Cartesian coordinates (the real part x and
the imaginary part y), the complex number being written as z = x + iy. This is the
natural labelling when we are dealing with the addition of two complex numbers,
because (1.5) says that the real and imaginary parts of A+B are obtained by adding
the real and imaginary parts of A and B.
In the case of multiplication, the Cartesian labelling no longer appears natural,
for it leads to the messy and unenlightening rule (1.6). The much simpler geometric
rule (1.2) makes it clear that we should instead label a typical point z with its polar
coordinates, r = |z| and θ = arg z. In place of z = x+iy we may now write z = r∠θ,
where the symbol ∠ serves to remind us that θ is the angle of z. [Although this
notation is still used by some, we shall only employ it briefly; later in this chapter
we will discover a much better notation (the standard one) which will then be used
throughout the remainder of the book.] The geometric multiplication rule (1.2) now
takes the simple form,
In common with the Cartesian label x + iy, a given polar label r∠θ specifies a
unique point, but (unlike the Cartesian case) a given point does not have a unique
8 Geometry and Complex Arithmetic
polar label. Since any two angles that differ by a multiple of 2π correspond to the
same direction, a given point has infinitely many different labels:
This simple fact about angles will become increasingly important as our subject
unfolds.
The Cartesian and polar coordinates are the most common ways of labelling
complex numbers, but they are not the only ways. In Chapter 3 we will meet
another particularly useful method, called “stereographic” coordinates.
1.1.4 Practice
Before continuing, we strongly suggest that you make yourself comfortable with
the concepts, terminology, and notation introduced thus far. To do so, try to
convince yourself geometrically (and/or algebraically) of each of the following facts:
p
Re(z) = 12 [z + z] Im(z) = 2i1
[z − z] |z| = x2 + y2
Im(z)
tan[arg z] = Re(z) z z = |z|2 r∠θ = r (cos θ + i sin θ)
Defining 1
z by (1/z) z = 1, it follows that 1
z = 1
r∠θ = r1 ∠(−θ).
R∠ϕ R 1 x y
= ∠(ϕ − θ) = 2 −i 2
r∠θ r (x + iy) x + y2 x + y2
√
(1 + i)4 = −4 (1 + i)13 = −26 (1 + i) (1 + i 3)6 = 26
√
(1 + i 3)3 (1 + i)5 √
= −4i √ = − 2 ∠ − (π/12) r∠θ = r∠(−θ)
(1 − i)2 ( 3 + i) 2
z1 + z 2 = z1 + z 2 z1 z 2 = z1 z2 z1 /z2 = z1 /z2 .
Lastly, establish the so-called generalized triangle inequality:
[1.5] A complex number can be interpreted as a rotation and expansion of the plane.
Geometrically, multiplication by a complex number A = R∠ϕ rotates the plane through
angle ϕ, and expands it by R, so, in particular, parallelograms are transformed into similar
parallelograms. It follows that A(B + C) = AB + AC.
First we will show how the symbolic rule may be derived from the geometric
rule. To do so we shall rephrase the geometric rule (1.7) in a particularly useful
and important way. Let z denote a general point in C, and consider what hap-
pens to it—where it moves to—when it is multiplied by a fixed complex number
A = R∠ϕ. According to (1.7), the length of z is magnified by R, while the angle of z is
increased by ϕ. Now imagine that this is done simultaneously to every point of the
plane:
• Both the rotation and the expansion are centred at the origin.
• It makes no difference whether we do the rotation followed by the expansion,
or the expansion followed by the rotation.
• If R < 1 then the “expansion” is in reality a contraction.
Figure [1.5] illustrates the effect of such a transformation, the lightly shaded
shapes being transformed into the darkly shaded shapes. Check for yourself that
√
in this example A = 1 + i 3 = 2 ∠ π3 .
It is now a simple matter to deduce the symbolic rule from the geometric
rule. Recall the essential steps taken by Bombelli in deriving (1.6): (i) i2 = −1;
(ii) brackets can be multiplied out, i.e., if A, B, C, are complex numbers then
10 Geometry and Complex Arithmetic
A(B + C) = AB + AC. We have already seen that the geometric rule gives us (i),
and figure [1.5] now reveals that (ii) is also true, for the simple reason that rotations
and expansions preserve parallelograms. By the geometric definition of addition, B + C
is the fourth vertex of the parallelogram with vertices 0, B, C. To establish (ii), we
merely observe that multiplication by A rotates and expands this parallelogram
into another parallelogram with vertices 0, AB, AC and A(B + C). This completes
the derivation of (1.6).
Conversely, we now show how the geometric rule may be derived from the sym-
bolic rule5 . We begin by considering the transformation z ÞÑ iz. According to the
symbolic rule, this means that (x + iy) ÞÑ (−y + ix), and [1.6a] reveals that iz is
z rotated through a right angle. We now use this fact to interpret the transformation
z ÞÑ A z, where A is a general complex number. How this is done may be grasped
sufficiently well using the example A = 4 + 3i = 5∠ϕ, where ϕ = tan−1 (3/4).
5
In every text we have examined this is done using trigonometric identities. We believe that the
present argument supports the view that such identities are merely complicated manifestations of the
simple rule for complex multiplication.
Euler’s Formula 11
See [1.6b]. The symbolic rule says that brackets can be multiplied out, so our
transformation may be rewritten as follows:
z ÞÑ A z = (4 + 3i)z
= 4z + 3(iz)
π
= 4z + 3 z rotated by 2 .
This is visualized in [1.6c]. We can now see that the shaded triangles in [1.6c] and
[1.6b] are similar, so multiplication by 5∠ϕ does indeed rotate the plane by ϕ, and
expand it by 5. Done.
1.2.1 Introduction
It is time to replace the r∠θ notation with a much better one that depends on the
following miraculous fact:
This result was discovered by Leonhard Euler around 1740, and it is called Euler’s
formula in his honour.
Before attempting to explain this result, let us say something of its meaning and
utility. As illustrated in [1.7a], the formula says that eiθ is the point on the unit
circle at angle θ. Instead of writing a general complex number as z = r∠θ, we can
now write z = r eiθ . Concretely, this says that to reach z we must take the unit
[1.7] [a] Euler’s formula states that eiθ has unit length and points at angle θ; expanding
this by r, we may reach any complex number. [b] The complex velocity V(t) (tangent to
the complex orbit Z(t)) is ultimately equal to M/δ, as δ Ñ 0.
12 Geometry and Complex Arithmetic
vector eiθ that points at z, then stretch it by the length of z. Part of the beauty of
this representation is that the geometric rule (1.7) for multiplying complex numbers
now looks almost obvious:
R eiϕ r eiθ = Rr ei(ϕ+θ) .
Put differently, algebraically manipulating eiθ in the same way as the real function
ex yields true facts about complex numbers.
In order to explain Euler’s formula we must first address the more basic ques-
tion, “What does eiθ mean?” Surprisingly, many authors answer this by defining
eiθ , out of the blue, to be (cos θ + i sin θ)! This gambit is logically unimpeachable,
but it is also a low blow to Euler, reducing one of his greatest achievements to a
mere tautology. We will therefore give two heuristic arguments in support of (1.10);
deeper arguments will emerge in later chapters.
d Z(t + δ) − Z(t) M
Z(t) = lim = lim = V(t).
dt δÑ0 δ δÑ0 δ
Thus, given a complex function Z(t) of a real variable t, we can always visualize Z
as the position of a moving particle, and dZ
dt as its velocity.
Euler’s Formula 13
[1.8] First explanation of Euler’s formula. The orbit Z(t) = eit is characterized by the
fact that its velocity V = iZ = (its position rotated by π2 ). Thus, the particle must always
move at right angles to its current position, and at unit speed. Therefore, after time t = θ
it will have travelled distance θ round the unit circle, subtending angle θ at the origin, so
eiθ = cos θ + i sin θ.
We can now use this idea to find the trajectory in the case Z(t) = eit . See [1.8].
According to (1.11),
velocity = V = iZ = position, rotated through a right angle.
Since the initial position of the particle is Z(0) = e0 = 1, its initial velocity is i,
and so it is moving vertically upwards. A split second later the particle will have
moved very slightly in this direction, and its new velocity will be at right angles to
its new position vector. Continuing to construct the motion in this way, it is clear
that the particle will travel round the unit circle.
Since we now know that |Z(t)| remains equal to 1 throughout the motion, it fol-
lows that the particle’s speed |V(t)| also remains equal to 1. Thus after time t = θ
the particle will have travelled a distance θ round the unit circle, and so the angle
of Z(θ) = eiθ will be θ. This is the geometric statement of Euler’s formula.
[1.9] Second explanation of Euler’s formula. Imagine the terms of the horizontal power
series for eθ to be rods connected by hinges. Now wrap the series into the illustrated
spiral, with each rod making a right angle with the one before it. The argument in the text
confirms Euler’s miraculous conclusion: the spiral ends at the point on the unit circle at
angle θ. So, imagining the rigid, vertical iθ-rod transformed into flexible string, we may
wrap it onto the unit circle, and its end will coincide with the end of the spiral!
Putting x equal to a real value θ, this infinite sum of horizontal real numbers is
visualized in [1.9]. To make sense of eiθ , we now cling to the power series and put
x = iθ:
(iθ)2 (iθ)3
eiθ = 1 + iθ + + + ··· .
2! 3!
As illustrated in [1.9], this series is just as meaningful as the series for eθ , but
instead of the terms all having the same direction, here each term makes a right
angle with the previous one, producing a kind of spiral.
This picture makes it clear that the known convergence of the series for eθ guar-
antees that the spiral series for eiθ converges to a definite point in C. However, it
is certainly not clear that it will converge to the point on the unit circle at angle θ.
To see this, we split the spiral into its real and imaginary parts:
where
θ2 θ4 θ3 θ5
C(θ) = 1 − + − ··· , and S(θ) = θ − + − ··· .
2! 4! 3! 5!
At this point we could obtain Euler’s formula by appealing to Taylor’s Theorem,
which shows that C(θ) and S(θ) are the power series for cos θ and sin θ. However,
we can also get the result by means of the following elementary argument that does
not require Taylor’s Theorem.
Euler’s Formula 15
We wish to show two things about eiθ = C(θ) + iS(θ): (i) it has unit length, and
(ii) it has angle θ. To do this, first note that differentiation of the power series C and
S yields
C ′ = −S and S ′ = C,
where a prime denotes differentiation with respect to θ.
To establish (i), observe that
d iθ 2
|e | = (C2 + S2 ) ′ = 2(CC ′ + SS ′ ) = 0,
dθ
which means that the length of eiθ is independent of θ. Since ei0 = 1, we deduce
that |eiθ | = 1 for all θ.
To establish (ii) we must show that Θ(θ) = θ, where Θ(θ) denotes the angle of
eiθ , so that
S(θ)
tan Θ(θ) = .
C(θ)
Since we already know that C2 + S2 = 1, we find that the derivative of the LHS of
the above equation is
′ S2 Θ′
[tan Θ(θ)] = (1 + tan Θ) Θ = 1 + 2 Θ ′ = 2 ,
2 ′
C C
and that the derivative of the RHS is
′
S S ′C − C ′S 1
= 2
= 2.
C C C
Thus
dΘ
= Θ ′ = 1,
dθ
which implies that Θ(θ) = θ + const. Taking the angle of ei0 = 1 to be 0 [would it
make any geometric difference if we took it to be 2π?], we find that Θ = θ.
Although it is incidental to our purpose, note that we can now conclude (with-
out Taylor’s Theorem) that C(θ) and S(θ) are the power series of cos θ and
sin θ.
1.3.1 Introduction
Often problems that do not appear to involve complex numbers are nevertheless
solved most elegantly by viewing them through complex spectacles. In this section
we will illustrate this point with a variety of examples taken from diverse areas
of mathematics. Further examples may be found in the exercises at the end of the
chapter.
The first example [trigonometry] merely illustrates the power of the concepts
already developed, but the remaining examples develop important new ideas.
1.3.2 Trigonometry
All trigonometric identities may be viewed as arising from the rule for complex
multiplication. In the following examples we will reduce clutter by using the
following shorthand: C ≡ cos θ, S ≡ sin θ, and similarly, c ≡ cos ϕ, s ≡ sin ϕ.
To find an identity for cos(θ + ϕ), view it as a component of ei(θ+ϕ) . See [1.11a].
Since
cos(θ + ϕ) + i sin(θ + ϕ) = ei(θ+ϕ)
= eiθ eiϕ
= (C + iS)(c + is)
= [Cc − Ss] + i[Sc + Cs],
we obtain not only an identity for cos(θ + ϕ), but also one for sin(θ + ϕ):
cos(θ + ϕ) = Cc − Ss and sin(θ + ϕ) = Sc + Cs.
This illustrates another powerful feature of using complex numbers: every complex
equation says two things at once.
Some Applications 17
[1.11] [a] Trigonometric identities for cos(θ + ϕ) and sin(θ + ϕ) are easily obtained
by taking the real and imaginary parts of ei(θ+ϕ) = eiθ eiϕ . [b] An identity for tan 3θ in
terms of T = tan θ follows easily from the illustrated fact that (1 + iT )3 has angle 3θ.
To simultaneously find identities for cos 3θ and sin 3θ, consider ei3θ :
cos 3θ + i sin 3θ = ei3θ = (eiθ )3 = (C + iS)3 = C3 − 3CS2 + i 3C2 S − S3 .
[1.12] A surprising property of quadrilaterals. [a] If squares are constructed on the sides
of a quadrilateral, the line-segments joining the centres of opposite squares are perpen-
dicular and of equal length! A short algebraic proof is made possible by taking the edges
of the quadrilateral to be complex numbers that sum to zero. [b] An alternative geometric
proof is based on the illustrated fact: If squares are constructed on two sides of an arbitrary
triangle, then the line-segments from their centres to the midpoint m of the remaining side
are perpendicular and of equal length.
1.3.3 Geometry
We shall base our discussion of geometric applications on a single example. In
[1.12a] we have constructed squares on the sides of an arbitrary quadrilateral. Let
us prove what this picture strongly suggests: the line-segments joining the centres of
opposite squares are perpendicular and of equal length. It would require a great deal
of ingenuity to find a purely geometric proof of this surprising result, so instead
of relying on our own intelligence, let us invoke the intelligence of the complex
numbers!
Introducing a factor of 2 for convenience, let 2a, 2b, 2c, and 2d represent complex
numbers running along the edges of the quadrilateral. The only condition is that
the quadrilateral close up, i.e.,
a + b + c + d = 0.
We wish to show that A and B are perpendicular and of equal length. These two
statements can be combined into the single complex statement B = iA, which says
that B is A rotated by (π/2). To finish the proof, note that this is the same thing as
A + iB = 0, the verification of which is a routine calculation:
A + iB = (a + b + c + d) + i (a + b + c + d) = 0.
6
This approach is based on a paper of Finney (1970).
20 Geometry and Complex Arithmetic
[1.14] Composition of rotations. [a] In general, two successive rotations (of θ and ϕ)
about different points (a and b) are equivalent to a single rotation of (θ + ϕ) about a
special point c. [b] But if (θ + ϕ) = 2π then the net result is instead a translation, as can
easily be seen here in the special case θ = ϕ = π.
Rϕ
b ◦ Ra = Tv ,
θ
where v = (1 − eiϕ )(b − a).
by twice the complex number connecting the first centre of rotation to the second.
That this is indeed true can be deduced directly from [1.14b].
The above result on the composition of two rotations implies [exercise] the
following:
Returning to our original problem, we can now give an elegant geometric expla-
(π/2) (π/2)
nation of the result in [1.12b]. Referring to [1.15a], let M = Rπ m ◦ Rp ◦ Rs .
According to the result just obtained, M is a translation. To find out what transla-
tion, we need only discover the effect of M on a single point. Clearly, M(k) = k, so
M is the zero translation, i.e., the identity transformation E. Thus
−1
R(π/2)
p ◦ Rs(π/2) = (Rπ
m) ◦ M = Rπ
m.
If we define s ′ = Rπ ′
m (s) then m is the midpoint of ss . But, on the other hand,
s ′ = Rp(π/2)
◦ Rs(π/2) (s) = R(π/2)
p (s).
Thus the triangle sps ′ is isosceles and has a right angle at p, so sm and pm are
perpendicular and of equal length. Done.
22 Geometry and Complex Arithmetic
(π/2) (π/2)
[1.15] [a] The total angle of rotation of M = Rπ m ◦Rp ◦Rs is 2π, so it is a translation.
But we see that M(k) = k, so the translation vanishes—M is the identity. It follows [see
text] that sm and pm are perpendicular and of equal length. [b] The orbit Z(t) = eat eibt
is a spiral: as eibt rotates round the unit circle with angular speed b, it is stretched by the
rapidly increasing factor eat .
1.3.4 Calculus
For our calculus example, consider the problem of finding the 100th derivative of
ex sin x. More generally, we will show how complex numbers may be used to find
the nth derivative of eax sin bx.
In discussing Euler’s formula we saw that eit may be thought of as the location
at time t of a particle travelling around the unit circle at unit speed. In the same
way, eibt may be thought of as a unit complex number rotating about the origin
with (angular) speed b. If we stretch this unit complex number by eat as it turns,
then its tip describes the motion of a particle that is spiralling away from the origin.
See [1.15b].
The relevance of this to the opening problem is that the location of the particle
at time t is
Z(t) = eat eibt = eat cos bt + i eat sin bt.
Thus the derivative of eat sin bt is simply the vertical (imaginary) component of
the velocity V of Z.
We could find V simply by differentiating the components of Z in the above
expression, but we shall instead use this example to introduce the geometric
approach that will be used throughout this book. In [1.16], consider the movement
M = Z(t + δ) − Z(t) of the particle between time t and (t + δ).
Recall that V is defined to be the limit of (M/δ) as δ tends to zero. Thus V and
(M/δ) are very nearly equal if δ is very small. This suggests two intuitive ways of
Some Applications 23
speaking, both of which will be used in this book: (i) we shall say that “V = (M/δ)
when δ is infinitesimal” or (ii) that “V and (M/δ) are ultimately equal” (as δ tends to
zero).
As explained in the new Preface to this 25th Anniversary Edition, both the concept
and the language of ultimate equality were introduced by Newton in his great Prin-
cipia of 1687, which is filled with inspired geometrical constructions, not tedious
calculus computations. Sadly, although it was the mathematical engine that pro-
pelled his extraordinary discoveries, Newton chose to emulate the expository style
of the ancient Greek geometers, writing out his proofs in words: he did not intro-
duce any symbol to represent his critical concept of ultimate equality. We, however,
shall employ the symbol ” ≍ ” to denote Newton’s concept, enabling us to write,
for example, V ≍ (M/δ).
We stress that here the words “ultimately equal” and “infinitesimal” are being
used in definite, technical senses; in particular, “infinitesimal” does not refer to
some mystical, infinitely small quantity.7 More precisely, following Newton’s lead
7
For more on this distinction, see the discussion in Chandrasekhar (1995).
24 Geometry and Complex Arithmetic
Using (1.13), it is now easy to take further derivatives. For example, the acceler-
ation of the particle is
d2 d
2
Z= V = (a + ib)2 Z = (a + ib) V.
dt dt
Continuing in this way, each new derivative is obtained by multiplying the previ-
ous one by (a + ib). [Try sketching these successive derivatives in [1.16].] Writing
√
(a + ib) = R eiϕ , where R = a2 + b2 and ϕ is the appropriate value of tan−1 (b/a),
we therefore find that
dn
Z = (a + ib)n Z = Rn einϕ eat eibt = Rn eat ei(bt+nϕ) .
dtn
Thus
dn at n
n
e sin bt = (a2 + b2 ) 2 eat sin bt + n tan−1 (b/a) . (1.14)
dt
1.3.5 Algebra
In the final year of his life (1716) Roger Cotes made a remarkable discovery that
enabled him (in principle) to evaluate the family of integrals,
Z
dx
n
,
x −1
where n = 1, 2, 3, . . .. To see the connection with algebra, consider the case n = 2.
The key observations are that the denominator (x2 − 1) can be factorized into
(x − 1)(x + 1), and that the integrand can then be split into partial fractions:
Z Z
dx 1 1 1 1 x−1
=2 − dx = 2 ln .
x2 − 1 x−1 x+1 x+1
As we shall see, for higher values of n one cannot completely factorize (xn − 1)
into linear factors without employing complex numbers—a scarce and dubious
commodity in 1716! However, Cotes was aware that if he could break down (xn −1)
into real linear and quadratic factors, then he would be able to evaluate the inte-
gral. Here, a “real quadratic” refers to a quadratic whose coefficients are all real
numbers.
For example, (x4 − 1) can be broken down into (x − 1)(x + 1)(x2 + 1), yielding a
partial fraction expression of the form
1 A B Cx D
= + + + ,
x4 − 1 x − 1 x + 1 x2 + 1 x2 + 1
and hence an integral that can be evaluated in terms of ln and tan−1 . More generally,
even if the factorization involves more complicated quadratics than (x2 + 1), it is
easy to show that only ln and tan−1 are needed to evaluate the resulting integrals.
26 Geometry and Complex Arithmetic
in which c and z are complex. Just as in real algebra, this series may be summed by
noting that zGm−1 and cGm−1 contain almost the same terms—try an example, say
m = 4, if you have trouble seeing this. Subtracting these two expressions yields
(z − c)Gm−1 = zm − cm , (1.15)
and thus
zm − cm
Gm−1 = .
z−c
If we think of c as fixed and z as variable, then (zm − cm ) is an mth -degree
polynomial in z, and z = c is a root. The result (1.15) says that this mth -degree
polynomial can be factored into the product of the linear term (z − c) and the
(m − 1)th -degree polynomial Gm−1 .
In 1637 Descartes published an important generalization of this result. Let Pn (z)
denote a general polynomial of degree n:
Pn (z) = zn + Azn−1 + · · · + Dz + E,
If we do not acknowledge the existence of complex roots (as in the early 18th
century) then this factorization will be possible in some cases (e.g., z2 − 1), and
impossible in others (e.g., z2 + 1). But, in splendid contrast to this, if one admits
complex numbers then it can be shown that Pn always has n roots in C, and the fac-
torization (1.16) is always possible. This is called the Fundamental Theorem of Algebra,
and we shall explain its truth in Chapter 7.
Each factor (z−ck ) in (1.16) represents a complex number connecting the root ck
to the variable point z. Figure [1.17a] illustrates this for a general cubic polynomial.
Some Applications 27
[1.17] Geometric evaluation of polynomials. [a] If the roots of polynomial P are known,
then its value at z can easily be visualized: Connect the roots to z, then add the angles
and multiply the lengths to obtain the direction and length of P(z). [b] First example of
Cotes’s geometrical factorization of xn − 1: Here, x2 − 1 = PC1 · PC2 .
Writing each of these complex numbers in the form Rk eiϕk , (1.16) takes the more
vivid form
Pn (z) = R1 R2 · · · Rn ei(ϕ1 +ϕ2 +···+ϕn ) .
Although the Fundamental Theorem of Algebra was not available to Cotes, let
us see how it guarantees that he would succeed in his quest to decompose xn − 1
into real linear and quadratic factors. Cotes’ polynomial has real coefficients, and,
quite generally, we can show that
If a polynomial has real coefficients then its complex roots occur in complex
conjugate pairs, and it can be factorized into real linear and quadratic factors.
For if the coefficients A, . . . , E of Pn (z) are all real then Pn (c) = 0 implies [exercise]
Pn ( c ) = 0, and the factorization (1.16) contains
(z − c)(z − c) = z2 − (c + c) z + cc = z2 − 2 Re(c) z + |c|2 ,
which is a real quadratic.
Let us now discuss how Cotes was able to factorize xn − 1 into real linear and
quadratic factors by appealing to the geometry of the regular n-gon. [An “n-gon” is an
n-sided polygon.] To appreciate the following, place yourself in his 18th century
shoes and forget all you have just learnt concerning the Fundamental Theorem of
Algebra; even forget about complex numbers and the complex plane!
For the first few values of n, the desired factorizations of Un (x) = xn − 1 are not
too hard to find:
U2 (x) = (x − 1)(x + 1), (1.17)
U3 (x) = (x − 1)(x2 + x + 1), (1.18)
28 Geometry and Complex Arithmetic
the last two “factors” being analogous to genuine linear factors. If we are to inter-
pret this expression (by analogy with the previous case) as the product of the
distances of P from four fixed points, then the points corresponding to the last
two “factors” must be off the line. More precisely, Pythagoras’ Theorem tells us that
√
a point whose distance from P is x2 + 12 must lie at unit distance from O in a
direction at right angles to the line OP.
Referring to [1.18a], we can now see that U4 (x) = PC1 · PC2 · PC3 · PC4 , where
C1 C2 C3 C4 is the illustrated square inscribed in the circle.
Since we have factorized U4 (x) with the regular 4-gon (the square), perhaps we
can factorize U3 (x) with the regular 3-gon (the equilateral triangle). See [1.18b].
8
Here, and in what follows, we shall suppose for convenience that x > 1, so that Un (x) is positive.
Some Applications 29
[1.19] The nth roots of unity: Each vertex of the regular n-gon is a solution of zn = 1.
= (x − 1)(x2 + x + 1),
which is indeed the desired factorization (1.18) of U3 (x)!
A plausible generalization for Un now presents itself:
If C1 C2 C3 · · · Cn is a regular n-gon inscribed in a circle of unit radius centred at
O, and P is the point on OC1 at distance x from O, then Un (x) = PC1 · PC2 · · ·
PCn .
This is Cotes’ result. Unfortunately, he stated it without proof, and he left no clue as
to how he discovered it. Thus we can only speculate that he may have been guided
by an argument like the one we have just supplied9 .
Since the vertices of the regular n-gon will always come in symmetric pairs that
are equidistant from P, the examples in [1.18] make it clear that Cotes’ result is
indeed equivalent to factorizing Un (x) into real linear and quadratic factors.
Recovering from our feigned bout of amnesia concerning complex numbers and
their geometric interpretation, Cotes’ result becomes simple to understand and to
prove. Taking O to be the origin of the complex plane, and C1 to be 1, the vertices
of Cotes’ n-gon are given by Ck+1 = eik(2π/n) . See [1.19], which illustrates the case
n = 12. Since (Ck+1 )n = eik2π = 1, all is suddenly clear: The vertices of the regular
n-gon are the n complex roots of Un (z) = zn − 1. Because the solutions of zn − 1 = 0
√
may be written formally as z = n 1, the vertices of the n-gon are called the nth roots
of unity.
9
Stillwell (2010) has instead speculated that Cotes used complex numbers (as we are about to), but
then deliberately stated his findings in a form that did not require them.
30 Geometry and Complex Arithmetic
zn − 1 = Un (z) = (z − C1 )(z − C2 ) · · · (z − Cn ),
Although the dot and cross product are meaningful for arbitrary vectors in space,
we shall assume in the following that our vectors all lie in a single plane—the
complex plane.
Given two vectors a and b, figure [1.20a] recalls the definition of the dot product
as the length of one vector, times the projection onto that vector of the other vector:
· ·
a b = |a| |b| cos θ = b a,
[1.20] [a] The dot product is defined to be the orthogonal projection of one vector onto
the other, times the length of the other vector. [b] The cross product is defined to be the
vector orthogonal to both (oriented according to the “right hand rule”), with length equal
to the area of the parallelogram they span.
Figure [1.20b] recalls the definition of the cross product: a × b is the vector per-
pendicular to the plane of a and b whose length is equal to the area A of the
parallelogram spanned by a and b. But wait, there are two (opposite) directions
perpendicular to C; which should we choose?
Writing A = |a| |b| sin θ, the area A has a sign attached to it. An easy way to see
this sign is to think of the angle θ from a to b as lying in range −π to π; the sign
of A is then the same as θ. If A > 0, as in [1.20b], then we define a × b to point
upwards from the plane, and if A < 0 we define it to point downwards. It follows
that a × b = −(b × a).
This conventional definition of a × b is intrinsically three-dimensional, and it
therefore presents a problem: if a and b are thought of as complex numbers, a × b
cannot be, for it does not lie in the (complex) plane of a and b. No such problem
·
exists with the dot product because a b is simply a real number, and this suggests
a way out.
Since all our vectors will be lying in the same plane, their cross products will
all have equal (or opposite) directions, so the only distinction between one cross
product and another will be the value of A. For the purposes of this book we will
therefore redefine the cross product to be the (signed) area A of the parallelogram spanned
by a and b:
a × b = |a| |b| sin θ = −(b × a).
Figure [1.21] shows two complex numbers a = |a| eiα and b = |b| eiβ , the angle
from a to b being θ = (β − α). To see how their dot and cross products are related
to complex multiplication, consider the effect of multiplying each point in C by
a. This is a rotation of −α and an expansion of |a|, and if we look at the image
32 Geometry and Complex Arithmetic
[1.21] The dot and cross products in terms of complex multiplication. Multiplication of C
by a is a rotation of −α and an expansion of |a|. It follows that the dot and cross products
can be expressed as the real and imaginary parts of a b.
under this transformation of the shaded right triangle with hypotenuse b, then we
immediately see that
·
a b = a b + i (a × b). (1.20)
a b = (|a| e−iα )(|b| eiβ ) = |a| |b| ei(β−α) = |a| |b| eiθ = |a| |b|(cos θ + i sin θ).
We end with an example that illustrates the importance of the sign of the area
(a×b). Consider the problem of finding the area A of the quadrilateral in [1.22a],
whose vertices are, in counterclockwise order, a, b, c, and d. Clearly this is just the
sum of the ordinary, unsigned areas of the four triangles formed by joining the
Transformations and Euclidean Geometry* 33
[1.22] Area of a quadrilateral in terms of signed areas of triangles. [a] The area of the
quadrilateral is the sum of the areas of the four triangles, and is therefore given by (1.21).
[b] The area of the quadrilateral is still the sum of the areas of the four triangles, because
the striped area is automatically subtracted from the areas of the other three triangles.
vertices of the quadrilateral to the origin. Thus, since the area of each triangle is
simply half the area of the corresponding parallelogram,
A = 1
2 [(a×b) + (b×c) + (c×d) + (d×a)]
1
= 2 Im a b + b c + c d + d a . (1.21)
Obviously this formula could easily be generalized to polygons with more than
four sides.
But what if 0 is outside the quadrilateral? In [1.22b], A is clearly the sum of the
ordinary areas of three of the triangles, minus the ordinary area of the striped tri-
angle. Since the angle from b to c is negative, 12 (b×c) is automatically the negative
of the striped area, and A is therefore given by exactly the same formula as before!
Can you find a location for 0 that makes two of the signed areas negative? Check
that the formula still works. Exercise 35 shows that (1.21) always works.
10
The excellent book by Nikulin and Shafarevich (1987) is the only other work we know of in which
a similar attempt is made.
34 Geometry and Complex Arithmetic
As the * following the title of this section indicates, the material it contains
may be omitted. However, in addition to “explaining” complex numbers, these
ideas are very interesting in their own right, and they will also be needed for an
understanding of other optional sections of the book.
Although the ancient Greeks made many beautiful and remarkable discover-
ies in geometry, it was two thousand years later that Felix Klein first asked and
answered the question, “What is geometry?”
Let us restrict ourselves from the outset to plane geometry. One might begin by
saying that this is the study of geometric properties of geometric figures in the
plane, but what are (i) “geometric properties”, and (ii) “geometric figures”? We
will concentrate on (i), swiftly passing over (ii) by interpreting “geometric figure”
as anything we might choose to draw on an infinitely large piece of flat paper with
an infinitely fine pen.
As for (i), we begin by noting that if two figures (e.g., two triangles) have
the same geometric properties, then (from the point of view of geometry) they
must be the “same”, “equal”, or, as one usually says, congruent. Thus if we had
a clear definition of congruence (“geometric equality”) then we could reverse
this observation and define geometric properties as those properties that are common
to all congruent figures. How, then, can we tell if two figures are geometrically
equal?
Consider the triangles in [1.23], and imagine that they are pieces of paper that
you could pick up in your hand. To see if T is congruent to T ′ , you could pick up
T and check whether it could be placed on top of T ′ . Note that it is essential that
we be allowed to move T in space: in order to place T on top of Te we must first
flip it over; we can’t just slide T around within the plane. Tentatively generalizing,
this suggests that a figure F is congruent to another figure F ′ if there exists a motion of
F through space that makes it coincide with F ′ . Note that the discussion suggests that
there are two fundamentally different types of motion: those that involve flipping
the figure over, and those that do not. Later, we shall return to this important
point.
It is clearly somewhat unsatisfactory that in attempting to define geometry in
the plane we have appealed to the idea of motion through space. We now rectify
this. Returning to [1.23], imagine that T and T ′ are drawn on separate, transparent
sheets of plastic. Instead of picking up just the triangle T , we now pick up the entire
sheet on which it is drawn, then try to place it on the second sheet so as to make
T coincide with T ′ . At the end of this motion, each point A on T ’s sheet lies over
a point A ′ of T ′ ’s sheet, and we can now define the motion M to be this mapping
A ÞÑ A ′ = M(A) of the plane to itself.
However, not any old mapping qualifies as a motion, for we must also capture
the (previously implicit) idea of the sheet remaining rigid while it moves, so that
Transformations and Euclidean Geometry* 35
[1.23] The triangles T ′ and Te are both “geometrically equal” (“congruent”) to T because
there exists a “motion” that carries T to each of these. But the motion that carries T to Te
requires a flip.
distances between points remain constant during the motion. Here, then, is our
definition:
A “motion” M is a mapping of the plane to itself such that the distance
between any two points A and B is equal to the distance between their (1.22)
images A ′ = M(A) and B ′ = M(B).
Note that what we have called a motion is often termed a “rigid motion”, or an
“isometry”.
Armed with this precise concept of a motion, our final definition of geometric
equality becomes
∼ F ′ , if there exists a motion M such that
F is congruent to F ′ , written F =
(1.23)
F ′ = M(F).
Next, as a consequence of our earlier discussion, a geometric property of a figure
is one that is unaltered by all possible motions of the figure. Finally, in answer to the
opening question of “What is geometry?”, Klein would answer that it is the study
of these so-called invariants of the set of motions.
One of the most remarkable discoveries of the 19th century was that Euclidean
geometry is not the only possible geometry. Two of these so-called non-Euclidean
geometries will be studied in Chapter 6, but for the moment we wish only to
explain how Klein was able to generalize the above ideas so as to embrace such
new geometries.
The aim in (1.23) was to use a family of transformations to introduce a concept
∼
of geometric equality. But will this =-type of equality behave in the way we would
like and expect? To answer this we must first make these expectations explicit. So
as not to confuse this general discussion with the particular concept of congruence
in (1.23), let us denote geometric equality by ∼.
(i) The family G must contain a transformation E (called the identity) that maps
each point to itself.
(ii) If G contains a transformation M, then it must also contain a transformation
M−1 (called the inverse) that undoes M. [Check for yourself that for M−1 to
exist (let alone be a member of G) M must have the special properties of being
11
Here G is the group of projections. If we do a perspective drawing of figures in the plane, then the
mapping from that plane to the “canvas” plane is called a perspectivity. A projection is then defined to
be any sequence of perspectivities. Can you see why the set of projections should form a group?
Transformations and Euclidean Geometry* 37
(a) onto and (b) one-to-one, i.e., (a) every point must be the image of some
point, and (b) distinct points must have distinct images.]
(iii) If M and N are members of G then so is the composite transformation N◦M =
(M followed by N). This property of G is called closure.
12
In more abstract settings it is necessary to add a fourth requirement of associativity, namely,
A ◦ (B ◦ C) = (A ◦ B) ◦ C. Of course for transformations this is automatically true.
38 Geometry and Complex Arithmetic
diverse areas of mathematics and physics, and the insights it continues to provide
lie on the cutting edge of contemporary research.
[1.25] Each point P is uniquely determined by its distances from the vertices A, B, C of
a triangle.
13
This is how earthquakes are located. Two types of wave are emitted by the quake as it begins:
fast-moving “P-waves” of compression, and slower-moving “S-waves” of destructive shear. Thus the
P-waves will arrive at a seismic station before the S-waves, and the time-lag between these events may
be used to calculate the distance of the quake from that station. Repeating this calculation at two more
seismic stations, the quake may be located.
Transformations and Euclidean Geometry* 39
[1.26] A rigid motion is uniquely determined by its effect on any triangle: If M maps ABC
to A ′ B ′ C ′ , then an arbitrary point P must be mapped to the illustrated point P ′ = M(P).
A big step towards classification is the realization that there are two funda-
mentally different kinds of motions. In terms of our earlier conception of motion
through space, the distinction is whether or not a figure must be flipped over before
it can be placed on top of a congruent figure. To see how this dichotomy arises in
terms of the new definition (1.22), suppose that a motion sends two points A and
B to A ′ and B ′ . See [1.27]. According to (1.24), the motion is not yet determined:
we need to know the image of any (non-collinear) third point C, such as the one
shown in [1.27]. Since motions preserve the distances of C from A and B, there are
e in the line L
just two possibilities for the image of C, namely, C ′ and its reflection C
′ ′ e
through A and B . Thus there are precisely two motions (M and M, say) that map
e sends C to C.
A, B to A ′ , B ′ : M sends C to C ′ , and M e
e
A distinction can be made between M and M by looking at how they affect
angles. All motions preserve the magnitude of angles, but we see that M also pre-
serves the sense of the angle θ, while Me reverses it. The fundamental nature of this
distinction can be seen from the fact that M must in fact preserve all angles, while
Me must reverse all angles.
To see this, consider the fate of the angle ϕ in the triangle T . If C goes to C ′
(i.e., if the motion is M) then, carrying out the construction indicated in [1.26],
the image of T is T ′ , and the angle is preserved. If, on the other hand, C goes to
e then the image of T is the reflection Te of T ′ in L, and
e (i.e., if the motion is M)
C
the angle is reversed. Motions that preserve angles are called direct, and those that
reverse angles are called opposite. Thus rotations and translations are direct, while
reflections are opposite. Summarizing what we have found,
[1.28] Every direct motion is a rotation, or else (exceptionally) a translation. In the excep-
tional case that the motion M maps a line segment AB to a parallel segment A ′ B ′ , M is
either [a] a translation, or [b] a rotation of π about the intersection point of the lines AA ′
and BB ′ . However, in the general case [c], M is a rotation of θ about O, where θ and O
are constructed as shown.
The sense in which translations are “exceptional” is that if the two segments are
drawn at random then it is very unlikely that they will be parallel. Indeed, given
AB, a translation is only needed for one possible direction of A ′ B ′ out of infinitely
many, so the mathematical probability that a random direct motion is a translation
is actually zero!
Direct transformations will be more important to us than opposite ones, so we
relegate the investigation of opposite motions to Exs. 39, 40, 41. The reason for the
greater emphasis on direct motions stems from the fact that they form a group
(a subgroup of the full group of motions), while the opposite motions do not. Can
you see why?
Note that the second sentence of (1.25) then implies that every opposite motion is the
composition of three reflections. See Ex. 39. In brief, every motion is the composition
42 Geometry and Complex Arithmetic
of either two or three reflections, a result that is called the Three Reflections
Theorem14 .
Earlier we tried to show that the set of motions forms a group, but it was not clear
that every motion had an inverse. The Three Reflections Theorem settles this neatly
and explicitly, for the inverse of a sequence of reflections is obtained by reversing
the order in which the reflections are performed.
In what follows, let ℜL denote reflection in a line L. Thus reflection in L1 followed
by reflection in L2 is written ℜL2 ◦ ℜL1 . According to (1.26), proving (1.28) amounts
to showing that every rotation (and every translation) is of the form ℜL2 ◦ ℜL1 . This
is an immediate consequence of the following:
14
Results such as (1.26) may instead be viewed as consequences of this theorem; see Stillwell (1992)
for an elegant and elementary exposition of this approach.
Transformations and Euclidean Geometry* 43
[1.30] Geometric reduction of two rotations to a single rotation. [a] The composi-
tion of two rotations Rϕ b ◦ Ra can be expressed as the composition of four reflections:
θ
(ℜL2′ ◦ ℜL1′ ) ◦ (ℜL2 ◦ ℜL1 ). [b] But if we choose L2 = L1′ then the middle two reflections
cancel, and we discover that Rϕ b ◦ Ra = Rc
θ θ+ϕ
.
point c, as in [1.30b]. Thus the middle two reflections cancel, and the composition
of the two rotations is given by
15
In the non-Euclidean geometries of Chapter 6 we will be drawing on curved surfaces, and the
amount of curvature in the surface will dictate an absolute unit of length.
Transformations and Euclidean Geometry* 45
[1.31] Dilative rotations. [a] A central dilation Dro leaves o fixed and radially
stretches each segment oA by r. [b] If this central dilation is followed by (or pre-
ceded by) a rotation Rθo with the same centre, then we obtain the dilative rotation
Dr,θ
o ≡ Ro ◦ Do = Do ◦ Ro .
θ r r θ
o ≡ Ro ◦ Do = Do ◦ Ro ,
Dr,θ θ r r θ
shown in [1.31b]. Note that a central dilation may be viewed as a special case of a
dilative rotation: Dro = Dr,0
o .
This figure should be ringing loud bells. Taking o to be the origin of C, (1.9) says
that Dr,θ
o corresponds to multiplication by r e :
iθ
Dr,θ
o (z) = r e
iθ
z.
Conversely, and this is the key point, the rule for complex multiplication may be viewed
as a consequence of the behaviour of dilative rotations.
Concentrate on the set of dilative rotations with a common, fixed centre o, which
will be thought of as the origin of the complex plane. Each Dr,θ o is uniquely deter-
mined by its expansion r and rotation θ, and so it can be represented by a vector
46 Geometry and Complex Arithmetic
DR,ϕ
o ◦ Dr,θ
o = Do ◦ Do
r,θ R,ϕ
= DRr,(θ+ϕ)
o ,
so the new vector is obtained from the original vectors by multiplying their lengths
and adding their angles—complex multiplication!
On page 19 we saw that if complex numbers are viewed as translations then
composition yields complex addition. We now see that if they are instead viewed
as dilative rotations then composition yields complex multiplication. To complete
our “explanation” of complex numbers in terms of geometry, we will show that
these translations and dilative rotations are fundamental to Euclidean geometry as
defined in (1.29).
To understand the general similarity Sr involved in (1.29), note that if p is an
(1/r)
arbitrary point, M ≡ Sr ◦ Dp is a motion. Thus any similarity is the composition of
a dilation and a motion:
Sr = M ◦ Drp . (1.30)
Our classification of motions therefore implies that similarities come in two kinds:
if M preserves angles then so will Sr [a direct similarity]; if M reverses angles then
so will Sr [an opposite similarity].
Just as we concentrated on the group of direct motions, so we will now concen-
trate on the group of direct similarities. The fundamental role of translations and
dilative rotations in Euclidean geometry finally emerges in the following surprising
theorem:
For us, at least, this fact constitutes one satisfying “explanation” of complex num-
bers; as mentioned in the Preface, other equally compelling explanations may be
found in the laws of physics.
To begin to understand (1.31), observe that (1.25) and (1.30) imply that a direct
similarity is determined by the image A ′ B ′ of any line-segment AB. First consider
the exceptional case in which A ′ B ′ are of equal length AB. We then have the three
cases in [1.28], all of which are consistent with (1.31). If A ′ B ′ and AB are parallel
but not of equal length, then we have the two cases shown in [1.32a] and [1.32b], in
both of which we have drawn the lines AA ′ and BB ′ intersecting in p. By appealing
to the similar triangles in these figures, we see that in [1.32a] the similarity is Dr,0
p ,
while in [1.32b] it is Dpr,π
, where in both cases r = (pA ′ /pA) = (pB ′ /pB).
Now consider the much more interesting general case where A ′ B ′ and AB are
neither the same length, nor parallel. Take a peek at [1.32d], which illustrates this.
Here n is the intersection point of the two segments (produced if necessary), and
Transformations and Euclidean Geometry* 47
[1.32] Every direct similarity is a dilative rotation (with the sole exception of an isomet-
ric translation). If a direct similarity S carries a line segment AB to a parallel segment
A ′ B ′ of different length, then either [a] S = Dr,0 r,π
p , or [b] S = Dp , where in both cases
r = (pA /pA) = (pB /pB). [c] For a given θ, there exists a Dq mapping A to A ′ if and
′ ′ r,θ
θ is the angle between them. To establish (1.31), we must show that we can carry
AB to A ′ B ′ with a single dilative rotation. For the time being, simply note that if
AB is to end up having the same direction as A ′ B ′ then it must be rotated by θ, so
the claim is really this: There exists a point q, and an expansion factor r, such that Dr,θ
q
carries A to A ′ and B to B ′ .
Consider the part of [1.32d] that is reproduced in [1.32c]. Clearly, by choosing
r = (nA ′ /nA), Dn r,θ
will map A to A ′ . More generally, you see that we can map A
to A ′ with Dr,θ ′
q if and only if AA subtends angle θ at q. Thus, with the appropriate
value of r, Dq maps A to A ′ if and only if q lies on the circular arc AnA ′ . The figure
r,θ
BnB ′ . Thus there are just two possibilities: q = n or q = m (the other intersection
point of the two arcs). If you think about it, this is a moment of high drama. We
have narrowed down the possibilities for q to just two points by consideration of
angles alone; for either of these two points we can choose the value of the expansion
r so as to make A go to A ′ , but, once this choice has been made, either B will map
to B ′ or it won’t! Furthermore, it is clear from the figure that if q = n then B does
not map to B ′ , so q = m is the only possibility remaining.
In order for Dm r,θ
to simultaneously map A to A ′ and B to B ′ , we need to have
r = (mA ′ /mA) = (mB ′ /mB); in other words, the two shaded triangles need to be
similar. That they are indeed similar is surely something of a miracle. Looking at the
angles formed at n, we see that θ + ⊙ + • = π, and the result follows immediately
by thinking of the RHS as the angle-sum of each of the two shaded triangles. This
completes our proof16 of (1.31).
The reader may feel that it is unsatisfactory that (1.31) calls for dilative rotations
about arbitrary points, while complex numbers represent dilative rotations about
a fixed point o (the origin). This may be answered by noting that the images of
AB under Dr,θ r,θ
q and Do will be parallel and of equal length, so there will exist a
translation [see Ex. 44 for details] Tv mapping one onto the other. In other words, a
general dilative rotation differs from an origin-centred dilative rotation by a mere
translation: Dr,θ r,θ
q = Tv ◦ Do . To sum up,
16
The present argument has the advantage of proceeding in steps, rather than having to be discov-
ered all at once. For other proofs, see Coxeter and Greitzer (1967, p. 97), Coxeter (1969, p. 73), and Eves
(1992, p. 71). Also, see Ex. 45 for a simple proof using complex functions.
Transformations and Euclidean Geometry* 49
these translations yields ordinary vector addition in space. Note that this vector
addition makes equally good sense in four-dimensional space, or n-dimensional
space for that matter.
Now consider the set Q of dilative rotations with a common, fixed centre O. Ini-
tially, the definition of multiplication goes smoothly, for the “product” Q1 ◦ Q2
of two such dilative rotations is easily seen to be another dilative rotation (Q3 ,
say) of the same kind. This follows from the above classification of direct sim-
ilarities by noting that Q1 ◦ Q2 leaves O fixed. If the expansions of Q1 and Q2
are r1 and r2 then the expansion of Q3 is clearly r3 = r1 r2 , and in Chapter 6 we
shall give a simple geometric construction for the rotation of Q3 from the rota-
tions of Q1 and Q2 . However, unlike rotations in the plane, it makes a difference
in what order we perform two rotations in space, so our multiplication rule is not
commutative:
Q1 ◦ Q2 ̸= Q2 ◦ Q1 . (1.32)
17
To see this, imagine a sphere centred at O. The direction of the axis can be specified by its inter-
section with the sphere, and this point can be specified with two coordinates, e.g., longitude and
latitude.
50 Geometry and Complex Arithmetic
essentially the same result; only much later18 was it recognized that Rodrigues’
geometry was equivalent to Hamilton’s algebra.
Hamilton and Rodrigues are just two examples of hapless mathematicians who
would have been dismayed to examine the unpublished notebooks of the great
Carl Friedrich Gauss. There, like just another log entry in the chronicle of his private
mathematical voyages, Gauss recorded his discovery of the quaternion rule in 1819.
In Chapter 6 we shall investigate quaternion multiplication in detail and find
that it has elegant applications. However, the immediate benefit of this discus-
sion is that we can now see what a remarkable property it is of two-dimensional
space that it is possible to interpret points within it as the fundamental Euclidean
transformations acting on it.
18
See Altmann (1989) for the intriguing details of how this was unravelled.
Exercises 51
1.5 Exercises
1 The roots of a general cubic equation in X may be viewed (in the XY-plane) as
the intersections of the X-axis with the graph of a cubic of the form,
Y = X3 + AX2 + BX + C.
√
x = 2 p cos 13 (ϕ + 2mπ) ,
√
where m is an integer and ϕ = cos−1 (q/p p).
(iii) Check that this formula gives the correct solutions of x3 = 3x, namely,
√
x = 0, ± 3.
4 Here is a basic fact about integers that has many uses in number theory: If two
integers can be expressed as the sum of two squares, then so can their product. With
the understanding that each symbol denotes an integer, this says that if M =
a2 + b2 and N = c2 + d2 , then MN = p2 + q2 . Prove this result by considering
|(a + ib)(c + id)|2 .
52 Geometry and Complex Arithmetic
5 The figure below shows how two similar triangles may be used to construct
the product of two complex numbers. Explain this.
6 (i) If c is a fixed complex number, and R is a fixed real number, explain with a
picture why |z − c| = R is the equation of a circle.
(ii) Given that z satisfies the equation |z + 3 − 4i| = 2, find the minimum and
maximum values of |z|, and the corresponding positions of z.
7 Use a picture to show that if a and b are fixed complex numbers then |z − a| =
|z − b| is the equation of a line.
8 Let L be a straight line in C making an angle ϕ with the real axis, and let d be
its distance from the origin. Show geometrically that if z is any point on L then
d = Im[e−iϕ z] .
20 Gaussian integers are complex numbers of the form m + in, where m and n are
integers—they are the grid points in [1.1]. Show that it is impossible to draw an
equilateral triangle such that all three vertices are Gaussian integers. [Hints: You
may assume that one of the vertices is at the origin; try a proof by contradiction;
if a triangle is equilateral, you can rotate one side into another; remember that
√
3 is irrational.]
21 Make a copy of [1.12a], draw in the diagonal of the quadrilateral shown in
[1.12b], and mark its midpoint m. As in [1.12b], draw the line-segments con-
necting m to p, q, r, and s. According to the result in [1.12b], what happens
to p and to r under a rotation of (π/2) about m? So what happens to the
line-segment pr? Deduce the result shown in [1.12a].
22 Will the result in [1.12a] survive if the squares are instead constructed on the
inside of the quadrilateral?
23 Draw an arbitrary triangle, and on each side draw an equilateral triangle lying
outside the given triangle. What do you suspect is special about the new trian-
gle formed by joining the centroids (cf. Ex. 19) of the equilateral triangles? Use
complex algebra to prove that you are right. What happens if the equilateral
triangles are instead drawn on the inside of the given triangle?
24 From (1.15), we know that
zn − 1
1 + z + z2 + · · · + zn−1 = .
z−1
(i) In what region of C must z lie in order that the infinite series 1 + z + z2 + · · ·
converges?
(ii) If z lies in this region, to which point in the plane does the infinite series
converge?
(iii) In the spirit of figure [1.9], draw a large, accurate picture of the infinite
series in the case z = 12 (1 + i), and check that it does indeed converge to
the point predicted by part (ii).
25 Let S = cos θ + cos 3θ + cos 5θ + · · · + cos(2n − 1)θ. Show that
sin 2nθ sin nθ cos nθ
S= or equivalently S= .
2 sin θ sin θ
[Hint: Use Ex. 24, then Ex. 18 to simplify the result.]
26 (i) By considering (a + ib)(cos θ + i sin θ), show that
p
b cos θ + a sin θ = a2 + b2 sin θ + tan−1 (b/a) .
27 Show that the polar equation of the spiral Z(t) = eat eibt in [1.15b] is r =
e(a/b)θ .
28 Reconsider the spiral Z(t) = eat eibt in [1.15b], where a and b are fixed real
numbers. Let τ be a variable real number. According to (1.9), z ÞÑ Fτ (z) =
eaτ eibτ z is an expansion of the plane by factor eaτ , combined with a rotation
of the plane through angle bτ.
(i) Show that Fτ [Z(t)] = Z(t + τ), and deduce that the spiral is an invariant
curve (cf. p. 42) of the transformations Fτ .
(ii) Use this to give a calculus-free demonstration that all rays from the origin
cut the spiral at the same angle.
(iii) Show that if the spiral is rotated about the origin through an arbitrary
angle, the new spiral is again an invariant curve of each Fτ .
(iv) Argue that the spirals in the previous part are the only invariant curves of
Fτ .
29 (i) If V(t) is the complex velocity of a particle whose orbit is Z(t), and dt is
an infinitesimal moment of time, then V(t) dt is a complex number along
the orbit. Thinking of the integral as the (vector) sum of these movements,
Rt
what is the geometric interpretation of t12 V(t) dt?
1
(ii) Referring to [1.15b], sketch the curve Z(t) = a+ib eat eibt .
(iii) Given the result (1.13), what is the velocity of the particle in the previous
part.
R1
(iv) Combine the previous parts to deduce that 0 eat eibt dt =
1
at ibt 1
a+ib e e 0
, and draw in this complex number in your sketch
for part (ii).
(v) Use this to deduce that
Z1
a (ea cos b − 1) + b ea sin b
eat cos bt dt = ,
0 a2 + b2
and
Z1
b (1 − ea cos b) + a ea sin b
eat sin bt dt = .
0 a2 + b 2
30 Given two starting numbers S1 , S2 , let us build up an infinite sequence S1 , S2 ,
S3 , S4 , . . . with this rule: each new number is twice the difference of the previous two.
For example, if S1 = 1 and S2 = 4, we obtain 1, 4, 6, 4, −4, −16, −24, . . .. Our
aim is to find a formula for the nth number Sn .
(i) Our generating rule can be written succinctly as Sn+2 = 2(Sn+1 − Sn ).
Show that Sn = zn will solve this recurrence relation if z2 − 2z + 2 = 0.
56 Geometry and Complex Arithmetic
(ii) Use the quadratic formula to obtain z = 1 ± i, and show that if A and B
are arbitrary complex numbers, Sn = A(1 + i)n + B(1 − i)n is a solution
of the recurrence relation.
(iii) If we want only real solutions of the recurrence relation, show that B = A,
and deduce that Sn = 2 Re[A(1 + i)n ].
(iv) Show that for the above example A = −(1/2)
h − i, and by iwriting this in
√ (n+4)π
polar form deduce that Sn = 2 n/2
5 cos 4 + tan−1 2 .
(v) Check that this formula predicts S34 = 262144, and use a computer to
verify this.
[Note that this method can be applied to any recurrence relation of the form
Sn+2 = pSn+1 + qSn .]
31 With the same recurrence relation as in the previous exercise, use a computer
to generate the first 30 members of the sequence given by S1 = 2 and S2 = 4.
Note the repeating pattern of zeros.
(i) With the same notation as before, show that this sequence corresponds to
A = −i, so that Sn = 2 Re[−i(1 + i)n ].
(ii) Draw a sketch showing the locations of −i(1 + i)n for n = 1 to n = 8, and
hence explain the pattern of zeros.
(iii) Writing A = a + ib, our example corresponds to a = 0. More generally,
explain geometrically why such a repeating pattern of zeros will occur if
and only if (a/b) = 0, ±1 or b = 0.
(iv) Show that SS12 = 12 1 − a b , and deduce that a repeating pattern of zeros
will occur if and only if S2 = 2S1 (as in our example), S1 = S2 , S1 = 0, or
S2 = 0.
(v) Use a computer to verify these predictions.
32 The Binomial Theorem says that if n is a positive integer,
n
X
n n n!
(a + b)n = an−r br , where =
r r (n − r)! r!
r=0
are the binomial coefficients [not vectors!]. The algebraic reasoning leading to
this result is equally valid if a and b are complex numbers. Use this fact to show
that if n = 2m is even then
2m 2m 2m 2m
− + − · · · + (−1) m+1
= 2m sin mπ2 .
1 3 5 2m − 1
33 Consider the equation (z − 1)10 = z10 .
(i) Without attempting to solve the equation, show geometrically that all 9
solutions [why not 10?] must lie on the vertical line, Re (z) = 21 . [Hint:
Ex. 7.]
Exercises 57
(ii) Dividing both sides by z10 , the equation takes the form w10 = 1, where
w = (z − 1)/z. Hence solve the original equation.
(iii) Express these solutions in the form z = x+iy, and thereby verify the result
in (i). [Hint: To do this neatly, use Ex. 18.]
34 Let S denote the set of 12th roots of unity shown in [1.19], one of which is ξ =
ei(π/6) . Note that ξ is a primitive 12th root of unity, meaning that its powers yield
all the 12th roots of unity: S = {ξ, ξ2 , ξ3 , . . . , ξ12 }.
(i) Find all the primitive 12th roots of unity, and mark them on a copy of [1.19].
(ii) Write down, in the form of (1.16), the factorization of the polynomial
Φ12 (z) whose roots are the primitive 12th roots of unity. [In general, Φn (z)
is the polynomial (with the coefficient of the highest power of z equal
to 1) whose roots are the primitive nth roots of unity; it is called the nth
cyclotomic polynomial.]
(iii) By first multiplying out pairs of factors corresponding to conjugate roots,
show that Φ12 (z) = z4 − z2 + 1.
(iv) By repeating the above steps, show that Φ8 (z) = z4 + 1.
(v) For a general value of n, explain the fact that if ζ is a primitive nth root
of unity, then so is ζ. Deduce that if n > 2 then Φn (z) always has even
degree and real coefficients.
(vi) Show that if p is a prime number then Φp (z) = 1 + z + z2 + · · · + zp−1 .
[Hint: Ex. 24.]
[In these examples it is striking that Φn (z) has integer coefficients. In fact it
can be shown that this is true for every Φn (z)! For more on these fascinating
polynomials, see Stillwell (1994).]
35 Show algebraically that the formula (1.21) is invariant under a translation by k,
i.e., its value does not change if a becomes a + k, b becomes b + k, etc. Deduce
from [1.22a] that the formula always gives the area of the quadrilateral. [Hint:
Remember, (z + z) is always real.]
(θ+ϕ)
36 According to the calculation on p. 20, Rϕ
b ◦ Ra = Rc
θ
, where
aeiϕ (1 − eiθ ) + b(1 − eiϕ )
c= .
1 − ei(θ+ϕ)
Let us check that this c is the same as the one given by the geometric construc-
tion in [1.30b].
(i) Explain why the geometric construction is equivalent to saying that c
satisfies the two conditions
c−b 1 c−a
arg = 2 ϕ and arg = − 12 θ.
a−b b−a
58 Geometry and Complex Arithmetic
(ii) Verify that the calculated value of c (given above) satisfies the first of these
conditions by showing that
" #
c−b sin θ2
= (θ+ϕ)
eiϕ/2 . (1.33)
a−b sin 2
iα
[Hint: Use 1 − e = −2i sin(α/2) eiα/2 .]
(iii) In the same way, verify that the second condition is also satisfied.
37 Deduce (1.33) directly from [1.30b]. [Hint: Draw in the altitude through b of
the triangle abc, and express its length first in terms of sin θ2 , then in terms of
sin (θ+ϕ)
2 .]
38 On page 20 we calculated that for any non-zero α, Tv ◦ Rα
0 is a rotation:
Tv ◦ Rα
0 = Rc , where c = v/(1 − e ).
α iα
However, if α = 0 then Tv ◦ Rα
0 = Tv is a translation. Try to reconcile these facts
by considering the behaviour of Rαc in the limit that α tends to zero.
(v) Hence describe (in geometric terms) the glide reflection represented by
G(z) = i z + 4i. Check your answer by looking at the images of −2, 2i,
and 0.
e
41 Let M(z) be the representation of a general opposite motion as a complex
function.
e
(i) Explain why M(z) e
is a direct motion, and deduce from (1.27) that M(z) =
iα
e z + w, for some α and w.
(ii) Using the previous exercise, deduce that every opposite motion is a glide
reflection.
42 On p. 20 we calculated that if (θ + ϕ) = 2π then
Rϕ
b ◦ Ra = Tv ,
θ
where v = (1 − eiϕ )(b − a).
(i) Let Q = (b − a) be the complex number from the first centre of rotation to
the second. Show algebraically that v has length 2 sin(θ/2) |Q|, and that its
direction makes an angle of π−θ
2 with Q.
(ii) Give direct geometric proofs of these results by redrawing figure [1.30b] in
the case (θ + ϕ) = 2π.
43 On p. 20 we calculated that
Tv ◦ Rα
0 = Rc ,
α
where c = v/(1 − eiα ).
(i) Show algebraically that the complex number from the old centre of rotation
|v|
(the origin) to the new centre of rotation (c) has length 2 sin(α/2) , and that
π−α
its direction makes an angle of 2 with v.
(ii) Representing both Rα 0 and Tv as the composition of two reflections, use the
idea in [1.30b] to give direct, geometric proofs of these results.
44 Just as in [1.13b], a dilative rotation Dr,θ
p centred at an arbitrary point p may be
performed by translating p to the origin, doing Dr,θo , then translating o back to
p. Representing these transformations as complex functions, show that
Dr,θ iθ
p (z) = r e z + v, where v = p(1 − r eiθ ).
Conversely, if v is given, deduce that
Tv ◦ Dr,θ
o = Dp ,
r,θ
where p = v/(1 − r eiθ ).
45 In the previous exercise you showed that an arbitrary dilative rotation or trans-
lation can be written as a complex function of the form f(z) = az + b, and,
conversely, that every such function represents a unique dilative rotation or
translation.
(i) Given two pairs of distinct points {A, B} and {A ′ , B ′ }, show [by finding
them explicitly] that a and b exist such that f(A) = A ′ and f(B) = B ′ .
(ii) Deduce the result (1.31).
CHAPTER 2
2.1 Introduction
Of course we may also describe the w-plane with polar coordinates so that
w = f(z) = R eiϕ , where R(z) and ϕ(z) are real functions of z. With the same
example as before, the transformation becomes
Visual Complex Analysis. 25th Anniversary Edition. Tristan Needham, Oxford University Press.
© Tristan Needham (2023). DOI: 10.1093/oso/9780192868916.003.0002
62 Complex Functions as Transformations
[2.1] A complex function f(z) is most commonly viewed as a mapping of points in one
copy of C (the “z-plane”)
√ to points in another copy of C (the “w-plane”): z ÞÑ w = f(z).
Here, f(z) = (1 + i 3)z = 2 eiπ/3 z is the combined effect of an expansion by factor 2
and a rotation by angle (π/3). Most commonly, we write z = x + iy and w = u + iv.
We shall find that we can gain considerable insight into a given f by drawing
pictures showing its effect on points, curves, and shapes. However, it would be
nice if we could simultaneously grasp the behaviour of f for all values of z. One
such method is to instead represent f as a vector field, whereby f(z) is depicted as a
vector emanating from the point z; for more detail, the reader is invited to read the
beginning of Chapter 10.
Yet other methods are based on the idea of a graph. In the case of a real function
f(x) of a real variable x we are accustomed to the convenience of visualizing the
overall behaviour of f by means of its graph, i.e., the curve in the two-dimensional
xy-plane made up of the points (x, f(x)). In the case of a complex function this
approach does not seem viable because to depict the pair of complex numbers
(z, f(z)) we would need four dimensions: two for z = x+iy and two for f(z) = u+iv.
Actually, the situation is not quite as hopeless as it seems. First, note that
although two-dimensional space is needed to draw the graph of a real function f,
the graph itself [the set of points (x, f(x))] is only a one-dimensional curve, mean-
ing that only one real number (namely x) is needed to identify each point within it.
Likewise, although four-dimensional space is needed to draw the set of points with
coordinates (x, y, u, v) = (z, f(z)), the graph itself is two-dimensional, meaning that
only two real numbers (namely x and y) are needed to identify each point within
it. Thus, intrinsically, the graph of a complex function is merely a two-dimensional
surface (a so-called Riemann surface), and it is thus susceptible to visualization in
ordinary three-dimensional space. This approach will not be explored in this book,
though the last three chapters in particular should prove helpful in understand-
ing Riemann’s original physical insights, as expounded by Klein (1881). See also
Introduction 63
Springer (1981, Ch. 1), which essentially reproduces Klein’s monograph, but with
additional helpful commentary.
There is another type of graph of a complex function that is sometimes useful.
The image f(z) of a point z may be described by its distance |f(z)| from the origin,
and the angle arg[f(z)] it makes with the real axis. Let us discard half of this infor-
mation (the angle) and try to depict how the modulus |f(z)| varies with z. To do so,
imagine the complex z-plane lying horizontally in space, and construct a point at
height |f(z)| vertically above each point z in the plane, thereby producing a surface
called the modular surface of f . Figure [2.2] illustrates the conical modular surface of
f(z) = z, while [2.3] illustrates the paraboloid modular surface of f(z) = z2 .
A note on computers. Beginning in this chapter, we will often suggest that you
use a computer to expand your understanding of the mathematical phenomenon
[2.2] The height above C of the modular surface of a complex function f(z) is its modulus,
|f(z)|. Here, f(z) = z, producing a cone.
under discussion. However, we wish to stress that the specific uses of the computer
that we have suggested in the text are only a beginning. Think of the computer as a
physicist would his laboratory—you may use it to check your existing ideas about
the construction of the world, or as a tool for discovering new phenomena which
then demand new ideas for their explanation. In the Preface we make concrete
suggestions (probably of only fleeting relevance) as to how your laboratory should
be equipped.
2.2 Polynomials
[2.5] As z rotates around the unit circle, starting at 1, its image under z ÞÑ z3 rotates
three times as fast, executing a complete revolution every time z executes one third of
a revolution. Thus the cube roots of 1 form an equilateral triangle. More generally, the
same reasoning explains why the nth roots of unity form a regular n-gon.
[2.6] Extending the reasoning of the previous figure, the cube roots of an arbitrary complex
number c will form an equilateral triangle, and the nth roots will form a regular n-gon.
of a regular n-gon. Thus the complete set of solutions will be obtained if we let m
take any n consecutive values, say m = 0, 1, 2, . . . , (n − 1).
[2.7] The solution of the cubic x3 = 3px + 2q was long thought to have two com-
p solutions. If q ⩽ p , Cardano’s 1545 formula says that x = s + s, where
2 3
pletely different
3 3 2
s = q+i p
√ 1 − q . On the
other hand, Viète’s
−1
1591 “angle trisection method” yields
√
x = 2 p cos 3 (ϕ + 2mπ) , where ϕ = cos (q/p p). The figure reveals that the two
methods are in fact the same.
of the three values of s must therefore be paired with the conjugate value of t. We
can now see how Cardano’s formula becomes Viète’s formula:
√
xm = sm + tm = sm + sm = 2 p cos 13 (ϕ + 2mπ) .
In Ex. 4 the reader is invited to consider the case q2 > p3 .
[2.8] [a] Ellipses with foci a1 and a2 have equation r1 + r2 = const. In 1687 Newton
proved that the planets orbit the sun in such ellipses, with the sun at one focus. [b] Seven
years earlier, however, Giovanni Cassini had instead proposed that they orbit the sun in
what are now called Cassinian curves, with equation r1 r2 = const. The figure eight curve
is called the lemniscate, and it played a pivotal role in understanding elliptic integrals and
elliptic functions.
These curves are illustrated in [2.8b]; they are called Cassinian curves, and the
points a1 and a2 are again called foci.
The following facts will become clearer in a moment, but you might like to think
about them for yourself. If k is small then the curve consists of two separate pieces,
resembling small circles centred at a1 and a2 . As k increases, these two compo-
nents of the curve become more egg shaped. When k reaches a value equal to half
the distance between the foci then the pointed ends of the egg shapes meet at the
midpoint of the foci, producing a figure eight [shown solid]. Increasing the value
of k still further, the curve first resembles an hourglass, then an ellipse, and finally
a circle.
Although Cassinian curves turned out to be useless as a description of planetary
motion, the figure eight curve proved extremely valuable in quite another context.
In 1694 it was rediscovered by James Bernoulli and christened the lemniscate—it
then became the catalyst in unravelling the behaviour of the so-called elliptic inte-
grals and elliptic functions. See Stillwell (2010) and Siegel (1969) for more on this
fascinating story.
Cassinian curves arise naturally in the context of complex polynomials. A gen-
eral quadratic Q(z) = z2 + pz + q will have two roots (say, a1 and a2 ) and so can
be factorized as Q(z) = (z − a1 )(z − a2 ). In terms of [2.8b], this becomes
[2.9] On the left, the Cassinian curves with foci ±1 have equation |(z−1)(z+1)| = const.
It follows that their images under z ÞÑ z2 are the circles illustrated on the right, centred at
1. It then follows immediately that the equation of the lemniscate is r2 = 2 cos 2θ.
70 Complex Functions as Transformations
[2.10] The form of the Cassinian curves in [2.8b] may be grasped intuitively as horizontal
sections of the modular surface of Q(z) = (z − a1 )(z − a2 ).
z ÞÑ w = Pn (z) = (z − a1 )(z − a2 ) · · · (z − an ).
Equivalently, the Cassinian curves are the cross-sections of the modular surface of
Pn (z). This surface has n cone-like legs resting on C at a1 , a2 , . . . , an , and for large
values of |z| it resembles the axially symmetric modular surface of zn .
Power Series 71
[2.11] The Spiric Sections of Perseus (150 BCE). The intersection of a torus with a plane
whose distance from the axis of symmetry equals the radius of the generating circle turns
out to be a Cassinian curve. If the plane touches the inner rim, the lemniscate makes a
surprise appearance!
where the cj ’s are real constants. Of course, this infinite series will normally only
converge to F(x) in some origin-centred interval of convergence −R < x < R. But how
is R (the radius of convergence) determined by F(x)?
It turns out that this question has a beautifully simple answer, but only if
we investigate it in the complex plane. If we instead restrict ourselves to the real
line—as mathematicians were forced to in the era in which such series were first
employed—then the relationship between R and F(x) is utterly mysterious. Histori-
cally, it was precisely this mystery1 that led Cauchy to several of his breakthroughs
in complex analysis.
1
Cauchy was investigating the convergence of series solutions to Kepler’s equation, which describes
where a planet is in its orbit at any given time.
72 Complex Functions as Transformations
To see that there is a mystery, consider the power series representations of the
functions
1 1
G(x) = and H(x) = .
1 − x2 1 + x2
The familiar infinite geometric series,
X ∞
1
= xj = 1 + x + x2 + x3 + · · · if and only if − 1 < x < 1, (2.3)
1−x
j=0
immediately yields
∞
X ∞
X
G(x) = x2j and H(x) = (−1)j x2j ,
j=0 j=0
where both series have the same interval of convergence, −1 < x < 1.
It is easy to understand the interval of convergence of the series for G(x) if
we look at the graph [2.12a]. The series becomes divergent at x = ±1 because
these points are singularities of the function itself, i.e., they are places where |G(x)|
becomes infinite. But if we look at y = |H(x)| in [2.12b], there seems to be no reason
for the series to break down at x = ±1. Yet break down it does.
To begin to understand this, let us expand these functions into power series cen-
P
tred at x = k (instead of x = 0), i.e., into series of the form ∞ j
j=0 cj X , where
X = (x − k) measures the displacement of x from the centre k. To expand G we first
generalize (2.3) by expanding 1/(a − x) about k:
1 1 1 1
= = X
,
a−x a − (X + k) (a − k) 1 − a−k
1
[2.12] [a] The convergence of the power series for G(x) = 1−x 2 is readily understood:
clearly it must end when we arrive at the singularities x = ±1. [b] In contrast to this,
2 also stops converging at ±1 is utterly
1
the fact that the power series for H(x) = 1+x
mysterious!
Power Series 73
[2.13] The radius of convergence is the distance to the nearest singularity. [a] It is intu-
1
itively clear that the radius of convergence R of the power series for G(x) = 1−x 2 centred
at k is√the distance from k to the nearest singularity. [b] The previously mysterious formula
1
R = 12 + k2 for the convergence of the real power series for H(x) = 1+x 2 is explained
and so
X ∞
1 Xj
= , if and only if |X| < |a − k|. (2.4)
a−x (a − k)j+1
j=0
To apply this result to G, we factorize (1−x2 ) = (1−x) (1+x) and then decompose
G into partial fractions:
X∞
1 1 1 1 1 1 1
=2 − =2 − Xj ,
1 − x2 1 − x −1 − x (1 − k)j+1 (−1 − k)j+1
j=0
where |X| < |1 − k| and |X| < |1 + k|. Thus the interval of convergence |X| < R is
given by
R = min {|1 − k|, |1 + k|} = (distance from k to the nearest singularity of G).
This readily comprehensible result is illustrated in [2.13a]; ignore the shaded disc
for the time being.
In the case of H(x), I cannot think of an elegant method of finding the expansion
using only real numbers, but see Ex. 9 for an attempt. Be that as it may, it can be
shown that the radius of convergence of the series in X is given by the strange
√
formula R = 1 + k2 . As with Cotes’ work in the previous chapter, we have here a
result about real functions that is trying to tell us about the existence of the complex
plane.
If we picture the real line as embedded in a plane then Pythagoras’ Theorem
√
tells us that R = 12 + k2 should be interpreted as the distance from the centre k
74 Complex Functions as Transformations
of the expansion to either of the fixed points that lie off the line, one unit from 0 in a
direction at right angles to the line. See [2.13b]. If the plane is thought of as C, then
these points are ±i, and
R = (distance from k to ±i).
The mystery begins to unravel when we turn to the complex function
1
h(z) = ,
1 + z2
which is identical to H(x) when z is restricted to the real axis of the complex plane.
In fact there is a sense—we cannot be explicit yet—in which h(z) is the only complex
function that agrees with H on this line.
While [2.12b] shows that h(z) is well-behaved for real values of z, it is clear that
h(z) has two singularities in the complex plane, one at z = i and the other at z = −i;
these are shown as little explosions in [2.13b]. Figure [2.14] tries to make this more
vivid by showing the modular surface of h(z), the singularities at ±i appearing
as “volcanoes” erupting above these points. We will sort through the details in a
moment, but the mystery has all but disappeared: in both [2.13a] and [2.13b], the
radius of convergence is the distance to the nearest singularity.
If we intersect the surface in [2.14] with a vertical plane through the real axis
then we recover the deceptively tranquil graph in [2.12b], but if we instead slice
1
[2.14] Twin volcanoes erupt in the modular surface of h(z) = 1+z 2 , directly above its
singularities at z = ±i. The graphs in [2.12] can now be understood in a unified way, as
the vertical sections of this surface in the imaginary and real directions. The mysterious
radius of convergence of the real power series for h(x) is thereby explained!
Power Series 75
the surface along the imaginary axis then we obtain the graph in [2.12a]. That this
is no accident may be seen by first noting that G(x) is just the restriction to the real
axis of the complex function g(z) = 1/(1 − z2 ). Since g(z) = h(iz), h and g are
essentially the same: if we rotate the plane by (π/2) and then do h, we obtain g. In
particular the modular surface of g is simply [2.14] rotated by (π/2), the volcanoes
at ±i being rotated to ±1.
where the cj ’s are complex constants, and z is a complex variable. The partial sums
of this infinite series are just the ordinary polynomials,
X
n
Pn (z) = cj zj = c0 + c1 z + c2 z2 + c3 z3 + · · · + cn zn .
j=0
For a given value of z = a, the sequence of points P1 (a), P2 (a), P3 (a), . . . is said to
converge to the point A if for any given positive number ϵ, no matter how small, there
exists a positive integer N such that |A − Pn (a)| < ϵ for every value of n greater
than N. Figure [2.15a] illustrates that this is much simpler than it sounds: all it says
is that once we reach a certain point PN (a) in the sequence P1 (a), P2 (a), P3 (a), . . .,
all of the subsequent points lie within an arbitrarily small disc of radius ϵ centred
at A.
In this case we say that the power series P(z) converges to A at z = a, and we write
P(a) = A. If the sequence P1 (a), P2 (a), P3 (a), . . . does not converge to a particular
point, then the power series P(z) is said to diverge at z = a. Thus for each point z,
P(z) will either converge or diverge.
Figure [2.15b] shows a magnified view of the disc in [2.15a]. If n > m > N then
Pm (a) and Pn (a) both lie within this disc, and consequently the distance between
them must be less than the diameter of the disc:
|cm+1 am+1 + cm+2 am+2 + · · · + cn an | = |Pn (a) − Pm (a)| < 2ϵ. (2.6)
Conversely, it can be shown [exercise] that if this condition is met then P(a) con-
verges. Thus we have a new way of phrasing the definition of convergence: P(a)
converges if and only if there exists an N such that inequality (2.6) holds (for arbitrarily
small ϵ) whenever m and n are both greater than N.
76 Complex Functions as Transformations
[2.15] [a] The sequence Pn (a) is said to converge to A if it inevitably enters every disc
centred at A, no matter how small we make its radius ϵ, and the disc then acts like a
black hole: having entered the disc, the sequence can never escape it! [b] The direct
route from Pm (a) to Pn (a) has length |Pn (a) − Pm (a)| and is shorter than the indirect
en (a) − P
route that goes via Pm+1 (a), Pm+2 (a), etc., which has length P em (a). But if P(a)
is absolutely convergent then the indirect route must have a length that tends to zero for
sufficiently large m and n, and therefore the direct route must, too. Therefore, absolute
convergence implies convergence.
en (a) − P
P em (a) = |cm+1 am+1 | + |cm+2 am+2 | + · · · + |cn an |
is the total length of the roundabout journey from Pm (a) to Pn (a) that goes via
Pm+1 (a), Pm+2 (a), etc. Since |Pn (a) − Pm (a)| is the length of the shortest journey
from Pm (a) to Pn (a),
Power Series 77
[2.16] [a] If P(z) converges at a then it also converges everywhere inside the shaded disc
|z| < |a|. And if it diverges at d then it also diverges in the striped region |z| > |d|. Its
behaviour is therefore unknown only in “the ring of doubt”, |a| ⩽ |z| ⩽ |d|. Testing a point
q half way across this ring then halves the width of the ring. [b] Iterating this q-algorithm,
the ring of doubt shrinks and converges to a definite circle, the circle of convergence, the
interior of which is the disc of convergence, and the radius of which, R, is the radius of
convergence.
en (a) − P
|Pn (a) − Pm (a)| ⩽ P em (a).
Thus |Pn (a) − Pm (a)| must also become arbitrarily small for sufficiently large m
and n. Done.
We can now establish the following fundamental fact:
If P(z) converges at z = a, then it will also converge everywhere inside
(2.8)
the disc |z| < |a|.
See [2.16a]. In fact we will show that P(z) is absolutely convergent in this disc; the
result then follows directly from (2.7).
If P(a) converges then the length |cn an | of each term must die away to zero
as n goes to infinity [why?]. In particular, there must be a number M such that
|cn an | < M for all n. If |z| < |a| then ρ = |z|/|a| < 1 and so |cn zn | < Mρn . Thus,
around in the disc of convergence. Clearly, since Em (0) = 0, the error will be
extremely small if z is close to the origin, but what if z approaches the circle of
convergence? The answer depends on the particular power series, but it can hap-
pen that the error becomes enormous! [See Ex. 12.] This does not contradict the above
result: for any fixed z, no matter how close to the circle of convergence, the error
Em (z) will become arbitrarily small as m tends to infinity.
This problem is avoided if we restrict z to the disc |z| ⩽ r, where r < R, because
this prevents z from getting arbitrarily close to the circle of convergence, |z| = R.
In attempting to approximate P(z) within this disc, it turns out that we can do the
following. We first decide on the maximum error (say ϵ) that we are willing to
put up with, then choose (once and for all) an approximating polynomial Pm (z) of
sufficiently high degree that the error is smaller than ϵ throughout the disc. That
is, throughout the disc, the approximating point Pm (z) lies less than ϵ away from
the true point, P(z). One describes this by saying that P(z) is uniformly convergent
on this disc:
If P(z) has disc of convergence |z| < R, then P(z) is uniformly convergent
(2.12)
on the closed disc |z| ⩽ r, where r < R.
Although we may not have uniform convergence on the whole disc of conver-
gence, the above result shows that this is really a technicality: we do have uniform
convergence on a disc that almost fills the complete disc of convergence, say
r = (0.999999999) R.
To verify (2.12), first do Ex. 12, then have a good look at (2.9).
2.3.4 Uniqueness
If a complex function can be expressed as a power series, then it can only be done
so in one way—the power series is unique. This is an immediate consequence of the
Identity Theorem:
If
c0 + c1 z + c2 z2 + c3 z3 + · · · = d0 + d1 z + d2 z2 + d3 z3 + · · ·
for all z in a neighbourhood (no matter how small) of 0, then the power
series are identical: cj = dj .
Putting z = 0 yields c0 = d0 , so they may be cancelled from both sides. Dividing by
z and again putting z = 0 then yields c1 = d1 , and so on. [Although this was easy,
Ex. 13 shows that it is actually rather remarkable.] The result can be strengthened
considerably: If the power series merely agree along a segment of curve (no matter how
small) through 0, or if they agree at every point of an infinite sequence of points that con-
verges to 0, then the series are identical. The verification is essentially the same, only
instead of putting z = 0, we now take the limit as z approaches 0, either along the
segment of curve or through the sequence of points.
80 Complex Functions as Transformations
We can perhaps make greater intuitive sense of these results if we first recall that
a power series can be approximated with arbitrarily high precision by a polynomial
of sufficiently high degree. Given two points in the plane (no matter how close
together) there is a unique line passing through them. Thinking in terms of a graph
y = f(x), this says that a polynomial of degree 1, say f(x) = c0 + c1 x, is uniquely
determined by the images of any two points, no matter how close together. Like-
wise, in the case of degree 2, if we are given three points (no matter how close
together), there is only one parabolic graph y = f(x) = c0 + c1 x + c2 x2 that can be
threaded through them. This idea easily extends to complex functions: there is one,
and only one, complex polynomial of degree n that maps a given set of (n + 1) points to
a given set of (n + 1) image points. The above result may therefore be thought of as
the limiting case in which the number of known points (together with their known
image points) tends to infinity.
Earlier we alluded to a sense in which h(z) = 1/(1 + z2 ) is the only complex
function that agrees with the real function H(x) = 1/(1 + x2 ) on the real line. Yet
clearly we can easily write down infinitely many complex functions that agree with
H(x) in this way. For example,
cos[x2 y] + i sin[y2 ]
g(z) = g(x + iy) = .
ey + x2 ln(e + y4 )
Then in what sense can h(z) be considered the unique generalization of H(x)?
P
We already know that h(z) can be expressed as the power series ∞ j 2j
j=0 (−1) z ,
and this fact yields [exercise] a provisional answer: h(z) is the only complex func-
tion that (i) agrees with H(x) on the real axis, and (ii) can be expressed as a power
series in z. This still does not completely capture the sense in which h(z) is unique,
but it’s a start.
More generally, suppose we are given a real function F(x) that can be expressed
as a power series in x on a (necessarily origin-centred) segment of the real line:
P P∞
F(x) = ∞ j
j=0 cj x . Then the complex power series f(z) =
j
j=0 cj z with the same
coefficients can be used to define the unique complex function f(z) that
(i) agrees with F on the given segment of the real axis, and (ii) can be expressed as
a power series in z.
For example, consider the complex exponential function, written ez , the geometry
P
of which we will discuss in the next section. Since ex = ∞ j
j=0 x /j!,
ez = 1 + z + 1 2
2! z + 1 3
3! z + 1 4
4! z + ··· .
Two power series with the same centre can be added, multiplied, and divided
(2.13)
in the same way as polynomials.
If the two series P(z) and Q(z) have discs of convergence D1 and D2 , then the result-
ing series for [P +Q] and PQ will both converge in the smaller of D1 and D2 , though
they may in fact converge within a still larger disc. No such general statement is
possible in the case of (P/Q) = P(1/Q), because the convergence of the series for
(1/Q) is limited not only by the boundary circle of D2 , but also by any points inside
D2 where Q(z) = 0.
Let us illustrate (2.13) with a few examples. Earlier we actually assumed this
result in order to find the series for 1/(1 − z2 ) centred at k. Using the partial fraction
decomposition
1 (1/2) (1/2)
2
= + ,
1−z 1−z 1+z
we obtained two power series for the functions on the RHS, and then assumed that
these power series could be added like two polynomials, by adding the coefficients.
In the special case k = 0 we can check that this procedure works, because we
already know the correct answer for the series centred at the origin:
1
= 1 + z2 + z 4 + z6 + · · · .
1 − z2
Since
1
= 1 + z + z 2 + z3 + z4 + z 5 + · · ·
1−z
1
and = 1 − z + z2 − z3 + z 4 − z5 + · · · ,
1+z
we see that adding the coefficients of these series does indeed yield the correct
series for 1/(1 − z2 ).
Since
1 1 1
= ,
1 − z2 1−z 1+z
we can recycle this example to illustrate the correctness of multiplying power series
as if they were polynomials:
[1 + z + z2 + z3 + z4 + z5 + · · · ][1 − z + z2 − z3 + z4 − z5 + · · · ]
= 1 + (1–1) z + (1–1+1) z2 + (1–1+1–1) z3 + (1–1+1–1+1) z4 + · · · ,
Next, let’s use (2.13) to find the series for 1/(1 − z)2 :
[1 + z + z2 + z3 + z4 + z5 + · · · ][1 + z + z2 + z3 + z4 + z5 + · · · ]
= 1 + (1+1) z + (1+1+1) z2 + (1+1+1+1) z3 + (1+1+1+1+1) z4 + · · · ,
P
and so (1 − z)−2 = ∞ j
j=0 (j + 1) z .
You may check for yourself that the above series for (1 − z)−1 and (1 − z)−2 are
both special cases of the general Binomial Theorem, which states that if n is any real
number (not just a positive integer), then within the unit disc,
n(n−1) 2 n(n−1)(n−2) 3 n(n−1)(n−2)(n−3) 4
(1 + z)n = 1 + nz + 2! z + 3! z + 4! z + · · · . (2.14)
Historically, this result was one of Newton’s key weapons in developing calculus,
and later it played an equally central role in the work of Euler.
In Exs. 16, 17, 18, we show how manipulation of power series may be used to
demonstrate the Binomial Theorem, first for all negative integers, then for all ratio-
nal powers. Although we shall not discuss it further, the case of an irrational power
ρ may be treated by taking an infinite sequence of rational numbers that converges
to ρ. Later we shall use other methods to establish a still more general version of
(2.14) in which the power n is allowed to be a complex number!
Next we describe how to divide two power series P(z) and Q(z). In order to
P
find the series P(z)/Q(z) = cj zj , one multiplies both sides by Q(z) to obtain
P
P(z) = Q(z) cj zj , and then multiplies the two power series on the right. By the
uniqueness result, the coefficients of this series must equal the known coefficients
of P(z), and this enables one to calculate the cj ’s. An example will make this process
much clearer.
P
In order to find the coefficients cj in the series 1/ez = cj zj , we multiply both
sides by ez to obtain
z 2 z3 z4
1 = [1 + z + + + + · · · ][c0 + c1 z + c2 z2 + c3 z3 + c4 z4 + · · · ]
2! 3! 4!
c c1 c2 2 c0 c1 c2 c3 3
0
= c0 + (c0 + c1 ) z + + + z + + + + z ··· .
2! 1! 0! 3! 2! 1! 0!
By the uniqueness result, we may equate coefficients on both sides to obtain an
infinite set of linear equations:
1 = c0 ,
0 = c0 + c1 ,
0 = c0 /2! + c1 /1! + c2 /0!,
0 = c0 /3! + c1 /2! + c2 /1! + c3 /0!, etc.
Successively solving the first few of these equations [exercise] quickly leads to
the guess cn = (−1)n /n!, which is then easily verified [exercise] by considering
Power Series 83
1/ez = 1 − z + 1 2
2! z − 1 3
3! z + 1 4
4! z − 1 5
5! z + · · · = e−z ,
provided this limit exists. For example, if we first recall [we will discuss this later]
that the real function ex may be written as
x n
ex = lim 1 + ,
nÑ∞ n
then applying the root test to the series
X∞ j2
j−3
P(z) = zj ,
j
j=1
yields [exercise] R = e3 .
84 Complex Functions as Transformations
On occasion both the ratio and root tests will fail, but there exists a slightly
refined version of the latter which can be shown to work in all cases. It is called
the Cauchy-Hadamard Theorem, and it says that
1
R= p .
lim sup n |cn |
We will not discuss this further since it is not needed in this book.
The above examples of power series were plucked out of thin air, but often our
starting point is a known complex function f(z) which is then expressed as a power
series. The problem of determining R then has a conceptually much more satisfying
answer. Roughly2 ,
Figure [2.17a] illustrates this, the singularities of f(z) being represented as explo-
sions. To understand which functions can be expanded into power series we need
deep results from later in the book, but we are already in a position to verify that
a rational function [the ratio of two polynomials] can be, and that the radius of
convergence for its expansion is given by (2.15).
[2.17] [a] Here, explosions mark the locations of singularities of a function f(z) that we
wish to expand into a power series centred at k, i.e., a series in powers of Z ≡ (z − k).
Then the radius of convergence R = (the distance from k to the nearest singularity.)
[b] Here we see the three discs of convergence that arise from three different choices of
the centre k, assuming that a is the nearest singularity for all three centres.
2
Later [p. 107] we shall have to modify the statement in the case that f(z) is a “multifunction”,
having more than one value for a given value of z.
Power Series 85
To begin with, reconsider [2.13a] and [2.13b], both of which are examples of
√
(2.15). Recall that in [2.13b] we merely claimed that R = 1 + k2 for the series expan-
sion of h(z) = 1/(1+z2 ) centred at the real point k. We now verify this and explicitly
find the series.
To do so, first note that (2.4) easily generalizes to
X ∞
1 Zj
= , if and only if |Z| < |a − k|, (2.16)
a−z (a − k)j+1
j=0
where a and k are now arbitrary complex numbers, and Z = (z − k) is the complex
number connecting the centre of the expansion to z. The condition |z − k| < |a − k|
for convergence is that z lie in the interior of the circle centred at k and passing
through a. See [2.17b], which also shows the discs of convergence when we instead
choose to expand 1/(a − z) about k1 or k2 . Since the function 1/(a − z) has just one
singularity at z = a, we have verified (2.15) for this particular function.
Earlier we found the expansion of 1/(1 − x2 ) by factorizing the denominator
and using partial fractions. We are now in a position to use exactly the same
approach to find the expansion of h(z) = 1/(1 + z2 ) centred at an arbitrary complex
number k:
1 1 1 1 1
= = − .
1 + z2 (z − i)(z + i) 2i −i − z i − z
Applying (2.16) to both terms then yields
∞
X
1 1 1 1
= − Zj . (2.17)
1 + z2 2i (−i − k)j+1 (i − k)j+1
j=0
Again, we have here a result concerning real functions that would be very difficult
to obtain using only real numbers.
86 Complex Functions as Transformations
∞
X
F(θ) = 12 a0 + [an cos nθ + bn sin nθ] ,
n=1
Taylor series and Fourier series of real functions are merely two different ways of
viewing complex power series.
3
Later it was found that some restrictions must be placed on F, but they are astonishingly weak.
4
In many areas of mathematics it is hard to find even one really enlightening book, but Fourier
analysis has been blessed with at least two: Lanczos (1966), and Körner (1988).
Power Series 87
If z moves outward from the origin along a ray θ = const. then v(r eiθ ) becomes
a function of r alone, say Vθ (r). For example,
r
V π4 (r) = √ .
2(1 + r2 ) − 2r
If z instead travels round and round a circle r = const. then v becomes a function
of θ alone, say Ver (θ). For example,
e 1 (θ) = 2 sin θ
V .
2 5 − 4 cos θ
Note that this is a periodic function of θ, with period 2π. The reason is simple
and applies to any V er (θ) arising from a (single-valued) function f(z): each time z
makes a complete revolution and returns to its original position, f(z) travels along
a closed loop and returns to its original position.
Now, to see the unity of Taylor and Fourier series, recall that (within the unit
disc) f(z) = 1/(1 − z) can be expressed as a convergent complex power series:
In particular,
r
√ = V π4 (r) = √1 r + r2 + √1 r3 − √1 r5 − r6 − √1 r7 + √1 r9 + · · · .
2(1 + r2 ) − 2r 2 2 2 2 2
Once again, consider how difficult this would be to obtain using only real numbers.
From this we find, for example, that
d98 r
√ = 98!
dr98 2(1 + r2 ) − 2r r=0
e 1 (θ):
If we instead put r = (1/2), we immediately obtain the Fourier series for V
2
2 sin θ e 1 (θ) =
=V 1
2 sin θ + 1
22
sin 2θ + 1
23
sin 3θ + 1
24
sin 4θ + · · · .
5 − 4 cos θ 2
The absence of cosine waves in this series correctly reflects the fact that Ve 1 (θ) is an
2
odd function of θ.
This connection between complex power series and Fourier series is not merely
aesthetically satisfying, it can also be very practical. The conventional derivation
e 1 (θ) requires that we evaluate the tricky integrals in (2.19),
of the Fourier series of V
2
88 Complex Functions as Transformations
whereas we have obtained the result using only simple algebra! Indeed, we can
now use our Fourier series to do integration:
Z 2π
2 sin θ sin nθ π
dθ = n .
0 5 − 4 cos θ 2
Further examples may be found in Exs. 21, 37, 38.
We end with a premonition of things to come. The coefficients in a Taylor series
may be calculated by differentiation, while those in a Fourier series may be cal-
culated by integration. Since these two types of series are really the same in the
complex plane, this suggests that there exists some hidden connection between dif-
ferentiation and integration that only complex numbers can reveal. Later we shall
see how Cauchy confirmed this idea in spectacular fashion.
ez = 1 + z + 1 2
2! z + 1 3
3! z + 1 4
4! z + ··· ,
ex+iy = ex eiy .
This is a consequence of the fact that if a and b are arbitrary complex numbers,
then ea eb = ea+b . To verify this we simply multiply the two series:
ea eb = 1 + a + 2!1 a2 + 3!1 a3 + · · · 1 + b + 2!1 b2 + 3!1 b3 + · · ·
2 3
a + 2ab + b2 a + 3a2 b + 3ab2 + b3
= 1 + (a + b) + + + ···
2! 3!
= 1 + (a + b) + 1
2! (a + b)2 + 1
3! (a + b)3 + · · ·
= ea+b .
The Exponential Function 89
[2.18] The power series for ez can be visualized as a spiral journey, the angle between
successive legs of the journey being fixed and equal to • = arg(z) = arg(x+iy). In the text
we show that the power series has the property that ea eb = ea+b , so that ex+iy = ex eiy .
Thus the spiral ends at the point that has length ex and angle y, as illustrated.
Here we have left it to you to show that the general term in the penultimate line is
indeed (a + b)n /n!.
• If z travels upward at a steady speed s, then w rotates about the origin at angular
speed s. After z has travelled a distance of 2π, w returns to its starting position.
Thus the mapping is periodic, with period 2πi.
• If z travels westward at a steady speed, w travels towards the origin, with ever
decreasing speed. Conversely, if z travels eastward at a steady speed, w travels
away from the origin with ever increasing speed.
• Combining the previous two facts, the entire w-plane (with the exception of
w = 0) will be filled by the image of any horizontal strip in the z-plane of
height 2π.
• A line in general position is mapped to a spiral of the type discussed in the
previous chapter.
• Euler’s formula eiy = cos y + i sin y can be interpreted as saying that ez wraps
the imaginary axis round and round the unit circle like a piece of string.
90 Complex Functions as Transformations
[2.19] The geometry of the complex exponential follows from the fact that ex+iy has
length ex and angle y. Thus horizontal lines y = const. map to rays, and vertical lines
x = const. map to circles. Mysteriously, for now, a computer can be used to empirically
confirm that small squares are ultimately mapped to small squares, in the limit that they
shrink and vanish. This is a central mystery of complex analysis, and it will begin to unravel
in Chapter 4.
• The half-plane to the left of the imaginary axis is mapped to the interior of the
unit circle, and the half-plane to the right of the imaginary axis is mapped to the
exterior of the unit circle.
• The images of the small squares closely resemble squares, and (related to this)
any two intersecting lines map to curves that intersect at the same angle as the
lines themselves.
[2.20] [a] The fact that (ex ) ′ = ex implies that the shaded triangle has unit base. It follows
that ynew ≍ (1 + δ) yold as δ vanishes. [b] Now start with x = 0 and yold = 1. Taking
= (x/n)
δ and repeatedly moving
δ to the right, the height is ultimately multiplied by
x n
1+ n x
each time, so ex ≍ 1 + n .
meaning of the series. We now describe a different approach in which the geometry
lies much closer to the surface. The idea is to generalize the real result,
x n
ex = lim 1 + . (2.20)
nÑ∞ n
Here is one way of understanding (2.20). As we discussed in Chapter 1, f(x) = ex
may be defined by the property f ′ (x) = f(x). Figure [2.20a] interprets this in terms
of the graph of y = f(x). Drawing a tangent at an arbitrary point, the base of the
shaded triangle is always equal to 1. As you see from the figure, it follows that if
the height is yold at some point x, then moving x an infinitesimal distance δ to the
right yields a new height given by
ynew = (1 + δ) yold .
To find the height ex at x, we divide the interval [0, x] into a large number n
of very short intervals of length (x/n). Since the height at x = 0 is 1, the height
at (x/n) will be approximately [1 + (x/n)] · 1, and so the height at 2(x/n) will be
approximately [1 + (x/n)] · [1 + (x/n)] · 1, and so......., and so the height at x =
n(x/n) will be approximately [1 + (x/n)]n . [For clarity’s sake, [2.20b] illustrates
this geometric progression with the small (hence inaccurate) value n = 3.] It is now
plausible that the approximation [1+(x/n)]n becomes more and more accurate as n
tends to infinity, thereby yielding (2.20). Try using a computer to verify empirically
that the accuracy does indeed increase with n.
Generalizing (2.20) to complex values, we may define ez as
z n
ez = lim 1 + . (2.21)
nÑ∞ n
92 Complex Functions as Transformations
First we should check that this is the same generalization of ex that we obtained
using power series. Using the Binomial Theorem to write down the first few terms
of the nth degree polynomial [1 + (z/n)]n , we get
z n h z i n(n − 1) h z i2 n(n − 1)(n − 2) h z i3
1+ = 1+n + + + ···
n n 2! n 3! n
1 − n1 2 1 − n1 1 − n2 3
= 1+z+ z + z + ··· ,
2! 3!
which makes it clear that we do recover the original power series as n tends to
infinity.
Next we turn to the geometry of (2.21). In deciphering the power series for ez we
felt free to assume Euler’s formula, because in Chapter 1 we used the power series
to derive that result. However, it would smack of circular reasoning if we were to
assume Euler’s formula while following our new approach to ez , based on (2.21).
Temporarily, we shall therefore revert to our earlier notation and write r∠θ instead
of r eiθ ; the fact we wish to understand is therefore written ex+iy = ex ∠ y.
With n = 6, figure [2.21] uses Ex. 1.5, p. 52, to geometrically construct the suc-
cessive powers of a ≡ [1 + (z/n)] for a specific value of z. [All six shaded triangles
are similar; the two kinds of shading merely help to distinguish one triangle from
the next.] Even with this small value of n, we see empirically that in this particular
case [1 + (z/n)]n is close to ex ∠ y. To understand this mathematically, we will try
to approximate a = [1 + (z/n)] when n is large.
[2.21] Begin with the lightly shaded triangle with vertices 0, 1, and a ≡ [1 + (z/n)], here
illustrated with n = 6. Now successively construct triangles that are all similar to the orig-
inal, as illustrated. We have thereby constructed the successive powers of a, culminating
in an = [1 + (z/n)]n . Our earlier computations demonstrate that this must ultimately
be equal to ex ∠ y, and this concrete example seems to confirm this. But what is the
geometrical explanation?!
The Exponential Function 93
Now set ϵ = (z/n) = (x + iy)/n. With the same values of z and n as in [2.21],
figure [2.23] shows the approximation b ≡ 1 + n
x
∠ n
y
to a, together with its
successive powers.
5
Once again, “ ≍ ” denotes Newton’s concept of ultimate equality; see the new Preface.
94 Complex Functions as Transformations
Returning to the general case, the geometry of (2.21) should now be clear. As n
goes to infinity,
h z in h x y in x n
1+ ≍ 1+ ∠ = 1+ ∠ y.
n n n n
Thus, in this limit as n goes to infinity, and using (2.20), we deduce that
ex+iy = ex ∠ y,
as was to be shown. In particular, if we put x = 0 then we recover Euler’s formula,
eiy = 1∠ y, and so we are entitled to write ex+iy = ex eiy .
For a slightly different way of looking at (2.21), see Ex. 22.
It is not hard to show that all the familiar identities for cos x and sin x continue
to hold for our new complex functions. For example, we still have
cos2 z + sin2 z = (cos z + i sin z)(cos z − i sin z) = eiz e−iz = e0 = 1,
despite the fact that this identity no longer expresses Pythagoras’ Theorem. Simi-
larly, we will show that if a and b are arbitrary complex numbers then
cos(a + b) = cos a cos b − sin a sin b (2.23)
sin(a + b) = sin a cos b + cos a sin b, (2.24)
despite the fact that these identities no longer express the geometric rule for
multiplying points on the unit circle. First,
cos(a + b) + i sin(a + b) = ei(a+b)
= eia eib
= (cos a + i sin a)(cos b + i sin b)
= (cos a cos b − sin a sin b) + i(sin a cos b + cos a sin b),
exactly as in the previous chapter. However, in view of the warning above, we do
not obtain (2.23) and (2.24) simply by equating real and imaginary parts. Instead
[exercise] one first finds the analogous identity for cos(a + b) − i sin(a + b), then
adds it to (or subtracts it from) the one above.
[2.24] [a] The graph of cosh x may be visualized as the average (i.e., midpoint) of the
graphs of ex and e−x . [b] The graph of sinh x may be visualized in the same way, as the
average of ex and −e−x .
the surface rises above the origin in the form of a cone. Also, | sin(z + π)| = | sin z|,
so there is an identical cone at each multiple of π along the real axis. These are the
only points [exercise] at which the surface hits the plane. Figure [2.25]—which we
have adapted from Markushevich (2005)—shows a portion of the surface. Notice
that this surface also yields the cosh graph, for if we restrict z = (3π/2) + iy to the
line x = (3π/2), for example, then | sin z| = cosh y.
A practical benefit of this unification is that if you can remember (or quickly
derive using Euler’s formula) a trig identity involving cosine and sine, then you can
immediately write down the corresponding identity for the hyperbolic functions.
For example, if we substitute a = ir1 and b = ir2 into (2.23) and (2.24), then we
obtain (2.25) and (2.26).
Cosine and Sine 97
[2.25] Cutting open the modular surface of the complex function sin z reveals the
previously hidden unity of the circular and hyperbolic functions.
The connection between the circular and hyperbolic functions becomes stronger
still if we generalize the latter to complex functions in the obvious way:
ez + e−z ez − e−z
cosh z ≡ and sinh z ≡ .
2 2
Since we now have
cosh z = cos(iz) and sinh z = −i sin(iz),
the distinction between the two kinds of function has all but evaporated: cosh is the
composition of a rotation through (π/2), followed by cos; also, sinh is the compo-
sition of a rotation through (π/2), followed by sin, followed by a rotation through
−(π/2).
eastward at unit speed, whose position at time t is z = t − ic. See [2.26], in which
the line is shown heavy and unbroken. As z traces this line, −z traces the line y = c,
but in the opposite direction. Applying the mapping z ÞÑ iz (which is a rotation of
2 ), the image particles trace the vertical lines x = ±c, again with unit speed and in
π
opposite directions. Finally applying z ÞÑ 21 ez , the image particles orbit with equal
and opposite angular speeds in origin-centred circles of radii 12 e±c .
The image orbit under z ÞÑ w = cos z of the original particle travelling on the
line y = −c is just the sum of these counter-rotating circular motions. This is clearly
some kind of symmetrical oval hitting the real and imaginary axes at a = cosh c
and ib = i sinh c. It is also clear that cos z executes a complete orbit of this oval
with each movement of 2π by z; this is the geometric meaning of the periodicity of
cos z.
I haven’t found a simple geometric explanation, but it’s easy to show symboli-
cally that the oval traced by cos z is a perfect ellipse. Writing w = u + iv, we find
from the figure [exercise] that u = a cos t and v = b sin t, which is the familiar
parametric representation of the ellipse (u/a)2 + (v/b)2 = 1. Furthermore,
p q
a2 − b2 = cosh2 c − sinh2 c = 1,
so the foci are at ±1, independent of which particular horizontal line z travels along.
Cosine and Sine 99
Try mulling this over. How does the shape of the ellipse change as we vary c?
How do we recover the real cosine function as c tends to zero? What is the orbit
of cos z as z travels eastward along the line y = c, above the real axis? What is the
image of the vertical line x = c under z ÞÑ cosh z? What is the orbit of sin z as z
travels eastward along the line y = c; how does it differ from the orbit of cos z; and
is the resulting variation of | sin z| consistent with the modular surface shown in
[2.25]?
Before reading on, try using the idea in [2.26] to sketch for yourself the image
under z ÞÑ cos z of a vertical line.
As illustrated in [2.27], the answer is a hyperbola. We can show this using the
addition rule (2.23), which yields
On a horizontal line, y is constant, so (u/ cosh y)2 + (v/ sinh y)2 = 1, as before. On
a vertical line, x is constant, so (u/ cos x)2 − (v/ sin x)2 = 1, which is the equation
of a hyperbola. Furthermore, since cos2 x + sin2 x = 1, it follows that the foci of the
hyperbola are always ±1, independent of which vertical line is being mapped.
Figure [2.27] tries to make these results more vivid by showing the image of a
grid of horizontal and vertical lines. Note the empirical fact that each small square in
[2.27] Using the same reasoning as the previous figure, vertical lines are mapped by
cos z to hyperbolas, and calculation reveals that these are also confocal, with the same
foci as the ellipses, ±1! Next, recall the reflection property of the conic sections: a ray of
light emitted from a focus is reflected directly towards the other focus by the ellipse, and
directly away from the other focus by the hyperbola, as illustrated. It follows that these
two sets of conic sections are the orthogonal trajectories of each other. Furthermore—and
this is an additional, surprising, and presently mysterious fact—as the size of the squares
in the grid on the left tends to zero, their images under cos z are ultimately squares, too,
as illustrated!
100 Complex Functions as Transformations
the grid is mapped by cos z to an image shape that is again approximately square. This is
the same surprising (and visually pleasing) phenomenon that we observed in the
case of z ÞÑ ez .
We hope your curiosity is piqued—later chapters are devoted to probing this
phenomenon in depth. In the present case of z ÞÑ cos z we can at least give a
mathematical explanation of part of the result, namely, that the sides of the image
“squares” do indeed meet at right angles; in other words, each ellipse cuts each
hyperbola at right angles.
This hinges on the fact that these ellipses and hyperbolas are confocal. To prove
the desired result [exercise], think of each curve as a mirror, then appeal to the
familiar reflection property of the conic sections: a ray of light emitted from a focus
is reflected directly towards the other focus by the ellipse, and directly away from
the other focus by the hyperbola, as illustrated in [2.27].
2.6 Multifunctions
two (b and c) using the fact that as z = r eiθ orbits round an origin-centred circle,
z3 = r3 ei3θ orbits with three times the angular speed, executing a complete revo-
lution each time z executes one third of a revolution. Put differently, reversing the
direction of the mapping divides the angular speed by three. This is an essential
√
ingredient in understanding the mapping z ÞÑ 3 z, which we will now study in
detail.
√ √ √
Writing z = r eiθ , we have 3 z = 3 r ei(θ/3) . Here 3 r is uniquely defined as the
real cube root of the length of z; the sole source of the three-fold ambiguity in the
formula is the fact that there are infinitely many different choices for the angle θ of
a given point z.
Think of z as a moving point that is initially at z = p. If we arbitrarily choose θ
√
to be the angle ϕ shown in [2.28], then 3 p = a. As z gradually moves away from p,
√ √
θ gradually changes from its initial value ϕ, and 3 z = 3 r ei(θ/3) gradually moves
Multifunctions 101
away from its initial position a, but in a completely determined way—its distance from
the origin is the cube root of the distance of z, and its angular speed is one third
that of z.
Figure [2.29] illustrates this. Usually we draw mappings going from left to right,
but here we have reversed this convention to facilitate comparison with [2.28].
√
As z travels along the closed loop A (finally returning to p), 3 z travels along
the illustrated closed loop and returns to its original value a. However, if z instead
√
travels along the closed loop B, which goes round the origin once, then 3 z does not
return to its original value but instead ends up at a different cube root of p, namely
b. Note that the detailed shape of B is irrelevant, all that matters is that it encircles
√
the origin once. Similarly, if z travels along C, encircling the origin twice, then 3 z
ends up at c, the third and final cube root of p. Clearly, if z were to travel along a
√
loop [not shown] that encircled the origin three times, then 3 z would return to its
original value a.
√ √
The premise for this picture of z ÞÑ 3 z was the arbitrary choice of 3 p = a, rather
√
than b or c. If we instead chose 3 p = b, then the orbits on the left of [2.29] would
√
simply be rotated by (2π/3). Similarly, if we chose 3 p = c, then the orbits would
be rotated by (4π/3).
√
The point z = 0 is called a branch point of 3 z. More generally, let f(z) be a mul-
tifunction and let a = f(p) be one of its values at some point z = p. Arbitrarily
choosing the initial position of f(z) to be a, we may follow the movement of f(z) as
z travels along a closed loop beginning and ending at p. When z returns to p, f(z)
102 Complex Functions as Transformations
√ √
[2.29] Writing z = r eiθ , we have 3 z = 3 r ei(θ/3) . Once again, let us arbitrarily start at
√3 p = a, and then let p move along the three illustrated paths, A, B, and C. As you see,
which cube root we end up at does not depend on the detailed shape of the path, it only
depends on the net change in θ—how many complete revolutions have been executed
when p returns home. The cube root rotates one third of a revolution each time p loops
around the branch point at the origin.
will either return to a or it will not. A branch point z = q of f is a point such that
f(z) fails to return to a as z travels along any loop that encircles q once.
√
Returning to the specific example f(z) = 3 z, we have seen that if z executes three
revolutions round the branch point at z = 0 then f(z) returns to its original value.
If f(z) were an ordinary, single-valued function then it would return to its original
value after only one revolution. Thus, relative to an ordinary function, two extra
revolutions are needed to restore the original value of f(z). We summarize this by
√
saying that 0 is a branch point of 3 z of order two.
More generally, if q is a branch point of some multifunction f(z), and f(z) first
returns to its original value after N revolutions round q, then q is called an alge-
braic branch point of order (N − 1); an algebraic branch point of order 1 is called
a simple branch point. We should stress that it is perfectly possible that f(z) never
returns to its original value, no matter how many times z travels round q. In this
case q is called a logarithmic branch point—the name will be explained in the next
section.
√
By extending the above discussion of 3 z, check for yourself that if n is an integer
then z(1/n) is an n-valued multifunction whose only (finite) branch point is at z = 0,
the order of this branch point being (n − 1). More generally, the same is true for
any fractional power z(m/n) , where (m/n) is a fraction reduced to lowest terms.
Multifunctions 103
[2.30] Here is the terminology that is used to describe different types of regions in the
plane.
104 Complex Functions as Transformations
[2.31] If we chose a simply connected region S that does not contain the branch point at
0, then no path within S can encircle the branch point at 0, and therefore it is possible to
√ √
define a single-valued branch of 3 z within S. For example, if we arbitrarily take 3 p = a,
√
then let z = p move
√ about within S and arrive at z = Z, say, then 3 z arrives at a unique
3
value f1 (Z) = Z that is independent of the path. But since there are three values of
√ √
3 p, there are three distinct branches of 3 z. Incidentally, note that these three branches
which we will call f1 (Z). Since the path is irrelevant, f1 is an ordinary, single-valued
√
function of position on S; it is called a branch of the original multifunction 3 z.
Figure [2.31] illustrates such a set S, together with its image under the branch
√
f1 of 3 z. Here we have reverted to our normal practice of depicting the mapping
√
going from left to right. If we instead choose 3 p = b then we obtain a second
√ √
branch f2 of 3 z, while 3 p = c yields the third and final branch f3 . Notice, inci-
dentally, that all three branches display the by now ubiquitous (yet mysterious)
preservation of small squares.
We now describe how we may enlarge the domain S of the branches so as to
obtain the cube roots of any point in the plane. First of all, as illustrated in [2.32], we
draw an arbitrary (but not self-intersecting) curve C from the branch point 0 out to
infinity; this is called a branch cut. Provisionally, we now take S to be the plane with
the points of C removed—this prevents any closed path in S from encircling the
branch point. We thereby obtain on S the three branches f1 , f2 , and f3 . For example,
the figure shows the cube root f1 (d) of d.
What about a point such as e on C? Imagine that z is travelling round an origin-
centred circle through e. The figure illustrates the fact that f1 (z) approaches two
different values according as z arrives at e with positive or negative angular speed.
Multifunctions 105
√
[2.32] In order to define a single-valued branch of f(z) = 3 z throughout the entire com-
plex plane, we begin by drawing an arbitrary curve C that starts at the branch point
0 and extends out to infinity, as illustrated, and define S to be C with the points of √C
removed. The curve C is called a branch cut. Out of the three possible choices for 3 d,
let us arbitrarily choose the illustrated point f1 (d). Then f1 (Z) is single-valued throughout
S, because it is impossible for z to encircle the branch point at 0 as it travels from d to
Z. But we still need to define f1 (e) at a point such as e that lies on C. By convention, it
is defined to be the value of f1 (z) when z arrives at e travelling counterclockwise, i.e.,
arriving from the right of C. The two other branches are defined in exactly the same way.
A common choice for C is the negative real axis. If we do not allow z to cross the
cut then we may restrict the angle θ = arg(z) to lie in the range −π < θ ⩽ π. This is
called the principal value of the argument, written Arg (z); note the capital first letter.
√
With this choice of θ, the single-valued function 3 r ei(θ/3) is called the principal
√
branch of the cube root; let us write it as [ 3 z ]. Note that the principal branch agrees
106 Complex Functions as Transformations
with the real cube root function on the positive real axis, but not on the negative
√
real axis; for example, [ 3 −8 ] = 2 ei(π/3) . Also note that the other two branches
associated with this choice of C can be expressed in terms of the principal branch
√ √
as ei(2π/3) [ 3 z ] and ei(4π/3) [ 3 z ].
It should be clear how the above discussion extends to a general fractional
power.
1
[2.33] [a] The real function f(x) = (1+x) 3 can be expressed as a power series, but only for
−1 < x < 1; the dashed curve shows the 30th degree polynomial obtained by truncating
the binomial series at x30 . But why does convergence break down here?! There are no
1
singularities in sight, and, unlike the resolution of the convergence mystery of 1+x 2 , the
1
complex generalization f(z) = (1 + z) also does not have any singularities! [b] But f(z)
3
does have a branch point at −1. Suppose the series were to converge at the illustrated
point z outside the unit disc. Then the two illustrated paths could encircle the branch
point, yielding two different values of f(z). But the power series is single-valued, so this
is impossible!
z = 0 the series equals 1. But while f(x) was an ordinary single-valued function
of x, the LHS of the above equation is a three-valued multifunction of z, with a
2π
second order branch point at z = −1. For example, f(0) takes three values: 1, ei 3 ,
2π
and e−i 3 . We now recognize that the power series represents just one branch of
f(z), namely the one for which f(0) = 1.
This solves the mystery. For suppose that the series were to converge inside the
larger circle in [2.33b], and in particular at the illustrated point z. Starting at z = 0
with the value f(0) = 1, then travelling along the two illustrated paths to z, we
clearly end up with two different values of f(z), because together the two paths
enclose the branch point at −1. But the power series cannot mimic this behaviour
since it is necessarily single-valued—its only way out is to cease converging outside
the unit disc. We have demanded the impossible of the power series, and it has
responded by committing suicide!
This example shows that a branch point is just as real an obstacle to convergence
as a singularity. Quite generally, this argument shows that if a branch of a multi-
function can be expressed as a power series, the disc of convergence cannot be large
enough to contain any branch points of the multifunction. This strongly suggests
a further generalization of the (unproven) statement (2.15):
If a complex function or a branch of a multifunction can be expressed as
a power series, the radius of convergence is the distance to the nearest (2.27)
singularity or branch point.
108 Complex Functions as Transformations
Much later in the book we will develop the tools necessary to confirm this
conjecture.
The divergence of the series beyond this interval is vividly conveyed by the dashed
curve, which is the graph of the 20th degree polynomial obtained by truncating the
binomial series at x20 .
As before, the explanation lies in C, where f(x) becomes the two-valued mul-
√ p
tifunction f(z) = z2 + 1. This can be rewritten as f(z) = (z − i)(z + i), which
makes it clear that f(z) has two simple branch points, one at i and the other at −i.
These branch points obstruct the convergence of the corresponding complex series,
limiting it to the unit disc shown in [2.34b].
In greater detail, the notation of [2.34b] enables us to write
√
f(z) = r1 r2 ei(θ1 +θ2 )/2 . (2.28)
Here we must bear in mind that the figure illustrates only one possibility (out of
infinitely many) for each of the angles θ1 and θ2 . To see that i is indeed a branch
1
[2.34] [a] The real function f(x) = (1 + x2 ) 2 can be expressed as a power series, but only
for −1 < x < 1; the dashed curve shows the 20th degree polynomialp obtained by trun-
cating the binomial series at x20 . [b] In the complex plane, f(z) = (z − i)(z + i) =
√
r1 r2 ei(θ1 +θ2 )/2 , which makes it clear that f(z) has branch points at ±i, and these
obstruct the convergence of the complex power series beyond the unit disc. If z loops
around i along L, as illustrated, then the net changes in θ1 and θ2 are 2π and 0,
respectively. So f(z) ⇝ eiπ f(z) = −f(z).
Multifunctions 109
point, suppose we start with the value of f(z) given by the illustrated values of θ1
and θ2 . Now let z travel round the illustrated loop L. As it does so, (z + i) rocks
back and forth, so θ2 merely oscillates, finally returning to its original value. But
(z − i) undergoes a complete revolution, and so θ1 increases by 2π. Thus when z
returns to its original position, (2.28) shows that f(z) does not return to its original
value, but rather to
√ √
f new (z) = r1 r2 ei(θ1 +2π+θ2 )/2 = eiπ r1 r2 ei(θ1 +θ2 )/2 = −fold (z).
Of course the same thing happens if z travels along a loop that goes once round −i,
instead of round +i.
In order to dissect f(z) into two single-valued branches, we appear to need two
branch cuts: one cut C1 from i to infinity (to prevent us encircling the branch point
at i), and another cut C2 from −i to infinity, for the same reason. Figure [2.35a]
illustrates a particularly common and important choice of these cuts, namely, rays
going due west. If we do not allow z to cross the cuts then we may restrict the angle
θ1 = arg(z − i) to its principal value, in the range −π < θ1 ⩽ π. For example, the
angle in [2.34b] is not the principal value, while the one in [2.35a] is. If θ2 is like-
wise restricted to its unique principal value then (2.28) becomes the single-valued
principal branch of f(z), say F(z). The other branch of f(z) is simply −F(z).
Let us return to the previous situation in which we allowed θ1 and θ2 to take
general values rather than their principal values. Figure [2.35b] illustrates the fact
that it is possible to define two branches of f(z) using only a single branch cut
C that connects the two branch points. If z is restricted to the shaded, multiply-
connected region S, then it cannot loop around either branch point singly. It can,
√ √
[2.35] [a] We can define two single-valued branches of f(z) = z2 + 1 = r1 r2 ei(θ1 + θ2 )/2
by making branch cuts going due west from both branch points, and then taking θ1 and
θ2 to be their principal values, as illustrated. [b] In fact we only need a single branch cut
connecting the two branch points, such as C. For then a loop L cannot encircle either
branch point singly, and if it encircles both, as illustrated, then f(z) does not change its
value.
110 Complex Functions as Transformations
however, travel along a loop such as L that encircles both branch points together.
But in this case both θ1 and θ2 increase by 2π, so (2.28) shows that f(z) returns to
its original value. Thus we can define two single-valued branches on S. Finally, we
may expand S until it borders on C.
Since arg(z) takes infinitely many values, differing from each other by multiples
of 2π, we see that log(z) is a multifunction taking infinitely many values, differing
from each other by multiples of 2πi. For example,
√
log(2 + 2i) = ln 2 2 + i(π/4) + 2nπi,
[2.36] Here we see that 0 is a branch point of log(z). However, this branch point is
quite unlike that of z(1/n) , for no matter how many times we loop around the origin (say
counterclockwise), log(z) never returns to its original value, rather it continues moving
upwards forever. If we make a branch cut along the negative real axis, we can restrict
the angle to be its principal value, in which case we obtain the principal value of log(z),
written Log (z) ≡ ln |z| + i Arg (z).
defined by −π < Arg (z) ⩽ π. This yields the principal branch or principal value of
the logarithm, written Log (z), and defined by
Log (z) ≡ ln |z| + i Arg (z).
√
For example, Log (− 3 − i) = ln 2 − i (5π/6), Log (i) = i(π/2), and Log (−1) = iπ.
Note that if z = x is on the positive real axis, Log (x) = ln(x).
Figure [2.37] illustrates how the mapping z ÞÑ w = Log (z) sends rays to hor-
izontal lines, and circles to vertical line-segments connecting the horizontal lines
at heights ±π; the entire z-plane is mapped to the horizontal strip of the w-plane
bounded by these lines. Study this figure until you are completely at peace with
it. You can see the price we pay for forcing the logarithm to be single-valued: it
becomes discontinuous at the cut. As z crosses the cut travelling counterclock-
wise, the height of w suddenly jumps from π to −π. If we wish w to instead move
continuously, then we must switch to the branch Log (z) + 2πi of the logarithm.
Another problem with restricting ourselves to the principal branch is that the
familiar rules for the logarithm break down. For example, Log (ab) is not always
equal to Log (a) + Log (b); try a = −1 and b = i, for example. However, if we keep
all values of the logarithm in play then it is true [exercise] that
log(ab) = log(a) + log(b) and log(a/b) = log(a) − log(b),
112 Complex Functions as Transformations
[2.37] The geometry of Log (z). Note that the price we pay for having a single-valued
logarithm is that it is discontinuous: as we cross the branch cut, its value jumps.
in the sense that every value of the LHS is contained amongst the values of the RHS,
and vice versa.
6
Once again, “ ≍ ” denotes Newton’s concept of ultimate equality; see the new Preface.
The Logarithm Function 113
1
There are n branches of (1 + z) n within the unit disc, but since L(0) = 0 we need
1
the branch of (1 + z) n that equals 1 when z = 0. Appealing to the Binomial series
for this principal branch, we obtain
1 1 1 1
L 1 n(n − 1) 2 n(n − 1)( n1 − 2) 3
1+ ≍ 1+ z+ z + z + ··· ,
n n 2! 3!
and hence
( n1 − 1)z2 ( n1 − 1)( 2n
1
− 1)z3 ( n1 − 1)( 2n
1 1
− 1)( 3n − 1)z4
L(z) ≍ z + + + + ··· .
2 3 4
Finally, since this ultimate equality becomes equality in the limit that (1/n) tends
to zero, we obtain the following logarithmic power series:
z2 z3 z4 z5 z6
Log (1 + z) = z − + − + − + ··· . (2.29)
2 3 4 5 6
For other approaches to this series, see Exs. 31, 32.
Using the ratio test, you can check for yourself that this series does indeed con-
verge inside the unit circle. In fact it can be shown [see Ex. 11] that the series also
converges everywhere on the unit circle, except obviously at z = −1. This yields
some very interesting special cases. For example, putting z = i and then equating
real and imaginary parts, we get
√ 1 1 1 1 1 1
ln 2 = − + − + − + ···
2 4 6 8 10 12
1 1 1 1 1
and π = 4 1− + − + − + ··· .
3 5 7 9 11
√
Try checking the first series by noting that if z = 1, then ln 2 = 12 ln(1 + z). The
remarkable second series for π was first published by Leibniz in 1676, though it was
already known to Newton and Gregory. It is usually obtained from the power series
for tan−1 x, evaluated at x = 1. Astonishingly, that power series was discovered
centuries before Newton, Gregory, and Leibniz roamed the Earth, by the Indian
mathematician, Mˉadhava of Sangamagrˉama (c.1340 – c.1425); see (Stillwell, 2010,
p. 166).
For other interesting applications of the logarithmic series, see Exs. 36, 37, 38.
zk = ek log(z) . (2.30)
114 Complex Functions as Transformations
Let z = r eiθ , where θ is chosen to be the principal value, Arg (z). Then
But the most general branch of log(z) is simply Log (z) + 2nπi, where n is an
integer, so
is true irrespective of which branch of the logarithm is chosen. Clearly, by the same
argument, (2.30) is true for all integer values of k.
1 1
Next, consider the three branches of z 3 . Recalling that the principal branch [z 3 ]
√
of this function is 3 r ei(θ/3) , where θ again represents the principal angle, you can
1 1
easily check that e 3 Log (z) = [z 3 ]. Thus the general branch of the logarithm yields
1 2nπ 1
e 3 log(z) = ei 3 [z 3 ].
Thus we have again confirmed (2.30), in the sense that the infinitely many branches
1 1
of log(z) yield precisely the three branches of the cube root: [z 3 ], ei(2π/3) [z 3 ], and
1
ei(4π/3) [z 3 ]. By the same reasoning, if (p/q) is a fraction reduced to lowest terms
p p
then e q log(z) yields precisely the q branches of z q .
Finally, note that the RHS of (2.30) is still meaningful if k = (a + ib) is a complex
number. Emboldened by the above successes, we now take (2.30) as the definition of
a complex power. If we use Log (z) in (2.30) then we find that the principal branch
of z(a+ib) is given by [exercise]
If z now travels along a closed loop encircling the origin n times, then log(z) moves
along a path from Log (z) to Log (z) + 2nπi, and z(a+ib) moves along a path from
[z(a+ib) ] to
If b ̸= 0 then the factor e−2πnb makes it obvious that z(a+ib) never returns to
its original value, no matter how many times we go round the origin. Thus z = 0
is a logarithmic branch point in this case. This is still true even if b = 0, provided
[exercise] that the real power a is irrational. Only when a is a rational number does
za return to its original value after a finite number of revolutions, and only when
a is an integer does za become single-valued.
We end with an important observation on the use of “ez ” to denote the single-
valued exponential mapping. Reversing the roles of the constant and variable in
(2.30), we are forced to define f(z) = kz to be the “multifunction” [see Ex. 29] f(z) =
ez log(k) . But if we now put k = e = 2.718 . . . then we are suddenly in hot water: the
exponential mapping “ez ” is merely one branch [what are the others?] of the newly
defined multifunction (2.718 . . .)z . To avoid this confusion, some authors always
Averaging over Circles* 115
write the exponential mapping as exp(z). However, we shall retain the notation
“ez ”, which is both convenient and rooted in history, with the understanding that
ez always refers to the single-valued exponential mapping, and never to the multifunction
(2.718 . . .)z .
If we imagine the plane to be massless, Z is the point at which we could rest the
plane on a pin so as to make it balance.
Throughout this section we shall take the masses of the particles to be equal, in which
case the centroid becomes the average position of the particles:
1 X
n
Z= zj .
n
j=1
∑n
j=1 mj zj
[2.38] [a] The centroid Z of masses mj located at zj is Z ≡ ∑n . In the case that
j=1 mj
the masses are all equal, join Z to the locations of the masses, and sum these connecting
complex numbers (zj − Z) . . . [b] . . . and discover that their sum vanishes! Furthermore,
this vanishing sum characterizes the location of Z.
Our next result will play a minor role at the end of this section, but later we
shall see that it has other interesting consequences. The convex hull H of the set of
particles {zj } is defined to be the smallest convex polygon such that each particle
lies on H or inside it. More intuitively, first imagine pegs sticking out of the plane
at each point zj , then stretch an imaginary rubber band so as to enclose all the pegs.
When released, the rubber band will contract into the desired polygon H shown in
[2.39a]. We can now state the result:
The centroid Z must lie in the interior of the convex hull H. (2.32)
For if p is outside this set, we see that the complex numbers from p to the particles
cannot possibly cancel, as they must do for Z. More formally, we take it as visually
evident that through any exterior point p we may draw a line L such that H and its
shaded interior lie entirely on one side of L. The impossibility of the complex num-
bers cancelling now follows from their lying entirely on this side of L, for they all
must have positive components in the direction of the illustrated complex number
N normal to L. Except when the particles are collinear (in which case H collapses
to a line-segment), the same reasoning forbids Z from lying on H.
As illustrated in [2.39b], an immediate consequence of (2.32) is that
If all the particles lie within some circle then their centroid must also lie
(2.33)
within that circle.
The main result we wish to derive in this section is based on the following fact.
Defining the centre of a regular n-gon to be the centre of the circumscribing circle,
[2.39] The centroid must lie within the convex hull. [a] If p lies outside the convex hull
H of the masses, then the connecting complex numbers from p to the masses cannot
vanish, therefore the centroid Z must lie within H. [b] Therefore, if all the particles lie
within some circle then their centroid must also lie within that circle.
By virtue of (2.31), we may as well choose the n-gon to be centred at the origin,
in which case the claim is that the sum of the vertices vanishes. As illustrated in
[2.40a], this is obvious if n is even since the vertices then occur in opposite pairs.
The explanation is not quite so obvious when n is odd; see [2.40b], which illus-
P
trates the case n = 5. However, if we draw zj systematically, taking the vertices
zj in counterclockwise order, then we obtain [2.40c], and the answer is suddenly
clear: the sum of the vertices of the regular 5-gon forms another regular 5-gon. The figure
explains why this happens. Since the angle between successive vertices in [2.40b] is
[2.40] The sum of the vertices of a regular n-gon vanishes. [a] If n is even, it is obvious
that the sum of the vertices of any regular, origin-centred n-gon must vanish, for the
vertices occur in diametrically opposite pairs. [b] In fact the sum of the vertices must also
vanish when n is odd, but this is no longer immediately obvious. Begin by noting that
the angle between successive vertices is (2π/n). [c] Drawing the sum of the vertices in
counterclockwise order, all is suddenly clear: the sum itself forms a regular n-gon, and
therefore it vanishes.
118 Complex Functions as Transformations
(2π/5), this is also the angle between successive terms of the sum in [2.40c]. Clearly
this argument generalizes to arbitrary n (both odd and even), thereby establishing
(2.34). For a different approach, see Ex. 40.
Note that if f(z) = c is constant, then its average over any set of points is equal
to c.
Henceforth, we shall restrict ourselves to the case where {zj } are the vertices of
a regular n-gon; correspondingly, ⟨f(z)⟩n will be understood as the average of f(z)
over the vertices of such a regular n-gon. Note that if we write f(z) = u(z) + iv(z),
then
⟨f(z)⟩n = ⟨u(z)⟩n + i ⟨v(z)⟩n . (2.35)
Initially, we consider only origin-centred polygons.
Consider, then, the average of f(z) = zm over the vertices of such a regular n-
gon. Figure [2.41] illustrates the case n = 6. In the centre of the figure is a shaded
regular hexagon, and on the periphery are the images of its vertices under the map-
pings z, z2 , . . ., z6 . Study this figure carefully, and see if you can understand what’s
going on. If we take still higher values of m, then this pattern repeats cyclically: z7
is like z1 , z8 is like z2 , and so on.
For us the essential feature of this figure is that unless m is a multiple of 6, the
image under zm of the regular 6-gon is another regular polygon. [Note that we
count two equal and opposite points as a regular 2-gon, but we do not count a
single point as a regular polygon.] More precisely, and in general,
Unless m is a multiple of n, the image under zm of an origin-centred reg-
ular n-gon is an origin-centred regular N-gon, where N = (n divided by
(2.36)
the highest common factor of m and n). If m is a multiple of n, then the
image is a single point.
Check that this agrees with [2.41]. Try to establish the result on your own, but see
Ex. 41 if you get stuck.
Combining this result with (2.34), we obtain the following key fact: If n >
m then ⟨zm ⟩n = 0. This is easy to generalize. If
Pm (z) = c0 + c1 z + c2 z2 + c3 z3 + · · · + cm zm
Averaging over Circles* 119
is a general polynomial of degree m, then its average over the vertices of the
n-gon is
In other words, the centroid of the image points is the image of the centroid. Expressing
this result in the language of averages,
Finally, let us generalize to regular n-gons that are centred at an arbitrary point k,
instead of the origin. Of course when we apply zm to the vertices of such a regular
polygon, the image points do not form a regular polygon. See [2.42], which shows
the effect of z4 on the vertices of a regular hexagon H centred at k, together with
the image of the entire circle on which these vertices lie. Nevertheless, the figure
also illustrates the surprising and beautiful fact that, once again, the centroid of the
image points is the image of the centroid of H. Figure [2.43a] confirms this empirically
120 Complex Functions as Transformations
[2.42] It is a surprising and beautiful fact (here illustrated with n = 6 and m = 4) that the
centroid of the image points of the vertices of a regular n-gon H under z ÞÑ zm is the
image of the centroid of H.
[2.43] [a] Empirical verification of the claim made in the previous figure: the sum of
the connecting complex numbers from k4 to images of the vertices of H vanishes, so k4
is indeed the centroid of these image points. [b] In order to mathematically explain this
e see main text
phenomenon, we view H to be the translation by k of the origin-centred H;
for the argument.
by showing that the sum of complex numbers connecting k4 to the image points is
indeed zero.
Extending our notation slightly, we may write the average of zm over the vertices
of H as ⟨zm ⟩H , so what we must show is that ⟨z4 ⟩H = k4 . It is no harder to treat the
general case of zm acting on the vertices of a regular n-gon H centred at k. First
note that H can be obtained by translating an origin-centred n-gon H e by k. See the
e translates to a vertex zj + k of H, it follows
example in [2.43b]. Since a vertex zj of H
that
P
But (z + k)m = m m j m−j
j=0 j z k is just an mth degree polynomial which maps
0 to k . Using (2.37), we conclude that if n > m then ⟨zm ⟩H = km , as was to be
m
shown.
Generalizing the argument that led to (2.37), we see that (2.37) is a special case
of the following result:
If n > m then the average of an mth degree polynomial Pm (z) over the
vertices of a regular n-gon centred at k is its value Pm (k) at the centre (2.38)
of the n-gon.
The fact that (2.40) holds for polynomials of arbitrarily high degree immediately
suggests that it might also hold for power series. We shall show that it does.
As usual we will only give the details for origin-centred power series, the gen-
P
eralization to arbitrary centres being straightforward. Let P(z) = ∞ j
j=0 cj z be the
Pm
power series, so that Pm (z) = j=0 cj zj are its approximating polynomials. If the
circle C lies inside the disc of convergence of P(z), then (2.12) implies the following.
No matter how small we choose a real number ϵ, we can find a sufficiently large m
such that Pm (z) approximates P(z) with accuracy ϵ throughout C and its interior. If
we write E(z) for the complex number from the approximation Pm (z) to the exact
answer P(z), then
Let m be chosen so that Pm (z) approximates the power series P(z) with
accuracy ϵ on and within a circle C centred at k. If a regular (m+1)-gon
(2.41)
is inscribed in C, then the average ⟨P(z)⟩m+1 of P(z) over its vertices
will approximate P(k) with accuracy 2ϵ.
We have thus transformed an exact result concerning the approximation Pm (z) into
an approximation result concerning the exact mapping P(z).
For example, let C be the unit circle, and let P(z) = ez . If we desire an accuracy
of ϵ = 0.004 everywhere on the unit disc then it turns out that m = 5 is sufficient,
i.e., the approximating polynomial of lowest degree that has this accuracy is
1 2 1 3 1 4 1 5
P5 (z) = 1 + z + 2! z + 3! z + 4! z + 5! z .
Averaging over Circles* 123
[2.44] Applying z ÞÑ ez to the unit circle and to the vertices of the inscribed regular
hexagon on the left yields the figure on the right. The centroid of the six image points is
almost the image of the centre of the original hexagon, namely, e0 = 1. As the number
of vertices goes to infinity, the result is ultimately exact.
[2.45] Empirical confirmation of the claim made in the previous figure: the connecting
complex numbers from 1 to the images of the vertices of the hexagon do indeed sum to
zero, at least to within the accuracy of this drawing.
124 Complex Functions as Transformations
In the limit that ϵ tends to zero and m tends to infinity, (2.41) yields a form of
Gauss’s Mean Value Theorem:
If a complex function f(z) can be expressed as a power series, and a circle
C (radius R and centre k) lies within the disc of convergence of that
power series, then
R2π
⟨f(z)⟩C = 2π1 iθ
0 f(k + R e ) dθ = f(k).
2.9 Exercises
1 Sketch the circle |z−1| = 1. Find (geometrically) the polar equation of the image
of this circle under the mapping z ÞÑ z2 . Sketch this image curve, which is called
a cardioid.
2 Consider the complex mapping z ÞÑ w = (z − a)/(z − b). Show geometrically
that if we apply this mapping to the perpendicular bisector of the line-segment
joining a and b, then the image is the unit circle. In greater detail, describe the
motion of w round this circle as z travels along the line at constant speed.
3 Consider the family of complex mappings
z−a
z ÞÑ Ma (z) = (a constant).
az − 1
[These mappings will turn out to be fundamental to non-Euclidean geometry.]
Do the following problems algebraically; in the next chapter we will provide
geometric explanations.
(i) Show that Ma [Ma (z)] = z. In other words, Ma is self-inverse.
(ii) Show that Ma (z) maps the unit circle to itself.
(iii) Show that if a lies inside the unit disc then Ma (z) maps the unit disc to
itself.
Hint: Use |q|2 = q q to verify that
4 In figure [2.7] we saw that if q2 ⩽ p3 then the solutions of x3 = 3px + 2q are all
real. Draw the corresponding picture in the case q2 > p3 , and deduce that one
solution is real, while the other two form a complex conjugate pair.
5 Show that the mapping z ÞÑ z2 doubles the angle between two rays coming
out of the origin. Use this to deduce that the lemniscate (see [2.9] on p. 69)
must self-intersect at right angles.
6 This question refers to the Cassinian curves in [2.9] on p. 69.
(i) On a copy of this figure, sketch the curves that intersect each Cassinian
curve at right angles; these are called the orthogonal trajectories of the
original family of curves.
(ii) Give an argument to show that each orthogonal trajectory hits one of the
foci at ±1.
(iii) If the Cassinian curves are thought of as a geographical contour map of
the modular surface (cf. [2.10]) of (z2 − 1), then what is the interpretation
of the orthogonal trajectories in terms of the surface?
126 Complex Functions as Transformations
(iv) In Chapter 4 we will show that if two curves intersect at some point p ̸=
0, and if the angle between them at p is ϕ, then the image curves under
z ÞÑ w = z2 will also intersect at angle ϕ, at the point w = p2 . Use this
to deduce that as z travels out from one of the foci along an orthogonal
trajectory, w = z2 travels along a ray out of w = 1.
(v) Check the result of the previous part by using a computer to draw the
√
images under w ÞÑ w of (A) circles centred at w = 1; (B) the radii of
such circles.
(vi) Writing z = x + iy and w = u + iv, find u and v as functions of x and
y. By writing down the equation of a line in the w-plane through w = 1,
show that the orthogonal trajectories of the Cassinian curves are actually
segments of hyperbolas.
7 Sketch the modular surface of C(z) = (z + 1)(z − 1)(z + 1 + i). Hence sketch the
Cassinian curves |C(z)| = const., then check your answer using a computer. To
answer the following questions, recall that if R(z) is a real function of position
in the plane, then R(p) is a local minimum of R if R(p) < R(z) for all z in the
immediate neighbourhood of p. A local maximum is defined similarly.
(i) Referring to the previous exercise, what is the significance of the orthog-
onal trajectories of the Cassinian curves you have just drawn?
(ii) Does |C(z)| have any local maxima?
(iii) Does |C(z)| have any non-zero local minima?
(iv) If D is a disc (or indeed a more arbitrary shape), can the maximum of |C(z)|
on D occur at a point inside D, or must the maximum occur at a boundary
point of D? What about the minimum of |C(z)| on D?
(v) Do you get the same answers to these questions if C(z) is replaced by
an arbitrary polynomial? What about a complex function that is merely
known to be expressible as a power series?
8 On page 69 we saw that the polar equation of the lemniscate with foci at ±1 is
r2 = 2 cos 2θ. In fact James Bernoulli and his successors worked with a slightly
different lemniscate having equation r2 = cos 2θ. Let us call this the standard
lemniscate.
(i) Where are the foci of the standard lemniscate?
(ii) What is the value of the product of the distances from the foci to a point
on the standard lemniscate?
(iii) Show that the Cartesian equation of the standard lemniscate is
(x2 + y2 )2 = x2 − y2 .
Exercises 127
[By virtue of (2.29), note that the second series is −Log (1 − z).]
P∞ j
12 Consider the geometric series P(z) = j=0 z , which converges to 1/(1 − z)
inside the unit disc. The approximating polynomials in this case are Pm (z) =
Pm j
j=0 z .
(i) Show that the error Em (z) ≡ |P(z) − Pm (z)| is given by
|z|m+1
Em (z) = .
|1 − z|
128 Complex Functions as Transformations
(ii) If z is any fixed point in the disc of convergence, what happens to the error
as m tends to infinity?
(iii) If we fix m, what happens to the error as z approaches the boundary point
z = 1?
(iv) Suppose we want to approximate this series in the disc |z| ⩽ 0.9, and fur-
ther suppose that the maximum error we will tolerate is ϵ = 0.01. Find the
lowest degree polynomial Pm (z) that approximates P(z) with the desired
accuracy throughout the disc.
13 We have seen that if we set Pn (z) = zn , then the representation of a com-
P
plex function f(z) as an infinite series ∞ n=0 cn Pn (z) (i.e., a power series) is
unique. This is not true, however, if Pn (z) is just any old set of polynomials.
The following example is taken from Boas and Boas (2010). Defining
zn−1 zn
P0 (z) = −1, and Pn (z) = − (n = 1, 2, 3, . . .),
(n − 1)! n!
show that
where the (complex) errors E1,2 (z) both have lengths less than ϵ. Use this to
show that by taking a sufficiently high value of n we can approximate [P(z) +
Q(z)] and P(z) Q(z) with arbitrarily high precision using [Pn (z) + Qn (z)] and
Pn (z) Qn (z), respectively.
15 Give an example of a pair of origin-centred power series, say P(z) and Q(z),
such that the disc of convergence for the product P(z)Q(z) is larger than either of
the two discs of convergence for P(z) and Q(z). [Hint: think in terms of rational
functions, such as [z2 /(5 − z)3 ], which are known to be expressible as power
series.]
16 Our aim is to give a combinatorial explanation of the Binomial Theorem (2.14)
for all negative integer values of n. The simple yet crucial first step is to write
n = −m and to change z to −z. Check that the desired result (2.14) now takes
Exercises 129
P∞
the form (1 − z)−m = r=0 cr zr , where cr is the binomial coefficient
m+r−1
cr = . (2.42)
r
[Note that this says that the coefficients cr are obtained by reading Pascal’s tri-
angle diagonally, instead of horizontally.] To begin to understand this, consider
the special case m = 3. Using the geometric series for (1−z)−1 , we may express
(1 − z)−3 as
[1 + z + z2 + z3 + · · · ] • [1 + z + z2 + z3 + · · · ] • [1 + z + z2 + z3 + · · · ],
p
powers, we must show that if p and q are integers then B(z, q ) is the principal
p
branch of (1 + z) q .
(i) With a fixed value of n, use the ratio test to show that B(z, n) converges
in the unit disc, |z| < 1.
(ii) By multiplying the two power series, deduce that
∞
X r
X
r n m
B(z, n) B(z, m) = Cr (n, m) z where Cr (n, m) = .
j r−j
r=0 j=0
21 Do the following problems by first substituting z = r eiθ into the power series
for ez , then equating real and imaginary parts.
P
(i) Show that the Fourier series for [cos(sin θ)] ecos θ is ∞n=0
cos nθ , and
n!
write down the Fourier series for [sin(sin θ)] ecos θ .
R2π
(ii) Deduce that 0 ecos θ [cos(sin θ)] cos mθ dθ = (π/m!), where m is a
positive integer.
√
(iii) By writing x = (r/ 2), find the power series for f(x) = ex sin x.
(iv) Check the first few terms of the series for f(x) by multiplying the series
for ex and sin x.
Exercises 131
(v) Calculate the nth derivative f(n) (0) using (1.14) on p. 25 of Chapter 1. By
using these derivatives in Taylor’s Theorem, verify your answer to part
(iii).
22 Reconsider the formula,
z n
ez = lim Pn (z), where Pn (z) = 1 + .
nÑ∞ n
26 For each function f(z) below, find and then plot all the branch points and sin-
gularities. Assuming that these functions may be expressed as power series
centred at k [in fact they can be], use the result (2.27) on p. 107 to verify the
stated value of the radius of convergence R.
(i) If f(z) = 1/(eπz − 1) and k = (1 + 2i), then R = 1.
√
(ii) If f(z) is a branch of 5 z4 − 1 and k = 3i, then R = 2.
√ √
(iii) If f(z) is a branch of z − i/(z − 1) and k = −1, then R = 2.
27 Until Euler cleared up the whole mess, the complex logarithm was a source
of tremendous confusion. For example, show that log(z) and log(−z) have no
common values, then consider the following argument of John Bernoulli:
32 Here is another approach to the logarithmic power series. As before, let L(z) =
Log (1 + z). Since L(0) = 0, the power series for L(z) must be of the form L(z) =
az + bz2 + cz3 + dz4 + · · · . Substitute this into the equation
1 + z = eL = 1 + L + 1 2
2! L + 1 3
3! L + 1 4
4! L + ··· ,
(ii) Consider the periodic “saw tooth” function F(θ) whose graph is shown
below. By substituting z = eiθ in the logarithmic series (2.29), use the
previous part to deduce the following Fourier series:
sin 2θ sin 3θ sin 4θ
F(θ) = sin θ − + − + ··· .
2 3 4
(iii) Check this Fourier series by directly evaluating the integrals (2.19).
(iv) Use a computer to draw graphs of the partial sums of the Fourier series.
As you increase the number of terms, observe the magical convergence of
this sum of smooth waves to the jagged graph below. If only Fourier could
have seen this on the screen, not just in his mind’s eye!
39 Show that (2.32) is still true even if the (positive) masses of the particles are not
all equal.
Exercises 135
40 Here is another simple way of deriving (2.34). If the vertices of the origin-
centred regular n-gon are rotated by ϕ, then their centroid Z rotates with them
to eiϕ Z. By choosing ϕ = (2π/n), deduce that Z = 0.
41 To establish (2.36), let z0 , z1 , z2 , . . . , zn−1 be the vertices (labelled counterclock-
wise) of the regular n-gon, and let C be the circumscribing circle. Also, let
wj = zm j be the image of vertex zj under the mapping z ÞÑ w = z . Think
m
3.1 Introduction
1
Also known as a “linear”, “bilinear”, “linear-fractional”, or “homographic” transformation.
2
According to Coxeter (1967), this connection was first recognized by H. Liebmann in 1905, the very
year that Einstein discovered Special Relativity!
Visual Complex Analysis. 25th Anniversary Edition. Tristan Needham, Oxford University Press.
© Tristan Needham (2023). DOI: 10.1093/oso/9780192868916.003.0003
138 Möbius Transformations and Inversion
Te2 − (X
e2 + Ye2 + Z
e2 ) = T 2 − (X2 + Y 2 + Z2 ).
The connection exhibited in (3.2) is deep and powerful. Just for starters, it means
that any result we establish concerning Möbius transformations will immediately
yield a corresponding result in Einstein’s Theory of Relativity. Furthermore, these
Möbius transformation proofs turn out to be considerably more elegant than direct
spacetime proofs.
To really understand the above claims, we strongly recommend that after read-
ing this chapter you consult Penrose and Rindler (1984, Ch. 1), as well as Shaw
(2006, Ch. 23) and Needham (2021, §6.4).
3.2 Inversion
[3.1] [a] Geometric inversion in the unit circle C: z ÞÑ IC (z) ≡ (1/z) leaves the points
of C fixed, and swaps the interior and exterior. When composed with conjugation, it
yields complex inversion, z ÞÑ (1/z). [b] Definition of geometric inversion z ÞÑ IK (z) in
a general circle K.
While stage (ii) is geometrically trivial, we shall see that the mapping in stage (i)
is filled with surprises; it is called3 geometric inversion, or simply inversion. Clearly,
the unit circle C plays a special role for this mapping: the inversion interchanges
the interior and exterior of C, while each point on C remains fixed (i.e., is mapped
to itself). For this reason we write the mapping as z ÞÑ IC (z) = (1/z), and we call
IC (a little more precisely than before) “inversion in C”.
This added precision in terminology is important because, as illustrated in [3.1b],
there is a natural way of generalizing IC to inversion in an arbitrary circle K (say
with centre q and radius R). Clearly, this “inversion in K”, written z ÞÑ e z = IK (z),
should be such that the interior and exterior of K are interchanged, while each point
on K remains fixed. If ρ is the distance from q to z, then we define e z = IK (z) to be
the point in the same direction from q as z, and at distance (R2 /ρ) from q. [Check for
yourself that this definition does indeed perform as advertised.]
As usual, we invite you to use a computer to verify empirically the many results
we shall derive concerning inversion. However, in the case of this particular map-
ping, you can also construct (fairly easily) a mechanical instrument that will carry
out the mapping for you; see Ex. 2.
Although we shall not need it for a while, it is easy enough to obtain a formula for
IK (z). Because the connecting complex numbers from q to z and to e z both have the
3
In older works it is often called “transformation by reciprocal radii”.
Inversion 141
[3.2] [a] Ordinary reflection in a line. [b] If the radius of K is large, so that it starts to look
straight, then geometric inversion z ÞÑ ez = IK (z) looks very much like ordinary reflection.
More precisely, as z approaches K, then IK (z) is ultimately its reflection in the tangent
to K. [c] Under inversion in K, the triangle aqb is similar to the triangle bqee a.
same direction, and their lengths are ρ and (R2 /ρ), it follows that (e
z−q)(z − q) = R2 .
Solving for e
z,
R2 q z + (R2 − |q|2 )
IK (z) = +q= . (3.4)
z−q z−q
For example, if we put q = 0 and R = 1, then we recover IC (z) = (1/z).
There is a very interesting similarity (which will deepen as we go on) between
inversion IK (z) in a circle K and reflection ℜL (z) in a line L. See [3.2a] and [3.2b].
First, L divides the plane into two pieces, or “components”, which are interchanged
by ℜL (z); second, each point on the boundary between the components remains
fixed; third, ℜL (z) is involutory or self-inverse, meaning that ℜL ◦ ℜL is the identity
mapping, leaving every point fixed. To put this last property differently, consider
a point z and its reflection e z = ℜL (z) in L. Such a pair are said to be “mirror
images”, or to be “symmetric with respect to L”. The involutory property says that
the reflection causes such a pair of points to swap places.
Check for yourself that IK (z) shares all three of these properties. Furthermore,
the black triangle in [3.2b] illustrates the fact that if K is large then the effect of IK on
a small shape close to K looks very much like ordinary reflection. [We will explain
this later, but you might like to check this empirically using a computer.] For these
reasons, and others still to come, IK (z) is often also called reflection in a circle, and
the pair of points z and e z = IK (z) are said to be symmetric with respect to K.
We end this subsection with two simple properties of inversion, the first of which
will serve as the springboard for the investigations that follow. Let us use the sym-
bol [cd] to stand for the distance |c − d| between two points c and d. We hope that
142 Möbius Transformations and Inversion
no confusion will arise from this, the square brackets serving as a reminder that
[cd] is not the product of the complex numbers c and d.
e = IK (b) are their
e = IK (a) and b
In [3.2c], a and b are two arbitrary points, and a
images under inversion in K. By definition, [qa][qe 2
a ] = R = [qb][qbe ], and so
e
[qa]/[qb] = [qb]/[qe
a].
Noting the common angle ∠aqb = ∠e e we deduce that
aqb,
e
e and b,
If inversion in a circle centred at q maps two points a and b to a
e a are similar. (3.5)
then the triangles aqb and bqe
Lastly, let us find the relationship between the separation [ab] of two points, and
the separation [eabe ] of their images under inversion. Using (3.5),
[e
abe ]/[ab] = [qb
e ]/[qa] = R2 /[qa][qb],
[3.3] The geometric inverse of a line is a circle through the centre of inversion. By virtue
of [3.2c], the illustrated right triangles are similar, so we see that the geometric inverse of
a line L is a circle that passes through the centre of the inversion, and that its tangent
there is parallel to L.
circle K1 (radius R1 ) centred at q, then it will hold for any other circle K2 (radius R2 )
centred at q.
Let z be an arbitrary point, and let e
z1 = IK1 (z) and e z2 = IK2 (z). Obviously e
z1 and
e
z2 are both in the same direction from q as z, and you can easily check that the ratio
of their distances from q is independent of the location of z:
[q e
z2 ]/[q e
z1 ] = (R2 /R1 )2 ≡ k, say.
Thus
where the “central dilation” Dkq [see p. 45] is an expansion (centred at q) of the
plane by a factor of k. It follows [exercise] that if (3.7) holds for K1 then it also holds
for K2 .
Look again at [3.3]. Since IK is involutory, it simply swaps the line and circle, and
so the image of any circle through q is a line not passing through q. But what happens to
a general circle C that does not pass through q? Initially, suppose that C does not
contain q in its interior. Figure [3.4] provides the beautiful answer:
[3.4] Geometric inversion maps circles to circles! This beautiful and important fact is
readily confirmed by first observing that [3.2c] implies that the two grey angles are equal
to each other, and, likewise, the two black angles are equal to each other. But this means
that the right angle at c, being the difference of the grey and black angles, implies the
illustrated right angle at e
c. Done!
it, as illustrated. Here a and b are the ends of a diameter of C, and they therefore
subtend a right angle at a general point c on C. To understand (3.9), first use (3.5)
to check that both the shaded angles are equal, and that both the black angles are
equal. Next look at the triangle abc, and observe that the external shaded angle at
a is the sum of the two illustrated internal angles: the right angle at c and the black
angle at b. It follows that ∠e aecbe = (π/2), and hence that a e are the ends of a
e and b
diameter of a circle through e c. Thus we have demonstrated (3.9) in the case where
C does not contain q. We leave it to you to check that the same line of reasoning
establishes the result in the case where C does contain q.
The result (3.7) is in fact a special limiting case of (3.9). Figure [3.5] shows a line
L, the point p on L closest to the centre q of the inversion, and a circle C tangent to
L at p. As its radius tends to infinity, C tends to L, and the image circle C e = IK (C)
tends to a circle through q.
Later we will be able to give a much cleaner way of seeing that (3.7) and (3.9) are
two aspects of a single result.
[3.5] Inversion of a line is the limit of inversion of a circle. Here we see that [3.3] is in
fact merely a special limiting case of the general result shown in [3.4]. As the radius of
C tends to infinity, so that C becomes the line L, the image circle C e = IK (C) becomes a
circle through q.
[3.6] Constructing the geometric inverse using orthogonal circles. [a] If C is orthogonal
to K, then it is mapped to itself, because the image must be another circle that again
passes through the fixed points a and b, and the tangent T at a is mapped into itself.
Thus it can only be C itself. [b] It follows that the geometric inverse e
z of a point z may
be constructed as the second intersection point of any two circles through z that are
orthogonal to K.
Figure [3.6a] illustrates two immediate consequences of this result. First, the disc
bounded by C is also mapped to itself, the shaded and hatched regions into which
K divides it being swapped by the inversion. Second, a line from q through a point
z on C intersects C for the second time at the inverse point e z.
Another consequence (the key result of this subsection) is the geometric con-
struction shown in [3.6b], the verification of which is left to you.
The inverse ez of z in K is the second intersection point of any two circles
that pass through z and are orthogonal to K.
146 Möbius Transformations and Inversion
[3.7] Inversion (reflection) in a line is the limit of inversion in a circle. [a] The reflection
e
z of z in the line L can be constructed as the second intersection point of any two circles
through z that are orthogonal to L. [b] The same construction for geometric inversion in
the circle K. As the radius of K tends to infinity, this becomes reflection in the tangent L.
Note that the construction of e z in [3.6a] is the special limiting case in which the
radius of one of the circles tends to infinity, and so becomes a line through q. For
other, less important, geometric constructions of inversion, see Ex. 1.
The previously mentioned analogy between inversion in K and reflection in a
line L now deepens, for the reflection e z = ℜL (z) of z in L can be obtained using
precisely the same construction; see [3.7a]. Note that the line-segment joining z and
e
z is orthogonal to L, and that its intersection p with L is equidistant from z and e z:
[pz]/[p e
z ] = 1.
As illustrated in [3.7b], the segment of L in the vicinity of p can be approximated
by an arc of a large circle K tangent to L at p. Here e z = IK (z) is the image under
inversion in K of the same point z as before. As you can see, there is virtually no
difference between the two figures. More precisely, as the radius of K tends to infin-
ity, inversion in K becomes reflection in L. In particular, [pz]/[p e
z ] tends to unity, or
equivalently, [p e
z ] is “ultimately equal” to [pz]. We can now understand what was
happening in figure [3.2b].
We can also check this result algebraically. First, though, observe that from the
geometric point of view it is sufficient to demonstrate the result for a single choice
of the line L and a single point p on it. Let us therefore choose L to be the real axis,
and let p be the origin. The circle K of radius R centred at q = iR is therefore tangent
to L at p. Using (3.4), we obtain [exercise]
z
IK (z) = .
1 − (iz/R)
[3.8] Conformal and anticonformal mappings preserve and reverse angles, respectively.
Again, we can also get this algebraically using the above equation. If R is fixed
and |z| < R, then [exercise]
iz2 z3
IK (z) = z + − 2 + ··· .
R R
Thus as z approaches p = 0, IK (z) is ultimately equal to ℜL (z) = z, which is
reflection in the tangent to K at p.
[3.9] Through any given point z not on K, there is precisely one circle orthogonal to K
that passes through z in any given direction.
are unable to say whether or not it preserves their sense, then we call it an isogonal
mapping.
It is easy enough to think of concrete mappings that are either conformal or anti-
conformal. For example, a translation z ÞÑ (z + c) is conformal, as is a rotation and
expansion of the plane given by z ÞÑ az. On the other hand, z ÞÑ z is anticonformal,
as is any reflection in a line. The analogy between such a reflection and inversion
in a circle now gets even deeper, for
To see this, first look at [3.9]. This illustrates the fact that given any point z not on
K, there is precisely one circle orthogonal to K that passes through z in any given
direction. [Given the point and the direction, can you think how to construct this
circle?]
As in [3.8], suppose that two curves S1 and S2 intersect at p, and that their tan-
gents there are T1 and T2 , the angle between them being θ. To find out what happens
to this angle under inversion in K, let us replace S1 and S2 with the unique circles
orthogonal to K that pass through p in the same directions as directions S1 and S2 ,
i.e., circles whose tangents at p are T1 and T2 . See [3.10a]. Since inversion in K maps
each of these circles to themselves, the new angle at p e = IK (p) is −θ. Done.
Figure [3.10b] illustrates the effect of z ÞÑ (1/z) on angles. Since this mapping
is equivalent to reflection (i.e., inversion) in the unit circle followed by reflection
in the real axis (both of which are anticonformal), we see that their composition
reverses the angle twice, restoring it to its original value:
Inversion maps any pair of orthogonal circles to another pair of orthogonal circles.
Of course if one of the circles passes through the centre of inversion then its image
will be a line. However, if we think of lines as merely being circles of infinite radius
then the result is true without qualification.
150 Möbius Transformations and Inversion
[3.11] Geometric inversion preserves symmetry. [a] If a and b are symmetric with respect
e in a line M are still symmetric with respect to
e and b
to a line L, then their reflections a
e
the reflection L of L: we say that symmetry is preserved. [b] Since geometric inversion
preserves circles and orthogonality, we see that if a and b are symmetric with respect to
e in a circle J are still symmetric with respect to
e and b
a circle K, then their “reflections” a
the reflection Ke of K: here, too, symmetry is preserved!
The preservation of symmetry result is now easily understood. See [3.11b]. Since
the two dashed circles through a and b are orthogonal to K, their images under
e and they therefore intersect in a pair of
inversion in J are likewise orthogonal to K,
e
points that are symmetric with respect to K.
[3.12] Inversion in a sphere sends a plane to a sphere through the centre of inversion.
Rotating [3.3] (in space) about the line through q and a, we see that inversion in the
sphere S maps the plane Π to a sphere that contains q and whose tangent plane there is
parallel to Π.
By the same token, if we rotate figure [3.4] about the line through q and a, then
we find that
Under inversion in a sphere, the image of a sphere that does not contain
the centre of inversion is another sphere that does not contain the centre of
inversion.
This result immediately tells us what will happen to a circle in space under inver-
sion in a sphere, for such a circle may be thought of as the intersection of two
spheres. Thus we easily deduce [exercise] the following result:
Under inversion in a sphere, the image of a circle C that does not pass
through the centre q of inversion is another circle that does not pass through
(3.12)
q. If C does pass through q then the image is a line parallel to the tangent
of C at q.
[3.13] Let S1 and S2 be intersecting spheres, and let C1 and C2 be the great circles in
which these spheres intersect any plane Π passing through their centres. Then S1 and S2
are orthogonal if and only if C1 and C2 are orthogonal. Also illustrated is the fact that if we
restrict attention to Π then geometric inversion of space in S1 induces two-dimensional
inversion in the circle C1 .
We now describe the steps leading to this result; they are closely analogous to the
steps leading to the two-dimensional preservation of symmetry result.
If we rotate figure [3.6a] about the line joining the centres of K and C, we deduce
that
Under inversion in a sphere K, every sphere orthogonal to K is mapped
(3.14)
to itself.
When we say that spheres are “orthogonal” we mean that their tangent planes
are orthogonal at each point of their circle of intersection. However, in order to
be able to easily draw on previous results, let us rephrase this three-dimensional
description in two-dimensional terms:
Let S1 and S2 be intersecting spheres, and let C1 and C2 be the great circles
in which these spheres intersect any plane Π passing through their centres.
Then S1 and S2 are orthogonal if and only if C1 and C2 are orthogonal.
See [3.13]. This figure is also intended to help you see that if we restrict attention to
Π then the three-dimensional inversion in S1 is identical to the two-dimensional
inversion in C1 . This way of viewing inversion in spheres allows us to quickly
generalize earlier results.
Three Illustrative Applications of Inversion 153
For example, referring back to [3.6b], we find—make sure you see this—that if p
e = IS1 (p) may be constructed as the second intersection point of any
lies in Π then p
two circles like C2 that (i) lie in Π, (ii) are orthogonal to C1 , and (iii) pass through p.
Next, suppose that S1 and S2 in [3.13] are subjected to inversion in a third sphere
K. Choose Π to be the unique plane passing through the centres of S1 , S2 , K, and let
C be the great circle in which K intersects Π. Since IC maps C1 and C2 to orthogonal
circles, we deduce [exercise] that (3.14) is a special case of the following result:
[3.14] Given any two circles A and B that touch at q, as shown, construct touching
circles, Ci , as shown. Then, very surprisingly, (1) the points of contact of the chain Ci all
lie on a circle that also touches A and B at q, and (2) the height of the centre of Cn above
L is 2n times the radius of Cn , as illustrated for n = 3.
154 Möbius Transformations and Inversion
[3.15] Referring to the previous figure, consider the illustrated (unique) circle K centred
at q that cuts C3 at right angles. Geometric inversion in K maps C3 to itself, and it maps A
and B to parallel vertical lines. The truth of both of the mysterious, bulleted facts below
suddenly becomes obvious!
The figure illustrates two remarkable facts about this chain of circles:
• The points of contact of the chain C0 , C1 , C2 , etc., all lie on a circle [dashed]
touching A and B at q.
• If the radius of Cn is rn , then the height above L of the centre of Cn is 2nrn . The
figure illustrates this for C3 .
Before reading further, see if you can prove either of these results using conven-
tional geometric methods.
Inversion allows us to demonstrate both these results in a single elegant swoop.
In [3.14], we have drawn the unique circle K centred at q that cuts C3 at right angles.
Thus inversion in K will map C3 to itself, and it will map A and B to parallel
vertical lines; see [3.15]. Check for yourself that the stated results are immediate
consequences of this figure.
4
I am grateful to my friend Paul Zeitz for challenging me with this problem, which appeared in the
USA Mathematical Olympiad.
Three Illustrative Applications of Inversion 155
[3.17] Geometric explanation of the previous figure. First, on the left, we use the con-
struction in [3.7a] to represent the reflection of q in an edge as the second intersection
point of any two circles through q whose centres lie on that edge. Here we have chosen
the centres of these circles to be the vertices of the quadrilateral. Shown on the right is
the result of performing an inversion in any circle centred at q. The orthogonal circles
become orthogonal lines, intersecting in a rectangle, so these four points lie on a circle.
But this means that the original four points were also on a circle. Done!
To demonstrate the result using inversion, we first use the construction in [3.7a]
to represent the reflection of q in an edge as the second intersection point of any two
circles through q whose centres lie on that edge. More precisely, let us choose the
centres of these circles to be the vertices of the quadrilateral; see the LHS of [3.17].
156 Möbius Transformations and Inversion
Note that, because the diagonals are orthogonal, a pair of these circles centred
at the ends of an edge will intersect orthogonally both at q and at the reflection of
q in that edge.
It follows that if we now apply an inversion in any circle centred at q, then a
pair of such orthogonal circles through q will be mapped to a pair of orthogonal
lines (parallel to the diagonals of the original quadrilateral); see the RHS of [3.17].
Thus the images of the four reflections of q are the vertices of a rectangle, and they
therefore lie on a circle. The desired result follows immediately. Why?
[3.18] [a] A cyclic quadrilateral is subject to Ptolemy’s Theorem: [ad] [bc] + [ab] [cd] =
[ac] [bd]. [b] Here is the result of inverting figure [a] in any circle K centred at a. Clearly,
ee
[b c ] + [e
cde ] = [b
ede ]. But (3.6) dictates how the separations of these inverted points
are related to the separations of the original points, and Ptolemy’s Theorem then follows
immediately!
The Riemann Sphere 157
Inverting figure [3.18a] in a circle K centred at one of the vertices (say a), we
obtain [3.18b], in which
ee
[b c ] + [e
cde ] = [b
ede ].
Recalling that (3.6) tells us how the separation of two inverted points is related to
the separation of the original points, we deduce that
[bc] [cd] [bd]
+ = .
[ab][ac] [ac][ad] [ab][ad]
Multiplying both sides by ([ab] [ac] [ad]), we deduce Ptolemy’s Theorem.
5
Some works instead define Σ to be tangent to the complex plane at its south pole.
6
This concept of “curvature” will be defined more precisely in Chapter 6.
The Riemann Sphere 159
[3.19] Stereographic Projection maps the surface of the (clear glass) sphere to its equa-
b on the surface of
torial plane. Standing at the north pole, N, fire a laser beam at a point p
the sphere: it goes on to hit the plane at p, called the stereographic projection (or image)
of pb . Conversely, p
b is symmetrically called the stereographic projection (or image) of p.
If p moves along a line in the plane, the laser beam sweeps out a plane, so the beam cuts
the sphere in a circle that passes through N, and its tangent there is parallel to the line
in the plane. This is because the equatorial plane and the tangent plane to the sphere at
N are parallel—both are horizontal—so the plane swept out by the laser beam intersects
these parallel planes in parallel lines.
complex numbers and points of Σ, we can imagine that the points of Σ are the com-
plex numbers. For example, S = 0 and N = ∞. Once stereographic projection has
been used to label each point of Σ with a complex number, Σ is called the Riemann
sphere.
We have already discussed the fact that a line in C may be viewed as a circle pass-
ing through the point at infinity. The Riemann sphere now transforms this abstract
idea into a literal fact:
To see this, observe that as p moves along the line shown in [3.19], the line con-
necting N to p sweeps out a plane through N. Thus p b moves along the intersection
of this plane with Σ, which is a circle passing through N. Done. In addition, note
that the tangent to this circle at N is parallel to the original line. This is explained in the
caption.
160 Möbius Transformations and Inversion
[3.20] Stereographic Projection is conformal. If two lines intersect in the plane, their
stereographic projections on the sphere are two circles that intersect at the same angle
at N, for the tangents to these circles at N are parallel to the lines in the plane. But, by
symmetry, the angle between these circles at p b is the same as their angle at N, which
we have just seen is the same as the angle between the lines in the plane. Done!
From this last fact it follows that stereographic projection preserves angles. Con-
sider [3.20], which shows two lines intersecting at p, together with their circular,
stereographic projections. By symmetry, the magnitude of the angle of intersec-
tion between the circles is the same at their two intersection points, p b and N. Since
their tangents at N are parallel to the original lines in the plane, it follows that the
illustrated angles at p and p b are of equal magnitude. But before we can say that
stereographic projection is “conformal”, we must assign a sense to the angle on the
sphere.
According to our convention, the illustrated angle at p (from the black curve to
the white one) is positive, i.e., it is counterclockwise when viewed from above the
plane. From the perspective from which we have drawn [3.20], the angle at p b is
negative, i.e., clockwise. However, if we were looking at this angle from inside the
sphere then it would be positive. Thus
for another 1500 years! This was first done around 1590 by Thomas Harriot.7 In
Chapter 6 we shall meet Harriot again, for in 1603 he made another profoundly
important discovery about the geometry of the sphere, this time relating the angles
in a spherical triangle to the triangle’s area—(6.9) on page 317.]
Clearly, any origin-centred circle in the plane is mapped to a horizontal circle
on Σ, but what happens to a general circle? The startling answer is that it too is
mapped to a circle on the Riemann sphere! This is quite difficult to see if we stick to our
original definition of stereographic projection, but it suddenly becomes obvious
if we change our point of view. Look again at [3.12], and observe how closely it
resembles the definition of stereographic projection.
To make the connection precise, let K be the sphere centred at the north pole N
of Σ that intersects Σ along its equator (the unit circle of C). Figure [3.21a] shows a
vertical cross section (through N and the real axis), of K, Σ, and C. The full three-
dimensional picture is obtained by rotating this figure about the line through N
and S. We now see that
√
If K is the sphere of radius 2 centred at N, then stereographic projection
is the restriction to C or Σ of inversion in K.
7
See Stillwell (2010, §16.2). For a short sketch of Harriot’s life, see Stillwell (2010, §17.7). The first
full biography of this very remarkable, unsung hero of mathematics and science—born within a mile
of Stephen Hawking’s place of birth (and mine!)—has at last been published: Arianrhod (2019).
162 Möbius Transformations and Inversion
Note that (3.16) could also have been derived from (3.12) in this way.
For our next example, consider z ÞÑ e z = (1/z), which is inversion in the unit
circle C. Figure [3.21b] shows a vertical cross section of Σ taken through N and the
point z in C. This figure also illustrates the very surprising result of transferring
this inversion to Σ:
Here is an elegant way of seeing this. First note that not only are the pair of points
z and e z symmetric (in the two-dimensional sense) with respect to C, but they
are also symmetric (in the three-dimensional sense) with respect to the sphere Σ.
Now apply the three-dimensional preservation of symmetry result (3.13). Since z
and ez are symmetric with respect to Σ, their stereographic images b z = IK (z) and
b
e
z = IK ( e
z ) will be symmetric with respect to IK (Σ). But IK (Σ) = C. Done! A more
elementary (but less illuminating) derivation may be found in Ex. 6.
By combining the above results, we can now find the effect of complex inversion
on the Riemann sphere. In C, we know that z ÞÑ (1/z) is equivalent to inversion
in the unit circle, followed by complex conjugation. The induced mapping on Σ
is therefore the composition of two reflections in perpendicular planes through the real
axis—one horizontal, the other vertical. However, it is not hard to see (perhaps
with the aid of an orange) that the net effect of successively reflecting Σ in any
The Riemann Sphere 163
two perpendicular planes through the real axis is a rotation of Σ about the real axis
through angle π. Thus we have shown that
The mapping z ÞÑ (1/z) in C induces a rotation of the Riemann sphere
(3.19)
about the real axis through an angle of π.
Recall that the point ∞ was originally defined by the property that it be swapped
with 0 under complex inversion, z ÞÑ (1/z). The result (3.19) vividly illustrates the
correctness of identifying N with the point at infinity, for the point 0 in C corre-
sponds to the south pole S of Σ, and the rotation of π about the real axis does indeed
swap S with N.
[3.22] The conformality of z ÞÑ zn . In the next chapter we will begin to address the
remarkable and mysterious fact that z ÞÑ zn is conformal, here illustrated in the case
n = 2 [TOP]. We will also see that conformality suffices to explain the illustrated fact
that a grid of small squares is ultimately mapped to a grid of squares, in the limit that
they vanish. Note, however, that it is not conformal at 0: the angles between the rays are
doubled. We see this too at the south pole of the Riemann sphere, and at the north pole,
corresponding to ∞. Since stereographic projection has been proven to be conformal,
it follows that the conformality of z ÞÑ z2 is transferred to the induced mapping of the
Riemann sphere [BOTTOM], and therefore it, too, ultimately maps squares to squares.
8
This is the American convention; in my native England the roles of θ and ϕ are the reverse of
those stated here.
The Riemann Sphere 167
[3.23] Two stereographic formulae. [a] From the similarity of the illustrated triangles, we
immediately obtain a Cartesian formula for z = x + iy = X+iY 1−Z in terms of the Cartesian
coordinates (X, Y, Z) of the point b z on the sphere. This can be inverted to find b
z in terms
of z; see (3.21). [b] If we instead use geographical coordinates (θ, ϕ) to describe b
z on the
sphere, the figure shows that we obtain a particularly elegant and useful polar formula:
z = cot(ϕ/2) eiθ .
Put differently, q = −IC (p), where C is the unit circle. Note that the relationship
between p and q is actually symmetrical (as clearly it should be): p = −(1/q). To
b has coordinates (ϕ, θ) then q
verify (3.23), first check for yourself that if p b has
coordinates (π − ϕ, π + θ). The remainder of the proof is almost identical to the
previous calculation. For an elementary geometric proof, see Ex. 6.
We know that a circle C will map to a circle—of course lines are now included as
“circles”—but what will happen to the disc bounded by C? First we give a useful
way of thinking about this disc. Imagine yourself walking round C moving coun-
terclockwise; your motion gives C what is a called a positive sense or orientation. Of
the two regions into which this positively oriented circle divides the plane, the disc
may now be identified as the one lying to your left.
Now consider the effect of the four transformations in (3.3) on the disc and on
the positively oriented circle bounding it. Translations, rotations, and expansions
all preserve the orientation of C and map the interior of C to the interior of the
image C e of C. However, the effect of complex inversion on C depends on whether
or not C contains the origin. If C does not contain the origin, then Ce has the same
e This is easily
orientation as C, and the interior of C is mapped to the interior of C.
understood by looking at [3.24].
If C does contain the origin then Ce has the opposite orientation and the interior
e
of C is mapped to the exterior of C. If C passes through the origin then its interior
e See [3.25].
is mapped to the half-plane lying to the left of the oriented line C.
9
Remarkably, Carathéodory (1937) proved that this property actually characterizes the Möbius
transformations: they alone have this property!
Möbius Transformations: Basic Results 169
[3.25] A limiting form of the previous figure has the image disc expand to become a half-
plane, but it is still true that the region to the left of C is mapped to the region to the left
e In the next chapter we will see that this a more universal phenomenon, having little
of C.
to do with the specific geometry of Möbius transformations.
To summarize,
A Möbius transformation maps an oriented circle C to an oriented circle
e in such a way that the region to the left of C is mapped to the region to
C (3.24)
e
the left of C.
the Möbius transformation. In geometric terms, this would mean that to specify a
particular Möbius transformation we would need to know the images of any four
distinct points. This is wrong.
If k is an arbitrary (non-zero) complex number then
az + b kaz + kb
= M(z) = .
cz + d kcz + kd
In other words, multiplying the coefficients by k yields one and the same mapping,
and so only the ratios of the coefficients matter. Since three complex numbers are suf-
ficient to pin down the mapping—(a/b), (b/c), (c/d), for example—we conjecture
(and later prove) that
There exists a unique Möbius transformation sending any three points
(3.25)
to any other three points.
In the course of gradually establishing this one result we shall be led to further
important properties of Möbius transformations.
If you read the last section of Chapter 1, then (3.25) may be ringing a bell: the
similarity transformations needed to do Euclidean geometry are also determined
by their effect on three points. Indeed, we saw in that chapter that such similarities
can be expressed as complex functions of the form f(z) = az + b, and so they actu-
ally are Möbius transformations, albeit of a particularly simple kind. However, for
such a similarity to exist, the image points must form a triangle that is similar to the
triangle formed by the original points. But in the case of Möbius transformations
there is no such restriction, and this opens the way to more flexible, non-Euclidean
geometries in which Möbius transformations play the role of the “motions”. This
is the subject of Chapter 6.
Let us make a further remark on the non-uniqueness of the coefficients of a
Möbius transformation. Recall from the beginning of this chapter that the inter-
esting Möbius transformations are the non-singular ones, for which (ad − bc) ̸= 0.
For if (ad − bc) = 0 then M(z) = az+b cz+d crushes the entire plane down to the
single point (a/c). If M is non-singular, then we may multiply its coefficients by
√
k = ±1/ ad − bc, in which case the new coefficients satisfy
(ad − bc) = 1;
the Möbius transformation is then said to be normalized. When investigating the
properties of a general Möbius transformation, it turns out to be very convenient to
work with this normalized form. However, when doing calculations with specific
Möbius transformations, it is usually best not to normalize them.
is also one-to-one and onto. This means that if we are given any point w in the w-
plane, there is one (and only one) point z in the z-plane that is mapped to w. We
can show this by explicitly finding the inverse transformation w ÞÑ z = M−1 (w).
Solving the above equation for z in terms of w, we find [exercise] that M−1 is also
a Möbius transformation:
dz − b
M−1 (z) = . (3.26)
−cz + a
Note that if M is normalized, then this formula for M−1 is automatically normalized
as well.
If we look at the induced mapping on the Riemann sphere, then we find that a
Möbius transformation actually establishes a one-to-one correspondence between
points of the complete z-sphere and points of the complete w-sphere, including
their points at infinity. Indeed you may easily convince yourself that
Using (3.26), you may check for yourself that M−1 (a/c) = ∞ and M−1 (∞) =
−(d/c).
Next, consider the composition M ≡ (M2 ◦ M1 ) of two Möbius transformations,
a2 z + b 2 a1 z + b1
M2 (z) = and M1 (z) = .
c2 z + d2 c1 z + d1
A simple calculation [exercise] shows that M is also a Möbius transformation:
(a2 a1 + b2 c1 )z + (a2 b1 + b2 d1 )
M(z) = (M2 ◦ M1 )(z) = . (3.27)
(c2 a1 + d2 c1 )z + (c2 b1 + d2 d1 )
It is clear geometrically that if M1 and M2 are non-singular, then so is M. This is
certainly not obvious algebraically, but later in this section we shall introduce a
new algebraic approach that does make it obvious.
If you have studied “groups”, or if you read the final section of Chapter 1, then
you will realize that we have now established the following: The set of non-singular
Möbius transformations forms a group under composition. For, (i) the identity mapping
E(z) = z belongs to the set; (ii) the composition of two members of the set yields a
third member of the set; (iii) every member of the set possesses an inverse that also
lies in the set.
or that it “remains fixed”. Note that under the identity mapping, z ÞÑ E(z) = z,
every point is a fixed point.
By definition, then, the fixed points of a general Möbius transformation M(z)
are the solutions of
az + b
z = M(z) = .
cz + d
Since this is merely a quadratic in disguise, we deduce that
From the above result it follows that if a Möbius transformation is known to have
more than two fixed points, then it must be the identity. This enables us to establish
the uniqueness part of (3.25). Suppose that M and N are two Möbius transforma-
tions that both map the three given points (say q, r, s) to the three given image
points. Since (N−1 ◦ M) is a Möbius transformation that has q, r, and s as fixed
points, we deduce that it must be the identity mapping, and so N = M. Done.
We now describe the fixed points explicitly. If M(z) is normalized, then the two
fixed points ξ+ , ξ− are given by [exercise]
p
(a − d) ± (a + d)2 − 4
ξ± = . (3.28)
2c
In the exceptional case where (a + d) = ±2, the two fixed points ξ± coalesce into
the single fixed point ξ = (a − d)/2c. In this case the Möbius transformation is
called parabolic.
[3.26] Classification of Möbius and Lorentz transformations. Each of the four types
of transformation has two names, depending on whether it is viewed as acting on C
[name on the left] or on spacetime [name on the right]. The mathematical classification
depends on analysing the fixed points, and this is done in detail in Section 3.7. For
the technical details of the spacetime interpretation via Lorentz transformations, see
Needham (2021, §6.4).
through these fixed points (which are orthogonal to the invariant circles) are per-
muted among themselves. This pure rotation is the simplest, archetypal example
of a so-called elliptic Möbius transformation.
With ρ > 1, figure [3.26b] illustrates the induced transformation on Σ corre-
sponding to the origin-centred expansion of C, z ÞÑ ρz. If ρ < 1 then we have a
contraction of C, and points on Σ move due South instead of due North. Again it
is clear that the fixed points are 0 and ∞, but the roles of the two families of curves
in [3.26a] are now reversed: the invariant curves are the great circles through the
fixed points at the poles, and the orthogonal horizontal circles are permuted among
themselves. This pure expansion is the simplest, archetypal example of a so-called
hyperbolic Möbius transformation.
Figure [3.26c] shows the combined effect of the rotation and expansion in [3.26a]
and [3.26b]. Here the invariant curves are the illustrated “spirals”; however, the two
families of circles in [3.26a] (or [3.26b]) are both invariant as a whole, in the sense
that the members of each family are permuted among themselves. This rotation
and expansion is the archetypal loxodromic Möbius transformation, of which the
elliptic and hyperbolic transformations are particularly important special cases.
Finally, [3.26d] illustrates a translation. Since the invariant curves in C are the
family of parallel lines in the direction of the translation, the invariant curves on
Σ are the family of circles whose common tangent at ∞ is parallel to the invariant
lines in C. Since ∞ is the only fixed point, a pure translation is an example of a
parabolic Möbius transformation.
Note the following consequence of the above discussion:
Later we will use this to show that each Möbius transformation is equivalent, in a
certain sense, to one (and only one) of the four types shown in [3.26].
M = M−1
ee
q e ◦ Mqrs
rs
Möbius Transformations: Basic Results 175
that
(z − q)(r − s)
[z, q, r, s] = .
(z − s)(r − q)
This is not quite so rabbit-like as it appears. Two hundred years prior to Möbius’
investigations, Girard Desargues had discovered the importance of the expression
[z, q, r, s] within the subject of projective geometry, where it was christened the
cross-ratio of z, q, r, s (in this order10 ). Its significance in that context is briefly
explained in Ex. 14, but the reader is urged to consult Stillwell (2010) for greater
detail and background. We can now restate (3.25) in a more explicit form:
The unique Möbius transformation z ÞÑ w = M(z) sending three points
e, er, e
q, r, s to any other three points q s is given by
e)(er − e
(w − q s) (z − q)(r − s)
e, er, e
= [w, q s] = [z, q, r, s] = . (3.30)
(w − e e)
s)(er − q (z − s)(r − q)
Although we have not done so, in any concrete case one could easily go on to solve
this equation for w, thereby obtaining an explicit formula for w = M(z).
The result (3.30) may be rephrased in various helpful ways. For example, if a
Möbius transformation maps four points p, q, r, s to p e, q
e, er, e
s (respectively) then the
cross-ratio is invariant: [e e, er, e
p, q s] = [p, q, r, s]. Conversely, p, q, r, s can be mapped
e, q
to p e, er, e
s by a Möbius transformation if their cross-ratios are equal.
Recalling (3.24), we also obtain the following:
Let C be the unique circle through the points q, r, s in the z-plane,
oriented so that these points succeed one another in the stated order.
e be the unique oriented circle through q
Likewise, let C e, er, e
s in the w-
e (3.31)
plane. Then the Möbius transformation given by (3.30) maps C to C,
and it maps the region lying to the left of C to the region lying to the
e
left of C.
10
Different orders yield different values; see Ex. 16. Unfortunately, there is no firm convention as
to which of these values is “the” cross-ratio. For example, our definition agrees with Carathéodory
(1964), Penrose and Rindler (1984), and Jones and Singerman (1987), but it is different from the equally
common definition of Ahlfors (1979).
176 Möbius Transformations and Inversion
[3.27] There exists a unique Möbius transformation that maps three ordered points on
e and the region to the left of C
the circle C to three other ordered points on the circle C,
e
is mapped to the region to the left of C.
Im [p, q, r, s] = 0. (3.32)
Möbius Transformations as Matrices* 177
Since the coefficients of the Möbius transformation are not unique, neither is the
corresponding matrix: if k is any non-zero constant, then the matrix k[M] corre-
sponds to the same Möbius transformation as [M]. However, if [M] is normalized
by imposing (ad − bc) = 1, then there are just two possible matrices associated
with a given Möbius transformation: if one is called [M], the other is −[M]; in other
words, the matrix is determined “uniquely up to sign”. This apparently trivial fact
turns out to have deep significance in both mathematics and physics; see Penrose
and Rindler (1984, Ch. 1) and Penrose (2005).
At this point there exists a strong possibility of confusion, so we issue the follow-
ing WARNING: In linear algebra we are—or should be!—accustomed to thinking
of a real 2 × 2 matrix as representing a linear transformation of R2 . For example,
!
0 −1
represents a rotation of the plane through (π/2). That is, when we apply it
1 0
to a vector yx in R2 , we obtain
! ! ! !
0 −1 x −y x
= = rotated by (π/2) .
1 0 y x y
a b
In stark contrast, the matrix c d
corresponding to a Möbius transformation
generally has complex numbers as its entries, and so it cannot be interpreted as a
linear transformation of R2 . Even if the entries are real, it must not be thought of in
178 Möbius Transformations and Inversion
!
this way. For example, the matrix 01 −10 corresponds to the Möbius transforma-
tion M(z) = −(1/z), which is certainly not a linear transformation of C. To avoid
confusion, we will adopt the following notational convention: We use (ROUND)
brackets for a real matrix corresponding to a linear transformation of R2 or of C, and
we use [SQUARE] brackets for a (generally) complex matrix corresponding to a Möbius
transformation of C.
Despite this warning, we have the following striking parallels between the
behaviour of Möbius transformations and the matrices that represent them:
[M−1 ] = [M]−1 .
But look at (3.27)! This is simply the matrix of the composite Möbius transforma-
tion (M2 ◦ M1 )(z). Thus multiplication of Möbius matrices corresponds to composition
of Möbius transformations:
Sir Roger Penrose. See Penrose and Rindler (1984, Ch. 1), Penrose (2005), and Shaw
(2006, Ch. 23).
Note that one immediate benefit of this approach is that there is no longer any
real distinction between a finite fixed point and a fixed point at ∞, for the latter
z1
merely corresponds to an eigenvector of the form . For example, consider how
0
Möbius Transformations as Matrices* 181
elegantly we may rederive the fact that ∞ is a fixed point if and only if M(z) is a
similarity transformation. If ∞ is a fixed point then
" # " #" # " #
z1 a b z1 a z1
λ = = .
0 c d 0 c z1
{k[M]} z = kλ z.
Since the eigenvalue does depend on the arbitrary choice of k, it appears that its
value can have no bearing on the geometric nature of the mapping M(z). Very
surprisingly, however, if [M] is normalized then the exact opposite is true! In the
next section we will show that the eigenvalues of the normalized matrix [M] completely
determine the geometric nature of the corresponding Möbius transformation M(z). In
anticipation of this result, let us investigate the eigenvalues further.
Recall the fact that the eigenvalues of [M] are the solutions of the so-called
char-
1 0
acteristic equation, det{[M] − λ[E]} = 0, where [E] is the identity matrix 0 1 . Using
the fact that [M] is normalized, we find [exercise] that the characteristic equation is
λ2 − (a + d)λ + 1 = 0,
1
λ+ = a + d. (3.34)
λ
The first thing we notice about this equation is that there are typically two eigen-
values, λ1 and λ2 , and they are determined solely by the value of (a + d). By
inspecting the coefficients of the quadratic we immediately deduce that
Aficionados of linear algebra will recognize (3.35) as a special case of the fol-
lowing general result on the eigenvalues λ1 , λ2 , . . . , λn of any n × n matrix
N:
λ1 λ2 . . . λn = det N and λ1 + λ2 + · · · + λn = tr N,
where tr N ≡ (the sum of the diagonal elements of N) is called the trace of N. For
future use, recall the following nice property of the trace function: If N and P are
both n × n matrices, then
tr {NP} = tr {PN}. (3.36)
In the case of 2 × 2 matrices (which is all that we shall ever need) this is easily
verified by a direct calculation [exercise].
⟨p , q⟩ = p1 q1 + p2 q2 = 0.
What does this “orthogonality” mean in terms of the points p = (p1 /p2 ) and
q = (q1 /q2 ) whose homogeneous coordinate vectors are p and q? The answer is
surprising. As you may easily check, the above equation says that q = −(1/p), and
so from (3.23) we deduce that
Two vectors in C2 are orthogonal if and only if they are the homogeneous
coordinates of antipodal points on the Riemann sphere.
In particular, [R] maps every pair of orthogonal vectors to another such pair, and
R(z) therefore maps every pair of antipodal points on Σ to another such pair. We
shall not attempt a real proof, but since the transformation of Σ is also known to be
continuous and conformal11 , it can only be a rotation of Σ.
The desired invariance of the inner product (3.37) may be neatly rephrased using
an operation called the conjugate transpose, denoted by a superscript ∗. This opera-
tion takes the complex conjugate of each element in a matrix and then interchanges
the rows and columns:
" #∗ " #∗ " #
∗ p1 ∗ a b a c
p = = [p1 , p2 ] and [R] = = .
p2 c d b d
Since the inner product can now be expressed in terms of ordinary matrix multi-
plication as ⟨p , q⟩ = p∗ q, and since [exercise] {[R]p}∗ = p∗ [R]∗ , we find that (3.37)
takes the form
p∗ {[R]∗ [R]} q = p∗ q.
11
If it were not continuous then it could, for example, exchange points on two antipodal patches of
Σ while leaving the remainder fixed. If it were continuous but anticonformal, then it could map each
point to its antipodal point, or to its reflection in a plane through the centre of Σ.
184 Möbius Transformations and Inversion
matrices, we can easily find the most general unitary matrix [R] by re-expressing
(3.38) as [R]∗ = [R]−1 :
" # " # " #
a c d −b a b
= =⇒ [R] = .
b d −c a −b a
Although we have left some unsatisfactory gaps in the above reasoning, we have
nevertheless arrived at an important truth: The most general rotation of the Riemann
sphere can be expressed as a Möbius transformation of the form
az + b
R(z) = . (3.39)
−bz + a
[3.29] The geometric idea behind the classification of the Möbius transformations. Given
a Möbius transformation M(z) with two fixed points, ξ+ and ξ− , two families of circles
naturally arise: the dashed family C1 passing through the fixed points, and the orthogonal
family C2 , such that ξ+ and ξ− are symmetric with respect to each member of C2 . Note
that M(z) must map members of C1 amongst themselves, and likewise members of C2
are mapped amongst themselves. The key idea is to now apply a Möbius transformation
z−ξ+
F(z) = z−ξ −
that sends one fixed point to the origin (south pole of the Riemann sphere)
and the other to ∞ (the north pole). Thus, on the right, the circles C1 become rays, and
the circles C2 become origin-centred circles. Then the induced mapping M f = F◦M◦
F clearly has fixed points 0 and ∞, and must therefore take the form M(
−1 f ez) = me z,
iα
where m = ρ e is called the multiplier; in the illustrated case, ρ = 1 and α = (π/3).
This multiplier m completely characterizes the original Möbius transformation, M, and
therefore serves to classify it. In this manner, we obtain the first three archetypes shown
in [3.26].
RHS of [3.29] shows the image of the LHS under such a Möbius transformation,
the simplest example of which is
z − ξ+
F(z) = .
z − ξ−
[Note that we have not bothered to write this in normalized form.] Since F is
a Möbius transformation, it must map the members of C1 to the circles passing
through 0 and ∞, i.e., to lines through the origin [shown dashed]. Furthermore,
since F is conformal, two such lines must contain the same angle at 0 as the corre-
sponding C1 circles do at ξ+ . We have tried to make this easy to see in our picture
by drawing C1 circles passing through ξ+ in evenly spaced directions, each one
making an angle of (π/6) with the next.
As an aside, observe that we now have a second, simpler explanation of the
existence of the family C2 of circles orthogonal to C1 . Since the illustrated set of
origin-centred circles are orthogonal to lines through 0, their images under F−1
must be circles orthogonal to each member of C1 .
Next, let e e = F(w) be the images under F of z and w = M(z). We
z = F(z) and w
may now think of F as carrying the original Möbius transformation z ÞÑ w = M(z)
186 Möbius Transformations and Inversion
and so
f = F ◦ M ◦ F−1 .
M (3.40)
12
Shading inspired by Ford (1929).
Visualization and Classification* 187
Just as in the case of a hyperbolic transformation, note that one fixed point is
repulsive while the other is attractive. In this figure we have taken α > 0 and ρ > 1;
how would it look if α were negative, or if ρ were less than one?
Visualization and Classification* 189
less than one. However, it is not so clear that this number is precisely (1/ρ), as we
know it must be. This too can be demonstrated geometrically, but let us instead
content ourselves with showing how our original algebraic argument may be re-
interpreted geometrically in terms of the “local effect” of M in the vicinity of each
of the fixed points.
Let us write Z = (z−ξ+ ) and W = (w−ξ+ ) for the complex numbers emanating
from ξ+ connecting that point to z and to its image w = M(z). We have claimed
(and partially verified) that if Z is infinitesimal then the effect of M is to rotate Z by
α and to expand it by ρ: in other words, W = m Z. To verify this, note that (3.42)
can be rewritten as
W w − ξ−
=m .
Z z − ξ−
As Z tends to zero, both z and w tend to ξ+ , and so the fraction on the right is
ultimately equal to m . Thus W is ultimately equal to m Z, as was to be shown.
After you have read the next chapter, you will be able to look back at what we
have just done and recognize it as an example of differentiating a complex function.
(by symmetry) orthogonal at their second intersection point. The RHS illustrates
what happens when we send ξ to ∞ by means of the Möbius transformation
1
G(z) = .
z−ξ
Clearly [exercise], the two orthogonal families of circles become two orthogonal
families of parallel lines. Conversely, if we apply G−1 to any two orthogonal fam-
ilies of lines on the right, then on the left we get two orthogonal families of circles
through ξ.
As before, let e
z = G(z) and w e = G(w) be the images on the RHS of z and w =
M(z). Thus the Möbius transformation z ÞÑ w = M(z) on the LHS induces another
Möbius transformation e z ÞÑ w f e
e = M( z ) on the RHS, where
f = G ◦ M ◦ G−1 .
M
f we deduce that M
Since ∞ is the sole fixed point of M, f can only be a translation:
f e
M( z) = e
z + T.
Now suppose that the arrows on the RHS of [3.33] represent the direction of the
translation T . As illustrated, we now draw a grid aligned with T , each shaded
square being carried into the next by M. f On the LHS of [3.33] we thus obtain a
vivid picture of the action of the original parabolic Möbius transformation M: each
solid circle is carried into itself; each dashed circle is carried into another dashed
circle; and each shaded region is carried into the next in the direction of the arrows.
If M(z) = az+b
cz+d is normalized, then we know from (3.28) that it is parabolic if
and only if (a + d) = ±2, in which case ξ = (a − d)/2c. Now let us determine the
corresponding translation T in terms of the coefficients. Since (G ◦ M) = (M f ◦ G),
the so-called normal form of M is given by
1 1
= + T.
w−ξ z−ξ
Since M maps z = ∞ to w = (a/c), we deduce that
1
T= = ±c,
(a/c) − ξ
where the “±” is the arbitrarily chosen sign of (a + d).
Thus, regardless of whether or not [F] is normalized, [M] is normalized if and only
f is normalized. Since M(z)
if [M] f f =
= m z, its normalized matrix is [exercise] [M]
√
m 0
√ . Recalling (3.36), we deduce that
0 1/ m
√ 1
m + √ = tr [F] [M] [F]−1 = tr [F]−1 [F][M] = tr [M] = a + d,
m
as was to be shown.
Möbius transformation M(z). We also claimed that if [M] is normalized then the
eigenvalues completely determine the character of M(z). We can now be more
precise:
Before proving this result, we illustrate it with the example of complex inver-
sion, z ÞÑ (1/z). We already know that the fixed points are ±1, that the associated
multipliers are both
given
by m = −1, and we easily find [exercise] that the nor-
0 i
malized matrix is i
. If we choose the homogeneous coordinate vector of a
0
" #
z
finite point z to be , then the eigenvectors corresponding to the fixed points
1
" #
±1
z = ±1 are . Since
1
" #" # " # " #" # " #
0 i 1 1 0 i −1 −1
=i and = −i ,
i 0 1 1 i 0 1 1
we see that the eigenvalues are given by λ = ±i, in agreement with (3.46).
√
Returning to the general case, comparison of (3.34) and (3.44) reveals that m
and λ satisfy the same quadratic, so we immediately deduce most of (3.46): the two
reciprocal values of m are equal to the two reciprocal values of λ2 . However, this
does not tell us which value of λ2 yields which value of m , nor is this line of attack
very illuminating. Here, then, is a more transparent approach.
We begin by recalling a standard result of linear algebra, which is valid for n × n
matrices:
Let us return to [3.29], in which the fixed point ξ+ of M (with associated mul-
tiplier m + ) was mapped to the fixed point 0 of M f = (F ◦ M ◦ F−1 ) by means of
z ÞÑ e z = F(z) = z−ξz−ξ− . In terms of linear transformations of C , the eigenvector
+ 2
ξ+ 0
of [M] is being mapped by [F] to the eigenvector of
1 1
" #
ξ+
The linear algebra result now tells us that if λ+ denotes the eigenvalue of ,
1
then
" # " #
f 0 0
[M] = λ+ .
1 1
This is true irrespective of whether or not any of the matrices in the above equation
are normalized.
Now suppose that [M] is normalized, as demanded in (3.46). Irrespective of
whether or not [F] is normalized, we have already noted that [M] is normalized
f is normalized. Since the normalized matrix of M(
if and only if [M] f e z) = m+e z is
√
f = m+ 0
given by [M] √ , we deduce that
0 1/ m +
" # "√ #" # " #
0 m+ 0 0 1 0
λ+ = √ =√ .
1 0 1/ m + 1 m+ 1
Thus m + = 1/λ2+ , as was to be shown.
3.8.1 Introduction
Recall from (3.4) that the formula for inversion or “reflection” in a circle K has the
form
Az + B
IK (z) = .
Cz + D
It follows easily that the composition of any two reflections (in circles or lines) is
a Möbius transformation. Since the composition of two Möbius transformations is
another Möbius transformation, it follows more generally that the composition of an
even number of reflections is a Möbius transformation.
Conversely, in this section we will use the Symmetry Principle [see p. 168] to
show that
Every non-loxodromic Möbius transformation can be expressed as the
composition of two reflections, and every loxodromic Möbius transfor-
mation can be expressed as the composition of four reflections.
In the following, it would be helpful (but not essential) for you to have read the
final section of Chapter 1.
f = ℜe ◦ ℜ e .
M B A
In particular, ℜA e
e maps the dark “rectangle” abutting the line A to the light “rect-
angle”, then ℜBe maps this to the dark “rectangle” abutting the line B. e The figure
tries to make this clear by also showing the successive images of both a point and
a diagonal circular arc of the original dark “rectangle”.
Now think what this means on the LHS of [3.34]. The Symmetry Principle tells us
that if two points are symmetric with respect to the line Ae then their images under
the Möbius transformation F−1 are symmetric with respect to the circle A = F−1 (A) e
through the fixed points. [Recall that in [3.29] the family of such circles was called
e on the RHS becomes reflection (i.e., inversion) in A on the
C1 .] Thus reflection in A
LHS. Of course the same goes for the second reflection in B. e Thus we have shown
the following:
As in [3.34], the RHS of [3.35] illustrates the successive effect of these two reflec-
tions on a dark rectangle abutting A. e Just as before, the Symmetry Principle
−1
applied to F tells us that the original Möbius transformation on the LHS can be
expressed as
M = IB ◦ IA .
Recall from [3.29] that A and B belong to the family C2 of circles orthogonal
to the family C1 of circles through the fixed points. At the time, we pointed out
an equivalent property of C2 , namely, that the fixed points ξ± are symmetric with
respect to each member of C2 ; this enables us to explain how it is that (IB ◦ IA )
leaves ξ+ and ξ− fixed. In the case of [3.34], this was obvious because each reflection
separately left those points fixed; in the present case, however, IA swaps the points,
then IB swaps them back again, the net effect being to leave them fixed.
In the case of an elliptic transformation, (3.47) describes how to pick out a pair
of C1 circles corresponding to any given angle α. In the present case of a hyperbolic
transformation, how are we to pick out a pair of C2 circles corresponding to any
given value of ρ? The answer depends on a third characterizing property of the C2
circles: they are the circles of Apollonius with limit points ξ± .
This terminology reflects Apollonius’ remarkable discovery (c. 250 BCE) that if
a point z moves in such a way that the ratio of the distances of z from two fixed
points ξ± remains constant, then z moves on a circle. Figure [3.35] makes this easy
to understand. As z travels round A, e z = F(z) travels round the origin-centred circle
e of radius rA . But this constant rA is none other than the ratio of the distances of
A
z from two fixed points ξ± :
|z − ξ+ |
rA = | e
z | = |F(z)| = .
|z − ξ− |
Note that this also explains the “limit point” terminology: as the ratio rA tends to
0, the corresponding Apollonian circle A shrinks down towards the limit point ξ+ ;
as rA tends to infinity, A shrinks down towards the other limit point ξ− . Another
bonus of our discussion is a result that is frequently not mentioned in geometry
texts: the limit points defining a family of Apollonian circles are symmetric with respect to
each of these circles.
Since the quantities rA and rB occurring in (3.48) are now expressible purely in
terms of the geometry of the LHS of [3.35], we have solved the problem of picking
an appropriate pair of C2 circles:
3.8.5 Summary
Lest the details obscure the simplicity of what we have discovered, we summarize
our results as follows:
A non-loxodromic Möbius transformation M can always be decomposed
into two reflections in circles A and B that are orthogonal to the invari-
(3.49)
ant circles of M. Furthermore, M is elliptic, parabolic, or hyperbolic
according as A and B intersect, touch, or do not intersect.
Recalling (3.41), we also deduce that a loxodromic Möbius transformation M can
always be decomposed into four reflections in circles:
M = {IB ′ ◦ IA ′ } ◦ {IB ◦ IA } = {IB ◦ IA } ◦ {IB ′ ◦ IA ′ } ,
where A and B both pass through the fixed points, and where A ′ and B ′ are both orthogonal
to A and B.
We should stress that these results concern the least number of reflections into
which a Möbius transformation can be decomposed. Thus if a particular Möbius
transformation is expressible as the composition of four reflections, this does not
necessarily imply that it is loxodromic—one might be able to reduce the num-
ber of reflections from four to two. For example, if A and B are lines containing
angle (π/12) at 0, and A ′ and B ′ are lines containing angle (π/6) at 0, then the
Automorphisms of the Unit Disc* 199
Figure [3.37a] gives one way of seeing this. Here q, r, s may be viewed as having
fixed locations on C, while q e, er, e
s are thought of as freely movable. Provided (as
e, er, e
illustrated) that q s induce the same orientation of C as q, r, s, we know from
200 Möbius Transformations and Inversion
[3.37] [a] Möbius automorphisms of D have three real degrees of freedom. Think of q,
r, s as fixed on C, but q e, er, e
s as freely movable, their locations specified by three angles.
Then we know from (3.31) that the unique Möbius automorphism of D mapping q, r, s
e, er, e
to q s, is given by z ÞÑ e z = M(z), where [e e, er, e
z, q s] = [z, q, r, s].
[b] Finding the most general Möbius automorphism via the Symmetry Principle. If a
Möbius automorphism of D sends a to 0, then the Symmetry Principle tells us that it sends
(1/a) to ∞. It then follows
easily [see text] that the most general Möbius automorphism
of D is Mϕ a (z) = e
iϕ z−a
a z−1 , which does indeed have three real degrees of freedom.
e, er, e
(3.31) that the unique Möbius automorphism of D mapping q, r, s to q s, is given
by z ÞÑ ez = M(z), where
[e e, er, e
z, q s] = [z, q, r, s].
e, er, e
Since three real numbers are needed to specify q s—their angles, for example—
this establishes (3.50).
If two Möbius automorphisms M and N map two interior points to the same
(3.51)
image points, then M = N.
Note that Mϕ iϕ
0 (z) = −e z = e
i(π+ϕ)
z simply rotates D about its centre 0 through
angle (π + ϕ). The general Möbius automorphism Mϕ a may be interpreted as Ma
0
followed by a rotation of ϕ, and from this point of view the really interesting part
of the transformation is M0a , which we will now abbreviate to Ma . This is the same
Ma whose properties you were asked to investigate algebraically in Chapter 2,
Ex. 3.
[3.38] Geometric meaning of the most general Möbius automorphism. The most general
Möbius automorphism Mϕ a (z) (above) is the composition of a rotation with the funda-
mental mapping Ma (z) = az−a
z−1 . To discover the geometric meaning of Ma (z), we first
observe that Ma swaps a and 0. The Symmetry Principle then guides us [see text] to the
answer: Ma = ℜL ◦ IJ , the order of the reflections being immaterial. The fact that Ma
swaps a and 0 can now be recognized as a special case of the fact that Ma is involutory:
(Ma ◦ Ma ) = E, and every pair of points {z, Ma (z)} is swapped by Ma .
As was explained earlier in [3.6] on page 145, the reflection IJ in any circle J
orthogonal to C will map D to itself, the two regions into which D is divided by J
being swapped. See [3.38]. At this point the obvious thing to do is to find the circle
J such that IJ swaps a and 0. Clearly the centre q of J must lie on the line L through
a and 0, but where?
We can answer this question with the same symmetry argument that we used
earlier. Since a and (1/a) are symmetric with respect to C, their images under IJ
are symmetric with respect to IJ (C) = C. Because we want IJ (a) = 0, we deduce
that IJ (1/a) = ∞. But the point that is mapped to infinity by IJ is the centre of J, so
q = (1/a).
Of course IJ is an anticonformal mapping; to obtain a conformal Möbius auto-
morphism we must compose it with another reflection. However, we have already
successfully swapped a and 0, so this second reflection must leave these points
fixed. The obvious (and only) choice is thus reflection in L. Here, then, is our
geometric interpretation of Ma :
Ma = ℜL ◦ IJ .
Incidentally, observe [exercise] that the order of these reflections doesn’t matter:
we may also write Ma = IJ ◦ ℜL .
Clearly the fixed points ξ± are the intersection points of J and L, and so they
are symmetric with respect to C. Since the reflections occur in orthogonal circles
Automorphisms of the Unit Disc* 203
through these points, Ma is elliptic, and the multipliers associated with ξ± are both
given by m = eiπ = −1. Thus the effect of Ma on an infinitesimal neighbourhood
of the interior fixed point ξ+ is a rotation of π. The fact that Ma swaps a and 0 can
now be recognized as a special case of the fact that Ma is involutory: (Ma ◦Ma ) = E,
and every pair of points z, Ma (z) is swapped by Ma . Finally, note that we can also
express Ma as (IL ′ ◦ IJ ′ ), where J ′ and L ′ are any two orthogonal circles through
ξ+ that are orthogonal to C. All this is illustrated in [3.39], which also shows some
of the invariant circles, together with the effect of Ma on a “square”.
The deeper meaning of all this will become clear once we have encountered
Hyperbolic Geometry in Chapter 6. We will then recognize that for inhabitants of
the hyperbolic plane, Ma is nothing more than an ordinary rotation of the plane about
ξ+ through angle π! See Section 6.3.11.
While we will return to the geometry of the general Möbius automorphisms Mϕ a
in Chapter 6, we remark here that they can only be elliptic, parabolic, or hyperbolic.
This is because (by construction) they leave C invariant, while a loxodromic Möbius
transformation has no invariant circles. To be more precise, in Chapter 6 we will
use the above interpretation of Ma to show geometrically that
204 Möbius Transformations and Inversion
Any simply connected region R (other than the entire plane) may be
(3.55)
mapped one-to-one and conformally to any other such region S.
In Chapter 12 we shall discuss this in detail, but for the time being we merely wish
to point out some connections between Riemann’s result and what we have learnt
concerning automorphisms of the disc.
First note that to establish (3.55) in general, it is sufficient to establish it in the
special case that S is the unit disc D. For if FR is a one-to-one conformal mapping
from R to D, and FS is likewise a one-to-one conformal mapping of S to D, then
S ◦ FR is a one-to-one conformal mapping of R to S, as required.
F−1
If M is an arbitrary automorphism of D, then M◦FR is clearly another one-to-one
conformal mapping from R to D. In fact every such mapping must be of this form.
For if e
FR were any other such mapping, then e FR ◦ F−1
R would be some automorphism
e
M of D, in which case FR = M ◦ FR .
Thus the number of one-to-one conformal mappings from R to S is equal to the
number from R to D, which in turn is equal to the number of automorphisms of D.
As we have already said, in Chapter 7 we will show that these automorphisms are
the Möbius transformations Mϕ a , which form a 3-parameter family. Thus (3.55) in
fact implies that the one-to-one conformal mappings from R to S form a three-parameter
family.
Exercises 205
3.10 Exercises
1 In each of the figures below, show that p and p e are symmetric with respect to
the circle. The dashed lines are not strictly part of the constructions, rather they
are intended to be helpful or suggestive.
3 Let S be a sphere, and let p be a point not on S. Explain why IS (p) may be con-
structed as the second intersection point of any three spheres that pass through
p and are orthogonal to S. Explain the preservation of three-dimensional
symmetry in terms of this construction.
4 Deduce (3.23), p. 167 directly from (3.18), p. 162.
5 Consider the following two-stage mapping: first stereographically project C
onto the Riemann sphere Σ in the usual way; now stereographically project Σ
206 Möbius Transformations and Inversion
back to C, but from the south pole instead of the north pole. The net effect of this
is some complex mapping z ÞÑ f(z) of C to itself. What is f?
6 Both figures below show vertical cross sections of the Riemann sphere.
(i) In figure [a], show that the triangles p0N and N0q are similar. Deduce
(3.23).
(ii) Figure [b] is a modified copy of [3.21b]. Show that the triangles z0N and
N0 e
z are similar. Deduce (3.18).
(i) If A and B are the two circles in question, show that there exists a pair of
points ξ± that are symmetric with respect to both A and B.
(ii) Deduce that if F(z) = (z − ξ+ )/(z − ξ− ), then F(A) and F(B) are concentric
circles, as was desired.
11 This exercise yields a more intuitive proof of the result of the previous exercise.
Using different colours for each, draw two non-intersecting, non-concentric cir-
cles, A and B, then draw the line L through their centres. Label as p and q the
intersection points of B with L.
(i) Using corresponding colours, draw a fresh picture showing the images A, e
e e e of A, B, L, q under inversion in any circle centred at p. To get you started,
B, L, q
note that Le = L.
(ii) Now add to your figure by drawing the circle K, centred at q e
e, that cuts A
at right angles, and let g and h be the intersection points of K and L.
208 Möbius Transformations and Inversion
Steiner discovered, very surprisingly, that if the chain closes for one choice of
C1 , then it closes for every choice of C1 , and the resulting chain always contains
the same number of touching circles. Explain this using the result of Ex. 10.
13 (i) Let P be a sphere resting on the flat surface Q of a table. Let S1 , S2 , . . . be a
string of spheres touching one another successively and all the same size
as P. If each S-sphere touches both P and Q, show that S6 touches S1 , so
that we have a closed “necklace” of six spheres around P.
(ii) Let A, B, C be three spheres (not necessarily of equal size) all touching one
another. As in the previous part, let S1 , S2 , . . . be a string of spheres (now
of unequal size) touching one another successively, and all touching A, B,
C. Astonishingly (cf. previous exercise), S6 will always touch S1 , forming
a closed “necklace” of six spheres interlocked with A, B, C. Prove this by
first applying an inversion centred at the point of contact of A and B, then
appealing to part (i).
Exercises 209
The chain of six spheres in part (ii) is called Soddy’s Hexlet, after the ama-
teur mathematician Frederick Soddy—fellow graduate of Merton College,
Oxford!—who discovered it (without inversion!). For further information on
Soddy’s Hexlet, see Ogilvy (1990). Soddy’s full time job was chemistry—in 1921
he won the Nobel Prize for his discovery of isotopes!
14 The figure below shows four collinear points a, b, c, d, together with the (nec-
essarily coplanar) light rays from those points to an observer. Imagine that
the collinear points lie in the complex plane, and that the observer is above
the plane looking down. Show that the cross-ratio [a, b, c, d] can be expressed
purely in terms of the directions of these light rays; more precisely, show that
sin α sin γ
[a, b, c, d] = − .
sin β sin δ
Suppose the observer now does a perspective drawing on a glass “canvas plane”
C (arbitrarily positioned between himself and C). That is, for each point p in C
he draws a point p ′ where the light ray from p to his eye hits C. Use the above
result to show that although angles and distances are both distorted in his
drawing, cross-ratios of collinear points are preserved: [a ′ , b ′ , c ′ , d ′ ] = [a, b, c, d].
15 Show that in both of the figures below, Arg [z, q, r, s] = θ + ϕ. Hence deduce
(3.32), p. 156.
210 Möbius Transformations and Inversion
17 Show geometrically that if a and c lie on a circle K, and b and d are symmetric
with respect to K, then the point [a, b, c, d] lies on the unit circle. [Hints: Draw
the two circles through a, b, d and through b, c, d. Now think of [z, b, c, d] as a
Möbius transformation.]
18 The curvature κ of a circle is defined to be the reciprocal of its radius. Let M(z) =
a z+b
c z+d be normalized. Use (3.3) to show geometrically that M maps the real line
to a circle of curvature
2 d
κ = 2c Im .
c
az+b
19 Let M(z) = cz+d be normalized.
(i) Using (3.3), draw diagrams to illustrate the successive effects of these trans-
formations on a family of concentric circles. Note that the image circles are
generally not concentric.
(ii) Deduce that the image circles are concentric if and only if the original family
of circles are centred at q = −(d/c). Write down the centre of the image cir-
cles in this case. [Note that this is not the image of the centre of the original
circles: M(q) is the point at infinity!]
(iii) Hence show geometrically that the circle IM with equation |c z + d| = 1 is
mapped by M to a circle of equal size. Furthermore, show that each arc of
IM is mapped to an image arc of equal size. For this reason, IM is called
the isometric circle of M.
For applications of the isometric circle, see Ford (1929) and Katok (1992).
Exercises 211
(iii) Explain why the Symmetry Principle implies that N(1/z) = N(z).
(iv) Show by direct calculation that the formula for N in part (ii) does indeed
satisfy the equation in part (iii).
25 Let M(z) be the general Möbius automorphism of the upper half-plane.
(i) Observing that M maps the real axis into itself, use (3.30) to show that the
coefficients of M are real.
(ii) By considering Im[M(i)], deduce that the only restriction on these real
coefficients is that they have positive determinant: (ad − bc) > 0.
(iii) Explain (both algebraically and geometrically) why these Möbius transfor-
mations form a group under composition.
(iv) How many degrees of freedom does M have? Why does this make sense?
26 Reconsider (3.53), p. 201.
(i) Use (3.45), p. 192 to show that Ma is elliptic.
(ii) Use (3.44), p. 192 to show that both multipliers are given by m = −1.
(iii) Calculate the matrix product [Ma ] [Ma ], and thereby verify that Ma is
involutory.
(iv) Use (3.28), p. 172 to calculate the fixed points of Ma .
(v) Show that the result of the previous part is in accord with figure [3.38].
27 Use (3.45), p. 192 to verify (3.54), p. 204.
CHAPTER 4
4.1 Introduction
Having studied functions of complex numbers, we now turn to the calculus of such
functions.
To know the graph of an ordinary real function is to know the function com-
pletely, and so to understand curves is to understand real functions. The key insight
of differential calculus is that if we take a common or garden curve, place it under a
microscope and examine it using lenses of greater and greater magnifying power,
each little piece looks like a straight line. When produced, these infinitesimal pieces
of straight line are the tangents to the curve, and their directions describe the local
behaviour of the curve. Thinking of the curve as the graph of f(x), these directions
are in turn described by the derivative, f ′ (x).
Despite the fact that we cannot draw the graph of a complex function, in this
chapter we shall see how it is still possible to describe the local behaviour of a
complex mapping by means of a complex analogue of the ordinary derivative—the
“amplitwist”.
Visual Complex Analysis. 25th Anniversary Edition. Tristan Needham, Oxford University Press.
© Tristan Needham (2023). DOI: 10.1093/oso/9780192868916.003.0004
214 Differentiation: The Amplitwist Concept
[4.1] The tip of the iceberg: small squares map to small squares—but why? Clearly the
mapping z ÞÑ z2 preserves the orthogonality of origin-centred rays and circles. So the
small, ultimately vanishing “squares” on the left must be mapped to small, ultimately van-
ishing rectangles on the right. Very surprisingly, though, squares are ultimately mapped to
squares! As we shall see, this is a consequence of the fact that the mapping is conformal.
is that the right angle of intersection between such circles and rays in the z-plane
is preserved by the mapping, which is to say that their images in the w-plane also
meet at right angles. As illustrated in [4.1], a grid of infinitesimal squares formed
from such circles and rays must therefore be mapped to an image grid composed of
infinitesimal rectangles. However, this does not explain why these image rectangles
must again be squares.
As we will explain shortly, the fact that infinitesimal squares are preserved is
just one consequence of the fact that z ÞÑ w = z2 is conformal everywhere except
at the two critical points z = 0 and z = ∞, where angles are doubled. In par-
ticular, any pair of orthogonal curves is mapped to another pair of orthogonal
curves. In order to give another example of this, we first dismember our map-
ping into its real and imaginary parts. Writing z = x + iy and w = u + iv, we
obtain
u + iv = w = z2 = (x + iy)2 = (x2 − y2 ) + i 2xy.
Thus the new coordinates are given in terms of the old ones by
u = x2 − y2 ,
(4.1)
v = 2xy.
We now forget (temporarily!) that we are in C, and think of (4.1) as simply rep-
resenting a mapping of R2 to R2 . If we let our point (x, y) slide along any of the
rectangular hyperbolas with equation 2xy = const., then we see from (4.1) that its
A Puzzling Phenomenon 215
[4.2] Again, small squares map to small squares—but why? The pre-images of the hori-
zontal and vertical lines on the right are the two families of hyperbolas on the left, which a
computation confirms are orthogonal to each other, as illustrated. Thus the squares on the
right must have pre-images that are rectangles on the left. But, once again, surprisingly,
the pre-images of the squares are actually squares!
image (u, v) will move on a horizontal line v = const. Likewise, the preimages of
the vertical lines u = const. will be another family of rectangular hyperbolas with
equations (x2 − y2 ) = const. Since their images are orthogonal, the claimed con-
formality of z ÞÑ z2 implies that these two kinds of hyperbolas should themselves
be orthogonal.
Figure [4.2] makes it clear that they are indeed orthogonal. We may verify
this mathematically by recalling that two curves are orthogonal at a point of
intersection if the product of their slopes at that point is equal to −1. Implicitly
differentiating the equations of the hyperbolas, we find that
x2 − y2 = const. ⇒ x − yy ′ = 0 ⇒ y ′ = +(x/y),
2xy = const. ⇒ y + xy ′ = 0 ⇒ y ′ = −(y/x).
Thus the product of the slopes of the two kinds of hyperbola at a point of
intersection is −1, as was to be shown.
Clearly we could carry on in this way, analysing the effect of the mapping on
one pair of curves after another, but what is really needed is a general argument
showing that if two curves meet at some arbitrary angle ϕ, then their images under
(4.1) will also meet at angle ϕ. To obtain such an argument, we shall continue to
pretend that we are living in the less rich structure of R2 (rather than our own
home C) and investigate the local properties of a general mapping of the plane to
itself.
216 Differentiation: The Amplitwist Concept
4.3.1 Introduction
Referring to [4.3], it’s clear that to find out whether any given mapping is confor-
mal or not will require only a local investigation of what is happening very near
to the intersection point q. To make this clearer still, recognize that if we wish to
measure ϕ, or indeed even define it, we need to draw the tangents [dotted] to both
curves and then measure the angle between them. We could draw a very good
approximation to one of these tangents simply by joining q to any nearby point p
on the curve. Of course the nearer p is to q, the better will the chord qp approx-
imate the actual tangent. Since we are only concerned here with directions and
angles (rather than positions) we may dispense with the tangent itself, and instead
use the infinitesimal vector qp that points along it. Likewise, after we have per-
formed the mapping, we are not interested in the positions of the image points
Q and P themselves; rather, we want the infinitesimal connecting vector QP that
describes the direction of the new tangent at Q. We will call this infinitesimal vector
QP the image of the vector qp. However natural this terminology may seem, note
that this really is a new sense of the word “image”.
Let us now summarize our strategy. Given formulae such as (4.1), which describe
the mapping of the points to their image points, we wish to discover the induced
mapping of infinitesimal vectors emanating from a point q to their image vectors
emanating from the image point Q. In principle, we could then apply the latter
mapping to qp and to qs, yielding their images QP and QS, and hence the angle
of intersection of the image curves through Q.
[4.3] Conformal mappings. The angle ϕ between the curves passing through q on the
left is defined to be the angle between their tangents there. But as p and s merge with q,
this angle ϕ ≍ ∠pqs. Let the images of p, q, s be P, Q, S, respectively. Then to prove
conformality, we need to show that ∠PQS ≍ ϕ ≍ ∠pqs. To do so, we need to show that
the transformation that carries the initial connecting vectors on the left to their “images” on
the right (ultimately) does not alter the angle between them. This (linear) transformation
is called the Jacobian, and its coordinate description in the case of a general mapping is
derived via the next figure.
Local Description of Mappings in the Plane 217
where Bx = B/Bx etc. Likewise, we find that the vertical component is given by the
formula
dv = (Bx v) dx + (By v) dy.
Since these expressions are linear in dx and dy, it follows (assuming that not all the
partial derivatives vanish) that the infinitesimal vectors are carried to their images
by a linear transformation. The general significance of this will be discussed later,
but for the moment it means that the local effect of our mapping is completely
described by a matrix J called the Jacobian. Thus,
! ! !
dx du dx
Þ−Ñ =J ,
dy dv dy
The geometric effect of this matrix is perhaps more clearly seen if we switch to
polar coordinates. At the point z = r eiθ —or rather (r cos θ, r sin θ), since for the
moment we are still in R2 —we have
!
cos θ − sin θ
J = 2r .
sin θ cos θ
The effect of the 2r is merely to expand all the vectors by this factor. This clearly
does not affect the angle between any two of them. The remaining matrix is
probably familiar to you as producing a rotation of θ, and hence it too does
not alter the angle between vectors. Since both stages of the transformation pre-
serve angles, we have in fact verified the previous claim: the net transformation is
conformal.
with “a (direct) similarity”, except that the former refers to the transformation of
infinitesimal vectors, whereas “a similarity” has no such connotation.
[We remind the reader of the discussion in both Prefaces: “infinitesimal” is
being used here in a definite, technical sense—small and ultimately vanishing,
the relationship to other infinitesimals being expressed via Newtonian “ultimate
equalities”.]
We can illustrate the new terminology with reference to the concrete case we
have just analysed: The mapping z ÞÑ z2 is locally an amplitwist with amplification
2r and twist θ. See [4.5]. Quite generally, this figure makes it clear that if a mapping
is locally an amplitwist then it is automatically conformal—the angle ϕ between the
infinitesimal complex numbers is preserved.
Returning to [4.1] and [4.2], we now understand why infinitesimal squares were
mapped to infinitesimal squares. Indeed, an infinitesimal region of arbitrary shape
located at z will be “amplitwisted” (amplified and twisted) to a similar shape at
z2 . Note that here we are extending our terminology still further: henceforth we
will freely employ the verb “to amplitwist”, meaning to amplify and to twist an
infinitesimal geometric object.
All we really have at the moment is one simple mapping that turned out to be
locally an amplitwist. In order to appreciate how truly fundamental this amplitwist
concept is, we must return to C and begin from scratch to develop the idea of
complex differentiation.
f ′ (x) · Ñ =
−Ñ . (4.3)
If f ′ (x) > 0 (as in [4.6b]) then the image of positive dx is a positive df, but if
′
f (x) < 0 then the infinitesimal image vector df is negative and points to the left,
as illustrated in [4.7]. In this case, df can be obtained by first expanding dx by
|f ′ (x)|, then rotating it by π. If we think of f ′ (x) as a point on the real axis of C,
then arg[f ′ (x)] = 0 when f ′ (x) > 0, and arg[f ′ (x)] = π when f ′ (x) < 0. Thus,
[4.6] The real derivative re-examined. [a] The traditional picture of the derivative f ′ as the
slope of the tangent to the graph y = f(x). [b] Let us instead break apart the two copies of
the real axis, as we are forced to break apart the two copies of C in the case of a complex
function, z ÞÑ f(z). Then an infinitesimal movement dx is mapped to df = f ′ dx: in other
words, the local transformation is an amplification by f ′ .
The Complex Derivative as Amplitwist 221
[4.7] If we think of f ′ (x) as a point on the real axis of C, then arg[f ′ (x)] = 0 when
f ′ (x) > 0, and arg[f ′ (x)] = π when f ′ (x) < 0. Thus, regardless of whether f ′ (x) is
positive or negative, we see that the local effect of f on an infinitesimal vector dx at x is
to expand it by |f ′ (x)| and to rotate it by arg[f ′ (x)].
regardless of whether f ′ (x) is positive or negative, we see that the local effect of
f on an infinitesimal vector dx at x is to expand it by |f ′ (x)| and to rotate it by
arg[f ′ (x)].
With all this fresh in our minds, we now attempt to generalize the notion of
“derivative” to mappings of C.
[4.8] The local effect of a general complex mapping. The effect of a complex function on
an infinitesimal arrow is to rotate it and expand it, which is equivalent to multiplication
by a complex number = (expansion factor)ei(rotation angle) = 2 ei(3π/4) , in the illustrated
example.
222 Differentiation: The Amplitwist Concept
f ′ (z) · Õ
Ñ
= . (4.4)
In order to produce the correct effect, the length of f ′ (z) must be the magnification
factor, and the argument of f ′ (z) must be the angle of rotation. For example, at the par-
ticular point shown in [4.8] we would have f ′ (z) = 2 ei(3π/4) . In fact, in the spirit
of Chapter 1, we need not even distinguish between the local transformation and
the complex number that represents it.
To find f ′ (z) we have looked at the image of a specific arrow at z, but (unlike the
case of R) there are now infinitely many possible directions for such arrows. What
if we had looked at an arrow in a different direction from the illustrated one?
We are immediately in trouble, because a typical mapping1 will do what you
see in [4.9]. Clearly the magnification factor differs for the various arrows, and
likewise each arrow needs to be rotated a different amount to obtain its image.
While we could still use a complex number in (4.4) to describe the transformation
of the arrows, it would have to be a different number for each arrow. There would
therefore be no single complex number we could assign to this point as being the
derivative of f at z. We have arrived at an apparently gloomy impasse: a typical
mapping of C simply cannot be differentiated.
[4.9] Infinitesimal complex numbers in different directions are generally rotated and
expanded by different amounts. In general, a complex function will map an infinitesi-
mal circle to an infinitesimal ellipse, as shown. But this means that infinitesimal radii in
different directions each undergo different rotations and expansions. The mapping is not
locally an amplitwist, and is therefore not “analytic”: f ′ (z) does not exist.
1
Shortly we will justify certain details of [4.9], such as the fact that an infinitesimal circle is mapped
to an infinitesimal ellipse.
The Complex Derivative as Amplitwist 223
Analytic mappings are precisely those whose local effect is an amplitwist: all the
infinitesimal complex numbers emanating from a single point are amplified and
twisted the same amount.
In contrast to [4.9], the effect of an analytic mapping can be seen in [4.10]. For such
a mapping the derivative exists, and simply is the amplitwist, or, if you prefer, the
complex number representing the amplitwist.
At this point you might quite reasonably fear that however interesting such map-
pings might be, they would be too exotic to include any familiar or useful functions.
However, a ray of hope is held out by the humble-looking mapping z ÞÑ z2 , for we
have already established that it is locally an amplitwist, as illustrated in [4.5], and
so it now gains admittance into the select set of analytic functions. In fact, quite
amazingly, we will discover in the next chapter that virtually every function we
have met in this book is analytic! Of course we have already seen plenty of empir-
ical evidence of this in our many pictures showing small “squares” being mapped
to small “squares”.
It should perhaps be stressed that all our recent pictures have been concerned
with local properties, and hence with infinitesimal arrows and figures. For example,
it’s clear from [4.10] that any analytic mapping will send infinitesimal circles
to other infinitesimal circles; however, this does not mean that such mappings
typically send circles to circles. Figure [4.11] (which contains [4.10] at its centre)
illustrates the fact that if we start with an infinitesimal circle and then expand it,
its image will generally distort out of all semblance of circularity. Of course, an
important exception to this is provided by the Möbius transformations, for these
[4.10] Analytic = locally an amplitwist, expanding and rotating all infinitesimal complex
numbers equally. By definition, an analytic mapping is locally an amplitwist, and the
derivative f ′ (z) simply is the amplitwist.
224 Differentiation: The Amplitwist Concept
[4.11] Analytic mappings only send infinitesimal circles to circles. While an analytic
mapping must send infinitesimal circles to circles, it will generally deform an expanding
circle into a shape that becomes less and less circular. The only analytic mappings that
preserve circles of all sizes are the Möbius transformations.
precisely do preserve circles of all sizes. In fact it can be shown that the Möbius
transformations are the only2 ones with this property.
2
This was proved by Carathéodory (1937).
Some Simple Examples 225
much easier to explain. However, the student meeting this subject for the first time
should be made aware of the fact that in all other books only the word deriva-
tive is used. Also, note that the two words are synonymous only to the extent
(cf. Chapter 1) that a complex number can be identified with the similarity trans-
formation it produces when each point is multiplied by it. Thus, for example,
“to differentiate” will not mean the same as “to amplitwist”: the former refers to
the act of finding the derivative of a function, while the latter refers to the act of
“amplifying and twisting” an infinitesimal geometric figure.
(z + c) ′ = amplitwist of (z + c) = 1 ei0 = 1.
Notice how this is in complete accord with the familiar rule of real calculus, namely,
d
that dx (x + c) = 1.
z ÞÑ Az.
While the meaning is richer, this is once again formally identical to the familiar
d
result dx (Ax) = A.
z ÞÑ z2 .
Our earlier investigation revealed that at the point z = r eiθ this mapping is
locally an amplitwist with amplification 2r and twist θ, as illustrated in [4.5]. Hence,
Once again, note that this result is formally identical to the formula (x2 ) ′ = 2x
of ordinary calculus. In the next chapter we shall obtain a directly complex and
geometrical demonstration of this fact.
226 Differentiation: The Amplitwist Concept
[4.12] The complex amplitwist formula agrees with the real derivative. [a] The transla-
tion z ÞÑ z + c has amplification 1 and twist 0, so its amplitwist is 1, in accordance with
the real formula, (x + c) ′ = 1. [b] The mapping z ÞÑ Az = a eiα z has amplification a and
twist α, so its amplitwist is a eiα = A, in accordance with the real formula, (Ax) ′ = A.
z ÞÑ z.
4.6.1 Introduction
In [4.5] we saw clearly that any mapping that is locally an amplitwist is also auto-
matically conformal. In terms of complex differentiation, we can now rephrase
this by saying that all analytic functions are conformal. The question then nat-
urally arises as to whether the converse might also be true. Is every confor-
mal mapping analytic, or, in other words, is the local effect of every conformal
mapping nothing more complicated than an amplitwist? If this were the case
then the two concepts would be equivalent and we would have a new way
of recognizing, and perhaps reasoning about, analytic functions. A tempting
prospect!
To dismiss this as a possibility would only require the discovery of a single
function that is conformal and yet whose local effect is not an amplitwist. The
example of complex conjugation, illustrated in [4.13b], shows how important it is
that we take into account the fact that the mapping preserves not only the mag-
nitude of angles, but also their sense. For z ÞÑ z is not analytic, but it is also not a
counterexample to the conjecture, because it is anticonformal.
We have seen that although conjugation does possess an amplification, it fails to
be analytic because it doesn’t have a twist. Let us now consider instead a function
that does possess a twist, but which again fails to be analytic, this time by virtue of
not having an amplification. The effect of such a mapping at a particular point is
illustrated in [4.14]. The three curves on the LHS intersect at equal angles of (π/3),
and on the RHS their images do too. But the picture clearly shows that we are not
dealing with an amplitwist. Imagine that the infinitesimal tangent complex num-
bers to the curves are first twisted, but then rather than being amplified, as they
would be by an analytic function, they are expanded by different factors. Despite
this, however, the initial twist ensures that the angle between two curves is pre-
served both in magnitude and sense: the mapping is genuinely conformal at this
point.
of (ii)], and all the while the angles will remain the same as the original’s. Thus,
any infinitesimal triangle in this region is mapped to another infinitesimal similar
triangle. Since the image triangle merely has a different size and alignment on the
page, it is indeed obtained by amplitwisting the original.
We have thus established the sought-after equivalence of conformal and analytic
mappings:
A mapping is locally an amplitwist at a point p if it is conformal throughout
an infinitesimal neighbourhood of p.
For this reason, the conventional definition of f being “analytic” at p is that f ′ exist
at p and at all points in an infinitesimal neighbourhood of p.
From this result we can immediately deduce, for example, that complex inver-
sion z ÞÑ (1/z) is analytic, for we have already demonstrated geometrically that
it is conformal. By the same token, it follows more generally that all Möbius
transformations are analytic.
For no extra charge, we can obtain a further equivalence simply by concentrating
on distances rather than angles. What we have just seen is that a mapping can-
not possess a twist throughout a region without also having an amplification. In
order to investigate the converse, suppose that a mapping is only known to possess
an amplification throughout a region. Re-examine [4.15] from this point of view.
Unlike the previous case, there is no longer any a priori reason for the image ABC
to betray any features common to the original. However, as we carry out the same
shrinking process as before, the local existence of amplifications begins to reveal
itself.
As the triangle becomes very small, we may consider two of its sides, for
example ab and ac, to be infinitesimal arrows emanating from a vertex. While
we may not yet know anything of angles, we do know that these arrows both
undergo the same amplification to produce their images AB and AC. But if we
now apply this reasoning at one of the other vertices, we immediately find that
in order to be consistent, all three sides must undergo the same3 amplification.
Once again we have been able to deduce that the image triangle is similar to the
original.
However, this time all we know is that the magnitude of the angles in the infinites-
imal image triangle are the same as those in the original. If the sense of the angles
also agree, then the image is obtained by amplitwisting the original, just as before.
But if the angles are reversed, then we must flip the original triangle over as well
3
We only mean “same” in the sense that the variations in amplification are of the same infinitesimal
order as the dimensions of abc. If the amplifications were precisely the same, then extending our
argument to a whole network of closely spaced vertices, we would conclude that the amplification
was constant throughout the region.
230 Differentiation: The Amplitwist Concept
[4.16] Analytic mappings of induce conformal mappings of the Riemann sphere, for
the simple reason that stereographic projection is conformal.
actually see the disc with it. The lens L2 is even more remarkable in that it magni-
fies by 1/ϵ2 , so that even a small part of our microscopic disc now completely fills
the viewing screen4 .
Let’s switch back to L1 so that we can see the whole disc again, and watch what
happens to it when we apply the transformation z ÞÑ z2 . It disappears! At best we
might see a single dot sitting at the image of the critical point. It is in this sense that
the mapping is crushing. However, if we now attach L2 instead, we can see our
mistake: the dot isn’t a dot, in fact it’s another disc of radius ϵ2 .
For this particular mapping, L2 was sufficient to see that the disc had not been
completely crushed. However, at a critical point of another mapping, even this
might not provide sufficient magnification, and we would require a stronger lens,
say Lm , to reveal that the image of the disc isn’t just a point. The integer m measures
the degree of crushing at the critical point.
4
In terms of this analogy we could say that most of the diagrams in this chapter, indeed in the rest of
the book, show views of the image complex plane taken through L1 . For example [4.10] depicts a tiny
circle being amplitwisted to produce another circle. However, if we viewed part of this image ‘circle’
with L2 instead of L1 , then deviations from circularity would become visible. Of course the smaller
we make the preimage circle, the smaller these deviations will be.
5
The reason we define it to be (m − 1), rather than m, is that this properly reflects the multiplicity
of the root of the derivative.
Critical Points 233
only so long as we focus on the effect of the function on separate chunks of its
domain. When one studies Riemann surfaces, one tries to fit all this partial infor-
mation into a global picture of the mapping, and in achieving this the critical points
will play a crucial role. They do so by virtue of yet another aspect of the peculiar
behaviour of a mapping in the vicinity of such points, and it is to this feature that we
now turn.
In the previous chapter we discussed the possibility of critical points being
located at infinity. In particular, we considered z ÞÑ zm . On the Riemann sphere
we drew two straight lines passing through the origin, and we thereby saw that
angles at both z = 0 and z = ∞ were multiplied by m. We therefore conclude that
∞ is a critical point of zm of order (m − 1), just like the origin. Actually, except
for m = 2, we don’t yet know if zm is conformal anywhere! However, in the next
chapter we will see that it is conformal everywhere except at the two critical points
we have just discussed.
4.8.1 Introduction
To end this chapter we will try to gain a better perspective on where the analytic
functions lie within the hierarchy of mappings of the plane. A benefit of this will be
the discovery of another way (the third!) of characterizing analytic functions, this
time in terms of their real and imaginary parts.
The first thing to do is realize that the “general” mappings (x, y) ÞÑ (u, v) that
we considered earlier were not really as general as they could have been. Picture
part of the plane as being a rolled out piece of pastry on a table. A general mapping
corresponds to “doing something” to the pastry, thereby moving its points to new
locations (the images) on the table. For example, we might cut the pastry in half
and move the two pieces away from each other. This is much more general than
anything we contemplated earlier, for it does not even possess the rudimentary
quality of continuity. That is, if two points are on either side of the cut, then no
matter how close we move them together, their images will remain far apart.
Even if we do insist on continuity, the resulting mappings are still more general
than those we have considered. For example, imagine pressing down the rolling-
pin somewhere in the middle of the pastry, and, in a single roll, stretching the far
side to twice its former size. This certainly is continuous, for bringing two points
together always brings their images together. The problem now lies in the fact that
if two infinitesimal, diametrically opposed arrows emanate from a point beneath
The Cauchy–Riemann Equations 235
the starting position of the pin, then they each undergo a quite different transfor-
mation. Thus, in an obvious sense (not a subtle complex-differentiation sense) the
mapping isn’t differentiable at this point. Nevertheless, provided we stay away
from this line, the mapping is differentiable in the real sense, and hence subject to
our earlier analysis using the Jacobian matrix.
Another interesting kind of mapping arises from the commonplace operation of
folding the pastry. Suppose we fold it like a letter being placed in an envelope [two
creases]. Three different points will end up above a single point of the table, and
the mapping is thus three-to-one. However, at the creases themselves the mapping
is only one-to-one, and furthermore, differentiability also breaks down there. Nev-
ertheless, provided that we only look at the fold-free portions of the pastry, such
many-one functions are still subject to our previous analysis.
Suppose we play with the pastry in an ordinary way, rolling it (not necessarily
evenly) now in this direction, now in another, then turning it, folding it, rolling
it again, and so forth; then, provided we suitably restrict the domain, we can
still apply our old analysis. While such a mapping is indeed very general, we
hope that this discussion has revealed that (being continuous and differentiable
in the real sense) it is, in fact, already quite high up the evolutionary ladder. It
will therefore not come as such a surprise to learn that the local geometric effect
of such a mapping is remarkably simple, though naturally not as simple as an
amplitwist.
[4.18] Singular Value Decomposition. A linear transformation maps the circle C to the
ellipse E. Let d be the diameter of C that is mapped to the major axis D. Consider the
chords of C [dashed] that are perpendicular to d. Since these are all bisected by d, their
images must be a family of parallel chords of E such that D is their common bisector.
They must therefore be the family perpendicular to D. We thereby deduce the Singular
Value Decomposition: A linear transformation is a stretch in the direction of d, another
stretch perpendicular to it, and finally a twist.
This result also makes sense at the level of counting degrees of freedom. Just as
the matrix has four independent entries, so the specification of our transformation
also requires four bits of information: the direction of d, the stretch factor in this
direction, the perpendicular stretch factor, and the twist.
The Cauchy–Riemann Equations 237
6
See Stewart (1993).
7
See Needham (2021, §15.4).
8
For a different proof of this fact, see Sommerville (1958, p. 237).
238 Differentiation: The Amplitwist Concept
f ′ = Bx u + i Bx v = Bx f, (4.8)
and
f ′ = By v − i By u = −i By f. (4.9)
3
By way of example, consider z ÞÑ z . Multiplying this out we obtain a rather
haphazard looking mess:
u + iv = (x3 − 3xy2 ) + i (3x2 y − y3 ).
However, differentiating the real and imaginary parts, we obtain
Bx u = 3x2 − 3y2 = +By v,
Bx v = 6xy = −By u,
and so the Cauchy–Riemann equations are satisfied. Thus, far from being haphaz-
ard, the special forms of u and v have ensured that the mapping is analytic. Using
(4.8) we can calculate the amplitwist:
(z3 ) ′ = 3(x2 − y2 ) + i 6xy = 3z2 ,
just as in ordinary calculus. Check that (4.9) gives the same answer.
In the next chapter we will sever our umbilical cord to R2 and discover how the
above results can be better understood by directly appealing to the geometry of the
complex plane.
Exercises 239
4.9 Exercises
5 Consider f(x + iy) = (x2 + y2 ) + i (y/x). Find and sketch the curves that are
mapped by f into (a) horizontal lines, and (b) vertical lines. Notice from your
answers that f appears to be conformal. Show that it is not in two ways: (i) by
explicitly finding some curves whose angle of intersection isn’t preserved; and
(ii) by using the Cauchy–Riemann equations.
6 Continuing from the previous exercise, show that no choice of v can make
f(x + iy) = (x2 + y2 ) + iv analytic.
7 (i) If g(z) = 3 + 2i then explain geometrically why g ′ (z) ≡ 0.
(ii) Show that if the amplification of an analytic function is identically zero (i.e.,
f ′ (z) ≡ 0) on some connected region, then the function is constant there.
240 Differentiation: The Amplitwist Concept
(iii) Give a simple counterexample to show that this conclusion does not follow
if the region is instead made up of disconnected components.
8 Use pictures to explain why if f(z) is analytic on some connected region, each of
the following conditions forces it to reduce to a constant.
(i) Ref(z) = 0
(ii) |f(z)| = const.
(iii) Not only is f(z) analytic, but f(z) is too.
9 Use the Cauchy–Riemann equations to give rigorous computational proofs of
the results of the previous two exercises.
10 Instead of writing a mapping in terms of its real and imaginary parts (i.e. f =
u+iv), it is sometimes more convenient to write it in terms of length and angle:
f(z) = R eiΨ ,
where R and Ψ are functions of z. Show that the equations that characterize an
analytic f are now
Bx R = R By Ψ and By R = −R Bx Ψ.
(i) Sketch a typical S for which b < a. Now sketch its image S e under the
z
mapping z ÞÑ e .
e from your sketch, and write down the ratio
(ii) Deduce the area of S
!
e
area of S
Λ≡ .
area of S
(iii) Using the results of the previous two exercises, what limit should Λ
approach as b shrinks to nothing?
(iv) Find limbÑ0 Λ from your expression in part (ii), and check that it agrees
with your geometric answer in part (iii).
16 Consider the complex inversion mapping I(z) = (1/z). Since I is conformal,
its local effect must be an amplitwist. By considering the image of an arc of an
origin-centred circle, deduce that |(1/z) ′ | = 1/|z|2 .
17 Consider the complex inversion mapping I(z) = (1/z).
(i) If z = x + iy and I = u + iv, express u and v in terms of x and y.
(ii) Show that the Cauchy–Riemann equations are satisfied everywhere
except the origin, so that I is analytic except at this point.
(iii) Find the Jacobian matrix, and by expressing it in terms of polar coordi-
nates, find the local geometric effect of I.
(iv) Use (4.8) to show that the amplitwist is −(1/z2 ), just as in ordinary calcu-
lus, and in accord with the previous exercise. Use this to confirm the result
of part (iii).
18 Recall Ex. 19, p. 210, where you showed that a general Möbius transformation
az + b
M(z) = ,
cz + d
maps concentric circles to concentric circles if and only if the original family
(call it F) is centred at q = −(d/c). Let ρ = |z − q| be the distance from q to z,
so that the members of F are ρ = const.
(i) By considering orthogonal connecting vectors from one member of F to
an infinitesimally larger member of F, deduce that the amplification of M
is constant on each circle of F. Deduce that |M ′ | must be a function of ρ
alone.
(ii) By considering the image of an infinitesimal shape that starts far from q
and then travels to a point very close to q, deduce that at some point in
the journey the image and preimage are congruent.
(iii) Combine the above results to deduce that there is a special member IM
of F such that infinitesimal shapes on IM are mapped to congruent image
242 Differentiation: The Amplitwist Concept
shapes on the image circle M(IM ). Recall that IM is called the isometric
circle of M.
(iv) Use the previous part to explain why M(IM ) has the same radius as IM .
(v) Explain why IM−1 = M(IM ).
(vi) Suppose that M is normalized. Using the idea in Ex. 16, show that the
amplification of M is
1
|M ′ (z)| = 2 2 .
|c| ρ
19 Consider the mapping f(z) = z4 , illustrated below. On the left is a particle p
travelling upwards along a segment of the line x = 1, while on the right is the
image path traced by f(p).
(i) Copy this diagram, and by considering the length and angle of p as it
continues its upward journey, sketch the continuation of the image path.
(ii) Show that A = i sec4 (π/8).
(iii) Find and mark on your picture the two positions (call them b1 and b2 ) of
p that map to the self-intersection point B of the image path.
(iv) Assuming the result f ′ (z) = 4z3 , find the twist at b1 and also at b2 .
(v) Using the previous part, show that (as indicated at B) the image path cuts
itself at right angles.
(iv) Let s represent the length of the segment of the lemniscate connecting the
origin to the point z. Deduce from the previous two parts that
Zr
dr
s= p ,
0 1 − (r4 /4)
hence the name, lemniscatic integral.
5.1.1 Introduction
In the previous chapter we began to investigate the remarkable nature of analytic
functions in C by studying mappings in the less structured realm of R2 . In particu-
lar, the Jacobian provided us with a painless way of deriving the Cauchy–Riemann
characterization of analytic functions, and also of computing their amplitwists.
However, this approach was rather indirect.
In this chapter we will instead study differentiation directly in the complex
plane, primarily through the use of ultimate equality—the powerful geometrical
tool that Newton unleashed in his great Principia of 1687; see both the original and
the new Preface. However, in the original edition these arguments were expressed
mainly in the language of infinitesimals, and we have left this mode of expression
largely intact. But in the captions, all of which are new to this edition, we have felt
free to employ the ≍ –notation.
Our first application of this approach will be the rederivation of the Cauchy–
Riemann (henceforth “CR”) equations, and the discovery of new forms that they
can take on.
Visual Complex Analysis. 25th Anniversary Edition. Tristan Needham, Oxford University Press.
© Tristan Needham (2023). DOI: 10.1093/oso/9780192868916.003.0005
246 Further Geometry of Differentiation
Zoom in on an individual square and its image, as depicted in the bottom half of
[5.1]. Suppose, as drawn, that the initial square has side ϵ. If we start at z and then
move a distance ϵ in the x-direction, the image will move along a complex number
given by
(change in x)·(rate of change with x of the image f) = ϵ Bx f.
Similarly, if the point moves along the vertical edge by going ϵ in the y-direction,
then its image will move along ϵ By f. Now since these two image vectors span a
square they must be related by a simple rotation of π/2, that is by multiplication
with i. After cancelling ϵ, we thus obtain
iBx f = By f, (5.1)
et voilà! That this is indeed a compact form of the CR equations may be seen by
inserting f = u + iv:
iBx (u + iv) = By (u + iv),
and then equating real and imaginary parts to yield
Bx u = By v and Bx v = −By u, (5.2)
just as before.
Cauchy–Riemann Revealed 247
To obtain the amplitwist itself, we recall that each infinitesimal arrow is taken to
its image by multiplication with f ′ . Now, since we know what the images are for
the two sides of the square, we easily deduce
ϵ Þ−Ñ ϵ f ′ = ϵ Bx f
⇒ f′ = Bx f
and
iϵ Þ−Ñ iϵ f ′ = ϵ By f
⇒ f′ = −i By f.
These are the same formulae (4.8) and (4.9) that we previously obtained, but now
we understand their geometrical meaning.
radial edge. If, on the other hand, we increase θ by dθ, then the point will move
in the perpendicular direction given by ieiθ . As dθ tends to zero, this edge is ulti-
mately equal to an infinitesimal arc of circle of length r dθ; the complex number
describing it will therefore be i eiθ r dθ = i z dθ. It’s also clear from our picture that
initially square ⇐⇒ dr = r dθ. (5.3)
Now look at the image. Just as before, if we increase r by dr then the image will
move along dr · Br f; likewise, changing θ by dθ will move the image along dθ · Bθ f.
If the mapping is analytic then these again span a square, and so the latter must be
i times the former:
dθ · Bθ f = idr · Br f.
Substituting (5.3) into this, and cancelling dθ, we obtain
Bθ f = ir Br f (5.4)
as the new compact form of CR. By inserting f = u + iv, the reader may verify that
(5.4) is equivalent to the following pair of Polar–Cart. equations:
Bθ v = +r Br u (5.5)
Bθ u = −r Br v. (5.6)
By examining the amplitwist that carries each arrow to its image we can also obtain
two expressions for the derivative:
eiθ dr Þ−Ñ eiθ dr · f ′ = dr · Br f
′
=⇒ f = e−iθ Br f (5.7)
and
i z dθ Þ−Ñ i z dθ · f ′ = dθ · Bθ f
=⇒ f ′ = −(i/z) Bθ f. (5.8)
As a simple example let’s take z3 = r3 e3iθ . From (5.7) we obtain
(z3 ) ′ = e−iθ 3r2 e3iθ = 3r2 e2iθ = 3z2 ,
while from (5.8) we obtain
(z3 ) ′ = −(i/z) r3 3ie3iθ = −(i/z) 3i z3 = 3z2 .
In obtaining the same answer from both these expressions we have also verified
that z3 actually was analytic in the first place.
Of the four possible ways of writing CR, only one now remains to be found,
namely the Polar-Polar form. We leave it to the reader to verify that if we write
f = R eiΨ (cf. Ex. 10, p. 240) then CR takes the form
Bθ R = −rR Br Ψ and R Bθ Ψ = r Br R.
An Intimation of Rigidity 249
[5.3] Cauchy’s Formula. The structure of an analytic function is so strong that given even
the meagre knowledge of how it affects a closed curve (just the points on the curve, mind
you), we can predict precisely what happens to each point inside! The precise recipe is
given by Cauchy’s Formula, (9.1), page 485.
250 Further Geometry of Differentiation
they are thus horizontal lines. In fact, if we swing the ray around counter-clockwise,
we can even tell whether its image line will move up or down. Look at [5.5], which
depicts the fate of a infinitesimal square bounded by two circles and two rays. We
know that the infinitesimal radial arrow connecting the two circles must map to
a connecting arrow between the lines going from left to right. But since the square
is to be amplitwisted, its image must be positioned as shown. Thus we find that a
positive rotation of the ray will translate the image line upwards.
We have made some good progress, but that we cannot yet have fully cap-
tured the consequences of analyticity can be seen from Ex. 5, p. 239. Despite not
being analytic, the mapping (x + iy) ÞÑ (x2 + y2 ) + i(y/x) was there shown to
possess all the above desiderata. Indeed it would be easy to write down an infini-
tude of nonanalytic functions that would be consistent with the known facts. Very
remarkably, and in stark contrast to this, if we restrict ourselves to analytic map-
pings, then we will prove that there is only one! To show this we must turn to the
CR equations.
In [5.4] we are mapping natural polar objects to natural Cartesian objects, so it’s
clear that we should employ the Polar–Cart. form, namely (5.5) and (5.6). In order
to put them to use, we must first translate [5.4] into ‘Equationspeak’. We could
describe the figure by saying that rotating the point only moves the image up and
down, not side to side; in other words, varying θ produces no change in u: Bθ u = 0.
It follows from (5.6) that Br v = 0. This says that moving the point radially outwards
does not affect the height of the image, and thus that rays are mapped to horizontal
[5.4] There is only one analytic mapping that does this. Suppose that a mapping sends
origin-centred circles to vertical lines, and the larger the circle, the further to the right
is the image line, but with no restriction on how the lines are spaced. It is easy to write
down an infinitude of different non-analytic mappings that do this, but, very remarkably,
if we restrict ourselves to analytic mappings, then we can prove that there is only one!
An Intimation of Rigidity 251
[5.5] Geometric proof of the claim made in the previous figure. Let us now restrict our-
selves to analytic mappings possessing the property in the previous figure. Since rays are
orthogonal to origin-centred circles, their images under a conformal mapping must be
orthogonal to vertical lines—so the images of rays are horizontal lines. Furthermore, we
know that the infinitesimal outward radial arrow connecting the two illustrated circles
must map to a connecting arrow between vertical lines going from left to right. Since the
square is amplitwisted to its image, we deduce that as the ray rotates counterclockwise,
its horizontal image line moves upwards.
lines. This is old news to us seasoned geometers, but fortunately we have another
equation left:
and hence
U + iV = A (ln r + iθ) + B,
252 Further Geometry of Differentiation
where B = const. But we recognize this special combination as none other than
the complex logarithm! Thus
f(z) = A log z + B. (5.10)
Suppose, more generally, that an analytic function g(z) is known to send circles
with centre c to parallel lines making a fixed angle ϕ with the imaginary axis. That
there is no fundamental difference between this and the previous case may be seen
by considering
z ÞÑ e−iϕ g(z + c);
for you may convince yourself that this possesses property [5.4], and hence it too
must equal (5.10). The rigidity of analytic functions has thus led to the rather
remarkable conclusion that the complex logarithm is uniquely defined (up to
constants) as the conformal mapping sending concentric circles to parallel lines.
A fringe benefit of the previous section was the discovery that log(z) actually is
analytic. Since this multifunction finds its simplest representation in Polar–Cart.
form, namely
log z = ln r + i(θ + 2mπ),
we can easily find its derivative using (5.7) or (5.8). For purposes of illustration, we
will now use them both:
(log z) ′ = e−iθ Br log z = e−iθ (1/r) = 1/z,
and
(log z) ′ = −(i/z) Bθ log z = −(i/z) i = 1/z. (5.11)
You notice, of course, how this is formally identical to the case of the ordinary, real
logarithm.
You may be wondering how our previous discussion of the branches of this mul-
tifunction affects all this. For example, it’s interesting how m (which labels the
different branches) does not appear in the result (5.11). The basic philosophy of
this book is that while it often takes more imagination and effort to find a picture
than to do a calculation, the picture will always reward you by bringing you nearer
to the Truth. In this spirit, we now find a visual explanation of (5.11) that will also
make it clear that the answer does not depend on m.
Equations (5.7) and (5.8) were derived by examining the infinitesimal geometry
of a general analytic mapping. Why not then apply this idea to the geometry of a
specific mapping, and thereby evaluate its amplitwist directly?
Visual Differentiation of log(z) 253
[5.6] Geometric demonstration that [log(z)] ′ = (1/z). Rotating z by the small, ultimately
vanishing angle δ moves it along the white complex number, which ultimately has length
rδ. All of the infinitely many images under log(z) move upwards together, along iδ. Thus,
for all branches, the amplification = (1/r), the twist = −θ, and therefore the amplitwist =
(1/r)e−iθ = (1/z).
Consider [5.6], which shows a typical point z and a few of its infinitely many
images under log. In order to find the amplitwist we need only find the image of
a single arrow emanating from z. The easiest one to find is shown in [5.6], namely
an arrow perpendicular to z. Notice how if z makes an angle θ with the horizontal,
then the perpendicular vector will make an angle θ with the vertical. Also, if it
subtends an infinitesimal angle δ at the origin, then—because it is like a small arc
of a circle—its length will be rδ. Now look at the images of z. Since we have purely
rotated z, its images will all move vertically up through a distance equal to the
angle of rotation δ. To make it easier to see what amplitwist carries the arrow at z
into its image, we have drawn copies of the original arrow at each image point. It
is now evident from the picture that
amplification = 1/r
twist = −θ
=⇒ amplitwist = (1/r) e−iθ = 1/z.
Although all the image vectors emanate from different points in the different
branches, they are all identical as vectors, and so it is clear that the amplitwist does
not depend on which branch we look at.
254 Further Geometry of Differentiation
We already know how to differentiate z2 and also log z, so how would you use this
knowledge to find, for example, the derivative of log(z2 log z)? Your immediate
reaction (chain and product rules) is quite correct, and in this section we merely
verify that all the familiar rules of real differentiation carry over into the complex
realm without any changes, at least in appearance.
5.4.1 Composition
The composite function (g ◦ f)(z) = g[f(z)] of course just means “do f, then do
g”. If both f and g are analytic then each of these two steps conserves angles, and
therefore the composite mapping does too. We deduce that g[f(z)] is analytic, and
we now show that the net amplitwist it produces is correctly given by the chain
rule.
Let f ′ (z) = A eiα and g ′ (w) = B eiβ , where w = f(z). Consider [5.7]. An infinites-
imal arrow at z is amplitwisted by f to produce an image at w; then this, in its turn,
is amplitwisted by g to produce the final image at g(w). It is clear from the picture
that
net amplification = AB
net twist = α + β
=⇒ net amplitwist = AB ei(α+β) ,
[5.7] Geometry of the Chain Rule. The local effect of the composition of two analytic
mappings is an amplitwist that is the composition of their separate amplitwists, so we
obtain the familiar chain rule.
Rules of Differentiation 255
As an example of this we may put g(z) = kz. In the last chapter we showed that
g ′ (z) = k, and so we now conclude from (5.12) that
[k f(z)] ′ = k f ′ (z).
[5.8] The amplitwist of the inverse of an analytic mapping has the opposite twist and
the reciprocal amplification, and so we obtain the familiar rule [f−1 (w)] ′ = 1/f ′ (z).
256 Further Geometry of Differentiation
[5.9] If f and g are analytic, then by considering the image of the small, ultimately van-
ishing complex number ξ under first their sum and then their product, we deduce that
(f + g) ′ = f ′ + g ′ , and (fg) ′ = f ′ g + f g ′ , just as in elementary real calculus.
Polynomials, Power Series, and Rational Functions 257
5.5.1 Polynomials
We can look at the rules of the previous section from a slightly different point of
view. Take rule (5.16), for example. In a way, what is on the RHS is less important
than the fact that there is a RHS. By this we mean that we have here a recipe for
creating new analytic functions: ‘given two such functions, form their product’.
Likewise, each of our other rules can be thought of as a means of producing new
analytic functions from old. The analytic functions are indeed the aristocrats of the
complex plane, but provided they only mate with their own kind, and only in ways
sanctioned by the rules (which allow many forms of incest!), their offspring will
also be aristocrats. For example, suppose we start with only the mapping z ÞÑ z,
which is known to be analytic. Our rules now quickly generate z2 , z3 , . . ., and thence
any polynomial.
Consider a typical polynomial of degree n:
Sn (z) = a0 + a1 z + a2 z2 + · · · + an zn .
We have just seen that this is analytic, and thus it maps an infinitesimal disc at p to
another at Sn (p). Furthermore, the amplitwist that transforms the former into the
latter is, according to (5.15),
Sn′ (z) = (a0 ) ′ + (a1 z) ′ + (a2 z2 ) ′ + · · · + (an zn ) ′ .
We already know how to differentiate the first four terms, and in the next section
we will confirm that in general (zm ) ′ = m zm−1 , as you no doubt anticipated. Thus
Sn′ (z) = a1 + 2a2 z + 3a3 z2 + · · · + nan zn−1 . (5.17)
1
For simplicity’s sake we shall use a power series centred at the origin. However, as we pointed out
in Chapter 2, this does not involve any loss of generality.
258 Further Geometry of Differentiation
[5.10] Power series are analytic. On the left, we see the images of a small, ultimately
vanishing disc D centred at p under the higher and higher degree polynomial approxi-
mations Sn (z) of a power series, S(z). On the right, we see these same image discs, now
translated to have their common centre at S(p). All these polynomial approximations are
analytic, so three equally spaced points on the rim of D will be amplitwisted to their
images on the rims of the image discs, and so it becomes clear that S(z) is indeed ana-
lytic, and that its amplitwist is the limit of the amplitwists of Sn as n goes to infinity. [For
ease of visualization, here the amplification and twist are both depicted as monotonically
increasing with n, but this will not be the case in general.]
amplitwists these vectors to three equally spaced image vectors. The figure shows
the gradual evolution2 of these images towards their final state (given by S) as we
successively apply S10 , S100 , etc. The amplitwist that carries the arrows of D into
these images therefore undergoes a corresponding evolution towards a final value.
The amplitwist S ′ that carries the original vectors of D to their ultimate images is
thus mimicked with arbitrarily high accuracy by Sn′ , as n increases. Therefore
S ′ (z) = a1 + 2a2 z + 3a3 z2 + 4a4 z3 + · · · . (5.19)
We have reached a very important conclusion. Any power series is analytic within
its disc of convergence, and its derivative is obtained simply by differentiating
the series term by term. Since the result of this process (5.19) is yet another con-
vergent power series with the same radius of convergence, there is nothing to
stop us differentiating again. Continuing in this manner, we discover that a power
series is infinitely differentiable within its disc of convergence. The reason this is
so important is that we will be able to show later that every analytic function can
be represented locally as a power series, and thus analytic functions are infinitely
differentiable.
This result is in sharp contrast to the case of real functions. For example, the
mileage displayed on the dash of your car is a differentiable function of the time
displayed on the clock. In fact the derivative is itself displayed on the speedometer.
However, in the instant that you hit the brakes, the second derivative (acceleration)
does not exist. More generally, consider the real function that vanishes for nega-
tive x, and that equals xm for non-negative x. This is differentiable (m − 1) times
everywhere, but not m times at the origin. Our complex aristocrats will be shown
to be quite incapable of stooping to this sort of behaviour.
2
For ease of visualization, we have taken both the amplification and the twist to be steadily increas-
ing with n. In general they could exhibit damped oscillations as they settled down to their final
values.
260 Further Geometry of Differentiation
We saw in the last section that z2 , z3 , z4 , . . . were all analytic. Composing with com-
plex inversion, it follows that z−2 , z−3 , z−4 , . . . are too. Since the inverse functions
(in the sense of [5.8]) are branches of the multifunctions z±1/2 , z±1/3 , . . . discussed
in Chapter 2, it follows that these too are analytic. Composing zp with z1/q (p, q
integers), it follows that any rational power is analytic. Furthermore, since the
geometric effect of any real power can be reproduced with arbitrary accuracy by
rational powers, it follows that these real powers are also analytic.
The calculation of the derivative of a real power za is similar to the example z3
given on p. 248. We find that
(za ) ′ = a za−1 , (5.20)
just as in ordinary calculus. In fact the real formula (xa ) ′ = a xa−1 can be thought
of as the specialization of (5.20) that results when both z and the infinitesimal arrow
emanating from it are taken to be on the real axis (cf. [4.7], p. 221).
Just as in the case of the complex logarithm, we do not rest at the result (5.20) of a
calculation, but rather we stalk the thing to its geometric lair. Since the amplitwist
is the same for all arrows, we need only find the image of a single arrow in the
direction of our choice. As a first (ill-fated) attempt, consider [5.11], in which we
have chosen an arrow parallel to z. To facilitate comparison, we have drawn a copy
of the initial arrow at the image point. You can see from the picture that
twist = (a − 1)θ BUT amplification = ???????
We are thus half-thwarted, for we cannot see how long the image arrow is. In fact
to figure this out would require precisely the same calculation (general Binomial
Theorem) as is needed in the real case. Oh well, “If at first you don’t succeed, …”
“Try, try an arrow perpendicular to z !” From [5.12], we see that this arrow orig-
inally makes an angle θ with the vertical, and so after magnifying the angle of
z by a, it will make an angle aθ with the vertical. Once again we see that the
twist = (a − 1)θ. However, this time we can see the amplification, simply by rec-
ognizing that each arrow is an infinitesimal arc of a circle. The angle subtended by
the arc has been magnified by a, while the radius of the circle has been magnified
by ra−1 . The net amplification of the arc is therefore a ra−1 . Thus
amplification = a ra−1
twist = (a − 1)θ
=⇒ amplitwist = a ra−1 ei(a−1)θ = a za−1 .
Visual Differentiation of the Power Function 261
To the best of our knowledge there is no3 direct, intuitive way of understanding the
real result (xa ) ′ = a xa−1 . It is therefore particularly pleasing that with the greater
generality of the complex result (5.21) comes the richer geometry of [5.12] needed
to see its truth.
We have already seen that (ez ) ′ = ez by calculation, and we will now explain it
geometrically. In [5.13] we have written a typical point z = x + iθ to make it easier
to remember that w = ez = ex eiθ has angle θ. Moving z vertically up through a
distance δ will rotate the image through an angle δ. Being an infinitesimal arc of
circle of radius ex , the image vector has length ex δ; its direction is θ to the vertical.
As usual, we have copied the original arrow at the image so that we may more
clearly see the amplitwist:
amplification = ex
twist = θ
=⇒ amplitwist = ex eiθ = ez .
Actually, we have been a little hasty. We haven’t really shown yet (at least not
geometrically) that ez is analytic: we don’t know if all arrows undergo an equal
amplitwist. Figure [5.13] tells us that if it’s analytic, then (ez ) ′ = ez . To establish
analyticity we need only see that one other arrow is affected in the same way.
In [5.14] we move z an infinitesimal distance δ in the x-direction, thereby mov-
ing the image radially outwards. Now, from ordinary calculus, the amplification
produced by ex along the real axis is ex (cf. [4.6], p. 220), so the length of this image
3
In special cases there are ways. For example, consider a cube of side x. It is easy to visualize that
if we increase the separation of one of the three pairs of faces by δ, we add a layer of volume x2 δ. The
result (x3 ) ′ = 3x2 follows.
Visual Differentiation of exp(z) 263
vector is ex δ. It is now clear that this new arrow in [5.14] has indeed undergone
precisely the same amplification and twist as that in [5.13], thus establishing the
analyticity of ez .
E ′ = E. (5.22)
Of course this doesn’t quite pin it down since k ex also obeys (5.22); however, if
we insist that the real solution of (5.22) also satisfy E(0) = 1, then no ambiguity
remains.
The object of this section is to show that the complex exponential function can
be characterized in exactly the same way. If a complex-analytic function E(z) is to
generalize ex then it must satisfy (5.22) on the real axis. We will now show geomet-
rically that (5.22) uniquely propagates ex off the real axis into the plane to produce
the familiar complex exponential mapping. The plan will be essentially to reverse
the flow of logic associated with [5.13] and [5.14].
A typical point z is being mapped to an unknown image w, where w = E(z)
is subject to (5.22). Decoding this equation, we find that it says that vectors ema-
nating from z undergo an amplitwist equal to the image point w. From this alone
we will figure out where w must be! In what follows, try to free your mind from
assumptions based on your previous knowledge of ez .
Consider what happens to the little (ultimately vanishing) square of side ϵ
shown in [5.15]. Because it’s twisted by the angle of w, its horizontal edge becomes
parallel to w, while its vertical edge becomes orthogonal to w. Thus horizontal
movement of z results in radial movement of the image, while vertical movement
results in rotation of the image. The question that now remains is exactly how swift
these radial and rotational motions are. Having used the twist, we now turn to the
amplification.
If z moves at unit speed in the x-direction, then since the amplification is r, the
image moves radially with speed equal to its distance from the origin. But this
Geometric Solution of E′ = E 265
is just the familiar property of the ordinary exponential function. Thus E maps
horizontal lines exponentially onto rays. If we now insist that E(0) = 1, then the
real axis maps to the real axis, and we thereby recover the ordinary exponential
function. We also know that translating a horizontal line upwards will rotate its
image ray counter-clockwise, but we don’t yet know how fast. In [5.15] dθ is the
infinitesimal rotation produced by moving z through a distance ϵ along the vertical
edge of the square. But since the amplification is r, we know that the image of this
edge has length rϵ, and consequently dθ = ϵ. In other words,
An infinitesimal vertical translation produces a numerically equal rotation. (5.23)
We can now completely describe the mapping produced by E(z). Imagine watch-
ing the image as we move from the origin to a typical point z = x + iθ in a
two-legged journey: first along the real axis to x, then straight up to z. See [5.16]. As
we move to x, the image moves along the real axis from 1 to ex . Repeated applica-
tion of (5.23) then tells us that moving up a distance θ will rotate the image through
an angle θ. For example, we find that E(z) wraps the imaginary axis around the unit
circle in such a way that
E(iθ) = cos θ + i sin θ.
This is our old friend, the celebrated Euler Formula. It also follows directly from
this geometry that the mapping has the property
E(a + b) = E(a) · E(b).
It is now entirely logical to define “ez ” to be E(z), and our work is done.
266 Further Geometry of Differentiation
[5.16] Geometry of the exponential. If we insist that E(0) = 1, then combining the two
facts we deduced in the previous figure, we see that we have arrived at the familiar
geometry of z ÞÑ ez .
As we indicated at the start of this section, there is in fact an even more com-
pelling explanation than the above. We have just used a very natural differential
equation to propagate ex off the real axis; however, it will turn out that even
this equation is superfluous. The rigidity of analytic functions is so great that
merely knowing the values of ex on the real axis uniquely determines its “analytic
continuation” into the complex realm.
5.9.1 Introduction
Earlier we alluded to the remarkable fact that analytic functions are infinitely dif-
ferentiable. In other words, if f is analytic then f ′′ exists. In this section we seek to
shed geometric light on the meaning and existence of this second derivative f ′′ . We
shall do so by answering the following question:
If an analytic mapping f acts on a curve K of known curvature κ at p,
then what is the curvature e e at f(p)?
κ of the image curve K
In the next section we shall see that the solution to this problem provides a novel
insight into (of all things!) the elliptical orbits of the planets round the sun at one
focus.
At the risk of ruining the suspense, here is the answer to our question:
" # !
1 f ′′ (p) b
ξ
e
κ= ′ Im +κ , (5.24)
|f (p)| f ′ (p)
An Application of Higher Derivatives: Curvature* 267
where bξ denotes the unit complex number tangent to the original curve at p. Before
explaining this result, let us simply test it on an example.
On the left of [5.17] we have drawn three line-segments, and on the right their
images under f(z) = ez . The segments are distinguished by the value of the angle
ϕ that each makes with the horizontal: ϕ = 0 for the dotted one; ϕ = (π/2) for
the dashed one; and the solid one represents a general value of ϕ. Now look at the
curvature of their images: e κ = 0 for the dotted one; e κ = e−a for the dashed one;
and on the solid image, e κ starts out large and then dies away as we spiral out from
the origin.
In order to compare these empirical observations with our formula, write the
unit tangent as b
ξ = eiϕ and note that if f(z) = ez then f ′′ = f ′ = ez . With z = x+iy,
formula (5.24) therefore reduces to
e
κ = e−x (sin ϕ + κ).
Using the fact that κ = 0 for our line-segments, and that ϕ is constant on each, you
may now easily check the accord between this formula and figure [5.17].
[5.18] Geometric derivation of the transformation law for curvature. The analytic map-
ping f transforms the curve K into the curve K. e Curvature is the rate of rotation of the
tangent, so κ ≍ (ϵ/|ξ|) at p is transformed into eκ ≍ (eϵ/|e
ξ|) ≍ (eϵ/|f ′ (p)ξ|) at p
e = f(p).
If the twists at p and at q were the same, then the turning angle ϵ would not change,
but in fact (as we see on the right) the amplitwists at p and q differ by χ ≍ f ′′ (p)ξ, so
there is an extra twist σ ≍ (angle subtended by χ). To geometrically determine this extra
′
twist σ, divide the figure onhthe right
i by f h(p) to obtain
i the figure at the bottom. We now
f ′′ (p)ξ
see that σ ≍ Im(ν) = Im f ′ χ(p) ≍ Im f ′ (p) , thereby completing the proof of the
curvature transformation law, (5.24).
An Application of Higher Derivatives: Curvature* 269
image curve K e under the mapping f; note that its sense is determined by that of K.
It is the curvature of K e at p
e = f(p) that (5.24) purports to describe.
As illustrated, ξ is a small (ultimately infinitesimal) complex number tangent to
K at p. With centre at p we have drawn a circle through the tip of ξ cutting K in
q. At q we have drawn another small (ultimately infinitesimal) tangent complex
number ζ, and we have marked the angle ϵ of rotation from ξ to ζ. Recall that the
curvature κ at p is, by definition, the rate of rotation of the tangent with respect to
distance along K. Since for infinitesimal ξ the arc pq equals |ξ|, the curvature at p
is therefore
ϵ
κ= . (5.25)
|ξ|
Likewise, at the image points p e and qe on the image curve K e we have drawn the
image complex numbers e e the rotation from e
ξ and ζ, ξ to ζe being e
ϵ. Thus the image
curvature is
e
ϵ
e
κ= . (5.26)
|e
ξ|
Our problem therefore reduces to finding e ϵ and |e
ξ|.
e
Since |ξ| is the length of the amplitwisted image of ξ,
|e
ξ| = (amplification) · |ξ| = |f ′ (p)| · |ξ|. (5.27)
The more interesting and difficult part of the problem is to find e
ϵ.
If ξ and ζ both underwent precisely the same twist, then the turning angle e
ϵ for
the images would equal the original turning angle ϵ. However, the twist at q will
differ very slightly, say by σ, from that at p. Thus
e
ϵ = ϵ + (extra twist) = ϵ + σ. (5.28)
′′
This is how f enters the picture, for it describes how the amplitwist varies.
The function f ′ is a perfectly respectable mapping in its own right, and it may be
drawn like any other. The right-hand side of [5.18] is precisely such a picture. Each
point z is mapped to the complex number that amplitwists infinitesimal complex
numbers emanating from z. In particular, we have drawn the images f ′ (p) and
f ′ (q) of p and q. The statement about infinite differentiability can now be recast in
a more blatantly astonishing form: if f is locally an amplitwist, then f ′ automatically is
too. We have indicated this in the picture by showing the disc at p being mapped
by f ′ to another disc at f ′ (p). This startling fact will now yield to us the value of σ.
The amplitwist that carries the disc at p to the disc at f ′ (p) is f ′′ (p). In particular,
ξ is amplitwisted to
χ = f ′′ (p) ξ.
But looking at the triangle on the right, the sides of which are the known quantities
f ′ (p) and χ, we see that the angle at the origin is precisely the extra twist σ that we
seek.
270 Further Geometry of Differentiation
It is easier to obtain an expression for this angle if we first rotate the triangle
to the real axis. This rotation is achieved quite naturally (see the bottom figure) by
dividing by f ′ (p); the sides of the triangle now become 1 and ν = [χ/f ′ (p)]. Because
σ equals the (ultimately) vertical arc through 1, the figure tells us that
′′
χ f (p) ξ
σ = arc = Im(ν) = Im ′ = Im .
f (p) f ′ (p)
Thus, from (5.26), (5.27), and (5.28), and taking evaluation at p as understood, we
obtain
h ′′ i
Im f ′ξ + ϵ
e
κ= f .
|f ′ | |ξ|
Finally, using (5.25) and noting that b ξ = (ξ/|ξ|) is the unit tangent at p, we do
indeed obtain formula (5.24).
Finally, we note that, alternatively, a very short (but unilluminating) computa-
tional proof can be obtained [exercise] by differentiation of the equation,
Twist = Im log f ′ .
Now consider the fate of all the curves that pass through p in the direction b ξ . The
general formula says that f will not only scale their curvatures by (1/|f ′ |) (as previ-
ously explained), but it will also increase their curvatures by the fixed amount k(b ξ).
In this sense, the first term corresponds to an intrinsic property of the mapping f.
4
The circle that touches the curve at the point in question, and whose curvature κ = (1/radius)
agrees with that of the curve at that point.
An Application of Higher Derivatives: Curvature* 271
i f ′′
K≡ . (5.29)
f ′ |f ′ |
We propose to call this complex function K (which does not appear to have been
investigated previously) the complex curvature5 of f.
To see that the complex curvature is indeed a natural quantity, picture K(p) as a
vector emanating from p. We will show that
The projection of K(p) onto a line through p is the curvature of the image
(5.30)
of that line at f(p).
See [5.19], in which K has also been drawn at two additional points. Note how
the increasing length of the projection of K onto the line corresponds to increasing
curvature along the image.
[5.19] Geometric interpretation of the complex curvature. The projection of the complex
curvature K(p) onto a line through p is the curvature of the image of that line at f(p).
Here, the projection grows as we move along the line, so the image curve bends more
tightly.
5
NOTE added in this 25th Anniversary Edition: I am delighted that my discovery of K and its math-
ematical properties has since found several applications (particularly in physics) that go beyond the
two applications that I originally discovered and published here in VCA, namely, to dual central force
fields [this chapter] and to the geometry of harmonic functions [Chapter 12]. For example, see Vitelli
and Nelson (2006) and Mughal and Weaire (2009).
272 Further Geometry of Differentiation
To prove (5.30), recall how the scalar product in R2 can be expressed in terms of
complex multiplication:
" #
h i h i ′′ b
K ξ·
b = Re K b ξ = Im i K b
f ξ
ξ = Im ′ ′ = k( b
f |f |
ξ ),
as was to be shown. This result yields a neater and more intelligible form of (5.24):
e
κ=K ξ ·
b + κ .
|f ′ |
(5.31)
K = κ1 + iκ2 .
[5.20] Geometric interpretation of the real and imaginary parts of the complex curva-
ture. The real and imaginary parts of the complex curvature K(p) are the curvatures of
the images of the real and imaginary line segments through p.
An Application of Higher Derivatives: Curvature* 273
e is just the angle σ on the RHS of [5.18]. This rotation is clearly greatest when
of Q
χ is perpendicular to f ′ (p), pointing counterclockwise along the circle |f ′ | = const.
This occurs when ξ is in the direction of K, for then
χ ∝ f ′′ K ∝ i/f ′ ∝ if ′ .
b.
R=K ξ ·
This achieves its maximum value Rmax = |K| when Q moves in the direction of K.
e Let E denote the rate of increase of the size6
Similarly, consider the expansion of Q.
6 e For example, if Q
Here we mean the linear dimensions of Q. e were a disc then we could take its
“size” to be its radius.
274 Further Geometry of Differentiation
Take a small weight W and suspend it just above a point o of a horizontal table
using several feet of thread, perhaps attached to the ceiling. If you pull W to the
side by just an inch or two then, because the thread is long, W barely rises above the
table’s surface and we may idealize this to a movement on the table. Furthermore,
although the forces acting on W in this displaced position are actually gravity and
the tension in the thread, the net effect [exercise] is as though o were magically
pulling W towards it with a force proportional to r, as was required. To avoid the
possibility of confusion later, we stress that gravity is playing absolutely no essen-
tial role here; it is merely providing one particularly convenient way of simulating
a linear force field.
Now pull W a little bit away from o and give it a gentle flick in a random
direction. You see that the orbit of W is a closed curve traversed again and
again—a beautifully symmetrical oval shape centred at o. But exactly what is
this oval?
It is an ellipse! To demonstrate this, take the tabletop to be C with o as its origin.
Once again take W to have unit mass, and let its location at time t be z(t). For
simplicity’s sake, let the force directed towards the origin equal the distance |z|.
The differential equation governing the motion of W will therefore be z̈ = −z,
the two basic solutions of which are z = e±it . These represent counter-rotating
motions of unit speed around the unit circle. [Try launching W so as to produce
these solutions.] The general solution is then obtained as a linear combination of
these motions:
where p and q may, without any real loss of generality, be taken as real and
satisfying p > q.
As is illustrated in [5.22], the addition of such counter-rotating circular motions
results in elliptical motion with the attracting point at the centre. This becomes clear
if we rewrite (5.33) as
z = a cos t + ib sin t ,
E = 12 (v2 + r2 ).
As the particle orbits round the ellipse in [5.22], we see that this expression always
equals 12 (a2 + b2 ).
276 Further Geometry of Differentiation
7
Newton assumed that the force varies as a power of the distance, but it has since been discovered
that the result is still true if we drop this requirement.
Celestial Mechanics* 277
The first two terms correspond to an origin-centred ellipse with foci at ±2pq; the
last term therefore translates the left-hand focus to the origin.
Expressed in dynamical terms, this geometric result states that while leaving the
attracting point fixed at the origin, z ÞÑ z2 transforms an orbit of the linear field into
an orbit of the inverse square field. However, we are only in a position to state the
result in this way because we already know what the orbits in the two fields look
like. Is there instead some a priori reason why z ÞÑ z2 should map orbits of the linear
field to orbits of the gravitational field? If there were such a reason then [5.23] could
be viewed as a novel derivation, or explanation, of the elliptical motion of planets
about the sun as focus.
That there is indeed such a reason was discovered in the first decade of the 20th
century. Several people deserve credit for this beautiful result which, at the time
of this writing, is still not widely known. Apparently Bohlin (1911) was the first to
publish it, not knowing that Kasner (1913) had already discovered a more general
result in 1909. Finally, knowing only of Bohlin’s work, Arnol’d (1990) rediscovered
Kasner’s general theorem.
[5.23] Transformation of the linear force field into the inverse-square force field. Very
remarkably, squaring an origin-centred ellipse results in another ellipse. Furthermore, one
of the foci of the image ellipse is located at the origin! As the text explains, this provides
a novel explanation of the elliptical orbits of the planets around the Sun at one focus.
278 Further Geometry of Differentiation
[5.24] Geometric determination of the central force law from the orbit. If h is the (con-
stant) angular momentum of the orbiting particle, the magnitude ofthe central
force field
κ sec3 γ
that holds the particle in its orbit of curvature κ is given by F = h2 r2
.
Celestial Mechanics* 279
5.10.5 An Explanation
As z describes an arbitrary orbit, (5.35) tells us the force F needed to hold it in that
orbit. Now apply the mapping z ÞÑ z2 , and let a tilde denote a quantity associated
with the image, e.g., er = r2 . The force e
F needed to hold the image in its orbit is
e e 2 e
κ sec3 γe
F=h ,
er 2
and we now seek to relate this to the original force F.
First, to find e
κ, simply put f(z) = z2 into (5.24) and thereby obtain [exercise]
1 h cos γ κ i
e
κ= + .
2 r2 r
Next, observe that since the ray from 0 to z maps to the ray from 0 to z2 , the
conformality of the mapping implies that γ e = γ.
e
Putting these facts into the formula for F, and substituting for the original speed
and force from (5.34) and (5.35), we get
!2
e
h 1 2
v + 1
rF
e
F= 2 2
. (5.36)
h er 2
Even if F is a simple power law, generally this e F will not be. However, if and only
8
if the original force field is linear , the numerator in the above expression magically
8
However, as we shall see in a moment, it could be the repulsive linear field F = −r instead of the
attractive one F = +r.
280 Further Geometry of Differentiation
becomes the constant total energy E of the particle in the original field:
!2
e
h E
e
F= ! (5.37)
h er 2
Associated with each power law F ∝ rA there is precisely one power law
e e
F ∝ er A that is dual in the sense that orbits of the former are mapped to
orbits of the latter by z ÞÑ zm , and the relationships between the forces
and the mapping are:
e + 3) = 4 (A + 3)
(A + 3)(A and m = .
2
To their result we add the following point of clarification on the role of energy:
5.11.1 Introduction
Throughout this book we have stressed how functions may be viewed as geometric
entities that need not be expressed (nor even be expressible) in terms of formulae.
As an illustration of the limitations of formulae, consider
G(z) = 1 + z + z2 + z3 + · · · .
This power series converges inside the unit circle |z| = 1, and consequently it is
analytic there. Figure [5.25] shows a grid of little squares inside this circle being
amplitwisted by z ÞÑ w = G(z) to another such grid lying to the right of the verti-
cal line Re (w) = 12 , which itself is the image of the circle. Now this circle is certainly
a barrier to the power series formula for G, since it clearly diverges at 1: geomet-
rically, the image of the circle extends to ∞. However, the circle is not a barrier to
the geometric entity that the formula is attempting to describe.
Consider a somewhat different-looking power series centred at −1:
" 2 #
1 z+1 z+1
H(z) = 1+ + + ··· .
2 2 2
G and H within their circles of convergence; it thus constitutes the complete ana-
lytic continuation of the mapping. [We encourage you to use this fact to check the
details of the figure.] The simplicity of this example is perhaps misleading. Usu-
ally one cannot hope to capture the entire geometric mapping within a single closed
expression such as 1/(1 − z).
When one stares at a figure like [5.25] one starts to sense the rigid growth of the
mapping due to the analytic requirement that an expanding mesh of tiny squares
must map to another such mesh. It also becomes clear how the mapping itself is
oblivious to the different formulae with which we try to describe it. Indeed we have
seen that the two circles—such formidable and impenetrable barriers to the power
series—have only a slight significance for the mapping itself: both hit z = 1, so both
images extend to ∞.
Analytic Continuation* 283
5.11.2 Rigidity
The essential character of analytic rigidity is captured in the following result:
In short, the higher the derivative that vanishes at p, the greater the degree of
crushing at p.
We now apply this insight to the given situation. Let s be the (possibly) tiny seg-
ment that is crushed by f(z). The amplification of f at a point of s may be read off
by looking in any direction. By choosing to look along s we find that the amplifica-
tion vanishes at each point of s. The entire segment is therefore made up of critical
points for which f ′ = 0. Now think of f ′ as an analytic mapping in its own right, just
as we did in [5.18]. We have just seen that this mapping automatically possesses the
same property as f did: it crushes s to a point. We conclude that its derivative must
also vanish on s. Clearly there is no end to this; all the derivatives of f must vanish,
and, correspondingly, infinitesimal discs centred on s must be totally crushed.
This means that there is at least a sheathlike region surrounding s which is com-
pletely crushed by f. But if we take a new curve lying in this region, the whole line
284 Further Geometry of Differentiation
of thought may be repeated to deduce that f must crush a still larger region. The
collapse of the function therefore proceeds outwards (at the speed of thought!) to
the entire domain.
5.11.3 Uniqueness
Suppose that A(z) and B(z) are both analytic functions defined on a region that
happens to be the same size and shape as California. Suppose, further, that A and
B both happen to have the same effect on a tiny piece of curve, say a fallen eyelash
lying in a San Francisco street. This tiny measure of agreement instantly forces them
into total agreement, even hundreds of miles away in Los Angeles! For (A − B) is
analytic throughout California, and since it crushes the eyelash to 0, it must do the
same to the entire state.
We can express this slightly differently. If we arbitrarily specify the image points
of a small piece of curve s, then in general there will not exist an analytic function
that sends s to this image. However, the previous paragraph assures us that if we
can find such a function on a domain including s, then it is unique.
This is the “compelling reason” we referred to earlier in connection with the
uniqueness of the generalization of ex to complex values. For if an analytic gener-
alization E(z) exists, then we see that it will be uniquely determined by the values
of ex on even a small piece of the real axis. Of course knowing this does not help in
the least to find out what E(z) actually is. The value of our previous derivations of
explicit expressions for E(z) therefore remains undiminished. On the other hand,
the new knowledge is not without practical implications. Consider these three very
different-looking expressions:
z n
lim 1 + , ex (cos y + i sin y) , 1 + z + z2 /2! + z3 /3! + · · · .
nÑ∞ n
They are all analytic, and they all agree with ex when z is real. Thus, without further
calculation, we know they must all be equal to each other, for they can only be
different ways of expressing the unique analytic continuation of ex .
New and important aspects of uniqueness emerge when we consider domains
that merely overlap, rather than coincide. Let g(z) and h(z) be analytic functions
defined on the sets P and Q shown in [5.26a]. If they agree on even a small segment
s in P ∩ Q then they will agree throughout P ∩ Q. If we imagine that we initially
only know about g on P, then we may think of h as describing the same geometric
mapping as g but with the domain P extended to encompass Q. We are encouraged
in this view by the fact that g uniquely determines this analytic continuation. For
suppose h∗ were another continuation of g into Q. On s we would then have h∗ =
g = h, but this forces h∗ = h throughout their common domain Q.
The functions G(z) and H(z) of the introduction furnish a concrete example of
the above, where P happens to lie wholly within Q. The function 1/(1 − z) then
constitutes the analytic continuation of H to the rest of the plane.
Analytic Continuation* 285
[5.26] Analytic continuation along alternative routes. [a] If g(z) and h(z) are analytic
functions defined in overlapping regions P and Q, and they agree on even a small segment
s in this overlap, then they must agree throughout P ∩ Q, and we are entitled to think of
h(z) as the unique analytic continuation of g(z) from P into Q. [b] Although the analytic
continuations along PQR · · · S and along PQ eR
e · · · S are both uniquely determined, they
need not agree on S: in this case we have a multifunction.
f(x) = a + bx + c x2 + d x3 + · · · .
F(z) = a + bz + c z2 + d z3 + · · · .
is convergent and hence analytic. But since F(x) = f(x) on the real axis, it follows
that F is the unique analytic continuation of f to complex values. In other words, the
transition from f to its analytic continuation does not change the formula (series).
For our next example we consider a real identity involving two variables: ex ·ey =
ex+y . It will help to appreciate the argument if you can be temporarily stricken with
amnesia, so that the complex function ez and its associated geometry suddenly
mean nothing to you. Suppose that an analytic continuation of ex to complex values
exists, and call it E(z). We can now show that E must be subject to precisely the same
law, and without even knowing what E is!
Let Fζ (z) ≡ E(ζ)·E(z), and let Gζ (z) ≡ E(ζ+z). First note that for fixed ζ both F(z)
and G(z) are analytic functions of z. Now suppose that ζ is real, so that E(ζ) = eζ . If
z now moves on a segment of the real axis then it follows from the real identity that
F(z) = G(z); but from our recent results we know this implies that they are equal
everywhere. If we hold z fixed instead, then analogous reasoning yields Fζ = Gζ ,
and we conclude that
for complex values of both ζ and z. It should be clear that this reasoning extends
to any identity, even one involving more than two variables.
[5.28] Extending an analytic mapping via reflection. If the analytic function f maps P to
Q, then we can use it to construct an analytic mapping from P to Q, given by f⋆ (z) ≡ f(z).
As we see, the two reflections reverse angles twice, resulting in a conformal mapping.
f⋆ (z) ≡ f(z).
The figure explains why f⋆ is conformal, and hence analytic. All three stages,
a ÞÑ a ÞÑ f(a) ÞÑ f(a), preserve the magnitude of an angle at a; the first reflec-
tion reverses the sense, then f preserves the reversed sense, and finally the second
reflection undoes the damage, restoring the angle to pristine condition at f⋆ (a).
In general this mapping f⋆ will not be a continuation of f in any sense, rather
it is an entirely new mapping. This should become clear if you imagine moving P
downwards until some of it crosses the real axis. P and P now overlap so that P ∩ P
constitutes a common domain for f and f⋆ , but we hope you can see that there is no
reason for them to agree with each other. This is clearer still if we take an example:
exercise
f = (rotation of ϕ) =⇒ f⋆ = (rotation of − ϕ).
[5.29] Analytic continuation via reflection. Suppose that the analytic function f defined
in P is the complex extension of a real function defined on L. Then the function f⋆ (z) ≡
f(z) introduced in the previous figure is the analytic continuation of f into P. Generalizing,
if f maps a line-segment L (not necessarily real) to another line-segment L b , then we
can analytically continue from one side of L to the other by using the fact that points
symmetric in L map to points symmetric in L b.
Chapter 12 we will show that it is intimately connected with the so-called method
of images of electrostatics and fluid dynamics.
We now turn to the special circumstance under which f⋆ is the analytic continu-
ation of f. Suppose that f is itself the complex generalization of a real function, and
let P have a part L of its boundary along the real axis, as in [5.29]. Since f is real on L,
the image set Q will also border on the real axis. Unlike the general situation previ-
ously considered, f and f⋆ will now automatically agree on their common domain
P ∩ P = L, for if z is real then
We can now think of f and f⋆ as being two parts of a single analytic mapping F
on P ∪ P. Indeed, by considering what happens to the two halves of an infinites-
imal disc centred on L, it’s clear that F is analytic there, for the image is another
infinitesimal disc. [What happens if we are at a critical point?] Once again, notice
how different this is from the case of real functions, for we could easily join two
pieces of graph together with a kink at the join; their values would then agree,
while their derivatives would not.
Of course if f is already defined in P (as well as P) then f⋆ must simply repro-
duce the mapping that’s already there. For example, the formula for the complex
generalization sin z is valid everywhere, so it should be subject to the symmetry
f⋆ (z) = f(z). Indeed if we follow the three steps of a ÞÑ f⋆ (a) then we do find that
eia − e−ia e−ia − eia
a ÞÑ a ÞÑ ÞÑ = sin a .
2i −2i
Analytic Continuation* 289
We can rephrase our result in a more symmetric and slightly generalized [exer-
cise] form. If f maps a line-segment L (not necessarily real) to another line-segment
b , then we can analytically continue from one side of L to the other by using the
L
b.
fact that points symmetric in L map to points symmetric in L
This sounds very reminiscent of the conservation of symmetry by Möbius trans-
formations that we discovered in Chapter 3, and indeed by fusing these two
symmetry principles we can obtain a significant generalization of our result. Sup-
pose that instead of mapping a particular9 line to a line, f sends a part C of a circle
to a part C b of another circle. We can reduce this to the previous case by using
two Möbius transformations to send C ÞÑ L, and C b ÞÑ Lb . We deduce that points
symmetric in C map to points symmetric in C b.
As a mixed example, imagine that f maps part of the unit circle to part of the
real axis. If f is only known inside the circle then the above result tells us [exercise]
that there is an analytic continuation to the exterior given by
† 1
f (z) ≡ f .
z
The complete analytic function F is then defined to be f inside the circle, and f†
outside the circle. By construction, this function sends symmetric pairs of points to
conjugate images: F† (z) = F(z).
Using what is now known as Schwarzian reflection10 , in 1870 H. A. Schwarz was
able to generalize his Reflection Principle beyond lines and circles to more general
curves. We end this chapter with a description of Schwarz’s simple, yet fascinating,
idea. The key is to use an analytic function to fake conjugation.
We know that reflecting every point across the real axis (z ÞÑ z) is not an analytic
function. However, given a sufficiently smooth11 curve K, it is possible to find an
analytic function SK (z) that selectively sends just the points of K to their conjugates:
z∈K =⇒ SK (z) = z .
Davis and Pollak (1958) christened SK the Schwarz function of K. We can now define
the Schwarzian reflection of z across K to be e
z = ℜK (z), where
ℜK (z) ≡ SK (z) .
To see why this is a good idea, consider [5.30]. First note that points on K are
unaffected, in accord with the ordinary notion of reflection, e.g.,
e = SK (q) = (q) = q .
q
9
We stress “particular”, because if a general line were sent to a line then the mapping could only be
linear. Similarly, in the new case, if a general circle were sent to a circle then the mapping would have
to be a Möbius transformation.
10
Schwarz (1972a).
11
The curve must in fact be “analytic”. On this point, see Davis (1974), which also contains many
interesting applications of the Schwarz function. A more advanced work is Shapiro (1992).
290 Further Geometry of Differentiation
where ϕ is the angle that the tangent to K makes with the horizontal. It is now clear
from the symmetry of the figure that if the point a is on the infinitesimal circle, then
e is indeed its reflection across the tangent of K. Thus, at least very close to K, z ÞÑ e
a z
is a reasonable generalization of the reflection concept. Furthermore, reflecting in
K twice yields the identity mapping, as it should. For since ℜK is anticonformal,
ℜK ◦ ℜK is conformal, i.e., analytic. But since this function maps each point of K to
itself, and since an analytic function is determined by its values on a curve, ℜK ◦ℜK
must be the identity mapping.
We leave it to the exercises for you to show that if K is a line or a circle then e z
is just the ordinary reflection, even if z is far from K. For example, the unit circle C
may be written as zz = |z|2 = 1, so that on C we have z = (1/z). Thus its Schwarz
function is SC (z) = (1/z), and so ℜC (z) = (1/z), which is just inversion in C.
Analytic Continuation* 291
Let us give a less trivial example, namely, reflection in the ellipse E with equation
(x/a)2 + (y/b)2 = 1. Writing x = 12 (z + z) and y = 2i 1
(z − z), then solving for z in
terms of z, we find [exercise] that
1 h p i
SE (z) = 2 (a 2
+ b 2
) z − 2ab z 2 + b 2 − a2 .
a − b2
With a = 2 and b = 1, for example, Schwarzian reflection is given by
h p i
ℜE (z) = 13 5 z − 4 z2 − 3 ,
which is illustrated in [5.31]. We encourage you to verify this figure with your
computer, as well as to examine the effect of ℜE on other shapes.
With the proper concept of reflection in hand, we may now generalize the above
method of analytic continuation across lines and circles to more general curves.
Let f be an analytic mapping defined in a region P bordering on a curve L that is
smooth enough to possess a Schwarz function, and let L b = f(L) be the image of L
under f. Much as in figures [5.28] and [5.29], we may now analytically continue f
across L by demanding that points that are symmetric in L map to points that are
symmetric in Lb . Thus [draw a picture!] the continuation f‡ of f to P
e = ℜL (P) is
f‡ = ℜLb ◦ f ◦ ℜL .
e for it is the compo-
By the same argument as in [5.28], this is indeed analytic in P,
sition of one conformal mapping with two anticonformal mappings. Also, f‡ = f
e is then subject to
on L. The complete analytic function F given by f in P and f‡ in P
‡
the symmetry F = F. This is Schwarz’s Symmetry Principle.
292 Further Geometry of Differentiation
Our previous results are just special cases of this construction. For example, if L
and Lb are segments of the real line then ℜL (z) = ℜ b (z) = z, so f‡ = f⋆ , as before.
L
b is a segment of
Similarly, if L is an arc of the unit circle (so that ℜL (z) = 1/z) and L
‡ †
the real line (so that ℜLb = z), then f = f , as before.
Exercises 293
5.12 Exercises
·
1 Show that if f = u + iv is analytic then (∇u) (∇v) = 0, where ∇ is the
“gradient operator” of vector calculus. Explain this geometrically.
2 Show that the real and imaginary parts of an analytic function are harmonic,
i.e., they both automatically satisfy Laplace’s equation:
∆Φ = 0 ,
(i) Show that the critical points of P(z) are the solutions of
1 1 1
+ + ··· + =0.
z − a1 z − a2 z − an
(ii) Let K be a circle with centre p. By considering the conjugate of the
equation in (i), deduce that p is a critical point if and only if it is the
centre of mass of the inverted points IK (aj ).
(iii) Show that the equation in (i) is equivalent to
z − a1 z − a2 z − an
+ + ··· + = 0,
|z − a1 |2 |z − a2 |2 |z − an |2
and by interpreting the LHS as a (positively) weighted sum of the vectors
from z to the roots of P(z), deduce Lucas’ Theorem: The critical points of a
polynomial in C must all lie within the convex hull of its zeros. This is a com-
plex generalization of Rolle’s Theorem in ordinary calculus. [Hint: Use the
fact that (2.32) on page 116 is still valid even if the masses are not equal.]
11 Use (ez ) ′ = ez to show that the derivatives of all the trig functions are given
by the familiar rules of real analysis.
12 Provided it is properly interpreted, show that (zµ ) ′ = µ zµ−1 is still true even
if µ is complex.
13 (i) If a is an arbitrary constant, show that the series
a(a − 1) 2 a(a − 1)(a − 2) 3
f(z) = 1 + az + z + z + ···
2! 3!
converges inside the unit circle.
(ii) Show that (1 + z)f ′ = af.
(iii) Deduce that [(1 + z)−a f] ′ = 0.
(iv) Conclude that f(z) = (1 + z)a .
14 As we pointed out in Chapter 3, stereographic projection has a very practical
use in drawing a conformal map of the world. Once we have this map we
can go on to generate further conformal maps, simply by applying different
analytic functions to it. One particularly useful one was discovered (using
other means) by Gerhard Mercator in 1569. We can describe it (though he
could not have) as the result of applying log(z) to the stereographic map.
(i) Look up both a stereographic map and a Mercator map in an atlas,
and make sure you can relate the changes in shape you see to your
understanding of the complex logarithm.
(ii) Imagine plotting a straight-line course on a Mercator map and then
actually travelling it on the high seas. Show that as you sail, the reading
of your compass never changes.
Exercises 295
15 (i) By noting that the unit tangent (in the counterclockwise direction) to an
origin-centred circle can be written as b ξ = i(z/|z|), show that formula
(5.24) for the curvature of the image of such a circle can be written as
h i
1 + Re z f ′′
′
e
κ= f .
|z f ′ |
(ii) What should this formula yield if f(z) = log z? Check that it does.
(iii) What should this formula yield if f(z) = zm ? Check that it does. What
is the significance of the negative value of e
κ when m is negative? [Hint:
Which way does the velocity complex number of the image rotate as z
travels counterclockwise round the original circle?]
16 As illustrated below, a region is called convex if all of it is visible from an arbi-
trary vantage point inside. Let an analytic mapping f act on an origin-centred
circle C to produce a simple image curve f(C), the interior of which is convex.
(i) From the formula of Ex. 15, deduce that if f maps the interior of C to the
interior of f(C), then the following inequality holds at all points z of C:
z f ′′
Re ⩾ −1.
f′
(ii) What is the analogous inequality when f maps the interior of C to the
exterior of f(C). [Hint: Ex. 4, p. 239.]
(iv) Repeat as much as possible of the above analysis in the cases f(z) = log(z)
and f(z) = zm , where m is a positive integer. [In neither of these cases
will you be able to see the exact value of |K(p)|.]
(v) According to the geometric reasoning in Ex. 18, p. 241, the amplification
of a Möbius transformation M(z) = az+b cz+d is constant on each circle
centred at −(d/c). Thus the complex curvature of M should be tangent
to these concentric circles. Verify this by calculating K.
(vi) Use a computer to verify figure [5.21] for all four mappings above.
18 Let two curves C1 and C2 emerge from a point p in the same direction. Two
examples are illustrated below.
(iv) Use the previous two parts to show that all Möbius transformations
have vanishing Schwarzian derivative. [Hint: Recall that the mappings
in part (ii) generate (via composition) the set of all Möbius trans-
formations.] Remark: Ex. 19, p. 481 shows that the converse is also
true: If {f(z), z} = 0 then f = Möbius. Thus Möbius transformations are
completely characterized by their vanishing Schwarzian derivative.
(v) Use the previous two parts to show that the Schwarzian derivative is
“invariant under Möbius transformations”, in the following sense: if M
is a Möbius transformation, and f is analytic, then
20 Think of the real axis as representing time t, and let a moving particle
w = f(t) (where f(z) is analytic) trace an orbit curve C. The velocity is then
v = ẇ = f ′ (t).
(i) Use (5.24) to show that the curvature of C is
Im(v̇/v)
e
κ= .
|v|
(ii) Argue that this result does not in fact depend on C being produced by
an analytic mapping, but is instead true of any motion for which the
velocity v and acceleration v̇ are well-defined.
(iii) Show that the formula may be rewritten as
Im(v v̇)
e
κ= .
|v|3
298 Further Geometry of Differentiation
D is symmetric in C ⇐⇒ C is symmetric in D.
Briefly, we may simply say that “C and D are symmetric”. Let’s see what
happens if we generalize C and D to intersecting arcs possessing Schwarz
functions, and generalize inversion to Schwarzian reflection.
(i) Explain why the statement “D is symmetric in C” is the same as “if
the point d lies on D then ℜD [ℜC (d)] = ℜC (d)”. Must the arcs be
orthogonal?
(ii) If D is symmetric in C, deduce that the mappings (ℜD ◦ ℜC ) and
(ℜC ◦ ℜD ) are equal at points of D.
(iii) Using the fact that these two mappings are analytic (why?), deduce that
C must also be symmetric in D. Thus, as with circles, we may simply say
that C and D are symmetric.
CHAPTER 6
Non-Euclidean Geometry*
6.1 Introduction
Visual Complex Analysis. 25th Anniversary Edition. Tristan Needham, Oxford University Press.
© Tristan Needham (2023). DOI: 10.1093/oso/9780192868916.003.0006
304 Non-Euclidean Geometry*
It turns out that the larger the figures examined, the larger the deviations from
the predictions of Euclidean geometry. However, it’s important to realize just how
small these deviations typically are for figures of reasonable size. For example, sup-
pose we measure the circumference of a circle having a radius of one meter. Even
if our measuring device were capable of detecting a discrepancy the size of a single
atom of matter, no deviation from Euclidean geometry would be found! Little won-
der, then, that for two thousand years mathematicians were seduced into believing
that Euclidean geometry was the only logically possible geometry.
It is a marvellous tribute to the power of human mathematical thought that non-
Euclidean geometry was discovered a full century before Einstein found that it was
needed to describe gravity. To locate the seeds of this mathematical discovery, let
us return to ancient Greece.
Euclid began with just five axioms, the first four of which never aroused con-
troversy. The first axiom, for example, merely states that there exists a unique line
passing through any two given points. However, the status of the fifth axiom (the
so-called parallel axiom) was less clear, and it became the subject of investigations
that ultimately led to the discovery of non-Euclidean geometry:
Parallel Axiom. Through any point p not on the line L there exists
(6.1)
precisely one line L ′ that does not meet L.
Figure [6.1a] illustrates the parallel axiom, and it also explains why this axiom can-
not be experimentally tested, at least as stated. As the line M rotates towards L ′ ,
the intersection point q moves further and further away along L. Our geometric
intuition is based on figures drawn in a finite portion of the plane, but to verify
that L ′ never meets L, we need an infinite plane. We can certainly try to imagine
what an infinite plane would be like, but we have no first hand experience to back
up our hunches.
These are very modern doubts we are expressing. Historically, mathematicians
fervently believed in (6.1), so much so that they thought it must be a logically nec-
essary property of straight lines. But in that case they ought to be able to prove it
outright, instead of merely assuming it as Euclid had done.
Many attempts were made to deduce (6.1) from the first four axioms, one of the
most penetrating being that of Girolamo Saccheri in 1733.1 His idea was to show
that if (6.1) were not true, then a contradiction would necessarily arise. He divided
the denial of (6.1) into two alternatives:
Spherical Axiom. There is no line through p that does not meet L. (6.2)
or
Hyperbolic Axiom. There are at least two lines through p that do not meet L. (6.3)
1
See Stillwell (2010).
Introduction 305
[6.1] Two forms of the Parallel Axiom. [a] The Parallel Axiom seems beyond empirical
verification: As the line M rotates towards L ′ , the intersection point q moves further and
further away along L, so to be certain that the supposedly unique parallel line L ′ never
intersects L, we need to go all the way to infinity! [b] However, an alternative formulation
of the Parallel Axiom is subject to empirical verification or refutation: The angles in every
triangle must sum to π, so that E = 0.
Our naming of (6.2) will become clear shortly, but the use of “hyperbolic” in
connection with (6.3) is more obscure, though standard.
In the case of (6.2), Saccheri was indeed able to obtain a contradiction, provided
“lines” are assumed to have infinite length. If we drop this requirement, then we
obtain a non-Euclidean geometry called spherical geometry. This is the subject of the
following section.
In the case of (6.3), Saccheri and later mathematicians were able to derive
very strange conclusions, but they were not able to find a contradiction. As we
now know, this is because (6.3) yields another viable non-Euclidean geometry,
called hyperbolic geometry. Of the two non-Euclidean geometries obtained from (6.2)
and (6.3), hyperbolic geometry is by far the more intriguing and important: it is
an essential tool in many areas of contemporary research. Furthermore, there is
even a sense (to be discussed later) in which hyperbolic geometry subsumes both
Euclidean and spherical geometry.
(Angle sum of T ) = π.
• A somewhat more natural way of defining the absolute unit of length is in terms
of the constant K. Since the radian measure of angle is defined as a ratio of
lengths, E is a pure number. On the other hand, the area A has units of (length)2 .
It follows that K has units of 1/(length)2 and so it can be written as follows in
terms of a length R: K = +(1/R2 ) in spherical geometry; K = −(1/R2 ) in hyper-
bolic geometry. Later we will see that this length R can be given a very intuitive
interpretation P.
• The smaller the triangle, the harder it is to distinguish it from a Euclidean tri-
angle: only when the linear dimensions are a significant fraction of R will the
difference become obvious. This is why Gauss chose the biggest triangle he
could in his experiment. Einstein’s theory explains why Gauss’s triangle was
nevertheless much too small: the weak gravitational field in the space surround-
ing the earth corresponds to a microscopic value of K and hence to an enormous
value of R. It would have been a different story if Gauss had been able to perform
his experiment in the vicinity of a small black hole!
2
This oversimplification does not do justice to Beltrami’s accomplishments. Later in this chapter
we shall see what Beltrami really did!
308 Non-Euclidean Geometry*
[6.2] The Pseudosphere. In 1868 Eugenio Beltrami finally removed the mystery and scep-
ticism that had overshadowed the strange, abstract laws of hyperbolic geometry ever
since their discovery, 40 years earlier. He did so by recognizing that the hyperbolic laws
simply are the intrinsic laws of geometry that hold true within a surface called the pseudo-
sphere (shown here) that has constant negative Gaussian curvature. The pseudosphere
extends upward indefinitely, but the base circle is a boundary beyond which it cannot be
extended.
To explain what we mean by this, let us first discuss how we may “do geometry”
on a more general surface, such as the surface of the strange looking vegetable3
shown in [6.3]. The idea of doing geometry on such a surface is essentially due to
Gauss and (in greater generality) to Riemann.
The first thing we must do is to replace the concept of a straight line with that
of a geodesic. Just as a line-segment in a flat plane may be defined as the short-
est route between two points, so a geodesic segment connecting two points on
a curved surface may be defined (provisionally) as the shortest connecting route
within the surface. For example, if you were an ant living on the surface in [6.3],
and you wanted to travel from a to b as quickly as possible, then you would fol-
low the illustrated geodesic segment. The figure also shows the geodesic segment
connecting another pair of points, c and d.
Here is a simple way you can actually construct such geodesic segments: take a
thread and stretch it tightly over the surface to connect the points a and b. Provided
3
European readers may think this an imaginary vegetable, but Americans can buy it in the
supermarket.
Introduction 309
that the thread can slide around on the surface easily, the tension in the thread
ensures that the resulting path is as short as possible. Note that in the case of cd,
we must imagine that the thread runs over the inside of the surface. In order to deal
with all possible pairs of points in a uniform way, it is therefore best to imagine the
surface as made up of two thinly separated layers, with the thread trapped between
them.
In fact there exists a wonderfully simple and practical (yet surprisingly little
known) method of constructing geodesics on any part of a surface. Cut a long,
straight, narrow strip of masking tape (aka painter’s tape) and start to lay one end
down on the surface in the direction you want your geodesic to go. Holding the
free end up away from the surface with one hand, use your other hand to gently
run your finger along the tape, allowing it to choose its own path as it sticks itself
to the surface as it goes. Behold! You have created a geodesic, and this construction
works equally well on ab and cd! In the case of ab, you can easily check the validity
of the construction by stretching a string over the surface: it will coincide with path
of your sticky tape! For the explanation of why this works (and many applications)
see Needham (2021).
It is now obvious how we should define distance in this geometry: the distance
between a and b is the length of the geodesic segment connecting them. Figure
[6.3] shows how we can then define, for example, a circle of radius r and centre p
as the locus of points at distance r from p. To construct this circle we may take a
piece of thread of length r, hold one end fixed at p, then (keeping the thread taut)
drag the other end round on the surface.
Given three points on the surface, we may join them with geodesics to form a
triangle; [6.3] shows two such triangles, ∆1 and ∆2 . Now look at the angles in ∆1 .
Clearly E(∆1 ) > 0, like a triangle in spherical geometry, while E(∆2 ) < 0, like a
triangle in hyperbolic geometry.
On the other hand, if you were an intelligent ant living on this patch, no geomet-
ric experiment you could perform within the surface would reveal that any change
had taken place whatsoever. We say that the intrinsic geometry has not changed.
For example, the curves into which the edges of ∆1 have been deformed are still
the shortest routes on the surface. Correspondingly, the value of E is unaffected by
stretch-free bending: E is governed by intrinsic (not extrinsic) curvature.
To highlight this fact, consider [6.4]. On the left is a flat piece of paper on which
we have drawn a triangle T with angles (π/2), (π/6), (π/3). Of course E(T ) = 0.
Clearly we can bend such a flat piece of paper into either of the two (extrinsically)
curved surfaces on the right4 . However, intrinsically these surfaces have undergone
no change at all—they are both as flat as a pancake! The illustrated triangles on
these surfaces (into which T is carried by our stretch-free bending of the paper) are
4
Of course the conical example on the far right cannot be obtained by bending a rectangle.
Introduction 311
[6.4] A bent piece of paper remains intrinsically flat. A flat piece of paper can be bent
into all kinds of extrinsically curved shapes in space, but its internal, intrinsic geometry
remains totally flat, and triangles constructed within the surface continue to obey all the
normal laws of Euclidean geometry, such as Pythagoras’s Theorem, and E = 0.
identical to the ones that intelligent ants would construct using geodesics, and in
both cases E = 0: geometry on these surfaces is Euclidean.
5
Gauss (1827).
6
Other names are intrinsic curvature, total curvature, or just plain curvature.
312 Non-Euclidean Geometry*
[6.5] The sign of the Gaussian curvature K(p) tells us the shape of the surface at p.
As we will now start to explain, it is no accident that we have used the same sym-
bol to represent Gaussian curvature as we earlier used for the constant occurring
in (6.4)—they are the same thing!
Gauss originally defined K(p) as follows. Let Π be a plane containing the nor-
mal vector n to the surface at p, and let κ be the (signed) curvature at p of the
curve in which Π intersects the surface. The sign of κ depends on whether the cen-
tre of curvature is in the direction n or −n. The so-called principal curvatures are
the minimum κmin and the maximum κmax values of κ as Π rotates about n. [Inci-
dentally, Euler had previously made the important discovery that these principal
curvatures occur in two perpendicular directions.] Gauss defined K as the product
of the principal curvatures:
K ≡ κmin κmax .
Note that this definition is in terms of the precise shape of the surface in space
(extrinsic geometry). However, Gauss (1827) went on to make the astonishing dis-
covery that K(p) actually measures the intrinsic curvature of the surface, that is, K is
invariant under bending! Gauss was justifiably proud of this result, calling it the The-
orema Egregium (Latin for “remarkable theorem”). As an example of the result, you
may visually convince yourself that K = 0 everywhere on each of the intrinsically
flat surfaces in [6.4].
The intrinsic significance of K is exhibited in the following fundamental result:
If ∆ is an infinitesimal triangle of area dA located at the point p, then
E(∆) = K(p) dA. (6.5)
Since E and dA are defined by the intrinsic geometry, so is K = (E/dA). Once again,
we refer you to works on differential geometry for a proof of (6.5).
It follows from (6.5) [see Ex. 1] that the angular excess of a non-infinitesimal
triangle T is obtained by adding up (i.e., integrating) the Gaussian curvature over
the interior of T :
Introduction 313
ZZ
E(T ) = K(p) dA. (6.6)
T
Figure [6.6] illustrates this using the simplest surfaces of each type. To obtain an
added bonus, recall that we previously associated an absolute unit of length R with
a non-Euclidean geometry by writing K = ±(1/R2 ). The bonus is that this length
R now takes on vivid meaning: in spherical geometry R is simply the radius of
the sphere, while in hyperbolic geometry it is the radius of the circular base of the
pseudosphere (called the radius of the pseudosphere). These two interpretations
will be justified later.
The requirement of constant curvature can be understood more intuitively by
reconsidering the discussion at the end of Chapter 1. There we saw that a central
idea in Euclidean geometry is that of a group of motions of the plane: one-to-one
mappings that preserve the distance between all pairs of points. For example, two
figures are congruent if and only if there exists a motion that carries the first into
coincidence with the second. In order that this basic concept of equality be available
in non-Euclidean geometry, we require that our surface admits an analogous group
of motions. If we take one of the triangles on the surface in [6.3], it’s clear that we
cannot slide it to a new location and still have it fit the surface snugly, because the
way in which the surface is curved at the new location is different: variation in the
curvature is the obstruction to motion.
314 Non-Euclidean Geometry*
[6.6] Euclidean, Spherical, and Hyperbolic geometry are the intrinsic geometries of
surfaces of constant Gaussian curvature.
and rotated freely, always fitting perfectly snugly against the surface. Exercise 15
shows how you can build your own pseudosphere; once built, you can verify this
surprising claim experimentally.
[6.7] Harriot’s 1603 proof that (∆) = (1/R2 ) (∆). Extending the sides of ∆ that
contain the angle α, we obtain the shaded wedge on the right, which occupies (α/2π)
of the complete sphere, so A(∆) + A(∆α ) = 2αR2 ; likewise for β and γ. Since each of
the four triangles on the left has an antipodal partner of equal area—as can be seen more
clearly on the right—it follows that together they occupy half the sphere: A(∆)+A(∆α )+
A(∆β ) + A(∆γ ) = 2πR2 , and the result follows immediately.
7
See Stillwell (2010, §17.6).
Spherical Geometry 317
On the other hand, it is clear in [6.7b] that ∆ and ∆α together form a wedge whose
area is (α/2π) times the area of the sphere:
Similarly,
A(∆) = (α + β + γ − π)R2 .
In other words,
as was to be shown.
[6.8] Reflection of the sphere in a line. [a] If Π is the plane through the line (great circle)
L, then reflection RΠ of space across Π induces the anticonformal reflection RL of the
sphere across L. [b] This reflection RL (a) can instead be expressed in intrinsic terms that
make sense to the inhabitants of the sphere: travel some distance d along the orthogonal
“line” M from a until you arrive at L, then travel an equal distance d again.
8
If L is thought of as the equator, then when a is one of the poles there are infinitely many M’s—
pick any one you like.
Spherical Geometry 319
[6.9] Rotation of the sphere via two reflections. [a] If the planes Π1 and Π2 meet at
angle (θ/2) then (RΠ2 ◦ RΠ1 ) is a rotation of space through angle θ about their line of
intersection. [b] The induced transformation of the sphere is therefore a rotation Rθp by θ
about the intersection point p of the lines on the sphere: RL2 ◦ RL1 = Rθp .
In Chapter 1 we saw that every direct motion of the plane was the composition
of two reflections: a rotation if the lines intersected; a translation if the lines were
parallel. We will now see that a similar phenomenon occurs on the sphere, but
because every pair of lines intersect, the composition of two reflections is always a
rotation—the sphere has no motions analogous to translations.
Figure [6.9a] illustrates the composition (RΠ2 ◦ RΠ1 ) of two reflections of space.
Here the planes Π1 and Π2 intersect in a line with direction vector v, and the angle
from Π1 to Π2 is (θ/2). Restricting attention to any one of the shaded planes orthog-
onal to v, we see that the transformation induced by (RΠ2 ◦ RΠ1 ) is (Rl2 ◦ Rl1 ),
where l1 and l2 are the lines in which Π1 and Π2 intersect the plane. Since (Rl2 ◦ Rl1 )
is a rotation of the plane through θ about the intersection point of l1 and l2 , it
is now clear that (RΠ2 ◦ RΠ1 ) is a rotation of space through angle θ about the
axis v.
Figure [6.9b] translates this idea into spherical terms. If Π1 and Π2 pass through
the centre of the sphere, and the lines (great circles) in which they intersect the
sphere are L1 and L2 , then
In other words,
Note that there is precisely one line P that is mapped into itself by Rθp . If we orient
P in agreement with the rotation (as illustrated) then we obtain a one-to-one corre-
spondence between oriented lines and points: P is called the polar line of p, and p
is called the pole of P.
In the case of the plane we used the analogue of (6.10) to show that the composi-
tion of two rotations about different points was equivalent (in general) to a single
rotation about a third point; exceptionally, however, two rotations could result in
a translation. As you might guess, in the case of the sphere there are no exceptions:
Figure [6.10a] shows how this may be established using exactly the same argument
that was used in the plane. In order to find the net effect of (Rϕ
q ◦ Rp ), draw the
θ
Rϕ
q ◦ Rp = (RN ◦ RM ) ◦ (RM ◦ RL ) = RN ◦ RL = Rr .
θ ψ
ψ = θ + ϕ − 2KA.
We may now complete the classification of the motions of the sphere. As we have
remarked, there is precisely one motion of the sphere that carries a given spherical
triangle abc to a given congruent image triangle. Figure [6.10b] helps to refine this
result. Using the same logic as was used in the plane, we see [exercise] that
There is exactly one direct motion M (and exactly one opposite motion
e that maps a given line-segment ab to another line-segment a ′ b ′ of
M)
e = (RL ◦ M ), where L is the line through (6.12)
equal length. Furthermore, M
′ ′
a and b .
Figure [6.10b] also shows how we may construct M. Draw the line P through a
and a ′ , and let p be its pole. With the appropriate value of θ, it’s clear that Rθp
will carry the segment ab along P to a segment of equal length emanating from
a ′ ; finally, an appropriate rotation Rϕ ′ ′ ′
a ′ about a will carry this segment into a b .
ϕ
Thus M = (Ra ′ ◦ Rp ), which is equivalent to a single rotation by virtue of (6.11).
θ
Every direct motion of the sphere is a rotation, and every opposite motion is
(6.13)
the composition of a rotation and a reflection.
As a simple test of this result (and your grasp of it) consider the antipodal mapping
that sends every point on the sphere to its antipodal point. Clearly this is a motion,
but how does it accord with the above result?
9
If the curvature is not constant, two surfaces can have equal curvature at corresponding points and
yet have different intrinsic geometry.
322 Non-Euclidean Geometry*
[6.11] Non-spherical surfaces of constant positive curvature. There exist surfaces of con-
stant positive curvature that are not a portion of the sphere, but Minding proved that their
intrinsic geometry is identical to that of the sphere, at least so long as we stay away from
the pointed tips on the left, or the edges on the right.
The sphere also has the advantage of making it obvious that its intrinsic geom-
etry admits a group of motions: in [6.11] it’s certainly not clear that figures can be
freely moved about and rotated on the surface without stretching them. Neverthe-
less, the above discussion shows that the actual shape of a surface in space is a
distraction, and it would be better to have a more abstract model that captured the
essence of all possible surfaces having the same intrinsic geometry.
By the “essence” we mean knowledge of the distance between any two points,
for this and this alone determines the intrinsic geometry. In fact—and this is a fun-
damental insight of differential geometry—it is sufficient to have a rule for the
infinitesimal distance between neighbouring points. Given this, we may determine
the length of any curve as an infinite sum (i.e., integral) of the infinitesimal seg-
ments into which it may be divided. Consequently, we may also identify the “lines”
of the geometry as shortest routes from one point to another, and we can likewise
[exercise] determine angles.
This leads to the following strategy for capturing the essence of any curved sur-
face S (not necessarily one of constant curvature). To avoid the distraction of the
shape of the surface in space, we draw a map (in the sense of a geographical atlas)
of S on a flat piece of paper. That is we set up a one-to-one correspondence between
points bz on S and points z on the plane, which we will think of as the complex plane.
Now consider the distance d b s separating two neighbouring points b b on
z and q
S. In the map, these points will be represented by z and q = z + dz, separated
by (Euclidean) distance ds = |dz|. Once we have a rule for calculating the actual
separation d b s on S from the apparent separation ds in the map, then (in principle)
we know everything there is to know about the intrinsic geometry of S.
Spherical Geometry 323
[6.12] Central projection maps the lines of the sphere to the lines of the map.
324 Non-Euclidean Geometry*
Next, consider (3.6) on p. 142, which describes the effect of inversion on the sepa-
ration of two points. By taking the limit in which the two points coalesce, we may
apply this result to [6.13b] to obtain [exercise]
2
db
s = ds.
[Nz]2
This can also be obtained more directly, without using (3.6), by [exercise] choosing
db
s parallel to C. Finally, applying Pythagoras’ Theorem to the triangle Nz 0, we
obtain
2
db
s = ds. (6.16)
1 + |z|2
This flat conformal map with metric (6.16) is the desired abstract depiction of all
possible surfaces of constant Gaussian curvature K = +1.
Λ( e
z)de
s = Λ(z) ds.
point of any two circles centred on L b and passing through bz . Note that these two
b
circles are orthogonal to L . Figure [6.14b] shows what this construction looks like in
the stereographic map. Since stereographic projection preserves circles and angles,
the two circles orthogonal to L b and passing through b z are mapped to two circles
orthogonal to L and passing through z. The second intersection point of these circles
is thus the reflection IL (z) of z in L! To sum up,
[For a different proof of (6.18), one that is perhaps even more natural than the one
above, see Ex. 2.] As an important special case, note that if Lb is the intersection of
Σ with the vertical plane through the real axis, then reflection of Σ in L b induces
complex conjugation, z ÞÑ z.
Now let’s find the complex functions corresponding to rotations of Σ. Figure
[6.15] illustrates a rotation R ψ b be
b . Let b
b of Σ through angle ψ about the point a
a
the antipodal point to a b , so that b = −(1/a) [see (3.23), p. 167]. These points a b
b
and b lie on the axis of the rotation and remain fixed; correspondingly, a and
b will be the fixed points of the induced transformation of C. Furthermore, it is
clear geometrically that the effect of the induced transformation on an infinites-
imal neighbourhood of a is a rotation about a [exercise], and, by virtue of our
conventions, the rotation angle is negative ψ.
Spherical Geometry 327
According to (6.10),
Rψ b 2 ◦ RL
b = RL
a b 1,
where L b 1 and L
b 2 are any two lines passing through a b (and hence also through b b)
such that the angle between them is (ψ/2). Since stereographic projection preserves
circles and angles, the images in C of these lines will be two circles L1 and L2 passing
through the fixed points a and b, and containing angle (ψ/2) there. It follows from
(6.18) that the transformation R ψa induced by the rotation R a
ψ
b is
a = IL2 ◦ IL1 .
Rψ
Note that this is in agreement with (3.39), p. 184: rotations of Σ induce Möbius
transformations of the form
Az + B
Rψa (z) = . (6.20)
−Bz + A
By virtue of (6.13), this formula represents the most general direct motion of Σ.
We have already noted that z ÞÑ z corresponds to a reflection of Σ, and it follows
[exercise] that the most general opposite motion is represented by a function of the
form
Az + B
z ÞÑ .
−B z + A
Figure [6.10b] provided a very elegant geometric method of composing rotations
of space. The above analysis now opens the way toan equally elegant method of
ψ
computing the net rotation produced by R ωb ◦ Ra
p b . All we need do is compose
the corresponding Möbius transformations:
ω ω ψ
Rp ◦ Rψ
a = Rp Ra .
An otherwise tricky problem has been reduced to multiplying 2 × 2 matrices!
Spherical Geometry 329
v = l i + m j + n k, l2 + m2 + n2 = 1.
Referring back to (3.20) on p. 166, we see that a and v are related as follows:
l + im 1+n
a= and |a|2 = .
1−n 1−n
Substituting these expressions into (6.19), and removing the common factor of
2/(1 − n), we obtain [exercise]
h i cos(ψ/2) + in sin(ψ/2) (−m + il) sin(ψ/2)
Rψv = . (6.21)
(m + il) sin(ψ/2) cos(ψ/2) − in sin(ψ/2)
h i
You may check for yourself that this matrix is “normalized”: det R ψ v = 1. This
makes life that much easier, for when we multiply two such matrices the resulting
matrix will be of precisely the same form. Thus, by comparing the result with (6.21),
we may read off the net rotation.
For example, suppose we perform a rotation of (π/2) about i, followed by a
rotation of (π/2) about j. The Möbius matrix of the net rotation will therefore be
" # " # " #
1 1 −1 1 1 i 1 1 − i −1 + i
√ √ = . (6.22)
2 1 1 2 i 1 2 1+i 1+i
Comparing this with (6.21), we see [exercise] that this is rotation of ψ = (2π/3)
about the axis v = √13 (i + j − k).
V = v1 + v1 I + v2 J + v3 K, (6.23)
where the coefficients are all real numbers. To define the product of two such
quaternions, Hamilton took 1 to be the identity, and he took I, J, K to be three
different square roots of −1, each analogous to i:
I2 = J2 = K2 = −1. (6.24)
These relations probably look familiar: they are formally identical to the vector
products of the basis vectors i, j, k in three-dimensional space. For example, i×j =
k = −j×i. We can use this analogy between i, j, k and I, J, K to express the product
of two quaternions in a particularly simple way.
First, let’s use the analogy to simplify the notation (6.23). As in ordinary algebra,
we suppress the identity 1 in the first term and write v 1 = v, which Hamilton called
the scalar part of V. Next we collect the remaining three terms into V ≡ v1 I + v2 J +
v3 K, which Hamilton called the vector part of V. Thus (6.23) becomes
V = v + V.
In the special case where the scalar part v vanishes, Hamilton called V = V a pure
quaternion. Historically, the concept of a pure quaternion was the forerunner of the
idea of an ordinary vector in space. In fact the very word “vector” was coined by
Hamilton in 1846 as a synonym for a “pure quaternion”.
If we multiply V by another quaternion W = w + W, then (6.24) and (6.25) imply
[exercise] that
·
V W = (v w − V W) + (v W + w V + V×W). (6.26)
·
V W = −V W + V×W. (6.27)
were viewed as merely two facets (the scalar and vector parts) of quaternion multi-
plication. However, it did not take physicists long to realize that the scalar product
and the vector product were each important in their own right, independently of
the quaternions from which they had both sprung.
Further results on quaternions will be derived in the exercises; here we wish
only to explain the connection between quaternions and rotations of space. This
connection hinges on the idea of a binary rotation, which means a rotation of space
though an angle of π. The appropriateness of the word “binary” stems from the
fact that if the same binary rotation is applied twice then the result is the identity.
According to (6.21), the Möbius transformation corresponding to the binary
rotation about the axis v = l i + m j + n k is
" #
in −m + il
π
[R v ] = .
m + il −in
Now, forgetting about quaternions for a moment, let us redefine 1 to be the identity
matrix, and I, J, K to be the binary rotation matrices about i, j, k, respectively. Thus
" # " # " # " #
1 0 0 i 0 −1 i 0
1= , I= , J= , K= .
0 1 i 0 1 0 0 −i
As a simple check, note that the Möbius transformation corresponding to the
Möbius matrix K is K(z) = −z. Make sure you can see why this is as it should
be.
Now we can state the surprising connection with quaternions: under matrix mul-
tiplication, these binary rotation matrices obey [exercise] exactly the same laws (6.24) and
(6.25) as Hamilton’s I, J, K. It follows that quaternion multiplication is equivalent
to multiplying the corresponding 2 × 2 matrices obtained by replacing Hamilton’s h i
1, I, J, K with the matrices above. Conversely, the general rotation matrix R ψ
v in
(6.21) can be expressed [exercise] as the quaternion
Rψ
v = cos(ψ/2) + V sin(ψ/2), (6.28)
Once again, but more easily than before, we deduce that this is rotation of ψ =
(2π/3) about the axis v = √13 (i + j − k).
Quaternions also yield a very compact formula for the effect of R ψ v on the posi-
tion vector P = X i + Y j + Z k of a point in space. Suppose that R ψ rotates e If
P to P.
v
332 Non-Euclidean Geometry*
R π become the pure quaternions P and P, e thereby proving the quaternion rotation
e
p
e = Rψ
formula discovered by Hamilton and rediscovered by Cayley: P −ψ
v P Rv .
10
NOTE for the 25th Anniversay Edition: There have no doubt been countless other important
applications of this idea over the past 25 years, but I have not researched them for this new edition.
Hyperbolic Geometry 333
angle with C. This justifies the illustrated effect w ÞÑ w ′ ÞÑ w ′′ ÞÑ w ′′′ of the three
b
rotations. Thus the net effect w ÞÑ w ′′′ is a rotation of θ about p
e:
−ψ
R θpbe = R ψ
b ◦ Rp
a b ◦ Ra
θ
b .
on the axis. This provides a good method of quickly sketching a fairly accurate
tractrix.
Returning to [6.18a], let σ represent arc length along the tractrix, with σ = 0
corresponding to the starting position X = R of the object we are dragging. Just as
the object is about to pass through (X, Y), let dX denote the infinitesimal change
in X that occurs while the object moves a distance dσ along the tractrix. From the
similarity of the illustrated triangles, we deduce that
dX X
=− =⇒ X = R e−σ/R . (6.30)
dσ R
The pseudosphere of radius R may now be simultaneously defined and constructed
as the surface obtained by rotating the tractrix about its axis. Remarkably, this sur-
face was investigated as early as 1693 (by Christiaan Huygens), two centuries prior
to its catalytic role in the acceptance of hyperbolic geometry.
[6.18] Newton’s tractrix. [a] Tie a small paperweight to a piece of string of length
R. On a table top, start with the string running along one edge, at right angles to
the other edge, the Y-axis. If you move the free end of the string along the Y-axis
edge of the table, the paperweight will be dragged along the illustrated curve, called
the tractrix, which Newton first investigated in 1676. We see that dX X
dσ = − R and so
−σ/R
X = Re . [b] The tractrix can also be constructed as an orthogonal trajectory
through the family of circles of radius R centred on the Y-axis.
Hyperbolic Geometry 335
[6.19] Geometric proof that the pseudosphere has constant negative Gaussian curva-
ture. [a] The principal directions of maximum and minimum curvature will always occur
in the orthogonal directions in which a surface has local mirror symmetry. [See Needham
(2021).] Clearly, one direction in which the pseudosphere has mirror symmetry is straight
up the pseudosphere, along a meridian tractrix generator. The other principal direction is
the orthogonal sideways direction along a circle of latitude. Thus, K = −1/(r er). [b] From
this figure we can deduce [see text] that the pseudosphere does indeed have constant
negative Gaussian curvature, K = −(1/R2 ).
11
For a full discussion of all these ideas, see Needham (2021).
336 Non-Euclidean Geometry*
By definition, the tractrix in this figure has tangents of constant length R. At the
neighbouring points P and Q, figure [6.19b] illustrates two such tangents, PA and
QB, containing angle •. The corresponding normals PO and QO therefore contain
the same angle •. Note that AC has been drawn perpendicular to QB.
Now let’s watch what happens as Q coalesces with P, which itself remains fixed.
In this limit, O is the centre of the circle of curvature, PQ is an arc of this circle, and
AC is an arc of a circle of radius R centred at P. Thus,12
PQ AC AC R
er ≍ OP and ≍ • ≍ =⇒ ≍ .
OP R PQ er
Next we appeal to the defining property PA = R = QB of the tractrix to deduce
[exercise] that as Q coalesces with P,
BC ≍ PQ.
Finally, using the fact that as Q coalesces with P the triangle ABC is ultimately
similar to the triangle T AP, we deduce that
r AC AC R
≍ ≍ ≍ .
R BC PQ er
Behold!13
1 1
K=− =− 2.
r er R
Once again, “ ≍ ” denotes Newton’s concept of ultimate equality; see the new Preface.
12
13
One of my most treasured possessions is an email from Bill Thurston saying that he liked my
Newtonian proof of this fundamentally important fact!
Hyperbolic Geometry 337
[6.20] Constructing a conformal map of the pseudosphere. [a] Having specialized to the
standard case, R = 1, we have X = e−σ . [b] Let us decide to use the angle x around the
axis of symmetry as the x-coordinate in this map. If we insist that the map be conformal
then all small (ultimately vanishing) distances must be scaled equally, regardless of direc-
tion. Since the sideways movement X dx on the left is divided by X in passing to the map
dy
on the right, all distances must be divided by X, so dσ = X1 = eσ and therefore y = eσ +k.
Choosing k = 0, we find that the distance db s on the pseudosphere is related
p to the dis-
tance ds in the map by the conformal metric formula, db s = ds/y = dx2 + dy2 /y.
generators of the pseudosphere [note that these are clearly geodesics], and the
curves σ = const. are circular cross sections of the pseudosphere [note that these
are clearly not geodesics]. Since the radius of such a circle is the same thing as the
X-coordinate in [6.18a], it follows from (6.30) that
In our map, let us choose the angle x as our horizontal axis, so that the tractrix
generators of the pseudosphere are represented by vertical lines. See [6.20b]. Thus
a point on the pseudosphere with coordinates (x, σ) will be represented in the map
by a point with Cartesian coordinates (x, y), which we will soon think of as the
complex number z = x + iy.
If our map were not required to be special in any way, then we could simply
choose y = y(x, σ) to be an arbitrary function of x and σ. In stark contrast to
338 Non-Euclidean Geometry*
this, our requirement that the map be conformal leaves (virtually) no freedom in
the choice of the y-coordinate. Let’s try to understand this.
Firstly, the tractrix generators x = const. are orthogonal to the circular cross
sections σ = const., so the same must be true of their images in our conformal map.
Thus the image of σ = const. must be represented by a horizontal line y = const.,
and from this we deduce that y = y(σ) must be a function solely of σ.
Secondly, on the pseudosphere consider the arc of the circle σ = const. (of radius
X) connecting the points (x, σ) and (x + dx, σ). By the definition of x, these points
subtend angle dx at the centre of the circle, so their separation on the pseudo-
sphere is X dx, as illustrated. In the map, these two points have the same height
and are separated by distance dx. Thus in passing from the pseudosphere to the
map, this particular line-segment is shrunk by factor X. [We say “shrunk” because
we’re dividing by X, but since X ⩽ 1 this is actually an expansion.] However, since
the map is conformal, an infinitesimal line-segment emanating from (x, σ) in any
direction must be multiplied by the same factor (1/X) = eσ . In other words, the
metric is
db
s = X ds.
Thirdly, consider the uppermost black disc on the pseudosphere shown in
[6.20a]. Think of this disc as infinitesimal, say of diameter ϵ. In the map, it will be
represented by another disc, whose diameter (ϵ/X) may be interpreted more vividly
as the angular width of the original disc as seen by an observer on the pseudo-
sphere’s axis. Now suppose we repeatedly translate the original disc towards the
pseudosphere’s rim, moving it a distance ϵ each time. Figure [6.20a] illustrates the
resulting chain of touching, congruent discs. As the disc moves down the pseudo-
sphere, it recedes from the axis, and its angular width as seen from the axis therefore
diminishes. Thus the image disc in the map appears to gradually shrink as it moves
downward, and the equal distances 8ϵ between the successive black discs certainly
do not appear equal in the map.
Having developed a feel for how the map works, let’s actually calculate the y-
coordinate corresponding to the point (x, σ) on the pseudosphere. From the above
observations (or directly from the requirement that the illustrated triangles be
similar) we deduce that
dy 1
= = eσ =⇒ y = eσ + const.
dσ X
The standard choice of this constant is 0, so that
y = eσ = (1/X).
Thus the entire pseudosphere is represented in the map by the shaded region lying
above the line y = 1 (which itself represents the pseudosphere’s rim), and the
metric associated with the map is
Hyperbolic Geometry 339
p
ds dx2 + dy2
db
s = = . (6.31)
y y
For future use, also note that an infinitesimal rectangle in the map with sides dx
and dy represents a similar infinitesimal rectangle on the pseudosphere with sides
(dx/y) and (dy/y). Thus the apparent area dx dy in the map is related to the true
area dA on the pseudosphere by
dx dy
dA = . (6.32)
y2
Beltrami pointed out that the first of these problems can be resolved as follows.
Imagine the pseudosphere covered by a thin stretchable sheet. To obtain the map in
[6.20b], we cut this sheet along a tractrix generator and unwrap it onto the shaded
340 Non-Euclidean Geometry*
region. Of course to make it lie flat and fit into this rectangular region, the sheet
must be stretched—the metric (6.31) tells us how much stretching must be applied
to each part. But now imagine the sheet as wrapping round and round the pseudo-
sphere infinitely many times14 , like an endless roll of cling film15 . By unwrapping
this infinitely long sheet (stretching as we go) we can now cover the entire region
above y = 1. According to this interpretation, a particle travelling along a horizon-
tal line in the map would correspond to a particle travelling round and round a
circle σ = const. on the pseudosphere, executing one complete revolution for each
movement of 2π along the line.
Now let us explain how the conformal map solves our second problem—the
pseudosphere’s edge. In terms of extrinsic geometry, this edge is an insurmountable
obstacle: we cannot extend the pseudosphere smoothly beyond this edge while
preserving its constant curvature. However, we only care about the pseudosphere’s
intrinsic geometry, and we have seen that if we measure distance using d b s = dsy ,
this is identical to the region y > 1 in [6.21].
Imagining yourself as a tiny two-dimensional being living in [6.21], walking
down a line x = const. is exactly the same thing as walking down a tractrix on
the pseudosphere. Of course on the pseudosphere your walk is rudely interrupted
at some point p b on the rim (σ = 0), corresponding to a point p on the line y = 1.
But in the map this point p is just like any other, and there is absolutely nothing
preventing you from continuing your walk all the way down to the point q on
y = 0.
Why stop at q? The answer is that you will never even get that far, because q
is infinitely far from p! Suppose that you are the illustrated small disc on the line
y = 2, and that I am standing outside your hyperbolic world, watching as you walk
at a steady pace towards y = 0. Of course you remain the same hyperbolic size as
you walk, but to me you appear to shrink. This is made particularly vivid by the
illustrated Euclidean interpretation [exercise] of your hyperbolic size d b s = ds
y :
14
Stillwell (1996) points out that this is probably the very first appearance in mathematics of what
topologists now call a universal cover.
15
For Americans, read “plastic wrap”.
Hyperbolic Geometry 341
[6.21] The pseudosphere has an edge, but the hyperbolic plane is uniform and infinite.
Suppose “You” start at the point on the pseudosphere corresponding to y = 2 and walk
down a tractrix generator at a steady rate of ln 2. Then in the conformal map you reach
y = 1 (the pseudosphere’s rim—the edge of your world!) after just one unit of time. But
if we imagine you living in the map, perceiving distances around you via the hyperbolic
metric, then you can continue your downward journey, arriving at y = (1/2) after two
units of time, y = (1/4) after three units of time, etc. Although you actually stay the same
size, in the map you appear to shrink, for your true size is the angle dbs that you subtend
at the horizon. Thus, viewed from outside your world, you shrink with each successive
unit of time, only halving your distance from the horizon, y = 0; therefore, you will never
reach it. Your world is infinite and uniform: no place or direction seems any different from
any other.
distance from y = 0, and therefore you will never reach it. [An appropriate name
for this phenomenon might be “Zeno’s Revenge”!]
We now possess a concrete model of the hyperbolic plane, namely, the entire
shaded half-plane y > 0 with metric d b s = dsy . The points on the real axis are
infinitely far from ordinary points and are not (strictly speaking) considered part of
the hyperbolic plane. They are called ideal points, or points at infinity. The complete
line y = 0 of points at infinity will be called the horizon16 .
Studying hyperbolic geometry by means of this map is like studying spheri-
cal geometry via a stereographic map, without ever having seen an actual sphere.
This is not as bad as it sounds. After all, by constructing geographical maps
through terrestrial measurements, man developed a good understanding of the
surface of the Earth centuries before venturing into space and gazing down on its
roundness!
Still, it would be nice to have the analogue of a globe instead of a mere atlas.
The pseudosphere only models a portion of the hyperbolic plane, but might there
16
For reasons that will be clear shortly, another name is the circle at infinity.
342 Non-Euclidean Geometry*
exist a different surface that is isometric to the entire hyperbolic plane? Sadly, in
1901 Hilbert proved17 that every pseudospherical surface necessarily has an edge
beyond which it cannot be smoothly extended while preserving its constant nega-
tive curvature. Thus the upper half-plane with metric (6.31) is as good a depiction
of the hyperbolic plane as we are going to get.
However, just as an atlas uses different kinds of maps to represent the surface of
the Earth, so we can and will use different types of maps to represent the hyperbolic
plane. The particular map we have obtained is conventionally called the Poincaré
upper half-plane, but there is also one called the Poincaré disc, and another called the
Klein disc. Poincaré obtained the first two models in 1882, while Klein obtained the
third in 1871.
We cannot let the names of these models pass without comment. Anyone with
even a passing interest in the history of mathematics will know that ideas are fre-
quently (usually?) named after the wrong person. In fact18 , the three models above
were all discovered by Beltrami! As we shall see, Beltrami obtained these three mod-
els (in 1868, 14 years before Poincaré) in a beautifully unified way, from a fourth
model consisting of a map drawn on a hemisphere. And in case you’re wondering,
yes, the hemisphere model is Beltrami’s, too!
In this 25th Anniversary Edition we shall dogedly attempt to set history straight,
as we have previously attempted to do in VDGF, by giving both men equal credit,
and renaming the conformal maps as the Beltrami–Poincaré half-plane and disc
models.
17
Hilbert (1965, Vol. 2, pp. 437–448), with English translation available in Hilbert (1902).
18
See Milnor (1982), Beltrami (1868a), Beltrami (1868b), and Stillwell (2010).
Hyperbolic Geometry 343
line in the map. Let us also define H{z1 , z2 } to be the h-distance (measured using
dbs = dsy ) between z1 and z2 . For example, if dz is infinitesimal, then
|dz|
H{z + dz, z} = .
Im z
Finally, let us define an h-circle of h-radius ρ and h-centre c to be the locus of
points z such that H{z, c} = ρ.
Since tractrix generators of the pseudosphere are clearly geodesic, vertical lines
in the map should also be geodesic, i.e., they should be examples of h-lines. Figure
[6.22a] confirms this directly by showing that
The (unique) shortest route between two vertically separated points is the
(6.34)
vertical line-segment L connecting them.
To see this, compare L with any other route, such as M. Let ds1 be an infinitesimal
segment of L at height y, and let ds2 be the corresponding element of M cut off by
horizontal lines through the ends of ds1 . Since
ds1 ds2
db
s1 = < = db
s 2,
y y
the total hyperbolic length of L is less than M’s. Done. From this we can deduce
that
[6.22] [a] Vertical lines are h-lines (geodesics), because d b s 1 = ds1 /y < ds2 /y = d b
s 2.
[b] Inversion in a semicircle orthogonal to the horizon preserves h-distance. Under
z ÞÑ ez = IK (z), we see that d be
s = de y = ds/y = d b
s/e s . Since distance is preserved in
this particular direction, it must be preserved in all directions.
344 Non-Euclidean Geometry*
[6.23] [a] Geometric proof that semicircles orthogonal to the horizon are h-lines
(geodesics). Construct L as shown, then invert it in any circle K centred at q, obtain-
ing IK (L) = L,e as shown. Since L e has been proven to be the shortest route from a e to
e and since IK preserves h-distance, L must be the shortest route from a to b. [b] If K
b,
is an h-line, then IK = RK is (literal) h-reflection in K. For if e z = IK (z) then all circles
through z and e z cut K orthogonally, and so the illustrated h-line P does, too. The h-lengths
of zm and me z are equal, for they are interchanged by the h-distance-preserving IK , so
e
z = RK (z), as claimed.
route from a to b, then Le would not be the shortest route from a e in violation
e to b,
of (6.34). Done.
Incidentally, note that this construction also enables us (in principle) to calculate
the h-distance between any two points in the hyperbolic plane:
e e
Im a
H{a, b} = H{e a, b} = ln ,
e
Im b
by virtue of (6.35). Later we shall be able to derive a more explicit formula.
The fact that a semicircle orthogonal to the real axis is an h-line strongly suggests
the following re-interpretation of (6.37): hyperbolic plane in the h-line
Inversion in a semicircle K orthogonal to the horizon is a reflection RK of the
(6.38)
hyperbolic plane in the h-line K.
In symbols, RK (z) = IK (z). Before proving this, let’s be clear what we mean by
reflection. Just as we would in Euclidean and spherical geometry, we begin the
construction of RK (z) by drawing the h-line P that passes through z and cuts K
perpendicularly, say at m. Then RK (z) is defined to be the point on P that is the
same h-distance from m as z.
To prove (6.38), consider [6.23b], in which e
z = IK (z). First recall that every circle
through z and e z is automatically orthogonal to K. In particular, the unique h-line
through z and ez must be orthogonal to K, and hence it is the desired “P” of the previ-
ous paragraph. Finally, recall that IK maps P into itself, swapping the segments zm
and ez m. Thus, since IK is a motion, these two h-line segments have equal h-length,
as was to be shown.
346 Non-Euclidean Geometry*
[6.24] The hyperbolic plane satisfies the Hyperbolic Axiom, for we see that there are
infinitely many h-lines [shown dashed] through p that fail to meet the h-line L: they are said
to be ultra-parallel to L. The two h-lines that meet L at the horizon are called asymptotic.
The h-distance of p from L is defined as in Euclidean geometry, as the length of the unique
line segment pq that meets L at right angles.
Conversely, if we are given any two points z and ez, then we may draw the per-
e
pendicular h-bisector K, and RK swaps z and z. Also note that z and its reflection
e
z = RK (z) are the same h-distance from every point k on K, just as in Euclidean and
spherical geometry. This is easily proved: since IK is a motion, and e k = IK (k) = k,
it follows that H{z, k} = H{e z, e
k} = H{e
z, k}.
It is becoming clear that hyperbolic geometry has much in common with
Euclidean geometry. However, now that we know what h-lines look like, [6.24]
shows that hyperbolic geometry really is non-Euclidean: there are infinitely many
h-lines through p [shown dashed] that do not meet the h-line L. Such h-lines are
said to be ultra-parallel to L.
Separating the ultra-parallels from the h-lines that do intersect L, we see that
there are precisely two h-lines that fail to meet L anywhere within the hyperbolic
plane proper, but that do meet it on the horizon. These two h-lines are called
asymptotic19 .
As in Euclidean geometry, the figure makes it clear that there is precisely one
h-line M passing through p that cuts L at right angles (say at q). In fact [exercise]
M may be constructed as the unique h-line through p and RL (p). The existence of
M makes it possible to define the distance of a point p from a line L in the usual
way, namely, as the h-length of the segment pq of M.
Since M and L are orthogonal, RM = IM maps L into itself, swapping the two
ends on the horizon. It follows [exercise] that RM swaps the two asymptotic lines,
19
Another commonly used name is parallel.
Hyperbolic Geometry 347
[6.25] The same geometry as the previous figure, but in the case where L is a vertical
half-line.
and that M bisects the angle at p contained by the asymptotic lines. The angle
between M and either asymptotic line is called the angle of parallelism, and is usually
denoted Π. As one rotates the line M about p, its intersection point on L moves off
towards infinity, and Π tells you how far you can rotate M before it starts missing
L entirely.
Finally, [6.25] merely serves to illustrate the same concepts and terminology as
[6.24], but in the case where the h-line L happens to be represented as a vertical
half-line instead of a semicircle.
and from this they were able to derive many of their other results. We now
give a simple geometric proof of this so-called Bolyai–Lobachevsky Formula. Green-
berg (2008) has called this “one of the most remarkable formulas in the whole of
mathematics”, but for us it will be of only incidental interest.
348 Non-Euclidean Geometry*
First note that it is sufficient to establish the formula using [6.25], rather than
[6.24]. This is because we may transform [6.24] into [6.25] by performing an inver-
sion (i.e., a hyperbolic reflection) in any semicircle centred at one of the ends of
L.
Figure [6.26] reproduces the essential elements of [6.25]. In order to find the h-
length D of the arc pq, let us apply the h-reflection z ÞÑ e z = RC (z), where C is
the illustrated semicircle that is centred at the end c of M, and that passes through
q. This carries the arc pq into the illustrated vertical line-segment peq. By virtue of
e, i.e., the ratio
(6.35), it only remains to find the ratio of the y-coordinates of q and p
of the Euclidean distances [qm] and [e pm].
From the fact that the radius pm is orthogonal to the circle M it follows [exercise]
that the angle pmc equals Π. It then follows [exercise] that the angle ce pm equals
(Π/2), as illustrated. Thus
[qm] [cm]
D = ln = ln = |ln tan(Π/2)| = − ln tan(Π/2),
[e
pm] [e
pm]
where the last equality follows from the fact that tan(Π/2) < 1, because Π is acute.
Thus tan(Π/2) = e−D , as was to be shown.
did in VDGF) calling it instead the Beltrami–Poincaré upper half-plane. What Poincaré
does deserve sole credit for—enormous credit!—is the realization that hyperbolic
geometry is intimately connected with complex analysis. The cornerstone of this
connection is the fact that the (direct) motions of the hyperbolic plane are Möbius
transformations. Let us outline how this comes about.
If L1 and L2 are two h-lines, then the composition
M ≡ RL2 ◦ RL1
of h-reflection in these lines will be a direct motion of the hyperbolic plane. Since
every h-reflection is represented in the map by inversion in a circle, we immediately
deduce that any direct motion of the form M is represented by a (non-loxodromic)
Möbius transformation M(z). Furthermore, later we will show that every direct
motion is of the form M; indeed, we will even give an explicit geometric construc-
tion for decomposing an arbitrary direct motion into two h-reflections. Supposing
this already done, we see that every direct motion is represented as a (non-loxodromic)
Möbius transformation.
Conversely, suppose that M(z) is an arbitrary Möbius transformation that maps
the upper half-plane to itself. Then it follows that M(z) must map the real axis (the
horizon) into itself. But a loxodromic Möbius transformation cannot possess such
an invariant line: its strangely shaped invariant curves were illustrated in [3.32] on
p. 188. Thus M(z) is non-loxodromic, and from (3.49), p. 198, we deduce that M(z)
is the composition of inversion in two circles orthogonal to the real axis. Thus the
most general Möbius transformation of the upper half-plane to itself represents a direct
hyperbolic motion of the type M above.
One way to discover the algebraic form of these Möbius transformations is to
use the formula (3.4), p. 141: inversion in a circle K centred at the point q on the
real axis, and of radius R, is given by
q z + (R2 − q2 )
IK (z) = .
z−q
Composing two such functions, we find [exercise] that a motion of type M corre-
sponds to a Möbius transformation
az + b
M(z) = , where a, b, c, d are real, and (ad − bc) > 0. (6.39)
cz + d
Recall that in Ex. 25, p. 212, you showed that this is the form of the most general
Möbius transformation of the upper half-plane to itself. Thus we have agreement
with the conclusion of the previous paragraph.
So much for the overview—now let’s look in detail at the direct motions M. We
know from [6.24] or [6.25] that there are just three possible configurations for the h-
lines L1 and L2 , and correspondingly M ≡ RL2 ◦ RL2 is one of three fundamentally
different types:
350 Non-Euclidean Geometry*
We can now reap the rewards of all our hard work in Chapter 3, for these three
types of motion are just the three types of non-loxodromic Möbius transformation:
(i) h-rotations are the “elliptic” ones; (ii) limit rotations are the “parabolic” ones;
and (iii) h-translations are the “hyperbolic”20 ones. At this point, you might find
it helpful to reread the discussion of these Möbius transformations at the end of
Chapter 3.
We already understand these Möbius transformations, so it only remains to look
at them afresh, through hyperbolic spectacles. That is, imagine that you belong
to the race of Beltrami–Poincarites—tiny, intelligent, two-dimensional beings who
inhabit the hyperbolic plane. To you and your fellow Beltrami–Poincarites, h-lines
really are straight lines, the real axis really is infinitely far away, etc. What will you
see if the above motions are applied to your world?
Let us begin with h-rotations. Figure [6.27] illustrates the elliptic Möbius trans-
formation—let’s call it Rϕ a —that arises in the case where the h-lines intersect at a,
and the angle from L1 to L2 is (ϕ/2). [We have chosen to illustrate ϕ = (π/3).] Thus
Rϕ iϕ
a has fixed points a and a, and the multiplier associated with a is m = e . As in
[6.27] A Hyperbolic Rotation, results from reflecting across intersecting h-lines, L1 and L2 .
20
Try not to be confused by this unrelated use of the word “hyperbolic”.
Hyperbolic Geometry 351
Chapter 3, each shaded “rectangle” is mapped by Rϕ a to the next one in the direction
of the arrows—some of these regions have been filled with black to emphasize this.
Consider how all this looks to you and your fellow Beltrami–Poincarites. For
example, you see each black “rectangle” as being exactly the same shape and size
as every other. To understand Rϕ a better, we begin by noting that (in terms of the
map) its effect on an infinitesimal neighbourhood of a is just a Euclidean rotation
of ϕ about a. But since the map is conformal, this implies that a Poincarite standing
at a will also see his immediate neighbourhood undergoing a rotation of ϕ.
More remarkably, however, the Poincarite at a will see the entire hyperbolic
plane undergoing a perfect rotation of ϕ. Every h-line segment ap he constructs
emanating from a is transformed by z ÞÑ e z = Rϕa (z) into another h-line segment
aep of equal length, making angle ϕ with the original. If the Poincarite gradually
increases ϕ from 0 to 2π, then he sees pe tracing out an h-circle centred at a, while in
the map we see p e miraculously tracing out a Euclidean circle! Thus the illustrated
Euclidean circles orthogonal to the h-lines through a are all genuine hyperbolic cir-
cles, and a is their common h-centre. Let us record this remarkable result, adding
a detail that is not too hard to prove [exercise]:
As a stepping stone to the limit rotations, [6.28] introduces a new type of curve
in the hyperbolic plane. On a line L in Euclidean geometry, let p be a fixed point,
let a be a moveable point, and let C be the circle centred at a that passes through p.
If we let a recede to infinity along L, then the limiting form of C is a line (through p
and perpendicular to L). Figure [6.28a] shows that it’s a different story in the hyper-
bolic plane. As a recedes towards the infinitely remote point A on the real axis, the
limiting form of C is a (Euclidean) circle that touches the real axis at A. This is nei-
ther an ordinary h-circle, nor an h-line: it is a new type of curve called a horocycle.
Figure [6.28b] shows that horizontal (Euclidean) lines are also horocycles. Note that
if K is any circle centred at A then the h-reflection RK = IK transforms [6.28a] into
[6.28b]. Thus the Beltrami–Poincarites cannot distinguish between these two types
of horocycle.
Now consider [6.29], which illustrates the parabolic Möbius transformation that
results from h-reflection in h-lines L1 and L2 that are asymptotic at A. Referring
to [6.27] and [6.28], you can now understand why this is called a limit rotation: it
may be viewed as the limit of the h-rotation Rϕ a as a tends to the point A on the
horizon. Note some of the interesting features of this picture: the invariant curves
are horocycles touching at A; each such horocycle is orthogonal to every h-line that
352 Non-Euclidean Geometry*
ends at A; and any two such horocycles cut off the same h-length on every h-line
that ends at A.
In terms of the map, the simplest limit rotation occurs when the asymptotic h-
lines L1 and L2 are represented as vertical Euclidean half-lines, say separated by
Euclidean distance (α/2). In this case, M = (RL2 ◦ RL1 ) is represented in the map
by the composition of two Euclidean reflections in parallel lines. Thus M is just a
Euclidean translation z ÞÑ (z + α) of the upper half-plane, and the invariant curves
are horizontal lines, which are again horocycles, but now of the form shown in
[6.28b]. Note that this Euclidean translation is not an h-translation. This is particu-
larly clear if we visualize the effect of M on the pseudosphere, where it becomes a
rotation through angle α about the pseudosphere’s axis.
Figure [6.30] illustrates the third and final type of motion, the h-translation
(hyperbolic Möbius transformation) resulting from h-reflection in two ultra-
parallel h-lines. First note that there is precisely one h-line L that is orthogonal
to both L1 and L2 . Unlike a Euclidean translation, this h-line L is the only h-line
that is mapped into itself; it is called the axis of the h-translation. Despite this dif-
ference, the name “h-translation” is appropriate, for every point on the h-line L
is moved the same h-distance (say δ) along L. If we assume that the axis L has
a direction assigned to it, then we may unambiguously denote this h-translation
by TLδ .
In Euclidean geometry, the invariant curves of a translation are the parallel lines
in the direction of the translation. However, [6.30] shows that the invariant curves
of TLδ are not h-lines, but rather arcs of Euclidean circles connecting the ends e1 and
Hyperbolic Geometry 353
[6.29] A Hyperbolic Limit Rotation results from reflecting across asymptotic h-lines, L1
and L2 . This transformation has no analogue in Euclidean geometry, but it can be thought
of as the limit of the h-rotation Rϕ
a as a tends to the point A on the horizon. Note that
the invariant curves are horocycles.
e2 of L. These are called the equidistant curves of L, because every point on such a
curve is the same h-distance from the h-line L. Make sure you can see this.
In terms of the map, the simplest h-translation occurs when the ultra-parallel h-
lines L1 and L2 are represented by concentric Euclidean semicircles, say centred at
the origin for convenience. In this case, the two h-reflections (i.e., inversions) yield
a central dilation z ÞÑ kz, where k is the real expansion factor. The axis of this h-
translation is the vertical line through the origin (the y-axis), and the equidistant
curves are all other (Euclidean) lines through the origin (cf. [6.20] and [6.21]). Note
that this Euclidean expansion is a similarity transformation of the map, but it is not
a similarity transformation of the hyperbolic plane—there are none!
Having completed our survey of these three types of direct motion, it’s impor-
tant to note that they not only look very different in terms of their effect on
the map, but they also have unique fingerprints in terms of the intrinsic hyper-
bolic geometry. To put this another way, Beltrami–Poincarites can tell these motions
apart. For example, of the three, only h-rotations have invariant h-circles, and only
h-translations have an invariant h-line.
Euclidean geometry, this will be established if we can show that the location of
a point p is uniquely determined by its h-distances from any three non-collinear
points a, b, c. Consider [6.31a], in which we have supposed (for simplicity’s sake
only) that the h-line L through a and b is represented by a vertical line in the map.
Through the point p, draw h-circles centred at a, b, and c. Since c does not lie on L
(by assumption), we see that p is the only point at which the three circles intersect.
Done.
Now suppose that an arbitrary motion carries two points a and b to the points
a and b ′ in [6.31b]. By the above result, the motion will be determined once we
′
know the image of any third point p not on the line L through a and b. Draw-
ing the illustrated h-circles with h-centres a ′ and b ′ and with h-radii H{a, p} and
H{b, p}, we see that the two intersection points p ′ and p e are the only possible
images for p. Furthermore, since the h-line L ′ through a ′ and b ′ is necessar-
ily orthogonal to the h-circles centred at those points, we also see that p ′ and
e are symmetric with respect to L ′ , i.e., p
p e = IL ′ (p ′ ) = RL ′ (p ′ ). Thus we have
shown that
There is exactly one direct motion M (and exactly one opposite motion
e that maps a given h-line segment ab to another h-line segment
M)
e = (RL ′ ◦ M), where L ′ is (6.40)
a ′ b ′ of equal h-length. Furthermore, M
′ ′
the h-line through a and b .
Hyperbolic Geometry 355
M = Rθw ◦ TLδ .
Implicitly, this formula decomposes M into four h-reflections, because TLδ and
Rθw can both be decomposed into two h-reflections. However, [6.32] illustrates that
we can always arrange for two of the four h-reflections to cancel. Defining m to
356 Non-Euclidean Geometry*
[6.32] Decomposing an arbitrary direct motion M into two h-reflections. Imagine two
points close together, z and z + dz, being mapped to their images w and w + dw. First
carry z to w with the unique h-translation TLδ (where δ = H{z, w}) along the h-line from
z to w. This carries dz conformally to de
z, still making the same angle with L. Now rotate
dez about w by θ to obtain dw. Thus, M = Rθw ◦ TLδ . This implicitly decomposes M
into four reflections, but we can always arrange for the middle two to cancel. Here,
M = (RC ◦ RB ) ◦ (RB ◦ RA ) = RC ◦ RA = Rϕ a , but clearly this construction may just as
easily yield an A and a C that are asymptotic or ultra-parallel, in which case M is a limit
rotation or an h-translation.
be the h-midpoint of the h-line segment zw, draw h-lines A and B orthogonal to
L and passing through m and w, respectively. Then TLδ = (RB ◦ RA ). If we now
draw an h-line C through w making angle (θ/2) with B, then Rθw = (RC ◦ RB ).
Thus, as we set out to show, every direct motion can be decomposed into two
h-reflections:
M = (RC ◦ RB ) ◦ (RB ◦ RA ) = RC ◦ RA .
In the illustrated example, it so happens that the h-lines A and C intersect, and
so the motion is an h-rotation: M = Rϕ a , where a is the intersection of A and C, and
(ϕ/2) is the angle between them. However, it is clear that this construction may
just as easily yield an A and a C that are asymptotic or ultra-parallel, in which case
M is a limit rotation or an h-translation.
Summarizing what we have shown, and recalling (6.39),
az + b
M(z) = , where a, b, c, d are real, and (ad − bc) > 0.
cz + d
Finally, returning to [6.32] and appealing to (6.40), the unique opposite motion
e carrying z to w and dz to dw is given by three h-reflections:
M
e = R L ′ ◦ RC ◦ RA .
M
Here L ′ is the illustrated h-line passing through w and (w + dw), i.e., passing
through w in the direction dw. This decomposition does not, however, yield the
e for that, and for the formula describing the
simplest geometric interpretation of M;
general opposite motion, see Ex. 24.
As we pointed out in the Introduction, this says (amongst other things) that the
angles of ∆ always add up to less than π, and that no matter how large we make
∆, its area can never exceed π. Referring to the differential geometry result (6.6),
we also see that in establishing this formula we will have provided an intrinsic21
proof of the fact that the hyperbolic plane is a surface of constant negative curvature
K = −1.
We have already remarked that Christiaan Huygens investigated the pseu-
dosphere as early as 1693, and to get acquainted with hyperbolic area we will
now confirm one of his surprising results: the pseudosphere has finite area. In
the upper half-plane the pseudosphere is represented by the shaded region
{0 ⩽ x < 2π, y ⩾ 1} shown in [6.20], and (6.32) implies that this region of infinite
Euclidean area does indeed have finite hyperbolic area:
ZZ Z 2π Z ∞ Z 2π Z∞
dx dy dy
A(pseudosphere) = dA = 2
= dx 2
= 2π,
x=0 y=1 y x=0 y=1 y
as Huygens discovered.
Figure [6.33a] illustrates a triangle on the pseudosphere. If the uppermost ver-
tex moves up the pseudosphere indefinitely, then the angle at that vertex tends to
zero, and the edges meeting at that vertex tend to asymptotic lines, namely, trac-
trix generators meeting at infinity. Such a limiting triangle, two of whose edges
are asymptotic, is called an asymptotic triangle. In order to establish (6.41) for ordi-
nary triangles, we first establish it for asymptotic triangles. Figure [6.33b] illustrates
21
Recall that earlier we used the pseudosphere to give an extrinsic proof.
358 Non-Euclidean Geometry*
[6.33] [a] An asymptotic triangle results when the top vertex moves up the pseudosphere
indefinitely. [b] The angular excess of an asymptotic triangle. Suppose the finite edge of
the asymptotic triangle T is an arc of the unit circle. A simple calculation [see text] then
shows that A(T ) = π − α − β. Taking the third angle of T to be zero, this indeed accords
with E(T ) = (−1)A(T ).
such a triangle T in the upper half-plane, the asymptotic tractrix generators becom-
ing vertical half-lines. By Huygens’ result, T clearly has a finite area A(T ), and
because the asymptotic edges meet at angle zero, the result we wish to establish
is A(T ) = (π − α − β).
To simplify the derivation of this result, [6.33b] supposes that the finite edge of
T is an arc of the unit circle. This does not involve any loss of generality, because
an arc of a circle of radius r centred at x = X may be transformed into an arc of the
unit circle by applying the limit rotation z ÞÑ (z − X), followed by the h-translation
z ÞÑ (z/r). From [6.33b] we now deduce that
Z cos β Z ∞ Z cos β
dy dx
A(T ) = √ 2
dx = √ ,
x=cos(π−α) y= 1−x2 y x=cos(π−α) 1 − x2
[6.34] The angular excess of a general hyperbolic triangle. [a] A general hyperbolic trian-
gle of area A can be rotated so that one edge becomes vertical. [b] But now A is clearly
the difference of the areas of two asymptotic triangles: one with angles α and (β + θ); the
other with angles (π − γ) and θ. It follows immediately that E(T ) = (−1)A(T ), thereby
confirming that the hyperbolic plane has constant negative Gaussian curvature K = −1.
A = [π − α − (β + θ)] − [π − (π − γ) − θ]
= π−α−β−γ
= −E.
z ÞÑ e
z = IK (z),
where K is the illustrated circle centred at −i and passing through ±1. In order for
this disc to represent the hyperbolic plane, its metric must be inherited from the
upper half-plane. That is, we must define the h-separation H{e e of two points in
a, b}
the disc to be the h-separation H{a, b} of their preimages in the upper half-plane.
Note that this implies [exercise] that the h-lines of the disc are precisely the images
of h-lines in the upper half-plane.
360 Non-Euclidean Geometry*
Before moving on, try staring at [6.35a] until the following details become clear:
(i) ±1 remain fixed and i is mapped to 0; (ii) the entire shaded part of the upper
half-plane is mapped to the shaded bottom half of the unit disc; (iii) the remaining
part of the upper half-plane (i.e., the top half of the unit disc) is mapped into itself;
(iv) h-lines in the disc are the images of h-lines in the upper half-plane, and these are
arcs of circles orthogonal to the unit circle; (v) the entire horizon of the hyperbolic
plane is represented by the unit circle, with the common point at infinity of vertical
h-lines in the upper half-plane being represented by −i.
At this point we have obtained a map of the hyperbolic plane within the unit
disc. However, since IK (z) is anticonformal, so is our map: angles in the upper
half-plane are currently represented by equal but opposite angles in the disc. If we
now apply z ÞÑ z, which reflects the disc across the real axis into itself, then angles
are reversed a second time, and we obtain the conformal Beltrami–Poincaré disc.
The net transformation from the Beltrami–Poincaré upper half-plane to the
Beltrami–Poincaré disc is thus the composition of z ÞÑ IK (z) and z ÞÑ z, and this is
a Möbius transformation, say D(z). Since D(z) maps i to 0 and −i to ∞, it is clear
that D(z) must be proportional to (z − i)/(z + i). Finally, recalling that a Möbius
transformation is uniquely determined by its effect on three points, and noting that
±1 remain fixed, we deduce [exercise] that
Hyperbolic Geometry 361
iz + 1
D(z) = . (6.42)
z+i
Alternatively, this may be derived by brute force [exercise] using the formula for
inversion, (3.4), p. 141.
Since D(z) preserves angles and circles, it is easy to transfer the basic types of
curve in the hyperbolic plane from the Beltrami–Poincaré upper half-plane to the
Beltrami–Poincaré disc. Figure [6.35b] illustrates that h-lines are represented by
arcs of circles orthogonal to the unit circle (such as L, A, U), including diameters
such as I. Incidentally, since the horizon is now represented by the unit circle, you
can understand why the horizon is also called the circle at infinity.
The terminology for h-lines is the same as before: I intersects L, A is asymptotic
to L, U is ultra-parallel to L, and a Euclidean circular arc E connecting the ends of
L is an equidistant curve of L. It is also easy to see that a Euclidean circle C lying
strictly inside the unit disc represents an h-circle, though its h-centre a does not
generally coincide with its Euclidean centre. Finally, the horocycles in [6.28a] and
[6.28b] are represented in the Beltrami–Poincaré disc by circles such as H that touch
the unit circle.
Now let us find the metric in the Beltrami–Poincaré disc. Ex. 19 shows how this
may be done by brute calculation, but the following geometric approach22 is much
more enlightening and powerful. First, [6.36a] recalls the earlier observation (6.33):
if ds is the infinitesimal Euclidean length of a horizontal line-element emanating
from z, then the angle between L and E is its hyperbolic length d b s = [ds/ Im(z)].
Note that in purely hyperbolic terms, L is an h-line orthogonal to ds, and E is an
equidistant curve of L. If we apply an h-rotation Rϕ z then L is carried into another
h-line L , and E is carried into an equidistant curve E ′ of L ′ , and the angle between
′
L ′ and E ′ is the same as before. Thus we have the following general construction:
Through one end of ds, draw the h-line l orthogonal to ds, and through the
other end of ds draw the equidistant curve e. Then the h-length db
s of ds is (6.43)
the angle of intersection (on the horizon) of l and e.
Now the beauty of interpreting db s as an angle in this way is that the Möbius
transformation D to the Beltrami–Poincaré disc is conformal, and so the above
construction of db s is valid there too!
Figure [6.36b] illustrates an infinitesimal disc of Euclidean radius ds centred at
z = r eiθ in the Beltrami–Poincaré disc. Because the map is conformal, the h-length
dbs of ds is independent of the direction of ds, so we may simplify the construction
(6.43) by choosing ds orthogonal to the diameter l through z. The equidistant curve
e is then the illustrated arc of a Euclidean circle through the ends of l.
22
We merely rediscovered this angular interpretation of hyperbolic distance, which we believe
originates with Thurston (1997). However, our explanation (and our applications) of the idea differ
somewhat from Thurston’s.
362 Non-Euclidean Geometry*
[6.36] Geometric derivation of the disc metric. [a] If ds is the infinitesimal Euclidean
length of a horizontal line-element emanating from z, then the angle between the orthog-
onal h-line L and the equidistant curve E is its hyperbolic length d b s = [ds/ Im(z)]. But
performing a rotation about z we see that d b s may equally be measured as the angle
between the h-line L ′ and its equidistant curve E ′ . [b] Since the mapping from the half-
plane to the disc is conformal, this angular interpretation of hyperbolic distance applies
here, too! Choose ds orthogonal to the diameter l through z. The equidistant curve e
is then the illustrated arc of a Euclidean circle of radius ρ through the ends of l, and
ρdb s = 1. [c] AB = A ′ B ′ . [d] So, 2ρ ds = (1 − r)(1 + r) = 1 − |z|2 , and the hyperbolic
disc metric is therefore d b 2
s = 1−|z| 2 ds.
ρdb
s = 1.
Next recall (or prove) the familiar property of circles illustrated in [6.36c], namely,
that all chords passing through a fixed interior point are divided into two parts
whose lengths have constant product: AB = A ′ B ′ . Applying this result to the copy
of [6.36b] shown in [6.36d], we obtain [exercise]
2ρ ds = (1 − r)(1 + r) = 1 − |z|2 .
Hyperbolic Geometry 363
where L1 and L2 are h-lines, namely, arcs of circles orthogonal to the unit circle. As
in the upper half-plane, every direct motion is therefore a non-loxodromic Möbius
transformation. We already know that there are just three hyperbolically distin-
guishable types of direct motion, and the distinction between them in terms of L1
and L2 is the same as before: we get an h-rotation when they intersect, a limit rota-
tion when they are asymptotic, and an h-translation when they are ultra-parallel.
364 Non-Euclidean Geometry*
[6.37] A Hyperbolic Rotation results from reflecting across intersecting h-lines, L1 and L2 .
[a] The typical case. [b] If the rotation is about the centre, then it reduces to an ordinary
Euclidean rotation.
We will discuss the formula for these Möbius transformations in a moment, but
first let’s draw pictures of them.
Figure [6.37a] shows a typical h-rotation; note the appearance of h-circles with a
common h-centre. Figure [6.37b] illustrates the pleasant fact that if L1 and L2 inter-
sect at the origin (in which case they are Euclidean diameters) then the resulting
h-rotation manifests itself as a Euclidean rotation.
In this connection, we offer a word of warning. As Euclidean beings, we suffer
from an almost overwhelming temptation to regard the centre of the Beltrami–
Poincaré disc as being special in some way. One must therefore constantly remind
oneself that to the Beltrami–Poincarites who inhabit the disc, every point is indis-
tinguishable from every other point. In particular, the Beltrami–Poincarites do not
see any difference between [6.37a] and [6.37b].
Figure [6.38a] illustrates a typical limit rotation generated by an L1 and an L2
that are asymptotic at a point A on the horizon. Once again note that the invariant
curves are horocycles touching at A, and that these are orthogonal to the family of
h-lines that are asymptotic at A.
Finally, [6.38b] illustrates a typical h-translation. Once again, note that there is
precisely one invariant h-line [shown in bold], and that the invariant equidistant
curves are arcs of circles through the ends of this axis.
From our work in the upper half-plane we know that the three types of motion
pictured above are the only direct motions of the Beltrami–Poincaré disc, and we
now turn to the formula that describes them. We know that every direct motion
is a Möbius transformation that maps the unit disc into itself, and at the end of
Hyperbolic Geometry 365
[6.38] [a] A Limit Rotation results from reflecting across asymptotic h-lines,
L1 and L2 . [b] A Translation results from reflecting across ultra-parallel h-lines,
L1 and L2 .
An immediate benefit of this insight is that we can now easily find the formula
for the h-separation of any two points, a and z. The h-rotation Ma brings a to the
origin, and we already know the formula (6.45) for the h-distance of a point from
there:
366 Non-Euclidean Geometry*
1 + |Ma (z)|
H{a, z} = H{Ma (a), Ma (z)} = H{0, Ma (z)} = ln ,
1 − |Ma (z)|
and so
|a z − 1| + |z − a|
H{a, z} = ln . (6.46)
|a z − 1| − |z − a|
Figure [6.39b] shows how to compose these h-rotations, using the same idea as
was used in both Euclidean and spherical geometry. The h-rotation Rϕ 0 is the com-
position of h-reflections in any two h-lines through 0 (diameters) containing angle
(ϕ/2). Thus, choosing the first h-line to be B, and calling the second h-line C, we
deduce that
a = (RC ◦ RB ) ◦ (RB ◦ RA ) = RC ◦ RA .
Mϕ
ds
db
s = . (6.47)
Z
This formula only describes the h-separation of points on the hemisphere, but
there is nothing preventing us from using it to define the h-separation of any two
infinitesimally separated points in the three-dimensional region Z > 0. This region
lying above C, with h-distance defined by (6.47), is called the half-space model of
three-dimensional hyperbolic space, denoted H3 . Without going into detail, it is clear
from (6.47) that the points of C are infinitely h-distant from points that lie strictly
above C. Thus C represents the two-dimensional horizon or sphere at infinity of
hyperbolic space.
At the moment it is a mere tautology that the geometry induced on the hemi-
sphere by (6.47) is that of a hyperbolic plane. To begin to see that there is real meat
on this idea, let us consider some simple motions of hyperbolic space. Clearly d b s
is unaltered by a translation parallel to C, so this is a motion. It is also clear that
dbs is unaltered by a dilation (X, Y, Z) ÞÑ (kX, kY, kZ) centred at the origin. More
generally, a dilation centred at any point of C will preserve d b
s , so this too is a motion.
By applying these two types of motion to the origin-centred unit hemisphere
that we have been studying, we see that
In the half-space model, H3 , every hemisphere orthogonal to C is a
(6.48)
hyperbolic plane, H2 .
370 Non-Euclidean Geometry*
In Euclidean geometry the intersection of two planes is a line, and this suggests
that an h-line should be the intersection of two hyperbolic planes. Thus we antici-
pate that every semicircle orthogonal to C is an h-line, for every such semicircle is the
intersection of two hemispheres orthogonal to C. Note that this agrees with what
we already know: the h-lines of the hemisphere model are semicircles orthogonal to
C. Figure [6.41] attempts to make this all much more vivid,23 showing the vertical
H2 walls filled with circles that are all the same hyperbolic size; compare this to
[6.20], page 337.
Let us return to two-dimensional geometry for a moment. Figure [6.42] illus-
trates how Beltrami obtained the upper half-plane model from his hemisphere
model. From a point q on the rim of the hemisphere, we stereographically project
onto the tangent plane at the point antipodal to q—actually, any plane tangent to
this one would do equally well. Since this preserves circles and angles, we see that a
typical h-line of the hemisphere is mapped to a semicircle orthogonal to the bottom
edge of the half-plane, while an h-line passing through q is mapped to a vertical
line.
23
This figure was not in the original edition of VCA. I drew it about 15 years later, for VDGF, where
it appears as figure [6.6].
Hyperbolic Geometry 371
[6.42] How Beltrami discovered the upper half-plane model. From a point q on the rim
of the hemisphere, stereographically project onto the tangent plane at the point antipodal
to q. Since this preserves circles and angles, we see that a typical h-line of the hemisphere
is mapped to a semicircle orthogonal to the bottom edge of the half-plane, while an h-line
passing through q is mapped to a vertical line. The angular interpretation of hyperbolic
distance given in [6.36] applies here, too. We therefore recover the familiar hyperbolic
metric in the upper half-plane: d b s = (ds/Z).
Well, since these are the h-lines, it certainly looks like we have obtained the
Beltrami–Poincaré upper half-plane, but to make sure, let’s check that its metric
is really given by (6.31). Since stereographic projection is conformal, we may yet
again use the construction (6.43). Choosing l to be the image of an h-line through
q, the figure immediately reveals that the metric is d b s = (ds/Z). Apart from a
change of notation, this is indeed the same as (6.31).
We have thus returned to the half-plane that began our journey, but we have
returned wiser than when we left. Looking at (6.47) we now recognize this half-
plane orthogonal to C as a hyperbolic plane within hyperbolic space. This reveals
the true role of the stereographic projection in [6.42].
We know that stereographic projection from q is just the restriction to the hemi-
sphere of inversion IK in a sphere K centred at q. Using the same argument (figure
[6.22b]) as in the plane, we see that IK preserves the metric (6.47), so it is a motion of
hyperbolic space, carrying h-lines into h-lines and carrying h-planes into h-planes.
Furthermore, (6.48) tells us that K is a hyperbolic plane in this hyperbolic space,
and we therefore suspect that IK is reflection in this h-plane. This can be confirmed
[exercise] by generalizing the argument in [6.23b]. Thus we have the following
generalization of (6.38):
Inversion in a hemisphere K orthogonal to the horizon is reflection RK
of hyperbolic space in the h-plane K.
372 Non-Euclidean Geometry*
It is beyond the scope of this book to explore the motions of hyperbolic space24 .
However, let us at least describe one particularly beautiful result.
Just as an arbitrary direct motion of an h-plane is the composition of two h-
reflections in h-lines within it, so an arbitrary direct motion of hyperbolic space
is the composition of four reflections in h-planes within it. Thus, in the half-space
model with horizon C, such a motion is the composition of four inversions in
spheres centred on C. If we restrict attention to the points of C then inversion in
such a sphere K is equivalent to two-dimensional inversion of C in the equato-
rial circle in which K intersects C. Conversely, inversion of C in a circle k extends
uniquely to an inversion of space: simply construct the sphere with equator k.
Finally, then, every direct motion of hyperbolic space can be uniquely repre-
sented in terms of C (the horizon) as the composition of inversion in four circles,
and this is none other than the most general Möbius transformation,
az + b
z ÞÑ M(z) = ,
cz + d
of the complex plane! Poincaré discovered this wonderful fact in 1883; see Poincaré
(1985).
We have seen that the direct motions of the hyperbolic plane, the Euclidean plane
and the sphere are subgroups of this group of general Möbius transformations. As
we shall now see, this fact has a remarkable geometrical explanation.
Hilbert’s result on surfaces of constant negative curvature shows that three-
dimensional Euclidean geometry cannot accommodate a model of the hyperbolic
plane. Amazingly, however, three-dimensional hyperbolic space does contain sur-
faces whose intrinsic geometry is Euclidean! In fact these surfaces are the horospheres
that generalize the horocycles. Analogously to [6.28], horospheres are Euclidean
spheres that touch C, as well as planes Z = const. that are parallel to C.
Vertical planes orthogonal to C look flat in our model of hyperbolic space, but
in reality they are intrinsically curved hyperbolic planes. However, a horosphere
Z = const. not only looks flat, it really is flat. For its metric, inherited from the
metric (6.47) of the surrounding space, is just
db
s = (constant) ds,
24
Thurston (1997) is an excellent source of information on these motions.
Hyperbolic Geometry 373
6.4 Exercises
e = Rψ
p ÞÑ p v p,
where R ψ
v is being thought of as a 2 × 2 matrix.
(i) Show that in homogeneous coordinates, (3.21), p. 166, becomes
2 p1 p2 |p1 |2 − |p2 |2
X + iY = and Z = .
|p1 |2 + |p2 |2 |p1 |2 + |p2 |2
(ii) To simplify this, recall that all multiples of p describe the same point p in
√
C. We can therefore choose the “length” of p to be 2:
⟨p , p⟩ ≡ |p1 |2 + |p2 |2 = 2.
With this choice, show that the above equations can be written as
" # " # " #
1 + Z X + iY p1 p1 p1 p2 p1
= = [ p1 p2 ] = p p∗ .
X − iY 1 − Z p2 p1 p2 p2 p2
(iii) Verify that
" #
1+Z X + iY
= 1 − iP.
X − iY 1−Z
(iv) Deduce that
h i∗
e = R ψ (1 − i P) R ψ = 1 − i R ψ P R −ψ ,
1 − iP v v v v
from which (6.29) follows immediately.
9 (i) Figure [6.40a] gave a two-step process for carrying a point z in the
Beltrami–Poincaré disc to the corresponding point z ′ in the Klein model.
Explain why the net mapping z ÞÑ z ′ of the disc to itself is the one shown
in figure [a] below, where C is an arbitrary circle passing through z and
orthogonal to the unit circle U.
(ii) Figure [b] is a vertical cross section of [6.40a] through z and z ′ . Deduce that
|z ′ | 1 a 2
= and = .
a b |z| b
By multiplying these two equations, deduce that z ′ = 1+|z|
2z
2.
14 (i) The figure above superimposes the stereographic and projective [see
p. 323] images of a great circle on the unit sphere. Let zs and zp be the
stereographic and projective images of the point whose spherical-polar
coordinates are (ϕ, θ). Referring to (3.22), p. 167, zs = cot(ϕ/2) eiθ . Show
that zp = [− tan ϕ] eiθ , and deduce that
2zs
zp = .
1 − |zs |2
Compare this with Ex. 9!
(ii) Sketch the curves on the hemisphere that are centrally projected to the cir-
cles |zp | = const. and to the rays arg zp = const. Although angles are
generally distorted in the projective model, observe that these circles and
rays really are orthogonal, as they appear to be.
(iii) Now let the sphere have radius R, and write zp = r eiθ for the projective image
of the point (ϕ, θ). Thus r = −R tan ϕ. Show that if zp moves a distance
dr along the ray θ = const., then the corresponding point on the sphere
moves a distance d b s r given by
dr
db
sr = .
1 + (r/R)2
(iv) Likewise, show that the separation d b s θ of the points on the sphere corre-
sponding to the points (r, θ) and (r, θ + dθ) in the map is given by
r dθ
dbsθ = p .
1 + (r/R)2
(v) Deduce that the spherical separation d b s corresponding to the points (r, θ)
and (r + dr, θ + dθ) is given by
dr2 r2 dθ2
db
s2= + .
(1 + (r/R)2 )2 1 + (r/R)2
(vi) Here is a crazy idea, essentially due to Lambert; see Penrose (2005, § 2.6).
Perhaps we can get a surface of constant negative curvature K = −(1/R2 ) =
+1/(iR)2 by allowing the radius R of the sphere to take on the imaginary
value iR. Verify that if we substitute R = i into the above formula, then it
becomes the Beltrami metric (6.50) of the hyperbolic plane! [To make true
sense of this idea, one must turn to Einstein’s relativity theory; see Thurston
(1997) and especially Penrose (2005).]
15 Take a stack of ten sheets of paper and staple them together, placing staples
along three of the edges. Use a pair of compasses to draw the largest circle
that will fit comfortably inside the top sheet. Pierce through all ten sheets in
the centre of the circle. With heavy scissors, cut along the circle to obtain ten
380 Non-Euclidean Geometry*
identical discs, say of radius R. Repeat this whole process to double the number
of discs to 20.
(i) Cut a narrow sector out of the first disc, and tape the edges together to
form a shallow cone. Repeat this process with the remaining discs, steadily
increasing the angle of the sector each time, so that the cones get sharper
and taller. Ensure that by the end of the process you are making very
narrow cones, using only a quarter disc or less.
(ii) Stack these cones in the order you made them. Explain how it is that you
have created a model of a portion of a pseudosphere of radius R. Create weird
new (extrinsically asymmetric) surfaces of constant negative curvature by
holding the tip of your structure and moving it from side to side!
(iii) Use the same idea to create a disc-like piece of “hyperbolic paper”, such as
you would get if you could simply cut out a disc from your pseudosphere.
Press it against the pseudosphere and verify that you can freely move it
about and rotate it on the surface.
16 By holding a fairly short piece of string against the surface of the toy pseudo-
sphere of the previous exercise, draw a segment of a typical geodesic. Extend
this segment in both directions, one string-length at a time. Note the surpris-
ing way the geodesic only spirals a finite distance up the pseudosphere before
spiralling down again.
(i) Use the upper half-plane to verify mathematically that the tractrix genera-
tors are the only geodesics that extend all the way up to the top.
(ii) Let L be a typical geodesic, and let α be the angle between L and the trac-
trix generator at the point where L hits the rim σ = 0. Show that the
maximum distance σmax that L travels up the pseudosphere is given by
σmax = | ln sin α|.
17 Suppose we have a conformal map of a surface in the xy-plane, with the metric
given by (6.15):
p
db
s = Λ ds = Λ(x, y) dx2 + dy2 .
An elegant result from differential geometry [see Needham (2021, §4.4)] states
that the Gaussian curvature at any point on the surface is given by
1
K=− ∆(ln Λ),
Λ2
where ∆ ≡ (Bx2 + By2 ) is the Laplacian. Try this out on the metric (6.16) of the
stereographic map of the sphere, and on the metrics (6.31) and (6.44) of the
half-plane and disc models of the hyperbolic plane.
Exercises 381
18 Use the Beltrami–Poincaré disc to rederive the formula tan(Π/2) = e−D for the
angle of parallelism. [Hint: Let one of the h-lines be a diameter.]
19 To derive the metric (6.44), consider the mapping (6.42) z ÞÑ w = D(z) from
the upper half-plane to the Beltrami–Poincaré disc. An infinitesimal vector dz
emanating from z is amplitwisted to an infinitesimal vector dw = D ′ (z) dz
emanating from w, and (by definition) the h-length d b
s of dw is the h-length of
dz. Verify (6.44) by showing that
2 |dw| |dz|
= = db
s.
1 − |w|2 Im z
20 Consider the mapping z ÞÑ w = Mϕ a (z) of the Beltrami–Poincaré disc to itself.
Use the calculational approach of the previous exercise to show that z ÞÑ w is
a hyperbolic motion, i.e., it preserves the metric:
2 |dw| 2 |dz|
= db
s = .
1 − |w| 2 1 − |z|2
21 In the upper half-plane, the h-rotation Rϕ
i through angle ϕ about the point i is
given by the following Möbius transformation:
cz + s
Rϕi (z) = , where c = cos(ϕ/2) and s = sin(ϕ/2).
−s z + c
Prove this in three ways:
′
(i) Show that Rϕ ϕ
i (i) = i and Ri (i) = eiϕ . Why does this prove the result?
(ii) Use the formula for inversion [(3.4), p. 141] to calculate the composition
Rϕi = (RB ◦ RA ), where A and B are h-lines through i, and the angle from
A to B is (ϕ/2). [Hint: Take A to be the imaginary axis, and use a diagram
to show that the semicircle B has centre −(c/s) and radius (1/s).]
(iii) Describe and explain the geometrical effect of applying (D ◦ Rϕ
i ◦ D ) to
−1
the Beltrami–Poincaré disc, where D is the mapping (6.42) from the upper
half-plane to the Beltrami–Poincaré disc. Deduce that
(D ◦ Rϕ
i ◦ D )(z) = e
−1 iϕ
z.
Re-express this equation in terms of products of Möbius matrices, and
solve for the matrix [Rϕ
i ].
382 Non-Euclidean Geometry*
22 (i) Referring to figure [a] above, show that the h-separation of two points in
the upper half-plane may be expressed in terms of a cross-ratio as
e az + b
M(z) = , where a, b, c, d are real, and (ad − bc) > 0.
cz + d
(iii) Use this formula to show that M e has two fixed points on the horizon (the
real axis). If L is the h-line whose ends are these two points, explain why L
is invariant under M. e
(iv) Deduce that Me is always a glide reflection: h-translation along L, followed
by (or preceded by) h-reflection in L.
25 Given a point p not on the h-line L, draw an h-circle C of h-radius ρ centred at
p. Draw the h-line orthogonal to L through p, cutting L in q. Draw an h-circle
C ′ of h-radius ρ centred at q, and let q ′ be one of the intersection points with
L. Through q ′ draw the h-line orthogonal to L, cutting C at a and b. Show that
the h-lines joining p to a and b are the two asymptotic lines! [Hint: Take L to
be a vertical line in the upper half-plane.] What happens if we perform this
construction in the Euclidean plane?
26 Sketch a hyperbolic triangle ∆ with vertices (in counterclockwise order) a, b,
and c. Let ξ be an infinitesimal vector emanating from a and pointing along
the edge ab. Carry ξ to b by h-translating it along this edge. Now carry it to c
Exercises 383
along bc, and finally carry it home to a, along ca. In Euclidean geometry these
three translations would simply cancel, and ξ would return home unaltered.
Use your sketch to show that in hyperbolic geometry the composition of these
three h-translations is an h-rotation about vertex a through angle E(∆).
[Suppose that ∆ is instead a geodesic triangle on an arbitrary surface S of
variable curvature. If an inhabitant of S wants to translate ξ along a “straight
line” (a geodesic), all he has to do is keep its length constant, and keep the
angle it makes with the line constant. This is called parallel transport, and it
plays a central role in differential geometry; see Needham (2021, Act IV). The
above argument still applies, and so when ξ is parallel transported round ∆, it
returns home rotated through E(∆). By virtue of (6.6), this angle of rotation is
the total amount of curvature inside ∆.]
27 Generalize the transformation from the upper half-plane to the Beltrami–
Poincaré disc to obtain a model of hyperbolic space in the interior of the
unit sphere. Describe the appearance of h-lines, h-planes, h-spheres, and horo-
spheres in this model, and explain why an h-sphere is intrinsically the same as
a Euclidean sphere of different radius.
CHAPTER 7
Visual Complex Analysis. 25th Anniversary Edition. Tristan Needham, Oxford University Press.
© Tristan Needham (2023). DOI: 10.1093/oso/9780192868916.003.0007
386 Winding Numbers and Topology
[7.1] The winding number ν of a closed loop L counts the number of times L encircles
a point.
[7.2] Using the winding number to define the “inside” and the “outside” of a closed
loop. A non-simple loop L partitions the plane into multiple simply connected regions,
Dj , and the winding number is constant in each one, taking the value νj . The “inside”
of L is then defined to be the union D1 ∪ D3 of the Dj ’s with νj ̸= 0, and the “outside”
is the union D2 ∪ D4 of the Dj ’s with νj = 0.
Winding Number 387
A typical loop such as L will partition the plane into a number of sets Dj (four
in this case). If the point p wanders around within one of these sets then it seems
plausible that the winding number ν(L, p) remains constant. Let’s check this.
Concentrate on just a short segment of L. As z traverses it, the rotation of
(z − p) will depend continuously on p unless1 p crosses L. In other words, if we
move p a tiny bit then the rotation angle will likewise only change a tiny bit.
Since the winding number of L is just the sum of the rotations due to all its seg-
ments, it follows that it too depends continuously on the location of p: a tiny
movement of p to p e can only produce a tiny change [ν(L, pe) − ν(L, p)] in the
winding number. But since this small difference is an integer, it must be exactly
0. Done.
Since L winds round each point of Dj the same number of times, it follows that
we can attach a winding number νj to the set as a whole. Verify the values of νj
given in the figure.
The “inside” can now be defined to consist of those Dj for which νj ̸= 0, while
the remaining Dj constitute the “outside”. Thus in [7.2] we find that D1 ∪ D3 is the
inside, while D2 ∪ D4 is the outside.
The “correctness” of this definition will become apparent in the next chapter.
1
Consider the behaviour of the rotation due to a short segment of L as p crosses it.
388 Winding Numbers and Topology
[7.3] Crossing Rule: If the loop is travelling from our left to our right as we cross it,
ν ⇝ ν + 1. For imagine that K deforms and moves out of our way, so that we never cross
it, then the winding number does not change. But on the far right, K has evolved into
K and a clockwise circle, L, so ν(K, r) = ν(K, s) + ν(L, s) = ν(K, s) − 1, and therefore
ν(K, s) = ν(K, r) + 1, as claimed.
If K is moving from our left to our right [our right to our left] as we
(7.1)
cross it, its winding number around us increases [decreases] by one.
Using this result, it is incredibly quick and easy to find the νj ’s for even the most
complicated loop. Try it out on [7.2]. Starting your journey well outside L, where
you know that the winding number is zero, move from region to region, using
crossing rule (7.1) to add or subtract one at each crossing of L.
An immediate consequence of this idea is a connection between n = ν(K, p)
and the number of intersection points of K with a ray emanating from p. Suppose
that the ray is in general position in the sense that it doesn’t pass through any self-
intersection points of K, nor is it tangent to K. If a point q on this ray is sufficiently
distant from p then clearly K cannot wind around it; thus as we move along the
ray from p to q the winding number changes by n. But the winding number only
changes when we cross K, and only one unit per crossing. The ray must therefore
intersect K at least |n| times. However, in addition to these |n| necessary crossings
there may be additional cancelling pairs of crossings. In general, then, the number
of intersection points will be |n|, or |n| + 2, or |n| + 4, etc. Figure [7.4] illustrates
these possibilities for a case in which n = 2, each intersection point being marked
with ⊕ or ⊖ according as the winding number increases or decreases as it is crossed.
[7.4] A ray emanating from p intersects K at least |ν(K, p)| times. For as we move along
the ray towards infinity, the winding number must eventually drop to zero. But the initial
value n = ν(K, p) only changes when we cross K, and only one unit at a time, so there
must be at least |n| such crossings. But, in addition to these |n| necessary crossings, there
may be additional cancelling pairs of crossings.
that the same must be true of a fixed point and a continuously moving loop: the
winding number of the evolving loop can only change if it crosses the point, and it
changes by ±1 according to the same crossing rule as before. Thus if a loop K can
be continuously deformed into another loop L without ever crossing a point p, the
winding numbers of K and L around p will be equal.
It is natural to ask if the converse is also true: if K and L wind round p the same
number of times, is it always possible to deform K into L without ever crossing p?
This is certainly a more subtle question, but by drawing examples you will be led
to suspect that it is true. In this section we will confirm this hunch, so establishing
that
At the end of the next chapter, this will turn out to be the key to understanding one
of the central results of complex analysis.
The result in (7.2) is the simplest example of a remarkable topological fact,
called Hopf’s Degree Theorem, that is valid in any number of dimensions. In
the 2-dimensional complex plane, a point can be surrounded using a closed 1-
dimensional curve—a loop. In 3-dimensional space, a point can be surrounded
using a closed 2-dimensional surface. Just as a circle in the plane winds once
390 Winding Numbers and Topology
around its centre, so a sphere in space encloses its centre just once. More gener-
ally, self-intersecting loops in the plane may enclose a point several times, and this
is precisely what ν counts. Similarly, it is possible to define a more general con-
cept (degree) that counts the number of times a surface surrounds a point in space.
Hopf’s Theorem now says that one closed surface may be continuously deformed
into another, without ever crossing p, if and only if they enclose p the same number
of times. Indeed, Hopf’s Theorem says the same is true of n-dimensional surfaces
enclosing points in (n + 1)-dimensional space!
where R(θ) and Φ(θ) are continuous functions. By rotating L (if necessary) we can
ensure that L(ei0 ) is a positive real number, so that we may set Φ(0) = 0. The net
rotation of w after it has returned to its starting point is then given by Φ(2π) = 2πν.
Clearly, the varying length of w is something of a red herring when it comes to
understanding winding numbers, and we now remove this distraction by pulling
each point w of L radially onto the point w b = w/|w| on the unit circle, so obtaining
[7.5] A closed loop L may be viewed as the image of the unit circle C under a continuous
mapping L. As θ goes from 0 to 2π, z = eiθ traverses C once, and w = L(eiθ ) =
R(θ) eiΦ(θ) traverses L once. If Φ(0) = 0 then the net rotation of w after it has returned
to its starting point is given by Φ(2π) = 2πν.
Hopf’s Degree Theorem 391
[7.6] [a] The loop L can be gradually pushed onto the unit circle by moving w along
b − w). As s increases from 0 to 1, Ls (z) = w + s (w
(w b − w) moves radially from w to
b on the unit circle. [b] With a value of s close to 1, the loop has almost evolved into its
w
b , given by w
final form, L b =L b (eiθ ) = eiΦ(θ) .
b − w) .
Ls (z) = w + s (w (7.3)
w b (eiθ ) = eiΦ(θ) .
b =L (7.4)
In this context, it is common to speak of the degree of the mapping L b which pro-
duces Lb , rather than of the “winding number” of L b (or L). The single real function
b
Φ(θ) completely describes the mapping L, and [7.7] shows how we can immedi-
ately read off the degree of L b (i.e. ν) from the graph of Φ(θ). Make sure you are
comfortable with the meaning of such a graph. For example, if z moves at unit
speed round C, what does the slope (including the sign) of the graph represent?
[7.7] Gradual evolution of a loop into the archetypal loop of the same winding number.
If we imagine Lb in the previous figure to be an elastic string wrapped around a cylin-
der, the slack will automatically be taken up, and the actual mapping L b of degree ν
b
will evolve into the archetypal mapping of degree ν, namely, J ν (z) = zν , for which
Φ(θ) = νθ. This can be achieved explicitly by following the illustrated evolution of
Φt (θ) = Φ(θ) + t [νθ − Φ(θ)] from t = 0 to t = 1.
The explicit two-stage deformation given above [(7.3) followed by (7.5)] allows
us to deform any loop of winding number ν into the archetypal loop c Jν , and with-
out the origin ever being crossed. Conversely, by reversing these steps, cJν may be
deformed into any loop of winding number ν. This demonstrates (7.2), for if K and
L both have winding number ν, we may first deform K into c Jν , and then deform
c
Jν into L.
Let A, B, and C be the complex numbers from the fixed points a, b, and c to the
variable point z. Figure [7.8] shows a circle Γ and its image f(Γ ) under the cubic
mapping
f(z) = (z − a) (z − b) (z − c) = ABC .
Notice that Γ encircles two zeros of the mapping, while f(Γ ) has a winding number
of 2 about zero. This is no accident. Since angles add when we multiply complex
numbers, the number of revolutions executed by ABC is just the sum of the revolu-
tions executed separately by each of A, B, and C. But as z goes round Γ once, A and
B both execute a complete revolution, while the direction of C merely oscillates.
Thus ν[f(Γ ), 0] = 2.
If we enlarged Γ so that it encircled c, then C would also execute a complete
revolution, and the winding number would increase to 3. Once again, the number
of points inside Γ that are mapped to 0 is the winding number of the image about
that point.
[7.8] If a simple loop Γ winds once around m roots of a polynomial P(z), then
ν[P(Γ), 0] = m. In the illustrated cubic example, the number of revolutions executed
by f(z) = ABC is just the sum of the revolutions executed separately by each of A, B,
and C. But as z goes round Γ once, A and B both execute a complete revolution, while
the direction of C merely oscillates, so f(z) winds around 0 twice.
394 Winding Numbers and Topology
It is clear that this result is independent of the circularity of Γ , and that it gener-
alizes to the case of a polynomial P(z) of arbitrary degree: If a simple loop Γ winds
once around m roots of P(z), then ν[P(Γ ), 0] = m.
Roots are simply preimages of 0, and from the geometric viewpoint there is noth-
ing special about this particular image point. Consequently, in future we will look
at the preimages of a general point p and we will call these preimages p-points of
the mapping.
The Argument Principle is a tremendous extension of the above result. Not
only does it apply to general analytic mappings but it also contains the converse
statement that the winding number tells us the number of preimages:
The meaning of the expression “counted with their multiplicities” will be explained
in the next section.
We wish to stress that this result is only peripherally connected with the confor-
mality that has been so central to all our previous thinking. In fact the Argument
Principle is a consequence of a still more general topological fact concerning map-
pings that are merely continuous. Our main effort will therefore be directed
towards understanding the general result (due to Poincaré), of which the Argument
Principle is merely a special case.
[7.9] When roots coalesce, the derivative must vanish. The same applies in C, for if
f ′ (a) ̸= 0 then an infinitesimal disc centred at a is amplitwisted to an infinitesimal disc
centred at 0, so that points close to a cannot map to 0.
f ′ (a) f ′′ (a) 2
f(z) − p = f(a + ∆) − f(a) = ∆+ ∆ + ··· .
1! 2!
The first nonzero term on the right is the one that dominates the local behaviour of
f(z) − p and decides what the multiplicity of a should be. Typically a will not be a
critical point [f ′ (a) ̸= 0] and so this local behaviour is like ∆ to the first power; we
say that a is a simple root with multiplicity +1.
Now consider the rarer case in which a is a critical point. If the order of the
critical point is (n − 1), so that f(n) is the first nonvanishing derivative at a, then
the dominant first term is proportional to ∆n , and we correspondingly define the
396 Winding Numbers and Topology
algebraic multiplicity2 of a to be n. The analogy between this definition and that for
polynomials may be brought to the fore by setting
where f(n) (a) is the first nonvanishing derivative. The previous equation can now
be written in “factorized” form as
where Ω(a) ̸= 0. From this point of view, the only difference between a general
analytic mapping and a polynomial is that the latter has a single, “once and for all”
factorization, while the former generally requires a different factorization of type
(7.7) in the neighbourhood of each p-point.
with Ω(a) ̸= 0. Thus the basic explanation is that as ∆ revolves round Ca once,
∆n rotates n-times as fast, and therefore f(z) completes n revolutions round p. If
2
Also known as order or valence.
A Topological Argument Principle* 397
Ω were a constant then this argument would be beyond reproach, and we can now
do a little calculation to show that a variable Ω does not disturb the conclusion:
As we shrink Ca down towards a, Ω(Ca ) will shrink down towards Ω(a), but
since Ω(a) ̸= 0 this implies that the image of a sufficiently small Ca will not wind
round 0: ν[Ω(Ca ), 0] = 0. Since ν[∆, 0] = 1, we conclude that ν[f(Ca ), p ] = n, as
claimed.
We may now broaden our horizon and use the above idea to define “multiplic-
ity” for a mapping h(z) that is merely continuous. Let Γa be any simple loop round
a that does not contain other p-points. Figure [7.10] shows such a loop as well as
some other p-points b, c, etc. If we continuously deform Γa into another such loop
e
Γa without crossing a (or any other p-point) then h(Γa ) will continuously deform
into h(eΓa ) without ever crossing p, and so
ν[h(e
Γa ), p ] = ν[h(Γa ), p ] .
ν(a) ≡ ν[h(Γa ), p ] .
3
Also known as the local degree of h at a.
398 Winding Numbers and Topology
In the case of the mapping in [7.10] we see that ν(a) = −2. If h happened to be
analytic then a would also possess an algebraic multiplicity n, but by deforming
Γa into the infinitesimal circle Ca , we find that the two kinds of multiplicity must
agree: ν(a) = n.
The local linear transformation at a is encoded by the Jacobian matrix J(a), and
we can use its determinant det[J(a)] to give a more practical formula for the topo-
logical multiplicity. We know from linear algebra that the determinant of a constant
2 × 2 matrix is the factor (including a sign for orientation) by which the area of a
A Topological Argument Principle* 399
figure is expanded. Likewise, det[J(a)] measures the local expansion factor for area
at a, and this is just (ξa ηa ). Thus
4
When interpreted in terms of vector fields (as we shall do in Chapter 10) this is the key to a very
surprising and beautiful fact called the Poincaré–Hopf Theorem.
400 Winding Numbers and Topology
[7.11] The winding number ν[h(Γ), p] equals the sum of the topological multiplicities
of the p-points inside Γ. For this winding number will not change value if we deform Γ
as shown into a sum of loops around the individual p-points within.
2π ν[h(Γ ), p ] = 2π ν[h(e
Γ ), p ]
= R(αβγδγβα)
= R(αβ) + R(βγ) + R(γδ) + R(δγ) + R(γβ) + R(βα)
= R(αβα) + R(βγβ) + R(γδγ)
= R(Γa ) + R(Γb ) + R(Γc )
= 2π[ν(a) + ν(b) + ν(c)] .
Clearly this idea extends to any number of p-points a1 , a2 , etc. lying inside Γ :
X
ν[h(Γ ), p ] = ν(aj ) . (7.11)
aj inside Γ
[7.12] Folding a piece of paper across the x-axis is a continuous mapping h, explicitly
given by h(x + iy) = x + i|y|. If Im(p) > 0, as shown, then p has two preimages,
a1 = p and a2 = p, so ν(a1 ) = +1 and ν(a2 ) = −1. Therefore, (7.11) predicts that
ν[h(Γ ), p ] = ν(a1 ) + ν(a2 ) = (+1) + (−1) = 0, which is indeed true!
402 Winding Numbers and Topology
In general, note that ν[h(Γ ), p ] = 0 merely implies that either there are no preim-
ages inside Γ or the preimages have cancelling topological multiplicities, as above.
However, if f is analytic and ν[f(Γ ), p ] = 0 then the conclusion is quite definite:
there are no preimages inside Γ . Later we shall return to this important point.
Returning to the example, observe that if p = X is real then there is only one
preimage, namely X, and ν(X) = 0. We can look at this in a nice way: as we
move p towards X, the two preimages a1 and a2 also move towards X, and when
they finally coalesce at X their opposite multiplicities annihilate. As we previously
pointed out, such points of vanishing multiplicity can only exist for nonanalytic
mappings.
Figure [7.13] shows a second more elaborate example, in which we subject a unit
disc of pastry to a three-stage transformation H that leaves the boundary Γ (the unit
circle) fixed: H(Γ ) = Γ . Here are the three stages: (A) form a “hat” by lifting up the
part of the disc lying inside the dashed circle, some of the pastry dough outside the
dashed circle being stretched to form the side of the cylinder. (B) radially stretch
the disc forming the top of this “hat” till its radius is greater than one. (C) press
down flat, i.e., project each point vertically down onto the plane.
If we pick a point p from the image set [bottom left] then the number of preim-
ages (counted naively) lying in the original disc is the number of layers of pastry
lying over p. In the final picture [bottom left] we have used the degree of shading
to indicate the number of these layers: one over the lightly shaded inner disc; two
[7.13] (A) Form a “hat” by lifting up the part of the disc lying inside the dashed circle,
some of the pastry dough outside the dashed circle being stretched to form the side of the
cylinder. (B) Radially stretch the disc forming the top of this “hat” till its radius is greater
than one. (C) Press down flat, i.e., project each point vertically down onto the plane. The
final number of layers [bottom left] is indicated by the shading: one layer over the lightly
shaded inner disc; two over the darkly shaded outer ring; and three over the black ring.
You can now geometrically verify that (7.11) works in all three regions.
Rouché’s Theorem 403
over the darkly shaded outer ring; three over the black ring. Make sure you can see
this.
We can now check (7.11). For example, if p lies in the darkly shaded outer ring,
ν[H(Γ ), p ] = ν[Γ , p ] = 0 and so the multiplicities of the preimages of such points
should sum to zero. By following the effect of the transformation on little loops
round each of the two preimages, we confirm this prediction: one preimage has
multiplicity +1 while the other has multiplicity−1.
Check for yourself that (7.11) continues to work if p instead lies in the inner disc
or in the black ring.
If |g(z)| < |f(z)| on Γ , then (f + g) must have the same number of zeros
inside Γ as f.
of degree n always has n roots. The basic explanation is simple: if |z| is large, the first
term dominates the behaviour of P(z) and the image of a sufficiently large origin-
centred circle C will therefore wind n times round 0; the Argument Principle then
says that P(z) must have n roots inside C.
Rouché’s Theorem merely allows us to make the above idea more precise. Let
f(z) = zn be the first term of P(z) and let g(z) be the sum of all the rest, so
that f + g = P. Now let C be the circle |z| = 1 + |A| + |B| + · · · + |E|. Using the
fact that |z| > 1 on C, it is not hard to show [exercise] that |g(z)| < |f(z)| on C, and
since f has n roots inside C (all at the origin), Rouché says that P must too.
Notice that we have not only confirmed the existence of the n roots, but have also
narrowed down their location: they must all lie inside C. In the exercises you will
see how Rouché’s Theorem can often be used to obtain more precise information
on the location of the roots of an equation.
m(z) = g(z) − z .
A fixed point then corresponds to no movement: m(z) = 0. Now let f(z) = −z. On
the boundary of D (the unit circle) we have
and so Rouché’s Theorem says that m(z) = g(z) + f(z) has the same number of
roots inside D as f has, namely, one.
If g is merely continuous then there can actually be several fixed points, some of
which will necessarily have negative multiplicities, while if g is analytic then there
can literally only be one.
Let’s see why. The Argument Principle tells us that the sum of the multiplicities
ν(aj ) of the p-points inside Γ is ν[f(Γ ), p ], but if p is outside f(Γ ) then (by definition)
this is zero. Since ν(aj ) is strictly positive for an analytic function, we conclude that
points outside f(Γ ) have no preimages inside Γ . On the other hand, if p lies inside
f(Γ ) then ν[f(Γ ), p ] ̸= 0, and so there must be at least one preimage inside Γ . [Unlike
(7.12), this is also true of nonanalytic mappings.]
Figure [7.14] illustrates an analytic f sending the shaded interior of Γ strictly
to the shaded interior of f(Γ ). Since f(Γ ) winds round the darker region twice, its
points have two preimages in Γ ; we can think of this as arising from the overlap of
two lightly shaded regions, one preimage per lightly shaded point.
One aspect of the “overspill” produced by H in [7.13] is that the points z which
end up furthest from the origin (i.e., for which the modulus |H(z)| is maximum) lie
inside Γ . Conversely, the absence of overspill for an analytic f means that
to guess that the maximum will occur at the centre of the square, but this is wrong.
Since F(z) = |(z − a) (z − b) (z − c) (z − d)| is the modulus of an analytic mapping,
the maximum must in fact occur somewhere on the edge of the square. The exact
location can now be found [exercise] using nothing more than ordinary calculus.
Returning to matters of theory, recall from Chapter 2 that the “modular surface”
of f is the surface obtained by lifting each point z vertically to a height |f(z)| above
the complex plane. If we look at the portion of this surface lying above Γ and its
interior, the result says that the highest point always lies on the edge, never inside.
Although the absolute maximum of the height always occurs on the edge, could
there perhaps be a local maximum of |f| at an interior point a, so that the surface
would have a peak above a? No! For if we cut out the piece of the surface lying
above the interior of any small loop γ round a, the highest point will fail to lie on
the edge. Thus a modular surface has no peaks. Further aspects of the modular
surface are investigated in the exercises.
This absence of local maxima is re-explained in [7.14]. The Argument Principle
says that since γ contains a, f(γ) must wind round A = f(a) at least once. This
makes it clear that there are always points on γ which have images lying further
from the origin than A. More formally (cf. [7.4]), any ray emanating from A must
intersect f(γ), and by choosing the ray to point directly away from the origin, the
intersection point is guaranteed to lie further from the origin than A. Thus |f| cannot
have a local maximum.
The Schwarz–Pick Lemma* 407
where a lies inside the disc. These one-to-one mappings of the disc to itself act as
rigid motions, for they preserve non-Euclidean distance.
Apart from a digression on Liouville’s Theorem, this section continues the work
(begun in Chapter 6) of exhibiting the beautiful pre-existing harmony that exists
between non-Euclidean geometry and the theory of conformal mappings. Our first
new piece of evidence that these two disciplines somehow “know” about each
other is the following:
Rigid motions of the hyperbolic plane are the only one-to-one analytic
(7.14)
mappings of the disc to itself.
408 Winding Numbers and Topology
There are of course many other kinds of analytic mapping of the disc to itself, but
according to (7.14) they must all fail to be one-to-one. For example, z ÞÑ z3 maps
the disc to itself, but it is three-to-one.
Observe that this result establishes a claim we previously made [see p. 204] in
connection with Riemann’s Mapping Theorem. There we explained that there are
as many mappings of one region to another as there are automorphisms of the
disc. We already knew that these automorphisms included 3-parameter’s worth of
Möbius mappings, and (7.14) now tells us that there are no more.
To verify (7.14) we will first establish a lemma (of great interest in itself) due to
Schwarz:
If an analytic mapping of the disc to itself leaves the centre fixed,
then either every interior point moves nearer to the centre, or else the
transformation is a simple rotation.
The example f(z) = z2 shows that the mapping need not be a rotation in order
for boundary points to keep their distance from the centre. However, at an interior
point we have |z| < 1, and so |f(z)| = |z|2 < |z|, in accord with the result.
Let f be any analytic mapping of the disc to itself leaving the centre fixed, so
that |f(z)| ⩽ 1 on the disc, and f(0) = 0. We wish to show that either |f(z)| < |z|
at interior points, or else f(z) = eiϕ z. To this end, consider the ratio F of image to
preimage:
f(z)
F(z) ≡ .
z
At first sight this may look undefined at 0, but a moment’s thought shows that as
z approaches the origin, F(z) approaches f ′ (0).
From the previous section we know that the maximum modulus of an analytic
function on the disc can only occur at an interior point if the function is constant,
otherwise it’s on the boundary circle |z| = 1. Thus if p is an interior point and z
varies over the unit circle C, then
Thus it is certainly true that no interior point can end up further from the centre. But
if even a single interior point q remains at the same distance from the centre then
|F(q)| = 1, which means that F has achieved its maximum modulus at an interior
point. In this case F must map the entire disc to a single point of unit modulus, say
eiϕ , so that f(z) = eiϕ z is a rotation. Done.
The result is illustrated in [7.15]. If f is not a rotation then every point z on a circle
such as K is mapped to a point w = f(z) lying strictly inside K, and the shaded
region is compressed as shown. If we shrink K down towards the origin then f
will amplitwist it to another infinitesimal circle centred at the origin, but having a
The Schwarz–Pick Lemma* 409
[7.15] Schwarz’s Lemma: If an analytic mapping f(z) of the disc to itself leaves the
centre fixed, then either every interior point moves nearer to the centre, or else the trans-
formation is a simple rotation. As explained in the text, this follows from applying the
Maximum-Modulus Theorem to F(z) ≡ f(z)/z.
smaller radius. Thus the amplification of f at the origin must be less than one. We
can only have |f ′ (0)| equal to one in the case of a rotation.
We can now return to (7.14). As in Schwarz’s Lemma, first suppose that the
mapping f leaves the centre fixed, but now take f to be one-to-one, so that it has
a well-defined analytic inverse f−1 which also maps the disc to itself and leaves the
centre fixed. By Schwarz’s Lemma, f sends an interior point p to a point q that is
no further from centre than p. But f−1 is also subject to Schwarz’s Lemma, and so
p = f−1 (q) must be no further from the centre than q. These two statements are
only compatible if |q| = |f(p)| = |p|. Thus f must be a rotation, which is indeed a
rigid motion of type (7.13).
Finally, suppose that the one-to-one mapping f does not leave the centre fixed,
but instead sends it to c. We can now compose f with the rigid motion Mc which
sends c back to 0. We thereby obtain a one-to-one mapping (Mc ◦ f) of the disc
to itself which does leave the centre fixed, and which must therefore be a rotation
Mϕ 0 . But this means [exercise] that
f = Mc ◦ Mϕ
0
is the composition of two rigid motions, and so is itself a rigid motion. Done.
the entire plane down to a region lying inside a finite circle, without going to the
extreme of completely crushing it to a point.
If we merely demand that the mapping be continuous then this can happen. For
example,
z
h(z) =
1 + |z|
maps the entire plane to the unit disc. Returning to analytic mappings, we notice
that complex inversion, z ÞÑ w = (1/z), manages to conformally compress the
infinite region lying outside the unit circle of the z-plane into the unit disc of
the w-plane. This looks quite hopeful: of the original plane only a puny unit disc
remains to be mapped.
To think like this is to completely forget the rigidity of analytic mappings. Hav-
ing decided to use complex inversion to map the region outside the unit circle, we
cannot change the rules when it comes to mapping the remaining disc: the mapping
z ÞÑ (1/z) acting in the exterior can only be analytically continued to the interior
in one way, namely as z ÞÑ (1/z). The requirement of analyticity thereby forces the
“puny” disc to explode, producing an image of infinite size.
We will now show that
An analytic mapping cannot compress the entire plane into a region
lying inside a disc of finite radius without crushing it all the way down
to a point.
This is Liouville’s Theorem. To understand this we must generalize Schwarz’s
Lemma slightly. Suppose that an analytic function w = f(z) leaves the origin fixed
and compresses the disc |z| ⩽ N to a region lying inside the disc |w| ⩽ M. By the
same reasoning as before, we find that if p lies inside the original disc (boundary
circle K) then
|f(z)| M
|F(p)| ⩽ [max |F(z)| on K] = max on K ⩽ .
N N
Hence,
M |p|
|f(p)| ⩽ .
N
But if f compresses the whole plane to a region lying inside the disc of radius M,
then the above result will continue to hold true no matter how large we make N.
Therefore f(p) = 0 for all p, and we are done.
Finally, if f does not leave the origin fixed, but instead sends it to c, we may apply
the previous argument to the function [f(z) − c]. This is the composition of f with
the translation which sends c back to 0. Since the image of the plane under f lies
inside the disc |w| ⩽ M, the translation of −c will produce a region lying inside the
disc |w| ⩽ 2M. The previous inequality then becomes
The Schwarz–Pick Lemma* 411
2M |p|
|f(p) − c| ⩽ .
N
Once again letting N tend to infinity, we conclude that f(p) = c for all p. Done.
5
Recall from Chapter 6 that this is the race of beings who inhabit the Beltrami–Poincaré model of
the hyperbolic plane.
6
Georg Alexander Pick [1859–1942] is perhaps best known for his wonderfully simple but com-
pletely unexpected Pick’s Theorem, which determines the area A of a lattice polygon in terms of the
number of lattice points Inside it (I), and on its Boundary (B), namely, A = I + (B/2) − 1. However,
his greatest contribution to science may have been the fact that he fought to appoint Einstein at the
German University of Prague in 1911, became close friends with him there—playing violin together
in a musical quartet—and finally introduced Einstein to the tensor calculus of Ricci and Levi-Civita,
without which Einstein could not have formulated General Relativity in 1915! [This is documented in
Goodstein (2018, p. 90), though, strangely, Einstein himself only seems to have publicly credited his
friend Marcel Grossmann for this critical piece of guidance.] Pick died at the hands of the Nazis at
Theresienstadt concentration camp on the 26th of July, 1942, at the age of 82.
412 Winding Numbers and Topology
Because this result contains Schwarz’s Lemma as a special case [we shall clarify
this shortly] it is often called the Schwarz–Pick Lemma. Despite the startling nature
of the result, we can actually understand its essence very simply; we need only ask
the question, “How do the Beltrami–Poincarites view [7.15]?”
Because their concept of angle is identical to ours, it follows that their concept
of an analytic function is also the same as ours—f appears conformal both to us
and to them. In addition, we both agree that rays emanating from the origin are
straight lines along which we may measure distance. Consequently, the Beltrami–
Poincarites willingly concede that w is closer to 0 than z is, although they violently
disagree with our quantitative determination of exactly how much closer it is. Now
recall that there is a small (psychological) flaw in the Beltrami–Poincaré model:
0 appears special to us because it is the centre of the disc, but to the Beltrami–
Poincarites who inhabit an infinite, homogeneous plane 0 is utterly indistinguishable
from any other point of their world.
The above explanation is formalized in [7.16]. In the top left figure we see that
the Beltrami–Poincarites have marked an arbitrary point a, drawn a few concen-
tric circles centred there, and on the outermost of these they have marked a second
point b. They (and we) now consider the effect of an analytic mapping f of their
world to itself. The point a is sent to some image point A = f(a) [top right] and
likewise b is sent to B = f(b). In order to compare the separation of A and B
with that of a and b, the Beltrami–Poincarites perform a rigid motion (MA ◦ Ma
would do) that moves the circles centred at a to circles (of equal hyperbolic size)
centred at A. Consequently, the hyperbolic separation of a and b will have been
decreased [increased] by f according as B lies inside [outside] the outermost of
these circles. In anticipation of Pick’s result we have drawn it inside, corresponding
to a decrease in hyperbolic separation. However, observe that as in our previous
numerical example, to us Euclideans it looks as though the separation has been
increased.
In order to show us poor blind Euclideans that the circles centred at a and A
really are concentric and of equal sizes (so enabling us to see that B really has got-
ten closer) the Beltrami–Poincarites perform the illustrated rigid motions Ma and
MA . These respectively move a and A to the origin [bottom left and bottom right
figures], yielding circles that are as concentric to us as they always were to them.
Ma moves b to z = Ma (b), while MA moves B to
w = MA (B) = (MA ◦ f)(b) = (MA ◦ f ◦ Ma )(z).
We shall abbreviate (MA ◦ f ◦ Ma ) to F, so that w = F(z).
We can now see that the following are all equivalent:
H{A, B} < H{a, b} ⇐⇒ H{0, w} < H{0, z} ⇐⇒ |w| = |F(z)| < |z|.
But F is an analytic mapping of the disc to itself which leaves the origin fixed, and
so it is subject to Schwarz’s Lemma. Thus unless F is a rotation—in which case
The Schwarz–Pick Lemma* 413
f = (MA ◦ F ◦ Ma ) is a rigid motion—we must have |w| = |F(z)| < |z|, as depicted.
Done.
Finally, let us express the Schwarz–Pick Lemma in symbolic form. If f is not a
rigid motion then |F(z)| < |z|, which may be written out more explicitly as
Thus
B−A b−a
< .
AB − 1 ab − 1
|dA| |da|
< ,
1 − |A| 2 1 − |a|2
which we may interpret [cf. (6.44), p. 363] as saying that, provided f is not a rigid
motion, the hyperbolic length of dA is less than that of its preimage da. This is the
infinitesimal version of (7.15).
(z − a) (z − b) 1
f(z) = = AB · . (7.16)
(z − c) C
ν [f(Γ3 ), 0] = 2 − m.
The Generalized Argument Principle 415
As with counting zeros, we could say in this case that c was a singularity [or pole,
as we shall now call such places] of multiplicity m.
The previous equation is an example of the Generalized Argument Principle:
Let f be analytic on a simple loop Γ and analytic inside except for a finite
number of poles. If N and M are the number of interior p-points and (7.17)
poles, both counted with their multiplicities, then ν [f(Γ ), p ] = N − M.
Simply by allowing an arbitrary number of factors on the top and bottom of (7.16)
we see that this result is certainly true when f is any rational function.
Before explaining why it works in general, let us develop a more vivid under-
standing of how it works in the case of our example (7.16). We have certainly shown
that as Γ crosses c the winding number drops from 2 to 1, but exactly how does this
unwinding occur?
If we look at the image plane just as Γ crosses c then f(Γ ) undergoes a sudden
and violent change of shape as it leaps to infinity and then returns, but this leaves
us none the wiser. However, if we instead watch its evolution on the Riemann sphere
then we gain a new and delightful insight into the process.
Figure [7.18] (which should be scanned like a comic strip) illustrates this. At
the time of the first picture [top left] Γ has already enclosed the two roots, and its
image is seen to wind round the origin twice. Now follow the evolution of f(Γ )
416 Winding Numbers and Topology
[7.18] Reducing the winding number by passing over the point at infinity. The Riemann
sphere provides a vivid explanation of the previous figure. As Γ expands from Γ2 to Γ3 ,
crossing the pole at c, the loop on the Riemann sphere simply slides over the north pole,
unwinding the loop and reducing its winding number from 2 to 1.
through the remaining pictures. As Γ crosses c [top right] there is no longer any
excitement—f(Γ ) merely slides across the north pole, and this is how the unwinding
is achieved. Try using a computer to animate the evolution of the image f(Γ ) on the
Riemann sphere as Γ expands through the roots and poles of a rational function f
of your choosing.
thinking of the modular surface of f, for there will be an infinitely high spike or
“pole” above the point a. Figure [2.14] on p. 74 is an example of this.
Since f is analytic, it follows that F(z) ≡ [1/f(z)] is also analytic and has a root at
a. If this root has multiplicity m then the factorization (7.7) of F is
where Ω is analytic and nonzero at a; in fact we know that Ω(a) = F(m) (a)/m!.
The local behaviour of f near a is therefore given by
e
Ω(z)
f(z) = , (7.19)
(z − a)m
e
where Ω(z) = [1/Ω(z)] is analytic and nonzero at a. This expression brings out the
analogy with rational functions and enables us to identify m as the algebraic mul-
tiplicity or order of the pole at a. We call a pole simple, double, triple, etc., according
as m = 1, 2, 3, etc.
Note that we have also found a way of calculating the order of a pole, namely,
as the order of the first nonvanishing derivative of (1/f). Once you have identified
the locations of the poles, you may use this method [exercise] to find the orders of
the poles of the following functions:
1 cos z 1
P(z) = ; Q(z) = ; R(z) = .
sin z z2 (ez − 1)3
You should have found that P has a simple pole at each multiple of π; Q has a
double pole at 0; and R has a triple pole at each multiple of 2πi.
One more piece of terminology. If the only singularities in some region of an
otherwise analytic function are poles, the function is called meromorphic in that
region.
In addition to poles, it is also possible for an otherwise analytic function to pos-
sess what are called essential singularities. We shall postpone detailed discussion of
such places to a later chapter, but it is clear that the behaviour of a function f in
the vicinity of an essential singularity s must be very strange and wild. If f were
bounded in the vicinity of s then s would not be a singularity at all, but on the
other hand f(z) cannot approach ∞ as z approaches s from all directions, for then
s would only be a pole.
Consider the standard example g(z) = e1/z , which clearly has a singularity of
some type at the origin. If we write z = r eiθ then
cos θ
|g(z)| = e r .
Figure [7.19] depicts the modular surface. If z approaches 0 along the imaginary
axis then |g(z)| = 1. But if the approach is instead made along a path lying to the
left of the imaginary axis (where cos θ < 0) then g(z) tends to 0. Finally, if the
418 Winding Numbers and Topology
[7.19] The modular surface of g(z) = e1/z reveals its essential singularity at the origin.
cos θ
Since |g(z)| = e r , the behaviour at 0 depends drastically upon the direction in which
we approach it. If cos θ < 0 as we approach, then the height |g(z)| drops to zero, and
the surface hits C. If we approach precisely along the tightrope of the imaginary axis (for
which cos θ = 0) then our height holds precariously steady at 1. But if cos θ > 0 as we
approach, then the height explodes to infinity faster than any polynomial!
approach path lies to the right of the imaginary axis then g(z) tends to ∞. In fact,
not only will |g(z)| become infinite in this case, but the rate at which it zooms off to
∞ is quite beyond the ken of any pole.
To see this, reconsider (7.19). The greater the order m, the faster the growth of f
as the pole at a is approached. However, no matter how great the order happens to
be, we know that (z − a)m dies away fast enough to kill this growth, in the sense
that its product with f remains bounded. Indeed, the order of a pole can be defined
as the smallest power of (z − a) which will curb the growth of f in this way.
Compare this with the growth of g(z) as its essential singularity is approached,
say along the positive real x-axis. To confirm that g grows faster than any mero-
morphic function, we need only recall from ordinary calculus that
eλ
lim xm e1/x = lim = ∞,
xÑ0 λÑ∞ λm
e
f(z) = (z − a)−m Ω(z),
Finally, reconsider figure [7.11]. You may now easily convince yourself that the
argument leading to (7.11) remains valid if some of the aj ’s are poles instead of
p-points. Let’s call these singular points sj . Thus
X X
ν[f(Γ ), p ] = ν [f(Γaj ), p ] + ν [f(Γsj ), p ] .
p-points poles
as was to be shown.
420 Winding Numbers and Topology
7.9 Exercises
1 A “simple” loop can get very complicated (see diagram). However, if we imag-
ine creating this complicated loop by gradually deforming a circle, it is clear
that it will wind round its interior points precisely once. Let N(p) be the num-
ber of intersection points of the simple loop with a ray emanating from p
(cf. [7.4]). What distinguishes the possible values of N(interior point) from
those of N(exterior point)? In place of the crossing rule (7.1), you now possess
(for simple loops) a much more rapid method of determining whether a point
is inside or outside.
You can use this result to play a trick on a friend F: (1) So that foul play cannot be
suspected, get F to draw a very convoluted simple loop for himself; (2) choose a
random point in the thick of things and ask F if it’s inside or not, i.e., starting at
this point, can one escape through the maze to the outside?; (3) after F has been
forced to recognize the time and effort required to answer the question, get him
to choose a point for you; (4) choosing a ray in your mind’s eye, scan along it
and count the intersection points. Amaze F with your virtually instantaneous
answer!
2 Reconsider the mapping L b in (7.4) of the unit circle to itself, and the associated
graph of Φ(θ) in [7.7]. If Φ ′ (a) > 0 then the graph is rising above the point θ =
a, and small movement of z will produce a small movement of the w having the
same sense. We say that Φ is orientation-preserving at a and that the topological
multiplicity ν (a) of z = eia as a preimage of w = eiΦ(a) is +1. Similarly, if
Φ ′ (a) < 0 then the mapping is orientation-reversing and ν (a) = −1. In other
words,
ν (a) = the sign of Φ ′ (a).
Compare this with the 2-dimensional formula (7.9).
(i) In [7.7], explain how the complete set of preimages of w = eiA can be found
by drawing the family of horizontal lines Φ = A, A ± 2π, A ± 4π, etc.
Exercises 421
(ii) If the set of preimages is typical in the sense that Φ ′ ̸= 0 at any of them,
what do we obtain if we sum their topological multiplicities? Thus to say
that the degree of Lb (the winding number of L) is ν is essentially to say
b is ν-to-one. [Hint: In [7.4], consider a ray as describing the location
that L
of w.]
3 For each of the following functions f(z), find all the p-points lying inside the
specified disc, determine their multiplicities, and by using a computer to draw
the image of the boundary circle, verify the Argument Principle.
(i) f(z) = e3πz and p = i, for the disc |z| ⩽ (4/3).
(ii) f(z) = cos z and p = 1, for the disc |z| ⩽ 5.
(iii) f(z) = sin z4 and p = 0, for the disc |z| ⩽ 2.
4 Reconsider [7.8].
(i) Use a computer to draw the image under a cubic mapping
f(z) = (z − a) (z − b) (z − c)
ν (a) = (−1)n .
dn Q dn−1 Q dQ
cn n
+ cn−1 n−1
+ · · · + c1 + c0 Q = 0.
dt dt dt
Recall that one solves this equation by taking a linear superposition of special
solutions of the form Qj (t) = esj t . Substitution into the previous equation
shows that the sj ’s are the roots of the polynomial
Note that Qj (t) will decay with time if sj has a negative real part. The issue of
whether or not the general solution of the differential equation decays away
with time therefore reduces to the problem of determining whether or not all n
roots of F(s) lie in the half-plane Re(s) < 0. Let R be the net rotation of F(s) as s
traverses the imaginary axis from bottom to top. Explain the following result:
The general solution of the differential equation will die away if and only if
R = nπ.
This is called the Nyquist Stability Criterion, and an F that satisfies this condition
is called a Hurwitz polynomial. [Hints: Apply the Argument Principle to the loop
consisting of the segment of the imaginary axis from −iR to +iR, followed by
one of the two semicircles having this segment as diameter. Now let R tend to
infinity.]
Exercises 423
zn e a = e z
ν [e
Γ , 0] = ν [p(Γ ), 0] + ν [q(Γ ), 0].
(ii) Write
g(z)
f(z) + g(z) = f(z) 1 + = f(z) H(z).
f(z)
If |g(z)| < |f(z)| on Γ , sketch a typical H(Γ ). Deduce that
ν [H(Γ ), 0] = 0.
ν [f(Γ ), 0] = ν [f ′ (Γ ), 0] + 1.
(iv) Deduce from the Argument Principle that f has one more root inside Γ
than f ′ has. This is sometimes called Macdonald’s Theorem, though I believe
its essence goes back as far as Riemann.
(v) From this we deduce, in particular, that f has at least one root inside Γ .
Derive this fact directly by considering the portion of the modular surface
lying above Γ and its interior.
14 In contrast to analytic mappings, it is perfectly possible for a continuous non-
analytic mapping to completely crush pieces of curve or even areas without
crushing the rest of its domain. Let us give a concrete example to show that the
Topological Argument Principle does not apply to this case. With r = |z|, the
mapping h(z) = ϕ(r) z will be a continuous function of z if ϕ(r) is a continu-
ous function of r. Consider the continuous mapping h of the unit disc to itself
corresponding to
ϕ(r) = 0, 0 ⩽ r < (1/2);
ϕ(r) = 2r − 1, (1/2) ⩽ r ⩽ 1.
But clearly this normal vector undergoes one positive revolution as z traces C.
Deduce that the vector m is also dragged round one revolution.
16 Let f(z) be an odd power of z, and consider its effect on the unit circle C. Note
two facts: (1) if p is on C then f(−p) points in the opposite direction to f(p); (2)
ν [f(C), 0] = odd, in particular it cannot vanish. This is only one example of a
general result. Show that (1) always implies (2), even if f is merely continuous.
[Hints: If f is subject to (1), what can we deduce about the net rotation R of f(z)
as z traverses the semicircle from p to −p? How is the rotation produced by the
remaining semicircle related to R ?]
17 Consider a spherical balloon S resting on a plane. If we gradually deflate S, each
point will end up on the plane so that we have a continuous mapping H of S into
the plane. Observe that the north and south poles, which are antipodal, have
the same image. The Borsuk-Ulam Theorem says that any continuous mapping
H of S into the plane will map some pair of antipodal points to the same image.
Consider the mapping F(p) = H(p) − H(p⋆ ), where p⋆ is antipodal to p. The
theorem then amounts to showing that F has a root somewhere on S. Prove this.
[Hints: It is sufficient to examine the effect of F on just the northern hemisphere.
By taking the boundary of this hemisphere (the equator) to be the circle C of
the previous exercise, deduce that ν [H(C), 0] ̸= 0.]
18 Let f be analytic on a simple loop Γ , and let p be a preimage of a point on f(Γ )
at which |f| is maximum. If ξ is a tangent complex number to Γ at p, and in the
same counterclockwise sense as Γ , show geometrically that ξf ′ (p) points in the
same direction as if(p). What is the analogous result at a positive minimum of
|f| ?
19 (i) If p is not a critical point of an analytic function f, show geometrically that
the modulus of f increases most rapidly in the direction [f(p)/f ′ (p)].
(ii) In terms of the modular surface above p, this direction lies directly beneath
the tangent line to the surface having the greatest “slope” (i.e., tan of the
angle it makes with the complex plane). Show that the slope of this steep-
est tangent plane is |f ′ (p)|, and note the analogy with the slope of the
ordinary graph of a real function.
(iii) What does the modular surface look like at a root of order n ?
(iv) What does the modular surface look like above a critical point of order
m ? Using the case m = 1, explain why such places are called saddle points.
(v) Rephrase Macdonald’s Theorem [Ex. 13] in terms of the number P of pits
and the number S of saddle points in the portion A of the modular surface
lying above Γ and its interior.
426 Winding Numbers and Topology
[Hint: If an edge of a cell does not form part of Γ then it is also an edge of an
adjacent cell, but traversed in the opposite direction. What is the net rotation of
f(z) round p as z traverses this edge once in one direction, then in the opposite
direction?]
CHAPTER 8
8.1 Introduction
In the last few chapters our efforts to extend the idea of differentiation to complex
mappings have been amply rewarded. By innocently attempting to generalize the
real derivative we were quickly led to the amplitwist concept, and the subject then
came to life with a character all its own. While many of the results cast familiar
shadows onto the world of the reals, many did not, and striking indeed was the
flavour of the arguments used to grasp them. The ability of z to freely roam the
plane unleashed in us a degree of visual imagination that had to remain dormant so
long as we could only watch the real number x forlornly pacing its one-dimensional
prison.
This little hymn to the glory of the complex plane can be sung again in the context
of integration, only louder. If differentiation breathed life into the subject, then inte-
gration could be said to give it its soul. Only after we have understood this soul will
we be able to demonstrate such fundamental facts as the infinite differentiability
of analytic mappings1 .
Rb
In ordinary calculus the symbol a has a clear meaning. However, if we wish to
generalize this to C then the need for new ideas is immediately apparent, for how
are we to get from a to b? In R there was only one way, but a and b are now points in
the plane, so we must specify some connecting path (called a contour) “along which
to integrate”. It is then natural to ask whether the value of the integral depends
upon the choice of this contour.
In general the value of the integral will depend on the route chosen. For example,
we will shortly meet an integral of a complex mapping that yields, when evaluated
for a closed contour, the area enclosed by the contour—a flagrant dependence of
1
Since the 1960’s it has actually become possible to do such things without integration, thanks to
pioneering work by G. T. Whyburn, and others. Nevertheless, integration still appears to provide the
simplest approach.
Visual Complex Analysis. 25th Anniversary Edition. Tristan Needham, Oxford University Press.
© Tristan Needham (2023). DOI: 10.1093/oso/9780192868916.003.0008
430 Complex Integration: Cauchy’s Theorem
value on contour. It should be made clear from the outset that while differentiation
only made sense for the strictly limited set of analytic functions, this is not the case
for integration. Indeed, the example just cited involves the integration of a non-
analytic function.
The principal aim of this chapter (beyond the mere construction of an integral
calculus) will be the discovery of conditions under which the value of an integral
does not depend on the choice of contour. One such result is an analogue of the
Fundamental Theorem of real analysis, and in deference to that subject it bears the
same name. However, in the complex realm this is actually a misnomer, for there
exists a still deeper result which has no counterpart in the world of the reals. It is
called Cauchy’s Theorem.
As we have said, it is not only possible, but sometimes useful to integrate non-
analytic functions. However, it should come as no surprise to learn that new
phenomena arise if we concentrate on the integrals of mappings that are analytic.
Cauchy’s Theorem is the essence of these new phenomena. Essentially it says that
any two integrals from a to b will agree, provided that the mapping is analytic
everywhere in the region lying between the two contours. Almost all the fundamental
results of the subject (including some already stated) flow from this single horn of
plenty.
Pn
[8.1] A Riemann Sum R ≡ j=1 f(xj ) ∆j approximates the area under a curve with
rectangles, the exact answer being obtained in the limit that n goes to infinity and each
∆j goes to zero.
freedom in the choice of xj becomes more and more limited, and the influence of
the choice on the area of the rectangle likewise diminishes. However, if we are
unwilling or unable to actually carry out the limiting process, then, as we shall
now see, we can ill afford to be so blasé in our choice of xj .
You probably dimly remember some professor showing you (8.1) before, and
perhaps you even evaluated a couple of examples by means of it. However, this
was no doubt quickly forgotten once you set eyes upon the Fundamental Theorem
of Calculus. In order to integrate x4 , why bother with taking the limit of some
complicated series when we know that the answer must be that function which
differentiates to x4 , namely, 51 x5 ?
The Fundamental Theorem is a wonderful thing, but one must remem-
ber that many quite ordinary functions simply do not possess an antideriva-
tive that is expressible in terms of elementary functions. To take a simple
example, the Normal Distribution of statistics requires a knowledge of the area
2
under the curve e−x , and this can only be computed numerically, perhaps via a
Riemann sum.
When doing a numerical calculation with (8.1), it would require an infinite
amount of time to find even a single area with perfect precision. It is therefore
important to be able to obtain good approximations to limnÑ∞ R while using only
a finite value of n. Several such methods exist: Simpson’s rule and the Trapezoidal
432 Complex Integration: Cauchy’s Theorem
rule, to name just two that may be familiar. Since the Trapezoidal rule will most
readily lend itself to complex generalization, we will now review it.
[8.2] The Trapezoidal Rule replaces the rectangles with trapezoids, and clearly yields an
accurate approximation to the integral while using only a modest value of n.
The Real Integral 433
[8.3] The exact same area as the previous figure, but now represented as rectangles whose
heights are given by the midpoints of the illustrated chords.
[8.4] The Midpoint Riemann Sum, RM , mimics the previous figure, but now uses the
height of the graph at each midpoint, instead of the height of the chord. As an approxi-
mation to the integral, a typical Riemann sum that uses the height at a random point has
a total error that dies away as ∆, but we will show that the Midpoint Riemann Sum is
much more accurate, with an error that dies away as ∆2 .
434 Complex Integration: Cauchy’s Theorem
[8.5] [a] Start of geometric determination of the accuracy of the Midpoint Riemann Sum.
The area of the rectangle occuring in RM is the area under D e C,
e but this is the same as
the area under the tangent, DC, so the error is the shaded region between DC and the
curve. This error is certainly less than the area of ABCD, which is (PQ) · ∆. [b] As the
chord PQR revolves about the fixed point Q, the product PQ · QR remains constant.
2
While many of the arguments in this book were merely inspired by Newton’s mode of thought in
the Principia, we have here an example that is very close to his actual methods.
The Real Integral 435
eC
D e [the RM rule]. However, note that if we rotate D eCe about P (keeping the ends
glued to the verticals) until it becomes tangent at DC, the area beneath it will remain
constant [why?]. Thus we are instead free to visualize each term of RM as being the
area lying beneath a tangent such as DC. It is now clear that the actual area lies
between the two values furnished by AB and DC, and that the error induced by
using either rule cannot exceed the area of the small quadrilateral ABCD, namely
[exercise], (PQ) · ∆. In order to find this area we will use the elementary property
of circles that is illustrated in [8.5b]: As the chord PQR revolves about the fixed point
Q, the product PQ · QR remains constant.
Over a sufficiently tiny distance we can consider any segment of curve to be
interchangeable with its tangent. However, over somewhat larger distances (or if
we simply require greater accuracy) we must instead replace it by a segment of its
circle of curvature, that is, the circle whose curvature κ agrees with that of the curve
at the point in question. In [8.6] we have drawn this circle for the segment at P. The
above result now informs us that
PQ · QR = (AQ)2 . (8.2)
As ∆ shrinks, both (AQ/DP) and (QR/PR) tend to unity, so in this limit we may
substitute DP for AQ and PR for QR. But if the tangent at P makes an angle θ
with the horizontal (in which case OP makes angle θ with the vertical) then DP =
1
2 ∆ sec θ, and PR = (2/κ) cos θ. Substituting these into (8.2), we obtain the result
area (ABCD) = PQ · ∆ ≍ 18 κ sec3 θ ∆3 . (8.3)
If M denotes the maximum of 81 κ sec3 θ over the integration range (which we take
to be of length L) then each such error will be less than M ∆3 . Since the number of
these error terms is (L/∆), we conclude that
total error < (LM) ∆2 ,
and this indeed dies away in the manner originally claimed. [At this point you may
care to look at Ex. 1]
Because the order of the induced error is the same for both RM and the Trape-
zoidal rule, we will tend not to distinguish between them when it comes to their
complex generalizations. This said, there remains one curious pedagogical point
still to be made.
Figure [8.5a] makes it clear that curves deviate less from their tangents than from
their chords, and thus one would anticipate that while the order of the error is the
same for both rules, RM would actually yield the more accurate value of the two.
This is indeed the case, and in fact [see Ex. 2] one can show that it is twice as accurate.
In addition to this accuracy, RM is, if anything, easier to remember and use than the
Trapezoidal formula. It is therefore doubly puzzling that the Trapezoidal formula
is taught in every introductory calculus course, while it appears that the midpoint
Riemann sum RM is seldom even mentioned.
R
[8.7] Complex Riemann Sums. To approximate K f(z) dz, we approximate K with many
small complex chords, ∆j , evaluate f(z) at an arbitrary point zj within each segment
P
of K, then evaluate the sum, f(zj ) ∆j . Given the established superior accuracy of the
Midpoint Riemann Sum (RM ) in the real case, here we have immediately chosen each
zj to be the midpoint of each segment. As we shall explain, the Riemann sum can be
visualized more easily if we choose all the ∆j ’s to have the same length. [The turning
angles ϕj illustrated here are for later use.]
[8.8] The values of the function along the contour of integration: wj = f(zj ). [The
angles τj between wj and wj+1 are for later use.]
Just as in the real case, we may obtain an accurate estimate of the integral without
passing to the limit, simply by choosing the zj to be at the midpoints of the segments
of K, rather than at random points. In fact this is the choice that we have illustrated
in [8.7]. Once again, this especially accurate Riemann sum will be denoted RM .
To begin to understand the geometry of RM , consider [8.8]. This shows the image
of K under the mapping z ÞÑ w = f(z), and in particular the image wj of the zj that
was singled out in [8.7]. The corresponding term of RM is then ∆ e j ≡ wj ∆j , and
we will choose to think of this as the arrow that results when wj “acts on” ∆j ,
expanding it by |wj | and rotating it by arg(wj ).
Having obtained each ∆ e j in this manner, we go on to join all these little arrows
together (tail to tip), as in [8.9]. The value of RM , and hence the approximate value
of the integral, is then the connecting complex number between the start and the
finish3 . Notice that since the answer is a connecting arrow, the point at which we
begin drawing RM is irrelevant.
While [8.9] is intended primarily to convey the general idea, it is in fact a faith-
ful evaluation of the specific RM corresponding to [8.7] and [8.8], and you may
now begin to convince yourself of this. This is perhaps most easily achieved by
concentrating on the lengths of the ∆ e j separately from their angles. As w traces
out the image curve in [8.8], its length diminishes, and this produces a corre-
sponding shrinking of the ∆ e j in [8.9]. Likewise, the increasing angle of w results
in progressively greater rotations of the ∆j .
3
The great physicist Richard Feynman used a similar kind of picture to explain his quantum-
mechanical “path integrals”, which are also complex, though they differ from contour integrals. See
Feynman (1985).
The Complex Integral 439
[8.9] Visualizing the Midpoint Riemann Sum, Rm . The angle of each term ∆ e j ≡ wj ∆j
Pe
in RM = ∆j is the sum of the angles of wj and ∆j , so the illustrated turning angle
e j between successive steps along the Riemann sum is therefore the sum of the turning
ϕ
e j = ϕ j + τj .
angles ϕj and τj in the previous two figures: ϕ
Let us spell this out in detail with reference to the concrete example furnished by
[8.7] and [8.8]. In [8.9], we get RM started in the right direction by rotating ∆1 by α,
thereby obtaining ∆ e 1 which points at angle α + β. We can now draw the rest of RM
using only (8.4). To lay down the next ∆ e we need to know ϕ e 1 = ϕ1 + τ1 . The small
positive τ1 clearly kills off just a fraction of the negative ϕ1 , resulting in a slightly
smaller negative bend in RM . Much the same happens when we lay down ∆ e 2 . The
angle ϕ3 at the next bend is positive, and it is therefore increased by τ3 , which itself
is about twice as big as τ1 and τ2 were. You should now be in a position to follow
the rest of RM ’s progress in far greater detail than you could before.
Although the above idea will shortly prove its worth on a theoretical level, it
is clearly not terribly practical. However, in Chapter 11 we will use an entirely
different approach to obtain a second, less strenuous, means of visualizing com-
plex integrals. We will thereby make a double fallacy of an assertion that is to be
found in most texts—assuming they even consider it worthy of note!—namely, that
complex integrals possess no geometric interpretation. Perhaps the mere frequency
with which this myth has been reiterated goes some way to explaining how it has
acquired the status of fact.
But the sum on the right is just the length of the polygonal approximation to K,
and hence it cannot exceed the actual length of K. Passing to the limit where RM
becomes the integral, we deduce that
Z
f(z) dz ⩽ M · (length of K) . (8.5)
K
For example, if f(z) = (1/z)2 and K is the circle |z| = r, then (8.5) implies that
R R
K f(z) dz ⩽ (2π/r). In particular this implies that limrÑ∞ K f(z) dz = 0. This is a
typical (albeit simplistic) application of (8.5): quite often one wishes to demonstrate
4
The method of visualizing complex integrals in Chapter 11 will enable us to express this condition
for equality in a particularly simple form. See Ex. 6, p. 574.
The Complex Integral 441
[8.10] [a] If L begins where K left off, then to integrate along K + L means to integrate
along K and then to continue integrating along L, and the resulting integral is then just the
sum of the two separate integrals. [b] Changing K to −K reverses the sign of the integral,
for −K is defined to be the same curve as K, but traversed in the opposite direction. [c] If
e then the
the integral from a to b along K equals the integral along an alternative route K,
e e
integral along the closed loop (K − K) = (K followed by − K) must be zero. Conversely,
if the integral vanishes for all closed loops then all curves between a and b will yield the
same value for the integral. In brief: Path independence is equivalent to vanishing loop
integrals.
Thus equality of the two integrals is equivalent to the vanishing of the integral
taken along the closed loop (K − K) e = (K followed by − K). e Conversely, if the
integral vanishes for all closed loops then all curves between a and b will yield the
same value for the integral. In brief: path independence is equivalent to vanishing loop
integrals. The centrepiece of complex analysis is the link between this phenomenon
and analyticity. Cauchy’s Theorem consists in recognizing that the vanishing of
loop integrals is the nonlocal manifestation of a local property of the mapping,
namely, that it is an amplitwist everywhere inside the loop.
[8.11] The integral of (1/z) along a circle. Since ∆1 and w1 are ultimately vertical and
horizontal respectively, it follows that RM (in [c]) initially heads off in the vertical direc-
tion. But since the turning angle τ of wj = (1/zj ) in [b] is the opposite of the turning angle
ϕ of ∆ in [a], ϕe = ϕ + τ = 0. In other words, RM has no bends, and so it continues on
in the imaginary direction. We wanted to illustrate the turning-angle approach for later
use, but here is a simpler explanation. Since the ∆j located at angle θ on the circle itself
points at θ to the vertical (as illustrated) and has length Aϕ, we see that multiplying it by
w = (1/A)e−iθ rotates it by −θ, producing a vertical ∆ e = w∆ = iϕ.
origin is also given by ϕ, it follows that |∆| = Aϕ. As z travels round the circle its
image w = 1/z travels round a circle of radius 1/A in the opposite direction (see
[8.11b]), and thus w shrinks each ∆ to produce a ∆ e of length ϕ.
Since ∆1 and w1 are ultimately vertical and horizontal respectively, it follows that
RM (which we choose to begin drawing at the origin in [8.11c]) initially heads off in
a vertical direction. But now we observe that τ = −ϕ, and consequently that ϕ e = 0.
In other words RM has no bends, and so it continues on in the imaginary direction5 .
Thus, irrespective of the radius, the integral equals i times the total angle Ψ through
5
We have used the turning angle idea in order to make the subsequent generalization to other
powers of z straightforward, but there is actually no need for it in the present case. Since the ∆j located
at angle θ on the circle itself points at θ to the vertical and has length Aϕ, we see that multiplying it
by w = (1/A)e−iθ rotates it by −θ, producing a vertical ∆ e = w∆ of length ϕ.
444 Complex Integration: Cauchy’s Theorem
which z turned on its journey along K. Convince yourself that this formulation of
the result remains valid even if K begins at a random point of the circle instead of
on the real axis.
In particular, and of crucial importance, is the case where z continues all the way
round the circle to form a closed loop. The value of the integral is then 2πi. The alert
reader will immediately be perplexed by this result. Why? Because it appears to fly
in the face of Cauchy’s Theorem. We have previously demonstrated geometrically
that complex inversion is analytic, so how can its loop integral fail to vanish?! The
resolution lies in the fact that Cauchy’s Theorem requires that the mapping be ana-
lytic everywhere inside the loop. But our loop encloses the origin, and just at this
one point the analyticity of complex inversion breaks down.
[8.12] The integral of (1/z) along a general contour. [a] The movement ∆ along a general
contour can be broken down into a radial component dr and an orthogonal component
rdθ making angle θ with the vertical. [b] Multiplication by w = (1/z) therefore yields
e = w∆ = (dr/r) + idθ.
∆
Complex Inversion 445
[8.13] The integral of (1/z) along a closed loop. [a] Instead of choosing ∆j ’s of equal
length, divide the path up, as shown, using closely spaced concentric circles centred at
the origin. Consider the pair ∆j and ∆k cut off by the same two circles. From the previous
figure, we deduce that real components of ∆ e j and ∆e k in the Riemann sum will cancel.
But every ∆ belongs to such a pair (or pairs), because if the closed loop L passes from the
interior of a circle to the exterior, then in order to join up with itself back in the interior,
it must recross the circle in the opposite direction somewhere else. Thus the integral
of (1/z) along a closed loop is always imaginary. The previous figure also implies that
this imaginary answer = i (net angle of rotation as the loop is traversed). [b] The
illustrated loop L does not enclose the origin, so the integral vanishes. [c] If the loop does
wind around 0 once, then the integral is 2πi.
the vertical and horizontal directions, as well as shrinking their lengths to dθ and
(dr/r), respectively.
Let us now see what happens if we stick all these ∆ e j together for a closed loop
such as L (see [8.13a]) that does not encircle the origin. In order to accomplish this
we will forsake our previous choice of equal lengths for all the ∆j , and instead
divide the path up, as shown, using closely spaced concentric circles centred at the
origin6 . Consider the illustrated pair ∆j , ∆k lying between adjacent circles. That
the ∆’s always do occur in such pairs is a consequence of L being a loop. For if
6
This only fails if part of L coincides with such a circle, but in that event we already know that the
contribution to the integral is iΨ.
446 Complex Integration: Cauchy’s Theorem
L passes from the interior of a circle to the exterior, then in order to join up with
itself back in the interior, it must recross the circle in the opposite direction some-
where else. Of course L may weave back and forth across a circle many times (e.g.
at a, a⋆ , b, b⋆ , c, c⋆ ), but the crucial point is that these crossings always occur in
oppositely directed pairs.
In [8.13b] we see the consequence of this for R. From [8.12b] it’s clear that for a
pair such as ∆ e j and ∆ e k , the horizontal components cancel. Since we have seen that
every ∆ belongs to such a pair, it follows that R will have no horizontal component
for any closed loop, whether or not the origin is encircled. It also follows from
[8.12b] that the height of this vertical Riemann sum is obtained by adding up all
the signed angles that the ∆’s subtend. For a loop such as [8.13a], which does not
encircle the origin, this sum is zero: as z traces out L its direction merely oscillates,
rather than executing a complete revolution. Thus, as illustrated in [8.13b], R closes
up on itself. On the other hand, if we translated L to any location where it encircled
the origin then z would execute a complete revolution, and [8.13b] would change
into [8.13c].
where the integral sign with a circle through it (which is a standard symbol) serves
to remind us that we are integrating around a closed contour. Figure [8.14] shows
various loops and the corresponding value of the integral of (1/z) round each of
them. Finally, note that (8.6) can easily be generalized [exercise] to
I
1
dz = 2πi ν(L, p). (8.7)
z−p
L
Conjugation 447
H
[8.14] The previous figure proves that L (1/z) dz = 2πi ν(L, 0).
8.5 Conjugation
8.5.1 Introduction
In the introduction we stressed that integration makes sense for any continu-
ous complex mapping, regardless of whether or not it is analytic. However, the
relatively lawless non-analytic functions give rise to integrals that behave less pre-
dictably than their analytic counterparts. In particular, Cauchy’s Theorem has no
jurisdiction here, and we therefore have no reason to anticipate path independence
or, equivalently, vanishing loop integrals. As an example of this type of behaviour,
we will show presently that the loop integral of the non-analytic conjugation map-
ping z ÞÑ z yields the area enclosed by the loop. Assuming this result for the
moment, let us use the examples z and (1/z) to spell out more clearly the differences
between the non-analytic and analytic cases.
In the analytic case, provided that the special point z = 0 was not enclosed, the
loop integral vanished. Even when the integral of (1/z) did not vanish, its possible
values were still neatly quantized in units of 2πi; one unit for each time the special
point z = 0 was enclosed by the loop. As we will see later, this behaviour is typical,
although a more general mapping may well possess several special points (at which
analyticity breaks down) dotted about in the plane. Once again, the integral is not
sensitive to the precise shape of the loop. Provided that none of the special points
are enclosed by the loop, then the integral vanishes. However, if some of the points
are enclosed, then each one makes its own distinctive contribution (generally not
2πi) to the integral, one unit for each time it’s encircled. The value of the integral
is just the sum of these discrete contributions.
448 Complex Integration: Cauchy’s Theorem
Contrast all this with our non-analytic example. The area of the loop (and hence
the integral of z) will almost never vanish. Furthermore, instead of being deter-
mined by stable topological properties, the value of the integral is sensitive to the
detailed geometry of the loop. Finally, the value is not neatly quantized, but instead
varies continuously as the loop changes shape.
Adding these elements together, we obtain the imaginary part of the Riemann sum
corresponding to the integral of z. Thus we conclude that
I
Im z dz = 2 (area enclosed) .
L
This result can be further simplified by noticing that z and (1/z) both point in the
same direction. It follows that we could draw a picture very similar to [8.12], the
e we would multiply by r instead of dividing by
only difference being that to obtain ∆
H
[8.15] L z dz = 2i (area enclosed by L). [a] Using H(1.20) on page 32, we see that
2 (element of area) = Im[(z + ∆)z] = Im[z∆], so Im L z dz = 2 (area). But z has the
same direction as (1/z), and the argument given in [8.12] easily generalizes, so the inte-
gral is purely imaginary, thereby proving the result. [b] The geometric meaning of the
integral is unaltered if 0 lies outside L, for although the top triangle includes area that lies
outside L, this extra area is cancelled by the negative area of triangles like the one at the
bottom. Compare this to [1.22b] on page 33.
Conjugation 449
it. The argument that followed from [8.12] therefore remains valid, and we deduce
that the integral of z around a closed loop is purely imaginary. Thus
I
z dz = 2i (area enclosed) . (8.8)
L
Next we ask how this formula would change if the origin were outside the loop.
Figure [8.15b] shows that the pleasing answer is, “Not at all!” The point is that the
integral adds up the signed areas subtended by the ∆’s at the origin. On the far
side, ∆ carries z counterclockwise, yielding a positive element of area; but on the
near side z is moving clockwise, yielding a negative element of area. When these
are added, the unwanted area lying outside the contour simply cancels, leaving
behind just the area enclosed.
As a simple example, consider a circle C of radius r centred at a, the equation of
which is r2 = |z − a|2 = (z − a) (z − a). Solving this for z, and using (8.7), we find
that
I I I
2 1
z dz = a dz + r dz
z−a
C C C
2
= 0 + r 2πi
= 2i (area enclosed).
From what we have done so far you might be inclined to think that the integral
of z could never vanish for a nontrivial loop. That this is false can be seen from the
figure eight loop in [8.16a]. This may be thought of as the union of two separate
loops. The top one is traversed in a positive sense and correspondingly yields its
ordinary area A1 ; but the bottom one is traversed in a negative sense and yields the
H
[8.16] [a] If L is a figure eight, then L z dz = 2i(A
H 1 −A2 ). Note that if L were symmetrical,
P
then the integral would vanish. [b] In general, L z dz = 2i inside νj Aj = 2i [2A1 + A3 ]
in this example.
450 Complex Integration: Cauchy’s Theorem
negative (−A2 ) of its ordinary area. Thus the integral is 2i(A1 − A2 ), and if the loop
were symmetrical then this would vanish.
where Aj denotes the area of Dj . For example, in the case of [8.16b] we obtain
H
L z dz = 2i [2A1 + A3 ]. We hope you can already see why this example is correct,
but a full explanation of the general formula (8.9) will be provided later in this
chapter.
RB m+1
[8.17] A zm dz = m+1 1
B − Am+1 , because [a] |∆| = Aϕ, [b] τ = mϕ and
e = Am+1 ϕ, so [c] ϕ
|∆| e = τ + ϕ = (m + 1) ϕ, and therefore ρ = |∆|/
e ϕ e = Am+1 /(m + 1).
chosen to place at the origin in [8.17c]. We will now determine the angle subtended
by this arc, and also its radius.
The angle that each ∆ e subtends at the origin is the same as the turning angle ϕ,
e
namely, (m + 1) times the angle subtended by each ∆. Thus
angle of FINISH = (m + 1)Ψ.
Also, if ρ is the radius, we see from the figure that
e = |∆|
e Am+1
ρϕ =⇒ ρ= .
m+1
We therefore conclude that if m ̸= −1 then
RM = FINISH − START
1 h m+1 i(m+1)Ψ i
= A e − Am+1
m+1
1 m+1
= B − Am+1 (8.10)
m+1
452 Complex Integration: Cauchy’s Theorem
which, as promised, is formally identical to the real result. We hope you will agree
that it’s rather fascinating how we have been able to visualize this result in a way
that would not have been possible in the real case.
As we have said, [8.17] actually depicts the concrete case m = 2, and before
continuing you may care to sketch another case for yourself; m = −2 might be a
fun one.
[8.18] The integral of (1/z) as a limiting case of the integral of zm . Here, n ≡ m − (−1)
measures the “distance” of zm from the ultimate target of z−1 . But there is also a fun-
damental difference between (1/z) and all other powers: its integral around a complete
circle is 2πi, whereas the integral for all other powers is zero!
Power Functions 453
because RM will now go round in a complete circle |n| times [clockwise if n < 0;
counterclockwise if n > 0], thereby returning to its beginning.
[8.19] Deformation Theorem. [a] Deform L by pushing the segment pq outward to form
the illustrated bump. Since the mapping is analytic between the two paths connecting p
and q, it follows from Cauchy’s Theorem that both integrals are equal, and since the rest
of L hasn’t changed, it follows that integral with bump = integral without bump. [b] We
can continue to deform L until it becomes J, and the value of the integral will not change.
7
In Chapter 11 we will give a physical explanation for this difference between complex inversion
and the rest of the negative powers.
454 Complex Integration: Cauchy’s Theorem
to the integral round L that comes from the piece between p and q. Suppose that
we deform L slightly by replacing this segment by the bump in the figure. Since
the mapping is analytic between the two paths connecting p and q, it follows from
Cauchy’s Theorem that both integrals are equal. Also, since the rest of L hasn’t
changed, it follows that integral with bump = integral without bump. All we need do
now, to obtain the stated result, is to let the bump grow and change shape (see
[8.19b]) until L has evolved into J.
The crucial idea is this
Thus, if you imagine the contour to be a rubber band, and the singularity to be a
peg sticking out of the plane (thereby obstructing motion past it), the integral has
the same value for all shapes into which the rubber band can be deformed.
We can immediately apply this Deformation Theorem to our problem. For if the
mapping is a negative power of z other than z−1 then the established fact that the
integral vanishes for a circular loop implies it continues to vanish for any loop into
which the circle can be deformed without crossing the singularity at the origin.
Thus formula (8.10) is path independent even for negative powers.
The Deformation Theorem also provides us with a much simpler derivation
of the result (8.6) governing the general loop integral of the complex inversion
mapping. Imagine taking a length of elastic string and winding it around an origin-
centred circle ν times, finally joining the ends together to form a closed loop. From
our earlier work, it follows that the value of the integral is then 2πiν. But the Defor-
mation Theorem says that this will be the value of the integral for any loop into
which the elastic string may be deformed without being forced over the peg (sin-
gularity) at the origin. Finally, by the Hopf Degree Theorem, the loops into which
it can so be deformed are those with winding number ν.
We can evaluate the integral round any loop C by noting this alternative expression:
i i
f(z) = − .
z+i z−i
Applying (8.7) therefore yields
I
f(z) dz = 2π[ν(C, i) − ν(C, −i)].
C
Assuming (as in [8.20a]) that L encloses both singularities, use this formula to verify
(8.12) for this particular function.
8.6.5 Residues
Since we now possess a fairly complete understanding of the loop integrals of
power functions, it is relatively easy to integrate simple rational functions: we need
only find the decomposition into so-called partial fractions, and then integrate term
by term. Indeed, this is precisely what we did in the example of the last paragraph.
Here is a slightly more complicated example: the integral of f(z) = z5 /(z + 1)2
taken round the contour K in [8.21]. This has a second-order pole at z = −1, and by
writing the numerator as [(z + 1) − 1]5 , we quickly find that
1 1
f(z) = − +5 − 10 + 10 (z + 1) − 5 (z + 1)2 + (z + 1)3 .
(z + 1)2 z+1
456 Complex Integration: Cauchy’s Theorem
[8.21] The Residue of a singularity. Let f(z) = z5 /(z + 1)2 , which has a second-order
pole at z = −1. If f(z) is decomposed into partial fractions then the only part that is not
removed by integration is its “residue”—that which remains, defined to be the coefficient
of the complex inversion term, 1/(z+1). This is the only part that does not integrate away
into thin air. The value of the integral therefore has only two ingredients: the residue and
the winding number, and the result is 2πi times their product, which turns out to be
−20πi for this particular function and contour.
But we know that the loop integral of powers other than −1 is zero, and so only the
complex inversion term [in square brackets] can contribute. In detail,
I
f(z) dz = 5 · 2πi ν(K, −1) = −20πi.
K
Thus the value of the integral has been determined by just two factors: the winding
number of the loop, and the amount (i.e., coefficient) of complex inversion con-
tained in the decomposition of the mapping. Because this latter number is the only
part of the function that remains after we integrate, it is called the residue of the
function at the singularity. Quite generally, the residue of f(z) at a singularity s is
denoted Res [f(z), s]. Thus in the above example, Res [z5 /(z + 1)2 , −1] = 5.
In fact the residue concept has a significance that extends far beyond sim-
ple rational functions, as our next example will illustrate. We have previously
[page 259] alluded to the remarkable fact that analytic functions are infinitely dif-
ferentiable, or equivalently, that they can always be represented by a power series
(Taylor’s) in the vicinity of a nonsingular point. For example, the Taylor series
centred at the origin for sin z is
1 3 1 1
sin z = z − z + z 5 − z7 + · · · .
3! 5! 7!
Clearly no such expansion can be possible at a singular point of a mapping.
Nevertheless, we may recover an analogous result near singularities simply by
broadening our notion of a power series to include negative powers. Such a series
is called a Laurent series.
The Exponential Mapping 457
Consider (sin z)/z6 . This is singular at the origin, but by simple division of the
above Taylor series we obtain the following Laurent series in the vicinity of the
singularity:
sin z 1 1 1 1 1 1
= 5− + − z + z3 − · · · .
z6 z 3! z3 5! z 7! 9!
Once again, the residue of the function is defined to be the coefficient of the complex
inversion term: Res [(sin z)/z6 , 0] = 1/5! in this case. If a power series converges at
every point on a contour, then we may accept for the moment that it makes sense
to integrate the series term by term. Once again we see that for a closed loop the
sole contribution to the integral comes from the residue. For example, if K is the
contour in [8.21] then
I
sin z 1 2πi
dz = 2πi ν(K, 0) = − .
z6 5! 5!
K
In the case of the exponential mapping the easiest contour along which to integrate
is a vertical line-segment, say L (see [8.22a]). Once again we will find the result to
be formally identical to its real counterpart:
Z
ez dz = eB − eA . (8.13)
L
R
[8.22] Geometric evaluation of L ez dz. [a] Let L be the illustrated vertical segment
from A to B = A + iθ. [b] If we choose to begin drawing RM at eA then the illustrated
arc
R traced by w = ez is also the precise path taken by the Riemann sum, and therefore
e e
L e dz = e − e . First, all the ∆’s have the same length, namely, |∆| = e |∆|. Since L
z B A A
arc e
|∆|
radius = = = eA ,
angle e
ϕ
as required. Lastly, the total angle subtended by the arc at its centre is just the sum
e j = |∆j |, namely θ. The identity of the two arcs is thus established.
of all the ϕ
Since ez is singularity-free, Cauchy’s Theorem assures us that its loop integral
always vanishes. Thus (8.13) must in fact be valid for any path from A to B.
8.8.1 Introduction
Through specific geometric constructions, combined with the use of Cauchy’s
Theorem, we have already learnt a good deal about the integrals of some of the
most important functions. However, there are two immediate problems still to be
resolved: one is pragmatic, while the other is aesthetic.
The Fundamental Theorem 459
The pragmatic one is that the formulae (8.10) and (8.13) are only known to hold
for certain special configurations of the points A and B: the derivation of (8.10)
assumes that they are equidistant from the origin; while for (8.13) they are assumed
to be vertically separated. To be sure, our various forms of Cauchy’s Theorem guar-
antee us that the integrals in question will continue to be path-independent, no
matter what the locations of A and B. But the problem is that we haven’t yet estab-
lished that these path independent values will continue to be given by the same
formulae as before. In this section we shall see that they are.
The aesthetic concern lies in the manner in which we derived path-independence
for negative powers of z. Recall that we were only able to apply Cauchy’s Theorem
after having explicitly produced an example (a circle) of a loop integral that van-
ishes in spite of enclosing the singularity. Although this was neat enough in itself,
one is left with the feeling that Cauchy’s Theorem cannot be the most direct
way of understanding a loop integral that continues to vanish in the presence of
singularities.
A resolution of both these problems is provided by the so-called Fundamental
Theorem of Contour Integration—a result that is formally identical to its similarly
named counterpart in ordinary calculus. The naming of this theorem is not entirely
appropriate, at least in the context of complex analysis. After all, so far we have
managed quite well without it, suggesting that if this theorem is “Fundamental”,
then Cauchy’s must be “Super-Fundamental”!
8.8.2 An Example
As our first example of this theorem, let us return to the exponential mapping of
the last section in order to discover why (8.13) is valid for any pair of points, not
just ones that are vertically separated. As so often happens in mathematics, all that
is required is a very slight shift in viewpoint.
Figure [8.23] depicts a curve K (connecting a pair of typical points A and B)
being mapped by ez to the curve K e connecting eA and eB . Now let us forget (for a
moment) all about integration and Riemann sums, and instead look at the figure
from the point of view of differentiation.
All the little arrows emanating from a point on K will be mapped to images ema-
nating from a point on K. e In particular, if the arrow ∆ is a little chord of K [tangent,
e will likewise be a little directed chord of
in the limit that it shrinks], then its image ∆
e But for an analytic mapping, such as we are now considering, the original arrows
K.
are sent to their images by a simple amplitwist:
e = (amplitwist of ez ) · ∆ = ez ∆ .
∆ (8.14)
If we now add up all these vector chords of K e then we obtain the connecting vector
V between its start and its finish. But (8.14) tells us that this vector V may also be
460 Complex Integration: Cauchy’s Theorem
[8.24] The Fundamental Theorem in the general case. If F maps K to K e and F ′ = f then
e e simply is the
· ∆ = f(z) ∆, and therefore K
∆ is amplitwisted to ∆ = [amplitwist of F(z)] R
Riemann sum for the integral of f. Therefore, K f(z) dz = V = F(B) − F(A).
Just as before, we conclude that K e is actually the path taken by the Riemann sum
of f, and that the vector V is once again the path-independent value of the integral:
Z
f(z) dz = V = F(B) − F(A) . (8.15)
K
8
We exclude contours that actually pass through the singularity.
The Fundamental Theorem 463
R R
[8.25] [a] Given
R R a fixed point R A, suppose we evaluate M f(z) dz and L f(z) dz. [b] Then
M f(z) dz − L f(z) dz = −L+M f(z) dz is an integral along an indirect route from P to
Q. If the integral is path-independent,
RZ we can replace the indirect route
R with the direct
route S. Defining F(Z) ≡ A f(z) dz, we deduce that F(Q) − F(P) = S f(z) dz.
464 Complex Integration: Cauchy’s Theorem
Z
e=
F : ∆ ÞÑ ∆ f(z) dz .
∆
But if ∆ is infinitesimal then the above integral equals f(P) ∆, thereby establishing
the original claim:
e = f(P) ∆ .
F : ∆ ÞÑ ∆
for all positions of A and B. This was apparently a clear improvement on [8.17]
where such formulae were merely established in the special case that A and
B were equidistant from the origin. However, we will now see that analytic-
ity makes it possible, paradoxically, for this special case to contain the general
case.
Consider these two functions:
ZZ
1
F(Z) = z2 dz G(Z) = (Z3 − A3 ) ,
A 3
both of which we now recognize as being analytic. From [8.17] we know that if Z
moves along the origin-centred circle passing through A then
F(Z) = G(Z) .
But by the uniqueness property of analytic functions [page 284] this identity must
continue to hold even if Z wanders off the circle, thereby establishing the general
result.
By applying exactly similar reasoning to the exponential mapping, we may like-
wise extrapolate the validity of (8.13) for vertically separated points (established
by [8.22]) to deduce that
ZZ
ez dz = eZ − eA ,
A
just as in real analysis. In a sense, this is correct, but a little care is required.
The subtlety is, of course, that the singularity at the origin causes the integral
of (1/z) not to be single-valued. Thus we must specify the contour K from 1 to
Z before the integral in (8.17) becomes well defined. On the other hand, until we
choose one of the infinitely many values θ(Z) for the angle of Z, the RHS of (8.17)
is also not well defined. These two difficulties now cancel each other out in the
following way.
√
In [8.26] we have drawn three different contours for the specific case Z = 1+i 3.
If we let θK (Z) stand for the net rotation as we follow K, then
θK0 (Z) = (π/3)
θK1 (Z) = (π/3) + 2π
θK2 (Z) = (π/3) + 4π .
In a sense, including the contour in the definition of angle has rendered it single-
valued. Notice that this definition does not depend on the precise shape of K, but
only on how many times the origin is encircled.
RZ
[8.26] Why 1 (1/z) dz = log Z is a multifunction. From [8.12] we deduce that
RZ
1 (1/z) dz = logK (Z) = ln |Z| + i θK (Z), where θK (Z) denotes the net change in the
angle as z traverses K. So the infinitely many values of log Z arise from the number of
times K loops around 0 before finally arriving at Z, each extra revolution adding 2πi to
the value of log Z .
466 Complex Integration: Cauchy’s Theorem
In this way, we may absorb the means of reaching Z into the definition of log(Z)
in order to obtain a single-valued answer:
Of course, the multiple-valued nature of log has merely been disguised, not done
away with. Nevertheless, by pursuing the above idea one is led to consider so-
called Riemann surfaces, whereby multifunctions can be rendered single-valued. But
that is a story for another day.
When more elegant means are not available it is nevertheless possible (in principle)
to evaluate a contour integral by expressing it in terms of ordinary real integrals.
We shall now briefly describe and illustrate this method.
The basic idea is to think of the contour L as being traced by a moving particle
whose position at time t is z(t). Next, instead of building the Riemann sum (hence
the integral) from very small vectors that are chords of L, we may equally well use
very small vectors that are tangent to L. This is done using the tangential complex
velocity v = dzdt : the chord representing the movement during the instant of time
δt may be replaced by the tangential vector v δt. Thus if L is traced out during the
time interval a ⩽ t ⩽ b, then
Z Zb
f[z] dz = f[z(t)] v dt .
L a
For example, suppose that L is one counterclockwise circuit of the circle with
radius ρ and centre q, and that f[z] = z. We know from our earlier work that the
answer should be 2πiρ2 . Since z(t) = q + ρ eit (0 ⩽ t ⩽ 2π) and v = i ρ eit , we
obtain
Z Z 2π
z dz = (q + ρ e−it ) iρ eit dt
L 0
Z 2π Z 2π
= iρq (cos t + i sin t) dt + iρ2 dt
0 0
= 2πiρ2 ,
as anticipated.
Naturally, the point of this method is not to confirm previously known results,
but rather to evaluate integrals that we couldn’t do before. For example, with the
Cauchy’s Theorem 467
same contour, but with f[z] = z2 , the answer can no longer be guessed. However,
you should now find it easy to discover that the answer is 4πiqρ2 .
By way of contrast with the non-analytic examples above, and as further practice
R
with this method, confirm (using the same contour L) that L z2 dz = 0, as predicted
by either Cauchy’s Theorem or the Fundamental Theorem. Likewise, confirm that
R it
E z dz = 0, where E is an origin-centred ellipse. [Hint: recall that z(t) = p e +
−it
qe moves on such an ellipse.]
For our last examples, take the contour to be a section of the parabola y = x2
between 0 and 1 + i; in temporal terms this can be represented as z(t) = t + it2 (0 ⩽
t ⩽ 1). Integrate z along this contour, first using the Fundamental Theorem, then
parametrically. Likewise, use the Fundamental Theorem to evaluate the integral
for ez . By equating the imaginary part of your answer with the imaginary part of
the parametric evaluation, deduce that
Z1
(2t cos t2 + sin t2 ) et dt = e sin 1 .
0
This result can be verified easily [exercise] without using complex numbers. Later,
though, we shall meet real integrals that cannot readily be evaluated by such ordi-
nary means, but which suddenly do become easy when viewed as arising from a
complex integral.
[8.27] Cauchy’s Theorem (step one): Fill the interior of C with a grid of small (ultimately
vanishing) squares, of side length ϵ, aligned with the real and imaginary axes. The sum
of the integrals around all these squares yields the integral around the boundary, K, for
every interior edge is matched with an oppositely directed abutting edge that cancels it.
The investigation of the integral of f along C has thus been reduced to the study of
the local effect of f on infinitesimal squares in the interior region.
It should be stressed that the discussion thus far is equally applicable to non-
analytic and analytic mappings. For example, with f(z) = z, (8.18) simply says
[see (8.8)] that the area inside K is the sum of the areas of the shaded squares. In
order to understand Cauchy’s Theorem, we must specialize to the case (illustrated
in [8.27]) where the local effect of f is an amplitwist throughout the interior of C.
First, though, let us try to guess how the magnitude of a typical integral in the
above summation will depend on ϵ (as the squares shrink) for a general mapping.
Experience with real integration, as well as the inequality (8.5), might lead one
to guess that the integral round an infinitesimal square would die away at the same
rate as its perimeter, that is, as ϵ. This is false. The fact that the square is a closed
contour, together with the fact that complex integration is a type of vectorial sum-
mation, implies that the integral must decay much faster than this. In the above
example of conjugation, we know that the exact value of each term is 2iϵ2 , and this
Cauchy’s Theorem 469
leads us to the correct guess, namely, that the terms die away as the square of ϵ. We
shall verify this in detail shortly, but for the moment the following rough argument
will suffice.
We know that for a general mapping, the integral round K—hence the summa-
tion in (8.18)—will be nonzero and finite. This leads us to believe that each term
must die away with the reciprocal dependence on ϵ as governs the growth of the
number of terms in the series. But the number of terms grows as (fixed area inside C,
divided by the area of each square), that is as (1/ϵ2 ). Thus the magnitude of each term
is expected to die away as ϵ2 . If our original guess had been correct, the order of the
sum in (8.18) would have been ϵ (1/ϵ2 ), yielding an infinite result as the squares
shrunk. Conversely, any contributions to the terms involving powers of ϵ greater
than two, cannot have any influence on the final result.
[8.28] Cauchy’s Theorem explained. According to our especially accurate midpoint Rie-
mann sum, RM , the integral along the bottom edge of this square can be approximated by
the single term Aϵ: the image of the midpoint a, times the number along this edge. If we
add this to the integral along the opposite edge, the answer is A ϵ+C (−ϵ) = (A−C) ϵ =
pϵ . Likewise, the contribution from the remaining two edges is also of order ϵ2 , namely,
(B − D) iϵ = iqϵ. But since f is locally an amplitwist,H the image is a square, and so
iq = (q rotated through a right angle) = −p. Therefore, □ f(z) dz = ϵ (p + iq) = 0.
470 Complex Integration: Cauchy’s Theorem
A ϵ + C (−ϵ) = (A − C) ϵ = pϵ .
Even if f is merely differentiable in the real sense, rather than locally an amplitwist,
|p| will still be proportional to ϵ, and the magnitude of pϵ will therefore be pro-
portional to ϵ2 , as anticipated. Likewise, the contribution from the remaining two
edges is also of order ϵ2 , namely, (B − D) iϵ = iq ϵ.
Perhaps you have already seen the light: if f is locally an amplitwist, the image
is a square, and so
iq = q rotated through a right angle = −p
I
=⇒ f(z) dz = ϵ (p + iq) = 0 . (8.19)
□
We conclude from (8.18) that the vanishing of loop integrals for analytic mappings
is indeed the nonlocal manifestation of their local amplitwist property!
Contrast this with non-analytic mappings. See [8.29]. Provided that a mapping is
differentiable in the real sense, we know [see page 236] that its local effect is expan-
sion (by different factors) in two perpendicular directions, followed by a twist.
Thus the image of an infinitesimal square will generally be a parallelogram; p and
q will not have equal length, nor will they be orthogonal. As we see, p and iq no
longer cancel, and ϵ (p + iq) is of order ϵ2 . When we add up the terms of (8.18), of
order (1/ϵ2 ) in number, the answer will therefore be nonzero and finite.
Conjugation provides a particularly striking example of this noncancellation for
non-analytic mappings. See [8.30]. In the terms of the previous paragraph we could
say that its expansion factors are everywhere 1 and −1 (in the horizontal and verti-
cal directions), and that its twist is zero. The image of the square is again a square,
but there is a crucial difference between [8.28] and [8.30]. Because this mapping is
anticonformal, it reverses the orientation of the square, and we see that
[8.29] The non-analytic case. The shrinking square is ultimately mapped to a (non-
square) parallelogram. As we see, p and iq no longer cancel, and ϵ(p + iq) is of order
ϵ2 . When we add up the terms of (8.18), of order (1/ϵ2 ) in number, the answer will
therefore be nonzero and finite.
Cauchy’s Theorem 471
H
[8.30] □ z dz = 2i(area of □), because ϵ(p + iq) = ϵ(iϵ + iϵ) = 2i ϵ2 .
I
z dz = ϵ (p + iq) = ϵ (iϵ + iϵ) = 2i ϵ2 = 2i (area of □).
□
[8.31] As in Section 7.1.2, we define the “inside” of a closed loop to be the set of points
for which the winding number is not zero. Then if an analytic mapping has no singularities
inside a loop, its integral round the loop vanishes, as it does in this example.
The General Cauchy Theorem 473
the process of understanding this particular instance of the theorem we shall be led
to a completely general argument for its validity.
Now comes the crucial observation. The integral round K can be expressed as a
linear combination of the integrals round the Cj ’s that bound the interior Dj ’s. In
the case of [8.32],
I I I I
f(z) dz = f(z) dz − f(z) dz + 2 f(z) dz . (8.22)
K C1 C2 C3
Consider, for example, the contour C3 , for which the counterclockwise sense hap-
pens to agree with the direction of K. On the other hand, C1 traverses this portion of
K in the opposite direction. Consequently, in (8.22), we end up integrating twice in
the correct direction, and once in the opposite direction; the net result is to integrate
P
[8.32] K = j νj Cj = C1 − C2 + 2 C3 in this example. [The explanation is in the next
figure.]
474 Complex Integration: Cauchy’s Theorem
P
[8.33] Proof that K = j νj Cj . Using the “crossing rule” (7.1) on page 388, we see
that νj = νk + 1. Thus the contour Cj in the direction of K will always occur precisely
one more time than the contour Ck in the opposite direction—the net result is that K is
traversed once in the correct direction.
along this part of K once in the correct direction. You should check for yourself that
all of K is correctly accounted for in this way. Substituting (8.21) into (8.22), we
have confirmed the prediction of the general theorem for this particular contour.
Since (8.22) is clearly true for any function f, we may abstract it away and write
the equation as
K = C1 − C2 + 2 C3 .
Notice that the coefficient of Cj in this sum is none other than the winding number
νj of K about Dj , and that we may therefore rewrite the previous equation as
X
K= νj Cj . (8.23)
j
From the above example, it is clear that to prove the general version of Cauchy’s
Theorem we need only show that (8.23) is true for any K.
Consider [8.33], which shows a portion of an arbitrary contour K sandwiched
between two of the regions (Dj and Dk ) into which it partitions the plane; also
shown is the counterclockwise sense of their boundaries (Cj and Ck ). Using the
“crossing rule” (7.1) on page 388, we deduce that νj = νk + 1. Thus, in (8.23),
we find that the contour Cj in the direction of K will always occur precisely one
more time than the contour Ck in the opposite direction—the net result is that K is
traversed once in the correct direction. Done.
As previously explained, in establishing (8.23) we have also deduced the General
Cauchy Theorem (8.20).
[8.34] If a closed contour can be shrunk to a point without crossing a singularity, the
integral round it vanishes. Hopf’s Degree Theorem tells us that this shrinking process is
possible if and only if the contour does not wind around any singularities.
Suppose that a contour can be deformed and shrunk down to a point without
ever crossing a singularity of an otherwise analytic function. By inequality (8.5),
the value of the integral will be zero at the end of this shrinking process. But by the
Deformation Theorem, the value of the integral remains constant throughout this
process. In other words,
If a closed contour can be shrunk to a point without crossing a singularity,
the integral round it vanishes. (8.24)
To wrap this up, we clearly need a way of recognizing when this shrinking pro-
cess is possible. For example, is it possible for the contour K in [8.31]? Figure [8.34]
shows that it is. Therefore (8.24) implies that the integral along K vanishes, in
agreement with the general theorem.
The two theorems are clearly very closely related. In fact we can now deduce the
General Cauchy Theorem from (8.24) by observing that the winding number of the
final shrunken loop vanishes; Hopf’s Degree Theorem then tells us that
The shrinking process in (8.24) is possible if and only if the contour does not
wind around any singularities.
H P
[8.35] The General Residue Theorem: KHf(z) dz = 2πi j ν(K, H sj ) Res [f(z), sj ]. We
P P
know that K = j νj Cj , and therefore K f(z) dz = j νj Cj f(z) dz. The Deforma-
H
tion Theorem then tells us that each Cj f(z) dz can be expressed as a sum of integrals
around the illustrated
H loops σj encircling the singularities within each Cj . Finally, each
such integral σk f(z) dz = 2πi Res [f(z), sk ].
By virtue of the Deformation Theorem (8.11), we know that the integral Ij has a
characteristic value that does not depend on the size or shape of σj . Furthermore,
as we saw in [8.20], if a simple loop contains several singularities, then the inte-
gral round that loop is the sum of I-values of the singularities it contains. In our
example,
I I
f(z) dz = I1 + I2 + I5 and f(z) dz = I3 + I4 .
C1 C2
in which each I-value has been multiplied by the number of times K winds round the
corresponding singularity. Since (8.23) is valid for arbitrary loops it follows that this
conclusion is too. As the grand finale to this chapter, we have thus obtained the
following completely general formula for the loop integral of an analytic function:
I X
f(z) dz = ν(K, sj ) Ij .
K j
The final icing on the cake is an efficient method of computing the Ij ’s. In the
next chapter we will verify our previous claim that in the neighbourhood of each
sj there exists a unique Laurent series [see page 456], the coefficient of the complex
inversion term being (by definition) the residue Res [f(z), sj ]. Granted this, we see
that Ij = 2πi Res [f(z), sj ]. Thus
I X
f(z) dz = 2πi ν(K, sj ) Res [f(z), sj ]. (8.26)
K j
This is the General Residue Theorem. Note that it contains the General Cauchy
Theorem as the special case in which each ν(K, sj ) = 0.
We will also see in the next chapter that it is possible to find the residues in this
formula directly, without going to the trouble of finding the whole Laurent series.
Thus, even before exemplifying its use, it should be clear that in (8.26) we have a
result of great practical and theoretical power.
478 Complex Integration: Cauchy’s Theorem
8.13 Exercises
(i) With the centre of the fixed coin at the origin, show that the epicycloid can
be represented parametrically as
h i
z(t) = B (n + 1) eit − ei(n+1)t .
(ii) By evaluating the integral in (8.8) parametrically, show that
area of epicycloid = πB2 (n + 1) (n + 2) .
8 The figure below shows four simple loops, and in each case we have indi-
cated how much shaded area is enclosed. Use parametric evaluation to verify
equation (8.8) for each of the four loops.
9 What is the generalization of (8.8) to the case where the contour is not closed?
10 Use (8.23) to verify (8.9).
11 The perfect symmetry of figure [8.18] results from integration round the unit
circle. Roughly how would this figure look if we instead used a somewhat
larger circle?
12 Let K be the contour in [8.21].
(i) Evaluate the following integral by factoring the denominator and putting
the integrand into partial fractions:
I
z
dz .
z2 − iz − 1 − i
K
(ii) Write down the Laurent series (centred at the origin) for (cos z/z11 ). Hence
find
I
cos z
dz .
z11
K
13 This exercise illustrates how one type of difficult real integral may be evaluated
easily using a complex integral.
Let L be the straight contour along the real axis from −R to +R, and let J be
the semi-circular contour (in the upper half plane) back from +R to −R. The
complete contour L + J is thus a closed loop.
480 Complex Integration: Cauchy’s Theorem
(i) Using the partial fraction idea of the previous exercise, show that the
integral
I
dz
4
(z + 1)
L+J
(ii) Equate the answer with the one obtained by parametric evaluation along
the straight contour from 0 to (a + ib), and deduce that
Z1
a (ea cos b − 1) + b ea sin b
eax cos bx dx = ,
0 a2 + b 2
and
Z1
b (1 − ea cos b) + a ea sin b
eax sin bx dx = .
0 a2 + b 2
(iii) Prove the results in (ii) by ordinary methods.
Exercises 481
16 (i) Show that when integrating a product of analytic functions, we may use
the ordinary method of “integration by parts”.
(ii) Let L be a contour from the real number −θ to +θ. Show that
Z
z eiz dz = 2i (sin θ − θ cos θ),
L
and verify this by taking L to be a line-segment and integrating parametri-
cally.
17 Let
n
1 1
f(z) = z+ ,
z z
where n is a positive integer.
(i) Use the Binomial Theorem to find the residue of f at the origin when n is
even and when n is odd.
(ii) If n is odd, what is the value of the integral of f round any loop?
(iii) If n = 2m is even and C is a simple loop winding once round the origin,
deduce from part (i) that
I
(2m)!
f(z) dz = 2πi .
(m!)2
C
(iv) By taking C to be the unit circle, deduce the following result due to Wallis:
Z 2π
(2m)!
cos2m θ dθ = 2m−1 π.
0 2 (m!)2
(v) Similarly, by considering functions of the form zk f(z) where k is an
integer, evaluate
Z 2π Z 2π
cos θ · cos kθ dθ and
n
cosn θ · sin kθ dθ
0 0
18 Let E be the elliptical orbit z(t) = a cos t + ib sin t, where a and b are positive
and t varies from 0 to 2π. By considering the integral of (1/z) round E, show
that
Z 2π
dt 2π
2 cos2 t + b2 sin2 t
= .
0 a ab
19 Let us verify the claim of Ex. 19, p. 297, that if a function has vanishing
Schwarzian derivative, then it must be a Möbius transformation. Following
Beardon (1984), suppose that {f(z), z} = 0, and define F ≡ (f ′′ /f ′ ).
(i) Show that 1/F(z) − 1/F(w) = −(z − w)/2.
(ii) Deduce that d
dz log f ′ (z) = −2/(z − a), for some constant a.
(iii) Perform two further integrations to conclude that f(z) is a Möbius trans-
formation.
482 Complex Integration: Cauchy’s Theorem
[The existence of such a series for any analytic function is derived in the next
chapter.]
(i) By integrating this series along K, show that the difference between the
1 ′′
exact integral and the RM value is roughly 24 f (a) ϵ3 . Verify that the
results of the first two parts of the previous exercise are in accord with
this finding.
(ii) Use the series to show that the complex number from the image of the
midpoint of K to the midpoint of the images of the ends of K is roughly
1 ′′ 2
4 f (a) ϵ . As ϵ shrinks, are these two types of midpoint distinguishable
under the magnifying lens that produces figure [8.28]?
(iii) From the Fundamental Theorem, deduce that the existence of such a series
implies the vanishing of the integral of f round loops within the disc of
convergence.
Exercises 483
23 Let f(z) be analytic throughout a region which contains a triangle with vertices
a, b, c, and hence with edges A ≡ (c − b), B ≡ (a − c), C ≡ (b − a). Given a
pair of point p and q, let us define wpq as a kind of average of f(z) along the
line-segment pq:
Zq
1
wpq ≡ f(z) dz.
(q − p) p
Show that this complex average mapping sends the sides of the triangle abc
to the vertices wab , wbc , wca of a similar triangle! We merely rediscovered this
result, which is apparently due to Echols (1923).
[Hint: Show that Awbc + Bwca + Cwab = 0, and use A + B + C = 0.]
24 Let K be a closed contour, and let ν be its winding number about the point a.
Show that
I z
e
dz = 2πi ν ea .
z−a
K
(z−a)
z
[Hint: Write e as e e a
, and expand e(z−a) as a power series.] This is a
special case of Cauchy’s Integral Formula (explained in the next chapter), which
states that if f is analytic inside K, then
I
f(z)
dz = 2πi ν f(a) .
(z − a)
K
25 Consider the image of the disc |z| ⩽ R under the mapping z ÞÑ k zm . As
the radius sweeps round the disc once, its image sweeps m times round the
image disc of radius |k| Rm . Thus we may sensibly define the area of the image
to be m π (|k| Rm )2 . With this understanding, show that if a mapping has a
convergent power series
f(z) = a + b z + c z2 + d z3 + · · · ,
then the area of the image is just the sum of the areas of the images under each
of the separate terms of the series:
area of image = π (|b|2 R2 + 2 |c|2 R4 + 3 |d|2 R6 + · · · ) .
This is Bieberbach’s Area Theorem.
Hint: Recall that the local area expansion factor is |f ′ |2 , so the image area is
ZZ Z R Z 2π
′2 ′ ′
|f | dx dy = iθ iθ
f (r e ) f (r e ) dθ r dr.
|z|⩽R 0 0
26 (i) Show that if f is an analytic function without singularities or p -points on
a loop L, then
I
1 f ′ (z)
ν [f(L), p ] = dz.
2πi f(z) − p
L
484 Complex Integration: Cauchy’s Theorem
9.1.1 Introduction
One of the principal objectives of this brief chapter is to tie up various loose ends
from previous chapters. In particular, we have previously claimed (but have not
yet explained) three important properties of an analytic function f(z):
The classical explanation1 of these facts hinges on the following result. If f(z) is
analytic on and inside a simple loop L, and if a is a point inside L, then
I
1 f(z)
dz = f(a). (9.1)
2πi z − a
L
This is called Cauchy’s Formula—it constitutes the precise statement of the “rigidity”
of analytic functions that we depicted in [5.3], p. 249. That is, the formula says that
the values of f on L rigidly determine its values everywhere inside L.
We will give two explanations of (9.1), both of which are firmly rooted in
Cauchy’s Theorem.
1
In the late 1950s a new approach was developed using topological ideas like those in Chapter 7,
and it was our original intention to employ that approach here. However, having lacked both the time
and the imagination to reduce the idea to its visual essentials, we have reluctantly fallen back on an
integral-based approach. For more on the topological approach, see Whyburn (1955), Whyburn (2015)
and Beardon (1979).
Visual Complex Analysis. 25th Anniversary Edition. Tristan Needham, Oxford University Press.
© Tristan Needham (2023). DOI: 10.1093/oso/9780192868916.003.0009
486 Cauchy’s Formula and Its Applications
The virtue of this transformation is that the integral round Cr turns out to have a
simple and helpful interpretation.
First recall that the average value ⟨f⟩Cr of f(zθ ) as zθ = a + r eiθ travels round
the circle Cr is defined by
Z
1 2π
⟨f⟩Cr ≡ f(zθ ) dθ.
2π 0
In the previous chapter we saw geometrically that if θ increases by dθ, causing
zθ to move dz along the circle, then dz/(z − a) = i dθ. Substituting this into (9.2),
[9.1] First explanation of Cauchy’s Formula. [a] If f(z) is analytic inside L then so is
f(z)/(z − a), except at z = a, so the Deformation Theorem implies that we can deform L
into the circle Cr of radius r without changing the value of the integral. But on Cr we can
interpret the integral as the average value ⟨f⟩Cr of f(z) on Cr , and, by the Deformation
Theorem again, this average value is independent of the radius. If zj are n equally spaced
points around Cr then ⟨f⟩Cr can be visualized as the limit as n Ñ ∞ of the centroid of
the image points f(zj ). [b] As Cr shrinks towards a, these image points cluster more and
H f(z)
L z−a dz = ⟨f⟩Cr = f(a).
1
more tightly around f(a), so 2πi
Cauchy’s Formula 487
we find that the original integral round L may be interpreted as the average value
of f on any of the circles Cr :
I
1 f(z)
dz = ⟨f⟩Cr .
2πi z − a
L
“inside” = { p | ν[L, p] ̸= 0 }.
Suppose in [9.2] that f has no singularities “inside” L. Then the only interior
singularity of [f(z)/(z − a)] will be the one at z = a. Here, L winds round a twice,
and it is clear that L may be deformed into a small circle centred at a and traversed
ν[L, a] = 2 times. By virtue of Cauchy’s Formula for simple loops, we deduce that
1
H f(z)
2πi L z−a dz = 2 f(a).
More generally, this line of reasoning suggests the following General Cauchy
Formula: If f(z) is analytic on and “inside” a general loop L, then
I
1 f(z)
dz = ν[L, a] f(a). (9.3)
2πi z − a
L
That this is always true is not quite clear from the above line of reasoning. Certainly
Hopf’s Theorem [(7.2), p. 389] guarantees that without crossing the singularity at
a, L may be deformed into a circle centred at a and traversed ν[L, a] times. But the
singularities of f may be scattered in the midst of L, although (by assumption) none
lie “inside” L. So is it clear that this deformation can always be performed without
crossing any of these singularities?
1
H f(z)
[9.2] Second explanation and the General Cauchy Formula: 2πi L z−a dz = ν[L, a] f(a).
Define Fa (z) to be the non-infinitesimal analogue of the amplitwist, expanding and
e so that V
rotating V to its “image” V, e = Fa (z) V. Indeed, as V vanishes, the analogy
becomes exact: Fa (a) ≡ limzÑa Fa (z) = f ′ (a). Since Fa (z) = [f(z) − f(a)]/[z − a]
is analytic inside L, its integral
H f(z) H dz vanishes, by Cauchy’s Theorem, and therefore
H f(z)
1 1 1
0 = 2πi L z−a dz − f(a) 2πi L z−a = 2πi L z−a dz − ν[L, a] f(a), proving the result.
Infinite Differentiability and Taylor Series 489
We encourage you to pursue this idea, but we shall now present a different
approach which yields (9.3) cleanly and directly. Consider the mapping z ÞÑ e
z=
f(z) in [9.2], and let us define
f(z) − f(a) e e
z−a
Fa (z) ≡ = .
z−a z−a
If V ≡ (z − a) is pictured as a vector emanating from a, and V e ≡ (ez−a e) is pic-
tured as its image emanating from a e, then Fa (z) describes the amount of rotation
e
and expansion that carries V into V = Fa (z) V. Thus Fa (z) is the non-infinitesimal
analogue of the amplitwist f ′ (a) that carries an infinitesimal vector ξ into its image
e
ξ = f ′ (a) ξ, and
Fa (a) ≡ lim Fa (z) = f ′ (a).
zÑa
In other words,
I I
1 f(z) 1 dz
0 = dz − f(a)
2πi z−a 2πi z−a
L L
I
1 f(z)
= dz − ν[L, a] f(a),
2πi z−a
L
as was to be shown.
Our first step is to obtain a neat expression for f ′ (a) in terms of the values of f(z)
on L. Applying Cauchy’s Formula to the analytic function Fa (z), we deduce that
I
′ 1 Fa (z)
f (a) = Fa (a) = dz
2πi z − a
L
I I
1 f(z) f(a) dz
= 2
dz − .
2πi (z − a) 2πi (z − a)2
L L
Now let’s use this to find the image e z = f ′ (z) of a short vector ξ
ξ under z ÞÑ e
emanating from a. Ignoring a term proportional to ξ2 , we find [exercise] that
I
e 2 f(z)
ξ ≡ f ′ (a + ξ) − f ′ (a) = dz ξ.
2πi [z − (a + ξ)]2 (z − a)
L
both (9.4) and (9.5) are precisely what we would get if we simply differentiated the
formula
I
1 f(z)
f(a) = dz
2πi z − a
L
with respect to a. Continuing in this way, we are led to conjecture that the nth
derivative f(n) may be represented as
I
(n) n! f(z)
f (a) = dz. (9.6)
2πi (z − a)n+1
L
f(z) = c0 + c1 z + c2 z2 + c3 z3 + · · · .
f(n) (0)
cn = , (9.7)
n!
so the power series is actually a Taylor series, and the coefficients do not depend
on R:
See [9.3]. Since z is inside the circle on which Z lies, |z| < |Z| = R, and |(z/Z)| < 1.
1
Thus 1−(z/Z) may be viewed as the sum of an infinite geometric series, and
I
1 f(Z)
f(z) = 1 + (z/Z) + (z/Z)2 + (z/Z)3 + · · · dZ.
2πi Z
C
Provided it makes sense to integrate this infinite series term by term, we deduce
that f(z) can indeed be expressed as a power series:
∞
X I
n 1 f(Z)
f(z) = cn z where cn = dZ. (9.8)
2πi Zn+1
n=0 C
Since
1 (z/Z)N
− [1 + (z/Z) + (z/Z)2 + · · · + (z/Z)N−1 ] = ,
1 − (z/Z) 1 − (z/Z)
it follows that
I
1 (z/Z)N f(Z)
f(z) − fN (z) = dZ.
2πi (Z − z)
C
Finally recall [see (8.5), p. 440] that the modulus of an integral cannot exceed the
product of the length of the path and the maximum modulus of the integrand at
points on the path. If M stands for the maximum value of |f(Z)/(Z − z)| on C, then
it follows that
One final point. We chose the origin as the centre of the expansion in order to
avoid algebraic clutter, but this choice really involves no loss of generality. For sup-
pose we instead choose the centre to be at a, meaning that we wish to expand f(z)
in powers of ξ ≡ (z−a). If f(z) is analytic at a then F(ξ) ≡ f(a+ξ) = f(z) is analytic
at the origin of the ξ-plane, and so it possesses an origin-centred Taylor expansion,
∞
X ∞
X
F(n) (0) n f(n) (a)
F(ξ) = ξ =⇒ f(z) = (z − a)n .
n! n!
n=0 n=0
c0 c1 cm−1
f(z) = + + ··· + + cm + cm+1 (z − a) + · · · .
(z − a)m (z − a)m−1 (z − a)
Recall that the coefficient of 1/(z − a) is called the “residue” of f(z) at a, denoted
Res [f, a]. Also recall the crucial significance of the residue in evaluating integrals:
if L is a simple loop containing a but no other singularities of f, then
I
f(z) dz = 2πi Res [f, a].
L
More generally, suppose that L is not required to be simple, and that f(z) has several
poles, at a1 , a2 , etc. The existence of the Laurent series was the missing ingredient
in our discussion of this situation in the previous chapter. Having established that
f does indeed possess a Laurent expansion in the vicinity of each of its poles, we
have also verified the General Residue Theorem [(8.26), p. 477]:
I X
f(z) dz = 2πi ν[L, an ] Res [f, an ]. (9.9)
n
L
From this general result one can derive other results that speed up the calculation
of residues in commonly encountered special cases. For example, suppose that f =
(P/Q) has a “simple” (i.e., order 1) pole at a as a result of Q having a simple root
at that point. In that case,
P(z)
Res [f(z), a] = lim (z − a)f(z) = lim h i.
zÑa zÑa Q(z)−Q(a)
z−a
P(z)
Thus, If f(z) = , and a is a simple root of Q, then
Q(z)
P(a)
Res [f(z), a] = ′ . (9.11)
Q (a)
Calculus of Residues 495
For example, consider f(z) = ez /(z4 − 1), which has simple poles at z = ±1, ±i.
If L is the circle |z − 1| = 1 then z = 1 is the only pole inside L, so (9.11) yields
I
ez ez
dz = 2πi Res [f, 1] = 2πi = 12 πie.
z4 − 1 4z3 z=1
L
(z4 − 1) = (z − 1)(1 + z + z2 + z3 ),
we may write f(z) = F(z)/(z − 1), where F(z) ≡ ez /(1 + z + z2 + z3 ). Since F(z) is
analytic inside L,
I I
ez F(z)
dz = dz = 2πi F(1) = 12 πie,
z4 − 1 z−1
L L
just as before.
2
Here, “today” refers to 1997! Today, in 2022, my claim is only more true than it was then.
496 Cauchy’s Formula and Its Applications
I
d 1 −2 π
f(z) dz = 2πi Res [f, i] = 2πi = 2πi = .
dz (z + i)2 z=i (2i) 3 2
L+J
But
I Z +R Z
dx
f(z) dz = + f(z) dz,
−R (x + 1)2
2
J
L+J
and, as you showed in the original exercise, the integral along J tends to zero as R
tends to infinity. Thus
Z +∞
dx π
2 2
= .
−∞ (x + 1) 2
The famous physicist Richard Feynman once bet3 his colleagues, “I can do by
other methods any integral anybody else needs contour integration to do.” It is a
tribute to complex analysis that Feynman lost this bet. Nevertheless, we can check
the above integral using a trick that frequently did enable Feynman to dispense
with residues: differentiation of a simpler integral with respect to a parameter.
Consider the elementary result,
Z +∞ +∞ π
dx 1 −1 x
2 2
= tan = .
−∞ x + a a a −∞ a
Differentiating this with respect to a yields
Z +∞
2a π
2 + a2 )2
dx = 2 ,
−∞ (x a
and substituting a = 1 then confirms our residue calculation.
For our second example, we will evaluate
Z 2π
dθ
I≡ , a > 1,
0 cos θ+a
by rewriting it as a contour integral round the unit circle C. See [9.4b]. As illustrated,
cos θ is the midpoint of z and (1/z), and dz is perpendicular to z and has length dθ:
in symbols, cos θ = 21 [z + (1/z)] and dz = iz dθ. Substituting into I,
I I
(dz/iz) dz
I= 1 = −2i 2 + 2az + 1
.
2 [z + (1/z)] + a z
C C
Since the singularities p and q of the integrand satisfy pq = 1, only one of them
lies inside C—in fact p and q are geometric inverses. Thus [exercise],
1 4π 2π
I = 4π Res ,q = =√ .
(z − p)(z − q) (q − p) a2 − 1
3
See Feynman (1997).
Calculus of Residues 497
R+∞
[9.4] Using residues to evaluate real integrals. [a] To evaluate (x2 + 1)−2 dx,
H −∞
consider the integral of f(z) = (z2 + 1)−2 . Then, L+J f(z) dz = 2πi Res [f, i] = π2 .
H R+∞
But as R Ñ ∞, J f(z) dz Ñ 0, and therefore −∞ (x2 + 1)−2 dx = π2 . [b] To evaluate
R2π
I ≡ 0 cosdθ iθ 1
θ+aH, where a > 1, let z = e , so that cos θ = 2 [z + (1/z)] and dz = izdθ.
dz
Then I = −2i C z2 +2az+1 . Factorizing the denominator into (z −p)(z −q), we obtain
h i
1 4π
I = 4π Res (z−p)(z−q) , q = (q−p) = √a2π
2 −1
.
series. We just want the (1/z) term of g, which will come from the z term of cot πz,
so that’s as far as we need go:
h i
(πz)2
cos πz 1 − 2! + · · ·
cot πz = =h i
sin πz πz − (πz) 3
+ ··· 3!
−1
1 (πz)2 (πz)2
= 1− + ··· 1 − + ···
πz 2 6
1 (πz)2 (πz)2
= 1− + ··· 1 + + ···
πz 2 6
1 πz
= − + ··· .
πz 3
In particular, note for future use that Res [cot(πz), 0] = (1/π).
Returning to the original function g, we find that
1 π
g(z) = 3
− + ··· ,
πz 3z
and so the origin is a triple pole with Res [g, 0] = −(π/3). Again, try checking this
using formula (9.10) instead.
Continuing with this example, it’s clear that g(z) also has a singularity at each
integer n. To find the residue at n, we could write z = n + ξ and expand g as a
Laurent series in powers of ξ. However, this is unnecessary. Since (1/z2 ) is non-
singular at n, and since cot[π(n + ξ)] = cot πξ,
Res [(1/z2 ) cot(πz), n] = (1/n2 ) Res [cot(πz), n]
= (1/n2 ) Res [cot(πz), 0]
= 1/(πn2 ).
More generally, note that if f(z) is any analytic function that is non-singular at n,
then
1
Res [f(z) cot(πz), n] = f(n). (9.12)
π
This may also be verified [exercise] using (9.11).
4
See Ex. 13 and Stillwell (2010).
Calculus of Residues 499
P∞ 1
[9.5] Evaluation of −∞ h using iRes [f(z)cot (πz), n] = π f(n). If we
f(n)
take f(z) = 1/z2 , then Res cot(πz)z2
, n = π 1n2 . Also, one finds [see text] that
h i H cot(πz) P
Res cot(πz)
z2
1
, 0 = − π3 . Therefore, 2πi S z2
dz = − π3 + π2 N 1
n=1 n2 . But as
N Ñ ∞, one finds [see text] that this integral goes to zero, and we thereby obtain
P
Euler’s extraordinary discovery of 1734: ∞ 1 π2
n=1 n2 = 6 .
Today such results can be derived in a systematic way using residues. Recon-
sider the function g(z) = (1/z2 ) cot(πz) above. With N a positive integer, let S be
the origin-centred square with vertices (N + 21 )(±1 ± i) shown in [9.5]. Adding up
the residues of the illustrated singularities inside S,
I X−1 X
N
1
g(z) dz = Res [g(z), 0] + Res [g(z), n] + Res [g(z), n]
2πi
S n=−N n=1
π 2X 1
N
= − + .
3 π n2
n=1
As we will now see, the integral on the LHS tends to zero as N tends to infinity,
and from this fact we immediately deduce Euler’s result.
To show that the integral of g(z) = (1/z2 ) cot(πz) does indeed tend to zero as
S expands, we must show that the size of the integrand dies away faster than the
perimeter (8N + 4) of S grows. First the easy part: |g(z)| = 1/z2 · | cot(πz)|, and on
S we clearly have |z| > N, so 1/z2 < (1/N2 ).
Next we must examine the size of
eiπz + e−iπz
| cot(πz)| = iπz
e − e−iπz
on the four edges of S. We begin with the horizontal edges, y = ±(N + 12 ). Since
e±iπz = e∓πy , it is not hard to see [exercise] that if N is reasonably large then
| cot(πz)| is very close to 1. Thus for sufficiently large N, | cot(πz)| will certainly be
less than 2, for example.
500 Cauchy’s Formula and Its Applications
Finally, on the vertical edges we have z = ±(N + 21 ) + iy, and it follows [exercise]
that
1 − e−2πy
| cot(πz)| = ⩽ 1.
1 + e−2πy
For sufficiently large N, we have established that | cot(πz)| < 2 everywhere on
S, so by virtue of (8.5), p. 440,
I
2
g(z) dz ⩽ (Max |g| on S)(perimeter of S) < (8N + 4).
N2
S
∞
X X
f(n) = −π Res [f(z) cot(πz)]. (9.13)
n=−∞ poles of f(z)
Of course if any of the poles of f(z) happen to be integers, then these values of n
are understood to be excluded from the LHS of (9.13).
P
Note that while symmetry enables us to calculate sums like ∞ (1/n2 ) and
P∞ 4
P∞
n=1
3
n=1 (1/n ) using (9.13), we cannot use (9.13) to calculate a sum like n=1 (1/n ).
What, you might ask, is the sum of this last series? The answer is that nobody knows!5
5
This was written in 1997, but, 25 years later, it remains a mystery.
Annular Laurent Series 501
9.4.1 An Example
We have seen that the Laurent series is the natural generalization of the Taylor
series when the centre of the expansion is a pole rather than a non-singular point.
However, this is by no means the only situation in which Laurent series are needed.
For example, consider
1
F(z) = ,
(1 − z)(2 − z)
whose simple poles are illustrated in [9.6a]. Since F is analytic within the unit
disc, it possesses a Taylor series in powers of z. This may be found most easily by
splitting F into partial fractions:
1 1
F(z) = −
(1 − z) (2 − z)
for |z|<1 for |z|<2
z }| { z }| {
X∞ ∞
X
1 1
= − = zn − 12 (z/2)n
(1 − z) 2[1 − (z/2)]
n=0 n=0
n
2 + 4 z + 8 z + · · · + 1 − (1/2) z + ··· , for |z| < 1.
1 3 7 2 n+1
=
502 Cauchy’s Formula and Its Applications
The pole at z = 1 means that outside the unit disc F cannot be expressed as a
power series in z. However, in the shaded annulus 1 < |z| < 2 it can be expressed
as a Laurent series in z:
for |z|>1 for |z|<2
z }| { z }| {
X∞ ∞
X
1 1
F(z) = − − =− (1/z)n+1 − 12 (z/2)n
z[1 − (1/z)] 2[1 − (z/2)]
n=0 n=0
2 3
1 1 1 1 z z z
= ··· − 3
− 2− − − − − − ··· , for 1 < |z| < 2.
z z z 2 4 8 16
Finally, in the region |z| > 2 beyond the annulus we obtain [exercise] a different
Laurent series:
1 3 (2n−1 − 1)
F(z) = 2
+ 3 + ··· + + ··· , for |z| > 2.
z z zn
1
[9.6] Laurent Series. [a] Let F(z) = (1−z)(2−z) . Then F(z) has a Taylor series within the unit
disc, but the singularity at z = 1 means that no power series can exist beyond. However, in
the shaded annulus, 1 < |z| < 2, it is possible to express F(z) as a so-called Laurent series
2 3
that has positive and negative powers of z: F(z) = · · ·− z13 − z12 − z1 − 12 − z4 − z8 − z16 −· · · .
Moving outside the annulus, a Taylor series is again impossible because of the two poles
closer in, but, here too, a Laurent series exists. However, it is now a different Laurent series:
n−1
F(z) = z12 + z33 + · · · + (2 zn−1) + · · · . [b] Laurent’s Theorem. If f(z) is analytic everywhere
P H
within the annulus, then f(z) = ∞ n 1 f(Z)
n=−∞ cn (z − a) , where cn = 2πi K (Z−a)n+1 dZ.
Annular Laurent Series 503
∞
X I
1 f(Z)
f(z) = cn (z − a)n , where cn = dZ. (9.14)
n=−∞
2πi (Z − a)n+1
K
Before establishing this result, which is called Laurent’s Theorem, we make the
following observations regarding its significance:
• The surprising thing about the result is the existence of a Laurent series, not the
fact that it converges in an annulus. Since we know that a power series in (z − a)
will converge inside a disc centred at a, it follows [exercise] that a power series
in 1/(z − a) will converge outside a disc centred at a. Since a Laurent series is (by
definition) the sum of a power series in (z − a) and a power series in 1/(z − a),
it follows that it will converge in an annulus.
• Previously we were able to deduce the existence of a Laurent series only in the
vicinity of a pole. The present result is much more powerful: as indicated by
the question marks in [9.6b], we make no assumptions at all concerning the
behaviour of f(z) in the disc D bounded by the inner edge of the annulus. In
practice, the outer edge of the annulus may be expanded until it hits a singu-
larity s of f(z), and the inner edge may likewise be contracted until it hits the
outermost singularity lying in D.
• If there are no singularities in D, then the inner edge of the annulus may be
completely collapsed, thereby transforming the annulus into a disc. In this case,
(9.14) does not contain any negative powers. For if n is negative then f(z)/(z −
a)n+1 is analytic everywhere inside K, and so cn = 0. In this way we recover the
existence of Taylor’s series as a special case of Laurent’s Theorem.
• Suppose that a is a singularity and that for a sufficiently small value of ϵ there
are no other singularities within a distance ϵ of a. In this case one says that a is an
isolated singularity of f(z). Applying Laurent’s Theorem to the annulus 0 < |z −
a| < ϵ, we find that there are just two fundamentally different possibilities: the
principal part of the Laurent series either has finitely many terms, or infinitely
many terms. Recall that in the latter case we have (by definition) an “essential
singularity”. See p. 417, where we considered the example
1 1 1
e1/z = 1 + + 2
+ + ··· .
1! z 2! z 3! z3
To sum up,
Now let us establish (9.14). In order to simplify the calculations, we will only
treat the case a = 0, illustrated in [9.7a]. Here, z is a general point in the annulus, C
and D are counterclockwise circles such that z lies between them, and L is a simple
loop round z, lying within the annulus.
504 Cauchy’s Formula and Its Applications
[9.7] Proof of Laurent’s Theorem. [a] The initial key step mimics theH rewriting of
1 f(V)
Cauchy’s Formula that was used to prove Taylor’s Theorem: f(z) = 2πi L V−z dV =
H h i H h i
1 f(Z) 1 1 f(W) 1
2πi C Z 1−(z/Z) dZ + 2πi D z 1−(W/z) dW. The theorem then follows
quickly by expanding both square brackets into geometric series, in z and (1/z), respec-
tively. [b] This picture justifies the previous equation by showing that L can indeed be
deformed within the annulus into (C) + (−D), without changing the value of the integral.
where the second equality follows from the fact that L may be deformed within
the annulus into (C) + (−D), as indicated in [9.7b].
Next, we rewrite the above equation as
I I
1 f(Z) 1 1 f(W) 1
f(z) = dZ + dW.
2πi Z 1 − (z/Z) 2πi z 1 − (W/z)
C D
The significance of this is that |(z/Z)| < 1 and |(W/z)| < 1, so both integrands on the
RHS can be expanded into geometric series, very much as we did in the example
of [9.6a].
Referring back to the derivation of the Taylor series (9.8), the integral round C
can be expressed as
I X∞ I
1 f(Z) 1 1 f(Z)
dZ = dZ zn .
2πi Z 1 − (z/Z) 2πi Zn+1
C n=0 C
Annular Laurent Series 505
9.5 Exercises
p
(iii) By taking K to be a circle of radius |z2 − 1| centred at z, deduce that
Z p
1 π
Pn (z) = (z + z2 − 1 cos θ)n dθ.
π 0
(iv) Check that this last formula yields the same P1 (z) and P2 (z) as you obtained
in part (i).
5 If C denotes the unit circle, show that
Z 2π I
sin2 θ i (z2 − 1)2 π
dθ = − 2
dz = .
0 5 − 4 cos θ 4 z (z − 2)(2z − 1) 4
C
6 Let f(z) be an analytic function with no poles on the real axis, and such that
|f(z)| < (const.)/|z|2 for sufficiently large |z|. By integrating f(z) eiz along the
contour (L + J) shown in [9.4a], deduce that
Z +∞ Z +∞ X
f(x) cos x dx + i f(x) sin x dx = 2πi Res [f(z) eiz ].
−∞ −∞ upper half-plane
(i) Use the Residue Theorem to show that if n = 2, 3, 4, . . ., and R > 1 (as
illustrated), then
I
dz 2πi i(π/n)
n
=− e .
1+z n
K
(ii) Show that
I h i Z∞
dz i(2π/n) dx
lim = 1 − e .
RÑ∞ 1 + zn 0 1 + xn
K
(iii) Deduce that (9.15) is indeed valid for odd n as well as even n.
P
10 Use (9.13) to show that ∞ 4 4
n=1 (1/n ) = (π /90).
11 Show that if f(z) is an analytic function such that |f(z)| < (const.)/|z|2 for
sufficiently large |z|, then
∞
X X
(−1)n f(n) = −π Res [f(z) cosec(πz)].
n=−∞ poles of f(z)
(iv) By integrating along any path from 0 to z that avoids integers, and then
exponentiating both sides of the resulting equation, deduce that
z2 z2 z2
sin z = z 1 − 2 1− 2 2 1 − 2 2 ··· .
π 2π 3π
[Hint: Recall that limzÑ0 (sin z/z) = 1.]
This famous formula was discovered by Euler in 1734, in a much simpler
way, via an extraordinary leap of imagination; see Stillwell (2010) for the
details.
(vi) Multiplying out the brackets on the right, we see that the z3 terms arise
from taking a z2 term from a single bracket, and 1’s from all the others.
But, on the left hand side, the power series for sin z is known. Thus,
1 3 1 1 1 1
z − z + · · · = z − 2 + 2 2 + 2 2 + 2 2 + · · · z3 + · · · .
3! π 2π 3π 4π
Euler was thereby able to sum a series that had baffled the greatest math-
ematicians that had come before him, making the wonderful and startling
discovery that
1 1 1 π2
1+ + + + · · · = .
22 32 42 6
CHAPTER 10
Visual Complex Analysis. 25th Anniversary Edition. Tristan Needham, Oxford University Press.
© Tristan Needham (2023). DOI: 10.1093/oso/9780192868916.003.0010
512 Vector Fields: Physics and Topology
[10.1] Vector field representation of f(z): Draw the complex number f(z) emanating
from z. [a] The vector field of f(z) = z2 . [b] The vector field of f(z) = (1/z).
[10.2] A vector field defines a mapping. [a] This vector field corresponds to the mapping
z ÞÑ (1/2)z. [b] This vector field corresponds to the mapping z ÞÑ (1/2)iz.
Consider the examples in [10.2a] and [10.2b]. If z lies on a circle of radius r then
the vector field in [10.2a] is radial and has length (r/2); in [10.2b] the vector field
has the same length but is tangential instead of radial. Check that when viewed as
mappings, [10.2a] corresponds to an expansion of the plane by (1/2), while [10.2b]
corresponds to the same expansion followed by (or preceded by) a rotation through
a right angle.
Vector Fields 513
If the vectors in [10.2a] were instead directed inwards, what would the corre-
sponding mapping be? If the vectors in [10.2b] were flowing clockwise, what would
the corresponding mapping be?
[10.3] The Phase Portrait of V(z) is the set of streamlines along which V flows: mathe-
matically, they are the curves that have V as their tangent vectors. Given such a picture,
one can immediately deduce the direction of V(z) at z, but its magnitude is invisible.
However, in many physical examples it is possible to draw the phase portrait in a spe-
cial way, such that the closer together the streamlines, the stronger the flow. This is the
case here, where electrodes are attached to a metal plate at A and B, and electricity flows
between them—you see that the spacing is smallest, and therefore the flow greatest, along
AB. Note that the streamlines are arcs of circles; this is proved geometrically in the next
figure.
Take two wires and connect them to a battery, then touch the ends to two points
A and B of a thin copper plate. Almost instantly a steady flow of electric current
from one electrode to the other will be set up in the plate. See [10.3]. At each point
z of the plate we now represent this flowing current by a time-independent vector
in the direction of the flow, and with a length equal to the strength of the current
there. Picturing the plate as a portion of C, the flow is thus expressed as a complex
function V(z).
Rather than drawing the actual vector field in [10.3] we have instead shown the
paths along which the electricity flows. Such a picture is called the phase portrait
of the vector field, and the directed curves along which the flow occurs are called
the integral curves or streamlines of the vector field. As illustrated, the streamlines
of this example are in fact arcs of circles connecting the two electrodes. We shall
justify this shortly.
Phase portraits are easy to take in visually and are thus a common way of rep-
resenting vector fields. By definition the vector field is everywhere tangent to the
streamlines, and thus its direction can be recovered from the phase portrait. On the
other hand, it would seem that a phase portrait would necessarily fail to include
the information about the lengths of the vectors. This is true in general, but for
Vector Fields 515
many vector fields that arise in physics it will be shown that there exists a special
way of drawing the phase portrait so that the strength of the flow is manifested
as the crowding together of the streamlines: the closer together the streamlines, the
stronger the flow1 . Later we shall explain this idea in detail, but for the moment we
remark that [10.3] has actually been drawn in this special way. For example, as we
approach the line-segment connecting the electrodes the streamlines become more
and more crowded together, corresponding to a stronger and stronger current.
1
Faraday was the first to conceive of vector fields in this way; Maxwell then rendered the idea
mathematically precise and exploited it to the hilt.
516 Vector Fields: Physics and Topology
2πr |V| = S.
Writing z = r eiθ , we find that the vector field of the source is therefore
2
This is an idealization—like the earth, the sun is somewhat flattened at the poles.
Vector Fields 517
[10.4] [a] A Source. Fluid is pumped in at a constant rate at 0 and travels outward sym-
metrically. Its velocity at z = r eiθ must be proportional to (1/r), for the total amount of
fluid crossing a circle of radius r per unit of time (the flux) will be this velocity multiplied
by 2πr. [b] The streamlines of a doublet are arcs of circles that pass through the source
and the sink. To prove that D = V⊕ + V⊖ is tangent to the illustrated arc of the circle ApB,
we must prove that • = ⊙. The angles ApB and pst are clearly equal, but we also have
(ts/ps) = |V⊕ |/|V⊖ | = Bp/Ap. So, the two shaded triangles are similar, and therefore
• = ⊙. Done.
S eiθ S 1
V(z) = |V| e iθ
= = .
2π r 2π z
[We note without proof that this is also the electric field on a plane at right angles
to a very long wire carrying a uniform charge of S per unit length.] The source in
[10.3] is at A instead of at the origin, and so it is described by
S 1
V⊕ (z) = .
2π z − A
and a sink of equal strength is called a doublet.] The answer is perhaps slightly
clearer if we switch to the equivalent electrostatic problem of parallel charged wires
through A and B. A unit charge at z is repelled by A with force V⊕ (z), and attracted
by B with force V⊖ (z). The net force D(z) of the doublet acting on the charge is then
simply the vector sum of the two separate forces:
S 1 1 S (A − B)
D(z) = V⊕ (z) + V⊖ (z) = − = . (10.1)
2π z − A z − B 2π (z − A)(z − B)
We will now show geometrically that, as claimed in [10.3], the net force at p is
tangent to the circle through A, p, and B. Consider [10.4b]. It is easy to see [exercise]
that D will be tangent to the circle if and only if the angles marked • and ⊙ are
equal, so this is what we must demonstrate. As illustrated, the angles ApB and pst
are clearly equal. But we also have
ts |V⊕ | Bp
= = .
ps |V⊖ | Ap
Thus the two shaded triangles are similar, and therefore • = ⊙, as was to be shown.
3
Otherwise known as critical points or singularities—terms to which we have already attached
different meanings.
Winding Numbers and Vector Fields* 519
[10.5] The values of the index (I) for various singular points. The geometric definition
of I is explained in the next figure.
[10.6] The index of a singular point is the net number of revolutions executed by the
vector field as we traverse a simple (counterclockwise) loop around the singular point.
Here, I = −1.
asymptotes occurring in the graph of a real function. Notice how easy it is to zoom
in on the vector field to find their precise locations. But now we can do much better
than the real case, for we can also immediately read off the precise nature of a root
or singularity simply looking at its index!
The following example illustrates the fact that a vector field contains more
information than an ordinary graph. If we sketch the graphs of
(x − 1)2 (x − 1)4
F(x) = and G(x) =
(x + 2)3 (x + 2)7
the results will be qualitatively the same: both look something like a parabola near
x = 1; both have branches going to opposite ends of the vertical asymptote at
x = −2; both look something like (1/x) for large x.
Now use the computer to draw the corresponding vector fields when x is
replaced by z. Striking indeed are the differences! As we traverse a small loop
around the root at z = 1, F makes two positive revolutions while G makes four;
doing the same at the pole z = −2, F makes three negative revolutions while
G makes seven; and on a very large origin-centred circle, F makes one negative
revolution while G makes three.
Returning to the general significance of (10.2), consider the ordinary winding
number ν [L, 0] of a loop L. This can now be viewed as the index of L with respect
to the vector field of the identity mapping:
ν [L, 0] = Iz [L].
Figure [10.7] illustrates this result with Iz [L] = 1. The winding number of L around
a general point a is likewise just its index with respect to the vector field (z − a):
ν [L, a] = I(z−a) [L].
[10.7] Index of the vector field V(z) = z is the winding number of the loop: ν [L, 0] =
Iz [L].
522 Vector Fields: Physics and Topology
[10.8] Poincaré’s method for quickly finding the index. [a] First, choose an arbitrary
direction, then find all the places zj on L where V(zj ) points in this direction. Count the
number P of zj ’s where V(z) is rotating in the positive, counterclockwise direction as we
pass through zj , and likewise count the number N of places where it is rotating in the
negative direction. Then, IV [L] = P−N. In the illustrated case, P = 2, N = 1, and so the
formula correctly predicts that IV [L] = 2 − 1 = 1. [b] Consider the image V(L) of L under
the mapping corresponding to the vector field V. On the ray in the chosen direction, start
far away at q, where the winding number ν = 0. The intersection points of the ray with
V(L) are the images V(zj ) of the zj ’s, and a positive rotation of V at zj corresponds to a
left-to-right crossing, so the crossing rule (7.1) increases ν by one. Likewise, a negative
rotation of V reduces ν by one. Thus we have proved Poincaré’s formula.
Winding Numbers and Vector Fields* 523
P Pn
[10.9] The Index Theorem: I [C] − g I [Bj ] = I [sj ]. This is certainly true in the
illustrated example, for I [C] − I [B1 ] − I [B2 ] = 2 − 0 − 1 = 1, and I [s1 ] + I [s2 ] =
2 + (−1) = 1.
[10.10] Proof of the Index Theorem. Using the dashed curves, break the region into
curvilinear polygons in such a way that each one contains at most one singular point,
and let their counterclockwise boundaries be Kj . If we sum the indices of all the Kj ’s
then we obtain the RHS of the Index Theorem. On the other hand, the index of a single
Kj is obtained by looking at how much the vector field rotates as one travels along each
edge of Kj , then adding up these net rotation angles. But when we sum the indices of all
the Kj ’s, each interior edge [dashed] is traversed twice, once in each direction, and the
associated angles of rotation therefore cancel. The remaining edges of the Kj ’s together
make up C and −B1 , −B2 , etc. Summing the associated angles of rotation (divided by
2π) therefore yields the LHS of the Index Theorem. Done.
Flows on Closed Surfaces* 525
I (vortex) + I (vortex) = 1 + 1 = 2,
I (dipole) = 2.
Try drawing your own streamlines on an orange, then sum the indices of the
singular points. Is this a coincidence??
There are no coincidences in mathematics! In the case of the sphere, the Poincaré–
Hopf Theorem states that if we sum the indices of any vector field on its sur-
face, we will always get 2 for the answer. Indeed, it says that we will get this
answer for any surface that is topologically a sphere, that is to say, any surface
into which the sphere may be changed by a continuous and invertible trans-
formation. If we imagine the sphere to be made of rubber, examples of such
transformations and surfaces are given by stretching without tearing. The surfaces
of the plum and the wineglass in [10.12a] are two examples of such topological
spheres.
The sphere is the boundary of a solid ball, and other closed surfaces may likewise
be obtained as the boundaries of other solid objects. For example, the surface of a
doughnut is called a torus (top of [10.12b]), and it is clear that this surface is topo-
logically the same as the beach toy at the bottom. But it seems equally clear that no
amount of stretching and bending can turn these surfaces into a sphere—[10.12a]
and [10.12b] are topologically distinct types of surface. Figure [10.12c] shows yet a
third topologically distinct class. Obviously we could continue this list indefinitely
just by adding more holes.
We shall not develop the topological ideas4 necessary to prove it, but once again
it seems clear that these classes of topologically distinct closed surfaces can be
classified purely on the basis of their number of holes. This number g is called
the genus of the surface (see [10.12]). We can now formulate the general result:
4
See “Further Reading”, at the end of this chapter.
Flows on Closed Surfaces* 527
[10.12] The genus (g) and the Euler characteristic (χ) of a closed surface. The genus
counts the number of holes in a closed surface. The Euler characteristic χ ≡ (2 − 2g)
turns out to be ubiquitous in topology, and it is therefore a more natural way of classifying
surfaces. Furthermore, χ in fact has its own, independent geometrical definition, and it is
then a theorem that χ = (2 − 2g). See Ex. 12.
The number χ ≡ (2 − 2g) occurring in this theorem is called the Euler characteris-
tic of the surface, and it turns out to be ubiquitous in topology. It is therefore more
natural to classify our surfaces using χ rather than g. See [10.12]. Furthermore, χ in
fact has its own, independent geometrical definition, and it is then a theorem that
χ = (2 − 2g). See Ex. 12.
An immediate consequence of (10.4) is that a vector field without any singular
points can exist only on surfaces of vanishing Euler characteristic, i.e., the topo-
logical doughnuts. Even then, the theorem does not actually guarantee that such a
vector field exists, it merely says that if there are singular points then their indices
must cancel. However, you can see for yourself [draw it!] that on a doughnut there
do exist at least two topologically distinct vector fields without any singular points:
one that circulates around its axis of symmetry, and one that circulates through the
hole.
net rotation of the vector field as the loop is traversed. But wait, rotation relative to
what?
To answer this question, we first re-examine the familiar concept of rotation in
the plane. Figure [10.13a] shows that (in the plane) the rotation of V(z) along L
can be thought of as taking place relative to a fiducial vector field having horizontal
streamlines, say U(z) = 1. If we define ∠UV to be the angle between U and V, and
let δL (∠UV) be the net change in this angle along L, then our old definition of the
index is
[10.13] Definition of the index of a singular point on a curved surface. [a] The original
definition of the index in the plane can be thought of as the net rotation of V relative to a
horizontal fiducial vector field U: IV [L] = 2π1
δL (∠UV). [b] The definition and value of
the index does not change if we replace U with any vector field that is nonsingular on and
inside L. [c] Now imagine that [b] is drawn on a rubber sheet. If we continuously stretch it
1
into the form of the curved surface then not only will 2π δL (∠UV) remain well-defined,
but its value will not change.
Flows on Closed Surfaces* 529
1
IV [L] = δL (∠UV). (10.5)
2π
If we continuously deform the horizontal streamlines of U in [10.13a] to produce
those in [10.13b] then, by the usual reasoning, the RHS of (10.5) will not change.
Thus we conclude that this formula yields the correct value of the index if we
replace U with any vector field that is nonsingular on and inside L.
Now imagine that [10.13b] is drawn on a rubber sheet. If we continuously stretch
it into the form of the curved surface in [10.13c] then not only will the RHS of (10.5)
remain well-defined, but its value will not change. To summarize: if s is a singular
point of a vector field V on a surface S, we define its index as follows. Draw any
nonsingular vector field U on a patch of S that covers s but no other singular points;
on this patch, draw a simple loop L going round s; finally, apply (10.5), that is count
the net revolutions of V relative to U as we traverse L.
1 X
= δKj (∠VW)
2π
all polygons
= 0,
530 Vector Fields: Physics and Topology
[10.14] Proof that the sum of the indices is the same for all vector fields. Let V (with
singular points vj marked •) and W (with singular points wj marked ⊙) be two dif-
ferent vector fields on the surface S. Divide up S into curvilinear polygons (dashed)
such that each one contains at most one vj and one wj . Now consider one of these
polygons and its boundary Kj . Draw any nonsingular vector
field U on the polygon.
Then IW [Kj ] − IV [Kj ] = 2π1
δKj (∠UW) − δKj (∠UV) = 2π 1
δKj (∠VW), which is
explicitly independent of the local vector field U. If we now sum this over all polygons,
P
each edge is traced twice, in opposite directions, so δKj (∠VW) = 0, and therefore
P P
IV [Kj ] = IW [Kj ]. Done.
because every edge of every polygon is traversed once in each direction, producing
equal and opposite changes in ∠VW. We have thus completed the first step: the
sum of the indices is independent of the vector field.
Since the index sum for the example in [10.11a] is 2, we now know that this
is the value for any vector field on a topological sphere. The second step of the
general argument is likewise to produce an example on a surface of arbitrary genus
g, such that the sum is χ = (2 − 2g). Figure [10.15] is such an example for g = 3, the
generalization to higher genus being obvious. Here we imagine that honey is being
poured onto the surface at the top—it then flows over the surface, finally streaming
off at the very bottom. As the figure explains, and as was required, the sum of the
indices is indeed equal to χ.
Further Reading. These topological ideas—in combination with ideas in the
next two chapters—open the door to the important subject of Riemann surfaces. In
particular, we hope you will find it easier to read Klein (1881), which champions
Flows on Closed Surfaces* 531
[10.15] A specific vector field on a surface of genus g for which the sum of the indices is
χ. We imagine that honey is being poured onto the surface at the top—it then flows over
the surface, finally streaming off at the very bottom. As the figure explains, and as was
required, the sum of the indices is indeed χ. But we know that this sum is the same for
all vector fields on the surface, and we have therefore shown that the sum of their indices
must always equal χ, completing the proof of the Poincaré–Hopf Theorem.
10.4 Exercises
1 Show both algebraically and geometrically that the streamlines of the vector
field z2 are circles that are tangent to the real axis at the origin. Explain why
the same must be true of the vector field 1/z2 .
2 Use a computer to draw the vector field of 1/(z sin2 z). Use this picture to
determine the location and order of each pole.
3 Use a computer to draw the vector field of
Use this picture to factorize P(z), and check your answer by multiplying out
the brackets.
4 Suppose that one of the streamlines of a vector field V is a simple closed loop
L. Explain why L must contain a singular point of V.
5 Find the index of each of the three singular points shown below.
(i) Verify that the index of each of the three singular points in the previous
question is correctly (and painlessly) predicted by Bendixson’s Formula:
I = 1 + 12 (e − h).
11 Imagine the surface of the unit sphere divided up into F polygons, the edges
all being “straight lines on the sphere”, i.e., great circles. Let E and V be the
total number of edges and vertices that result from dividing up the sphere in
this way.
(i) Let Pn be an n-gon on the unit sphere. Use (6.9), p. 317, to show that
A(Pn ) = [angle sum of Pn ] − (n − 2)π.
[Hint: Join the vertices of Pn to a point in its interior, thereby dividing it
into n triangles.]
(ii) By summing over all polygons, deduce that
F − E + V = 2.
[This argument is due to Legendre (1794); the result itself is a special case of
the result in the following exercise.]
12 Let S be a smooth closed surface of genus g, so that its Euler characteristic is
χ(S) = 2 − 2g. As in [10.14], let us divide S into F polygons, and let E and V be
the total number of edges and vertices, respectively.
(i) Draw a simple example on the surface of an orange and convince yourself
(by drawing it) that we may obtain a consistent flow over the entire surface
whose only singular points are (1) a source inside each of the F polygons;
(2) a simple saddle point on each of the E edges; (3) a sink at each of the
V vertices. We call this a Stiefel vector field, in honour of Hopf’s student,
Eduard Stiefel [1909–1978]. See Frankel (2012, §16.2b). For an example of
a Stiefel vector field, see the front cover of VDGF, Needham (2021).
(ii) By applying the Poincaré–Hopf Theorem to such a flow on the general
surface S, deduce the following remarkable result, which we call the Euler-
Lhuilier Formula:
F − E + V = χ(S).
NOTES: (1) This is actually the normal definition of χ, and it is then a theorem
that χ(S) = 2−2g. (2) Euler discovered the original version of this formula
in 1750 for topological spheres, with χ = 2; this is called Euler’s Polyhedral
Formula. For a masterful, mathematically accurate, yet riveting account of
this history and the connected mathematical ideas, see Richeson (2008).
(iii) Verify this result for your example in (i), then try it out on a doughnut.
13 The figure below shows all the normals that may drawn from the point p to
the smooth surface S. Let R(q) be the distance from p to a point q of S, and let
us say that q is a critical point of R if the rate of change of R vanishes as q begins
to move within S; we need not specify the direction in which q begins to move
because we are assuming that S has a tangent plane at q.
Exercises 535
[This lovely result is essentially due to Reech (1858), though he did not
express it in terms of χ(S), nor did he use an argument like the one above.
With hindsight, Reech’s work is a clear harbinger of Morse Theory, which
it predates by some 70 years.]
(vi) Verify Reech’s theorem for a couple of positions of p in the case where S is
a torus (doughnut).
CHAPTER 11
We promised long ago that there was a more vivid way of understanding complex
integrals than the geometric Riemann sum of Chapter 8. In this section we lay the
foundations for this elegant new approach. If you are already familiar with vector
calculus then you can skip this section and go directly to Section 11.2.
11.1.1 Flux
In order to define the flux a little more carefully than before, consider [11.1]. At each
point of the directed path K we introduce a unit tangent vector T in the direction
of the path, and a unit normal vector N pointing to our right as we travel along K.
In terms of the corresponding complex numbers, this convention amounts to
T = iN.
The figure also shows how a vector field X (which we will first think of as the
velocity of a fluid flowing over the plane) can be decomposed into tangential and
normal components:
·
X = (X T ) T + (X N) N. ·
Only the second of these components carries fluid across K, and the amount
flowing across an infinitesimal segment ds of the path in unit time (i.e. its flux) is
·
thus (X N) ds. This is a refinement over our previous definition in that the flux
now has a sign: it is positive or negative according as the flow is from left to right
or from right to left. The total flux F [X, K] of X across K is then the integral of the
fluxes across its elements:
Z
F [X, K] = (X N) ds.
K
·
Check for yourself that the flux satisfies
F [−X, K] = F [X, −K] = −F [X, K].
Visual Complex Analysis. 25th Anniversary Edition. Tristan Needham, Oxford University Press.
© Tristan Needham (2023). DOI: 10.1093/oso/9780192868916.003.0011
538 Vector Fields and Complex Integration
[11.1] Decomposition of a vector into its tangential and normal components. Note that
the normal is defined to go to the right of the direction of travel, so that T = iN.
[11.2] If we picture X as the velocity of a fluid, then the flux, F [X, K], of X across K
represents the net flow of fluid out of R per unit time. [a] Mathematically, the flux is the
integral of the outward normal component of X. [b] The flux can be visualized as the limit
of the sum of the illustrated signed areas: positive if the flow is outward, and negative if it
is inward.
The flux concept is further illustrated in [11.2] for the case where K is a simple
closed loop bounding the shaded region R. Figure [11.2a] shows the normal com-
ponents of X, the signed magnitudes of which we must integrate to obtain F [X, K].
Figure [11.2b] shows how we might make an estimate of this flux. We replace K
by a polygonal approximation with directed edges ∆j , and at the midpoint of each
one we draw the normal component of X. The flux is then approximately given by
the algebraic sum of the signed areas of the shaded rectangles. In this case there is
clearly more positive area than negative, so the flux is positive. As the ∆j ’s become
shorter and more numerous, the approximation of course gets better and better.
Flux and Work 539
In the case of the simple loop K in [11.2a] there is another interesting way of
looking at the flux:
F [X, K] = [fluid leaving R per unit time] − [fluid entering R per unit time].
Henceforth we will always take our fluid to be incompressible. Thus, provided there
are no sources or sinks in R, what flows into R must also flow out of R:
F [X, K] = 0.
Indeed, we may turn this around and define a flow to be sourceless in a region if
all simple loops in that region have vanishing flux. The simplest example of such
a flow without any (finite) sources or sinks is X = const. If the loop does contain
a source, for example, then incompressibility says that the flux equals the strength
of the source.
Although we will only concern ourselves with two-dimensional flows, we
should at least mention the concept of flux in three dimensions. If a fluid is flowing
through ordinary space, it no longer makes sense to speak of the flux across a curve,
but it does make sense to speak of the rate at which the fluid crosses a surface. If
N now stands for the normal to this surface, then the flux across an infinitesimal
·
element of area dA is once again given by (X N) dA. The total flux is then obtained
by integrating this quantity over the whole surface. Just as in two dimensions, the
incompressibility of a three-dimensional flow is equivalent to the statement that all
closed surfaces (that do not contain sources or sinks) have vanishing flux.
Lastly, we should point out that although the word “flux” is Latin for “flow”,
it is standard practice to retain this terminology when applying our mathematical
definition to any vector field X, regardless of whether it actually is the velocity of
a flowing substance. For example, the electric field represents a force, but one of
the four fundamental laws of electromagnetism says that we can think of it as an
incompressible flow in which positive and negative electric charges act as sources
and sinks, so that its flux through a closed surface in space equals the net charge
enclosed.
11.1.2 Work
So far we have only studied the normal component of X; we turn next to its tan-
gential component. To do so, let us now imagine that X is a force field rather than a
flow.
If a particle on which a force acts is displaced infinitesimally then we know from
elementary physics that the work done by the field (i.e. the energy it expends) is
the component of the force in the direction of displacement, times the distance
moved. Thus if the particle moves an amount ds along K then the work done by X is
·
(X T ) ds. As with flux, this definition contains a sign, the physical significance of
540 Vector Fields and Complex Integration
which we will explain shortly. If the particle is moved along the entire length of K,
the total work done by the field is then
Z
W [X, K] = (X T ) ds.
K
·
Figure [11.3a] illustrates the tangential components of X on K, the signed magni-
tudes of which we must add up to obtain W.
Just as for F, check that W satisfies
[11.3] [a] If we picture X as a force field, then the integral of the tangential component
is the work, W [X, K], done in moving a particle along K. [b] If a puck of mass m is
fired with speed vin into a frictionless channel following K, and it emerges with speed vout ,
then the conservation of energy implies that the work is the change in the kinetic energy:
W = 21 mv2out − 12 mv2in .
Flux and Work 541
This formula also gives clear meaning to the sign of the work: it is the sign of the
change in speed. Thus if W is positive the field expends energy speeding up the
puck and increasing its kinetic energy, while if W is negative then the puck has to
give up some of its kinetic energy in doing work against the field.
Next, imagine that we bend K round so that the ends almost join to form
a closed loop. When the puck travels along the corresponding channel it will
therefore emerge at essentially the same place that it entered. Suppose it were to
emerge with greater speed than it entered. Joining the ends of the channel together,
the puck would therefore go round the loop faster and faster, gaining energy
with each circuit—energy that could be harnessed to solve the world’s energy
crisis!
Although we may construct mathematical examples for which this happens, if
no energy is supplied from outside the puck/field system then a physical force
field will not behave in this way; it will conserve energy so that the puck returns to
its starting point with exactly the same speed with which it was launched1 . Such a
field is called conservative. Mathematically, X is conservative if and only if
Just as we applied the concept of flux to vector fields that were not flows, so we
may apply the concept of work to vector fields that do not represent force. How-
ever, in this general setting it is standard practice to call W [X, K] the circulation of
X along K rather than the work. As with “flux”, this terminology originates from
thinking of X as representing a flow. To see why, take K to be a closed loop and
consider the following thought-experiment of Feynman (1963). Imagine that the
fluid flowing over the plane with velocity X is instantaneously frozen everywhere
1
If energy can be supplied from outside the system, then the work need not vanish for a closed loop.
In fact the operation of all electrical machines depends on the ability of a moving magnet to create an
electric field that can speed up our puck. However, there is still no violation of energy-conservation
since work is being done to move the magnet. See Feynman (1963).
542 Vector Fields and Complex Integration
except within the narrow strip where our channel used to be. The “circulation” is
then [exercise] the speed with which the unfrozen fluid flows (or circulates) round
K, times the length of K.
If this circulation vanishes for every closed loop then the flow is said to be irro-
tational. Just as “circulation” means W [X, K], irrespective of the physical nature of
X, so with equal generality “irrotational” is short for the mathematical statement
(11.1). Thus a conservative force field could also be described as irrotational.
[11.4] Local flux and local work. Accurate, ultimately exact estimates of F and W around
a small, ultimately vanishing square can be found by evaluating X at the midpoints (a,
b, c, d) of the sides, then summing the appropriate components. Writing X = P + iQ, we
find that F [X, □ ] ≍ (Bx P + By Q)(area of □ ), and W [X, □ ] ≍ (Bx Q − By P)(area of □ ).
Flux and Work 543
This expression can be simplified by considering the formal dot product of the
gradient operator ∇ with the vector field:
! !
∇ X= · Bx
By
P
Q
· = Bx P + By Q.
·
This quantity ∇ X is called the divergence of X, and in terms of it we have
·
F [X, □ ] ≍ [∇ X(z)] (area of □ ). (11.3)
In the next section we will see that (11.3) is true if □ is replaced by an infinitesimal
loop of arbitrary shape. This important result explains the term “divergence”, for
·
it says that ∇ X is the local flux per unit area flowing away from z, i.e., diverging
from z. In future we will abbreviate “local flux per unit area” to “flux density”.
Repeating the above analysis for the work, we find [exercise]
The quantity ∇×X is called the curl of X. Geometrically, it measures the extent
to which X ‘curls around’ the point z. Physically, in terms of force fields, the above
result says that the curl is the local work per unit area, or work density. There
is also a vivid interpretation in terms of flows. If we drop a small disc of paper
onto the surface of the flowing liquid at z, in general it will not only start to move
(translate) along the streamline through z with speed |X(z)|, but it will also rotate
about its centre with some angular speed ω(z). It can be shown that the aspect of
X which determines the rate of rotation ω is none other than the curl:
1
ω(z) = 2 [∇×X(z)].
For this reason “curl” is sometimes denoted “rot”, which is short for “rotation”.
[11.5] Divergence and curl in geometric form. Let P be the illustrated orthogonal
trajectory of the streamlines, let p measure distance along it, and let s denote dis-
tance along the streamline S. Also, let κP and κS denote the curvatures of P and S,
·
respectively. Then ∇ X = Bs |X| + κP |X| and ∇×X = −Bp |X| + κS |X|.
considering the flux out of an infinitesimal “rectangle” R, two sides of which are
segments of streamlines of X, while the other two sides are segments of orthogonal
trajectories through the streamlines. See [11.5].
Here z is the point down to which R will ultimately be collapsed in order to find
the divergence there, S and P are the streamline and orthogonal trajectory through
z, and s and p are arc-length along S and P, the direction of increasing p being
chosen to make a positive right angle with X.
The net flux out of R is the difference between the fluxes entering and leaving.
The flux entering is |X| dp, while the flux leaving is the same expression evalu-
ated on the opposite side of R. It is now clear that two factors contribute to more
fluid leaving than entering: (1) greater fluid speed |X| as the fluid exits; (2) greater
separation dp of the streamlines as the fluid exits.
The second factor is clearly governed by how much the direction of X changes
along dp, in other words, by the curvature κP of P at z. More precisely, if δ denotes the
increase in a quantity as we move ds along the streamline, then [exercise] δ(dp) =
κP ds dp. Thus
(Net flux out of R) = δ{|X| dp}
= (δ|X|) dp + |X| δ(dp)
= (Bs |X| + κP |X|) (area of R).
The flux density is therefore
·
∇ X = Bs |X| + κP |X|. (11.5)
Flux and Work 545
In fact [exercise] this formula is still true for a three-dimensional vector field,
provided that there exists2 a surface P orthogonal to the streamlines, and κP is taken
to be the sum of its principal curvatures.
Turning to the circulation round R, identical reasoning yields [exercise] an
equally neat formula for the curl:
∇×X = −Bp |X| + κS |X|, (11.6)
where κS is the curvature at z of the streamline S.
Although we suspect that (11.5) and (11.6) must have been known to the likes of
Maxwell, Kelvin, or Stokes, we have not found any reference to these formulae in
modern literature.
·
∇ X=0 and ∇×X = 0.
The vector field is then said to be divergence-free and curl-free in R.
For example, consider the vector field of a point source with strength S:
!
S S x/(x2 + y2 )
X(z) = ⇐⇒ X = .
2πz 2π y/(x2 + y2 )
This should have zero flux density (i.e. divergence) everywhere except at the ori-
gin, where it should be undefined. Check that this is so. Recall that we previously
claimed that this was also the electrostatic field of a long, uniformly charged wire.
We can now see that this makes physical sense in that the field is locally conser-
vative. Thus if we fire our puck (which must now carry electric charge in order to
experience the force) round an infinitesimal loop, it will return to its starting point
with its kinetic energy unchanged. To verify this statement you need only check
that the field is curl-free.
We have seen that a sourceless and irrotational field is divergence-free and curl-
free. To end this section we wish to establish the converse result: if the divergence
and curl vanish throughout a region, the flux and work vanish for all simple loops
in that region. We will then have,
A vector field is sourceless and irrotational in a simply connected region if and only
if it is divergence-free and curl-free there.
To understand this converse, consider [11.6] which essentially reproduces part of
[8.27], p. 468. Let us now recycle the line of reasoning associated with that figure.
2
The condition for existence is that the curl either vanish or be orthogonal to the vector field.
546 Vector Fields and Complex Integration
[11.6] Gauss’s Theorem and Stokes’s Theorem. Using the same logic as in the proof of
Cauchy’s Theorem, summing integrals of the normal or the tangential component of X
around the boundaries of the shaded squares yields the integral along the outer boundary
K. But then the flux and work around each square can be expressed in terms of the
divergence and curl of X, via (11.3) and (11.4). This immediately yields Gauss’s Theorem
(11.7) and Stokes’s Theorem (11.8).
We begin by noting that as the grid gets finer and finer, the flux or work for K
becomes the flux or work for C. Next we relate these quantities to the divergence
and curl inside K. Check for yourself that exactly the same mathematical reasoning
which previously yielded
I X I
f(z) dz = f(z) dz ,
K shaded squares □
now yields
X
F [ X, K] = F [ X, □ ]
shaded squares
and
X
W [ X, K] = W [ X, □ ].
shaded squares
However, in the present context these results become accessible to physical intu-
ition. The first says that the total amount of fluid flowing out of K is the sum of
fluxes out of the interior squares. What does the second one say?
Now let the squares of the grid shrink so as to completely fill the interior R of
C. Using (11.3) and (11.4) and replacing the sum over squares by a double integral
over infinitesimal areas dA, we obtain Gauss’s Theorem,
Complex Integration in Terms of Vector Fields 547
ZZ
F [ X, C] =
R
·
[∇ X] dA, (11.7)
From these we see that if the divergence and curl vanish everywhere in R then the
flux and work for C also vanish, as was required.
Again following the logic in Chapter 8, consider what happens to the flux and
work as we continuously deform a closed contour, or an open contour with fixed
end points. You should be able to see that (11.7) and (11.8) imply two deformation
theorems:
If the contour sweeps only through points at which the divergence
(11.9)
vanishes, the flux does not change.
If the contour sweeps only through points at which the curl vanishes,
(11.10)
the work does not change.
from the vector field point of view. See [11.7]. In forming a Riemann sum with
terms H dz we now have the minor advantage that H = |H| eiβ and dz = eiα ds
are not drawn in separate planes, as they were in Chapter 8. However, we still face
the problem that H dz = |H| ei(α+β) ds involves the addition of angles, which is not
easy to visualize. Just as it is more natural to subtract vectors [yielding connecting
vectors] than to add them, so it is also more natural to subtract angles, for this yields
the angle contained between two directions.
The simple and elegant solution to our problem is to consider a new vector field:
instead of drawing H(z) at z we draw its conjugate H(z) = |H| e−iβ . We shall call
this the Pólya vector field of H. Before showing how this solves our problem, let us
offer (i) a caution and (ii) a reassurance:
(i) The Pólya vector field of H is not obtained by reflecting the picture of the ordi-
nary vector field for H in the real axis, for this would attach H(z) to z instead of
z. This will become very clear if you (or your computer) draw the Pólya vec-
tor fields of z and z2 , for example. Comparison with [10.1], p. 512, reveals that
the resulting phase portraits (not the vector fields themselves) are identical to
548 Vector Fields and Complex Integration
those of (1/z) and (1/z2 ). This is because zn points in the same direction as
(1/zn ).
(ii) As we will see in a moment, much is gained by representing H by its Pólya
vector field, but we also wish to stress that nothing is lost: the new field con-
tains exactly the same information as the old one. For example, it is clear that
the index of a loop L merely changes sign when we switch to the Pólya vector
field:
IH [L] = −IH [L].
Thus an nth order root of an analytic H still shows up clearly in its Pólya vector
field as a singular point, but now with index −n instead of n. Likewise, a pole
of order m produces a singular point of index m instead of −m.
Returning to integration, the great advantage of the Pólya vector field is that the
angle θ that it makes with the contour (see [11.7]) is given by θ = α − (−β), and
this is precisely the angle we were trying to visualize—the angle of the term H dz
in the Riemann sum. Better still, we find that
H dz = |H| eiθ ds
= |H| cos θ + i |H| sin θ ds
= · ·
H T + i H N ds.
Thus the real and imaginary parts of each term in the Riemann sum are the work
and flux of the Pólya vector field for the corresponding element of the contour. We
have thus discovered a vivid interpretation (due to Pólya3 and first championed
R
[11.7] How the Pólya
vector fieldH makes it possible to visualize K H(z) dz. We can
· ·
see that H dz = H T +i H N ds, so this immediately yields the extremely impor-
tant
R and useful physical and geometrical interpretation of the complex contour integral:
K H(z) dz = W [H, K] + i F [H, K].
3
See Pólya and Latta (1974).
Complex Integration in Terms of Vector Fields 549
by Bart Braden4 .) of the complex integral of H in terms of the work and flux of its
Pólya vector field along the contour:
Z
H(z) dz = W [H, K] + i F [H, K]. (11.11)
K
and
! !
Bx u
∇×H = × = −(Bx v + By u).
By −v
Thus the divergence and curl of H will both vanish if and only if the Cauchy–
Riemann equations are satisfied. Note for future use that these two equations are
really two aspects of a single complex equation,
i Bx H − By H = ∇×H + i ∇ H, · (11.13)
the vanishing of the LHS being the compact form of the CR equations, (5.1) on
page 246.
4
See Braden (1985), Braden (1987)—which won the MAA’s Carl B. Allendoerfer Award in 1988—and
Braden (1991).
550 Vector Fields and Complex Integration
In other words, in each unit of time, 2 units of fluid are being pumped into each
unit of area. The flux of fluid out of K is therefore 2A. On the other hand the flow
is curl-free:
! !
Bx x
∇×H = × = 0,
By y
so there is no circulation round K. Inserting these facts into (11.14) we obtain (11.15).
Figure [11.8] is a concrete example of this new way of looking at (11.15), the
shape of K having been chosen so as to make the values of the circulation and flux
obvious.
Clearly H(z) = z has no circulation along either of the arcs, and it has equal
and opposite circulations along the line-segments. The total circulation round K
Complex Integration in Terms of Vector Fields 551
[11.8] Area as one half the flux of H = z. Clearly, H(z) = z has no circulation along
either of the arcs, and it has equal and opposite circulations along the line-segments.
The total circulation round K therefore vanishes. Equally clearly, there is no flux across
the line-segments, but there is across the arcs. The larger arc has length aϕ and the
speed of the fluid crossing it is a, so the flux across it is a2 ϕ; similarly, for the smaller
arc it is b2 ϕ. Thus, F [z, K] = (fluid out) − (fluid in) = 2 (shaded area). Therefore,
H
K z dz = W [z, K] + i F [z, K] = 2iA.
therefore vanishes. Equally clearly, there is no flux across the line-segments, but
there is across the arcs. The larger arc has length aϕ and the speed of the fluid
crossing it is a, so the flux across it is a2 ϕ; similarly, for the smaller arc it is b2 ϕ.
Thus,
F [z, K] = (fluid out) − (fluid in) = 2 12 a2 ϕ − 12 b2 ϕ = 2 (shaded area).
Before moving on, let us clear up a paradoxical feature of the vector field z: fluid
is being pumped in uniformly throughout the plane, and yet the flow appears to
radiate from one special place, namely, the origin. The resolution (see [11.9]) lies in
the trivial identity z = z0 + (z − z0 ), which says that the flow from the origin is the
superposition of the sourceless, irrotational field z0 and a copy of the original flow,
but now centred on the arbitrary point z0 instead of the origin.
According to (11.11),
I
1
dz = W [(1/z), L] + i F [(1/z), L].
z
L
552 Vector Fields and Complex Integration
[11.9] The identity z = z0 + (z − z0 ) says that the flow from the origin is the superposition
of the sourceless, irrotational field z0 and a copy of the original flow, but now centred on
the arbitrary point z0 instead of the origin.
But the Pólya vector field (1/z) is an old friend—it is a source of strength 2π located
at the origin.
Figure [11.10] illustrates the intuitive nature of the result from the new point of
view. If a loop does not enclose the source, just as much fluid flows out as in; if a
Complex Integration in Terms of Vector Fields 553
simple loop does enclose the source, it intercepts the full 2π of fluid being pumped
in at the origin; more generally, a loop will accrue 2π of flux each time it encircles
the source.
To finish the explanation of (11.16) we must show that a source is pure flux, i.e.
every loop has vanishing work or circulation. Since a source is curl-free except at
the origin, Stokes’ Theorem guarantees vanishing work for simple loops that do
not contain 0. If the loop does contain 0 then it’s not so obvious. However, it is
obvious for an origin-centred circle. You can now finish the argument for yourself
by appealing to the Deformation Theorem (11.10).
In connection with another matter, consider the shaded sector in [11.10]. The
same amount of fluid will cross each segment of a contour which passes through it,
but the sign of the flux will depend on the direction of the contour. Try meditating
on the connection between this fact and the crossing rule for winding numbers
[(7.1), p. 388].
1
= 2
[Bx H − i By H] z + 12 [Bx H + i By H] z.
This will become exact in the limit that |z| shrinks to nothing.
Turning to the Pólya vector field itself, and substituting (11.13), we find
· z
H(z) ≍ H(0) + (∇ H) + (∇×H) + C z,
2
iz
2
(11.17)
[11.11] Local behaviour of a general vector field. In the limit that z Ñ 0, the general Pólya
·
vector field is given by this ultimate equality: H(z) ≍ H(0)+ 12 (∇ H)z+ 12 (∇×H)iz+C z.
Unless H(0) = 0, the constant first term dominates: vectors near the origin differ little
from the vector at the origin. The remaining three terms correct this crude approximation.
The second term describes a vector field (cf. [11.8]) that is irrotational and has constant
divergence, equal to that of H at the origin. The third term describes a vector field that is
sourceless and has constant curl, equal to that of H at the origin. The final term is both
irrotational and sourceless.
irrotational and has constant divergence, equal to that of H at the origin. The third
term describes a vector field that is sourceless and has constant curl, equal to that
of H at the origin. The final term is both irrotational and sourceless.
Note that this decomposition is geometrically meaningful because the appear-
ance of each of the component vector fields is qualitatively unaffected by the
value of its coefficient5 . We hope these observations make the formula (11.17) both
plausible and meaningful.
Now let us return to the original problem. Let K be a small simple loop of arbi-
trary shape round the origin, and let A be the area it encloses. We wish to show that
the divergence and curl of H at 0 are the limiting values of the flux per unit area
and work per unit area as K shrinks to the origin. Using (11.17) in (11.11) we find
5
This is obvious for the source and vortex terms, but not for the last term; see Ex. 10.
Complex Integration in Terms of Vector Fields 555
W [H, K] + i F [H, K]
I
= H(z) dz
K
I I I
= H(0)
K
1
·
dz + 2 ∇ H − i ∇×H z dz + C z dz.
K K
This becomes exact as K shrinks to nothing. But even if K is not small, we know
that the exact values of these three integrals are
I I I
dz = 0, z dz = 2iA, z dz = 0.
K K K
Thus
W [H, K] + i F [H, K] = ∇×H + i ∇ H A. ·
Equating real and imaginary parts, we obtain the desired results.
Figure [11.12] illustrates this field for positive A and B, as well as showing the
geometric significance of the algebraic decomposition above.
The first term is familiar as a source at p of strength 2πA, a negative value for
A corresponding to a sink. The second term is a multiple of the less familiar field
i/(z − p) which represents a vortex6 at p. It is easy to see that the circulation round
one of its circular streamlines is 2π, so this will also be its value for any simple
loop round p—we say that the vortex has strength 2π. On the other hand its flux
vanishes for all loops. While a source is pure flux, a vortex is pure circulation.
6
We are now using this term in a narrow sense—previously “vortex” referred to all vector fields
of this topological form.
556 Vector Fields and Complex Integration
[11.12] A more physical view of Cauchy’s Formula. If f(p) = A + iB, then as z Ñ p, the
local behaviour of H(z) ≡ f(z)/(z−p) is H(z) ≍ A[1/(z−p)]−B[i/(z−p)], as illustrated.
The first term is a source (pure flux) of strength 2πA. The second term represents a vortex
(pure circulation) such that the circulation
H f(z) round one of its circular streamlines is −2πB. If
C is a simple loop around p, then C (z−p) dz = W [H, C] + i F [H, C] = −2πB + i 2πA =
2πi f(p).
7
In the particular case of z2 this has also been observed by Braden (1991), though he did not supply
the general argument which follows.
Complex Integration in Terms of Vector Fields 557
H
[11.13] Pólya vector field explanation of C zn dz = W [z n , C] + i F [z n , C] = 0. Draw
the Pólya vector field z n of zn on an origin-centred circle. Rotating this picture about the
origin through any angle cannot change either the work or the flux. But rotating the picture
by π/(n + 1) yields the negative of the original field, changing the sign of both the work
and the flux. Since the work and the flux must remain unchanged and reverse sign, they
must both vanish.
yields the negative of the original field and, correspondingly, the negative of the
original work and flux. Since W and F are simultaneously required to remain the
same and to reverse sign, they must both vanish.
The same argument applies to z2 under a rotation of (π/3), and to z n under a
rotation of π/(n + 1). Use your computer to check this for n = 3. To understand
this symmetry better, consult Ex. 10.
[11.14] The Dipole Pólya vector field (1/z 2 ). [a] If we draw the field on a circle, then
a rotation of π reverses the field, so by the same reasoning as the previous figure,
H 2 2 2 m
C (1/z ) dz = W [(1/z ),HC] + i F [(1/z ), C] = 0. In general, (1/z ) is reversed under a
m
rotation of π/(m − 1), so C (1/z ) dz = 0. [b] The full phase portrait of the dipole field.
Note that the streamlines are perfect circles.
which dies away as the source/sink separation 2ϵ tends to zero. The solution is
to increase the strength S in inverse proportion to the separation 2ϵ, so that 2ϵS
remains constant. If we call this real constant 2πk, the limiting doublet field (as
ϵ → 0) is
k e−iϕ
D(z) = ,
z2
i.e. the general dipole field obtained by rotating [11.14] by +ϕ and scaling up the
speed of the flow by k, which we may think of as the “strength” of the dipole. Thus
the Pólya vector field of (d/z2 ) is a dipole whose axis points in the direction of d,
and whose strength is |d|. The complex number d is called the dipole moment.
We created the dipole by coalescing equal and opposite sources, increasing their
strength so as to avoid mutual annihilation. Continuing this game, we ask, “What
will happen if we coalesce equal and opposite dipoles, increasing their strength so as
to avoid mutual annihilation?” Figure [11.15] reveals the pleasing answer. Figure
[11.15a] represents a pair of equal and opposite dipoles located at ±ϵ and hav-
ing real dipole moments ±d, while [11.15b] is the Pólya vector field of (1/z3 ). The
resemblance is striking, and we can show algebraically that [11.15b], which is called
a quadrupole, is indeed the appropriate limiting case of [11.15a].
The field for [11.15a] is
1 1 z
Q(z) = d − = 4dϵ 2 . (11.19)
(z − ϵ)2 (z + ϵ)2 (z − ϵ2 )2
[11.15] The Quadrupole Pólya vector field (1/z 3 ). [a] Dipoles of strength ±d at ±ϵ are
merged by letting ϵ Ñ 0. But in order to form a quadrupole, the dipole moments d must
grow in inverse proportion to the separation ϵ, so that dϵ remains constant, otherwise
the two opposite dipoles would simply annihilate each other, leaving no field. [b] The full
phase portrait of the resulting quadrupole.
560 Vector Fields and Complex Integration
Once again letting the strength d grow in inverse proportion to the separation, so
that k = 4dϵ remains constant, the coalescence of the dipoles yields the quadrupole:
k
Q(z) = 3 .
z
In general, the Pólya vector field of (q/z3 ) is called a quadrupole with quadrupole
moment q.
We have thus explained the vanishing circulation and flux of (1/z3 ): each of
the dipoles in [11.15a] is known not to generate any circulation or flux, so the
quadrupole in [11.15b] won’t either. You are invited to continue this line of thought
by showing (geometrically and algebraically) that the fusion of two quadrupoles
yields the so-called octupole field, (1/z4 ), and so on.
Dipoles, quadrupoles, octupoles, etc., are collectively known as multipoles. Sim-
ilarly, dipole moments, quadrupole moments, etc., are collectively known as
multipole moments.
q d ρ
f(z) = 3
+ 2 + +a + b z + c z2 + · · · . (11.20)
z
| z
{z z}
P(z)
In the vicinity of the singularity, the behaviour of f is governed by its principal part
P, the Pólya vector field of which is
q d ρ
P(z) = 3
+ 2+ .
z z z
This we now recognize to be the superposition of a quadrupole, a dipole, and
a source-vortex combination of the type shown in [11.12]. Thus the principal
part of the Laurent series amounts to what a physicist would call a multipole
expansion.
To visually grasp the meaning of such an expansion, consider [11.16] which
illustrates a typical P. Very close to the singularity, the field is completely dom-
inated by the quadrupole with its characteristic four loops, but as we move
slightly further away the quadrupole’s influence wanes relative to the dipole.
Indeed, at this intermediate range we clearly see the characteristic two loops
of a dipole. Finally, at still greater distances, both the quadrupole and the
dipole become insignificant relative to the source-vortex, the precise form of
which is determined solely by the residue ρ. Compare with [11.12], in which
ρ = A + iB.
Continuing our outward journey, now well beyond the unit circle, the entire
principal part becomes negligible relative to the remaining terms of (11.20). First
a becomes important, then bz takes over, and so on. Thus as we approach infinity
the field at first resembles a dipole, then a quadrupole, and so on. However, unlike
the approach to the pole, on the journey to infinity we may experience multipoles
of greater and greater order, without end.
Of course in general f may possess other singularities and (11.20) will cease to
be meaningful when |z| increases to the distance of the nearest one. Nevertheless,
in the region where it is valid, we may still think of the non-negative powers as
representing multipoles at infinity.
To recap, Laurent’s series and the Residue Theorem may be conceived of physi-
cally as follows. The only term capable of generating circulation and flux is (ρ/z),
which may itself be decomposed into a vortex of strength W = −2π Im(ρ) and
a source of strength F = 2π Re(ρ). All the other terms correspond to multipoles
which generate neither circulation nor flux; a finite collection of these reside at the
pole, while the rest are at infinity.
562 Vector Fields and Complex Integration
[11.16] Multipole expansions and the Residue Theorem. Consider the Pólya vector field
f(z) of an analytic function in the vicinity of a triple pole. The dominant behaviour
near the singularity is given by the conjugate of the principal part of the Laurent series:
P(z) = zq3 + zd2 + ρz , which we recognize to be the superposition of a quadrupole, a
dipole, and a source–vortex combination of the type shown in [11.12]. Thus the principal
part of the Laurent series amounts to what a physicist would call a multipole expansion.
Very close to the singularity, the field is completely dominated by the quadrupole with its
characteristic four loops, but as we move slightly further away the quadrupole’s influence
wanes and we see the characteristic two loops of a dipole. Finally, at still greater distances,
both the quadrupole and the dipole become insignificant relative to the source–vortex,
the precise form of which is determined solely by the residue ρ. Neither the quadrupole
nor the dipole generates any circulation or flux,H so if L encircles the pole, the circulation
and flux is entirely due to the source–vortex: L f(z) dz = 2πiρ.
11.3.1 Introduction
Phase portraits are so convenient that it is easy to forget that in general they cannot
represent the lengths of the vectors. In this section we shall see that if a vector field is
either sourceless or irrotational (or both) then there exists a special way of drawing
the phase portrait so that the lengths are represented.
The Complex Potential 563
[11.17] The Stream Function Ψ and the k-flux tubes of a sourceless flow. If H is source-
less, and K is any contour connecting an arbitrary fixed point a to a variable point
z, the flux across it is independent of K, yielding the well-defined stream function,
Ψ (z) ≡ F [H, K]. Now choose a number k and draw just those streamlines for which
Ψ = 0, ±k, ±2k, ±3k, . . . Any cross section of a channel between neighbouring stream-
lines therefore intercepts the same flux k, and so we call it a k-flux tube. This special phase
portrait (pioneered by Faraday and Maxwell) has the great advantage that the speed of
the flow is now visible via |H| ≍ (k/ϵ): the denser the streamlines, the faster the flow.
To summarize:
Let the phase portrait of a sourceless vector field be constructed using k-
flux tubes. If k is chosen small, the speed of the flow at any point will be
(11.22)
approximately given by k divided by the width of the tubes in the vicinity
of the point. For infinitesimal k, the result is exact.
However, since the number of k-flux tubes passing through a given region will
vary inversely with k, our phase portrait will get very cluttered if k is chosen too
small. In practice (cf. [10.3], p. 514) we get a good feel for the speed of the flow with
relatively few streamlines.
Let’s apply these ideas to the simple (non-analytic) example H(z) = i z. The
Pólya vector field is then
!
y
H = −i z ⇐⇒ H = ,
−x
the streamlines of which are clockwise circles round the origin, the speed of the
flow round each one being equal to its radius. See [11.18].
Although this vector field is not irrotational [∇×H = − 2], it is sourceless
·
[∇ H = 0], and thus it possesses a stream function. For convenience’s sake, let’s
choose a = 0. We already know that the streamlines are origin-centred circles, so
to find the value of Ψ on the streamline of radius R we must find the flux for any
The Complex Potential 565
path from the origin to any point on this circle. Choosing the path to be the portion
of the positive real axis from 0 to R, we see that
!
0
ds = dx and N = .
−1
Thus
Z ZR
·
Ψ = (H N) ds =
0
x dx = 12 R2 .
Knowing the stream function we are now in a position to draw the special
phase portrait. Choosing k = (1/2) we find that the radii of the streamlines
√ √ √
are 1, 2, 3, . . . Figure [11.18] illustrates these streamlines, and qualitatively
confirms the prediction of (11.22). As we move outward from the origin the
streamlines become more crowded together, reflecting the increasing speed of
the flow.
[11.19] The field in terms of the stream function: H = −i ∇Ψ. First, the direction of ∇Ψ
is the one that is orthogonal to the streamlines and along which Ψ increases. Thus −i ∇Ψ
points in the direction of H. Second, |∇Ψ| ≍ (k/ϵ) ≍ |H|. Done.
! !
dΨ = (Bx Ψ) dx + (By Ψ) dy =
Bx Ψ
By Ψ
· dx
dy
·
= ∇Ψ dz.
∆Ψ ≡ Bx2 Ψ + By2 Ψ = 0.
Solutions of this equation are called harmonic, so we may restate the result as
follows:
A sourceless field is irrotational if and only if its stream function is harmonic.
The verification is a simple calculation:
! !
Bx By Ψ
∇×H = × = −∆Ψ.
By −Bx Ψ
[11.20] The Potential Function Φ of a conservative force field. If K is any contour con-
necting an arbitrary fixed point a to a variable point z, the work done by the conservative
(irrotational) field in moving the particle along K is a well-defined function of z, indepen-
dent of K, called the potential function, Φ(z) = W [H, K]. Just as we did with the stream
function, we draw just those equipotentials for which Φ = 0, ±l, ±2l, ±3l, . . . In this
picture the same amount of work l is required to move the particle from each equipoten-
tial to the next, so we call the region lying between two such adjacent equipotentials an
l-work tube. The equipotentials are the orthogonal trajectories through the lines of force.
Also, |∇Φ| ≍ (l/δ) ≍ |H|, so H = ∇Φ.
Thus the magnitude of the force is represented by the crowding together of the
equipotentials:
we see that
A conservative force field is sourceless if and only if its potential function
(11.30)
is harmonic.
The LHS of [11.21] illustrates such a division into approximately square k-cells.
We have labelled Φ = 11k and Ψ = 3k, but we have left it to you [exercise]
to label the remaining streamlines and equipotentials; this can only be done in
one way.
Note that once such a special phase portrait (including the equipotentials) has
R
been drawn with a small value of k, the value of L H dz is easy to find. For if L
crosses m equipotentials and n streamlines, an accurate estimate of the integral
will be k(m + in). If L crosses an equipotential or streamline more than once, how
should m and n be counted?
We mention in passing that there is an interesting physical interpretation of the
k-cells which is due to Maxwell (1881). Suppose that the vector field represents the
flow of a fluid having unit mass per unit area. In the limit of vanishing k, the speed
v will be constant throughout any particular cell, and the kinetic energy of the fluid
in that cell will be
2
2 1 2 k
1
kinetic energy = 2 (area) v = 2 ϵ = 21 k2 .
ϵ
570 Vector Fields and Complex Integration
[11.21] The Complex Potential, Ω = Φ + iΨ. If H(z) is analytic, then H(z) is both irro-
tational and sourceless, so we may simultaneously define and superimpose l-work tubes
orthogonal to the streamlines with k-flux tubes along the streamlines, thereby dividing the
region of the flow into small “rectangles” (as l and k tend to zero). We now go one step
further and choose l = k, dividing the region into squares bounded by equipotentials
and streamlines. This makes it vividly clear that if we combine the potential and stream
functions into the single complex potential Ω = Φ + iΨ, then the mapping Ω is locally
an amplitwist, i.e., Ω ′ exists. Indeed, we will show that H = Ω ′ .
Thus
Each k-cell contains the same amount of energy, and the total energy in a
region is thus obtained by counting the number of k-cells contained within
it.
If we reinterpret the vector field as an electrostatic field, and correspondingly rein-
terpret “energy” as electrostatic energy, the result is still valid; this was the context
in which Maxwell discovered it.
The result (11.31) is intimately connected with ideas of complex analysis. To
see this, let us combine the potential and stream functions into a single complex
function Ω called the complex potential:
Ω(z) = Φ(z) + i Ψ(z).
Returning to the dominant point of view of this book, think of Ω as a mapping. The
RHS of [11.21] shows the image of the special phase portrait under this mapping:
The complex potential maps streamlines to horizontal lines and equipo-
tentials to vertical lines. Furthermore, each square k-cell is mapped to a
square of side k. Thus Ω is an analytic mapping.
We may check this symbolically. Equating (11.24) and (11.28) we obtain
! !
Bx Φ By Ψ
= ,
By Φ −Bx Ψ
which are the CR equations for Ω.
The Complex Potential 571
H = ∇Φ = Bx Φ + i By Φ = Bx Φ − i Bx Ψ = Bx Ω = Ω ′ .
[11.22] Ω(z) = z+ z1 represents the superposition of a uniform eastward flow and a dipole.
11.3.6 Examples
(1) We previously claimed that the streamlines of the dipole H = (1/z 2 ) in [11.14b]
were perfect circles, and we asked you to provide a simple geometric proof. A
second demonstration is obtained by finding the complex potential:
1 1 1
H= 2
=⇒ Ω ′ = 2 =⇒ Ω = − + c.
z z z
The streamlines are the images under Ω−1 (z) = −1/(z − c) of horizontal lines. The
result follows from the fact that inversion sends straight lines to circles through
the origin.
(3) A source of strength 2π at the origin has vector field H = (1/z). If we choose to
measure work and flux along a path L emanating from z = 1 then [see p. 465] the
complex potential is
The Complex Potential 573
[11.23] The complex potential of the source (1/z) is a multifunction: although the work
is single-valued, the flux increases by 2π with every extra revolution around the source.
But if we restrict attention to a simply connected region D that does not contain the
source, then Ω becomes single-valued.
While the work Φ is single-valued, the flux Ψ is a multifunction whose values differ
from each other by multiples of 2π. This makes perfect sense since each time L
encircles the source it intercepts the full 2π of fluid being pumped in there. Note
that the single-valued inverse function Ω−1 (z) = ez does indeed map horizontal
and vertical lines to the source’s streamlines and equipotentials.
If we wish to obtain a single-valued complex potential we may do so by con-
fining our attention to any simply connected region not containing the source. The
shaded region D in [11.23] is an example. Any two paths from 1 to z that lie wholly
within D may be deformed into each other without ever leaving D, hence without
crossing the source, hence without altering the flux. For example, we see that for
the particular choice of D in [11.23], the unique values of Ψ at (1 + i) and at (2 + 2i)
are (π/4) and (9π/4). However, a different choice of D might well yield different
values of Ψ at these two points.
More generally, if D is any simply connected region not containing any singular-
ities of an otherwise analytic H, the Pólya vector field H will possess a single-valued
complex potential in D.
574 Vector Fields and Complex Integration
11.4 Exercises
1 For each of the following vector fields X verify that the geometric formulae
(11.5) and (11.6) yield the correct values for the divergence and for the curl:
(i) X = (1/z).
(ii) X = z.
(iii) X = x2 , where z = x + iy.
(iv) X = y2 , where z = x + iy.
(v) X = i(1/r2 )eiθ , where z = reiθ .
2 For each of the following vector fields X, calculate F [X, C] and W [X, C] for
the given loop C, then check your answers by substituting the results of the
previous question into (11.7) and (11.8).
(i) X = x2 , and C is the edge of the rectangle a ⩽ x ⩽ b, −1 ⩽ y ⩽ 1, traversed
counterclockwise.
(ii) X = i(1/r2 )eiθ , and C is the edge of the region a ⩽ r ⩽ b, 0 ⩽ θ ⩽ π,
traversed counterclockwise.
3 Use a computer to draw the Pólya vector field of f(z) = 1/[z sin z] and thereby
identify the locations and orders of the poles of f(z). For each of the following
H
choices of C, numerically estimate C f(z) dz by making on-screen measure-
ments of the vectors, then estimating the flux and circulation round C. In each
case check your estimate by calculating the exact answer using residue theory.
(i) Let C be a small circle centred at −π.
(ii) Let C be a small circle centred at 0.
(iii) Let C be a small circle centred at π.
(iv) Let C be a small circle centred at 2π.
(v) Let C be the boundary of the rectangle 1 ⩽ x ⩽ 7, −1 ⩽ y ⩽ 1.
4 Repeat parts (i) and (ii) of the previous question using f(z) = z cosec 2 z.
5 Let L be a contour from the real number −θ to +θ. By choosing L to be a line-
segment, and then sketching the Pólya vector field at points along L, show that
R iz
L z e dz is purely imaginary. Verify this by calculating the exact value of the
integral.
6 All complex analysis texts recognize the great utility of the inequality
Z Z
L
f(z) dz ⩽
L
·
|f(z)| |dz|, (11.32)
Exercises 575
but none that we know of have sought to answer the question, “When does
equality hold?” This is probably because no elegant answer is forthcoming (cf.
our attempt in Chapter 8) without the concept of the Pólya vector field. How-
ever, armed with the Pólya vector field, we have what we shall call Braden’s
Theorem8 :
Equality holds in (11.32) if and only if the contour L cuts the streamlines
of the Pólya vector field of f at a constant angle.
Explain Braden’s Theorem.
7 Continuing from the previous question, suppose that f(z) = z.
(i) Show that if L is a segment of the spiral with polar equation r = eθ , then
the condition of Braden’s Theorem is met.
(ii) Verify by explicit calculation that equality does indeed hold in (11.32), as
predicted.
8 Consider the flow created by (2n + 1) sources, each of strength 2π, located at
(i) If Ωn (z) denotes the complex potential of this flow, show that
z2 z2 z2
Ωn (z) = ln z 1 − 2 1 − 2 2 ··· 1 − 2 2 + const.
π 2π nπ
(ii) Ignoring the constant, and referring to Ex. 13 on p. 508, deduce that as the
number of sources increases without limit, Ωn (z) tends to Ω(z) = ln[sin z].
(iii) Check that this answer makes sense by using a computer to draw the
velocity vector field, V = Ω ′ .
9 (i) Explain why the derivative of the complex potential of a source yields the
complex potential of a dipole.
(ii) Referring to the previous question, draw a sketch predicting the appear-
d
ance of the flow whose complex potential is Ω(z) = dz ln[sin z]. Check
your answer by getting the computer to draw this flow.
10 Reconsider the term C z in the local decomposition (11.17) of a general vector
field. See [11.11].
(i) Show that the visual appearance of the vector field C z is essentially inde-
pendent of the value of C. More precisely, show that if C = eiϕ then
increasing ϕ merely causes the entire picture of the vector field C z to
rotate, in fact exactly half as fast as eiϕ rotates.
8
See Braden (1987). We independently recognized this fact, probably at about the same time as
Braden himself.
576 Vector Fields and Complex Integration
(ii) To make the result vivid, create a computer animation of the vector field
eiϕ z as ϕ increases from 0 to π.
(iii) More generally, show that if n is an integer and F(z) stands for either zn
or z−n , then the vector field of eiϕ F is obtained by rotating the vector field
of F through ϕ/(n + 1). [Note that the n = −1 fields (including sources
and vortices) are exceptional.]
11 Consider a flow such that the inverse complex potential is Ω−1 (w) = w + ew .
Use a computer to draw the streamlines, and verify mathematically that the
picture may be interpreted as the flow out of a channel −π ⩽ Im(z) ⩽ π,
Re(z) ⩽ −1.
12 Consider the flow with complex potential
1 ez + 1
Ω(z) = .
2 ez − 1
Use a computer to draw the streamlines, and verify mathematically that the pic-
ture may be interpreted as the flow that results when the dipole with complex
potential Ω(z) = (1/z) is confined to the channel −π ⩽ Im(z) ⩽ π.
13 Continuing from the previous question, what would the new complex potential
be if fluid were flowing down the channel with speed v prior to the insertion of
the dipole? Check your answer by using a computer to draw the streamlines.
14 Suppose that the doublet consisting of a source of strength 2π at z = 1 and a
sink of equal strength at z = −1 is inserted into the uniform flow with real, pos-
itive velocity v. Locate the “stagnation points” (singular points of zero velocity)
of the net flow, and describe (perhaps with the aid of a computer animation)
how they move as v varies from 0 to 3.
15 If two sources are located at opposite corners of a square, and two sinks are
located at the other two corners, and all four are of equal strength, then show
that the circle through these four points is a streamline. Check this by getting
the computer to draw the complete flow.
16 Show that the streamlines produced by two vortices of equal strength are Cassi-
nian curves (figure [2.8b], p. 68) whose foci are the locations of the two vortices.
[Note that your reasoning immediately generalizes: Cassinian curves with n foci
are the streamlines of n equal vortices placed at the foci.]
17 Show that the streamlines [11.15b] and the equipotentials of a quadrupole are
lemniscates (see [2.9], p. 69).
CHAPTER 12
1
Linguistically, the common origin of both terms is the Latin word “conjugatus”, meaning joined
together.
Visual Complex Analysis. 25th Anniversary Edition. Tristan Needham, Oxford University Press.
© Tristan Needham (2023). DOI: 10.1093/oso/9780192868916.003.0012
578 Flows and Harmonic Functions
[12.1] Evolution of a source into its dual vortex: As ϑ goes from 0 to (π/2), the field
b ≡ Hπ/2 =
Hϑ ≡ eiϑ H = eiϑ (1/z) rotates from a pure source into its dual pure vortex, H
iH = (i/z).
Later we shall see that the concept of a dual flow is very useful. For example,
having found the flow of a fluid round an obstacle, the dual flow represents the
electric field which solves an analogous problem in electrostatics.
As interesting examples of dual flows, consider what happens in the vicinity
of a singularity. Figure [12.1] illustrates how, as ϑ varies from 0 to (π/2), a source
gradually evolves into a dual vortex of equal strength. Note (cf. [11.12], p. 556) that
the intermediate flow may also be viewed as a superposition of the original flow
and its dual. Indeed, this is true quite generally:
b.
Hϑ = (cos ϑ)H + (sin ϑ)H
Check for yourself that the type of qualitative change of flow exhibited in [12.1]
does not occur in the case of higher multipoles. For example, the dual of a dipole
is just another dipole. As ϑ varies from 0 to (π/2), are all the intermediate flows
dipoles, as well? See Ex. 10 of the previous chapter.
Observe that in passing from a flow to its dual the roles of the streamlines and
equipotentials are interchanged: the streamlines of the dual flow are the equipoten-
tials of the original, while the equipotentials of the dual flow are the streamlines
of the original. Symbolically, this interchange of roles is manifested in the fact that
the dual potential and stream functions are
b = +Ψ and Ψ
Φ b = −Φ.
[12.2] In passing from a flow to its dual, the streamlines and equipotentials are
interchanged.
fluid crosses it from left to right, so the original and dual stream functions increase
in the illustrated directions. We now see clearly that Φ b and Ψ increase in the same
direction, while Ψ b and Φ increase in opposite directions.
Given a complex potential Ω = Φ + iΨ, we may thus think of Ψ as either
the stream function, or as the dual of the potential function. Likewise, Φ may be
thought of as either the potential function, or as minus the dual of the stream func-
tion. Since any analytic function f = u + iv may be thought of as a complex
potential, we may extend this language and say that v is dual to u, and that −u
is dual to v.
Finally, we cannot resist at least mentioning two miraculous connections
between the above ideas and the study of soap films, also known as minimal
surfaces, characterized by vanishing mean curvature.
Imagine yourself standing on the tangent plane to a saddle-shaped surface. In
the case of a soap film, as you turn around, the average (or “mean”) curvature
vanishes: the surface bends away from the tangent plane equally and oppositely as
you turn around.
First miracle: Each complex analytic function H(z) describes the shape of a
minimal surface, and vice versa. Second miracle: Varying ϑ causes the minimal
surface corresponding to Hϑ (z) to undergo stretch-free bending: all these minimal
surfaces have identical intrinsic geometry. For example, if H corresponds to the so-
b corresponds to the so-called catenoid, and [12.3] illustrates
called helicoid, then H
the stretch-free bending of one into the other, each intermediate surface itself being
a minimal surface.
For an elementary introduction to the fascinating subject of minimal surfaces, see
Hildebrandt and Tromba (1985); for the mathematical details, see Nitsche (1989).
580 Flows and Harmonic Functions
Ψ = 3X2 Y − Y 3 .
Alternatively, since
The simplicity of the second method depended crucially on our ability to express
∇Φ(x, y) as a function Ω ′ (z) of z, but it is not always so obvious how to do this.
However, there does exist a systematic method of doing this in the case where Φ
is defined in a region containing a segment of the real axis.
Let V(x) be the vector field evaluated on the real x-axis, i.e., V(x) = ∇Φ(x, 0). If
Φ(x, y) is an explicit formula in terms of the familiar functions (powers, trigono-
metric, exponential) that possess complex analytic generalizations, then V(x) is
such a formula also. Hence if in the formula for V(x) we now replace the sym-
bol x with the complex variable z then we obtain an analytic function V(z) which
agrees with Ω ′ (z) when z is real. But, as we saw in Chapter 5, this implies that the
two functions must continue to agree when z becomes complex.
Thus our recipe for finding Ω ′ (z) as an explicit formula in z is to calculate
∇Φ(x, y), set y = 0, then substitute z for x:
where F stands for three terms which vanish when y = 0. Using (12.1) we get
Ω ′ (z) = esin z cos z, and hence Ψ = Im esin z .
Conformal Invariance 583
In other words, corresponding points in the two planes are assigned equal function
values. We will now show (first symbolically then geometrically) that
e
Harmonicity is conformally invariant: Φ(w) is harmonic if and only if
(12.3)
Φ(z) is harmonic.
e
As before, think of Φ(w) e ≡ ∇Φ.
as the potential of the vector field V e If and only
e e e e
if Φ is harmonic, V possesses an analytic complex potential Ω(w) = Φ(w)+i Ψ(w), e
where the stream function Ψ e is the harmonic dual of Φ.
e Since f is analytic, so is its
composition with Ω: e
e
Ω(z) ≡ Ω[f(z)] = Φ(z) + iΨ(z).
(i) With a small value of k, draw the equipotentials Φ = 0, ±k, ±2k, ±3k, . . ..
(ii) Choose one of the resulting k-work tubes [shaded in the figure] and draw
line-segments across it in such a way that the tube is divided into squares.
(iii) Extend these line-segments into lines of force [dashed] of V, i.e., orthogonal
trajectories through the equipotentials.
Then Φ is harmonic if and only if these lines of force divide each k-work tube into squares.
Figure [12.4a] illustrates this test for a Φ that is harmonic, while [12.4b] illustrates
it for one that is not. The result (12.3) can now be seen as nothing more than a
statement of the conformal invariance of this geometric test. Let us spell this out.
Equation (12.2) defines the potential of each point in the z-plane to be the same
as its image point (under f) in the w-plane. Thus f maps the k-work tubes of Φ to
584 Flows and Harmonic Functions
[12.4] Geometric test for harmonicity of Φ. STEP ONE: With a small value of k, draw
the equipotentials Φ = 0, ±k, ±2k, ±3k, . . . STEP TWO: Divide a k-work tube [shaded]
into squares. STEP THREE: Extend the edges of the squares into lines of force. Then Φ
is harmonic if and only if these lines of force divide each k-work tube into squares.
So [a] is harmonic, and [b] is not.
[12.5] Conformal invariance of the harmonicity test for Φ. If f is conformal, then the
image test grid will pass the harmonicity test if and only if the original test grid passes the
test. Here, the initial grid passes, so its image does, too.
e See [12.5]. Finally, since f is conformal, the constructed test
the k-work tubes of Φ.
grid for Φ will be composed of squares if and only if the image grid is composed
of squares. Figure [12.5] illustrates the case where the potentials are harmonic.
e 1
∆Φ(w) = ∆Φ(z). (12.4)
|f ′ (z)|2
Conformal Invariance 585
· ·
e 1
∇ V(w) = ∇ V(z). (12.5)
|f ′ (z)|2
Now consider [12.6], which illustrates a toy model of the phenomenon. The
potential Φ(z) = (S/4) |z|2 generates a vector field V = (S/2) z of uniform diver-
·
gence ∇ V = S. With a small value of k, the LHS of [12.6] shows the special
equipotentials Φ = 0, ±k, ±2k, ±3k, . . ., for which the strength of the field is
inversely proportional to the separation of the curves. Now apply the mapping
w = f(z) = cz, which is a rotation and an expansion by |c|. By definition, these
e
expanded circles are equal-valued equipotentials of Φ(w), so that
e S
Φ(w) = Φ(z) = 41 S |z|2 = 1
|w|2 .
4 |c|2
The field Ve = ∇Φ e therefore has uniform flux density ∇ V ·
e = S/|c|2 , proving (12.5)
for this case.
More intuitively still, in [12.6] compare the flux leaving the shaded disc on the
left with the flux leaving the shaded image disc on the right. Since the separation
of adjacent equipotentials is scaled up by |c|, the strength of the field on the rim of
586 Flows and Harmonic Functions
the image disc is scaled down by |c|, while the circumference of the rim is scaled
up by |c|. The net effect is that the flux of Ve out of the image disc is the same as the
flux of V out of the original disc. Finally, since the area of the disc is scaled up by
|c|2 , the flux density is scaled down by |c|2 .
To employ this idea in the general setting, it is only necessary to recognize
that the local behaviour of a general potential is very similar to our toy potential,
and that the local effect of a general analytic mapping f is very similar to our toy
mapping, with |f ′ | playing the role of |c|.
We will not spell this out completely because we will shortly be able to give a
second explanation which is even simpler. However, according to (11.17), p. 553,
the behaviour of V very near to z0 is expressible as
·
(z − z0 )
V(z) = [∇ V(z0 )] + Y(z),
2
where Y is sourceless and, of course, irrotational. Correspondingly, the local
behaviour of the potential is2
·
Φ(z) = 14 [∇ V(z0 )] r2 + Υ(z), (12.6)
Note that this result is in accord with Gauss’s Mean Value Theorem, which says
that if Φ is harmonic then ⟨Φ⟩−Φ(p) = 0 for circles of any size, not just infinitesimal
ones. In fact if you have already convinced yourself of (12.6) then [exercise] you
2
This may also be derived directly by taking the Taylor series for Φ and rewriting it in a rather
unobvious way.
A Powerful Computational Tool 587
may derive (12.7) by using the fact that the harmonic function Υ obeys Gauss’s
Mean Value Theorem.
Before giving a more direct derivation of (12.7), let us return to Gauss’s Mean
Value Theorem itself and rederive it without appealing to complex analysis3 . Let
V = ∇Φ be the vector field of the potential function Φ. The flux of V out of a
(non-infinitesimal) circle C of radius r is then [exercise]
Br ⟨Φ⟩ = 1
2 r ∆Φ,
which may be integrated to yield the formula in (12.7). To complete the explana-
tion of (12.7) it is only necessary to observe that the Laplacian of an arbitrary Φ is
constant within an infinitesimal circle.
Knowing the meaning of the Laplacian, it is a simple matter to understand its
conformal invariance as expressed in (12.4). The analytic mapping f amplitwists an
e centred at p
infinitesimal circle C centred at p to an infinitesimal circle C e, the new
′ e
radius being er = |f (p)| r. By definition, Φ(e
p) = Φ(p). Likewise, the values of Φ e at
points of Ce are the same as those of Φ at the preimages on C, so ⟨Φ⟩ e on C e equals
⟨Φ⟩ on C. Thus (12.7) implies
e p) = |f ′ (p)|2 r2 ∆Φ(e
r2 ∆Φ(p) = er 2 ∆Φ(e e p),
A zealot might wish for an ideal world in which calculation would always be
relegated to the confirmation of insights provided by geometry. Alas, even this
author must confess to occasional lapses in which calculation has preceded under-
standing! We now describe a powerful computational tool which in many areas of
complex analysis provides a considerable saving of labour. In the next section the
study of the “complex curvature” [cf. Chapter 5] will provide a good showcase for
its simplicity and elegance.
3
Previously we got it from ⟨f⟩ = f(p), which in turn came from Cauchy’s formula.
588 Flows and Harmonic Functions
and we are free (as we have previously done) to think of this as a complex function
∇R = Bx R + i By R.
From this we may abstract the complex gradient operator ∇, together with the
conjugate operator ∇:
∇ = Bx + i By and ∇ = Bx − i By .
These two operators open the way to an exciting new method of calculation.
Given a vector field
!
u
f= ,
v
we have seen how the real version of ∇ may be formally dotted or crossed with f to
·
yield its divergence ∇ f or its curl ∇×f. The interpretations of these quantities as
flux and work densities shows them to be truly geometric, that is to say, coordinate-
independent. However, there would seem to be no natural way of applying ∇
directly to f to obtain a new vector field ∇f. However, if we replace ∇ by its com-
plex version ∇, and replace the vector field f by the complex function f = u + iv,
then there is a natural definition:
• ∇(f + g) = ∇f + ∇g.
• ∇(fg) = f∇g + g∇f.
• If f is analytic then ∇f[g(z)] = f ′ [g(z)] ∇g. For example,
∇eg(z) = eg(z) ∇g.
A Powerful Computational Tool 589
·
∇f = ∇ f + i ∇×f.
Similarly,
·
∇f = ∇ f − i ∇×f,
which shows, once again, that a vector field is sourceless and irrotational if and
only if it is the Pólya vector field of an analytic function.
• The Laplacian operator ∆ can be expressed neatly as ∇∇ = ∆ = ∇∇.
In the next section, and in the exercises at the end of the chapter, you will see
the strength of the new technique. You may also find that exercises from previous
chapters are solved more readily by this method. For the moment, here are just two
examples of the use of the complex gradient.
The first is simply to observe how neatly the theorems of Gauss and Stokes
[p. 547] may be combined into a single complex result: If C is the boundary curve
of a simply connected region R then
I ZZ
f dz = i ∇f dA.
R
C
∇z = f ′ ∇w and ∇z = f ′ ∇w . (12.8)
For example,
[12.7] Geometric harmonicity test for a family of curves, E. STEP ONE: Choose two
members of E that are very close together. STEP TWO: Draw line-segments across the
region between them [shaded] so as to divide it into squares. STEP THREE: Extend these
line-segments into streamlines [dashed], i.e., orthogonal trajectories through E. STEP
FOUR: Choose one of the resulting flux tubes [darkly shaded] and draw line-segments
across it so as to divide it into squares. STEP FIVE: Extend these line-segments into mem-
bers of E. Then E is harmonic if and only if the resulting grid is composed of squares.
[a] Concentric circles are harmonic. [b] Concentric ellipses are not harmonic.
feeling that in order for a grid of squares to form, the bending of the curves of E
must be connected in some special way with the bending of the curves of S. By
examining the curvatures of the two types of curve we shall see that this is indeed
the case.
First let us attach directions to the curves of E and S, so that their curvatures will
have well-defined signs. Choose the direction for S arbitrarily, but then define the
direction of E to be that of a tangent vector to S rotated through a positive right angle.
Through any given point w0 there passes one member C1 of S and one member C2
of E. Let the curvatures of C1 and C2 be κ1 and κ2 , and let s1 and s2 be the arc lengths
along C1 and C2 . We then have the following striking result:
(12.9) is trivially satisfied. In [12.8b] it is clear that both curvatures are decreasing
at the indicated point, so (12.9) is violated there.
The result (12.9) seems to have been first published by Bivens (1992). We had
also hit upon the same result, but as a natural consequence of investigating our
complex curvature concept, which we introduced in Chapter 5. Here is the pertinent
result:
The complex curvature vector field of an analytic mapping f is
(12.10)
automatically sourceless, and its stream function is Ψ = 1/|f ′ |.
To see the connection between the two results, first observe that a family E in the
w-plane is harmonic if and only if it is the image set of vertical lines in the z-plane
under an analytic mapping w = f(z). For if E is harmonic then its potential function
Φ(w) is the real part of a complex potential z = Ω(w) which maps E to vertical lines,
and so f(z) ≡ Ω−1 (z) has the required property. Conversely, if E is the image set of
vertical lines under an analytic f then it is harmonic, indeed its harmonic potential
is the real part of the complex potential Ω(w) ≡ f−1 (w).
Thus if E is harmonic then the curves C1 and C2 through w0 = f(z0 ) are the images
under f of the horizontal and vertical lines through z0 . But referring to [5.20], p. 272,
we see that curvatures occurring in (12.9) are simply the real and imaginary parts
of the complex curvature of f at z0 :
K(z0 ) = κ1 + iκ2 .
curvature is
·
Bκ1 Bκ2 ′ Bκ1 Bκ2
∇ K= + = |f (z0 )| + .
Bx By Bs1 Bs2
The result (12.10) therefore implies that equation (12.9) is a necessary condition for
E to be harmonic; the question of sufficiency will be addressed shortly.
To prove (12.10) we will use the complex gradient technique of the previous
section; later we will give a proper geometric explanation. We must show that Ψ =
1/|f ′ | is the stream function for the complex curvature
i f ′′
K= ,
f ′ |f ′ |
in other words [cf. (11.24), p. 566], K = −i ∇Ψ.
Since
1 i
−i ∇Ψ = −i ∇ = ′ 2 ∇|f ′ |,
|f ′ | |f |
we need to know ∇|f ′ |. Because f is analytic, so is f ′ , and this implies ∇f ′ = 0 and
∇f ′ = 2f ′′ . Hence
i f ′ f ′′ i f ′′
−i ∇Ψ = = = K.
|f ′ |3 f ′ |f ′ |
The fact that harmonicity implies (12.9) can also be understood without appeal-
ing to ideas from complex analysis. Let S and E be the streamlines and orthogonal
trajectories of a vector field X. With the present notation, the results (11.5) and (11.6)
from the previous chapter [see p. 544] become
·
B|X| B|X|
∇ X= + κ2 |X| and ∇×X = − + κ1 |X|.
Bs1 Bs2
In order for E (or S) to be harmonic, X must be divergence-free and curl-free, so
B B
κ1 = ln |X| and κ2 = − ln |X|,
Bs2 Bs1
from which (12.9) immediately follows.
We conclude this section by establishing the converse result that equation (12.9)
is a sufficient condition for E to be harmonic. In the w-plane, let Θ(w) be the angle
that the curve C1 through w makes with the horizontal. The angle of the curve C2
through w is therefore Θ + (π/2). Since the curvatures of C1 and C2 are the rates of
594 Flows and Harmonic Functions
change of these two angles with respect to the distances along the curves, we then
have
BΘ BΘ
κ1 = and κ2 = .
Bs1 Bs2
Next we calculate the Laplacian of Θ; the reason will be clear in a moment.
Because the Laplacian is coordinate-independent, we may choose our coordinate
directions tangent to C1 and C2 , obtaining
B BΘ B BΘ Bκ1 Bκ2
∆Θ = + = + .
Bs1 Bs1 Bs2 Bs2 Bs1 Bs2
Thus equation (12.9) implies that Θ(w) is harmonic, in which case it is the real
part of an analytic function, say, G(w). We may now define an analytic function
H(w) = e−iG ∝ e−iΘ such that H ∝ eiΘ is everywhere tangent to S. Thus S
and E are the streamlines and equipotentials of the Pólya vector field of an analytic
function. Done.
Kf (z) = −i ∇z (1/|f ′ |)
i
= − ′
∇w |Ω ′ |
Ω
|Ω ′ |2
= − ′ [−i ∇w (1/|Ω ′ |)]
Ω
= −Ω ′ (w) KΩ (w). (12.11)
|Ω ′ | Ω ′′
κ1 + iκ2 = −i .
(Ω ′ )2
Next we turn to the curl of K. The complex curvature is the Pólya vector field of
the function (−if ′′ /f ′ |f ′ |), and the presence of |f ′ | in the denominator prevents this
from being analytic. Thus while K has been shown to be divergence-free, it cannot
also be curl-free. What then is its curl?
To find it, recall that
·
∇K = ∇ K + i ∇×K.
Since ∇f ′ = 0 = ∇f ′′ ,
i f ′′ f ′′ 1 f ′′ |f ′′ |2
∇K = ∇ = ′ i∇ = K = −i .
f ′ |f ′ | f |f ′ | f′ |f ′ |3
·
Thus ∇ K = 0, which we already knew, and
|f ′′ |2
∇×K = − .
|f ′ |3
Although we have just differentiated K, note that the result does not depend
on any higher derivatives of f than occur in K itself. Indeed, the result may be
re-expressed as
Since K is geometrically defined by f, and since the curl operator is also geomet-
ric, the curl of K must encode some (presumably simple) geometric information
about the mapping f. Unfortunately, we have not yet succeeded in decoding this
information.
[12.9] Geometric proof that the complex curvature K has stream function Ψ = 1/|f ′ |.
We know from (5.32) that the streamlines of the complex curvature are the level curves
of the amplification, |f ′ |. It follows readily that −i∇Ψ points in the direction of K. Next,
observe that under the mapping f ′ , the movement dz1 along K is mapped to a movement
tangent to the circle f ′ = const., and therefore dz2 is mapped to a radial movement of
d|f ′ | |f ′′ |
length d|f ′ | = |f ′′ (a) dz2 |. Finally, |∇Ψ| = − |dz
dΨ
2|
= |f1′ |2 |dz 2|
= |f ′ |2 = |K|. Done.
The geometrically derived result (5.32), p. 273, says, in part, that the stream-
lines of the complex curvature are the level curves of the amplification, |f ′ |. Figure [12.9]
illustrates this: sections of adjacent streamlines are mapped by f ′ to arcs of origin-
centred circles. It also shows how the infinitesimal complex number dz1 along the
streamline through the point a is amplitwisted by f ′′ (a) to a complex number at
f ′ (a) that points counterclockwise along the circle |f ′ (z)| = |f ′ (a)| = const. Cor-
respondingly, the orthogonal number dz2 = −i dz1 is amplitwisted to a complex
number at f ′ (a) that points radially outwards.
In order for K to be sourceless it must be of the form K = −i∇Ψ, so that ∇Ψ (the
direction of maximum increase of Ψ) must be directed as shown, and Ψ must be
a function of |f ′ |. Since |f ′ | increases in the direction dz2 , while ∇Ψ is in the oppo-
site direction, Ψ must be a decreasing function of |f ′ |. Thus if Ψ is any decreasing
function of |f ′ | then −i∇Ψ will be a sourceless vector field in the direction of K. It
only remains to show that for the particular function Ψ = 1/|f ′ |, the magnitudes also
agree, i.e., |∇Ψ| = |K|.
Let d|f ′ | and dΨ be the changes in |f ′ | and Ψ that result from the movement dz2 .
From the picture we see that d|f ′ | = |f ′′ (a) dz2 |, so that with Ψ = 1/|f ′ |,
dΨ 1 d|f ′ | |f ′′ |
|∇Ψ| = − = ′2 = ′ 2 = |K|,
|dz2 | |f | |dz2 | |f |
as was to be shown.
The Complex Curvature Revisited* 597
[12.10] Geometric proof that −KΩ (w) = f ′ (z) Kf (z). The horizontal line-segment at
z is mapped by f to a segment of curvature κ1 , with unit tangent b ξ . The inverse mapping
Ω unbends this piece of curve, so, by the general transformation law of curvature, (5.31),
KΩ ξ·′
b +κ1 /|Ω ′ | = 0. Therefore, the component of −KΩ in the direction of b
′
ξ is κ1 /|Ω ′ | =
κ1 |f |. Likewise, the orthogonal component is κ2 |f |. Thus −KΩ is obtained by expanding
Kf by |f ′ |, and rotating it by arg(bξ ). But arg(b
ξ ) = arg(f ′ ), so −KΩ (w) = f ′ (z) Kf (z).
Done.
Next we give a more geometric derivation of (12.12). Figure [12.10] shows a hor-
izontal line-segment at z being mapped by f to a segment of curve at w whose
curvature is κ1 and whose unit tangent is b
ξ . The inverse mapping Ω unbends this
piece of curve and sends it back to the straight line-segment at z. Hence, by the
general transformation law of curvature, (5.31), page 272,
·
b + κ1 /|Ω ′ | = 0.
KΩ ξ
as was to be shown.
It is also possible to give a more geometric derivation of the result (12.13). How-
ever, we shall not bother to do this since we have not yet been able to establish the
significance of that result.
598 Flows and Harmonic Functions
[12.11] Finding the flow around a given obstacle, B. Given a frictionless flow, imagine
that we instantly freeze the shaded region between streamlines, turning it into an obstacle,
B, around which the flow continues, undisturbed. The image of the boundary under the
conformal complex potential is a horizontal line, and, conversely, the problem of finding
the flow around a given B therefore amounts to finding a conformal mapping Ω with this
property.
12.5.1 Introduction
Consider a typical fluid flow such as [12.11]. From our assumption that the fluid
has no viscosity [this is most certainly an idealization] it follows that if we were to
suddenly freeze the fluid within a flux tube, such as the shaded region in [12.11],
then the unfrozen fluid would continue to flow in exactly the same way as before.
The same idea applies if [12.11] instead represents flowing heat: if the shaded region
of the metal plate were suddenly replaced with material which did not conduct
heat then there would be no disturbance to the flow of heat in the remainder of the
plate. This is much less of an idealization than in the case of fluid flow.
Conversely, if we insert an obstacle into a flow then the new disturbed flow must
be such that the boundary B of this obstacle is a streamline, or is made up of seg-
ments of streamlines. If we think of the complex potential Ω of this disturbed flow
as a mapping, this means that Ω maps B to a horizontal line, or to segments of a
horizontal line.
The problem of finding flows around a given B therefore amounts to finding
conformal mappings Ω with this property. In fact, since the complex potential of
a given flow is only defined up to a constant, we may further demand that the
horizontal image line be the real axis. Alternatively, this characterization may be
Flow Around an Obstacle 599
[12.12] The flow around a disc can be realized as the previously considered super-
position of a uniform flow with a dipole, with complex potential Ω(z) = z + (1/z).
12.5.2 An Example
As our first example of a flow in the presence of a barrier, reconsider the case of
Ω(z) = z+(1/z). See [12.12]. As we previously discussed, this represents a uniform
flow to the right into which has been inserted a dipole at the origin. On the other
hand it certainly appears that the picture can be interpreted in a second way, as
the flow around a circular obstacle (the shaded unit disc) inserted into a uniform
flow. [Later we shall see that it is no accident that the flow may be interpreted in
these two ways.] That both the top and bottom halves of the unit circle are indeed
segments of the streamline Ψ = 0 is easily seen from the fact that Ω maps them to
the segment −2 ⩽ x ⩽ 2 of the real axis:
Note the breakdown of the grid at ±1 and check that this corresponds to the fact
that these are the stagnation points of the flow.
While this is a possible flow round C that is uniform at infinity, it is not the only
such flow. First let us discuss this fact in terms of the flow round an obstacle of
arbitrary shape.
600 Flows and Harmonic Functions
For any flow round an obstacle, the boundary curve B is composed of stream-
lines, so B has zero flux. If the obstacle has been inserted into a uniform flow then
there are no singularities outside B (except at infinity), and it follows from the
deformation theorem for flux [(11.9), p. 547] that if B is deformed into any loop
round the obstacle then the flux through it continues to vanish. Loops that do
not enclose the obstacle may be contracted to a point without crossing any sin-
gularities, so their fluxes also vanish. Since all loops have vanishing flux, the flow
is not merely locally sourceless, it is totally sourceless. Put differently, this says
that the flux crossing any curve between two points in the flow is independent
of the path.
It is a different story for the circulation/work round B, for there is no a priori
reason for this to vanish. Let S denote the value of this circulation. By the Defor-
mation Theorem (11.10), p. 547, the circulation round any simple loop enclosing
the obstacle will also equal S, while the circulation round loops that do not enclose
the obstacle will vanish. Put differently, this says that the circulation along a path
between two points in the flow is dependent on the choice of that path. The circu-
lation Φ(z) along a path from a fixed point to a variable point z is a multifunction
of z: if the difference of two paths is a loop winding round the obstacle n times, the
difference of the two values of Φ(z) will be nS.
We may now state (without proof) the uniqueness property of flows round a
given obstacle when it is inserted into a uniform flow of given velocity: for each
value of the circulation S there is precisely one flow. In particular, there is a unique
flow that is totally irrotational, i.e., for which S = 0. For a circular obstacle, this is
the one shown in [12.12].
In the case of the disc it is easy enough to construct each of the flows for which
S ̸= 0. We need only superpose the totally irrotational flow in [12.12] with the flow
of a vortex of strength S at the origin. Figure [12.13] illustrates this flow for a small
value of S. The reader is strongly encouraged to use a computer to verify this figure.
As you gradually increase the value of S, notice how the stagnation points on the
circle move towards each other and finally coalesce at i. At what value of S does
this occur? Verify your empirical answer with an exact calculation. Increasing the
value of S still further, this single stagnation point moves off the circle and up the
imaginary axis, and we obtain the qualitatively different flow shown in [12.14].
Returning to the totally irrotational flow in [12.12], figure [12.15] is an attempt
to illustrate the geometry of its complex potential in greater detail than before. The
top half is essentially just a copy of [12.12], while the bottom half illustrates the
image under the complex potential mapping. Though the figure is intended to be
largely self-explanatory, we make the following observations:
• If we choose to measure circulation and flux along paths emanating from p, then
Ω(p) = 0.
Flow Around an Obstacle 601
[12.13] For each value of the circulation S around an obstacle, there exists a unique
flow. If S = 0 then the flow is given by the totally irrotational flow in the previous figure.
If S ̸= 0, the unique flow is the superposition of the S = 0 flow with an origin-centred
vortex of circulation S. This figure depicts the case of a small value of S.
[12.15] A more detailed view of the complex potential mapping of the irrotational
flow [12.12].
disc and the exterior of the line-segment connecting Ω(±1). [The latter region is
often described as a plane with a cut along this segment.]
• The bottom half of the figure illustrates two routes [black/white arrows] by
which the dot-filled shape may be moved to the brick-filled shape. The top illus-
trates their images under Ω−1 (as defined above). Both routes yield the same
image for the brick-filled shape, but the price we pay is that Ω−1 is not even
continuous (let alone analytic) on the cut: witness the fate of the black shape as
we follow the white route. However, consider the following alternative. Again
following the white arrows, sketch what happens if we analytically continue Ω−1
as we cross the cut. Hint: look at the flow inside the unit circle in [12.12].
• The fact that we have ended up with two different images for the brick-filled
shape reflects the fact that the two routes enclose a branch point at Ω(1). This
Flow Around an Obstacle 603
[12.16] Electrostatic interpretation of the dual of the irrotational flow around the
disc. A cross section of a copper cylinder inserted into a uniform electric field: the former
streamlines now become equipotentials. The surface of such a conductor must itself be
an equipotential, with the static electric field running perpendicular to it, or else charges
would experience a force acting within the surface, causing them to move.
We now turn to two more physical interpretations of [12.12]. First, we may view
the two-dimensional flow in [12.12] as being a cross section of a genuine three-
dimensional flow. Suppose we were to make a few thousand photocopies of this
figure and stack them neatly one on top of the other. The shaded discs would fit
together to form a cylinder perpendicular to the stack, and the streamlines would
represent the flow round this cylinder. Of course a real cylinder has ends, and,
when it is inserted into a uniform flow perpendicular to its axis, the flow on a plane
that is close to one of these ends will no longer look like [12.12].
To give the second interpretation we must explain a previous remark, namely,
that the dual of a flow round an obstacle is the electric field which solves an anal-
ogous problem in electrostatics. The dual of the flow in [12.12] is shown in [12.16].
Since C was a streamline of the original flow, it is now an equipotential of the dual
flow.
Suppose we insert a long copper cylinder into a uniform electric field perpen-
dicular to its axis. Almost instantly, the free electric charges within the cylinder
will settle themselves into an equilibrium distribution such that the electric field
becomes steady (“electrostatic”). Figure [12.16] then represents a cross section of
604 Flows and Harmonic Functions
this electrostatic field on a plane that cuts perpendicularly through the conductor
somewhere in the middle. [Again, the field near the ends will be different.]
Here’s why. Just as we require Ψ to be constant on the boundary of an obsta-
cle in a fluid flow, so we require Φ to be constant on the boundary of a conductor in an
electrostatic field. This is equivalent to demanding that the electric field be perpen-
dicular to the boundary of the conductor, and it is not hard to see why the latter
must be true. For if the electric field E = ∇Φ were not perpendicular, it would
have a nonzero component within the surface of the conductor. But this means
that the free charges there would experience a force and would move, contradicting
our assumption that the field has settled into an electrostatic one.
The figure also sketches the equilibrium distribution of negative ⊖ and positive
⊕ charge on the surface of the cylinder. This distribution is such that the charge
density is proportional to the (signed) strength of the electric field at the surface. If
a phase portrait has been divided into squares, as this one has, this implies [why?]
that the charge density is proportional to the density of field lines (per unit length)
leaving the conductor’s surface. To learn more of the physics of electrostatics, see
Feynman (1963); for a geometric approach, see Maxwell (1881).
Clearly the disturbed complex potential Ωd (which we seek) must have stream-
lines which look something like those in [12.18]: they closely resemble those of
Ωu near the singularity, and the barrier is a streamline. That [12.18] is actually the
exact solution follows from the illustrated fact that equipotentials [dashed] may be
drawn so as to divide this flow into squares. We now describe how this Ωd may be
found by the method of images.
Flow Around an Obstacle 605
[12.17] The flow resulting from inserting the unit disc into a dipole field.
[12.18] The flow of a source in the presence of a barrier along the real axis.
As we saw in Chapter 5, Schwarz’s Symmetry Principle (p. 286) says that a func-
tion which is analytic on one side of the real line, and which takes real values on that
line (e.g., Ωd ), may be analytically continued to the other side by taking conjugate
points to have conjugate images. This is vividly clear in [12.18]: the grid of squares
606 Flows and Harmonic Functions
[12.19] The Method of Images with a linear boundary. To find the flow depicted in
the previous figure, we superpose the original source with its image (of equal strength)
after reflection across the real axis barrier. More generally, if Ωu (z) is the undisturbed
complex potential of a set of multipoles in the upper half-plane, without any barrier, then
Ω⋆u (z) ≡ Ωu (z) represents the undisturbed flow of the mirror-image multipoles in the
lower half-plane. Then the method of images tells us that in the presence of the real axis
barrier, the disturbed flow is the superposition of the original flow and its mirror
image: Ωd (z) = Ωu (z) + Ω⋆u (z).
in the upper half-plane may be continued to the lower half-plane by reflecting the
grid in the real axis. The complete flow is thus given by [12.19].
Intuitively, we have found that the flow in the presence of the barrier may be
obtained by removing the barrier and instead inserting another source of the same
strength as the original, but located at the mirror image in the real axis. Thus Ωd is
the superposition of these two sources:
the superposition of the undisturbed flow of this set of multipoles and the undis-
turbed flow of their mirror images. Note that while the mirror image multipoles
will be of the same type and strength as the originals, the direction of their mul-
tipole moments will be different. For example, a dipole in the direction (3 − 2i)
reflects to one in the conjugate direction (3 + 2i). More generally, a multipole with
multipole moment Q reflects to one with moment Q.
Now let us turn this method into a formula. In Chapter 5 we showed that
from a given analytic mapping f(z) we can produce a new analytic mapping f⋆ (z)
according to the recipe
The physical significance of this new analytic function is easy to see: if Ωu again
represents the complex potential of a superposition of multipoles in the upper half-
plane, then Ω⋆u will be the undisturbed complex potential of their mirror images
in the lower half-plane. The formula for the disturbed complex potential is thus
Note that this formula does indeed satisfy Schwarz’s Symmetry Principle: Ω⋆d (z) =
Ωd (z), so the real axis is a streamline. Naturally, if Ωu instead represents a collec-
tion of multipoles in the lower half-plane, then (12.16) is again the solution in the
presence of the barrier.
Essentially the same method may be used to find the disturbed flow in [12.17],
in which the barrier is now a circle rather than a line. Reconsider that figure. We
have drawn the streamlines Ψ = const. at random, but had we instead chosen the
values of Ψ in arithmetic progression (as we did in [12.12]) then it would have been
possible to divide this flow into a grid of infinitesimal squares, with the unit circle
being comprised of edges of these squares. As we saw in Chapter 5, it is again
possible to extend this grid across the barrier (cf. [12.12]), but to do so we must
replace reflection in a line by its analogue for circles, namely, inversion.
Performing this inversion in the unit circle we obtain [12.20], the dipole at 2
inverting to another dipole at (1/2). It is now clear that to find Ωd we should
remove the barrier and superpose the undisturbed flows of these two dipoles. But
what is the undisturbed complex potential of this new dipole at (1/2)?
As we saw in Chapter 5, if reflection in the real line is replaced by inversion in
the unit circle then the recipe (12.15) may be modified to generate a new analytic
function f† given by
† 1
f (z) ≡ f . (12.17)
z
The physical significance of this new analytic function is much as before: if Ωu
represents the complex potential of a superposition of multipoles outside the unit
608 Flows and Harmonic Functions
circle, then Ω†u will be the undisturbed complex potential of their images under
inversion. The analogue of (12.16) is now
Let us apply this method to find the disturbed complex potential of [12.20] (and
hence of [12.17]). If Ωu is given by (12.14) then
(1 − i) (1 − i)z
Ω†u (z) = = .
(1/z) − 2 1 − 2z
That this is indeed a dipole at (1/2) may be seen by rewriting it as
(1 − i) 1 (1 − i)
Ω†u (z) = − − .
4 z − (1/2) 2
Because constants have no effect on the flow, this is a dipole at (1/2) with dipole
moment (1 − i)/4. Unlike the case of reflection across a line, note that it is not only
the direction of the dipole moment which is affected by the inversion, but also its
magnitude: here the strength of the inverted dipole is one quarter the strength of
the original. Superposing the two dipoles we obtain
(1 + i) (1 − i)z
Ωd (z) = Ωu (z) + Ω†u (z) = + ,
z−2 1 − 2z
and you may use a computer to verify that this formula does yield the flow in
[12.20].
Like [12.19], figure [12.20] has symmetry, but it is of a more subtle kind than
before: by construction, the figure reproduces itself under inversion in the unit cir-
cle. This symmetry can be made to leap from the page by projecting [12.20] onto the
Riemann sphere. As we learnt in Chapter 3, inverting in the unit circle is equivalent
to reflecting the Riemann sphere in its equatorial plane. The flow on the northern
and southern hemispheres should therefore be mirror images of each other in this
plane. Behold figure [12.21]! Note that on the sphere the strengths of the two dipoles
become equal.
We can now see that [12.12] is simply a limiting case of [12.21], for as the southern
dipole moves towards the south pole (0), its reflection moves towards the north
pole (∞), and a solitary dipole at ∞ projects to a uniform flow in the plane.
In Chapter 5 we saw that both reflection in a line and inversion in a circle were
special cases of Schwarzian reflection in a curve; see [5.30], page 290. We now use
this fact to generalize the method of images. For example, suppose that we insert
an ellipse E into a uniform flow. Clearly the flow will look something like [12.22],
and that this is in fact the exact flow is once again apparent from its divisibility into
squares. How did we find it?
The generalization of the formulae (12.15) and (12.17) to the case of Schwarzian
reflection RK in a curve K is
For example, if K is the real line (so that RK (z) = z) then f‡ = f⋆ , while if K is
the unit circle (so that RK (z) = 1/z) then f‡ = f† . Thus if Ωu is the undisturbed
610 Flows and Harmonic Functions
[12.21] Projection of the previous figure onto the Riemann sphere. By construction,
the previous figure is symmetric under inversion in the unit circle, which induces reflection
of the Riemann sphere in its equatorial plane. So the flow on the northern and southern
hemispheres should therefore be mirror images of each other, and indeed we see that they
are! Note that the unequal strengths of the original dipoles in the plane project to dipoles
of equal strength on the sphere. Also observe that [12.12] may now be understood as a
limiting case of this figure, for as the southern dipole moves towards the south pole (0),
its reflection moves towards the north pole (∞), and a solitary dipole at ∞ projects to a
uniform flow in the plane.
and you may use a computer to verify that this formula yields figure [12.22].
Note, however, that everything is not quite as it seems. While Ωu is a uniform
Flow Around an Obstacle 611
[12.22] The Method of Images with a general boundary via Schwarzian reflection.
To find the illustrated flow around the ellipse, E, we superpose a steady eastward flow
with its Schwarzian reflection RE across E; see [5.31]. More generally, if Ωu (z) is the
undisturbed complex potential of a set of multipoles on one side of a curve K, without
any barrier, then Ω‡u (z) ≡ Ωu [RK (z)] represents the undisturbed flow of the Schwarzian
reflections of the multipoles on the other side of K, and vice versa. Then the method of
images tells us that in the presence of the barrier, K, the disturbed flow is the super-
position of the original flow and its mirror image: Ωd (z) = Ωu (z) + Ω‡u (z). This
subsumes both previous methods of images as special cases, for if K is the real axis then
RK (z) = z and f‡ = f⋆ , while if K is the unit circle then RK (z) = 1/z and f‡ = f† .
flow to the right with speed 1, the behaviour of Ωd for large values of |z| is
a uniform flow to the right with speed (4/3). Of course if we wish the flow
to have unit speed far from the ellipse then we need only multiply this Ωd
by (3/4).
To demonstrate the practicality of mapping one flow onto another, let us return
to the problem of finding the flow round an obstacle, restricting our attention to the
case where the obstacle is inserted into a uniform flow. For example, let us rederive
the flow round the ellipse (x/2)2 + y2 = 1 in [12.22]. Suppose that we knew of a
one-to-one conformal mapping w = f(z) from the exterior of the unit circle C in
the z-plane to the exterior of this ellipse E in the w-plane. We already know the
flow round C when it is inserted into a uniform flow of velocity 1, so applying f
to these streamlines yields some flow round E. If we want the flow round E to be
uniform far from E then we must further demand that f(z) behave like a multiple
of the identity far from C: f(z) ≈ cz if |z| is large. Assigning equal values of Ψ to
the original and image streamlines, the image flow far from E will then be uniform,
with speed (1/|c|) and direction c [why?].
Recall figure [5.22], p. 276. As z = eit describes C, w = p eit + q e−it describes
the ellipse (x/a)2 + (y/b)2 = 1, where p = (a + b)/2 and q = (a − b)/2. Thus in
the case of (x/2)2 + y2 = 1, the mapping we seek is
1 1
w = f(z) = 2 3z + z . (12.19)
Let’s go through the details and check that we recover the flow in [12.22]. Since
the flow round C has complex potential Ω(z) = z + (1/z), and since the complex
e
potential Ω(w) at the image w = f(z) of z is defined to be the same as Ω at z, we
e
have Ω(w) = z + (1/z). To express this as an explicit function of w we solve (12.19)
for z and obtain [why the choice of +?]
p
z = 31 w + w2 − 3 .
e
Although we could immediately insert this into the formula for Ω(w), we may save
ourselves a little algebra by first noting that (12.19) implies (1/z) = 2w − 3z. Thus
p
e
Ω(w) = z + (1/z) = 2(w − z) = 32 2w − w2 − 3 .
Apart from the factor of (2/3), signifying that the velocity is (2/3) far from E
(as anticipated), this is the same formula we previously obtained by Schwarzian
reflection.
As another illustration of this idea, suppose that E were instead inserted into a
uniform flow in the direction of eiϕ . To find the flow round E we need only find
the flow round C when it is inserted into such a flow, and then apply f. Since the
undisturbed complex potential is Ωu (z) = e−iϕ z, the method of images says that
the flow round C is
eiϕ
Ωd (z) = Ωu (z) + Ω†u (z) = e−iϕ z + .
z
With ϕ = (π/4) this flow is illustrated on the left of [12.24]. On the right is
the desired flow round E obtained by applying f to the flow round C. Use your
computer to verify this figure.
In this manner, we may derive the flow round an infinite variety of obstacles:
choose any analytic mapping w = f(z) which is one-to-one outside C and which
e
behaves like cz for large |z|, then Ω(w) = f−1 (w) + [1/f−1 (w)] is the flow round an
obstacle whose boundary curve B is f(C).
12.6.1 Introduction
Recall that Riemann’s Mapping Theorem [p. 204] asserts that any simply connected
region R (other than the entire plane) may be mapped one-to-one and conformally
to any other such region S. Granted this, we saw that no loss of generality results
from taking S to be the unit disc D, and that there must be as many mappings from
R to S as there are automorphisms of D. Later [p. 407] we showed that these auto-
morphisms are the hyperbolic rigid motions, which have three degrees of freedom.
614 Flows and Harmonic Functions
[12.24] Conformally mapping one flow onto another. Using the mapping obtained in
the previous figure, the known flow on the left around the unit disc may be conformally
mapped to the flow around the ellipse on the right. [NOTE: In the text, we confirm that
this new approach yields the same answer that we formerly obtained via the method of
images.] In this manner, we may derive the flow round an infinite variety of obstacles:
choose any analytic mapping w = f(z) which is one-to-one outside C and which behaves
e
like cz for large |z|, so that the flow becomes uniform, then Ω(w) = f−1 (w) + [1/f−1 (w)]
is the flow round an obstacle whose boundary curve B is f(C).
Thus the complete result we seek to understand is this: there exists a three-parameter
family of one-to-one, conformal mappings between R and D.
There are at least two standard proofs of this fundamental result, and most
advanced books on complex analysis include one of these. Despite the fact that
one of the arguments (due to Koebe) is constructive in nature—and therefore, in
principle, comprehensible—we have not yet found a way to present it in a manner
consistent with the aims of this book. However, the interested reader will find an
excellent description of the idea underlying Koebe’s proof in Hilbert (1952), and a
clear description of the technical details in Nehari (1952) or Nevanlinna and Paatero
(2007). To our knowledge, the deepest investigation of this idea is that given by
Henrici (1991).
On a brighter note, the above ideas on flows will enable us to gain considerable
insight into both the existence of such mappings and the fact that they have three
degrees of freedom. We shall do so by reversing the idea of conformally mapping
one flow onto another. That is, by resorting to physical experiment one may obtain
flows, and these may then be used to construct conformal mappings.
The Physics of Riemann’s Mapping Theorem 615
[12.25] Constraint on the mapping between flows. Divide the flow around R and the
flow around the unit disc D into small, square k-cells. If the illustrated point z is mapped
by f to w = f(z), then if z moves four squares downstream and two squares “up” the
equipotentials, w must do the same. But the sought-after conformal mapping w = f(z)
has a constraint: it must map the stagnation points a and b to the stagnation points a e
e This is because the geometric signature of the critical point b is the angle of (π/4)
and b.
between the streamline and equipotential through b, and this property is preserved by a
conformal mapping.
616 Flows and Harmonic Functions
then, as we have previously remarked, the flow is unique. Next, we have arbitrarily
chosen a point [not shown] from which to measure circulation and flux, i.e., a point
at which the potential Φ and stream function Ψ both vanish. With a small value of k,
we have then constructed the k-flux tubes and k-work tubes, thereby dividing the
exterior of R into small, approximately square, k-cells. As usual, we may imagine
shrinking the value of k to zero, so that ultimately the k-cells are square.
As illustrated, let a and b be the two stagnation points on the boundary of R
(ordered so that the flow passes from a to b), and let S1 and S2 be the two segments
of boundary streamline connecting them. Since the flow is totally irrotational, the
circulations along S1 and S2 must be equal to each other, the common value being
the potential difference [Φ] between a and b:
Put geometrically, this says that the number of squares abutting S1 and S2 must be
equal to each other, this number being given by [Φ]/k.
Now insert the unit disc D into a uniform flow with velocity v eiϕ and divide up
the resulting flow into k-cells, using the same value of k as before. Let us employ
e
tildes to denote corresponding entities in the new flow: the stagnation points are a
e e e
and b, the segments of streamline connecting them are S1 and S2 , and the potential
e and Ψ.
and stream functions (relative to an arbitrarily chosen point) are Φ e
The method of conformally mapping points z in the exterior of R to points w =
f(z) in the exterior of D now seems clear: identify “corresponding” squares in the
two grids! Let us control our excitement and think this through. Presumably, by
“corresponding” we mean that once we know the image w of one grid point z then
the image of any other is determined by the pair of grids: if z moves four squares
downstream and two squares “up” the equipotentials, then w does the same. See
[12.25].
However, we cannot arbitrarily choose to map a particular z to a particular w,
for the sought-after conformal mapping w = f(z) must map the stagnation points a and b
e This is because the geometric signature of the critical
e and b.
to the stagnation points a
point b is the angle of (π/4) between the streamline and equipotential through b,
and this property is preserved by a conformal mapping.
Now the snag is this: defining the image of a to be a e if
e (as we must), b will map to b
e
and only if the number of squares abutting S1 is equal to the number abutting S1 . Looking
closely, we see this is not the case in [12.25]: there are fewer squares along S e1 than S1 .
To correct the situation we must slightly increase the speed v of the flow round the
disc. More precisely, v must be chosen so that the potential difference [Φ] e across
D equals the potential difference [Φ] across R. Since we know that [Φ] e = 4v, we
deduce that
A one-to-one, conformal mapping between the two grids is obtained if
and only if the disc is inserted into a flow of speed v = [Φ]/4.
The Physics of Riemann’s Mapping Theorem 617
Although we hope our use of grids has helped to make the mapping vivid (and
to show that it is both one-to-one and conformal), we should perhaps point out that
the grids are in no way essential to the definition of f. All we are doing is identifying
points by means of circulation and flux: the image w = f(z) of z is defined by
e
Ω(w) e a
= Ω(z) + Ω( e ) − Ω(a) ,
In other words, w and z correspond if and only if their circulation and flux are
equal.
In this case the mapping may be written as f = Ω e −1 ◦ Ω, which we may interpret
e
as follows. Refer to [12.25] and [12.15]. The complex potentials Ω(z) and Ω(w) map
points lying strictly outside R and, respectively, D to points in a plane that is slit
along the real axis from 0 to [Φ]. Thus f may be thought of as first mapping the
exterior of R to the slit plane by means of Ω, then mapping the slit plane to the
exterior of D by means of Ω e −1 .
We now return to the mappings themselves and ask, how “many” of them do we
obtain by means of this construction? We begin by explaining the illusory nature
of some of the apparent freedoms in the construction. First, why not insert R into a
uniform flow of arbitrary speed? Of course we can, but nothing new results from
doing so. For example, if we double this speed then we must also double v, because
we must maintain equality between the number of k-cells along S1 and S e1 . But this
yields the same mapping f as before. Second, in constructing the grid outside D,
why not use a different value of k, say e k, from that used outside R? In order to
maintain equality between the number of squares along S1 and S e1 we would then
e
have to change v to k[Φ]/4k [why?], and this would produce the same grid and
mapping as before.
Clearly, however, we do obtain new mappings by varying the direction ϕ of the
flow round D. If F denotes the particular mapping f corresponding to ϕ = 0, then
the mapping Fϕ corresponding to a general value of ϕ is Fϕ (z) = eiϕ F(z), namely,
F followed by a rotation.
618 Flows and Harmonic Functions
It would seem plausible that still other mappings could be obtained by varying
the direction—let us call it θ—of the flow into which R is inserted. Not so. This fol-
lows from the fact that the flow round R in this case may be obtained by applying
F−1 to the flow round the disc when it is inserted into a uniform flow in the direction
θ. Clarify this in your own mind by referring to figures [12.23] and [12.24]. Thus the
mapping between the two flows is only sensitive to the angle between their direc-
tions: inserting R and C into flows with directions θ and ϕ, respectively, yields the
mapping F(ϕ−θ) .
Since ϕ is the only genuine degree of freedom, the mappings we have con-
structed belong to a one-parameter family. We are therefore missing two degrees
of freedom. Though we will explain this mystery shortly, you may care to think
about it on your own before reading further.
Granted the generosity of Nature in providing such flows round obstacles, we
now have a physical method of determining f, but we lack a mathematical proce-
dure for doing so. This is a very hard problem. However, there does exist an explicit
formula for f in the case that R is bounded by a polygon. This is called the Schwarz-
Christoffel formula; see Nehari (1952) or Pólya and Latta (1974) for good discussions
of the result. We shall not enter into this here, except to say that the advent of high-
speed computers has opened the way to approximating f by approximating the
boundary of a given R with a polygon; see Trefethen (1986) and Henrici (1991) for
this and other algorithmic approaches to the problem.
4
Ideas related to those which now follow may be found in Bak and Newman (2010), Siegel (1969,
p. 148), and especially Courant (1950). Although Riemann himself employed physical reasoning, the
idea of relating his mapping theorem to dipoles seems to have originated with Hilbert (1965, Vol. 3,
pp. 73–80).
The Physics of Riemann’s Mapping Theorem 619
[12.26] Interior mappings and dipoles. Referring to the previous figure, inversion of R
in a circle centred at the arbitrarily selected interior point q yields [a], while inversion of D
in the unit circle yields [b]. In both cases, we obtain dipoles. In this way, the one-to-one,
conformal mapping between the exteriors of R and C now yields another such mapping
between the interior dipole flows of R† and C: for example, the black T-shape is mapped
to the black T-shape.
of fluid. In fact [exercise] since the original uniform flow had unit speed, the dipole
at q has unit strength.
In this way, the one-to-one, conformal mapping between the exteriors of R
and C now yields another such mapping between the interiors of R† and C. For
example, the stagnation point α maps to the stagnation point α e , q maps to 0, and
†
the black T-shape inside R maps to the one inside C. Although the speed of the
flow grows arbitrarily large as we approach q, there is nothing dramatic about the
behaviour of the mapping near q. For example, imagine sliding the T-shape along
the streamlines towards q, and consider its image.
We may look at this construction rather differently. Suppose that R† is a given,
fixed curve whose interior we wish to map to the unit disc. We may obtain such a
mapping f as follows:
(i) Insert a horizontal dipole of unit strength into the interior of R† at an arbi-
trary interior point p. Divide the interior into k-cells of the flow, and let N
denote the number of them abutting R† .
(ii) At the centre 0 of C, insert a dipole of strength d, and direction ϕ. Divide the
interior of C into the k-cells of this flow, and adjust the strength d until the
number of them abutting C is equal to N. If [Φ] again denotes the circulation
620 Flows and Harmonic Functions
This one-to-one, conformal mapping f from the interior of R† to the unit disc
is, in fact, the most general such mapping. This may be seen from the fact that
the construction has the full three degrees of freedom: two for the location p of the
dipole in R† , and one for the direction ϕ of the dipole in C. It is not hard to see
the geometric significance for f of these physical degrees of freedom: p is the point
which is mapped to the centre of the disc, i.e., f(p) = 0; and ϕ is clearly the twist
of f at p, i.e., ϕ = arg[f ′ (p)].
Note that while the twist of f at p is freely specifiable, the amplification |f ′ (p)| is
not. In fact [exercise, or read on] |f ′ (p)| is simply the strength d of the dipole in C,
and we have seen how the latter is fixed in the course of the above construction.
Put differently, |f ′ (p)| = [Φ]/4.
We can now return to the construction in [12.25] and see why it is missing two
degrees of freedom. To obtain the general mapping between the exteriors of R and
D, we may take this newly constructed general mapping between the interiors of
R† and C, then reverse (i.e., repeat) the inversions which took us from [12.25] to
[12.26].
In the above process of generalization the fundamental step was moving the
dipole at q to an arbitrary interior point p. Since the flow in [12.26b] did not undergo
any fundamental change, inversion returns it to a flow round D like that shown in
[12.25]. However, to send R† back to the boundary of R we must invert in the circle
of unit radius centred at q (which we now take to be the origin), and the flow inside
R† of the dipole at p inverts to the flow outside R of a dipole placed at (1/p). Figure [12.17]
illustrates such a flow when R is a disc.
The construction in [12.25] is now recognizable as the special p = q in which the
dipole outside R has been placed at ∞ rather than at a general point. Here are our
two missing degrees of freedom: there was nothing wrong with choosing to place
one of the dipoles (the one outside D) at ∞, but we then (unnecessarily) insisted
on placing the other dipole there too. In terms of the resulting mapping between
the exteriors, this amounted to insisting that ∞ map to ∞, whereas the general
construction allows us to map any point outside R to ∞ by placing a dipole there.
[12.27] Mapping interiors with vortices instead of dipoles. Place a source of strength
S at p inside B, and a source of strength S e at the centre of C. Dividing each flow into
small, square k-cells, we readily obtain a conformal mapping from one interior to the
other, with p mapping to 0. But this is only possible if the number N of k-cells abutting
B equals the number N e abutting C. This implies that the two vortices must be of equal
strength: S = Nk = Nk e = S. e Finally, we may specify which boundary point a e of C
corresponds to the boundary point a of B. Now the mapping is completely tied down,
and, for example, the black T-shape is mapped to the black T-shape.
5
Of course we could insert several sinks whose strengths summed to that of the source, but that
would be messier still.
622 Flows and Harmonic Functions
Firstly, in defining the correspondence between the grids we are forced to map
p to 0. Secondly, having chosen the strength S of the vortex in B, the strength S e of
the one in C must be chosen so that (as illustrated) the number N e of k-cells abut-
ting C is equal to the number N abutting B. This is the same idea as in the dipole
construction, but the answer in this case is much simpler: put S e equal to S.
The explanation is as simple as the answer. Concentrate on one of the vortices,
say the one in B. Its strength S is, by definition, the circulation along any simple
loop round p, and we may conveniently take this to be one of the streamlines. The
equipotentials cut this streamline into N segments, each having circulation k. Thus
S = Nk, and likewise S e = Nk.
e The condition S e = S follows immediately.
Finally, observe that we have not yet pinned down the mapping. In both [12.25]
and [12.26] there were stagnation points which we were forced to identify, thereby
tying down the mapping. However, here there are none, and we must do the job by
hand. A common procedure goes like this. We know that the streamline B maps to
the streamline C, and we may now insist that a particular point a on the first maps
to a particular point a e on the second.
If we wished to be more definite, we might do the following. Consider the
heavily-dashed equipotential in B which exits p travelling due east, and choose
a to be its intersection with B. See [12.27]. If we now choose a e = eiϕ then the
mapping f between the two grids is completely tied down. For example, the black
T-shape in B maps to the black T-shape in C. The only advantage of specifying a
and a e in this particular way is that ϕ then has a simple interpretation in terms of f,
for it is clearly the twist of f at p: ϕ = arg[f ′ (p)].
As with the dipole method, we have obtained the full three-parameter family of
mappings: two for the point p that f maps to the centre of C, and one for the twist
of f at p.
Now let us turn to other physical interpretations of this construction. Taking
the dual of the flows in [12.27], we obtain [12.28]. The equipotentials have become
streamlines, and the streamlines (B and C in particular) have become equipoten-
tials. By the same reasoning as before, the strengths of the two sources manifest
themselves geometrically as k times the number of streamlines emanating from
each, and these must be set equal to each other in order to construct the conformal
mapping between the two grids.
Let us digress briefly. Since the strength of a source is measured by the number
of k-flux tubes emanating from it, a conformal mapping sends a source to another
source of equal strength. Obviously, the same goes for a sink. Returning to [12.26],
we can understand why the strength d of the dipole inside C is the amplification
of f at p. The dipole of unit strength at p may be thought of as a doublet in which
the source and sink are an infinitesimal distance ϵ apart, and each has strength
(1/ϵ). The conformal mapping f sends this to another doublet at 0: the strengths of
The Physics of Riemann’s Mapping Theorem 623
[12.28] The source construction dual to the vortex construction of the previous
figure. In order to make physical sense, here we must replace the fluid flow with an
electrostatic field, the sources being electric charges. By the same reasoning as before,
the strengths of the two charges manifest themselves geometrically as k times the number
of electric field lines emanating from each, and these must be set equal to each other in
order to construct the one-to-one conformal mapping between the two grids.
the source and sink are preserved, but their separation (and hence the strength of
dipole) is amplified by |f ′ (p)|, as was to be shown.
Returning to the physical interpretation of [12.28], we have already observed
that such pictures of sources make no sense when thought of as fluid flows. How-
ever, they do make perfect sense when thought of as electrostatic fields. Imagine
that the dark regions in [12.28] represent cross-sections of blocks of copper through
which we have bored “cylindrical” holes with cross-sections given by B and C.
Now imagine that p and 0 are the cross-sections of two long, very thin, uniformly
(and equally) charged wires running down these holes. The electrostatic fields gen-
erated by these wires automatically have B and C as equipotentials, so [12.28]
faithfully represents these fields, with the streamlines being the electric field lines.
We may also interpret [12.28] in terms of heat flows. Imagine that the white
shapes bounded by B and C have been cut from a sheet of heat-conducting metal,
and imagine that the dark regions are filled with ice, thereby maintaining B and C
at constant temperature (i.e., potential). If heat is introduced at a steady rate at p
and at 0 then the flow of heat will eventually settle down to the one in [12.28], with
the dashed equipotentials being the isotherms.
It must be observed, however, that a point source of heat is a much less physical
concept than its electrostatic analogue, which may be realized (as we have said)
by a very thin charged wire. The reason is that the potential function Φ becomes
624 Flows and Harmonic Functions
[12.29] Using flows to construct the most general automorphism of the disc. Return-
ing to [12.27], suppose we choose B = C to be the unit circle. Then the vortex flow shown
here in [a] may be viewed as a one-to-one conformal mapping to the flow in [b], i.e., it is
an automorphism of the disc. Recall that we previously used the Symmetry Principle to
find the formula for the most general such automorphism: (3.52), on page 201. We can
now rederive this formula by finding the flow in [a], which we accomplish by means of
the method of images, in the next figure.
6
However, note that we are not using (12.18). See Ex. 14 to understand why.
The Physics of Riemann’s Mapping Theorem 625
[12.30] Using the method of images to construct the most general automorphism of
the disc. To obtain the flow in [12.29a], we superpose a vortex at p with an equal and
opposite vortex at its reflection, 1/p.
in C, namely, a fictitious vortex of equal and opposite strength at (1/p). See [12.30].
The complex potential of [12.29a] is thus
Of course we have not done anything really new, for we have used the method of
images, and this is merely a disguised form of the Symmetry Principle by means of
which the result was originally obtained. Nevertheless, we hope you have found it
instructive—and delightful!—to be able to understand these automorphisms from
a new, physical point of view.
[12.31] The Green’s Function Gp (z) of a region R is defined to be the harmonic equi-
librium temperature distribution when heat is supplied at constant rate 2π to the point p
while holding the temperature all around the boundary B at the constant value 0. We may
then construct a harmonic dual Hp (z), that is, a harmonic function whose level curves are
the paths along which the heat flows (orthogonally through the isotherms Gp = const.)
from p to the boundary. Thus knowledge of Gp is sufficient for the construction of the
complete complex potential, Ω(z) = −[Gp (z) + i Hp (z)], and the conformal mapping of
R to D is then explicitly given by w = f(z) = eΩ(z) = e−G e−iH .
Gp (z) ≈ − ln ρ, (12.23)
Gp (z) = − ln ρ + gp (z).
gp (z) = ln ρ.
Gp (z) = − ln |f(z)|.
To see this, consider the temperature distribution Gep (w) in D obtained by trans-
planting Gp with f. We know that this conformal transplantation preserves the
harmonicity of the original, that it sends the source at p to an equal source at
f(p) = 0, and that the vanishing temperature on B is transplanted to vanishing tem-
perature on C. Thus this temperature distribution G ep (w) = Gp [f−1 (w)] in D must
be the Green’s function with pole at 0, and we already know that this is − ln |w|.
Done.
Exactly the same reasoning yields the following generalization. Let J(z) be a
one-to-one conformal mapping of R to some other simply connected region S with
The Physics of Riemann’s Mapping Theorem 629
f(z) − f(s)
Gs (z) = − ln .
f(s) f(z) − 1
As a bonus, notice that this general formula establishes the previously claimed
“symmetry property” of the Green’s function:
Gs (z) = Gz (s).
For a more common approach to the symmetry property, see Ex. 15.
We end this section with a result which we will need later. The analogue for heat
flows of the velocity of a fluid flow is the heat flow vector H; in the present case,
H = −∇G. Let us call its magnitude Q ≡ |H| the local heat flux; this is the analogue
of fluid speed, and it represents the heat flux (per unit length) across a short line-
segment at right angles to the flow. Since B is an isotherm, H is orthogonal to B,
and so the local heat flux at a boundary point z may be expressed as
BG
Q(z) = − , (12.25)
Bn
where n measures distance in the direction of N, the outward unit normal vector
to B (see [12.31]). We can now state the result:
from 0 and hitting C at w = f(z). [Remember, f was originally defined to have this
property!] Let e
ϵ be the width of this image tube at w. Since the segment of B at z of
length ϵ is amplitwisted by f ′ (z) to the segment of C at w of length e
ϵ,
e
ϵ
|f ′ (z)| = .
ϵ
Next, recall that the width of a k-flux tube at any given point is equal to k divided
by the local heat flux [previously fluid speed] at that point. Since the local heat flux
is constant on C, its value at w is simply the ratio of the strength of the source at 0
to the perimeter of C, and by construction this ratio is (2π/2π) = 1. Since the local
heat flux at z is Q(z), we see that
e
ϵ = k/1 and ϵ = k/Q.
Thus
e
ϵ k
|f ′ (z)| = = = Q(z),
ϵ k/Q
as was to be shown.
12.7.1 Introduction
Consider a steady heat flow within a metal plate whose faces are insulated. Other
than at singularities, the temperature T (z) is then a harmonic function, and the
(locally) sourceless heat flow vector field is H = −∇T .
Let us measure the temperature around the circumference C of a circle of radius
R, the interior of which is free of sources and sinks, and the centre of which we may
conveniently choose to be the origin. As z = R eiθ moves round C, we may express
the measured temperature as a function of the angle: T = T (θ). We hope that it
may seem physically plausible that these values actually determine the tempera-
ture at any interior point a. Indeed, if a = 0 then we know [from Gauss’s Mean
Value Theorem] that the temperature at the centre of C is simply the average of the
temperatures on C:
Z
1 π
T (0) = ⟨T ⟩ = T (θ) dθ. (12.28)
2π −π
Eventually we will discover that the generalization of this result to the case a ̸= 0
is given by
Z
1 π R2 − |a|2
T (a) = T (θ) dθ. (12.29)
2π −π |z − a|2
Dirichlet’s Problem 631
Writing a = r eiα (r < R), and appealing to the cosine formula [exercise], this is
usually written as
Z
1 π R2 − r2
T (a) = T (θ) dθ.
2π −π R2 + r2 − 2Rr cos(θ − α)
This is called Poisson’s formula, and the quantity in square brackets is called the
Poisson kernel, which we shall write as Pa (z).
Formula (12.29) says that T (a) is a weighted average of T on C, the temperature
of each element of C contributing to T (a) in proportion to its weight Pa (z). Notice
that Pa (z) dies away inversely with the square of the distance between a and the
element of C, so that if one element is twice as far from a as another, its influence
on the temperature at a is only one quarter as great. If a = 0 then all parts of C have
equal influence (for all are equally far from a) and you can see that we do recover
(12.28).
Poisson’s formula is connected with the following important and difficult
issue. Instead of dealing with a pre-existing harmonic function, Dirichlet’s prob-
lem demands that we arbitrarily (but piecewise continuously) assign values to the
boundary of a simply connected region R and then seek a continuous harmonic
function in R which takes on these values as the boundary is approached.
In the case of the disc, H. A. Schwarz demonstrated that not only does the
solution to Dirichlet’s problem exist, but it is explicitly given by (12.29). If we
are handed the piecewise continuous values T (θ) on C then we may construct
a function T (a) in the interior according to Poisson’s recipe. Schwarz’s solution
then amounted to showing that T (a) is automatically harmonic, and that as a
approaches a boundary point at which T (θ) is continuous, T (a) approaches the
given value T (θ). Let us begin to explain all this.7
7
Much of the following material previously appeared in Needham (1994).
8
Schwarz (1972b).
632 Flows and Harmonic Functions
Being close to the cold side, we would expect a to be cool. Figure [12.32b] shows the
new temperature distribution obtained by projection through a. It is now vividly
clear how the distant hot semicircle is ‘focused’ through a onto a much smaller arc,
yielding a low average temperature on C and hence a low temperature at a itself.
To begin to establish (12.30), recall the conformal invariance of harmonic func-
tions: if T (z) is any harmonic function and h(z) any conformal mapping, then T (z⋆ )
is automatically harmonic, with z⋆ = h(z).
Suppose now that h(z) maps the disc to itself . If z = R eiθ lies on C then so does
⋆
z⋆ = R eiθ , and since we suppose that we have measured the temperature all round
C, we therefore know the temperature T (θ⋆ ) at z⋆ . Having the values of T (θ⋆ ), we
may now compute the integral in (12.28) for the harmonic function T [h(z)] to obtain
Zπ
⋆ 1
T (0 ) = T (θ⋆ ) dθ , (12.31)
2π −π
in which it should be stressed that the averaging is still taking place with respect
to the angle of z, not its image z⋆ .
We may interpret (12.31) as follows: the temperature at 0⋆ is the average of
the new temperature distribution on C obtained by transplanting the temperature
measured at each z to the new location z⋆ . We are now half way to Schwarz’s result.
To find the temperature at a we must find a conformal mapping of the disc to itself
such that 0 is sent to a, then take the average of the new temperature distribution.
But viewing the disc as the Beltrami–Poincaré model of the hyperbolic plane,
we are already very familiar with such mappings! Peek at [12.33b], to which we
Dirichlet’s Problem 633
shall return in a moment. If m is the midpoint (in the hyperbolic sense) of the line-
segment 0a, then the half-turn9 z ÞÑ z⋆ = Ma (z) of the hyperbolic plane about m
interchanges 0 and a: 0⋆ = a (as we desire) and a⋆ = 0. Thus to establish (12.30) we
need only demonstrate the illustrated fact that if z lies on C then z⋆ lies at the end of
the (Euclidean) chord B passing from z through a. In Chapter 3 we derived a formula
for Ma (z), so we could easily obtain this result by calculation; however, we prefer
a direct geometric approach.
First we need a simple result which is explained in [12.33a]. Consider the family
N ≡ {circular arcs passing from z to z⋆ },
where for the moment z⋆ may be thought of as any given point of C. The figure
shows three members of N: the arc A through 0, the Euclidean chord B, and the
hyperbolic line L. [Recall that in terms of hyperbolic geometry the members of N
consist of the equidistant curves of L.] The result we need is this:
The Euclidean chord B is the unique member of N such that L bisects the
(12.32)
angle contained by A and B.
Since each member of N is uniquely determined by the direction in which it
emerges from z, and since the radius z0 is tangent to L at z, this is equivalent to
9
See (3.53), p. 201.
634 Flows and Harmonic Functions
the following: if tz and t0 are tangent to A at z and 0, respectively, then the black
angle tz0 and the shaded angle z⋆ z0 are equal. The proof is immediate from the
figure and is left to the reader.
Now turn your attention to [12.33b], in which z⋆ is the image of z after a half-
turn about m. To finally establish Schwarz’s result we must prove the illustrated
fact that a lies on the chord B. To do so, consider the image A⋆ of A. Since the half-
turn interchanges the pair z and z⋆ and the pair 0 and a, A⋆ must be a member of N
passing through a. But since L⋆ = L, the conformality of the mapping also says that
the angle between A⋆ and L must equal the angle between A and L. We conclude
from (12.32) that this arc A⋆ through z, a, and z⋆ is none other than B. Done10 .
10
This argument is perhaps conceptually clearer than the more elementary one in Needham (1994).
Dirichlet’s Problem 635
[12.34] Solving Dirichlet’s Problem with Schwarz’s geometric construction. [a] Sup-
pose that the boundary temperature jumps from T1 to T2 as we pass from C1 to C2 .
If a arrives at z while travelling in a direction making an angle βπ with C, then T (a)
approaches [β T1 + (1 − β) T2 ]. [b] To derive Poisson’s formula from (12.33), con-
sider the movement R ∆θ⋆ ≍ t of z⋆ resulting from a movement R ∆θ ≍ s of z.
So (∆θ⋆ /∆θ) ≍ (t/s). But t and s are corresponding sides of two similar h i triangles
[shaded], so (t/s) = (σ ′ /ρ). And since (σ ′ /ρ) ≍ (σ/ρ), we obtain dθ
⋆
dθ
= σ ρ . Finally,
h i h i
ρ σ = ρ ′ σ ′ = (R + r)(R − r) = (R2 − r2 ), so dθ
⋆ 2 2
dθ
= σρ =
R −r
ρ2
= Pa (z). Done.
that (∆θ⋆ /∆θ) is ultimately equal to (t/s). But t and s are corresponding sides of
two similar triangles [shaded], so (t/s) = (σ ′ /ρ). Finally, since (σ ′ /ρ) is ultimately
equal to (σ/ρ), we obtain
dθ⋆ σ
= .
dθ ρ
Thus (12.33) becomes
Zπ
1 σ
T (a) = T (θ) dθ . (12.34)
2π −π ρ
Consequently, to derive Poisson’s formula we need only show that [σ/ρ] is the
Poisson kernel Pa (z). This was precisely how Schwarz, working in the opposite
direction, originally deduced his result from Poisson’s formula.
Since ρ σ = ρ ′ σ ′ is constant, we may evaluate it for the dotted diameter through
a to obtain ρ σ = (R2 − r2 ). Thus we do indeed find that
2
σ R − r2
= = Pa (z).
ρ ρ2
636 Flows and Harmonic Functions
[12.35] Geometric proof that Pa (z) is harmonic, and the derivation h i of Schwarz’s
formula. Since the angle at E is a right angle, we have Pa (z) = σ = |z+a| cos γ
|z−a| =
ρ
Re z−a . Since Pa (z) is the real part of an analytic function, it is indeed harmonic. Let
z+a
for this is analytic and has T (a) as its real part. This result is called Schwarz’s formula,
and it enables us to resurrect the complete analytic function f everywhere inside C
just from the ashes of its real part on C.
11
Incidentally, this means that the isotherms are the arcs of circles through p and q.
638 Flows and Harmonic Functions
Figure [12.37] is intended to make this result vivid.Turning one’s head succes-
sively through the same small angle marked • one would see the thermometers
located at the white dots on the boundary. The average of their temperatures is
then a good approximation [exact as • → 0] to ⟨⟨T ⟩⟩a , and hence to the average of
the temperature where we stand and the temperature at 0. Note how the white dots
become crowded together on the part of the boundary nearest us. As anticipated,
this part of the boundary therefore has the greatest influence on the temperature
where we stand.
To obtain our third and final interpretation of Poisson’s formula, imagine that the
disc is the Beltrami–Poincaré model of the hyperbolic plane, and that you are once
again standing at the point a looking out to K, which is now infinitely far away
on the horizon. How big does K appear to you in this distorted geometry? To a
Godlike observer looking down on this model of the hyperbolic plane, the straight
Dirichlet’s Problem 639
lines along which light travels to you now appear to be arcs of circles orthogonal
to C, and so you see the angular size of K as being
hyperbolic angle = λ + (• + ⊙) .
(• + ⊙) = 1
2 (K⋆ − K),
The temperature where you are is simply proportional to how big K looks! [The
result can also be obtained directly by appealing to the conformal and circle-
preserving nature of the hyperbolic half-turn Ma (z) considered earlier.]
Reinterpreting (12.33), we now see that dθ⋆ is simply the hyperbolic angle sub-
tended at a by the element of C: the temperature of each element of C contributes to
640 Flows and Harmonic Functions
the temperature at an interior point in proportion to its hyperbolic size as seen from that
point. Much as we did in the Euclidean case, let ≺ T ≻a denote the average of the
temperatures you see on the horizon of the hyperbolic plane as you turn your head
through a full revolution while standing at a. We have thus found what we shall
call Bôcher’s formula:
T (a) = ≺ T ≻a . (12.36)
This result (exceeding even the beauty of Schwarz’s) is due to Bôcher (1898), Bôcher
(1906). We have chosen to present (12.36) as a consequence of Schwarz’s result, but
at the end of the section we shall see that it can be understood in a much simpler
way.
The analogue of [12.37] is now [12.38]. Standing at the same point as before, and
again turning one’s head successively through the angle •, the figure shows the
new locations of the thermometers we see on the boundary. The average of their
temperatures is then a good approximation [exact as • → 0] to ≺ T ≻a , and hence to
the temperature where we stand. Note how the white dots again become crowded
together on the part of the boundary nearest us, so that this part of the boundary
has the greatest influence on the temperature where we stand.
From the vantage point of (12.36), the distinction between (12.28) and (12.33)
evaporates. Every point of the hyperbolic plane is on an equal footing with every
other, it is merely that the hyperbolic angle dθ⋆ happens to coincide with the more
familiar Euclidean angle dθ when a = 0.
Formulated in this way, we may carry the result over to the upper half-plane
model for hyperbolic geometry. [The full justification for this transition will be
explained at the end of the section.] The horizon is now the real axis and ‘straight
lines’ are now (for our Godlike observer) semicircles meeting the real axis at right
angles. The temperature where you stand is now the average (as • → 0) of the
temperatures at the white boundary points in [12.39].
Figure [12.40] analyses this in greater detail. It shows both the hyperbolic angle
∆θ⋆ and the Euclidean angle ∆θ subtended at a by the element ∆x of the horizon.
Thinking of ∆x as sufficiently small that T (x) is essentially constant on it, the con-
tribution to the temperature at a is (1/2π) T (x) ∆θ⋆ . Integrating along the entire
horizon we obtain
Z
1 x=∞
T (a) = T (x) dθ⋆ . (12.37)
2π x=−∞
In order to put this into precisely the same form as (12.35), we need to find
(dθ⋆ /dx). We shall do this via an attractive and rather surprising fact: The non-
Euclidean angle ∆θ⋆ is exactly double the Euclidean one ∆θ, even if ∆x is not small. To
see this, concentrate on the semicircle meeting the axis at p. The angle between the
dotted tangent at a and the vertical is clearly double that between the chord ap and
the vertical. The result then follows immediately.
Dirichlet’s Problem 641
Ω ∆θ ξ Y
≍ ≍ .
∆x ∆x Ω
We can now combine this with the result of the previous figure to obtain
dθ⋆ dθ Y Y
=2 =2 = 2 .
dx dx Ω2 (X − x)2 + Y 2
Inserting this into (12.37), we obtain (12.35).
While the precise form of the above argument may be new, the basic idea of
transferring Bôcher’s result from the disc to the half plane was given by Osgood
(1928). For a different approach to (12.35), see Lange and Walsh (1985). For more
on all three of the interpretations thus far obtained, see Perkins (1928).
where ds is an element of arc length along B. Thus Qa now plays the same role as
the Poisson kernel did in (12.29). Indeed, we previously calculated Qa for the unit
disc, and we now recognize the result (12.27) as the Poisson kernel Pa .
Although formula (12.38) is valuable both in theory and practice, we should
point out that it is less explicit than Poisson’s formula, for to find Qa we must first
find the Green’s function. But as we previously explained, the problem of finding
Ga is itself a Dirichlet problem: to construct
Ga (z) = − ln ρ + ga (z),
we must find the harmonic function ga with boundary values ga (z) = ln ρ. For-
mula (12.38) says that if we can just solve this particular boundary value problem
then we can solve them all.
To begin with, imagine that T (z) is a given harmonic function in R whose value
T (a) at an interior a we wish to determine from the boundary values. The idea
behind our explanation of (12.38) is very simple12 . The Green’s function Ga enables
12
For a beautiful physical explanation of (12.38) in terms of electrostatic energy, see Maxwell (1873)
or, better still, Maxwell (1881, Ch. III).
644 Flows and Harmonic Functions
result (12.26): the amplification |f ′ (z)| equals the local heat flux Qa (z). This concludes
the derivation of Green’s formula, (12.38).
This argument also explains the stronger result that (12.38) solves Dirichlet’s
problem for R. Using f to conformally transplant the given boundary values from B
to C, we know that Poisson’s formula allows us to construct the solution to Dirich-
let’s problem in D. Transferring this solution back from D to R with f−1 , we have
found the harmonic function T in R, and its value at a must then be given by (12.38).
You can now understand why we lavished so much attention on the special case
of the disc.
We end this section with the observation that Green’s formula (12.38) possesses
the beautiful geometric interpretation shown on the LHS of [12.43]. Just as in
[12.38], one imagines standing at a and turning one’s head successively through
the small angle •. But now suppose that light travels along the illustrated streamlines
of the heat flow H = −∇Ga associated with the Green’s function. We would then see
the thermometers at the illustrated points on the boundary. The general formula
(12.38) says that the average of these temperatures (as • → 0) is the temperature
where one stands! Bôcher’s interpretation is clearly just a special case13 .
The explanation essentially reiterates the derivation of (12.26). Let zθ be the
boundary point we see when we look in the direction θ. Green’s formula says that
the temperature T (zθ ) of the element ds contributes to the temperature at a in pro-
portion to Qa (zθ ) ds, which is the flux of H through ds. Now follow the shaded flux
tube back to the source at a, and let dθ be its angular width there. Since 2π of flux
emerges symmetrically from a, the flux Qa (zθ ) ds emitted into our tube is equal to
dθ. Thus (12.38) may be re-expressed as
I
1
T (a) = T (zθ ) dθ, (12.39)
2π B
namely, as the average of the boundary temperatures T (zθ ) over all directions.
Done.
We have presented [12.43] as a geometric interpretation of (12.38), but we may
instead use it to simplify and illuminate our derivation of that formula. The key
observation is that (even without passing to the limit of vanishing •) the average in
[12.43] of the observed temperatures on B is conformally invariant. As before, let J(z)
be a one-to-one conformal mapping of R to some other simply connected region
S with boundary Y. Just as we did with f, let us choose J so that the directions of
curves through a are preserved (i.e., arg [J ′ (a)] = 0). Let wθ ≡ J(zθ ) be the image
on Y of zθ on B.
By the conformal invariance of the Green’s function, the image of the streamline
leaving a at angle θ is the streamline leaving J(a) at the same angle. Thus wθ is not
13
If we define the distance between two infinitesimally separated points of R to be the hyperbolic
distance between their images in D, then R becomes a (non-standard) conformal model of the hyper-
bolic plane, and the geodesics emanating from a are the streamlines of [12.43]. The two results may
then be viewed as identical.
646 Flows and Harmonic Functions
[12.43] The geometric meaning of Green’s formula. Green’s formula (12.38) possesses
the beautiful geometric interpretation shown on the left. Just as in [12.38], one imagines
standing at a and turning one’s head successively through the small angle •. But now
suppose that light travels along the illustrated streamlines of the heat flow H = −∇Ga
associated with the Green’s function. We would then see the thermometers at the illus-
trated points on the boundary. The general formula (12.38) says that the average of these
temperatures (as • → 0) is the temperature where one stands! Bôcher’s interpretation is
clearly just a special case. Indeed, if we define the distance between two infinitesimally
separated points of R to be the hyperbolic distance between their images in D, then R
becomes a (non-standard) conformal model of the hyperbolic plane, and the geodesics
emanating from a are the streamlines. The two results may then be viewed as identical.
only the image of zθ , it is also the boundary point which an observer at J(a) sees
when he looks in the direction θ. But, by definition, the temperature at each point
zθ on B is transplanted to wθ on Y, so the observer at J(a) sees exactly the same
temperatures on Y as the original observer at a saw on B. Done.
Passing to the limit of vanishing •, the conformal invariance of this average may
expressed as
I I
1 1
T (zθ ) dθ = Te(wθ ) dθ.
2π B 2π Y
Figure [12.43] illustrates the particular case where J = f is the previously con-
structed function which maps R to D and a to 0. The virtue of this special case
is that the conformally invariant average may now be evaluated. By Gauss’s Mean
Value Theorem, the average of the transplanted temperatures Te(wθ ) ≡ T (zθ ) on C
is the temperature Te(0) ≡ T (a) at the centre:
I I
1 1
T (zθ ) dθ = Te(wθ ) dθ = T (a).
2π B 2π C
Dirichlet’s Problem 647
Thus, returning to R and passing to the limit of vanishing •, the average of the
observed temperatures is the temperature where you stand. Finally, the argument
associated with [12.43] shows that (12.39) is equivalent to (12.38).
As illustrated in [12.44], this idea of a conformally invariant hyperbolic average
lends unity to much of what we have done. Top centre is a depiction of Gauss’s
[12.44] Hyperbolic geometry unifies all methods of solving Dirichlet’s Problem. For
the Beltrami–Poincarites who inhabit the hyperbolic plane, the conformally invariant aver-
age that Gauss calls for at the centre of the disc [see top centre] is utterly indistinguishable
from the Poisson disc-formula [top left] at an arbitrary point. In both cases, the Beltrami–
Poincarite turns her head successively through the small angle •, and both figures show
the locations of the thermometers she sees on the boundary as light travels along hyper-
bolic straight lines (geodesics) to her eye. The average of their temperatures is then a
good approximation [exact as • → 0] to the temperature where she stands. But she
could also not detect any change if we transplanted her and the boundary temperatures
to the Beltrami–Poincaré upper half-plane: the temperature would still be the average
over all directions. Finally, if we conformally transplant both the boundary temperatures
and the hyperbolic metric to a general region R [top right], then it becomes a (non-
standard) conformal model of the hyperbolic plane. All four figures may then be viewed
as hyperbolically indistinguishable.
648 Flows and Harmonic Functions
12.8 Exercises
Bz = 1
2 ∇ and Bz = 1
2 ∇.
(ii) Deduce (at least formally) that an analytic function f depends on z but not on
z:
Bz f = f ′ and B z f = 0.
10 As usual, let b
ξ be the unit tangent to a curve of curvature κ, and let e
κ be the
curvature of the image under the analytic mapping f(z). Also let s and e
s denote
arc length of the original and image curves. Finally, let Ψ = (1/|f ′ |) be the
stream function of the complex curvature, K = −i∇Ψ.
(i) Use (5.31), p. 272 to show that B s̃ e b ].
κ = Ψ Bs [κΨ + K ξ ·
(ii) Show that Bs bξ = iκbξ and Bs Ψ = ξ ·
b (iK).
(iii) Deduce that
B s̃ e · b.
κ = Ψ2 Bs κ + Ψ (Bs K) ξ
· ·
[Hint: Remember (or prove) that (ia) (b) + (a) (ib) = 0.]
(iv) Recall Ex. 18, p. 296, in which we saw how Newton attempted to define
the “angle” Θ between two touching curves as the difference of their
curvatures: Θ ≡ (κ1 − κ2 ). Although this is not quite conformally
invariant, show that Θ2 /Bs Θ is conformally invariant. This geometri-
cally meaningful generalization of the concept of angle is called Kasner’s
Invariant.
11 With the same notation as in the previous exercise, let p measure distance in
the direction ib ξ perpendicular to the curve. By substituting K = −i∇Ψ into
(5.31), p. 272, show that the image curvature is given by the tidy formula,
e
κ = Bp Ψ + κ Ψ.
16 Let T (z) be the temperature distribution in the unit disc if the top semicircle
(Im(z) > 0) is kept at temperature +(π/2) and the bottom semicircle is kept at
temperature −(π/2). Show that
1+z
T (z) = Arg .
1−z
17 Let T (z) be a non-negative temperature distribution in the disc |z| ⩽ R. Writing
|a| = r, use Poisson’s formula (12.29) to derive Harnack’s inequality:
R−r R+r
T (0) ⩽ T (a) ⩽ T (0).
R+r R−r
18 Use Harnack’s inequality [previous exercise] to prove the following analogue
of Liouville’s Theorem: If T is harmonic in the whole plane and is bounded from
above (or below), then T is a constant.
19 Substitute the disc Green’s function (12.22) into (12.25), thereby confirming
the formula (12.27) for the heat flux at the boundary of the disc.
20 (i) Use the method of images to find the Green’s function for the upper half-
plane.
(ii) Use this to show that Green’s general formula (12.38) does indeed yield
the Poisson half-plane formula (12.35).
21 Use the idea behind the method of images to show that if 0 < Re (p) < 1 then
the Green’s function of the half-disc Re (z) ⩾ 0, |z| ⩽ 1 is
z−p z+p
Gp (z) = − ln + ln .
pz − 1 pz + 1
Check this by getting the computer to draw the level curves of Gp (z).
BIBLIOGRAPHY
Bôcher, M. 1898. Note on poisson’s integral. Bull. Amer. Math. Soc., 4(9):424–426.
Bôcher, M. 1906. On harmonic functions in two dimensions. Proceedings of the
American Academy of Arts and Sciences, 41(26):577–583.
Bohlin, K. 1911. Note sur le problème des deux corps et sur une integration nouvelle
dans le problème des trois corps. Bull. Astr., 28:113–119.
Braden, B. 1985. Picturing functions of a complex variable. The College Mathematics
Journal, 16(1):63–72.
Braden, B. 1987. Polya’s geometric picture of complex contour integrals. Mathemat-
ics Magazine, 60(5):321–327.
Braden, B. 1991. Visualization in teaching and learning mathematics: a project, volume
no. 19, chapter : The Vector Field Approach in Complex Analysis. Mathematical
Association of America.
Brieskorn, E. and H. Knörrer 2012. Plane Algebraic Curves, Modern Birkhäuser
classics. Basel: Birkhäuser.
Carathéodory, C. 1937. The most general transformations of plane regions which
transform circles into circles. Bulletin of the American Mathematical Society,
43(8):573–579.
Carathéodory, C. 1964. Theory of Functions of a Complex Variable, 2nd english ed
edition. New York: Chelsea Pub. Co.
Chandrasekhar, S. 1995. Newton’s Principia for the Common Reader. Oxford: The
Clarendon Press, Oxford University Press.
Courant, R. 1950. Dirichlet’s Principle, Conformal Mapping, and Minimal Surfaces,
volume 3. New York: Interscience Publishers.
Coxeter, H. S. M. 1946. Quaternions and reflections. The American Mathematical
Monthly, 53(3):136–146.
Coxeter, H. S. M. 1967. Proceedings of the International Conference on the Theory of
Groups, chapter : The Lorentz Group and the Group of Homographies, Pp. 73–77.
New York: Gordon and Breach Science Publishers.
Coxeter, H. S. M. 1969. Introduction to Geometry, 2nd edition. New York: Wiley.
Coxeter, H. S. M. and S. L. Greitzer 1967. Geometry Revisited, volume 19. Random
House.
Davis, P. and H. Pollak 1958. On the analytic continuation of mapping functions.
Transactions of the American Mathematical Society, 87(1):198–225.
Davis, P. J. 1974. The Schwarz Function and its Applications, volume no. 17. Mathe-
matical Association of America.
Bibliography 655
de Gandt, F. 1995. Force and Geometry in Newton’s Principia. Princeton, NJ: Princeton
University Press. Translated from the French original and with an introduction
by Curtis Wilson.
do Carmo, M. P. 1994. Differential Forms and Applications, Universitext. Berlin:
Springer-Verlag.
Dray, T. 2015. Differential Forms and the Geometry of General Relativity. CRC Press,
Taylor & Francis Group.
Duffin, R. J. 1957. A note on Poisson’s integral. Quarterly of Applied Math, 15:109–111.
Earl, R. 2019. Topology: A Very Short Introduction, Very short introductions. Oxford
University Press.
Echols, W. H. 1923. Some properties of a skewsquare. The American Mathematical
Monthly, 30(3):120–127.
Eves, H. 1992. Fundamentals of Modern Elementary Geometry. Boston: Jones and
Bartlett.
Feynman, R. P. 1963. The Feynman Lectures on Physics. Reading, Mass.: Addison-
Wesley Pub. Co.
Feynman, R. P. 1966. Nobel lecture: ”The Development of the Space-Time View of
Quantum Field Theory”. Physics Today, August:31–34.
Feynman, R. P. 1985. QED: the Strange Theory of Light and Matter. Princeton, N.J.:
Princeton University Press.
Feynman, R. P. 1997. Surely you’re joking, Mr. Feynman!: Adventures of a Curious
Character, Norton paberback edition. New York: W.W. Norton.
Finney, R. L. 1970. Dynamic proofs of Euclidean theorems. Mathematics Magazine,
43(4):177–185.
Ford, L. R. 1929. Automorphic Functions, 1st edition. New York: McGraw-Hill book
company, inc.
Frankel, T. 2012. The Geometry of Physics: an Introduction, 3rd ed edition. Cambridge:
Cambridge University Press.
Fulton, W. 1995. Algebraic Topology: A First Course, volume 153. New York: Springer-
Verlag.
Gauss, C. F. 1827. General Investigations of Curved Surfaces, Translated from the Latin
and German by Adam Hiltebeitel and James Moreh ead, 1965 edition. Raven
Press, Hewlett, N.Y.
Goodstein, J. R. 2018. Einstein’s Italian Mathematicians: Ricci, Levi-Civita, and the Birth
of General Relativity. AMS.
656 Bibliography