You are on page 1of 248

THE

PERFECT
SHAPE
SPIRAL
STORIES
The Perfect Shape
Øyvind Hammer

The Perfect Shape


Spiral Stories
Øyvind Hammer
Natural History Museum
University of Oslo
Oslo, Norway

ISBN 978-3-319-47372-7 ISBN 978-3-319-47373-4 (eBook)


DOI 10.1007/978-3-319-47373-4

Library of Congress Control Number: 2016959575

© Springer International Publishing AG 2016


This work is subject to copyright. All rights are reserved by the Publisher, whether the whole or part of the
material is concerned, specifically the rights of translation, reprinting, reuse of illustrations, recitation,
broadcasting, reproduction on microfilms or in any other physical way, and transmission or information
storage and retrieval, electronic adaptation, computer software, or by similar or dissimilar methodology now
known or hereafter developed.
The use of general descriptive names, registered names, trademarks, service marks, etc. in this publication does
not imply, even in the absence of a specific statement, that such names are exempt from the relevant protective
laws and regulations and therefore free for general use.
The publisher, the authors and the editors are safe to assume that the advice and information in this book are
believed to be true and accurate at the date of publication. Neither the publisher nor the authors or the editors
give a warranty, express or implied, with respect to the material contained herein or for any errors or omissions
that may have been made.

Printed on acid-free paper

This Copernicus imprint is published by Springer Nature


The registered company is Springer International Publishing AG
The registered company address is: Gewerbestrasse 11, 6330 Cham, Switzerland
Preface

Transient whorls in a turbulent mountain stream, always changing, never the


same; majestic swirls of galaxies; the curls of a child’s hair; exquisite seashells in
the sunlit tropical ocean; the perfect golden ornaments of bronze age Europe;
the labyrinth of the Minotaur, and his brutal but elegant horns; the hiero-
glyphic paths of subatomic particles in the vast detectors of high-energy
physics; volutes in a Greek temple dedicated to the glory of Poseidon; the
narwhal’s tusk (or rather the unicorn’s, as the ancients knew well); the violent
yet placid storms of Jupiter; the cochlea in the deep, dark recesses of your ears;
the mathematician’s curves in the complex plane; climbing plants in an
overgrown garden; the patterns of petals and seeds on the sunflower; winding
stairs in a Renaissance palace; the helix of DNA; the accretion disc around a
black hole; breaking ocean waves; a beach of golden sand behind a promon-
tory; a viper in its attack position; a coil of rope on the deck of a sailing ship;
the path traced by your fingertip as you unfold your finger.

Spirals.

No other shape evokes more strongly a sense of beauty, mystery and


eternity. Cyclic, but not repeating, endless, but not unbounded, the spiral
must surely be the Perfect Shape. Often one of the first figures drawn by a
child, and one of the most beloved elements of decorative art, the spiral seems
to draw you in through its coils, grabbing your attention, not letting you go. I
was seduced by its charms more than 20 years ago, studying spiral fossils, and
now and then I encounter other poor souls on the Internet where I recognize
my own ghastly symptoms, seeing spirals everywhere and desperately needing

v
vi Preface

to tell other people about it. Beware. This is a warning. If you are susceptible,
then maybe you should not read this book.
From the beginning of human culture, the spiral has been a spiritual
symbol, depicting the sun, perhaps, or a journey in winding circles where
the pilgrim closes in on enlightenment for every turn, like the path up Mount
Purgatory in Dante’s Divina Commedia. Several books have been written
about such “spirals of the soul”. There are also good books about the science of
spirals in physics or biology. This book is meant to be different. It is a
collection of essays organized somewhat like a spiral, circling around the
common theme while covering a wide spectrum of human knowledge. The
subject of spirals opens the door to a celebration of the richness of nature,
culture and the human intellect.
There will be a little math, not too difficult I hope, but you can safely skip
the equations if you are not particularly interested. I have included these
equations partly because they look pretty and partly to show that I am not
just telling you fairy tales. The literature on spirals, in books and on the
Internet, is a maze, full of mirages, myths and misconceptions, swirling stories
that are repeated endlessly without basis in reality. This mesmerizing web of
spiral legends is entertaining but also very frustrating. I have tried to check the
sources and do the math myself, but it would not surprise me if there are still
errors. Let me know if you find one!
Now starts the wild ride around the vortex. I hope you will enjoy it.

Oslo, Norway Øyvind Hammer


Acknowledgments

First of all thanks to my family: Marte, Cyrus and Eiel. Especially to Marte for
all our spiral discussions; what luck to live with a botanist! Thanks also to all
the brilliant photographers who have allowed me to use their work, most of
them without compensation. To my employer, the Natural History Museum
in Oslo, for being nice. To Springer Publishing for printing such a strange
book. And most of all, to the Laws of Nature, or God, or whatever it was, that
gave us the Perfect Shape.

vii
Contents

1 Spirals of the Abyss 1


2 The Spiral Zoo 3
3 A Bearded Man in a Bathtub 7
4 The Icon 11
5 The Golden Spiral Silliness 15
6 Spiral Energy 19
7 Curling Up 23
8 The King of Snails 29
9 Spira Mirabilis 33
10 Unfortunate Moths and Lopsided Falcons 39
11 Circular Tessellations 41
12 Ropes and Rifles 53
13 The Lost Sea of Spirals 57
14 The Great Spiral in the Sky 59
15 The Case of the Staircase 65
16 The Spiral of the Ancient Mariner 69

ix
x Contents

17 Gnomons, a Miracle, and Charles Babbage 73


18 Curls of Green 79
19 The Pendulum and the Galaxy 85
20 How to Grab a Can of Beer 89
21 An Interlude at the Beach 93
22 When Television Was Spiral 95
23 Thou Shalt Love Thy Neighbour 99
24 Spiral Jetty, Tatlin’s Tower 101
25 Now It Gets Complex 105
26 The Killer Spiral 109
27 The Friend 113
28 The Labyrinths of History 115
29 Newton’s Spiral Headache 119
30 Sculptures of the Sea 123
31 The Spiral of the Bird Priests 133
32 Squaring the Circle 135
33 The Daemon Beavers of Nebraska 139
34 Under the Mistletoe 141
35 Double Spirals, Twice the Fun 147
36 Maelstrom 151
37 Treasures in the Mud 159
38 Subatomic Squiggles 161
39 Nature Red in Blood and Claw 165
40 Coffee, Kepler and Crime 169
41 Dürer’s Dirty Secret 173
42 The Spiral from the Depth of Time 177
43 Propelling, the Archimedean Way 179
Contents xi

44 Unwrapping Mummies 183


45 Pagan Coils 189
46 A Note on Toilet Paper 201
47 A Delightful Nuclear Missile Disaster 203
48 Shaligram-Shilas and the Hands of Vishnu 207
49 The Quest for the Sublime Spiral 211
50 A Very Funny Fish 217
51 Spirals of the Mind 219
52 The Spider’s Spiral Spin 225
53 The Mystery of the Twisted Tree 229

Afterword 233
Appendix A: Mathematical Derivations 235
Appendix B: Program Code 243
Literature 249
Index 255
1
Spirals of the Abyss

In shales formed from deep-sea muds, all over the world, geologists keep
stumbling upon the intriguing trace fossil Spirorhaphe. A perfect spiral, a
foot or more in diameter, is imprinted upon the petrified sea floor like a
bronze-age ornament. Such spirals date back at least to the Ordovician period,
some 460 million years ago, and continue through the geological record almost
to the present day. But what are they? The organism responsible for these
fantastic feeding traces was believed to be extinct, and its identity forever lost
to science (Fig. 1.1).
Then, in 1962, when scientists lowered a camera into the Kermadec Trench
in the southwestern Pacific, beautiful, modern-day Spirorhaphe traces were
finally revealed (Bourne and Heezen 1965). One of the pictures even seemed
to capture the trace-maker in action. It looked like an acorn worm, a repre-
sentative of an enigmatic group that fits only uncomfortably into the System of
Animals but has been placed in the phylum Hemichordata. And it was huge:
with a diameter of 5 cm it was quite a monster compared with most of its
shallower-water brethren.
As more pictures were taken from the deep sea, these spirals turned out to be
relatively common. The famous photographic volume “The Face of the Deep”
(Heezen and Hollister 1971) contains several examples. However, it was not
until 2005 that a good video recording of the actual trace making was
announced, together with the spectacular capture of the organism (Fig. 1.2).
The story was sensational enough to make it to the pages of Nature, but
without reference to the fossil record (Holland et al. 2005).

© Springer International Publishing AG 2016 1


Ø. Hammer, The Perfect Shape, DOI 10.1007/978-3-319-47373-4_1
2 The Perfect Shape

Fig. 1.1 The trace fossil Spirorhaphe, ca. 20 million years old, Spain. Photo
Falconaumanni, Creative Commons Attribution-Share Alike 3.0 Unported license

Fig. 1.2 Acorn worm with its trail, North Atlantic. Scale bar 5 cm. Adapted by
permission from Macmillan Publishers Ltd: Nature, Holland et al., copyright (2005)

Almost since the conception of animal life, this slimy worm has been sitting
in the eternally dark and cold depths of the sea, silently spinning its spirals at a
rate of 5 mm per minute. Hundreds of thousands of millennia passed. Life
ventured onto land. Dinosaurs came and went; mammals and birds conquered
the dry world. For the deep-sea acorn worm, nothing of this mattered much. It
sat down there where the sun never shines, surviving, hardly moving.
How appropriate that it builds a perfect spiral, the symbol of eternity.
2
The Spiral Zoo

A spiral is usually defined as a curve in the plane that winds around a central
point, moving away from the point as it revolves. It is a somewhat imprecise
definition, perhaps, but it will do for our purposes. It is usually a good idea to
express spirals in terms of polar coordinates, where the radius r is a function of
rotation angle φ (phi) (Fig. 2.1):
r ¼ f ðφÞ:

According to the definition, as the curve winds around the origin, the angle φ
increasing, the radius r should also increase. I guess it could sometimes decrease
a little bit without the spiral crashing into itself, but let us be a little strict and
require that r increases all the time. It could also decrease all the time, so that
the spiral moves inwards instead of outwards. In other words, f (φ) is a
monotonic function. Now there are many monotonic functions, and each of
them will produce a spiral in polar coordinates. Mathematicians are fond of
putting names on curves, and spirals are no exception. Given any simple
monotonic function, chances are very high that the corresponding spiral has
a fancy name. Just a few of them are given in Figs. 2.2 and 2.3. They are
certainly all pretty, and most of them are really interesting as well. We will start
with the simplest of them all.

© Springer International Publishing AG 2016 3


Ø. Hammer, The Perfect Shape, DOI 10.1007/978-3-319-47373-4_2
4 The Perfect Shape

r
φ

Fig. 2.1 In polar coordinates, the position of the red dot is given as (φ, r)

A 20 B 6

15
4

Radius
Radius

10 3

2
5
1

0 0
-5 0 5 10 15 20 25 0 5 10 15 20 25
Angle (radians) Angle (radians)

C 1.0
D 30

0.8
20
0.6
Radius

Radius

0.4
10
0.2

0.0
0
0 5 10 15 20 25 30 0 5 10 15 20 25 30
Angle (radians) Angle (radians)

E
0.8
Radius

0.6

0.4

0.2

0 5 10 15 20 25 30
Angle (radians)

Fig. 2.2 Monotonic functions in Cartesian coordinates. (a) Exponential function,


r ¼ ekφ. (b) Square root function, r ¼ √φ. (c) Hyperbolic function, r ¼ 1/φ. (d) Linear
function, r ¼ kφ. (e) Inverse square root function, r ¼ 1/√φ
2 The Spiral Zoo 5

Fig. 2.3 The same monotonic functions as in Fig. 2.2, but plotted in polar
coordinates. (a) Logarithmic spiral, r ¼ ekφ. (b) Fermat spiral, r ¼ √φ. (c) Hyperbolic
spiral, r ¼ 1/φ. (d) Archimedes spiral, r ¼ kφ. (e) Lituus, r ¼ 1/√φ
3
A Bearded Man in a Bathtub

The true depth of the Greek contributions to human progress is under-


communicated. Classical Greek philosophers and artists are highly praised
more or less by convention. Still, there is an underlying sentiment that it
was all very good for its time, but nothing compared with modern stuff, of
course. Aristotle, did he not say a lot of silly things, retarding science until the
Renaissance? Pythagoras, was he not that hippie with quaint ideas about some
polyhedron or other? And Archimedes? His main claim to fame is that he
discovered that he floated in the bathtub, so he jumped up and shouted
Eureka. All well and good, but surely not quite Einstein.
Most of us do not learn Greek anymore, but luckily, there are translations of
the main classical works. Euclid's Elements is available in paperback. That
volume (a collection of 13 books) is a brutal revelation. The intellectual
determination, the level of intelligence, the modern way of thinking, and
not least the massive scale, are all astonishing. We may be familiar with
individual, clever theorems and proofs, such as the “Pythagorean” theorem,
but there are hundreds of those in Euclid, building each other up into an
enormous, elaborate but solid structure. It is hard to believe that this is a work
from Antiquity, and not some modern fake by a brilliant conman. The
Elements is not “good considering its age”. It is an enormous mathematical
achievement judged by modern standards.
Back to Archimedes. He was a real person, walking the Earth from about
287–212 BC, writing real books that we can still read. Among his many
triumphs, he more or less invented calculus, although within the customary
Greek framework of geometry, which appears cumbersome to us. And when it

© Springer International Publishing AG 2016 7


Ø. Hammer, The Perfect Shape, DOI 10.1007/978-3-319-47373-4_3
8 The Perfect Shape

comes to floating bodies, the bathtub story does not quite reflect his achieve-
ments, to say the least. His results in hydrostatics include calculations of the
centers of gravity and buoyancy of complicated geometric solids, something
that students find hard even with modern methods of calculus.
And of course, he was interested in spirals. Maybe the widespread use of
spirals in Greek art and architecture was a source of inspiration. Besides, his
friend, the mathematician Conon, had already worked on the subject. Archi-
medes’ book “On Spirals” was preserved in a copy made in the ninth or tenth
century AD. This manuscript disappeared at some point in the sixteenth
century, but luckily, backups had been made in the meantime. One of these
medieval copies was identified in the famous “Archimedes palimpsest”, made
in ca. 950 AD but not discovered until 1906 in a library in Constantinople. As
could be expected from this man, “On Spirals” is an impressive work—a
beautiful development of seemingly unconnected theorems that are combined
into grand conclusions. After proving 11 fundamental theorems, his definition
of a spiral follows:

If a straight line drawn in a plane revolve at a uniform rate about one extremity
which remains fixed and return to the position from which it started, and if, at
the same time as the line revolves, a point move at a uniform rate along the
straight line beginning from the extremity which remains fixed, the point will
λιξ) in the plane.
describe a spiral (ε
Transl. T.L. Heath (1897)

(Note that Archimedes uses the word ελιξ, “helix“, not spiral).
Imagine walking slowly outwards along the revolving hand of a clock (the
straight line). Your position will then describe the shape we now call an
Archimedes spiral, where the distance between succeeding whorls, as measured
in the radial direction, is constant (Fig. 3.1). In the modern language of

Fig. 3.1 Archimedes spiral


3 A Bearded Man in a Bathtub 9

Fig. 3.2 The principle of the Archimedes cam. A spring-loaded bar (red) moves
linearly as a function of the rotation angle of the spiral cam

analytic geometry, as invented by Descartes and Fermat, we can express this in


polar coordinates as follows:
r ¼ kφ;

where r is the radius and φ (phi) the rotation angle, and with k an arbitrary
constant of radial velocity.
Here, we only plot the spiral for positive angles. If we use both the positive
and negative angles, we produce a nice but confusing two-branched spiral. I
have made a choice in this book to only refer to the positive angles in the
Archimedes spiral, for simplicity.
In mechanical engineering the Archimedes spiral can be used as a cam, a
device that translates circular motion into linear translation. Simply let a
spring-loaded bar rest against the spiral, pointing towards the center. The
position of the bar will then be proportional to the rotation of the cam
(Fig. 3.2).
4
The Icon

As a first approximation, an animal is an invertebrate. Practically all animals,


both individuals and species, are invertebrates, making the backboned crea-
tures an interesting but vanishingly small minority. So disregarding for the
moment the big cats, the birds of prey and other such vertebrate curiosities,
there is one animal above all that symbolizes beauty, elegance, sophistication
and mystery. Its perfect spiral shape is abused everywhere in advertisements
and business logos—you see it on thousands of products from vitamin
pillboxes and contraceptive packages to exercising machines and computers.
Yet, the animal itself is fairly rare, almost never seen alive, and not many people
know what kind of beast it really is. It is indeed a strange and beautiful thing. It
is the Nautilus.
Nautilus is a cephalopod, the class of uncannily intelligent mollusks that
includes the squid, the octopus and the cuttlefish. The Nautilus is none of
those, but represents the single living representative of a subclass known as the
Nautiloidea. To be precise, there are at least three living species in two genera:
The common chambered nautilus Nautilus pompilius, the smaller bellybutton
nautilus Nautilus macromphalus and the rare crusty nautilus, Allonautilus
scrobiculatus. These odd animals have an external shell filled with nitrogen,
helping the creature to maintain buoyancy (Fig. 4.1). Somewhat paradoxically,
if it did not have that heavy gas container, built to withstand considerable
pressures, it would not need the extra uplift to begin with! The shell is
subdivided into a sequence of chambers, of which the outermost, large open
“body chamber” is not filled with gas but contains the bulk of the animal’s soft

© Springer International Publishing AG 2016 11


Ø. Hammer, The Perfect Shape, DOI 10.1007/978-3-319-47373-4_4
12 The Perfect Shape

Fig. 4.1 Nautilus shell and X-ray image. 14 cm across

Fig. 4.2 Ammonites (reconstruction)

parts. This must be the functional raison d’être of the shell: to provide a
protective armor against predators.
Nautiloids are extremely abundant in the fossil record, especially in the
Paleozoic era, ending some 250 million years ago. The extinct ammonoids
(including the ammonites) of the subsequent Mesozoic were their distant
relatives, with similar external, gas-filled shells (Fig. 4.2). Nautiloid and
ammonoid shells are perhaps the most exquisite and evocative fossils known,
not only because of their intrinsic geometric beauty but also because their
exotic shape seems so fitting considering the depth of time from which they
come. This latter aspect is also abundantly present in for example the trilobites,
but they are plain ugly in comparison. I like Pliny the Elder’s (23–79 AD)
note that ammonite fossils make us dream visions of the future:
4 The Icon 13

Hammonis cornu inter sacratissimas Aethiopiae [gemmas], aureo colore arietini


cornus effigiem reddens, promittitur praedivina somnia repraesentare.
The ‘Horn of Ammon’, which is one of the most sacred stones of Ethiopia,
has a golden yellow colour and is shaped like a ram’s horn. The stone is
guaranteed to ensure without fail dreams that will come true.

The shell of the Nautilus or an ammonite traces out a spiral, but clearly not
one of the Archimedes type. In the Archimedes spiral, the radius increases by a
constant increment for each whorl, in so-called arithmetic progression. In
contrast, the radius of the Nautilus shell is very close to increasing by a constant
percentage for each whorl—what is known as geometric progression. In polar
coordinates, the radius is an exponential function of rotation angle φ:

r ¼ aekφ :

The parameter a is simply a scaling factor, of little interest. The number e is


2.7128. . ., the Euler constant. Selecting this constant is a matter of mathe-
matical convention (albeit a highly convenient one)—any fixed positive num-
ber would do but would imply a rescaling of the parameters. The expansion
coefficient k controls the whorl expansion rate. The angle φ can be extended to
large negative values, taking the spiral inwards towards the origin.
An appropriate name for this shape would be an exponential spiral. How-
ever, when radius is an exponential function of rotation, then conversely
rotation is a logarithmic function of radius. From this somewhat inverse way
of reasoning, the spiral is called logarithmic.
The logarithmic spiral has a multitude of surprising and beautiful mathe-
matical properties, is ubiquitous in nature, and has occupied the minds of
many of the greatest human minds through history. It is the backbone of this
book—the Perfect Shape.
5
The Golden Spiral Silliness

There is an irritating thing we have to dispense with forthwith: The Golden


Spiral. The golden section, or golden ratio, is a funny number. We may define
it as the ratio Φ such that if a/b is Φ, then (a + b)/a is also Φ. To put it in a
quirky way that Euclid would approve of: The whole is to the larger part as the
larger is to the smaller:

a b

(a+b) / a = a/b = Φ

It is not difficult to calculate Φ. From a/b ¼ Φ, we get a ¼ bΦ.


Plugging into (a + b)/a ¼ Φ gives (bΦ + b)/bΦ ¼ Φ. The b cancels out nicely:
1 + 1/Φ ¼ Φ. Now multiply with Φ on both sides and rearrange, and you get
Φ2  Φ  1 ¼ 0. The positive solution to this quadratic equation is
pffiffiffi
5þ1
Φ¼ ¼ 1:6180339887498948482045868343656 . . .
2

This number has all kinds of strange properties—we will later see that it is
closely connected to the Fibonacci sequence, for example. It was well known in
Antiquity, and is discussed by Euclid, but contrary to popular belief, it was
probably not important in art and architecture at that time. However, starting
from the Renaissance, the golden ratio has been a holy number for architects.
Rectangles with the golden ratio proportion are common in old and new

© Springer International Publishing AG 2016 15


Ø. Hammer, The Perfect Shape, DOI 10.1007/978-3-319-47373-4_5
16 The Perfect Shape

Fig. 5.1 A trendy rectangle on a new business building down the street where I live
in Oslo. Guess what the height/width ratio is

buildings, in the shape of the façade as a whole and its parts, in the openings
for doors and windows (Fig. 5.1). The golden rectangle is believed by some to
be harmonious and beautiful, but I also suspect it is a kind of internal hallmark
among architects, a way of signaling knowledge of classical design principles.
I should mention that it has become a small cottage industry to debunk the
“myth” of the use of the golden section in architecture. It is certainly true that
there are countless outrageous claims about the golden section in Egyptian
pyramids or Greek temples, based on cherry-picking particular distances in
5 The Golden Spiral Silliness 17

0.618

Fig. 5.2 Construction of the Golden Spiral by successive inscription of golden


rectangles

particular buildings and disregarding the overwhelming number of counter-


examples. Nevertheless, starting with Luca Pacioli’s massive work De divina
proportione, printed in 1509 with illustrations by none other than Leonardo da
Vinci, the golden section has been such an important subject in the education
of architects that it would be very surprising if it were never used in practice.
The hero of modern architecture, Le Corbusier, was particularly explicit about
his use of the golden section.
Now if we inscribe a series of successively smaller golden rectangles inside
each other, we get a swirly figure where squares and golden sections and
Fibonacci numbers crop up everywhere (Fig. 5.2). It is possible to inscribe a
logarithmic spiral into this figure, as shown in the figure. A segment of this
logarithmic spiral inside one square can be approximated by one quarter of a
circle. This particular logarithmic spiral is called the Golden Spiral, or the
Fibonacci spiral. After one quarter revolution of the spiral, its size has increased
by a factor Φ. From this, we can calculate the expansion coefficient: k ¼ 2 ln
Φ/π ¼ 0.3063. . .
It is a nice spiral with some cool properties, but far too much has been made
of the Golden Spiral in the literature. It is only a special case of a logarithmic
spiral, and not even common in nature. Contrary to what you may read
elsewhere, the Nautilus is not a golden spiral, in fact it is not even close,
with k ¼ 0.177 or so. Some abalone snails (Haliotis) are closer, without this
signifying anything in particular (Fig. 5.3).
18 The Perfect Shape

Fig. 5.3 Haliotis clathrata, Philippines, 3 cm long. The logarithmic spiral expansion
coefficient k ¼ 0.25, slightly less than the Golden Spiral k ¼ 0.31

You might like to estimate the expansion coefficient of your favorite


spiral—maybe it is closer to the Golden Spiral than my Haliotis? There are a
couple of ways to do it. You can measure two radii 180 apart, and get k ¼ ln
(r2/r1)/π. A possibly more accurate method is to take a picture and find the x–y
coordinates of a number of points along the spiral with an image editing
program. These points can be fitted to a logarithmic spiral by computer, using
for example my free program Past (see Appendix B). It is not necessary to know
where the pole of the spiral is situated; the program will estimate it for you.
6
Spiral Energy

The giant turbines of hydroelectric power plants are among the largest and
most poetic machines made by humans, turning colossal water pressures and
velocities into some 16 % of the World’s electric energy. The most common
design is the Francis turbine, invented by James B. Francis in 1848. In the
Francis turbine, water enters tangentially as in an old water wheel, drives the
blades of the runner, and exits axially in the draft tube (Fig. 6.1). The largest
Francis turbines deliver up to 800 MW, about the power of one of the World’s
largest nuclear reactors.
In these turbines, water is delivered to the runner over a full 360 circle. As
water escapes through the runner and into the draft tube, pressure and velocity
are lost. In order to maintain the water velocity delivered to the runner, the
diameter of the inlet tube is progressively reduced, resulting in a so-called spiral
casing. This idea is not due to Francis himself, but probably originated in the
1880s—the first definite spiral casing I have found is in a turbine constructed
by Adolf Pfarr for the Voith company in Germany in 1886.

© Springer International Publishing AG 2016 19


Ø. Hammer, The Perfect Shape, DOI 10.1007/978-3-319-47373-4_6
20 The Perfect Shape

Fig. 6.1 Water flow in the spiral casing of a Francis turbine. The water passes
through the blades of the “runner” (light gray) and escapes through a “draft tube”
from the center of the turbine (dark gray)

In the classical design of a spiral Francis turbine casing, the outer radius of
the casing is a partial Archimedes spiral, not a logarithmic one. Still, the
reduction in diameter of the tube inwards in the spiral strongly evokes the
image of a giant snail or ammonite—a prehistoric monster brimming with
power. The spiral casing of the Francis turbine is a symbol of Man’s trium-
phant but destructive control of Nature (Fig. 6.2).
6 Spiral Energy 21

Fig. 6.2 Spiral casing for Francis turbine, 5.4 m diameter, J.M. Voith
Maschinenfabrik. Bundesarchiv, Bild 102-11144/Georg Pahl 1928
7
Curling Up

In much of the popular scientific literature on the subject, the logarithmic


spiral shape of the Nautilus remains a mysterious curiosity: is it not fantastic
that an animal can build itself according to such an “advanced” equation! But
science is not primarily about form—it is about process—admittedly the two
are nearly one and the same in many biological systems. Nobody knows how a
Nautilus, or almost any organism for that matter, constructs its body. Devel-
opmental biology, the science about how a leopard makes its spots (or, in other
words, why it has them—is not this the only proper way to explain biological
form?), is still in its infancy. But we can hypothesize, and as it turns out, the
logarithmic spiral is easy to produce indeed. In fact, it is difficult to avoid.
First of all, try to imagine life in a shell. You sit in there, cozy and protected,
but also a bit cramped. You want to grow, but the shell is rigid. What can you
do? One possibility would be to break or dissolve the shell, and build a new
and bigger one. Crabs and lobsters do that, but it is a real hassle and a risky
operation. Another strategy is to build more shell around the opening, and
move outwards. Imagine your shell being a box or a cylinder with one open
side. If you add to the opening, you will make the box or cylinder more
elongated. It will change shape, and you will have to deform with it in order to
fill it. This is not very smart, because you will depart more and more from the
most functional shape as you grow, and end up as a very long, thin thing. But
if your shell is instead like an ice cream cone, you can keep growing by adding
material to the edge, and magically keep your shape. Such a shape, staying
invariant under incremental addition, is called gnomonic, and the increment
itself is known as its gnomon.

© Springer International Publishing AG 2016 23


Ø. Hammer, The Perfect Shape, DOI 10.1007/978-3-319-47373-4_7
24 The Perfect Shape

Fig. 7.1 Orthoconic nautiloids (reconstruction)

Fig. 7.2 Top left: A conical shell with a trapezium as its gnomon. Bottom left:
Coiling it up produces a logarithmic spiral. Right: Crochet work by Maayke Koevoets

Some early nautiloids, and some peculiar ammonites called baculitids, did
make a long conical shell like this (Fig. 7.1). Straight-shelled nautiloids were
particularly common in the Ordovician and Silurian periods, some 400–480
million years ago. In many places around Oslo, where I live, there are spectac-
ular limestone surfaces packed with meter-long conical nautiloid shells. The
giant of the Ordovician, Cameroceras, grew to 6 m length or more. According to
simple calculation of the center of gravity and buoyancy of a cone, a straight-
shelled nautiloid should orient vertically with the tip up. But Cameroceras and
some related nautiloids secreted calcareous material inside the shell, perhaps
weighing down the rear end sufficiently to ensure a more horizontal position.
Now suppose you find that your long cone is getting cumbersome and
fragile. Why not coil it up, to make it more compact and robust? Strangely,
coiling a straight cone up tightly produces a logarithmic spiral (Fig. 7.2).
7 Curling Up 25

No fancy calculations are required by the animal—the logarithm is a natural


one! The almost unreasonable degree to which mathematics fit with nature has
yet again been demonstrated.
Christopher Wren (1632–1723), the master architect responsible for St
Paul’s Cathedral in London, may have been the first person to note the
similarity between spiral seashells and logarithmic spirals, and also to under-
stand that a logarithmic spiral can be made by coiling up a cone or pyramid.
The primary reference seems impossible to find, but John Wallis, in his
“Tractatus duo de cycloide etc.” (1659, pp. 107–108) mentions these matters,
giving credit to “our Wren”:

Hanc ipsam curvam, alia occasione, contemplatus item est Wrennius noster
This very curve has also been studied by our Wren.

You may ask what “coiling up a cone” is supposed to mean. It requires a


brutal deformation, as the outer side must be considerably stretched and the
inner side compressed. One way to proceed is to consider that the diameter of
a cone is proportional to its length. This is why it keeps it shape when adding a
trapezoidal gnomon. When coiling up the cone, this property should be
preserved, which will indeed happen if we deform the cone into a logarithmic
spiral. As shown in Appendix A.1, diameter D is then proportional to path
length s, as required, with a coefficient depending only on k:
 
k e2πk  1
D ¼ pffiffiffiffiffiffiffiffiffiffiffiffiffi s:
1 þ k2

For example, if the Nautilus, with expansion coefficient k ¼ 0.177, grows


forward 1 cm at the inner edge of the aperture, it will increase the diameter
of the aperture with 0.36 cm.
The gnomonic properties of the cone translate into self-similar properties of
the logarithmic spiral. Just as you can zoom in on the apex of a cone, and
always see a similar shape, you can scale a logarithmic spiral with a constant
factor c, and you will end up with exactly the same spiral only rotated by an
angle (ln c)/k:

crðφÞ ¼ cekφ ¼ elnc ekφ ¼ ekφþlnc ¼ ekðφþðlncÞ=kÞ ¼ rðφ þ ðlncÞ=kÞ:

Not only the snail and the Nautilus coil up their cones to save space and make
the structure more solid. The small but beautiful spiral shells of the polychaete
worm Spirorbis are very common on seaweeds and rocks along Atlantic coasts.
26 The Perfect Shape

Fig. 7.3 The B&W Nautilus loudspeaker. Image courtesy of Bowers & Wilkins

The near-conical tail of a chameleon forms a beautiful logarithmic spiral when


coiled up in resting position.
In loudspeaker design, it is important to reduce resonances caused by
reflections within the cabinet. By placing a long cone behind the speaker,
sound is reflected at low angle along the walls of the cone and directed
backwards, instead of reflected forwards again where it would cause acoustical
mayhem. A dampening material makes the sound die out before reaching the
end of the cone. The Bowers & Wilkins “Nautilus” loudspeaker works on this
principle (Fig. 7.3). For the high-frequency loudspeaker elements (the
tweeters), simple cones suffice. But for the bass element (the woofer), a straight
cone would be far too big for your living room. Coiling it up not only saves
space, but also makes the product look amazing. Form follows function.
The human cochlea, the sensory organ of the inner ear converting sound
waves to neural signals, works on a principle similar to the Nautilus loud-
speaker. Looking surprisingly like a snail shell, the cochlea is a near-logarithmic
spiral formed by almost three turns of a coiled-up cone filled with liquid.
Sound enters the opening of the cochlea and propagates through the liquid
towards the apex of the cone. On the way, the wave interacts with a parallel
wave set up in the basilar membrane, running inside the cone. The basilar
membrane is stiffest near the opening, causing it to be excited by the highest
frequencies (10–20 kHz) there. Becoming sloppier inwards, the membrane is
moved by progressively lower frequencies. Sensory hairs along the basilar
membrane are therefore reacting to different frequencies depending on their
7 Curling Up 27

Fig. 7.4 Mouflon (wild sheep), Ovis orientalis

position, producing an efficient spectrograph. The coiling allows the human


basilar membrane, which is almost 35 mm long, to be easily accommodated
within the skull.
Besides coiling up a cone there is also another way to form a logarithmic
spiral shell, perhaps even more elegant. If you let a cone grow a constant
percentage faster on one side than the other, it will curve into a logarithmic
spiral. This is probably the direct mechanism responsible for most logarithmic
spiral shapes in biology, including shells, teeth, horns (Fig. 7.4) and claws.
Wentworth Thompson’s beautiful book “On Growth and Form” (1917)
discusses this mechanism at length. But while the process is quite obvious in
2D, it is quite confusing in 3D. It is then necessary to specify not only how
much faster the outer edge grows compared with the inner edge, but also how
the growth rate varies in between, and the balance of growth rates in the radial
and the left–right directions around the edge. It is not easy to predict the final
shape from these growth parameters, but in most cases, the shape of the
opening will change through growth (Fig. 7.5). Such shape change through
growth, so-called allometry, is in fact very common in ammonite and snail
shells.
28 The Perfect Shape

Fig. 7.5 Three attempts at specifying radial and left–right growth rates as a
function of position around the growing edge of an ammonite. All three
simulations start with a circular cross section of the aperture. Adapted from
Hammer and Bucher (2005)
8
The King of Snails

Which way do snails coil? By convention, if you place a spiral shell with the tip
(apex) up and the opening facing you, a right-handed or dextral shell has the
opening to the right. Dexter is Latin for right. In a top view, the right-handed
shell coils down and clockwise. This is the normal coiling direction. Find a
snail and check! The opposite, left-handed coiling direction is called sinistral
(sinister being Latin for left). The nomenclature is logical and in accordance
with that used in mathematics. Another memory aid is that it easier to put your
right hand into a right-handed shell.
An old scheme, no longer in use, was to imagine walking into the opening
and up. If you walked to the left, the shell would be “leiotropic” (left-turning).
If you walked to the right, the shell would be “dexiotropic”. Hence, a
leiotropic or left-turning shell is a dextral or right-handed one, and a
dexiotropic or right-turning shell is sinistral or left-handed. This confusing
terminology was abandoned by conchologists ages ago, but still shows up now
and then in popular literature.
Jules Verne writes about these matters in “Twenty thousand leagues under
the sea”, in his usual stiff and delightful style:

“What is the matter, sir?” [Conseil] asked in surprise. “Has master been bitten?”
“No, my boy; but I would willingly have given a finger for my discovery.”
“What discovery?”
“This shell,” I said, holding up the object of my triumph.
“It is simply an olive porphyry.”
“Yes, Conseil; but, instead of being rolled from right to left, this olive turns
from left to right.”

© Springer International Publishing AG 2016 29


Ø. Hammer, The Perfect Shape, DOI 10.1007/978-3-319-47373-4_8
30 The Perfect Shape

“Is it possible?”
“Yes, my boy; it is a left shell.”
Shells are all right-handed, with rare exceptions; and, when by chance their
spiral is left, amateurs are ready to pay their weight in gold.

The naturalists’ reaction to their find is somewhat over-enthusiastic.


Left-handed (sinistral) shells are in fact found regularly, although many species
are almost exclusively right-handed.
The concept of left and right is, by the way, more complicated than we often
realize. Children learn the difference between up and down and front and back
quite early, but left and right is much trickier. How can this be? As explained
by the physicist Richard Feynman, Nobel laureate and legendary teacher, the
problem is well illustrated with a mirror. In the “mirror image”, left and right
are allegedly reversed. Point to your right cheek, and your double in the mirror
points to her left, right? But still, your head remains on top in the mirror, and
your feet at the bottom. Up and down are clearly not reversed. And, even more
strangely, if you point your finger onto the mirror at the right side, it meets the
mirror finger on the same side. Are then left and right not reversed after all?
What is going on?
The explanation is that the mirror is not really flipping anything at all,
except front and back. You are facing the mirror, but the mirror image is facing
you. Now left and right are defined with respect to up–down and front–back:
take your right hand, point your index finger forwards, your thumb up, and
your remaining fingers at right angles to both index finger and thumb. Your
remaining fingers now, by definition, point to the left. The left–right axis has a
secondary status compared with the primary axes of up–down and front–back.
It is all very complicated—no wonder it takes time to learn. Moreover, this
definition of left–right has a fundamental flaw: It refers to the right hand,
making the definition completely circular. It turns out that it is impossible to
communicate the concepts of left and right without pointing to some example
asymmetric object.
In German-speaking countries, there is a legend that a left-handed snail
specimen is a “snail king”—Schneckenkönig. The story applies especially to the
escargot Helix pomatia (also known as Roman snail or edible land snail) of
Europe. It is estimated that roughly one in a hundred thousand specimens of
H. pomatia is left-handed. Finding such a snail is believed to mean luck.
Hermit crabs live inside discarded snail shells, and have therefore evolved a
right-handed twist. Still, rarely, they are able to occupy sinistral shells. Kosuge
and Imafuku (1997) recorded five such cases in nature. I must also mention a
charming study by Imafuku (1994), where he looked at how hermit crabs
8 The King of Snails 31

knock sand out of snail shells. When encountering a dextral shell they always
turn the shell in the correct direction, which is to the left. Sinistral shells, on
the other hand, are sometimes turned to the left but usually to the right, until
the sand is dislodged.
We know that coiling direction in the snail shell is controlled by a single
gene. So why is coiling direction, seemingly of no functional consequence, not
equally distributed among left and right? The answer seems to lie in the
physiology of mating—it is well-nigh impossible for oppositely coiling snails
to copulate. Geometrically, the situation can be compared with hand shaking:
A right-handed person can easily shake hands with another right-handed, but a
left-handed and a right-handed person are much less compatible. This results
in a so-called symmetry breaking in the population, leading to the dominance of
one of the morphs. The dynamics may be similar to the symmetry breaking
that led to the dominance of matter over anti-matter in the universe (the two
cannot co-exist) or the symmetry breaking early in the history of life that led to
the dominance of one coiling direction, or chirality, in biomolecules such as
proteins and DNA (the two forms cannot function efficiently together).
It seems that whenever there is a rule in biology, there is an exception. A
very few snail species are amphidromine, meaning that there are significant
proportions of both coiling directions in the population. The most well-
known examples are in the genus Amphidromus, living in trees and shrubs
throughout Southeast Asia. Some species of Amphidromus are mostly left-
handed, some are mostly right-handed, and some, like Amphidromus palaceus
(Fig. 8.1), are amphidromine. Rebelling against the usual sexual convention
among snails, these animals do not have any trouble copulating with the
opposite chirality; in fact they positively prefer it (Schilthuizen et al. 2007).
Still, dominance of one handedness in a population is often seen in Nature
even when there is no functional explanation for it. A curious example is the
asymmetry in the human body, with the heart on the left side and the liver on
the right. Only one in about 10,000 people (Torgersen 1950) has it the other
way round, a condition called situs inversus. These mirror-image people
function absolutely normally, and there are no issues with their interaction
with others.
Really weird in this respect is the “thermometer foram” Neogloboquadrina
pachyderma, a marine planktonic foraminiferan (single-celled amoeba-like
organism) with a tiny spiral shell. That it exists in two versions, one left-
handed and one right-handed, is not so surprising. The odd thing is that the
right-handed form is mainly found in warm water, whereas the left-handed
one lives in cold water. The relentless coming and going of ice ages can be
traced through oceanic sediment cores by the ratio of left-handed to
32 The Perfect Shape

Fig. 8.1 A mixture of dextral (right-handed) and sinistral specimens of


Amphidromus palaceus, Java, Indonesia. This image is highly disturbing for a
conchologist. The shells are from 45 to 55 mm in length

right-handed fossil specimens. Needless to say, the reason remains a mystery,


but recently taxonomists have placed the two forms in two separate species,
implying a genetic basis for the phenomenon.
Oh, and one more thing. In Japan there is a snail-eating snake, Pareas
iwasakii, with an asymmetric jaw. Experiments have shown that this facilitates
attack on normally-coiled (right-handed) snails (Hoso et al. 2010). Conse-
quently, Japanese snails have co-evolved to have a much higher percentage of
left-handed shells than elsewhere in the world. What a terrible choice for the
snail: Should it be right-handed, making it easier to copulate, or left-handed,
reducing the risk of being eaten? And what about the snake: Should it consider
making left-handed jaws as a countermeasure?
9
Spira Mirabilis

The mathematical history of the logarithmic spiral involves an astonishing


array of great thinkers—a Who’s Who of seventeenth and eighteenth century
European academia. This fascinating shape acted like a memetic virus creeping
into the sharpest of minds. Indeed, the logarithmic spiral could have been an
appropriate graphic symbol for the Enlightenment.
As far as we know, the logarithmic spiral was not known to the Greeks,
although who can tell what was contained in the countless wonderful books
lost through the millennia. Apart from an obscure little glimpse of the
logarithmic spiral by Albrecht Dürer in 1525, it seems that the French
monk Marin Mersenne (1588–1648) started it all. Himself a talented math-
ematician, he was also a communication hub, or intellectual catalyst, keeping
an enormous correspondence with men such as Fermat, Pascal, Galileo,
Hobbes, Huygens and Torricelli. What a time! In a letter to the astronomer
Fabri de Peiresc (discoverer of the Orion Nebula), dated July 26, 1634, he
mentions that he is working on a “new curve”. But the first real work on the
logarithmic spiral is found in a letter back to Mersenne dated September
12, 1638 (Fig. 9.1), and then in a letter from October 11 the same year.
These letters are from none other than René Descartes. Now perhaps most
famous for less important contributions, such as being because he thought,
and dying from a cold he picked up in Sweden, Descartes was in fact one of the
most illustrious mathematicians of all time.
In his September 12 letter, Descartes is describing some results regarding
falling bodies, and then very briefly and without any proof, as if it were obvious
to anyone, brings together two important properties of the logarithmic spiral.

© Springer International Publishing AG 2016 33


Ø. Hammer, The Perfect Shape, DOI 10.1007/978-3-319-47373-4_9
34 The Perfect Shape

Fig. 9.1 Extract from Descarte’s letter to Mersenne of September 12, 1638, from an
1898 edition

The first is his starting point—that the total arc length from the origin (A in
his figure) along the spiral, say to a point C (ANBC in the figure), divided by
the radius length AC, is constant. This result illustrates the self-similar
(gnomonic) nature of the logarithmic spiral. The second property is that the
angle between the radius and the tangent vector is constant—the spiral is
equiangular (Fig. 9.2). This was a fairly humble beginning for our Perfect
Shape. No equation, no name, no proof. But there it was—a new curve for the
world to ponder. A bright future lay ahead.
It is difficult to track the history of the logarithmic spiral through the
following years. The idea seemed to spread very quickly, or was discovered
independently by many mathematicians. The next solid evidence is a paper
called ‘De infinitis spiralibus’ by Evangelista Torricelli (1644). In this text, he
works out both the rectification and the quadrature of the logarithmic spiral,
that is, he finds both the distance along the curve and the area that it encloses.
His arguments are ingenious but his infinitesimal methods are still geometrical
and reminiscent of Archimedes. True calculus was still some twenty years into
the future. In any case, the rectification of the logarithmic spiral demonstrated
9 Spira Mirabilis 35

α
α

Fig. 9.2 The equiangular property of the logarithmic spiral. The angles between
the spiral (black) and the radius vectors (red) are everywhere the same. Moreover,
the cotangent of this angle is equal to the expansion coefficient: k ¼ cotan α

that although the spiral coils around the origin infinitely many times, the total
length of the line inwards is finite.
In Appendix A.2, I have taken the time to repeat Torricelli’s achievement,
but this time with the modern methods of calculus as laid out by Leibniz and
Newton. Thanks to calculus, we can follow a standard recipe for this kind of
problem, almost without thinking. Thinking is hard work, so let us avoid it as
far as possible! The total distance s from the pole of the spiral (φ ¼ 1) to any
point, e.g., φ ¼ ϕ is then
pffiffiffiffiffiffiffiffiffiffiffiffiffi
a 1 þ k2 kϕ
sðϕÞ ¼ e :
k

The name “logarithmic spiral” was invented by Jakob Bernoulli (not by Pierre
Varignon as sometimes claimed). The earliest reference that I have found is in
the journal Acta eruditorum from 1691. The paper begins as follows:

Si in plano circuli BCH jaceat curva BDEIPC, quam secent, eodem angulo
obliquo, radii CB, CL &c. ex centro circuli C educti, dicetur Curva haec Spiralis
Logarithmica; quoniam sumptis arcubus LM, MN &c. inifinite parvis &
aequalibus, hoc est, ipsis BL, BM, BN, arithmetice proportionalibus, radii DC,
EC, IC, sunt geometrice proportionales, ob triangula similia DCE, ECI, &c.
36 The Perfect Shape

Fig. 9.3 Logarithmic spiral, from Jakob Bernoulli (1691), Acta eruditorum, p. 282

My attempt at translation into English follows (refer to Fig. 9.3):

If, in the plane circle BCH, the curve BDEIPC is laid down, meeting the radii CB,
CL etc. from the centre C of the circle by the same oblique angles, this curve is called
a logarithmic spiral; because supposing the arcs LM, MN etc. infinitely small and
equal, BL, BM, BN are arithmetically proportional and the radii DC, EC, IC
geometrically proportional, on account of the similar triangles DCE, ECI etc.

This much had already been known since Descartes. The first part refers to
the equiangular property of the spiral, while the second (in the lopsided,
geometrical language typical of the time) refers to the exponential form in
polar coordinates.
In a paper in Acta eruditorum in the following year (1692; p. 245), we find
the term “Spira mirabilis”, the wonderful spiral. Bernoulli was obviously
hypnotized by this curve. Among other things, he observed that the logarith-
mic spiral magically stays invariant under a number of geometric transforma-
tions, including the so-called pedal and evolute (Fig. 9.4).
Self-similar, equiangular, infinitely convoluted yet finitely long. To this list
of curious properties of the logarithmic spiral, let me add another. Some
mathematicians would claim that the title of the Perfect Shape should go to
the circle (or more generally the sphere), not the logarithmic spiral. After all,
the circle is the locus of equidistance from a point. It has maximal symmetry.
But apart from the many beautiful properties of the logarithmic spiral, most
notably its scale invariance, it can be argued that the circle is but a degenerate
logarithmic spiral where the parameter k is zero:

r ¼ aekφ ¼ ae0 ¼ a:
9 Spira Mirabilis 37

Fig. 9.4 Jacob Bernoulli’s tomb at Münster Cathedral, Basel, Switzerland. Eadem
mutata resurgo, “although transformed I stay the same”. Note the workman’s
failure to produce a logarithmic spiral. Photo: Wolfgang Volk, Berlin

Fig. 9.5 Logarithmic spiral, k ¼ 0.015. As k ! 0, the spiral will fill the plane

But here lurks a funny thing. No matter how small, but non-zero, you choose
k, the spiral remains exactly that: a spiral. In the limit, it gets more and more
tightly wound, filling the paper until it gets black, but there is no sign of it
approaching a circle (Fig. 9.5). Only when k is exactly zero does the shape flip
into this totally different state. Maybe we can call it a geometric discontinuity,
but no discontinuity at zero is readily apparent in the analytical equation itself.
Another surprising property of the logarithmic spiral is that if you roll it
along a straight line, the trajectory of the center of the spiral will also be a
straight line (Fig. 9.6). Such a trajectory is called a roulette. You can take a
Nautilus shell, for example, and roll it over a table. As shown in Appendix A.3,
the axis of the shell (the umbilicus) will travel along a straight line. The slope of
this line is equal to the expansion coefficient k. For the limiting case of a circle
(k ¼ 0), the line will be horizontal.
38 The Perfect Shape

1
2
3
4
5
6

Fig. 9.6 A logarithmic spiral rolls from left to right over a flat table. Positions are
shown for successive 60 rotations. The center of the spiral will travel along a
straight line (red)

One more thing. Different properties of the strange logarithmic spiral lead
to several different, equivalent definitions. We may call it logarithmic, or
equiangular, but we might also call it the spiral of radial acceleration. Consider
again its equation in polar coordinates, r ¼ aekφ. Now a peculiar property
of the exponential function is that it is its own derivative, up to a constant:
dr/dφ ¼ akekφ. In other words, dr/dφ ¼ kr. Now imagine that we are generat-
ing the spiral by rotating a line around the origin, with an angular velocity
v (in radians per second), so φ ¼ vt. Plugging into the equations above, and
with the kernel rule, we get a time derivative
dr dr dφ
¼ ¼ kvr:
dt dφ dt

In other words, the velocity along the rotating radius is proportional to the
radius. Wentworth Thompson (1917) and Ghyka (1946) therefore define the
logarithmic spiral as “a plane curve produced by a point moving on the line
(vector radius) with a speed proportional to its distance from the pole”. This is
analogous to Archimedes’ definition of his spiral, where the speed is constant.
Formally, I suppose, in order to establish the identity of definitions, we also
need to prove the opposite inference, namely that this property of radial
acceleration leads to the logarithmic spiral equation. We can obtain this
proof by integration.
10
Unfortunate Moths and Lopsided Falcons

It is amazing that we do not know why moths are attracted to lights. In the
beauty of the night they fly into open flames like kamikaze pilots, their reasons
eluding us. But the prize for the mathematically most elegant idea goes to the
theory of “transverse orientation”. Let us assume (we do not know this either)
that moths navigate by light. If so, we can imagine how they could maintain a
straight path by flying at a constant angle with respect to a natural light such as
the moon, or the sun under the horizon at dusk or dawn (Fig. 10.1, top). The
path will be straight at least for a short period of time, while the moon or sun
doesn’t move too far in the sky.
This would only work because the moon and sun are so very far away,
making their rays practically parallel. Now put up an artificial light very close
to the moth. Trying to keep a constant angle to the light, the poor creature will
succumb to math and enter a Spiral of Doom (Fig. 10.1, bottom). As we saw
in the previous chapter, the constant angle between the spiral line and the
radius vector is a defining property of the equiangular or logarithmic spiral.
If it is not true, at least it is a good story.
Here is another good story. Raptor birds (falcons, hawks and eagles) have
more acute vision about 45 to the side of the head. Tucker (2000) studied
carefully the head positions of falcons, hawks and eagles watching him. The
birds tended to twist their heads about 40 to the side, clearly to get a sharper
view. So what should the raptor do when diving towards its prey? It has two
options. It can twist the head to the side, but this increases aerodynamic drag.
Or it can fly in such a way that the prey is always 40 to the side of the tangent
of the line of motion, producing a logarithmic spiral path just like the moth

© Springer International Publishing AG 2016 39


Ø. Hammer, The Perfect Shape, DOI 10.1007/978-3-319-47373-4_10
40 The Perfect Shape

Fig. 10.1 Flight by transverse orientation. Top: Light source far away (natural
condition). Bottom: Light source at short distance. All angles indicated are equal

40˚

40˚

Fig. 10.2 The approach path (red) of a hawk keeping a constant 40 angle to the
side of the prey (black dot)

(Fig. 10.2). Tucker calculated that although this path is longer than the
straight line to the prey, it is faster because keeping the head straight minimizes
the drag. And this is not only theory. Tucker et al. (2000) set up an elaborate
optical tracking system in the Rocky Mountains, measuring the paths of
peregrines. The flight paths were indeed curved, approximating the theoretical
logarithmic spirals to a fair degree.
Now imagine that instead of travelling at a constant angle with respect to a
light source or prey, you go by magnetic compass at a constant bearing with
respect to North. Just like the moth, you will do just fine as long as the North
Pole is very far away compared with your distance of travel. But if you travel
very far, or if you are close to the Pole, you will move in a spiral, not in a
straight line, and you will inevitably end up on the Axis of the World. We will
come back to this in a later round.
11
Circular Tessellations

Some particularly beautiful examples of logarithmic spirals turn up if we cover


the plane with equally shaped geometric elements such that the number of
elements along concentric circles is constant. An excellent illustration is the
astonishing floor mosaic from a second century AD Roman villa in Corinth,
Greece, showing Dionysos (some say Apollo) in the center surrounded by a
circular tessellation of triangles (Fig. 11.1). The dizzying display with the god
of wine in the middle must have been quite a party piece! The question is what
the local Greek mathematician would have said if invited to dinner, for the
geometry here is worth a closer study.
Consider one colored triangle. The outward-pointing vertex is roughly
right-angled, while the two basal angles are about 45 . The base of the triangle
sits on a circle. The result is that the angle between a side of a triangle and the
straight radial line into the center is always approximately 45 , producing an
intricate set of crossing equiangular (logarithmic) spirals with expansion coef-
ficient k ¼ cotan 45 ¼ 1 (Fig. 11.2). In the mosaic, some of these spirals are
marked with black triangles.
The little halo of spirals around Dionysos and the double helix framing it all
are intriguing—is it possible that the artist was aware of the spiral properties of
the tessellation, and decided to let the other elements follow the same theme?
Calter (2000) called attention to a similar mosaic from Pompeii. Noting the
logarithmic spirals, he called it a logarithmic rosette. This construction is also
found in some of M. C. Escher’s work. In Fig. 11.3, there are 12 birds in each
concentric circle, and the resulting logarithmic spirals are marked with lines.

© Springer International Publishing AG 2016 41


Ø. Hammer, The Perfect Shape, DOI 10.1007/978-3-319-47373-4_11
42 The Perfect Shape

Fig. 11.1 Roman floor mosaic from Corinth, Greece, second century AD. Photo
Carlos Parada

45º

45º

Fig. 11.2 Detail of a circular tessellation of right-angled triangles. The sides of the
triangles make 45 angles with radial lines, producing an approximately equiangular
(logarithmic) spiral. The circle radii are in geometric progression

There is a strange tradition for decorating the insides of architectural domes


with such logarithmic rosettes. The enormous dome of the church in Mosta,
Malta is a good example. With 32 elements in each circle (like the Dionyosos
mosaic!), the diminishing sizes towards the pole exaggerate the perspective,
giving a magnificent impression (Fig. 11.4).
11 Circular Tessellations 43

Fig. 11.3 M.C. Escher’s Path of Life III, 1966. © 2016 The M.C. Escher Company—
The Netherlands. All rights reserved. www.mcescher.com

Fig. 11.4 Rotunda of Mosta, Malta, consecrated 1871. Photo Jean-Cristophe


Benoist. Creative Commons Attribution-Share Alike 3.0 Unported license
44 The Perfect Shape

Fig. 11.5 Sheikh Loth Allah Mosque, Isfahan, Iran, finished 1618. Photo Adam
Jones. Creative Commons Attribution-Share Alike 2.0 Generic license

An even more impressive case is the dome of the Sheikh Loth Allah Mosque,
Isfahan, Iran (Fig. 11.5). Again it has 32 elements in each circle—was the
Mosta dome inspired by the mosque in Isfahan?
More modern, but no less beautiful, is the Gemasolar power plant near
Seville in Spain (Fig. 11.6). Here, 2650 colossal mirrors called heliostats track
the Sun through its daily journey through the sky, reflecting the light onto a
140 m high tower where it heats up a reservoir of molten salt, driving a steam
turbine. The salt keeps driving the turbine through the night, making
Gemasolar a continuous, 24 h, 20 MW power source, saving some 30,000 t
of CO2 emissions every year. The heliostats are placed into three radial zones,
the outer two arranged as staggered circular tessellations with their k ¼ 1
logarithmic spirals. This association of the Sun with spirals resonates deeply
with ancient symbolism, as we will discuss later.
But the most spectacular example of circular tessellations must be the
pattern of eyes on the peacock’s tail (Fig. 11.7). In Fig. 11.8, the oldest and
therefore largest feathers are shown in light gray (outermost semicircle of eyes).
As all the feathers are growing, there will come a time when the youngest
feathers (black, innermost semicircle) have moved sufficiently far out that a
new generation of small feathers can be initiated, intercalated between the
previous generation. If this simple procedure is carried out with perfect
11 Circular Tessellations 45

Fig. 11.6 Aerial view of the Gemasolar power plant near Seville, Spain. The helio-
stat field is about 1.6 km across. Image courtesy of Gemasolar solar thermal plant,
owned by Torresol Energy ©SENER

Fig. 11.7 Indian peacock, Pavo cristatus. Photo N.A. Naseer, www.nilgirimarten.
com, Creative Commons Attribution-Share Alike 2.5 India
46 The Perfect Shape

Fig. 11.8 Cartoon of peacock tail development, with youngest feathers in black. A
logarithmic spiral trace is indicated

precision, the bird is rewarded with an astonishing, regular, circular tessella-


tion, with logarithmic spiral rays of eyes. It provides an easy-to-read gauge for
genetic quality in the male, as any small imperfection in the timing of
development will lead to an obvious pattern dislocation.

How to Brace a Circle


Structural rigidity was a major concern in the construction of the majestic
zeppeliners of the 1920s and 30s, and when the German engineer Karl
Arnstein designed the US naval airships USS Akron and USS Macon in
1929, the strength of the hull had top priority. The Akron was 239 m long,
with circular girders 22.5 m apart. Each of these girders was amply braced by
tension wires to make them stiff (Fig. 11.9). To reduce torsion, the wires took
up not only radial but also tangential forces. Everywhere the wires met at a
constant angle, both to each other and to the radial forces that they set up. The
constant angles made forces equally strong across the mesh. This can only be
done by circular tessellation, giving an approximately equiangular spiral tra-
jectory for each wire.
The Hindenburg did not have such an arrangement, but used a simpler
configuration with straight radial wires. The cause of the famous disaster in
1937 has never been identified, but the investigations at the time pointed out
that sharp turns made during the preparations for landing might have stressed
the hull, causing a bracing wire to snap and damage a hydrogen cell. Perhaps
35 lives had been saved if logarithmic spirals had been used instead.
11 Circular Tessellations 47

Fig. 11.9 The structural design of the circular girders on USS Akron. Photo pro-
vided by Dan Grossman, www.airships.net
48 The Perfect Shape

Sunflowers
The florets in a sunflower or coneflower head form beautiful spirals that have
been subject to centuries of research. Together with the Nautilus, the sun-
flower is an icon of mathematical biology. The florets start out as small
primordia in the central part (the apex), and migrate outwards as they grow.
If the florets prefer to maintain their shapes while staying closely packed, we
would expect a circular tessellation to result. This is almost the case, and quasi-
logarithmic spirals are indeed obvious, but there are kinks and irregularities
(try to follow a spiral from the outer edge and towards the center in the
coneflower, Fig. 11.10).
A famous and enigmatic property of sunflower heads is that the number of
spirals in one direction (known as parastichies) tends to be a Fibonacci number.
Not only in sunflowers by the way, but in many plants. The Fibonacci
numbers form a series such that each number is the sum of the two preceding
ones: 1, 1, 2, 3, 5, 8, 13, 21, 34, etc. There is no end to the strange stories and

Fig. 11.10 Top: Coneflower head. Bottom: Base of pine cone with 8 left-turning
and 13 right-turning parastichies. 8 and 13 are successive Fibonacci numbers
11 Circular Tessellations 49

facts pertaining to these numbers, there are books, journals, lives dedicated to
them. Fibonacci numbers in plants have given rise to a whole scientific
industry. This phenomenon has never been really understood, and only
recently have things started to clarify.
Let us assume that the plant “wishes” to pack the florets as densely as
possible. In addition, we recognize that the florets are initiated one at a time,
just outside the center of the apex, at a constant angular distance from the
previous floret. This is called spiral phyllotaxis, and is a fundamental principle
of plant growth. Finally, we make the simplification that the florets do not
grow after initiation (thereby removing ourselves slightly from the logarithmic
rosettes). Now the question is: What angular increment will give the optimal
packing? If we try 1/6 of 360 , or 1/8, or another unit fraction of the whole
circle, we will not do very well (Fig. 11.11).
The florets radiate in six or eight straight lines, and are not packed very
efficiently. It does not help much to use other simple fractions such as 2/3 or
3/4 of 360 . Irrational numbers such as the square root of two are a little
better. But since we have mentioned Fibonacci, what happens if we use ratios
of successive Fibonacci numbers? Let us try, first 5/8 and then 8/13
(Fig. 11.12).
In the middle of the flower, 5/8 works fairly well, but further out the florets
organize themselves in eight straight lines again. However, for 8/13 something
strange happens. The florets are more evenly distributed, and after some
squinting, we can follow the spiral parastichies. I see five spirals bending out
counterclockwise. If we make a sunflower where the angle between successive
florets is a ratio between successive Fibonacci numbers, the number of
parastichies will also be Fibonacci numbers (I guess this can be proven). In
these simulations, by the way, the radius is made proportional to the square
root of the angle.

Fig. 11.11 60 (1/6 circle) and 45 (1/8 circle) between successive florets. I have
drawn lines between successive florets, with the newest floret innermost
50 The Perfect Shape

5/8 8/13 8/13


With parastichies shown

Fig. 11.12 Packing of florets with angular distances of 5/8 and 8/13 of a circle

Continuing in this manner, we get an increasingly even packing the further


we proceed in the Fibonacci series (13/21, 21/34, etc.). It turns out that we
achieve an optimal packing if we continue to the limit, that is infinitely far out
in the sequence. The ratios then converge to (√5  1)/2, the inverse of the
Golden Ratio. Figure 11.13 shows what it looks like for the first 250 florets if
we use an angular increment of 360  (√5  1)/2, or about 222.49 . I count
34 parastichial spirals in each direction in the outer part. If I continue the
program and plot many more florets, the number of spirals is expected to
increase to the next Fibonacci number, which is 55.
From a geometric perspective, it is therefore possible to model the arrange-
ment of florets according to a simple model (Vogel 1979). In polar coordi-
nates, and with n the index of a floret, we can use:
pffiffiffi !
51
ϕ ¼ 2πn 1 
2
pffiffiffi
r¼c n

This formulation gives a 137.5 angle between consecutive florets. The model
does not represent the actual biological process, because the model florets are
added at the periphery and do not move. It is often noted that Vogel’s model
produces a close packing of similar-sized elements, but this does not quite
capture the situation in the sunflower, where the florets increase in size as they
move out. Vogel’s model based on the sunflower was suggested by Noone
et al. (2012) for arranging heliostats in solar power plants, giving a more
efficient packing than the circular tessellation zones used in the Gemasolar
plant (Fig. 11.6).
Vogel’s equation describes a parabolic spiral (also known as Fermat’s spiral),
where the radius increases as the square root of angle. It is important to
11 Circular Tessellations 51

Fig. 11.13 The first 250 florets in Vogel’s sunflower equation. Program code in
Appendix B

differentiate between such a spiral traced out by the sequential addition of


primordia (the generative spiral), and the apparent spirals (parastichies) seen in
the sunflower head.
There the matter stood until quite recently. To recapitulate, the story went
somewhat like this: The sunflower is constrained to use a certain machinery,
which we call spiral phyllotaxis, a product of its evolutionary history. To
achieve optimal packing within this framework, it measures up exactly
137.51 (the Golden Angle) between successive florets. Due to “deep” math-
ematical relationships, which we trust the experts to deal with, this leads to the
number of spiral parastichies becoming Fibonacci numbers.
But this story has long been considered unsatisfactory. The sunflower must
control the angle very accurately for the mechanism to work. How can a plant
measure angles with such precision? Some physicists and mathematicians
(particularly Douady and Couder 1996) started playing with entirely different
models. And then, in a series of publications from 2002 until now, several of
them in the prestigious journals Science and Nature, molecular biologists
achieved a breakthrough that turned the whole story upside down. A good
review of the present knowledge was given by Sassi and Vernoux (2013).
It turns out that there is a signaling molecule, auxin, which is broken down
by the cells at the site of an initiating floret. At the same time, the cells measure
the concentration of auxin in their surroundings. When the level becomes
sufficiently high, it means that the density of florets nearby is low. This triggers
52 The Perfect Shape

the initiation of a new floret. The result is a self-organizing system producing a


spacing pattern.
The plant does not sit there measuring angles like an engineer. Nature does
not work that way. The plant does not set out certain angles in order to achieve
optimal packing. On the contrary, the optimal packing self-organizes, and the
angle is but a byproduct. This explains the precision and robustness of the
system. In addition, this mechanism allows the florets to grow after initiation,
as we see in nature. Stevens put it quite bluntly already in 1974 (quoted by
Davis 1993):

It simply grows its stalks or florets in succession around the apex of the stem so
that each fits the gaps of the others. The plant is not in love with the Fibonacci
series; it does not even count its stalks; it just puts out stalks where they will have
the most room.
12
Ropes and Rifles

Modern mathematicians differentiate between spirals, which are basically


two-dimensional figures with increasing radius as a function of polar angle,
and helices, which are screw-shaped curves on the surface of a three-
dimensional cylinder or cone. From an aesthetical, intuitive point of view,
spirals and helices are in the same class, both being winding, open curves with
no beginning or end—finite yet infinite, swirling, mesmerizing. Archimedes
used the word “helix” for his spirals, and in everyday language, we often mix up
the two, as when we talk of spiral staircases (which are actually helices).
Like trochospiral shells, a helix can be left-handed or right-handed. If you look
at the helix end-on, a right-handed helix will turn away from you and clockwise.
A left-handed helix will turn away from you and counter-clockwise. It doesn’t
matter which way you look at it—a right-handed helix is right-handed as seen
from both ends. Another way to remember it (Fig. 12.1) is that the coils on the
near side of a right-handed helix go the same way, lower left to upper right, as the
pencil shading of a right-handed person. Of course, the most iconic helix is the
double helix of DNA. Normal DNA in a cell is a right-handed helix.
The handedness of a twisted rope is called the lay. You could make a twisted
rope by fixing the three strands at one end, and turning the other end around.
Turning the rope clockwise is natural for a right-handed person, producing a left-
handed triple helix (yes it is confusing, but try for yourself and you will see what I
mean). Some of the most ancient ropes known were most likely made in this
way, reflecting the handedness of the rope maker. The left-handed chords worn
by the Stone-Age “ice man” Ötzi found in the Alps in 1991, from about 3200
BC, is a spectacular example. However, mechanical devices for rope making were

© Springer International Publishing AG 2016 53


Ø. Hammer, The Perfect Shape, DOI 10.1007/978-3-319-47373-4_12
54 The Perfect Shape

Fig. 12.1 Left: A left-handed helix (near side shaded gray), and the shading of a left-
handed person. Right: A right-handed helix, and the shading of a right-handed person

Fig. 12.2 Left: The bryozoan Archimedes, left-handed form (reconstruction).


Right: Right-handed spiral chute for grain bags, ca. 1920, originally 16 m tall.
Moss Town and Industry Museum, Norway. Photo Marte H. Jørgensen

developed already in ancient Egypt, and since that time there has been no
particular preference for left-handed rope work (in contrast with the claims of
Cook 1914). Most modern twisted rope is right-laid, as a matter of convention.
If we “fill in” the helix in to the coiling axis, we produce a coiling screw
surface called a helicoid. A machine screw is an approximate helicoid but with a
cylindrical core. A fascinating helicoid-like organism is the extinct
(Carboniferous-Permian) bryozoan with the appropriate name Archimedes
(Fig. 12.2). This strange colonial animal could be either left-handed or
right-handed, and could reach a height of 10 cm or more. Thousands of
12 Ropes and Rifles 55

Fig. 12.3 The view down my Tikka T3 rifle with right-handed, four-grooved
rifling. The caliber is .30-06 (7.62 mm diameter), the barrel is 57 cm long and the
rifling completes one revolution in 28 cm (the twist rate is 1100 )

small individuals generated a water current entering the colony from the top,
moving down and outwards through the perforated whorls. Clearly, the
individuals living at the top or the periphery would get most of the food
share, while those poor old ones at the bottom center would slowly starve to
death as the colony grew (McKinney et al. 1986).
The rifling in the barrels of most modern firearms is also constructed as a set
of helices. Handguns and rifles commonly use a quadruple or sixfold helix
(Fig. 12.3). After firing, the bullet expands inside the barrel, and the helical
grooves of the rifling cut (“engrave”) into it. This forces the bullet into a rapid
spin which stabilizes it like a gyroscope, making it easier to hit your target.
Most spirals in this book are Good. The spirals of the rifle are Evil.
If you bend a helix around into a circle, and connect the two ends together,
you produce a continuous, beautiful structure called a toroidal helix. It makes a
handsome bracelet or finger ring, popular in the Bronze Age and the Iron Age.
56 The Perfect Shape

PLASMID

Fig. 12.4 Left: Ninth century bracelet in gold, southwest Russia. Walters Art
Museum, licence CC-BY-SA-3.0. Right: A plasmid is DNA connecting to itself, making
a double-toroidal helix

A short piece of double-helix DNA can also connect its two ends, making a
ring-shaped plasmid. The plasmid can itself twist into a super-helix (Fig. 12.4).
The toroidal helix is also important in fusion energy. In the exotic plasma
confinement device known as the stellarator, the ions follow a trajectory that is
sometimes on the inside, sometimes on the outside of a torus, ensuring that
inwards and outwards forces are cancelled.
13
The Lost Sea of Spirals

The Mesozoic, or the Age of Dinosaurs, was a time of magic, a lost era of
strange beasts on land and in the sea, of true monsters of incomprehensible
size, a time of nightmares but also of immense beauty. During this enormous
expanse of time, from 252 to 66 million years ago, there appeared the
Tyrannosaurus rex, pterosaurs as large as airplanes, thunderous sauropods,
ichthyosaurs, mosasaurs, plesiosaurs. But invertebrates were also wonderful
back then. In the oceans there were innumerable ammonites. Their fossils are
everywhere in Mesozoic rocks, turning up in great numbers even in unex-
pected places such as the limestone floors of banks and airport terminals.
Although the basic spiral shape of an ammonite is easily recognizable, they
were a diverse group with thousands of species, making them ideal for dating
sedimentary rocks. Their countless, perfect spiral shells must have washed up
in great piles on the beaches, their mother-of-pearl glittering in the sun,
crushed by the feet of passing dinosaur herds. Ammonite shells are among
the most beautiful things ever created by living organisms (Fig. 13.1).
The basic body plan of an ammonite is similar to the Nautilus. They are
both swimming cephalopods, with a (usually) spiral shell divided into cham-
bers, and with a long string, the siphon, running through it. Moreover, we
believe that the ammonites evolved from nautiloids, through a protracted
evolutionary sequence lasting for a hundred million years prior to the Meso-
zoic (to be pedantic, the true ammonites did not appear until the Jurassic,
about 200 million years ago; the earliest Mesozoic forms are properly called
ammonoids). But ammonites were quite distinct from nautiloids in many
respects. One obvious difference in the fossils is the shape of the walls—the

© Springer International Publishing AG 2016 57


Ø. Hammer, The Perfect Shape, DOI 10.1007/978-3-319-47373-4_13
58 The Perfect Shape

Fig. 13.1 The ammonite Dactylioceras, Early Jurassic, Yorkshire, England. 7 cm


diameter

septa—separating the gas chambers. The contact between the septum and the
outer shell defines a line called the suture. The septum of a nautiloid is a
simple, smooth, curved surface. The septum of an ammonite is not at all
simple. It is incredibly curly, with small folds upon larger folds, producing an
intricately fractal suture line.
Although ammonites are very common fossils, they remain mysterious. The
soft parts are preserved extremely rarely in fossils, and then only as vague
shadows that are debated endlessly among scholars. From time to time, radical
theories turn up, such as ammonites being sedentary animals bound to the sea
floor, or the shell being an internal rather than external structure, with some
large, unknown soft body surrounding it. These may be unlikely scenarios, but
they illustrate that the fossil evidence of ammonite morphology is limited.
Most ammonites were fairly small, usually with a diameter of 5–10 cm, but
there were also giants. Nobody knows what spiral behemoths could have
glided through the depths of the Jurassic or Cretaceous seas, like pelagic
Francis turbines, but we do have some extraordinary fossils that give us a
glimpse. Largest of them all is Parapuzosia seppenradensis from the late Creta-
ceous, about 75 million years old. The record specimen was found in 1895 in
Westfalen, northwest Germany. The diameter is 174 cm, but the outermost
chamber is not preserved and it is likely that the original shell would have been
a staggering 3.5 m in diameter (Teichert and Kummel 1960).
14
The Great Spiral in the Sky

The class structure of nineteenth century England and Ireland may not
conform to our modern ideals, but we cannot deny that it allowed for some
extraordinary scientific achievements. Take William Parsons, the third Earl of
Rosse, for example. Born in the year 1800, his father the 2nd Earl of Rosse and
owner of a large estate in Parsonstown (now Birr) in central Ireland, he was not
destined to experience the miserable living conditions of most of his compa-
triots. Combine that economic freedom with a brilliant mind and a talent for
practical engineering, and you have a recipe for scientific success.
William Parson’s passion was telescope construction. In 1842, he started on
a project of truly vast proportions: the “Leviathan of Parsonstown”, a gigantic
reflector of the type pioneered by Newton, with a 1.8 m diameter primary
mirror of bronze and a telescope tube 17 m long. Several assistants were
needed to swing the uncooperative construction about, while the observer
was suspended precariously some 20 m above ground. And, most astonishing
of all, despite appearances it was a precision instrument with excellent optics
(Fig. 14.1).
In April 1845, in the early days of the Great Famine, on a clear, moonless
night, the Earl looked up at the Great Dipper, followed its handle to the end,
and continued about 3 and 30 min to the southwest. Orders were given, and
the Leviathan was slowly and reluctantly rotated towards its target, a hazy spot
in the constellation of Canes Venatici. Charles Messier had discovered this
faint nebulosity in 1773, and given it the number 51 in his famous catalogue.
It must have been an exciting moment for the Earl. The nature of the
enigmatic, nebulous patches in the night sky, many of them visible to the

© Springer International Publishing AG 2016 59


Ø. Hammer, The Perfect Shape, DOI 10.1007/978-3-319-47373-4_14
60 The Perfect Shape

Fig. 14.1 The Leviathan of Parsonstown, a 183 cm aperture Newtonian telescope


built by the 3rd Earl of Rosse. Picture dates from before 1914

naked eye, had been debated for a century. Some of them could be resolved
into individual stars by modest optical equipment, and these star clusters did
not cause much controversy. Others remained ghostly clouds of indeterminate
shape. In the 1750s, Immanuel Kant and others had speculated that some of
the nebulae could be distant “island universes”—clouds of innumerable stars
like our own Milky Way, but much further away. In 1845, this fundamental
question of the structure of the universe had not yet been settled.
M51 is not visible without optical aid, although a pair of 7  50 binoculars
is sufficient to spot it under dark skies, barely visible as a fuzzy but quite large
patch of ghost-like luminosity. But when Earl Rosse looked at M51, he saw
something truly beautiful. He saw the Whirlpool Galaxy—a dramatic swirl of
dim light, of interlocking spirals (Fig. 14.2). And he could glimpse tiny points
of light within it, a glimmering of stars. In the years to come he would study a
large number of similar nebulae, some spiral, others elliptical.
Earl Rosse was the first observer on Earth to see the spiral structure of a
galaxy. Little did he know that he was looking at light that had started its
journey 23 million years ago, at the end of the Oligocene, when nimravids,
entelodonts, oreodonts and other fantastic mammals walked the Earth. It was
14 The Great Spiral in the Sky 61

Fig. 14.2 M51 as drawn by the Earl of Rosse (1850)

not until the 1920s that astronomers, most notably Edwin Hubble, demon-
strated the enormous distance of galaxies, confirming the old speculations of
Kant (Fig. 14.3).
At first sight, the spiral arms in galaxies look like long whisks of stars drawn
out by the higher angular velocity in the center, similar to splotches of milk in
rotating coffee. But this cannot be correct, because a typical galaxy will have
rotated hundreds of times since it was formed, and the arms should therefore
have been much more tightly wound. The most popular modern theory
involves a spiral density pattern that rotates with a velocity different from
that of the stars and dust (Bertin and Lin 1996). As it turns out, spiral arms in
galaxies tend to be of the logarithmic type. We might get back to this later on.
The “Cosmological Principle” is a hypothesis in cosmology saying that on
large scales, the Universe should be homogenous (similar everywhere) and
isotropic (similar in all directions). Another way of putting it is that the
Universe should look similar for all observers, no matter where they are
situated and in what direction they are looking. But there is something very
odd about spiral galaxies, seemingly in conflict with the Cosmological Princi-
ple: They do not seem to spin in random directions. This peculiar observation
62 The Perfect Shape

Fig. 14.3 M51 as seen by the Hubble space telescope. Counterclockwise (left)
rotation. Credit: NASA, ESA, S. Beckwith (STScI), and The Hubble Heritage Team
(STScI/AURA)

was first made by Michael Longo in 2007 (Longo 2011) and then confirmed
in later studies, including a massive automated survey of 126,501 spiral
galaxies by Shamir (2012). Spiral galaxies are classified as either clockwise
rotating (right handed) or counterclockwise (left handed). Now of course the
handedness depends on the position of the observer—a left galaxy would be
right if viewed from the opposite side. If galaxies rotate randomly, with no
preference for left or right, we should see an even distribution of handedness in
whatever direction we point our telescope. This is not what Longo and Shamir
found. If you look towards a point roughly in the constellation of Cancer,
there will be about 3 % more left handed spirals than right handed ones.
Conversely, if you look in the opposite direction, in the direction of Sagit-
tarius, there will be a couple percent more of the right-handed variety. The
pattern is similar whether you look at galaxies close by or far away. The signal is
weak and the statistics a little iffy, but the data seem to show that galaxies tend
to prefer a certain spin direction, which looks left or right to us depending on
the direction we look in (Fig. 14.4).
If true, this indicates an overall asymmetry in the universe. Such an asym-
metry cannot be reconciled with the Cosmological Principle.
14 The Great Spiral in the Sky 63

Looking in
this direction,

Earth
Looking in
this direction,
we see this:

we see this:

Fig. 14.4 If there is a preferred handedness for spiral galaxies throughout the
observable universe (black galaxies) we would see a different handedness (red)
depending on the direction we look
15
The Case of the Staircase

Spiral, or more precisely helical, staircases provide beautiful and space-efficient


access to upper floors. Although they are not at all trivial things to build, we
can find spiral staircases in some very old building structures. Noteworthy
examples are the tower of Agios Petros on the Greek island of Andros
(400–200 BC), ca. 20 m tall; the Broch of Mousa in Shetland (100 BC),
ca. 13 m; and the astonishing Trajan’s Column in Rome (113 AD), 30 m tall
plus a 5 m pedestal.
The handedness of spiral staircases is an interesting issue. According to Cook
in his classical book “The Curves of Life” (1914), medieval spiral staircases
were usually left-handed, meaning that you turn right as you are walking
up. The purpose was apparently to make it more difficult for a right-handed
attacker ascending the stairs to use the sword in his right hand. The defender at
the top, on the other hand, would be free to use his right hand in the outer part
of the staircase. This story is now commonly accepted, but exceptions abound
and there is perhaps reason to be skeptical. Guy (2011) reported that left-
handed (“clockwise”) staircases dominated in England and Wales in the
Norman period (1070–1200), but right-handed ones became increasingly
common thereafter. He lists 85 examples of right-handed staircases from
1070 to the 1500s. Agios Petros and Trajan’s Column are right-handed,
Broch of Mousa is left-handed. The handedness of modern spiral staircases
seems to be rather arbitrary—I suppose there isn’t much swash buckling
anymore so the advantage of sinistral stairs is dwindling.
The spiral staircase accommodates a long path within a small space, reduc-
ing the required slope to reach a certain height. The Cave of Swallows (Sotano

© Springer International Publishing AG 2016 65


Ø. Hammer, The Perfect Shape, DOI 10.1007/978-3-319-47373-4_15
66 The Perfect Shape

Fig. 15.1 The upper part of the Sotano de las Golondrinas cave shaft, seen from
the cave floor 333 m below the surface

de las Golondrinas) in Mexico is an interesting natural analogue. The deepest


single cave shaft in the world, it is shaped like a 333 m deep bottle, 300 m long
by 130 m wide in the lower part but narrowing to only a 50 m diameter
opening at the surface (Fig. 15.1). Staring down into this abyss is not for the
faint hearted. The cave got its name from the countless thousands of white-
collared swifts living there. Their morning exit is one of the World’s most
spectacular natural wonders. Unable to fly straight up, the birds are forced into
a helical ascent. A swirling mass of swifts is slowly rising, with groups of about
50 shooting out of the opening at a time, desperate to avoid the waiting birds
of prey. Clearly, the birds must agree on a handedness for the helix, or they
would collide with each other. It is another case of symmetry breaking. When I
was there the helix was right-handed (the birds circling counter-clockwise as
seen from above), and other visitors seem to always report the same
handedness.
Looking at a helical staircase from below or above produces a dizzying
spiral image, used to full effect in Hitchcock’s Vertigo. The importance of the
spiral in this movie is reflected in the official poster by Saul Bass (Fig. 15.2)
and the animations in the opening sequence. These famous designs were
made by a machine called a harmonograph, which we will get back to later in
the book.
The central-point projection of a helical staircase (Fig. 15.3) is reminiscent
of a logarithmic spiral. The same spirals appear when looking down the barrel
15 The Case of the Staircase 67

Fig. 15.2 Poster for Hitchcock’s Vertigo (Paramount Pictures Corporation 1958)

of a rifle (Fig. 12.3). But is there not something inelegant about it: the inner
coils do not seem to approach the center at quite a sufficient rate? Let us have a
closer look.
The apparent image size R of an object halves as the distance (i.e., height
above the viewer) d doubles. In other words, with an appropriate scaling the
image radius from the center of the spiral to the handrail should go as R(d) ¼
1/d. The helical shape means that the height should increase linearly with
rotational angle measured from the level of the viewer: d ¼ kφ, where k is the
pitch of the screw. Putting this together, we get R(φ) ¼ 1/kφ. When the
68 The Perfect Shape

Fig. 15.3 Left: Staircase in the Department of Physics, University of Oslo, built
1935. Photo Marte H. Jørgensen. Right: Hyperbolic spiral

handrail thus reaches the level of the viewer, the apparent length of the radius
goes to infinity. This curve is known as a hyperbolic spiral. So, alas, not a
logarithmic one this time.
16
The Spiral of the Ancient Mariner

If you sail in a constant, oblique direction with respect to north, you follow a
so-called rhumb line. You would be excused to think that you were following a
straight line, which would bring you to your target in the shortest possible
time. But then you would have forgotten that the Earth is round.
From ancient times, maps were drawn with a grid of crisscrossing rhumb
lines (Fig. 16.1). Going in a certain compass direction, you would be following
one of these lines (or a parallel to it), safely bringing you to your destination at
the end of the line. Or that was the idea, at least. In late medieval and early
Renaissance times, such maps (called portolans) were drawn directly from
navigational data giving directions and distances. The mathematical implica-
tions were not much discussed until the sixteenth century. It turned out that
you cannot project the spherical surface of the Earth onto a two-dimensional
map in such a way that you preserve both directions and distances accurately.
One, or both, has to yield. In Mercator’s famous map projection, presented in
1569, the directions are given priority so that rhumb lines are indeed straight
lines, at the cost of a ridiculous inflation of distances and areas at high latitudes.
What shape, then, does a rhumb line really have on the sphere? It is certainly
not a straight line. Instead, it is a beautiful spiral, the loxodrome (Fig. 16.2).
Close to the poles, it approximates to a logarithmic spiral. This has the curious
implication that if you start from Alaska and walk or swim in a constant
northeast direction, you would end up walking around the pole an infinite
number of times, but still get there in finite time (this is assuming that you
walk with constant speed, notwithstanding that your spin rate would go to
infinity).

© Springer International Publishing AG 2016 69


Ø. Hammer, The Perfect Shape, DOI 10.1007/978-3-319-47373-4_16
70 The Perfect Shape

Fig. 16.1 Map of the Caribbean, Jakobsz Theunis, 1666

Fig. 16.2 Left: Loxodrome. All angles with the latitudes are equal. Right:
M.C. Escher’s Sphere surface with fish, 1958. © 2016 The M.C. Escher Company—
The Netherlands. All rights reserved. www.mcescher.com
16 The Spiral of the Ancient Mariner 71

Fig. 16.3 Clockwise hair whorls. Left: Grid of line segments, all drawn 60 from the
radius vector. Program code in Appendix B. Right: My son Eiel 1 week after birth,
viewed from above

Clearly, the loxodrome cannot be the fastest route between two points on
the globe. This is the property of another curve, the great circle. Following the
great circle, as you do when you incongruously pass over Greenland (of all
places) flying from Europe to the US, means you have to constantly adjust
your bearing, not the simplest of tasks before computers became common at
sea and in the air.
If we cover the plane with short line segments all oriented at the same angle
to the radius vector, a whorl pattern emerges (Fig. 16.3). All the lines are
tangent to logarithmic spirals because of the equiangular property, and the eye
traces out these spirals by “connecting the dots”. We can do the same thing on
a hemisphere such as the top of a human head, covering it with hairs all
oriented at constant bearing with respect to a pole, giving the impression of a
set of packed loxodromes. This is called a hair whorl, and most people have one
(some even have two). Hair whorls are easiest to see in infants and when the
hair is cut short. For successful hair whorl spotting, I recommend a queue or a
bus, where the heads in front of you can be studied discretely at short distance.
The pole is usually close to the top of the head, or slightly behind or off to the
side.
The direction, or chirality, of human hair whorls has received a lot of
scientific attention, but with mixed results and there is currently not much
consensus on the matter. It is clear that the clockwise direction (when moving
outwards) is more common than the counterclockwise direction, with num-
bers varying across studies and populations from 51 % in Japan (Klar 2003) to
more than 90 % in the United States (Klar 2009). However, whether or how
the direction is inherited is unclear. McDonald (2011) gives a good overview
72 The Perfect Shape

of this strange research field, with many references. Some papers (Klar 2003)
have claimed a genetic link between hair whorl direction and handedness, with
left-handed people tending towards counterclockwise whorls, but this has not
been supported by later studies. Moreover, Klar (2004) secretly studied hair
whorls on a gay beach in Delaware (imagine that!), finding a much higher
incidence of the counterclockwise direction there than in the general popula-
tion, but again this result has proven difficult to reproduce.
17
Gnomons, a Miracle, and Charles Babbage

A logarithmic spiral shell has the great advantage that while each growth
increment is larger than the previous one, it has the same shape, and the
whole shell also maintains its shape during growth. The Nautilus animal has
the luxury of being able to live in a body chamber (the outermost chamber)
that never changes form, only increases in size. This idea was of great impor-
tance to the Greek mathematicians, through the idea of the gnomon.
Thomas Heath, in his 1925 commentary to Euclid’s Elements, traces the
word gnomon through ancient history. Passages from Herodotus and Suidas
indicate that Anaximander originally introduced the gnomon from Babylonia
to Greece in the sixth century BC, meaning the vertical staff used in a sundial.
The term slowly extended to describe the L-shaped tool (steel square) used to
draw a right angle.
If you take such an L shape away from a square, or add it to a square, you are
still left with a square (Fig. 17.1). Aristotle puts it quite clearly, among all his
unhelpful classifications in the Categories:

But there are some things which undergo increase but yet not alteration. The
square, for instance, if a gnomon is applied to it, undergoes increase but not
alteration, and so it is with all other figures of this sort.

Now it starts to get interesting.


Probably, already the Pythagoreans, in the fifth century BC, had observed
that the square numbers (1, 4, 9, 16 etc.) could be produced by successive
addition of odd numbers (1, 3, 5, 7 etc.):

© Springer International Publishing AG 2016 73


Ø. Hammer, The Perfect Shape, DOI 10.1007/978-3-319-47373-4_17
74 The Perfect Shape

Fig. 17.1 Left: A yellow square and its white gnomon. Right: A parallelogram and
its gnomon

Fig. 17.2 Odd numbers as gnomons to the square numbers

0þ1¼1
1þ3¼4
4þ5¼9
9 þ 7 ¼ 16
...

The Pythagoreans had made the connection between this arithmetic property
and the geometric property of gnomonic addition to squares, and therefore
referred to the odd numbers as gnomons with respect to the square numbers.
In Fig. 17.2, there is one blue dot. 1 is the first square number. Adding the
3 red dots, you get the second square number, which is 4. Adding the 5 green
dots, you get the third square number, which is 9, etc.
17 Gnomons, a Miracle, and Charles Babbage 75

Euclid generalized this concept to parallelograms. His second definition in


Book II reads:

Παντoζ δε παραλληλoγραμμoυ χωριoυ τϖν περι την διαμερρoν αυτoυ


παραλληλoγραμμαν εν oπoιoνoν συν τoιζ δυσι παραπληρωμασι γνωμων
καλεισθω.
And in any parallelogramic area let any one whatever of the parallelograms
about its diameter with the two complements be called a gnomon.

This definition is rather incomprehensible to a modern reader, but it


describes the generalization of a right-angled gnomon of a square to a skewed
L-shape as the gnomon of a parallelogram (Fig. 17.1).
Hero of Alexandria generalized even further, and defined a gnomon to mean
any shape that, when added to another shape, preserves the original shape. He
also mentions gnomonic numbers, such as the odd numbers being the gno-
mons of squares. (This whole story of successive addition of elements to a
concept, while keeping the general idea, is itself very gnomonish!)
Theon of Smyrna, ca. 100 AD, discussed gnomonic numbers with respect
to the square, but also the triangle and other polygons. The triangular numbers
are obtained by adding the natural numbers 1, 2, 3, . . . as gnomons to a single
point (Fig. 17.3).
The triangular numbers are therefore 1, 3, 6, 10, . . ., or in general n(n + 1)/2.
This was perhaps the equation used by the young Carl Friedrich Gauss,
when, according to legend, he astonished his teacher by seemingly adding
the numbers from 1 to 100 within seconds: 1 + 2 + 3 + . . . + 100 ¼ 100 
(100 + 1)/2 ¼ 5050.

Fig. 17.3 Natural numbers as gnomons to the triangular numbers


76 The Perfect Shape

It is time to consult the Holy Scripture. In John, 21:4–11 we read:

Early in the morning, Jesus stood on the shore, but the disciples did not realize
that it was Jesus. He called out to them, “Friends, haven’t you any fish?”
“No,” they answered.
He said, “Throw your net on the right side of the boat and you will find
some.” When they did, they were unable to haul the net in because of the large
number of fish.
. . . Jesus said to them, “Bring some of the fish you have just caught.” So
Simon Peter climbed back into the boat and dragged the net ashore. It was full of
large fish, 153, but even with so many the net was not torn.

Now the question is, of course, why did the Evangelist so specifically state
that there were 153 fishes? It seems an odd number in more than one way.
St. Augustine, around 400 AD, thought he had the answer: 153 is the 17th
triangular number (n ¼ 17 in the equation above), 17 being the sum of 7 (the
number of spiritual gifts), and 10 (the number of commandments)! The idea
demonstrates the level of mathematical knowledge at the time, but the
problem with such numerology is of course that any number has some
interesting mathematical property.
So, we can generate a table of the triangular numbers n(n + 1)/2 by
successive addition of the natural numbers as gnomons, and we can make a
table of n2 by addition of the odd numbers. We can also construct a second-
order gnomonic sequence—gnomons upon gnomons. For example, let us say
that we want to produce a table of the cubic numbers, n3. The sequence starts
as 13 ¼ 1; 23 ¼ 8; 33 ¼ 27; 43 ¼ 64; 53 ¼ 125. The differences between
successive cubic numbers are then 7, 19, 37, 61, etc. The differences between
these successive differences (second-order differences) are 12, 18, 24, etc.
The differences between these are a constant, 6. We can show these gener-
ations of differences as follows:

Hence, to produce our table of cubic numbers we can work this procedure
in reverse. We start with 12 and use the number 6 (the third-order difference)
as a gnomon, giving the sequence
17 Gnomons, a Miracle, and Charles Babbage 77

12 þ 6 ¼ 18
18 þ 6 ¼ 24
24 þ 6 ¼ 30
...

These are our required second-order differences. Then we use these numbers
as gnomons again, starting from 7:
7 þ 12 ¼ 19
19 þ 18 ¼ 37
37 þ 24 ¼ 61
...

It can be shown that the nth element in this sequence is 3n2 + 3n + 1. Then we
use this sequence as gnomons yet again. Starting with 1, we get
1þ7¼8
8 þ 19 ¼ 27
27 þ 37 ¼ 64
...

which are the required cubic numbers n3.


In fact, we can produce any kth degree polynomial by the successive
application of k gnomonic sequences, the first of which is always a constant
(6 in our example). And, since any well-behaved mathematical function can be
approximated by a polynomial, e.g., by Taylor expansion, it follows that the
method of gnomons can be used to generate any mathematical table.
This method is usually called the method of differences. It was used by the
mechanical calculation machine known as the Difference Engine, proposed by
Charles Babbage in 1822. The idea had occurred to him after hours of tedious
work together with the astronomer John Herschel, checking mathematical
tables made by underpaid human computers:

After a time many discrepancies occurred, and at one point these discordances
were so numerous that I exclaimed, “I wish to God these calculations had been
executed by steam,” to which Herschel replied, “It is quite possible.”
78 The Perfect Shape

Fig. 17.4 Difference Engine built after Babbage’s plans. Science Museum, London.
Photo Geni, Creative Commons Attribution-Share Alike 4.0 International licence

Before electronic calculating devices, mathematical tables were indispens-


able everywhere in science and engineering, and also for astronomical naviga-
tion (no, they did not have GPS). Errors in such tables could make bridges fall
down or ships sail into cliffs, and Babbage managed to raise a large sum of
money for the construction of his Difference Engine (Fig. 17.4). Unfortu-
nately, although the mathematical concept was sound, a full-sized, working
difference engine was never realized by Babbage. However, the Difference
Engine led him to the idea of the Analytical Engine, the precursor of the
modern programmable computer.
Through the ancient idea of the gnomon, the molluscan shell connects to
the heart of mathematics. But I fear the snails don’t care about that.
18
Curls of Green

In the summer of 1864, 5 years after the publication of The Origin of Species,
Charles Darwin was the most celebrated scientist in the world. Any lesser man
would rest on his laurels, but Darwin had retreated to continue with basic
research, the kind of unglamorous but indispensable science that is rarely
funded by research councils. Among all the wonders of Nature, he had chosen
to focus on the behavior of climbing plants. Bryonies and bellflowers, beans
and bindweeds, bushwillows and birthworts, hops, hoyas, honeysuckles,
ceropegias, cucumbers and climbing ferns, guinea flowers, glorybowers, morn-
ing glories, wisterias, leadworts, jasmines, almost every twining vine or liana
known to Victorian England filled all corners and snaked up every wall of his
house. The results were published in one of Darwin’s many less well-known
books, On the Movements and Habits of Climbing Plants (Darwin 1875).
A twining plant grows in a helix around its host (Fig. 18.1). Darwin noted
that most species have a definite handedness, meaning that all individuals
within one species wind the same way. He listed the handedness of 42 twining
species, and found that only 11 were left-handed, the rest were right-handed.
The probability of such a skewed ratio happening by chance is about one in
500.
A possible functional explanation for the definite handedness in twining
plants is suggested by the picture below of several bindweeds climbing on the
same host stem. It works well because they all wind the same way. If some were
left-handed and some were right-handed, they would cross and rub against
each other. Mountaineers learn that crossing ropes can be deadly—ropes in
tension can cut through each other quite easily. Perhaps this is another

© Springer International Publishing AG 2016 79


Ø. Hammer, The Perfect Shape, DOI 10.1007/978-3-319-47373-4_18
80 The Perfect Shape

Fig. 18.1 Left: Bindweed, Convolvulus arvensis, right-handed helix on grass. Mid-
dle: Bindweed bundle. Photo Marte H. Jørgensen. Right: Honeysuckle, Lonicera
periclymenum, left-handed helix on tree. Photo Marte H. Jørgensen

example of a symmetry breaking process, like the one responsible for the
definite coiling direction in snail shells.
Darwin also noted that the helical habit of many climbing plants results
from a simple mechanism called circumnutation. The hop or bindweed does
not feel the presence of the host stem in order to curl around it. Instead, the
growing tip just moves around in a circle, even when floating in free air before
having found a host. As soon as the tip hits another stem by chance, this
circular movement will ensure that the climber wraps around the host stem
like a constricting snake. It is not particularly intelligent, but it works.
Scientists are still debating the mechanism for circumnutation. One theory
held that gravity is required. In 1983, a spectacular experiment was carried out
to test this idea. On the sixth flight of the space shuttle Columbia, sunflower
seedlings were grown in microgravity. Sunflowers circumnutate on Earth, and
so they did in space, showing that gravity is not a requirement. Columbia
would make another 27 missions before she broke up on re-entry in 2003,
killing seven astronauts in one of the most heart-breaking tragedies in the
history of space travel.
Curly plants have long been design winners. A Mediterranean plant called
the Acanthus, with its fan-like, deeply incised leaves, was popular in Greek and
Roman art and architecture, most famous for decorating Corinthian columns.
As centuries passed, the acanthus design departed more and more from natural
realism, and evolved into the “acanthus scroll”, now a general term for any
18 Curls of Green 81

Fig. 18.2 Acanthus scroll from the Rococo. Alexis Peyrotte (1699–1769). Right:
Detail from harp, France, ca. 1770

Fig. 18.3 Left: Climbing plants on arabesque iron balustrade, late eighteenth
century, Bogstad manor, Norway. Photo Marte H. Jørgensen. Right: Detail from
iron gate, late seventeenth century, Germany

vaguely vegetation-looking curved and whorled design (Fig. 18.2). The spiral
shapes may have been inspired by dead, dried-up Acanthus leaves. Acanthus
scrolls are ubiquitous in Roman murals, medieval illuminated manuscripts, in
Rococo furniture and, most elegantly of all, in Art Nouveau. From the
sixteenth century, such curly flower patterns were also called arabesques.
Right now, the arabesque style is not in vogue among architects and fashion
designers, but it will come back. When we go to the stars, our ships may not
look like a twentieth century toilet inside, as we are used to from the movies,
but decorated with patterns of swirling leaves and coiling stems (Fig. 18.3).
Several varieties of the common ornamental houseplant Begonia rex make a
curious and decorative spiral twist on their leaves (Fig. 18.4). The most specta-
cular one is aptly named the “escargot begonia”. The mechanism must be
accelerated growth in the tangential direction along the edge of the base of the
leaf. Similar, but less extreme spirals are seen in the coltsfoot (Fig. 18.4), the
lower leaves of the burdock (Arctium) and many other plants.
82 The Perfect Shape

Fig. 18.4 Top left: Fern fiddlehead, Borneo. Photo Robert David Siegel, M.D.,
Ph.D., Stanford University. Top middle: Begonia rex leaf. Photo Marte
H. Jørgensen. Top right: Coltsfoot leaf, Tussilago farfara. Bottom left: Autumn
lady’s tresses, Spiranthes spiralis. Photo Marinella Zepigi, Acta Plantarum. Bottom
middle: Heliotropium indicum. Photo Alexey Sergeev. Bottom right: Cucumber
tendril, Cucumis sativus. Photo Marte H. Jørgensen

Among all spiral forms in plants, the most beautiful are perhaps the
“fiddleheads” formed by the coiled-up fronds of young ferns (Fig. 18.4).
They are usually approximately logarithmic spirals. Because of the “fractal”
shape of ferns, with lateral branches looking just like smaller versions of the
whole leaf, the fiddlehead can sometimes contain a series of smaller fiddleheads
inside it, producing a truly mesmerizing figure. Another beautiful feature of
some fiddleheads is the recurving shape of the stem below the spiral, producing
an elegant, smooth curve. We will later discuss a similar mathematical con-
struct, the lituus.
The tendrils of cucumbers are also smashingly photogenic. Gerbode et al.
(2012) studied helical cucumber tendrils in great detail, including the typical
reversal of coiling direction in the middle of the tendril. Ever since Darwin,
this phenomenon has been explained as a mechanism for ensuring that the
18 Curls of Green 83

twisting sums to zero. Cucumber tendrils usually start to coil only after touch-
ing a target (thigmotropic coiling). I am not sure how the gorgeous free-
standing, planispiral tendrils as shown in Fig. 18.4 form; perhaps they
represent misunderstandings by the cucumber, initiating coiling by slipping
off the target or by the touch of an insect.
19
The Pendulum and the Galaxy

Consider a pendulum swinging both in the x and the y directions. Because of


friction, the movement is dampened with time. As shown in Appendix A.4,
the path of such a pendulum is a “squashed” logarithmic spiral, with each
whorl looking a bit like an ellipse (Fig. 19.1).
In Chap. 15, we mentioned that the spirals in Hitchcock’s Vertigo were
made with a harmonograph. This interesting machine, invented in the mid
1800s, exists in many different versions. The simplest harmonograph consists
of a pendulum moving in the x direction, at the bottom of which is suspended
another pendulum moving in the y direction, drawing a pattern on a surface.
An alternative version has a single pendulum swinging in one plane, and the
drawing surface itself is swinging in the perpendicular direction. This will
produce so-called Lissajous patterns, well known to electronics engineers
playing with two oscillators connected to an x–y oscilloscope. When the
pendulums reduce their amplitudes over time because of friction, the result
is a damped Lissajous pattern. In contrast with the single pendulum shown
above, the harmonograph with two pendulums can have independent fre-
quencies in the x and y directions. This allows much more complex patterns
than just logarithmic spirals.
For the production of the Vertigo spirals, it seems a single pendulum was
used, swinging freely in both the x and the y plane. Disregarding frictional
decay, such a pendulum can produce a circular or elliptical motion. In
addition, the drawing surface was rotating. The rotation of the surface caused
the major axis of the ellipse to rotate over time (the Earth’s orbit around the
sun makes a similar movement called perihelion precession). Finally, the

© Springer International Publishing AG 2016 85


Ø. Hammer, The Perfect Shape, DOI 10.1007/978-3-319-47373-4_19
86 The Perfect Shape

Fig. 19.1 The path of a dampened pendulum

exponential decay of the pendulum made the ellipses decrease in size over
time, similarly to the squashed logarithmic spiral as we saw above. These
motions all combined to make a complex, beautiful curve. In this curve
there appeared bands of intersecting paths, shaped as two logarithmic spirals.
A pendulum swinging over a rotating plane is also the essence of the
Foucault pendulum. In 1851, Léon Foucault placed a 28 kg lead sphere at
the end of a 67 m long wire suspended from the dome of the Panthéon
building in Paris to illustrate the rotation of the Earth. At the latitude of Paris,
the swing plane of the pendulum rotated a full 360 every 32.7 h. There are
now many such pendulums in universities and science centers all over the
world. The geometry of the experiment becomes much simpler if the pendu-
lum is placed on the North or South Pole, and some hardy people at the Scott-
Amundsen base on the South Pole did exactly that in 2001. It made a full
circle about once every 24 h, as it should. Now instead of letting the pendulum
swing in a plane, give it a slight sideways push so it swings in a decaying ellipse.
The setup is now very much like the Vertigo harmonograph and the Foucault
pendulum describes a similar complex pattern.
I spent some time trying to simulate the Vertigo harmonograph on the
computer. The period of the pendulum, the amplitudes, phases, decay con-
stant and rotation speed must be chosen extremely precisely to produce a
desired pattern. There are simply too many parameters. But Fig. 19.2 is fairly
close to the Vertigo poster.
If we change the parameters slightly, we stop the paths from crossing, but
the double logarithmic spiral is still evident as bands of higher density
(Fig. 19.3).
In this case, the path is a continuous curve, but the visual appearance would
have been very similar if we had used a set of nested, rotated ellipses instead.
19 The Pendulum and the Galaxy 87

Fig. 19.2 Computer simulation of the Vertigo harmonograph image. Program


code in Appendix B

Fig. 19.3 Harmonograph image with density waves

Now imagine that each such ellipse is the orbit of a star around its galactic
center, and each ellipse is slowly rotating by precession. As in the figure, this
will produce zones of higher density of stars, shaped like logarithmic spirals.
These density waves are ghostly, dynamic structures—the stars pass straight
through them in their vast orbits—but still they are bright enough to define
88 The Perfect Shape

the arms of a spiral galaxy. Or that is the current theory anyway (e.g., Francis
and Anderson 2009).
So I think it is fair to say that in 1958, the graphic designers for Vertigo made
a working model for the formation of spiral galaxies, which is not a trivial
thing.
20
How to Grab a Can of Beer

Close your hand and fold your fingers in; then unfold as if you are releasing a
butterfly. It is a beautiful, elegant and smooth gesture. What path does your
fingertip follow? Experts seem to agree it is approaching a logarithmic spiral
(e.g., Kamper et al. 2003), but the devil is in the details.
The question is also relevant for other jointed structures, such as robotic
arms. Can we calculate the robot’s path to avoid collisions with obstacles? And
what, if any, is the connection with Ptolemaic cosmology?
The situation is sketched in Fig. 20.1. The longest, innermost part of the
finger is called the proximal phalanx, with length LA. As the finger uncoils, the
innermost joint (i.e., between hand and proximal phalanx) is opening with an
angle φA such that the joint at the outer end of the proximal phalanx moves in
a circle A with radius LA. Simultaneously, this outer joint is opening with an
angle φB such that the intermediate phalanx forms the radius vector with
length LB of a circle B. In effect, the movement on circle B is an epicycle on the
larger circle A. Finally, and still simultaneously, the outermost joint is opening
with an angle φC such that the distal phalanx forms the radius vector with
length LC of a circle C—a second order epicycle. In this way, your fingertip
moves as the celestial bodies of the Ptolemaic system—each planet, the Sun
and the Moon flying on a small circle, this circle itself gliding stately in a larger
circle centered on a point near the Earth.
The relative lengths of the phalanges is a matter of debate, some authors
have reported Fibonacci numbers here, while others disagree. The index finger
ratios LA/LC ¼ 2.5 and LB/LC ¼ 1.4 given by Buryanov and Kotiuk (2010) are
probably as good as any.

© Springer International Publishing AG 2016 89


Ø. Hammer, The Perfect Shape, DOI 10.1007/978-3-319-47373-4_20
90 The Perfect Shape

Circle B

Circle A Circle C

LA
LB

1
LC

Fig. 20.1 An uncoiling finger in positions 1, 2 and 3. Assuming equal angular


opening rates of the three joints, the fingertip will follow the black spiral path. In
position 2, the fingertip follows a small circular path C centered on the outermost
joint. The outermost joint itself follows a larger circular path B centered on the
intermediate joint, which again moves in the circle A

By aligning the coordinate system appropriately, the position of the finger-


tip is then
x ¼ LA sin ðφA Þ þ LB sin ðφA þ φB Þ þ LC sin ðφA þ φB þ φC Þ
y ¼ LA cos ðφA Þ þ LB cos ðφA þ φB Þ þ LC cos ðφA þ φB þ φC Þ:

In the figure above, I have assumed that all three joints start at right angles, and
then expand at equal rates until completely extended at 180 . This results in
the harmonic series
x ¼ LA sin φ þ LB sin 2φ þ LC sin 3φ
y ¼ LA cos φ þ LB cos 2φ þ LC cos 3φ:

It is difficult to see how we can connect these equations with a logarithmic


spiral. Still, it is intriguing that measurements of human finger paths seem to
show good fits to logarithmic spirals. Clearly, we have some freedom in how
the three joint angle rates are controlled. The angles need not increase at the
20 How to Grab a Can of Beer 91

Fig. 20.2 If the hand forms a circular arch at all times when opening, what is the
path of the fingertip? Reprinted from Littler (1973), with permission from Elsevier

same rate, nor do the ratios of rates need to be constant over time. If we wanted
to, we could optimize these parameters to produce an approximation to a
logarithmic spiral, as observed.
There is another, interesting way to approach this problem. Littler (1973)
speculated that a logarithmic spiral path could result from a functional
constraint: What would happen if the closing hand, at all times, attempts to
form a circular arch? This would make sense if the hand is constructed to grab
cylindrical objects, such as branches when climbing a tree. Or to hold a beer
can (an evolutionary biologist might say that the hand was pre-adapted for that
purpose). Could such a constraint result in a logarithmic spiral path for the
fingertip? Littler did not work out the mathematics, but he did make a
beautiful figure (Fig. 20.2).
Let us, for the sake of argument, simplify the geometry by assuming that the
finger is not jointed, but forms a continuous circular arch with increasing
radius r. We place the base of the hand, of length π, at the origin (0, 0), and let
the fully extended hand point vertically up (Fig. 20.3). Because the base of the
hand is kept fixed, the center of the circle is moving towards the left as the
radius is increasing.
92 The Perfect Shape

Path of (0, π)
finger tip

r=∞

r=2

r=1

(-2, 0) (0, 0)

Fig. 20.3 Model for the opening of the hand. The abstracted hand (thick black
line) keeps forming a circular arch while opening, here shown in three positions

It is fairly easy to calculate (Appendix A.4) that the tip of the finger, with
coordinates (x, y), follows a spiral-shaped curve as a function of a polar angle φ:
π 1
x¼ cos φ 
φ φ
π
y ¼ sin φ
φ

Hence, the curve is a variant of a hyperbolic spiral, but with the pole
(at φ ! 1) moving along the x axis.
The resulting spiral path is certainly not a logarithmic spiral judging from
the equations, but it does look a bit like one. If we try to adjust the parameters
of a logarithmic spiral to this theoretical path by computer, it turns out that we
can get an extremely good, but not perfect fit.
So it’s all a bit unsatisfactory. We can say, by looking at the equations with
epicycles, that the jointed finger cannot follow a logarithmic spiral exactly.
Also, an abstracted, continuous finger forming a circular arch in order to grasp
cylindrical objects will follow a sort of continuously translating hyperbolic
spiral that approximates to a logarithmic spiral numerically.
And that is about as far as I got on the matter.
21
An Interlude at the Beach

Beaches are dynamic systems. Sand is continuously on the move, driven by


waves, currents and wind. So how is it possible that these rivers of sand
maintain their shapes? Like so many physical systems, a beach tends towards
an equilibrium shape, where removal and deposition of sand balance each
other at every point. The system is terribly complicated, but let us for the sake
of illustration consider a minimal and simplistic model. In this model, sand is
transported by wave action alone; waves always come from the same oblique
direction; water depth outside the beach is constant so there is no refraction;
and there is no total loss or gain of sand (the mass is conserved). Into this
idealized world, we place the sunbather’s dream: a straight, infinitely long
beach. This beach will stay there forever, and it will stay straight. The waves
will transport sand along the beach, but at any position, the gain from
upstream balances the loss downstream. To explain informally why the
shape is stable, we can use the old physicist’s trick of considering the effects
of a small perturbation. What will happen if we make a little notch in the
beach? Sand will get trapped there, and the notch will fill up. Conversely, a
little protrusion will quickly erode.
Now introduce a thin, long rocky headland sticking out from the beach.
The waves will diffract at the tip of the headland, and fan out in all directions
(Fig. 21.1). Down shore from the headland, the waves will now meet the beach
at continuously changing angles. This is not a stable configuration, and the
system will start to evolve towards a steady state. Can we predict the equilib-
rium shape of the beach? What is required is that the waves now meet the

© Springer International Publishing AG 2016 93


Ø. Hammer, The Perfect Shape, DOI 10.1007/978-3-319-47373-4_21
94 The Perfect Shape

Fig. 21.1 A wave train from the left hits a barrier. The diffracted waves hit the
logarithmic spiral beach (yellow) with a constant angle

Fig. 21.2 Half Moon Bay, California, from Google Earth

beach at the same angle everywhere—otherwise the erosion would be uneven


and the shape would be changing.
The same angle everywhere—it’s the equiangular spiral again!
And yes, beaches down shore from headlands and wave breakers do seem to
form logarithmic spirals (Fig. 21.2). This was first observed by Krumbein in
1944 and has since been confirmed and discussed by many scientists and
coastal engineers, e.g., LeBlond (1979), although in recent years other geo-
metric models have also become popular.
22
When Television Was Spiral

John Logie Baird, the inventor of television, was a busy man. Born in 1888 in
Scotland, his early career was packed with absurd business projects. At the age
of 12 he supplied his home with electricity using a water wheel and home-
made batteries, causing permanent acid damage to his hands. In the same year
he made a telephone system with wires across the street, nearly cutting the
head off a horseman. During World War I he developed a medicine for
hemorrhoids, started the Baird Undersock Company, traded shoe polish,
chocolate and cigarettes, became a socialist, and blew up the Glasgow electrical
grid in an attempt to make diamonds. In 1919 he went to Trinidad, set up a
mango jam business that failed spectacularly, and caught a number of diseases
including dysentery and malaria on top of his several eye diseases and other
ailments. In 1920 he went back to London where his business turned to
honey, anti-wrinkle cream and soap. But despite all this nonsense, Baird was in
fact a very clever man. He decided to learn electronics, an area of endless
possibilities after the invention of the vacuum tube and radio transmission. His
mastery of electronics would change the world. By 1925, several inventors
were independently making crude attempts at transmitting live images, but it
was Baird who first succeeded in making a practical, semi-commercial system.
Baird’s televisor was an electromechanical device based on a principle for
image scanning invented by the German Paul Nipkow in 1884. The Nipkow
disk was simply a large, round metal plate perforated by holes in a spiral
pattern (Fig. 22.1). A powerful lamp was placed behind the disk. In Baird’s
first televisors, there were 30 holes (Fig. 22.2). The viewing area was a small
trapezoid window, sometimes enlarged using a magnifying glass, placed in

© Springer International Publishing AG 2016 95


Ø. Hammer, The Perfect Shape, DOI 10.1007/978-3-319-47373-4_22
96 The Perfect Shape

Fig. 22.1 A Nipkow disk with 30 scan lines, as used in Baird’s first televisors. The
50 cm diameter metal disk is perforated by 30 holes arranged in an Archimedes
spiral. Red and grey lines are only for reference. In this version, the disk is spinning
clockwise and the small viewing area (dark grey) is placed on the upper side of the
disk

front of the disk, with an angular width of 360/30 ¼ 12 . As the disk was
spinning, first the outermost hole in the spiral would pass across the top of the
window, tracing out a line of light called a scan line. Then the second hole
would enter the window one scan line further down, and so on. The 30 holes
made a total of 30 scan lines. At the same time the lamp would be flickering
very rapidly according to the incoming wireless signal, varying the intensity of
the light in synchrony with the rotation and thus building up an image. The
disk rotated five times a second, producing a new image frame every 0.2 s.
A similar Nipkow disk was operating in the broadcasting studio, but there
an even stronger lamp was shining steadily behind the disk. With the aid of
focusing lenses, this produced a spot of light flying across the singer or other
subject, scanning it line by line. A bank of selenium photocells recorded the
reflected light. The signal was then amplified and sent to the radio transmitter.
22 When Television Was Spiral 97

Fig. 22.2 A Baird televisor from 1928 in the collection of the Norwegian Museum
of Science and Technology. Rear view. Lamp and viewing window on the left. Note
the tiny holes near the edge of the disk. Image courtesy of the Norwegian Museum
of Science and Technology

The recording had to be carried out in a darkened room in order to obtain


sufficient contrast between the reflected spotlight and the ambient light.
The image quality in Baird’s first televisors was appalling. The image was
tiny and monochromatic, and the resolution of 30 scan lines was vastly
insufficient. For comparison the PAL system, used until recently for television
broadcasting in most of the world, used 625 scan lines, and even that did not
give very impressive quality. The image rate of five frames per second was also
far too low for human perception. Finally, the synchronization of the signal
with the rotation of the disk was poor, causing the image to roll and shimmer.
Baird developed his invention quickly, and although his televisor never
reached the mass market it was a considerable success, receiving great acclaim
from the general public, from the broadcasting industry and from fellow
inventors and scientists. He would remain an influential television engineer
until his death in 1946. In 1926 he increased the frame rate to 12.5 Hz,
98 The Perfect Shape

reducing the jerkiness to an acceptable level. In 1928 he made the first


transatlantic television transmission, demonstrated the first color television,
and even experimented with 3D television. An entry in the New York Times,
February 9th, 1928, described the transmission of images across the Atlantic in
blooming language:

When the televisor, a black box compact enough to be carried around in a taxi,
had done its work with this rhythmic rumble from across the sea, the visions
gradually built themselves up of tiny oblongs of light suspended in a whirling
rectangle of brilliance in the machine’s gaping mouth.

The number of scan lines was steadily increased, reaching a maximum of


240. In addition, Baird invented a hybrid camera system where a traditional
photographic camera recorded the pictures onto film which was developed
immediately and automatically, and then the film was scanned directly by a
Nipkow disk. This contraption allowed filming in broad daylight.
Despite all these improvements, by the mid-1930s it was becoming clear
that Baird’s mechanical television was not the future. New all-electronic
television sets based on the cathode ray tube were more reliable, less noisy,
and produced larger images of better quality. Still, the principle of building up
the picture with a dot of light flying along scan lines remained fundamental to
analog television technology until the 1990s, only now being replaced by the
complicated, pixel-based image coding of digital TV.
But for a few years television depended on the old spiral of Archimedes,
mesmerizing the world with magical, tiny, flickering, orange images of vaude-
ville singers, Charleston dancers and Shakespeare actors.
23
Thou Shalt Love Thy Neighbour

Consider four amorous mice in the corners of a square. Each mouse has fallen
in love with its neighbor in the counterclockwise direction, and starts running
towards it. It is a sad story of unreturned affection. What will be the paths of
the mice? This interesting little exercise has become popular in educational
mathematics, and it is variously known as the mouse problem, the bug
problem, the turtle problem or the dog problem, depending on the favorite
animal of the author (e.g., Lucas 1877; Gardner 1965; Nahin 2012). A less
nice variant is that of four homing missiles or pirate ships chasing each other.
The problem can be generalized to 3, 5 or any number of mice placed in the
corners of a regular polygon (for an odd number, one mouse must be
homosexual) (Fig. 23.1).
A proper analysis of the problem requires some calculus, not too difficult
but a little cumbersome, so we will skip it here. But it should not come as a
great surprise that the paths are logarithmic spirals. Since the four mice are in a
square configuration to begin with, and since the problem is quite symmetri-
cal, we might expect that they stay in a square, but rotating and shrinking as
the little rodents converge. Such a sequence of rotating and shrinking polygons
is sometimes called a whirl. And if the paths continue to be tangent to these
polygons, we might expect to see equiangular spirals. Thus, from the figures,
the radius vector seems to bisect the angles of the polygons. The angle between
the path and the radius vector should then be 45 for the four-mice problem,
giving an equiangular spiral with expansion coefficient k ¼ cotan 45 ¼ 1, i.e.,
the circular tessellation spiral; the angle is 30 for the three-mice problem; and

© Springer International Publishing AG 2016 99


Ø. Hammer, The Perfect Shape, DOI 10.1007/978-3-319-47373-4_23
100 The Perfect Shape

Fig. 23.1 Top left: Three mice starting in the corners of an equilateral triangle,
chasing their neighbors. Red triangles connect positions at equal time steps. Top
right, Four mice in a square. Bottom: 16 mice in random starting positions

90–180/N degrees for the N-mice problem. An infinite number of mice will
make a circle. See Nahin (2012) for a thorough and entertaining discussion.
The usual oddities of logarithmic spirals apply. As the total path length is
finite, the mice will meet in finite time. For the special case of N ¼ 4, the path
length is curiously equal to the side of the square. However, the mice will have
to rotate an infinite number of times to get there, and most likely get ripped
apart by the centrifugal force in the approach.
If the mice do not start in the corners of a regular polygon, more complex
patterns will arise. Random positions can produce some striking figures. This
general case for N ¼ 3 is discussed by Nahin (2012). Higher N is conveniently
studied numerically (see program code in Appendix B), or, better, with real
people running after each other, which would make a nice student project.
24
Spiral Jetty, Tatlin’s Tower

The spiral is ubiquitous in modern art, forming a link between primitivism


and constructivism (Israel 2015). The most famous piece of spiral art is surely
“Spiral Jetty” by Robert Smithson. This 460-m long jetty of mud and rock was
built in 1970 on the shore of the Great Salt Lake in Utah. It starts as a long
straight line projecting into the lake, then curving into a spiral. Like the
impressive spirals of the Nazca Lines in Peru (200–600 AD), Spiral Jetty is
best appreciated from the air.
In spite of Smithson’s scientific interest and knowledge, he does not seem to
have cared much about the mathematical properties of his spiral. In a sketch
for the work (Fig. 24.1), the spiral appears Archimedean in the inner part, then
increasing the whorl distance outwards towards the straight tail. However, the
finished work is more Archimedean throughout. Smithson did not describe
the details of construction. He says that “a string was then extended from a
central stake in order to get the coils of the spiral”, but he does not reveal how
the length of string was varied with angle.
Smithson’s 1972 essay on the work, and the accompanying movie
(Fig. 24.2), are truly beautiful. He connects his spiral to galaxies, to the
propeller on the helicopter he sits in while filming, to screw dislocations in
crystals:

This site was a rotary that enclosed itself in an immense roundness. From that
gyrating space emerged the possibility of the Spiral Jetty.

His linking of the spiral and the Sun, both in the text and in the movie, is
reminiscent of the symbolic connection between spirals and the Sun in

© Springer International Publishing AG 2016 101


Ø. Hammer, The Perfect Shape, DOI 10.1007/978-3-319-47373-4_24
102 The Perfect Shape

Fig. 24.1 Robert Smithson. Sketch for Spiral Jetty in Red Salt Water. 1970. Art
© Estate of Robert Smithson/VAGA, NY/BONO, Oslo 2016

Fig. 24.2 Spiral Jetty bathing in sun light. Smithson 1970. Art © Estate of Robert
Smithson/VAGA, NY/BONO, Oslo 2016

prehistoric Europe. The priest kings of the Bronze Age would have approved
of Spiral Jetty.
More geometric, but no less poetic, are the unrealized plans for a giant,
constructivist tower in Petrograd (St. Petersburg). Designed by Vladimir
Tatlin in 1919–1920, the Monument to the Third International was to be
some 400 m tall, consisting of a trochospiral, right-coiling double helix of steel
together with supporting structures (Fig. 24.3). Inside it, enormous buildings
24 Spiral Jetty, Tatlin’s Tower 103

Fig. 24.3 Model of Tatlin’s Monument to the Third International, 1919

in geometric shapes (cube, pyramid, cylinder and hemisphere) were to rotate


in daily, monthly or annual cycles. The connotations of a colossal astronomical
clock were reinforced by the whole tower being tilted 23.5 , like the axis of the
Earth. The daily and annual cycles of the Sun’s movement in the sky. A giant
spiral. There is that symbolic connection again . . . surely it must be a
coincidence?
25
Now It Gets Complex

Math was becoming a hot subject again in the Renaissance. A new breed of
brilliant mathematicians was enthusiastically building on the solid foundation
made in Antiquity. It is clear from their writings that they felt very cool and
proud, discovering that they could outperform the Greek! One of their early
areas of study was the algebraic solutions to polynomials (i.e., giving an explicit
formula for the roots), something the Greek had glossed over in their enthu-
siasm for geometry. Medieval mathematicians in India and elsewhere had
already figured out a formula for the second-order, or quadratic, equation,
ax2 + bx + c ¼ 0. Here it is (for me it brings back unhappy memories of
sweaty High School math classes):
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
b  b2  4ac
x¼ :
2a

Clearly, it is easy to make a second-order polynomial where the number under


the square root becomes negative. For example, consider x2 þ 2x þ 3 ¼ 0.
According to the equation, we should have x ¼ 1  √2. This was not
immediately a problem to the medieval mathematicians; they could just
shrug and say that you can’t take the square root of a negative number, so
the equation has no solution. But things got worse.
In the sixteenth century, a group of Italian mathematicians started to tackle
the difficult algebraic solutions to third- and fourth-order polynomials. Fore-
most among them were Niccolò Fontana Tartaglia (1499–1557), Gerolamo
Cardano (1501–1576), Lodovico Ferrari (1522–1565) and Rafael Bombelli
(1526–1572). For some of these polynomials, square roots of negative numbers

© Springer International Publishing AG 2016 105


Ø. Hammer, The Perfect Shape, DOI 10.1007/978-3-319-47373-4_25
106 The Perfect Shape

turned up in intermediate calculations, even if the final solution was a


completely respectable real number. There was no way around it anymore. It
was necessary to attack the square roots of negative numbers directly—to accept
them as useful constructions and to make rules for how to calculate with them.
pffiffiffiffiffiffiffi2
First of all, it was clear that 1 should be defined as equal to 1. This
is, after all, what the square root is all about. However, this leads to a subtle
pffiffiffiffiffiffiffi2 pffiffiffiffiffiffiffipffiffiffiffiffiffiffi pffiffiffiffiffiffiffi
problem. By definition, 1 ¼ 1 1 ¼ 1: But shouldn’t 1
pffiffiffiffiffiffiffi pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
1 also be equal to ð1Þð1Þ, i.e., þ1? It seems we cannot use all of the
pffiffiffiffiffiffiffi
usual algebraic rules for 1. To reduce the risk of such errors, Gauss
suggested to instead use the letter i for this quantity, i.e., i2 ¼ 1. That’s
i for imaginary, an unfortunate name (due to Descartes) as it implies some-
thing mysterious and non-existing.
What does all of this have to do with spirals? Quite a lot. We are getting there.
First, we need to define complex numbers. A complex number is complex but
not complicated; it is simply a number consisting of a real and an imaginary
part: z ¼ a + bi. We can plot a complex number as a point in the complex plane,
with the real part on the horizontal axis and the imaginary part on the vertical
axis (Fig. 25.1).
Perhaps the main reason for the popularity of complex numbers in science
and engineering is a very surprising but simple equation due to Euler (1743).
Although it has a firm theoretical basis, it is so strange that it is true almost as
much by definition as by derivation:

eiθ ¼ cos θ þ isin θ:

Im

2 z = 3 + 2i

1 2 3 Re

Fig. 25.1 A complex number z plotted in the complex plane. Re is the real axis, Im is
the imaginary axis
25 Now It Gets Complex 107

The special case for θ ¼ π gives what many regard as the most beautiful
identity in mathematics: eiπ ¼ 1.
Euler’s formula means that z ¼ eiθ describes a circle with radius 1 in the
complex plane. Now consider the equation z ¼ ekθeiθ, or z ¼ e(k+i)θ. It is simply
the exponential function with a complex exponent. Because of the imaginary
term in the exponent, the curve will rotate around the origin. Because of the
real term, the radius will increase exponentially. The curve described in the
complex plane must be a logarithmic spiral!
Considering the fundamental importance of the exponential function in
mathematics, it is not surprising that logarithmic spirals in the complex plane
turn up in all kinds of systems involving complex numbers. In the 1980s and
1990s, fractals were all the rage. These beautiful mathematical objects are
defined by being infinitely wrinkly, never getting smooth no matter how far
you zoom in on them. One example is the Julia set of the quadratic complex
map. Constructing it is quite simple. Start with a complex number z ¼ a + bi,
and plug it into the function f(z) ¼ z2 + c. The complex number c is a fixed
parameter—you will get different figures for different c. This will give you a
new number f(z), which you plug back into the equation as a new value for z. If
you repeat this feedback process, you will find that the resulting series of
numbers either remains bounded or goes to infinity. If it remains bounded
then plot the original point a + bi with a bright color in the complex plane. If it
diverges, paint it with a color depending on how fast it escapes. Now repeat the
whole operation for all starting points a + bi within some rectangle in the
complex plane, and you might get something like this (Fig. 25.2):

Fig. 25.2 Julia set for f(z) ¼ z2 þ 0.285 þ 0.01i, with zoom to the lower right quad-
rant. Program code in Appendix B
108 The Perfect Shape

The spirals are close to logarithmic. And you can zoom in to see new
beautiful spirals made of spirals inside spirals—it never ends. Although fractals
are now so totally out of fashion, the strangeness of such endlessly rich images
coming out of such a simple algorithm never gets old. A gallery of strange
spirals in Julia sets is given by Davis Philip (1992).
26
The Killer Spiral

A monolayer is a membrane with one layer. A bilayer is a membrane with two


layers. Lipid bilayers make up your cell walls and the walls around your cell
nuclei. If the bilayer is flat, the two layers must have the same area. But if one
layer then expands more than the other, the membrane will have to bend, and
then it gets interesting. Bending of membranes by expansion of one side is a
fundamental mechanism for producing form in Nature, responsible for the
formation of cells, for making all kinds of organs and vesicles in the embryo,
and for the coiling of materials during drying and heating. The bimetal,
invented by John Harrison (1693–1776), the legendary clockmaker, is a
bilayer of two different alloys with different thermal expansion coefficients,
causing the strip to bend when heated.
While a spruce or pine cone hangs on a tree, all the little scales are kept
moist. After falling to the ground, the cone starts drying, but the outer side of
each scale dries faster than the inner side. Hence, the scale contracts more on
the outer side than the inner, causing the scales to bend outwards and release
the seeds (Reyssat and Mahadevan 2009). The mechanism is very similar to
that of the bimetal, but driven by dehydration rather than temperature.
Curvature in one direction is conveniently measured as 1/r, where r is the
radius of curvature, or the radius of a circle inscribing the curved membrane. If
we know that a is the length of the outer layer, b the length of the inner layer,
and d the total thickness of the bilayer, we can compute the radius of curvature
(Fig. 26.1).
Taking the radius r to the centre line of the bilayer, and measuring the angle
φ spanned by the curved membrane in radians, we have

© Springer International Publishing AG 2016 109


Ø. Hammer, The Perfect Shape, DOI 10.1007/978-3-319-47373-4_26
110 The Perfect Shape

b
d

r
φ

Fig. 26.1 The radius of curvature r for a bilayer of thickness d, outer length a, and
inner length b

 
d
a¼φ rþ
2
 
d
b¼φ r
2

These equations give


a r þ d=2
¼ :
b r  d=2

Solving for r, we get


d ða þ bÞ
r¼ :
2ða  bÞ

Now if the length of a curved bilayer sheet is exactly 2πr, it will curl up into a
cylinder of radius r where the two edges of the sheet are just touching each
other. If the sheet is longer, it is forced to curl into a spiral scroll. The outer
layers of the scroll will not be able to coil up as much as they would like to,
being constrained by the inner layers. This happens with drying birch bark,
with old parchment, and with countless other things (Fig. 26.2).
There is also a mineral like that, one that you would not like to meet.
Chrysotile has a bilayer structure, with a brucite layer Mg(OH)3 and a tetra-
hedral silicate layer Si2O5. The thickness of the bilayer is about 0.8 nm. The
26 The Killer Spiral 111

Fig. 26.2 Cinnamon sticks. The dried bark folds into single and double scrolls

Fig. 26.3 Chrysotile fibril

unit cell of the brucite layer is slightly larger than that of the silica layer, causing
the sheet to bend into a spiral scroll with a typical diameter of 25 nm. The
scroll can be very long, making a fibril (Fig. 26.3). The fibrils join together into
the long, sharp, nasty fibers of asbestos. If you breathe in enough of these fibers
you will die.
27
The Friend

Mountains are full of cracks, which are nice to grab into when climbing. How
should we design an anchor for such cracks—a safe, strong attachment for the
climber’s rope? This engineering problem has a surprising solution.
In 1974, Ray Jardine and his companions were making astonishingly fast
climbs up the imposing granite walls of Yosemite, California. A previous 3-day
record was broken by a 20-h virtual run up the cliff. His peers were
dumbfounded—this could not be done! Jardine managed to keep the secret
inside his small blue bag for several years: a new, strange contraption that
would revolutionize climbing. The code word for the device was “the friend”.
In fact, the Russian mountaineer and inventor Vitaly Abalakov (1906–1986)
had already made a similar thing, but Jardine improved the design.
One possible crack anchor involves two metal bars, hinged in the middle
and with the rope pulling on the hinge (Fig 27.1, left middle). The good thing
about this device is that the harder you pull on the rope, the harder the bars
will stem onto the rock, increasing friction further. The bad thing is that it will
only work in a crack of a certain width. If the crack is slightly wider, the bars
will make a too large angle with the walls (or even worse, not even reach there),
and you die. If the crack is slightly narrower, the bars will make a too small
angle with the wall. This will make them slip, and you die. It turns out that for
typical rocks, the ideal angle, giving high friction but without putting too
much force on the device, is about 76 . What we need, then, is a shape that
makes a constant angle of 76 between the wall at the contact point and the

© Springer International Publishing AG 2016 113


Ø. Hammer, The Perfect Shape, DOI 10.1007/978-3-319-47373-4_27
114 The Perfect Shape

76º
76º

Fig. 27.1 Left: A hinged bar anchor would work for only a particular crack width.
Right: A logarithmic spiral anchor

force line from the contact point to the hinge, regardless of the length (radius)
of that line.
The same angle everywhere—it is the equiangular spiral again!
The resulting device is called a cam (Fig 27.1, right). These days you will
find countless versions of it in any sports shop selling climbing gear. It all
started with Descartes.
28
The Labyrinths of History

One of the oldest, deepest and most pervasive symbols of Western culture is
that of the labyrinth. With its winding, unfathomable roots extending deep
into the Neolithic, the labyrinth can be followed through Greek mythology
and Roman art and remains important in Christian architecture.
Very strangely, many of the early labyrinths, including some Neolithic
examples and most of the Greek ones, follow an identical scheme known as
the seven-course design (Fig. 28.1). There is only one possible path, taking you
from the entrance to the center.
The most mysterious structures from the Neolithic and Bronze Age of
Europe are perhaps the megalithic stone circles and the “henges” of Britain.
They come in a bewildering variety of shapes and sizes. Some of them may
represent domestic or defensive buildings, but most of them must have had
some ritual significance. Some, such as Stonehenge, were oriented according to
the movement of the Sun, and may have been sacred “astronomical observa-
tories” for celebrating and ensuring the passage of the seasons. It is not difficult
to understand why these fantastic constructions attract modern people seeking
mystery and a feeling of cosmic spirituality.
The “wood henges” and related structures can now be recognized mainly
from patterns of holes that originally held huge, wooden posts. Woodhenge
itself, situated some 3 km northeast of Stonehenge, can be dated very roughly
to about 2000 BC. The whole structure, including a surrounding ditch and
bank, is 85 m wide. In the center, archeologists discovered the skeleton of a
child, perhaps sacrificed. Another skeleton, of a teenager, was found nearby.

© Springer International Publishing AG 2016 115


Ø. Hammer, The Perfect Shape, DOI 10.1007/978-3-319-47373-4_28
116 The Perfect Shape

Fig. 28.1 The classical, seven-course labyrinth consists of two intertwining lines
(red and black)

Surrounding the central burial are six vaguely concentric circles of post holes.
A gap in the outermost bank is interpreted as the entrance.
The original appearance of these timber constructions (including whether
they held up a roof) is disputed, but the patterns of postholes inspire some
interesting speculation. In his book “Understanding the Neolithic” (1999),
Julian Thomas describes the confusing patterns of post holes at Woodhenge
and nearby Durrington Walls. If you try to retrace the steps of the Neolithic
pilgrim or shaman coming in through the entrance at Durrington Walls, you
find that after passing four of the concentric circles, your path is blocked by a
post. The interpretation is that the visitor now had to turn to the side, walking
between two concentric rows.
Now look at the schematic labyrinth above. The resemblance is uncanny.
Just as at Durrington Walls, you pass four concentric circles before turning to
the right. Add to this the opinion of some scholars that screens were suspended
between the posts, and the interpretation of wood henges as some sort of
labyrinth is not far-fetched. There is little evidence that Woodhenge
reproduced the classical labyrinth in detail, but it is not easy to explain the
arrangement of postholes in terms of engineering, for holding up a roof. It
seems too asymmetric and irregular. The idea of some kind of rite of passage,
where youths entered the labyrinth like Theseus, progressing through the
unfathomable circles to the sacred center, between sheets flapping in the
wind, to reappear in the outside world cleansed and spiritually elevated, is
appealing. A more sinister possibility, suggested by the Greek myths and the
finds of adolescent skeletons at Woodhenge, is that some horrific priest resided
28 The Labyrinths of History 117

Fig. 28.2 Man-in-the-maze, American southwest

in the center, demanding sacrifices of young men and girls like the Minotaur.
In any case, the fact that the wood henges were built during the height of
Minoan culture does not harm our speculative story.
The passage of purification or catharsis into a spiral-like labyrinth and out
again is deeply entrenched in the mythology of Christianity. Just think of
Dante’s Divine Comedy, with its descent through the circles of Hell, followed
by the mirror-imaged ascent up the terraces of the mountain of Purgatory. The
labyrinth in the Cathedral of Chartres is used by modern pilgrims in a similar
way, and it is not unreasonable to imagine that the same idea was prevalent in
medieval times. The connection to prehistory is illustrated by the fact that
there was originally a bronze plaque in the center, depicting Theseus slaying
the Minotaur.
But perhaps the most enigmatic fact of all is this: The very same geometric
construction is found in the ancient cultures of America and Asia. Labyrinths
with the seven-course classical design turn up at Precolumbian sites in North
America, and at Vedan sites in India (Fig. 28.2).
At this point things are getting quite incredible, and the temptation for
producing wild diffusionist theories uncomfortable, so let us leave it there!
29
Newton’s Spiral Headache

The Law of Universal Gravitation was perhaps Isaac Newton’s greatest achieve-
ment. Without it we would not have satellite communication, we would not
have reached the Moon or Mars, and, more importantly, we would have no idea
how the Universe works. The law says that any two bodies are attracted to each
other by a force that is proportional to the product of their masses and inversely
proportional to the square of the distance, with a coefficient G called the
gravitational constant, G ¼ 6.67408  1011 m3 kg1 s2:
m1 m2
F¼G :
r2

Although this law had already been suggested by Hooke, Wren (remember
him, the man who made snails by coiling up cones?) and others, it was Newton
who demonstrated it mathematically in his Principia Mathematica (1687). Or
did he? In Basel in Switzerland, there was a clever, irritating man who thought
otherwise. His name was Johann Bernoulli, and he was none other than the
brother of Jakob Bernoulli, the logarithmic spiral enthusiast whom we met
earlier. Spirals took center position also in Johann’s controversy with Newton,
starting in 1710. The battle would rage for nearly ten years, delaying the
general acceptance of the Law of Gravitation.
The bone of contention was the “inverse square force law”. Newton had
forcefully demonstrated that if a planet or comet follows the path of a conic
section (ellipse, parabola or hyperbola), as had been observed by keen-eyed
astronomers, the gravitational force would have to be inversely proportional to
the square of the distance. Bernoulli had no qualms with that. The problem

© Springer International Publishing AG 2016 119


Ø. Hammer, The Perfect Shape, DOI 10.1007/978-3-319-47373-4_29
120 The Perfect Shape

was the inverse statement, namely that if the square force law holds, then the
path would necessarily be a conic section (Corollary 1 to Proposition 13 in the
Principia). Newton had not proven this explicitly, at least not in the first
edition (the second edition arrived in 1713).
It could not be denied that Bernoulli had a point there. To explain why the
inverse law was a logical fallacy, he produced a counter-example. Newton
himself had calculated what the force law would have to be if a body moved in
a logarithmic spiral trajectory rather than a conic section. It does not do that, of
course, but it was an interesting exercise. The answer is that the force would
have to be inversely proportional to distance cubed, rather than distance
squared. But Bernoulli was able to show that if a body moved in a hyperbolic
spiral, the force would also follow an inverse cube law. The particular spiral
(logarithmic or hyperbolic) would depend on the initial velocity. In other

A B

E
C D

Fig. 29.1 The five “species” of Cotes’ spirals. (a) A secant spiral, r ¼ 1/cos(0.07φ).
(b) A hyperbolic cosecant spiral (Poinsot spiral), r ¼ 1/sinh(0.1φ). Program code in
Appendix B. (c) A hyperbolic secant spiral (the other Poinsot spiral), r ¼ 1/cosh(0.1φ).
(d) A hyperbolic spiral, r ¼ 1/φ. (e) A logarithmic spiral, r ¼ e0.1φ
29 Newton’s Spiral Headache 121

words, even if we show that a logarithmic spiral path implies an inverse cube
law, the converse does not follow, because a body moved by such a law could
also follow other paths. Bernoulli argued that if this problem arises for the
inverse cube law, why not also for the inverse square law? Maybe there are
other paths than the conic sections, compatible with Newton’s law of
gravitation?
Bernoulli’s point was made even clearer by Roger Cotes in 1714 (Harmonia
Mensurarum, pp. 30–35). He found that not only the logarithmic and hyper-
bolic spirals, but also three other spirals could result from an inverse cube law,
namely the secant spiral, the hyperbolic secant spiral, and the hyperbolic
cosecant spiral. These additional cases are now often called Cotes’ spirals
(Fig. 29.1).
Newton and his proponents were eventually able to clarify their position
and to explain why the conic sections are the only possible solutions under the
inverse square law. In 1720, Bernoulli finally gave up throwing spirals at
Newton and acknowledged that the conic sections are the only possible
paths of celestial bodies if the Law of Universal Gravitation holds true. Well
done. We would not have had GPS or space telescopes without these guys.
30
Sculptures of the Sea

Nautiloids and most ammonites are either straight-shelled or planispiral,


meaning that they coil in one plane. Because of this, they are bilaterally
symmetric: the left is a mirror image of the right. Now grab the center of
the logarithmic spiral and pull it out in the direction normal to the coiling
plane. The bilateral symmetry is broken, and the shell is now called
trochospiral. This additional feature opens up the full treasure chest of mollusk
shell shape—from chunky whelks and clams to the most elegant and delicate
conches. In fact, as a first approximation, it is possible to reduce most mollusk
shells (and some other shells too, such as those of brachiopods) to a mathe-
matical model involving only three parameters k, T and D.
The first parameter is k, the expansion coefficient, controlling the “tight-
ness” of the logarithmic spiral. Sometimes we use a related value called
W instead, giving the increase per whorl of the spiral radius (the whorl
expansion factor). With r ¼ ekφ, the equation for the logarithmic spiral, we
have that W ¼ e2πk.
The second parameter is T, the trochospirality. The first step in making the
shell is to produce the generating curve—a hypothetical spiral line running
through the center of the coiling, expanding tube (Fig. 30.1). This line is
simply a logarithmic spiral pulled along the z axis. In cylindrical coordinates:

r ¼ ekφ ;
z ¼ Tr:

To make a planispiral shell we set T ¼ 0.

© Springer International Publishing AG 2016 123


Ø. Hammer, The Perfect Shape, DOI 10.1007/978-3-319-47373-4_30
124 The Perfect Shape

Fig. 30.1 The generating curve (red ) of the seashell model, in cylindrical
coordinates (φ, r, z). In this case, the parameters of the model are k ¼ 0.1, T ¼ 1.6.
The aperture (gray) with increasing radius R is moved along the generating curve,
producing the surface of the shell

We also need to decide on a shape for the aperture, e.g., a circle. This choice
constitutes additional free parameters that do not concern us at this point. The
aperture is pulled along the generating curve, sweeping out the surface of the
shell, and as it goes, it is being scaled by increasing its radius R according to the
third parameter, D:
R ¼ Dr:

The parameter D, the whorl overlap, controls how much the shell tube over-
laps the previous whorl. A shell with small overlap is called evolute, a shell with
large overlap is involute.
That is all. The idea for this model was introduced in a thorough but highly
readable paper by Rev. Henry Moseley (also known as Canon Moseley)
already in 1838 (Fig. 30.2). Validating his mathematical models for growth
and form with measurements on real specimens, Moseley ranks among the first
mathematical biologists in the modern sense.
30 Sculptures of the Sea 125

Fig. 30.2 From Rev. H. Moseley (1838)

Fig. 30.3 Tower cranes

By the way, the hook of a tower crane (Fig. 30.3) follows a three-
dimensional conical path similar to Moseley’s generating curve but based on
an Archimedes spiral. Tower cranes have all kinds of fascinating mathematical
and physical properties, one of them being that they basically work in
126 The Perfect Shape

Fig. 30.4 Left: The trochospiral path (red ) followed by a rotating crane with
continuous hoisting and outwards movement of the trolley. Right: The Silurian
graptolite Spirograptus turriculatus, ca. 4 cm long, from the collections of the
Natural History Museum, University of Oslo

cylindrical coordinates. The operator uses a joystick controlling three motors:


One for rotation of the crane, one for the position of the trolley along the
working arm, and one for the height of the hook. These parameters, called
slew, trolley and hoist, are nothing but the polar angle, radius and elevation in
a cylindrical coordinate system. Consider first holding the hoist constant while
rotating the crane and increasing the trolley at constant speeds. The hook will
describe an Archimedes spiral. Now start hoisting. The load will make a helical
motion on the surface of a straight cone, tip down (Fig. 30.4).
The crane spiral is an Archimedes spiral pulled out linearly along the z axis,
so that the vertical distance between whorls is constant rather than geometri-
cally increasing as in the snail shell. It is more properly known as a conical helix
or conical spiral. The conical spiral is not gnomonic like Moseley’s logarithmic
seashell model, and not common in nature. The only example I know of
among the animals is the graptolite Spirograptus turriculatus from the Silurian,
some 440 million years old (Fig. 30.4). The graptolites were strange, colonial,
mainly planktonic organisms common from the Ordovician to the
Carboniferous.
Perhaps the most awe-inspiring conical helix is the external spiral ramp of
the Malwiya (“Spiral”) minaret of Samarra, Iraq, part of the Great Mosque
(848–852 AD). It was built by the great Caliph Ja’far al-Mutawakkil ‘ala
Allah, probably as much as a symbol of his immense power as an act of piety.
The Malwiya is a conical tower, 52 m tall, standing on a square base
30 Sculptures of the Sea 127

Fig. 30.5 The Malwiya minaret, Great Mosque, Samarra, Iraq, in 1973. IgorF,
vlastni foto, CC BY-SA 3.0 licence

(Fig. 30.5). In 2005, after American and Iraqi troops had seized Samarra,
U.S. soldiers ascended the holy helix, like al-Mutawakkil himself on his white
donkey more than a millennium before, like Dante up Mount Purgatory, like
King Nimrod up the Tower of Babel, and established a sniper base at the top.
Inevitably, the top floor was blasted to smithereens by the insurgents.
Moseley’s logarithmic shell model was made famous among scientists by the
computer simulations of paleontologist David Raup. In a paper in the journal
Science in 1962, he described how it is possible to plot 2D cross sections of
mathematical shell models using an IBM-7090 computer (one of the first
transistorized computers) connected to a plotter. From these cross sections, he
was able to make perspective drawings by hand. Three years later, in 1965, he
managed to automate the perspective drawing, including removal of hidden
points (Fig. 30.6, left). However, Raup noted that the procedure is “relatively
128 The Perfect Shape

Fig. 30.6 Raup’s computer renderings of mollusk shells. Left: Perspective plots
made with the IBM-7090 computer. Right: 3D plots made with a PACE TR-10 analog
computer connected to an oscilloscope with intensity control. From Raup and
Michelson (1965). Reprinted with permission from AAAS

costly in terms of computer time”. This may seem absurd to the modern
computer user, but the 7090 cost nearly three million USD in 1960, and could
carry out only about 0.4 million operations per second. Your little computer is
at least a thousand times faster.
To reduce cost, Raup also experimented with the use of analog computers
connected to analog cathode-ray tube displays (Fig. 30.6, right). The results
were quite elegant, and allowed him to investigate a whole range of possible
shell morphologies. An analog computer, by the way, is a thing of beauty. Now
completely obsolete of course, it is a calculator that uses simple electronic
circuits to compute arithmetic operations, and differentiation and integration,
on voltages, representing numbers. The PACE TR-10, used by Raup, was a
fairly small machine, about the size of a kitchen chair (Fig. 30.7). I find it
astonishing that you can generate the shell of a whelk with a few resistors and
transistors.
Nowadays it is trivial to draw Raupian shells by computer, as in Fig. 30.8.
The model makes it clear that mollusk shells exist within a continuous shape
space spanned out by the model parameters—a morphospace—such that
evolutionary change between the shapes should be relatively easy. The shells
in the figure are taken from a movie that I have placed on Youtube, showing
smooth transformation between the different forms: https://www.youtube.
30 Sculptures of the Sea 129

Fig. 30.7 PACE TR-10 analog computer, 1960. Photo Daderot. CC0 1.0 Universal
Public Domain Dedication

com/watch?v¼f2GYFUipkvw. You might also like to try my interactive 3D


shell generator for Windows, at http://folk.uio.no/ohammer/seashell.
Also recall that all these shapes are gnomonic and self-similar, so the shape of
the juvenile shell, still preserved near the apex, has the same shape as the
complete, adult shell. In nature, this is rarely exactly the case, because the
growth parameters vary somewhat over the life of the animal. Still, the
logarithmic spiral provides a remarkably good model for seashells. Perhaps
130 The Perfect Shape

Fig. 30.8 Some shell shapes made using the Raup model. Top left: Planispiral shell
(T ¼ 0), moderate k, small D, give an ammonite shape. Top right: Slightly
trochospiral (T > 0) with large k—a cockle shell with the left–right asymmetry
which is characteristic for the bivalves. Bottom left: Moderate T, a low-spired
snail. Bottom right: Large T and D give a high-spired (turriform) shell

this is part of the reason we find this curve so aesthetically appealing—the


Perfect Shape.

The Beauty of the Columella


Seashells are gorgeous on the outside. But they are just as stunning on the other
side, the inside. Many people have never seen the inside of a snail shell, which
is a pity. It feels horribly wrong to saw a conch in two, or even worse, to smash
it with a hammer. To use an X-ray CT scanner is a more civilized alternative
(Fig. 30.9). The most surprising feature inside a snail is the columella, a pillar
running along the central coiling axis. Basically just the inner wall of the
coiling cone, it is an architectural gem, sometimes embellished by helical
grooves making it look like a screw or spiral staircase.
30 Sculptures of the Sea 131

Fig. 30.9 Top left: The inside of Cymbiola, a volute snail, with its columella, a right-
handed helix. Ca. 5 cm long. 3D tomography and X-ray image. Top right: A terebrid.
Bottom left: This Conus has a tenuous, fragile columella. 3D and X-ray. Bottom
right: The top of the Conus shell has been removed by computer. The images were
obtained with an industrial CT scanner, with no harm done to the shells

The Little Lid


Many snails can close their shell with a door called the operculum, which is
Latin for ‘little lid’. For the winkles on rocky shores, this is a life-saver because
it stops desiccation at low tide. For other snails, the little lid provides a
formidable defense against crabs and other predators. And in the operculum
of many species, especially of the marine genus Turbo, we find yet another
magnificent spiral (Fig. 30.10). In right-coiling (dextral) shells, the spiral as
seen on the inside of the operculum always coils clockwise outwards. It is as if
the snail wants to decorate its door with a shape that symbolizes its whole way
of living.
132 The Perfect Shape

Fig 30.10 The inner side of the operculum of a Turbo snail. The active growth edge
is at upper left. Ca. 3 cm across

The construction of the door is amazing. The disk has to fit the opening of
the shell snugly at all times, while the size of the opening is increasing. This is
perhaps not very difficult if the operculum grows concentrically all around the
edge. Some species do exactly that, but the operculum of Turbo and many
other snails grows in gnomonic fashion by adding curved triangular incre-
ments along only part of the edge. This process follows a similar principle as
the growth of the entire shell, and gives a logarithmic spiral for the same
mathematical reason. But the weird and wonderful thing is how the whole
operculum can keep fitting the opening when growing in this way. To make it
work, the operculum has to be continuously rotated.
Henry Moseley commented on the surprising properties of the Turbo
operculum already in his 1838 paper:

That the same edge which fitted a portion of the first less section should be
capable of adjustment, so as to fit a portion of the next similar but greater section,
supposes a geometrical provision in the curved form of the chamber of great
apparent complication and difficulty. But God has bestowed upon this humble
architect the practical skill of a learned geometrician.
31
The Spiral of the Bird Priests

In ancient Rome, politicians did not listen to economists with computer


models, or technocratic advice from consultants and bureaucrats. Instead,
they watched the birds. Perhaps it worked just as well. A small collegium of
priests, the augurs, interpreted the will of the gods from the flight of birds,
their species, their sounds, their directions and velocities. And in his hand, the
augur would hold a lituus, a wand in the form of a spiral, a symbol of his
wisdom and power, a tool for marking out the four directions of the heavens.
The lituus is a common symbol on Roman coins, signaling the virtues of the
emperor.

In his Harmonia mensurarum, published posthumously in 1722, Roger


Cotes set out to construct a spiral such that all sectors between the x axis
and the curve have the same areas (Fig. 31.1). The result was a curve on the
polar form r ¼ a/√φ. The equality of sector areas follows immediately from the
definition. Squaring and re-arranging, we get φr2 ¼ a2, and the area φr2/2 of
the sector is then the constant a2/2.

© Springer International Publishing AG 2016 133


Ø. Hammer, The Perfect Shape, DOI 10.1007/978-3-319-47373-4_31
134 The Perfect Shape

r φ

Fig. 31.1 The lituus of Roger Cotes (1722). All sectors (e.g., the one shown in
orange) between the x axis and the curve have the same area

Noting the similarity with the augur’s wand, Cotes called this curve the
lituus (several web pages and books erroneously attribute this name to Colin
Maclaurin). I wonder how many of today’s mathematicians would have made
that connection. Classical scholarship is not what it used to be.
32
Squaring the Circle

After Archimedes defined his spiral in his book On Spirals, a further 17 theorems
are proven. Perhaps the most interesting result is hidden in Proposition
18, which reads as follows:

If OP be the initial line, P the end of the first turn of the spiral, and if the tangent
to the spiral at P be drawn, the straight line OT drawn from O perpendicular to
OP will meet the said tangent in some point T, and OT will be equal to the
circumference of the first circle.

Let us try to make sense of this. The theorem is illustrated in Fig. 32.1. The
line OP is the straight line referred to in the definition, used to generate the
spiral. This line revolves one full turn, at which point the spiral has reached
point P. Now draw the tangent to the spiral at P, and also a line perpendicular
to OP from O. These two lines meet at a point T. The theorem tells us that the
length OT is equal to the circumference of “the first circle”, which means
the circle with radius OP. In other words, OT ¼ 2πOP. Moreover, the area of
the circle can be found by a trivial combination of this theorem with an earlier
proposition by Archimedes (Proposition 1 in the work called Measurement of
the Circle):

The area of any circle is equal to a right-angled triangle in which one of the sides
about the right angle is equal to the radius, and the other to the circumference, of
the circle.

© Springer International Publishing AG 2016 135


Ø. Hammer, The Perfect Shape, DOI 10.1007/978-3-319-47373-4_32
136 The Perfect Shape

O P

Fig. 32.1 Proposition 18 in Archimedes’ On Spirals states that the circumference of


the red circle equals OT

This means that the area of the first circle is equal to the area of the triangle
OPT. In other words, if we can draw the spiral and its tangent at P, we know
the area of the circle with radius OP. Characteristically, Archimedes does not
state the triumphant proposition, but leaves it for the reader to figure out: He
had squared the circle. Today, this is just another of those mysterious,
old-fashioned idioms that we use off-hand without much contemplation,
like “acid test”, “buy a pig in a poke” and “cut the mustard”. Squaring the
circle refers to some vain attempt to accomplish an impossible task. To the
ancient Greeks, it simply meant finding the area of a circle. And since they
32 Squaring the Circle 137

insisted on doing everything geometrically from the axioms as set out by


Euclid, they felt they ought to construct this area by compass and ruler,
forging a square with the same area as the given circle. This problem became
a fashionable puzzle in classical times—something like the Rubik’s cube of the
nineteen eighties or the Sudoku of today. Renaissance scholars revived the
problem, but without much success. Eventually, mathematicians started to
doubt whether the project was at all possible, and then finally, in 1882,
Ferdinand von Lindemann terminated more than 2000 years of debate
through his proof that pi is a transcendental number, outside the reach of
Euclid’s geometry.
Archimedes’ solution cannot be reached by compass and ruler, because it
requires the construction of the spiral. After him, several solutions using
non-Euclidean contraptions were devised by other Greek mathematicians,
and also numerous ingenious approximations to pi were made using precursors
to modern methods of series. Archimedes himself worked in this latter field.
Still, the squaring of the circle by spirals still stands as one of the most elegant
achievements of Greek mathematics. In addition, it was the first scholarly
treatise on spirals in history—the beginning of more than 2000 years of
scientific fascination with these mesmerizing curves.
And by the way, if you found this chapter tough, try reading the original
work by Archimedes, where he actually proves it all. I promise it will cure you
of any misguided ideas about the intellectual superiority of the modern mind.
33
The Daemon Beavers of Nebraska

Frederick Courtland Kenyon would soon become a world authority on the


brain of the bee. But in 1893 he was braving the prairie of Nebraska, on a
mission to explore one of the strangest wonders of paleontology. In his colorful
field report (Kenyon 1895) he describes some bizarre fossils “known for some
time to the cowboys and ranchmen of the region . . . as devil’s corkscrews”.
They occur mostly in the Harrison Formation of early Miocene age, around
20 million years ago. The formation is named after the little town of Harrison,
described by Kenyon as “a brick court house, a church, a school-house, a hotel,
the almost invariable liquor-saloon, several stores, and some two dozen or
more dwellings”. Only a year previously, the corkscrews had been named
Daemonelix by the great Professor Barbour of Nebraska University. They are
found in the thousands over large areas of Nebraska and Wyoming, as giant,
perfectly formed sandstone helices, usually from 210 to 275 cm tall. From the
bottom of the screw, long, straight cylinders can extend horizontally or
diagonally for a considerable distance (Fig. 33.1).
True to their name, the daemon helices caused a scientific mayhem that would
continue for half a century. In a long row of papers in prestigious journals,
bearded scholars presented their theories with much passion and academic pomp,
ranging from giant freshwater sponges (a ridiculous suggestion for anyone who
has seen a sponge) and petrified lianas (how was that supposed to happen?) to
giant plant roots (perhaps not so unreasonable) and many other things.
The truth is very strange, but should really have been obvious in the light of
the evidence. The correct explanation had occurred to Barbour almost from
the beginning, although he would later retract from it. Inside some of the

© Springer International Publishing AG 2016 139


Ø. Hammer, The Perfect Shape, DOI 10.1007/978-3-319-47373-4_33
140 The Perfect Shape

Fig. 33.1 Daemonelix burrow, Nebraska, with Frederick C. Kenyon for scale.
The picture was probably taken in 1893. Courtesy of Agate Fossil Beds National
Monument and James St. John

structures, fitting them snugly, there were complete fossils of a rodent. The
helices are nothing but the spiral staircases of small beavers, leading down to
the living chambers far below ground. In the early Miocene, Nebraska was
covered by dry grassland much like today. Not quite the place where you
would expect beavers, but Palaeocastor lived a long time ago and beavers did
different things back then.
The daemon beaver theory was not generally accepted until after the Second
World War. Martin and Bennett (1977) made a thorough study, describing
many interesting details including the marks from teeth and claws on the
burrow walls. About half of the burrows are left-handed, the other half right-
handed; the old beavers were ambidextrous!
A curious twist to the story was provided by Smith (1987). He discovered a
similar, though less spectacular structure from the late Permian (ca. 255 million
years ago) of South Africa. This was before the mammals, before the dinosaurs.
Inside some burrows, there are skeletons of Diictidon, a dicynodont mammal-
like reptile with a body shape reminiscent of a burrowing rodent. It is a splendid
example of convergent evolution—the independent appearance of similar
organs, body shapes and behaviors in distantly related organisms living in similar
niches. Interestingly, all 50 burrows studied by Smith are right-handed helices.
34
Under the Mistletoe

Walk straight for a certain distance. Then turn to the left, making an angle θ
with your previous walk. Walk a certain percentage further than last time, and
then turn the same angle left. Continuing in this way, you will make a
polygonal spiral with segments in geometric progression.
Consider now such a polygonal spiral where the length of a segment is given
as ci ¼ gci1, with g > 1 an arbitrary constant and with constant turning
angle θ. All triangles made up of a spiral segment c and two successive spiral
radii r in Fig. 34.1 are similar (they have the same shape but different size).
This gnomonic property suggests that we can construct a logarithmic spiral
through the vertices of the polygonal spiral. We can calculate the expansion
rate k of this logarithmic spiral as follows.
The angle Δφ can be found directly from θ because the angles in a triangle
sum to π:
π ¼ Δφ þ α þ β ¼ Δφ þ θ
Δφ ¼ π  θ

Moreover, we have, from the similarities of triangles and the equation for the
logarithmic spiral:

ci ri aekφi
g¼ ¼ ¼ kφ ¼ ekðφi φi1 Þ ¼ ekΔφ ¼ ekðπθÞ
ci1 r i1 ae i1

© Springer International Publishing AG 2016 141


Ø. Hammer, The Perfect Shape, DOI 10.1007/978-3-319-47373-4_34
142 The Perfect Shape

ri

Δφ β
Δφ
r i -1 ci

α
α c i -1 β
θ = α+β

Fig. 34.1 Thick black line: Polygonal spiral with segment ratio g ¼ ci/ci1 and
turning angle θ. Red line: Corresponding logarithmic spiral

Rearranging, we get the desired k as a function of the two parameters of the


polygonal spiral, which were the ratio g of successive spiral segments and the
turning angle θ:
ln g
k¼ :
πθ

In an earlier chapter we met the American artist Robert Smithson, the maker
of Spiral Jetty. Smithson produced several other interesting pieces of spiral art.
One of them is his proposal for an airport in Dallas (1967), originally
constructed with a sequence of triangular mirrors (Fig. 34.2). Each isosceles
triangle has a base angle of θ ¼ 30 , and is scaled so that one base corner meets
the center of the baseline of the following triangle. It is not difficult to show
that each triangle is then scaled up with a factor g ¼ 2√3/3 relative to the
previous one. We can make similar constructions with other base angles,
giving g ¼ sec θ (ah, those poor old forgotten terms of trigonometry—the
secant, or 1/cos, has long passed its Golden Age). Smithson’s handsome
polygonal spiral, looking like the tail of a scorpion, can be approximated
with a logarithmic spiral with k ¼ 0.055, as calculated from the equations
above. He used the same construction for the steel sculpture Gyrostasis, now on
permanent display in the Hirshhorn Museum, Washington DC.
In fact, you can take just about any geometric figure and use it as a gnomon
in this fashion, and you will end up with a shape where consecutive vertices lie
on a logarithmic spiral. In an earlier chapter we saw that you can stack up
golden rectangles to produce the Golden Spiral (or Fibonacci spiral). In the
same way, you can stack up sheets of paper with a height/width ratio of √2,
34 Under the Mistletoe 143

2√3/3

Fig. 34.2 The geometry of Smithson’s proposal for Dallas, Fort Worth, regional
airport, 1967

A0

A4
A3

A5
A1

A2

Fig. 34.3 A sequence of international paper sizes, A0, A1, A2 etc., define a
polygonal spiral

where each sheet is the size of the previous sheet folded in two (Fig. 34.3). This
is the standard sequence of international paper sizes: A0, A1, A2, A3, A4 etc.
The diagonals of these rectangles make a polygonal spiral, with vertices on a
logarithmic spiral with k ¼ 0.221. It seems that any shape-preserving, additive
growth process will lead to a logarithmic spiral. This is a deep truth, bringing
144 The Perfect Shape

Fig. 34.4 Mistletoe in the Botanical Garden, Oslo. Polygonal spiral marked with
dashed line

us close to the root explanation for why logarithmic spirals are so common in
nature.
The European mistletoe Viscum album grows more or less in a plane. At
regular intervals each growth tip branches in two with a certain angle (botanists
call it dichotomous branching). The cycle repeats while the whole plant is
growing. This process gives a polygonal spiral at the edge of the mistletoe.
When I measured up the internodal lengths in Fig. 34.4, I found something
more like an arithmetic than a geometric progression. This would mean that in
equal time intervals the length of every segment increases with a constant
increment, regardless of length, instead of a constant percentage. This indi-
cates that the growth takes place only at the end of a segment, not throughout
it. The result is a shape more like an Archimedes than a logarithmic spiral.
Similar spirals can occur in the necks of long-necked vertebrate animals such
as ostriches, swans, sauropod dinosaurs (those very big ones such as
Brachiosaurus and Diplodocus) and plesiosaurs, and in the tails of snakes,
dinosaurs, cats and monkeys. The length of one vertebra is often a near-
linear function of its number position in the spine, although usually with
some departure at the ends of the neck and tail. The measurements of ostrich
34 Under the Mistletoe 145

neck vertebrae by Cobley and co-authors (2013) is an example. Now assume


that each joint can bend at a constant maximal angle, say 10 from the straight.
In reality, this maximal bending angle varies somewhat along the spine, but let
us keep things simple. The resulting polygonal spiral, with segment lengths in
arithmetic progression and constant angles of the joints, will approximate to an
Archimedes spiral just like the mistletoe.
Another polygonal spiral with a nice story behind it is the so-called Spiral of
Theodorus. The Pythagoreans famously discovered (to their horror, according
to legend) that the square root of two is irrational, meaning that it cannot be
written as a ratio between two integers. But what about the square roots of
other non-square integers? In the dialogue Thaetetus (ca. 369 BC), Plato claims
that the mathematician Theodorus had shown that the square roots of the
integers up to 17 are irrational. The actual proof is not given; the complete
passage in Plato is only the following:

Theodorus here was drawing some figures for us in illustration of roots, showing
that squares containing three square feet and five square feet are not commen-
surable in length with the unit of the foot, and so, selecting each one in its turn
up to the square containing seventeen square feet and at that he stopped.
Transl. Harold N. Fowler, 1921

Other translators give those last, all-important words differently. For exam-
ple, McDowell (1973) uses “at that point he somehow got tied up”. Anderhub
(see next paragraph) quoted the staggering number of 55 different translations
of this passage until 1936, with a bewildering range of meanings. The whole
thing has always been something of a riddle.
Plato goes on to explain that Thaetetus, together with a fellow known as
Socrates the Younger, generalized the proof to any non-square N. Now we may
ask, why did Theodorus stop at 17? We have no idea, but one ingenious theory
was suggested in a long but eloquent essay by a Jakob Heinrich Anderhub,
financial director of the German chemical company Kalle AG (part of IG
Farben) in 1941. The reference is utterly obscure (check it out). Anderhub
suggested that Theodorus started with a right-angled triangle with legs of
length 1, and hypotenuse √2. On the hypothenuse, he constructed another
right-angled triangle with opposed side of unit length, which would have a
hypotenuse of length √3. Continuing in this way, he produced a spiral all the
way to N ¼ 17, but then the figure would intersect itself (Fig. 34.5). This is
why Theodorus stopped at N ¼ 17, according to the theory of Anderhub: the
figure would simply get ugly if he continued (“aus zeichentechnischen
Gründen”). Anderhub does not provide any suggestions about how the proofs
146 The Perfect Shape

4 3
5 2
1
6
1
7 17

16
8
15
9
14
10 13
11 12

Fig. 34.5 The “Spiral of Theodorus”

of irrationality would have been obtained; in fact he suggests that Theodorus


did not prove it at all, only “illustrated” it.
Whether or not this construction was used by Theodorus, or whether or not
Theodorus even existed, it is an interesting spiral. Davis (1993) gives a
thorough and entertaining analysis.
35
Double Spirals, Twice the Fun

A logarithmic spiral shell is gnomonic—the snail or cephalopod inside it can


add gnomon increments to the edge and thus grow without changing shape.
But if one logarithmic shell is gnomonic, then surely the combination of two
such shells is gnomonic too! This opens up interesting possibilities. You can
connect the two shells with a hinge, making an enclosed space which you can
open for feeding and close for protection (Fig. 35.1). The whole structure can
be shaped like a wedge, which can dig down into the mud. The scallops,
Pectinidae, can even flap their shells and swim away with quirky movements.
This clever trick of duplication was made independently (we believe) in two
major animal groups, the brachiopods and the bivalves. The brachiopods, or
lamp shells, form a separate phylum, very common as fossils but relatively rare
in modern oceans. A brachiopod shell is like two connected, planispiral
Nautilus shells, but with a much higher expansion coefficient. This building
plan makes the shell symmetric around the mid plane. The bivalves, on the
other hand, are a class within the phylum of mollusks (Fig. 35.2). A bivalve
shell is more like two connected trochospiral snail shells, producing an asym-
metrical structure twisting to the side.
In most bivalves and brachiopods the two shells connect together very
precisely when closed. If the edge of the shell has folds on it, a convex fold
in one valve will fit into a concave fold in the other. This phenomenon, called
occlusion, raises an interesting question about the regulation of growth. It
would be hard to synchronize growth rates with sufficient precision to ensure
occlusion if the two shells grew independently. Presumably, there is some kind
of feedback control and communication between the two valves, so that if one

© Springer International Publishing AG 2016 147


Ø. Hammer, The Perfect Shape, DOI 10.1007/978-3-319-47373-4_35
148 The Perfect Shape

Fig. 35.1 Two logarithmic spirals connected with a hinge, closed and open

Fig. 35.2 CT images of bivalves. Top left: The common cockle, Cerastoderma edule.
Note the trochospiral twisting to one side (T > 0 in Raup’s model). Top right:
Longitudinal cross section of the cockle, showing the logarithmic spiral. Bottom:
Cross section of the blue mussel, Mytilus edulis, with a very high expansion
coefficient

grows ahead of the other, this will be sensed by some mechanism and the
relative growth rates adjusted.
Bivalves are among the most successful of marine animals. They are often
conspicuous, as in large oyster banks or giant Tridacna clams, but most of
them live out of sight, dwelling in their cold, dark burrows under the vast
muddy seafloor. Bivalves are important and common fossils, occurring in great
numbers especially in rocks younger than 250 million years (since the Trias-
sic). The largest known bivalve, Platyceramus platinus from the late Cretaceous
(ca. 85 million years) of North America, could sometimes reach a length of
3 m, longer than a horse (Kauffman et al. 2007).
When I work as a paleontologist with marine Jurassic rocks, I am always
struck by what we might call the Triumph of the Spiral. Everything is spiral.
The ammonites are logarithmic spirals. The bivalves and brachiopods are
35 Double Spirals, Twice the Fun 149

logarithmic spirals. The snails, the fish teeth and the squid hooks (more on
those later) are logarithmic spirals. The foraminifera are logarithmic spirals.
The scaphopods and many worm tubes are logarithmic spirals. And that pretty
much covers the fossil fauna, apart from the odd starfish, sea urchins, belem-
nites (another cephalopod fossil group) and large marine reptiles. Modern
oceans are not much different. When it comes to marine life, the logarithmic
spiral dominates almost any other shape, only the sphere comes close.
36
Maelstrom

The edge of the whirl was represented by a broad belt of gleaming spray; but no
particle of this slipped into the mouth of the terrific funnel, whose interior, as far
as the eye could fathom it, was a smooth, shining, and jet-black wall of water,
inclined to the horizon at an angle of some forty-five degrees, speeding dizzily
round and round with a swaying and sweltering motion, and sending forth to the
winds an appalling voice, half shriek, half roar, such as not even the mighty
cataract of Niagara ever lifts up in its agony to Heaven.
The mountain trembled to its very base, and the rock rocked. I threw myself
upon my face, and clung to the scant herbage in an excess of nervous agitation.

“This,” said I at length, to the old man—”this can be nothing else than the great
whirlpool of the Maelström.”
Edgar Allan Poe (1841)—A descent into the Maelström

Although the Maelstrom, or Moskstraumen, in Norway is indeed an


impressive phenomenon, it must be admitted that Poe was taking more than
a little poetic liberty. Still, vortices in the form of tornadoes and hurricanes
are some of the most powerful and awe-inspiring phenomena in nature
(Fig. 36.1).
How does a vortex form? The familiar bathtub drain swirl is the archetypical
example. Even before pulling the plug, the water mass will always be in some
complicated motion, however small. Therefore, the velocity field will always
contain a rotating component, just by chance. When the drain is opened, the
water is drawn in, accelerating radially. Now the principle of conservation of
angular momentum comes into play (at this point every textbook conjures up

© Springer International Publishing AG 2016 151


Ø. Hammer, The Perfect Shape, DOI 10.1007/978-3-319-47373-4_36
152 The Perfect Shape

Fig. 36.1 Spiral vortex at Saturn’s North Pole, 2000 km across. Cassini image
courtesy SSI/ISA/ESA/NASA

that pirouetting figure skater, pulling in her arms and thus picking up a
ridiculous rotation speed). The small circular flow is amplified, and a vortex
has been established.
The amplification of angular momentum in a vortex is the basis for a nice
trick played on tourists visiting the equator in South America and Africa.
Water is drained from a hole in a bucket, and a matchstick is placed on the
surface. South of the line, the stick rotates clockwise. North of the line, it
rotates counterclockwise. It is a delightful demonstration of the Coriolis Effect,
which controls the wind direction around low pressures and the direction of
oceanic eddy currents. Unfortunately, it is a harmless hoax. The Coriolis Effect
comes into play only at very large scales, and there is no way it could work in a
bucket. Clearly, the demonstrator is imposing an imperceptibly small rotation
of the fluid in the desired direction before the plug is pulled, perhaps when
filling, or by a discrete dip of the finger. As the vortex develops, this rotation
gets steadily faster and success is guaranteed.
Any particle in the vortex moves in a spiral. What kind of spiral is it? As we
have done several times in this book, we make simplifying assumptions. First,
we limit the analysis to movement in a horizontal plane. Secondly, we assume
that the particle is drawn towards the drain according to the laws of incom-
pressible, potential flow. This is a simplified methodology for hydrodynamics,
making the assumption that the flow is so-called irrotational. This means that
velocity is the gradient of a scalar velocity potential, which again is a solution to
the Laplace equation—a fairly simple partial differential equation. Moreover,
it is easy to derive stream functions and streamlines from the velocity potential
36 Maelstrom 153

(the streamlines are normal to the equipotential lines). The velocities are
tangent to the streamlines.
Because velocity potentials and stream functions can be combined linearly,
we can decompose the flow field into a simple sink flow with flow rate —Q
and a simple circular flow (i.e., zero radial velocity) with circulation Γ. The
stream function in polar coordinates (r,θ) is then (Guyon et al. 2001):
Qθ Γ r
Ψ ¼ Ψsink þ Ψvortex ¼   ln
2π 2π r 0

with r0 an arbitrary constant. A particular streamline is selected by setting Ψ


constant. Solving for r gives

r ¼ keQθ=Γ

where k selects the streamline. The particle moves in a logarithmic spiral! The
expansion coefficient is Q/Γ, meaning that if there is no drainage (Q ¼ 0) the
particles will move in circles.
Certainly, for real-world vortices, we would need to take into account many
other aspects, not least the third dimension, and the logarithmic shape would
be modified. Still, natural hurricanes often tend towards logarithmic spirals.
We are not yet done with the Maelstrom. The magnificent map of Scandi-
navia by Olaus Magnus, the Carta Marina (1539), is adorned with many a
fearful monster of the sea, devouring each other and the unfortunate, brave
ships venturing too far from shore. The Maelstrom is marked with the
ominous label “Hecest horrenda Caribdis”—here is the horrendous Charybdis
(Fig. 36.2).
In Greek mythology, Scylla and Charybdis were two sea monsters guarding
each side of a narrow strait. We meet them in Jason and the Argonauts, and in
Homer’s Odyssey. Charybdis is a poor creature so thirsty for salt water that she
swallows enough to make a horrific vortex:

You will find, Odysseus, the other rock lies lower, but they are so close together
that there is not more than a bowshot between them, and on it a great fig tree in
full leaf grows, and under it lies the sucking whirlpool of Charybdis. Three times
a day she belches forth the black waters, and three times she sucks them down
again; see that you be not there when she is sucking, for if you are, Poseidon
himself could not save you.
The Odyssey—Book 12
154 The Perfect Shape

Fig. 36.2 Detail from Carta Marina, Olaus Magnus, 1539, with the Maelstrom on
the coast of Norway

Clearly, Homer was not a very good oceanographer—a tidal Charybdis


should of course suck and belch twice a day, not thrice!

Breaking Waves: the Kelvin–Helmholtz Instability


Consider two liquids or gases, possibly of different densities, moving relative to
each other and separated by a sharp horizontal interface (Fig. 36.3). If there is a
small irregularity in this interface (and there will always be), it will be amplified
with time. Any small protrusion in the lower liquid will be pushed up and
pulled out downstream by the moving upper liquid. At the same time, a back
eddy forms behind each crest, moving liquid back in the opposite direction.
The result is a sequence of beautiful spiral waves with regular spacing, looking
like a classical “scroll” or Greek key pattern (Fig. 36.4). With time, these
waves will grow in size, and eventually they break up into turbulence. The
whole process, known as the Kelvin–Helmholtz instability, is responsible for
some spectacular natural phenomena; in particular the beautiful Kelvin–
Helmholtz clouds (Fig. 36.5). They are fairly common, but usually not very
36 Maelstrom 155

Fig. 36.3 Four stages in the formation of a Kelvin–Helmholtz instability. The


whorls will later break up into turbulence. Note the symmetry between the two
liquids

Fig. 36.4 Scroll pattern on the Norwegian Baldishol gobelin, ca. 1150 AD. Photo
Frode Inge Helland (cropped). CC-BY-SA-3.0 licence

Fig. 36.5 Left: Spectacular Kelvin–Helmholtz clouds over Boulder, Colorado, USA.
Photo Michael deLeon. Right: More typical Kelvin–Helmholtz formations over Lake
Mjøsa, Norway. A minute later the whorls had dissolved. Picture taken from paddle
steamer Skibladner (1856) by Marte Holten Jørgensen
156 The Perfect Shape

Fig. 36.6 Lava coils in the Athabasca region, Mars. Mars Reconnaissance Orbiter
(HIRISE) image, ca. 300 m wide. Courtesy of NASA/JPL/University of Arizona, public
domain

well formed. In rare cases, perfect but transient trains of Kelvin–Helmholtz


clouds can appear spontaneously in the sky, like enormous, breaking waves in a
stormy sea of air, a fitting theatre set for the Sun Ship. After a few minutes,
these prized trophies of weather photographers dissolve into chaos. Breaking
ocean waves are due to a related instability but where gravity also plays an
important role.
Some truly wonderful spirals turned up recently in images from Mars. Ryan
and Christensen (2012) identified 269 single and double spirals, 5–30 m wide,
in the Athabasca region close to the Martian equator (Fig. 36.6). This region
seems to be covered with solidified lava flows of fairly young age, <200 million
years old judging from the low density of impact craters.
The spirals from Mars look very odd, but similar ones have been observed
on Earth and are called lava coils. Ryan and Christensen explain them as
forming where underground lava flows of different speed are in contact,
producing a shear zone and a Kelvin–Helmholtz instability. This causes a
rotation in the overlying, solidifying lava, giving rise to spiral-shaped cracks.
Another idea might be that the spirals form in a similar way as in the sheets of
drying chemical precipitate studied by Leung et al. (2001). Small fragments of
the sheet detach from the substrate, starting from the edge. The resulting radial
tensile forces produce cracks in the tangential direction. The cracks propagate
inwards with the detachment front, producing a spiral.
36 Maelstrom 157

Leonardo’s Whirls
Leonardo da Vinci knew a lot about vortices in flowing water. Try to imagine
him, inserting planks of wood into the flow, creating eddies of the type we now
call Kármán vortex streets, observing them with the keenest eye and the brightest
mind that ever were. Among his many sketches of spiral eddies, some are
accurate scientific drawings while his late “Deluge” series is violent and
disturbing. Leonardo loved spirals in general, and they turn up everywhere in
his drawings of plants, human hair (which he compared with the movements of
water), clothing and mechanical devices, as well as in his writings (Fig. 36.7).

Fig. 36.7 Spirals in the art of Leonardo da Vinci. Top left: Study of falling water
(Royal Collection Trust/© Her Majesty Queen Elizabeth II 2016); Top right: Warrior
with helmet (©The Trustees of the British Museum. All rights reserved); Bottom left:
Study for the head of Leda, or is it Princess Leia? (Wikimedia Commons, public
domain); Bottom right: A deluge (Royal Collection Trust/© Her Majesty Queen
Elizabeth II 2016)
158 The Perfect Shape

Leonardo’s classification of spirals and eddies is a bit odd. His first type is
the level spiral, or a spiral in the plane. The second type is the convex spiral, a
helical spiral (trochospiral) with the tip pointing up. Conversely, there is also
the concave spiral, with the tip pointing down, as in a bathtub vortex. Finally,
the cylindrical spiral is a simple helix.
37
Treasures in the Mud

Take a cup of mud from the seafloor; your local beach will do nicely. Wash it
through a fine sieve, with a mesh of 0.1 mm or so (use a lot of water). Look at
the sandy residue through a stereo microscope. Some of the particles will be
mineral grains, but most of them will be exquisite pieces of biological archi-
tecture—the shells of foraminifera (Fig. 37.1). These are single-celled,
amoeba-like organisms building elaborate houses with many rooms, transpar-
ent like glass, white like porcelain, brown like leather, or grainy like sugar.
Fragile, ornamented bottles, long glassy needles, pearls on a string, paper
lanterns, branching trees, ugly lumps; the diversity in form among the thou-
sands of species is breathtaking, but the most common shape is the spiral.
Some of them are near-perfect logarithmic spirals (e.g., Planostegina), some
more Archimedes-like (e.g., Planispirillina) and many are complicated
trochospirals (Rotalinoides). The mathematical precision can rival that of an
ammonite or a snail. It is not easy to understand how these gorgeous shells can
be constructed by a single, slimy cell of so little wit.
Foraminiferan shells grow by successive addition of chambers. In the spiral
forms, these chambers are approximately gnomons, ensuring the overall sim-
ilarity of the shell during growth, much like a Nautilus or ammonite. In many
planktonic species, such as Globigerinella, the chambers are near spherical,
increasing in geometric progression and with a near constant angular
increment.
Paleontologists are very fond of foraminifera. They are extremely common
fossils, found in almost every little piece of marine sedimentary rock (except
very old ones). In modern petroleum and gas production, the well is often

© Springer International Publishing AG 2016 159


Ø. Hammer, The Perfect Shape, DOI 10.1007/978-3-319-47373-4_37
160 The Perfect Shape

Fig. 37.1 Some spiral foraminifera, most of them <1 mm long. Courtesy of Michael
Hesemann, foraminifera.eu, unless otherwise noted. Top row: Rotalinoides
gaimardi, Bulimina marginata (courtesy of Fabrizio Frontalini), Hyalinea baltica
(courtesy of Marianna Musco), Elphidium macellum, Globigerinella siphonifera.
Bottom row: Laevipeneroplis proteus, Planostegina costata (a giant, 3 mm long),
Planispirillina tuberculatolimbata, Spiroloculina excavata, Candeina nitida

drilled horizontally along the reservoir. At the rig, a paleontologist is busy


looking at the forams coming up. If the drill bit moves too far up or down, the
rocks will be a million years younger or older, the microfossils will be different,
and the drilling direction is adjusted. It is called biosteering—navigation by
evolution.
38
Subatomic Squiggles

Charles Thomson Rees Wilson liked clouds, and he liked to make them in his
laboratory. Wilson’s work was based on some curious experiments by the
Scotsman John Aitken in the 1880s. Aitken had injected water vapor into a
bottle. If the bottle contained filtered air, not much happened. But if the air
was unfiltered, water would condense on particles and form a fog. Aitken had
discovered cloud nucleation, a phenomenon that is much discussed in modern
climate research. He made even nicer clouds by expanding the volume of the
chamber with a piston. The rarefaction of the gas caused a temperature drop
and water would condense.
Wilson wanted to study effects of sunlight on clouds in the laboratory. He
started by repeating Aitken’s experiments, but Wilson was more thorough. He
discovered that when he increased the volume at a certain rate, fog would form
even in completely clean air. Rather than dismissing this as a fluke, Wilson had
the rare gift of serendipity. He guessed that the nucleation events were caused
by charged particles in the air. He started experimenting with X-rays and
radioactive materials, which increased the nucleation dramatically. In 1910, he
managed to photograph the paths of charged particles racing through the
chamber, leaving a trail of fog behind. Wilson just wanted to make a cloud
in a bottle, but now he had stumbled upon the cloud chamber, a tool for
studying the paths of individual subatomic particles. It started to smell like a
Nobel Prize in physics, which he received in 1927.
A further development was the bubble chamber, invented by Donald
A. Glaser in 1952. In a way, a bubble chamber is the opposite of a cloud
chamber. Instead of cooling a vapor, the bubble chamber allows subatomic

© Springer International Publishing AG 2016 161


Ø. Hammer, The Perfect Shape, DOI 10.1007/978-3-319-47373-4_38
162 The Perfect Shape

particles to vaporize a superheated liquid. The bubble chamber can detect


higher energy particles than the cloud chamber, and was the primary detector
of high-energy physics from the 1950s to the 1970s. Nowadays, the bubble
chamber has largely been superseded by vast and very complex electronic
particle detectors such as the ALICE and CMS of the Large Hadron Collider
at CERN.
Bubble chambers, like newer electronic detectors, are usually placed within
a strong magnetic field. The charged particles are deflected by this field
according to Lorentz’ law:
F ¼ qv  B

where F is the force vector imposed on the particle, q the charge of the particle,
v the velocity vector of the particle, and B the magnetic field vector. If you have
many electrons moving through a wire, and you put the wire into a magnetic
field, the wire will feel a force, and that is how you make an electric motor. The
interesting and odd part of this equation is the little x, which is the cross product
operator. The magnitude of the cross product depends on the magnitudes of
the input vectors and the angle between them, while the direction of the
product is given by the right hand rule: Use your right hand, point your index
finger in the direction of the first vector and bend it in the direction of the
second vector (you will probably have to rotate your hand). Now your thumb
points in the direction of the cross product, normal to the two input vectors.
That the force vector follows this rule, rather than a similar left hand rule,
reflects a fundamental asymmetry in the universe.
Now take a two-dimensional bubble chamber, look at it from above, apply
magnetic field lines in the downwards direction, and shoot a positron in from
the left (Fig. 38.1). According to Lorentz’ law, the positron will feel a force

Fig. 38.1 A positron with positive charge q (red path) enters a cloud chamber
(box) with a magnetic field directed perpendicularly into the paper (crosses).
The resulting Lorentz force F deflects the positron to the left, forcing it into a
circular path
38 Subatomic Squiggles 163

Fig. 38.2 In this bubble chamber experiment at CERN in 1960, negatively charged
pions were shot in from the left and interacted with protons in the fluid, producing
a spray of different particles. Image courtesy of CERN. CERN-EX-11465-1

deflecting it to the left. It will continue to move to the left, following a


counterclockwise circular path with the force always pointing towards the
center. A negatively charged electron, on the other hand, will be deflected to
the right and enter a clockwise circle. Because F ¼ ma, (Newton’s second law)
the radius of the circle will depend on the mass as well as the velocity of the
particle.
As the particle moves in its circular path, it will lose energy by so-called
bremsstrahlung and to a lesser extent by interactions with the fluid of the
bubble chamber. This causes the radius of the circle to decrease, resulting in a
delicate spiral (Fig. 38.2).
39
Nature Red in Blood and Claw

The menacing claws of eagles, lions or the dinosaur Deinonychus (meaning


“Terrible Claw”) are all near-perfect logarithmic spirals. The same goes for arm
hooks on squid, the teeth of saber-toothed cats, or just about any sharp,
piercing attack weapon used by predatory animals (Fig. 39.1).
Like for mollusk shells, an explanation can be sought in either functional or
developmental terms. Starting with the former, what could be the advantages
of a logarithmic spiral claw?
One intriguing answer was given by Mattheck and Reuss (1991). They
computed that a logarithmic spiral is the optimal shape for a claw in terms of
stresses under operation. The stress field in such a claw is completely even, so
that no particular point is more susceptible to failure.
Another advantage is rooted in the equiangular property. As the claw is
penetrating the prey, it would make sense to keep a constant angle of contact,
for several reasons including the minimization of lateral forces caused by
wiggling. Under this constraint, only a logarithmic spiral will maintain a
constant direction from a given point (the center of the spiral) to the contact
point as the claw moves into the victim. The force vector acting on the claw
can therefore maintain its direction. Of course, a straight claw or tooth has the
same property, but this is only a special case of the logarithmic spiral, with
infinite whorl expansion coefficient. In contrast, a circular shape would require
a continuous change in force direction, which would necessitate a much more
complex attack technique. The same arguments would apply to man-made
piercing weapons, but I have not yet found a convincing example of a

© Springer International Publishing AG 2016 165


Ø. Hammer, The Perfect Shape, DOI 10.1007/978-3-319-47373-4_39
166 The Perfect Shape

Fig. 39.1 Oviraptor dinosaur claw, ca. 15 cm long. Late Cretaceous, South Dakota.
Pole of the spiral located close to the tip (distally)

logarithmic spiral shape in any sword or dagger. Some variants of the Arabic
jambiya dagger come close, perhaps.
There is an interesting parallel with the construction of an old-fashioned
ship’s anchor, with a shank and two recumbent arms with flukes. One arm
digs in as the end of the shank is resting on the sea floor and dragged along by
the ship pulling away. The theory was worked out by none other than the
brilliant Swedish shipbuilder, vice admiral Fredrik Henrik af Chapman, in
1796 (Fig. 39.2). In a paper titled “Om rätta Formen på Skepps-Ankrar”
(On the Appropriate Shape for Ship’s Anchors) for the Swedish Royal Acad-
emy of Science, he calculated that the sum of the vertical force causing the
digging-in and the horizontal force resisting ploughing through the mud is
maximized if the angle between the sea floor (the direction of applied force)
and the arm is 67.5 . Such a constant angle of contact during penetration, as
the end of the shank is moved horizontally, can only be maintained by shaping
the arm like a logarithmic spiral. The expansion coefficient for Chapman’s
pffiffiffi
anchor is 2  1.
Most claws have the center of the spiral located distally (near the tip), and
the spiral then expands proximally towards the base. For this kind of spiral, the
appropriate action is by a pulling force, with the claw angled backwards and
attacking the prey from the side. The force exerted by prey trying to escape
away from the predator will then only cause the claws to dig in further. The
principle is very similar to that of Chapman’s anchor.
A few claws are oriented in the reverse direction, with the highest curvature
near the center of the logarithmic spiral, located proximally (near the base). A
dramatic example is the arm hooks of the Mesozoic squid-like cephalopods
known as belemnoids. Most of the hooks are small and numerous, but some are
enormous (Fig. 39.3) and occur in pairs. It is believed that these “mega-hooks”
39 Nature Red in Blood and Claw 167

Fig. 39.2 From F.H. af Chapman, 1796, Om ra€tta Formen pa° Skepps-Ankrar. Figures 3
and 4 show the anchor digging in as the end of the shank (A) is dragged horizontally
to the right

Fig. 39.3 A 4 cm long cephalopod arm hook from the Jurassic of Spitsbergen, with
the pole of the spiral near the base (proximally). From the collections of the Natural
History Museum, University of Oslo

belong to male specimens. Such hooks would be more efficient if moved by a


pushing action, thrusting the tip into the target. This is supported by the fact
that the tips of the hooks are directed away from the body of the belemnoid
animal, rather than back towards it (Hammer et al. 2013).
168 The Perfect Shape

A related grisly matter concerns the optimal shape of a rotary knife or a


slashing weapon such as a sabre. If the knife is in the shape of a logarithmic
spiral with the center coinciding with the axis of rotation, the blade will cut
with a constant shearing angle, which can be selected for best performance.
Developmentally, the logarithmic spiral claw or tooth is the easiest one to
make, as it only requires a difference in growth rate on opposite sides of the
base of the structure. In fact, often only the outer part of the claw is
logarithmic. The basal part can grow in a more complex way because it is
embedded in tissue that can resorb and remodel the structure.
40
Coffee, Kepler and Crime

Stir a cup of coffee and let the fluid settle into a quasi-steady rotation.
Then introduce a straight stripe of milk from the center to the wall of
the cup (very hard to do in practice—I have tried!). Now if the angular
velocity is the same everywhere, the stripe will not be deformed but just
rotate as a solid body like the minute hand of a clock. The coffee is then
said to be spun up. But as it is spinning down, the coffee will rotate more
slowly near the outer wall, because of friction. The stripe of milk will
then deform into a spiral, as the outer part lags behind the inner. You have
probably observed similar spirals in your coffee.
So let us say that the angular velocity, in radians per second, is some
function of radius r:

¼ f ðr Þ:
dt

Integrating in time, it is clear that the angle along which the milk stripe has
moved after a time T is
ϕðr; T Þ ¼ Tf ðr Þ:

Solving for r, we can give the equation for the resulting spiral:
 
ϕ
r ¼ f 1 :
T

© Springer International Publishing AG 2016 169


Ø. Hammer, The Perfect Shape, DOI 10.1007/978-3-319-47373-4_40
170 The Perfect Shape

However, it is not obvious what the velocity profile f (r) would be in the cup of
coffee. Instead of going down that difficult avenue, let us try a logarithmic
profile where velocity is increasing with radius:
f ðr Þ ¼ lnr, r > 0:

Inverting this function and plugging into the previous equation, we get the
shape of the milk stripe
1
r ¼ eTϕ ;

which is a logarithmic spiral where the expansion coefficient is large for small
T, but reduces with time so that the spiral gets more and more tightly wound
(Fig. 40.1). Similarly, a linear angular velocity profile, i.e., f (r) ¼ r, gives an
Archimedes spiral where the distance between whorls reduces with time,
r ¼ φ/T.
What we have learned from coffee can be applied to astrophysics. Consider
a so-called bar galaxy, where stars are concentrated along a straight line out
from the center. How will this line be deformed as the galaxy is rotating? It
would be reasonable to assume that the stars orbit around the galactic center
according to Kepler’s Third Law from 1619, which says that the square of the
orbital period is proportional to the cube of the orbital radius (or semimajor

Fig. 40.1 Left: An initial straight line (black) is moved by a rotating fluid where
angular velocity increases as the logarithm of radius. The line is progressively
deformed into logarithmic spirals with decreasing expansion coefficient. Right:
Angular velocity follows Kepler’s Third Law, with angular velocity decreasing
with radius
40 Coffee, Kepler and Crime 171

axis, to be precise). The orbital period is inversely proportional to angular


velocity f (r), so we can write
 
1 2
/ r3 ;
f ðr Þ

or

f ðr Þ / r 3=2 :

Using the machinery above, the resulting spiral will have the form
 2=3
T
r¼ :
ϕ

This spiral, similar to a hyperbolic spiral, is plotted for increasing T in


Fig. 40.1. Could a spiral galaxy form by such deformation of a bar galaxy?
There are problems. First, I don’t think these spirals look particularly
much like the arms of a spiral galaxy such as M51. Also, observations show
that orbits of stars around a galactic center do not follow the Third Law at all.
In fact, the velocity in km/s is almost constant for large radii (Fuchs et al.
1998), which is very strange and often explained as an effect of Dark Matter. I
leave it as an exercise for the reader (a singularly irritating phrase common in
mathematics and physics texts) to figure out the resulting spirals.
There is a distortion filter in Photoshop called “twirl”, working on exactly
this principle. It moves pixels in the image by an angular displacement as a
function of radius (Fig. 40.2). For a large pixel resolution this is effectively a
so-called bijective map, meaning that each point in the original corresponds to

Fig. 40.2 Three stages in a rotational deformation of Mona Lisa, made with
Photoshop
172 The Perfect Shape

exactly one point in the distorted image, and vice versa. Such a map is
invertible, so you can simply run it in reverse to reconstruct the original
image. In a famous case from 2007, the invertibility of the twirl map became
the downfall of child molester Cristopher Paul Neil. He posted more than
200 images of himself in incriminating situations on the Internet, with his face
obscured by the twirl filter. Experts in the German police could easily invert
the distortion, with the original face reappearing as if by magic. Neil was
arrested and sentenced to 6 years in prison because of his lack of skills in set
theory.
41
Dürer’s Dirty Secret

Albrecht Dürer (1471–1528) was the Leonardo of Germany—a brilliant artist


but also proficient in geometry. In the first book of his Underweysung der
Messung mit dem Zirckel und Richtscheyt (Instructions for Measuring with
Compass and Ruler), first published in 1525, he writes at length about
different spirals and helices, with beautiful illustrations. Most of them are
derived from Archimedes spirals. But there is also a small, very interesting
paragraph, largely overlooked by the historians of spirals, where Dürer con-
structs a logarithmic spiral, more than a century before Descartes. Hardly a
mathematical masterpiece, it is rather a creative but vague idea from the mind
of an artist. Here is the quaint German text:

Es mag ein ewige lini erdacht werden/die da stettiglich zu eim Centrum einwartz/
. . . /vnd nimer mehr zu keym end kombt/Dise lini kan man mit der hand der
vnentlichen grosse vnd kleyne halben nit machen/Dann ir anfang vnd end so sie
nit sind/ist es nit zu finden/das fast allein der verstand/Aber ich will sie vnden mit
eim anfang vnd end/so vil dan muglich ist anzeige/Ich heb an bey eim punckten.
a. vnd zeuch dise lini zirckelsweis hinein/als solt sie zu eim Centro lauffen/vnd so
offt sie in eynander laufft brich ich der weiten zwische der lini ein halbteyl ab/des
gleichen thu ich/so ich mit der lini vom.a. heraus lauff/so offt ich mit ir vber
eynander lauff/so offt gib ich der lini ein halbteyl zu/von der weyten/Also laufft
dise lini ye lenger ye enger hinein/vnnd lenger ye weyter heraus/vnd kumbt doch
nimer meer zu keim ende/weder hinein noch heraus wie ich das zu verstehen hie
vnden hab aussgeriessen.

© Springer International Publishing AG 2016 173


Ø. Hammer, The Perfect Shape, DOI 10.1007/978-3-319-47373-4_41
174 The Perfect Shape

In English:

One might imagine a never-ending line, which steadily moves towards a center
. . . and never comes to an end. Due to its infinite greatness and smallness we
cannot produce this line by hand, for its beginning and end are not to be found
as they do not exist, except in the mind. But below I will show it with a
beginning and end, as far as possible. From a point a, I draw this line in circles
therefrom as if running towards a center, and every time it runs into itself I
reduce the width between the lines by half. The same I do so that when drawing
the line outwards from a, every time it runs above itself I give the line another
half in the width. So this line gets narrower the further it runs inwards, and wider
the further it runs outwards, but never comes to an end, neither inwards nor
outwards as I have sketched below for understanding.
(Transl. the author and Franz-Josef Lindemann)

It is not very clear, is it, but my interpretation is that Dürer makes a point a,
draws a line inwards so that the distance between succeeding whorls is halved,
and outwards so that the distance between succeeding whorls is doubled. He
provides a sketch that seems to fit the bill roughly (Fig. 41.1).
It is not difficult to show that Dürer’s infinite line is a logarithmic spiral
with expansion coefficient k ¼ ln2=2π, or about 0.1103. The computer plot
does not quite resemble his sketch, though. Dürer pulled a fast one. He
thought he would get away with it, never suspecting his imprecise drawing
would be checked by an electronic calculating engine half a millennium into
the future.
Dürer is obviously thrilled by the fact that the spiral “never comes to an end,
neither inwards nor outwards”. He may have been the first victim of the
logarithmic spiral memetic virus that would return so epidemically a century
later. But we have seen earlier that although the spiral never ends, it has finite
length. Going inwards, Dürer’s spiral reduces in size by half for every whorl.

Fig. 41.1 Left: Dürer’s sketch of his “ewige lini”—infinite line. Right: The
logarithmic spiral r ¼ e0.1103φ, plotted following Dürer’s instructions
41 Dürer’s Dirty Secret 175

Let us say that the first whorl has length 1/2. The next whorl in has length 1/4,
the next has length 1/8, etc. The total length Ln of the n first inwards whorls is
then the sum of the geometric series
1 1 1 1
Ln ¼ þ þ þ ... þ n:
2 4 8 2

To find a closed expression for Ln, we use an old trick. First multiply by 2:
2 2 2 2 1 1 1 1
2Ln ¼ þ þ þ . . . þ n ¼ 1 þ þ þ . . . þ n1 ¼ 1 þ Ln  n :
2 4 8 2 2 4 2 2

Then subtract Ln from both sides:


1
Ln ¼ 1  :
2n

Now let the number n of inwards whorls go to infinity. The total length of the
spiral will then approach 1. An infinite number or whorls, but a finite total
path length. This is reminiscent of Zeno’s so-called dichotomy paradox. Zeno
(ca. 490–430 BC) argued that motion is impossible, and a runner cannot get
to the end of a racetrack, because first he must get half way there, then half way
of the remaining distance (i.e. 1/4), then 1/8, 1/16 etc. We don’t know exactly
why Zeno thought this impossible. Perhaps he believed that the sum of all
these terms is infinite, which is plain wrong, as we have seen. Or, more
interestingly, he believed that it is impossible to carry out an infinite number
of tasks, however small, in finite time. In any case, I quite like the answer of
Diogenes the Cynic (ca. 410–323 BC), although somewhat lacking in math-
ematical rigor: Saying nothing, he just stood up and moved!
42
The Spiral from the Depth of Time

Paleontology is the study of miraculously preserved remains of ancient


life. And the story of the spiriferid of the Barents Sea is as miraculous as
it gets.
About 260 million years ago, a small brachiopod was patiently filtering
seawater for tiny food particles in what is now the Arctic Barents Sea. The
brachiopods belong to an almost extinct phylum of animals, vaguely clam-like,
but perhaps even more beautiful. They prospered in the Paleozoic but are
relatively rare in modern oceans. It was warmer then, before the mammals,
before the dinosaurs, and the cataclysmic mass extinction at the end of the
Permian was not yet on the horizon.
This little brachiopod, only a few millimeters across, lived its little life like
trillions of other animals that have struggled for existence through the history
of life on Earth. For hundreds of millions of years the dead shell was buried
further and further down, deep under the floor of what much later became the
North Atlantic Ocean. Far into the Earth’s crust it found its grave, under the
stormy Barents Sea and under more than a kilometer of rock. Then, in 2015,
Lundin Petroleum was drilling for oil and gas with the semisubmersible rig
“Island Innovator”. A small plug from a depth of 1852 m (almost 1500 m
below the sea floor) was given to me for CT scanning. From inside this plug, a
brachiopod emerged on the computer screen; a virtual 3D reconstruction
made from 3000 high-resolution X-ray images and elaborate mathematical
calculations (Fig. 42.1).
This particular brachiopod was a spiriferid, which is why it gets a place in
our book. The spiriferids are characterized by their magnificent spiral-shaped

© Springer International Publishing AG 2016 177


Ø. Hammer, The Perfect Shape, DOI 10.1007/978-3-319-47373-4_42
178 The Perfect Shape

Fig. 42.1 Spiral-shaped feeding apparatus (brachidium) in a Permian spiriferid


brachiopod, well 7220/3-1, Barents Sea. Sample provided by Lundin Petroleum, CT
image by the author

Fig. 42.2 The spiriferid brachiopod Spiriferina pinguis from the Jurassic of France,
ca. 4 cm across. The shell is partially broken, revealing the brachidium within. From
the collections of the Natural History Museum, University of Oslo

brachidium, which supported their feeding apparatus (Fig. 42.2). The


brachidium is beautifully preserved in the specimen from the Barents Sea. It
is a spiral from the abyss of time. It comes from an animal of another age, from
a different world, recovered in a thin core from 1500 m under the ocean floor.
It is perhaps the most thought-provoking spiral in this book.
43
Propelling, the Archimedean Way

The Archimedes screw, consisting of a helix (or helicoid, to be precise) inside a


tube, is an efficient machine for lifting water (Fig. 43.1). Its elegant design has
a distinctly “modern” look, and it is still the mechanism of choice for some
engineering applications. Surprisingly, the Archimedes screw was in common
use already in classical times, by the Greeks, Romans, and Egyptians. It is
described in detail in several classical texts, including Vitruvius’ De Architectura,
and the invention was ascribed to Archimedes by Diodorus of Sicily (first
century BC).
Unfortunately, the screw is not mentioned in any surviving text by Archimedes
himself. The classical references are partly ambiguous, and may be totally wrong,
about the identity of the inventor. Perhaps it was a mix-up of several things—the
universal admiration of Archimedes at the time, the fact that he had written a paper
on spirals (although, as we have seen, a spiral is quite distinct from a helix), and not
least that he lived in Alexandria at a time when the “Archimedes” screw was in
common use there. This last fact is particularly interesting. Could it be that the
Egyptians already knew how to use the screw, and Archimedes merely learned it
from them? If so, the Egyptians may in turn have learned it from the Assyrians.
350 years before Archimedes, the Assyrian king Sennacherib described an irriga-
tion device that could have been an Archimedes screw, although the text is short
and open to interpretation. Also, the Greek geographer and historian Strabo
(ca. 64 BC–21 AD) writes that the Archimedes screw was used in the Hanging
Gardens of Babylon, but it is not clear whether he is talking about a later time
when the Archimedes screw might already have traveled from the West to the East.
These matters are discussed in an entertaining paper by Dalley and Oleson (2003).

© Springer International Publishing AG 2016 179


Ø. Hammer, The Perfect Shape, DOI 10.1007/978-3-319-47373-4_43
180 The Perfect Shape

Fig. 43.1 Derelict Archimedes’ screw from 1914 in Kirkøy, Norway. The outer cover
is missing, revealing the 6 m long screw. The device was originally driven by a wind
mill, and was built as part of a failed attempt to drain the small lake Arekilen in
order to reduce malaria. Photo Marte H. Jørgensen

A standard drill bit is an Archimedes screw. The helical grooves are not there for
excavating the hole, but for transporting the dust and chips out of it. The
Archimedes screw was well known in Antiquity, and the screw was one of the
five Simple Machines, the building blocks of mechanics, listed by Heron of
Alexandria. Still, there is only scattered evidence of twisted drill bits until the late
eighteenth century, and the first influential patent for a screw auger was filed by the
Connecticut shipbuilder Ezra L’Hommedieu as late as in 1809 (Mercer 1929). It is
a right-handed double helicoid (Fig. 43.2). Next time you use a drill, give a
moment’s thought to L’Hommedieu. Without him, your job would be a night-
mare. It is amazing that Watt’s steam engine, Harrison’s chronometer, Volta’s
battery and differential calculus were all invented before a practical hand drill!
The famous helicopter designed by Leonardo da Vinci, although clever, is
more funny than useful (Fig. 43.3). It might fly if it could be made much
lighter than allowed by materials available at the time, but only if powered by
an external operator. Lacking a reaction force, as provided by the tail rotor in a
modern helicopter, it would not be possible to operate it with an onboard
engine. Leonardo himself refers to the rotor as a helix, and it is fair to assume
that the Archimedes screw was a major inspiration—the helicopter pushing air
much like the screw lifts water.
The Archimedes screw was not to be used for practical propulsion until the
nineteenth century. The first proper steam ship to be fitted with a screw was
the aptly named SS Archimedes, in 1839 (Fig. 43.4). It was an immediate
success, and the propeller would soon outcompete the cumbersome and
43 Propelling, the Archimedean Way 181

Fig. 43.2 Screw auger, from US Patent 1,114X, L’Hommedieu 1809. Source: United
States Patent and Trademark Office, www.uspto.gov

Fig. 43.3 Flying machine, Leonardo da Vinci, 1493. The shading direction indicates
that Leonardo was left-handed. Public domain, Wikimedia Commons
182 The Perfect Shape

Fig. 43.4 The screw of SS Archimedes, 1839. Illustrierte Zeitung, Leipzig, 1843.
Public domain, Wikimedia Commons

Fig. 43.5 Left: Counterclockwise rotation of flagella, forward motion. Right:


Clockwise rotation, tumbling motion

inefficient paddlewheel. The original design was a proper, complete, single-


turn helicoid, but it was replaced after only a year with a more modern-looking
two-bladed screw to reduce vibration.
The modern ship or airplane propeller is not quite an Archimedes screw of
course; it has a much more complex shape. Still, the turboprop engine and the
computer cooling fan have deep roots through the millennia, back to the
beautiful water elevators of Rome.
Or further back. Much further back. Propulsion by a rotating helix was
invented billions of years ago, by bacteria. The working principle of the bacterial
flagellum, a left-handed helix rotating at up to 1000 rounds per minute, is the
same as that of the SS Archimedes screw. Escherichia coli and some other species
have a funny twist on it. Instead of one flagellum, they have six (Fig. 43.5).
When rotating them counterclockwise, the flagella combine into a bundle and
the bacterium swims forward. Clockwise rotation of some or all flagella forces
the flagella apart in all directions, causing the bacterium to tumble so that it can
turn to swim in another direction. It is not very elegant, but it works.
44
Unwrapping Mummies

A black, infinitely thin string is wrapped around a red circle (Fig. 44.1). Now
unwrap the string from the circle while holding it taut. The end of the string
will follow a spiral path outwards. This path is called the involute of the circle.
The Cartesian coordinates for the involute of a circle with radius a and a
polar angle φ are
x ¼ aðcos φ þ φsin φÞ
y ¼ aðsin φ  φcos φÞ

Now clearly, for large φ, these equations will converge on those describing an
Archimedes spiral, i.e.,
x ¼ aφsin φ
y ¼ aφcos φ

In other words, the involute of a circle becomes almost identical to an


Archimedes spiral far out from the pole (Fig. 44.2).
The Archimedes spiral has a constant distance between the whorls measured
along radii, while the involute spiral has constant distance as measured normal
to the curve. Almost the same thing, but not quite. The involute is more similar
to the shape of a coiled-up rope with constant thickness.
Coiling up long rope-like structures is a common source of involute spirals
in Nature. I know of no more stunning example than the colonial stage of the
salp Pegea of the Pacific Ocean (Fig. 44.3). The salps are weird, sac-like
tunicates swimming by jet propulsion. The tunicates are closely related to

© Springer International Publishing AG 2016 183


Ø. Hammer, The Perfect Shape, DOI 10.1007/978-3-319-47373-4_44
184 The Perfect Shape

Fig. 44.1 Involute of a circle, with some positions of the string

the vertebrates, and also include the “sea squirts” sitting on rocks around the
coasts of Europe. The salps switch between living solitary lives and arranging
themselves in large colonies called chains. A Pegea chain consists of two parallel
rows. New individuals are added continuously by asexual reproduction (bud-
ding). Now imagine such a chain wiggling in random twists and turns. The jet
engines would point in all directions, and the net thrust would be close to zero.
Much wiser then to coil up into a spiral, ensuring that all the water guns are
aligned, shooting normal to the plane of the spiral.
The involute of a circle is important in mechanical engineering. Gears are
fundamental to modern society, and the profile of gear teeth is critical to
ensure smooth, efficient, noise free operation. The challenge is to achieve a
constant angular velocity of the driven wheel as the tooth engages and releases.
The curve known as the cycloid has this property, but most modern gears use
instead a segment of the involute of a circle, giving a characteristic convex
profile to the sides of the teeth (Fig. 44.4). This important invention can be
traced back to none other than the brilliant mathematician Leonard Euler
(1707–1783). It is remarkable that even the nameless, mechanical genius who
constructed the incredible Antikythera mechanism, the Greek astronomical
calculator from 200 to 100 BC, used triangular gear teeth. The resulting
44 Unwrapping Mummies 185

Fig. 44.2 Archimedes spiral with a ¼ 1 (black) and the involute of a circle with
radius 1 (red) become indistinguishable after <1 whorl. Program code in
Appendix B

jerkiness must have contributed to inaccuracy in the predictions of the solar


and lunar movements.
Now, of course, there is no reason to stop with the involute of a circle. We
can take any shape and form its involute. The involute of a (two-dimensional)
mummy is described by the end of its bandage as you unwrap it. Or tie an
eager puppy with a long rope to the corner of her square dog house, and let her
run around the house a few times while keeping the rope taut (Fig. 44.5). As
the rope wraps around the house, the puppy will describe the involute of a
square, which consists of a sequence of connected quarter circles with reducing
radius.
The involute of a square is not a very pretty spiral is it, looking rather
lopsided and irregular. We will return to it later. Let us instead try something
crazy: What is the involute of a logarithmic spiral? The result is shown in
Fig. 44.6, and it is surprising. The involute is another logarithmic spiral—in
fact, the two spirals have the same whorl expansion rate and only differ by a
rotation. Jakob Bernoulli discovered this—yet another beautiful property of
the logarithmic spiral.
186 The Perfect Shape

Fig. 44.3 Chain of Pegea salps, colony ca. 30 cm across. Photo Mike Hallack

Fig. 44.4 Left: Involute gears in water mill, ca. 1900. Moss Town- and Industry
Museum, Norway. Photo Marte H. Jørgensen. Right: A gear wheel than cannot
work: The rectangular teeth would jam instantly. Sculpture by Tom Erik Andersen,
Øvre Eiker, Norway
44 Unwrapping Mummies 187

First quadrant,
r=7m
Second quadrant,
r=6m

1m

1m

7 m rope

Fig. 44.5 A puppy on a rope, running around her doghouse, describes the involute
of a square

Fig. 44.6 A logarithmic spiral (black) and its involute (red), with some positions of
the string
45
Pagan Coils

Neolithic and Bronze Age Europeans were crazy about spirals. The spiral is so
pervasive in rock art, jewelry and architecture that it characterizes and unites
cultures from Scandinavia to Egypt through several thousand years. It is a
super-symbol, turning up everywhere, in ever-changing configurations and
combinations. Scholars debate endlessly about its meaning—perhaps it
represented the sun, perhaps it had a botanical origin, perhaps it signified
ocean waves, or it was a symbol of eternity, or illustrated shamanistic experi-
ences, or was just for decoration. Quite likely, several of these played a role,
perhaps changing in relative importance over time. In any case, the massive use
of spirals, often in ritualistic settings, indicates that they often carried some
deep, mystical meaning, now forever lost by the passage of millennia and the
victories of newer religions.
Any visit to an archeological museum will demonstrate the importance of
the spiral in prehistory and antiquity, and I will give only a few examples.
The interconnected spirals at the Tarxien megalithic temple, Malta, dated
to about 3200 BC, are so beautifully carved that they seem anachronistic,
something you would rather expect from much younger, classical Greek or
Roman art (Fig. 45.1). The clever, balanced design contains both left-handed
and right-handed spirals, with embellishments reminiscent of plants. The
similarity to the much later Greek “key” or meander pattern is intriguing,
although the topology is different: the Tarxien pattern contains a series of
joined single spirals while the Greek key is a single meandering line (Fig. 45.2).
It is ironic that the Greek key, a symbol so strongly associated with culture,

© Springer International Publishing AG 2016 189


Ø. Hammer, The Perfect Shape, DOI 10.1007/978-3-319-47373-4_45
190 The Perfect Shape

Fig. 45.1 Tarxien temple, Malta, ca. 3200 BC. Photo Berthold Werner. Creative
Commons Attribution-Share Alike 3.0 Unported licence

Fig. 45.2 Greek key pattern

beauty and early democracy, has been turned into an ominous-looking Nazi-
style banner by the Greek fascist party, the Golden Dawn.
The spiral patterns in the Neolithic passage tomb mound at Newgrange,
Ireland (ca. 3200 BC) are far less refined than those at Tarxien, but not
without complexity (Fig. 45.3). Some of them occur in interlocking,
labyrinthic triplets known as triskeles.
The famous and mysterious Phaistos disk from the Minoan culture in
Crete, probably dating from the 2nd millennium BC, is another
Archimedes-like spiral (Fig. 45.4). The text has never been deciphered.
Archaeologists are debating whether the spiral was made inwards or outwards;
it would probably be easier to make the turns of equal width if starting from
the center, but the symbols are sometimes overlapping in a way that indicates
the opposite. Perhaps the spiral itself was made outwards, and then the writing
was filled in subsequently in the inwards direction.
45 Pagan Coils 191

Fig. 45.3 The entrance stone of the passage tomb at Newgrange, Ireland, ca. 3200
BC. Note the spiral triskele at the left. Photo Jal74. Creative Commons Attribution-
Share Alike 4.0 International license

Fig. 45.4 The Phaistos disk, side A. Photo C. Messier. Creative Commons
Attribution-Share Alike 4.0 International license
192 The Perfect Shape

Fig. 45.5 Left: Spiral bronzes, ca. 7 cm wide, Norwegian Bronze Age. Photo Per
E. Fredriksen, NTNU Vitenskapsmuseet, licence CC BY-SA 4.0. Right: Replica of
tutulus from Rogaland, Norway, ca. 1500–1100 BC. Photo Ørjan Engedal, www.
bronsereplika.no

The Bronze Age could also be called the age of spirals, at least in Scandina-
via, where they were absolutely everywhere. An example is the spectacle fibula,
a brooch made of a single bronze wire twisted into two near-Archimedes spirals
(involutes of a circle, to be precise). They were all the rage around the ninth
century BC (Fig. 45.5).
Spirals are also common in Scandinavian Bronze Age rock art. In Norway,
an interesting symbol with a pair of spirals inside a circle (Fig. 45.6), some-
times stowed on a ship, may represent the sun; or is it a face?
Near Sandefjord in Norway, the Bronze Age artist was so preoccupied with
spirals that he could not even depict humans without them (Fig. 45.7). The
eight large, irresistibly cute anthropomorphic figures at this site illustrate
spiralomania at its most extreme.
The bow of the Norwegian Oseberg ship from about AD 820 is formed as
an elegantly tapered spiral serpent, about 60 cm across, continuing seamlessly
into the hull (Fig. 45.8). The spiral is slightly stretched in the vertical
direction, including a widening of the snake’s body below the head, and
even a relative shortening of its segments there. When seen obliquely from
below, the distortion seems to disappear. Whether this illusion was inten-
tional is of course impossible to know, but if so then it reveals an impressive
attention to detail and an advanced understanding of perspective and
projection.
The Viking ships were painted red, sings Torbjørn Hornklove in his
Haraldskvæði, ninth century AD:
45 Pagan Coils 193

Fig. 45.6 Bronze-Age twin spiral, Hafslund, Norway

Fig. 45.7 Spiral Man, Haugen, Sandefjord, Norway


194 The Perfect Shape

Fig. 45.8 The dragon head of the Oseberg ship, partly reconstructed. The drawing
is copyright Kulturhistorisk museum, University of Oslo, CC BY-SA 4.0

dróttin Norðmanna, lord of Norway,


djúpum ræðr kjólum, commanding deep keels,
roðnum ro˛ngum, reddened stems,
ok rauðum skjo˛ldum, and red shields,
tjo˛rguðum árum, tarred oars,
ok tjo˛ldum drifnum. and tents with ocean foam.

These exquisite ship spirals were not a Viking invention. They are known
already from Scandinavian rock carvings from the Bronze Age, around 1000
BC. A particularly beautiful example from Begby in Norway shows a ship
with an elegantly tapering stern, continuing smoothly into a splendid spiral
(Fig. 45.9). The ship sails from left to right. This is the Day Ship, according
to current theory developed in detail by Danish archaeologist Flemming Kaul
based on a large database of Bronze Age iconography. The Day Ship navi-
gates the sky from East to West, carrying the Sun, or being the Sun;
sometimes the Day Ship is pulled by a horse, sometimes it is carried by a
giant, sometimes it takes turns with the Sun Horse pulling the Sun Chariot.
At sunset, the Divine Snake loads the Sun from the Day Ship onto the large,
sinister Night Ship, sailing under the dark sea from West to East, right to left,
perhaps carrying the souls of the dead. At dawn, the Sacred Fish loads the
Sun back onto the Day Ship, celestial herdsmen summon the Horse, and the
cycle repeats. It is a majestic, beautiful myth with deep roots. The parallel
45 Pagan Coils 195

Fig. 45.9 The stern of the Day Ship. Rock carving at Begby, Norway, ca. 1000
BC. Photo Marte H. Jørgensen

with the Egyptian solar barge of Ra seems hardly coincidental. Perhaps the
spiral itself symbolized the wheel of night and day, the incomprehensible but
comforting repeatability of the Cosmos, the eternal diurnal and annual
cycles, predictability and order.
This wonderful religion of spirals was celebrated in processions carrying
helical bronze lurs, glittering in the sun, in pairs coiling dextrally and
sinistrally, like the horns of the Sacred Bull of Crete, their mighty bass tones
traveling for kilometers across the open landscape. Spiral fibulas, gold wire
helices adorning the hair, the adoration of the spiral beauty of Nature, the
Infinite Cycle, death and resurrection. Those were the days.
And then there is Quetzalcoatl. The God of Wind, the God of Knowledge,
the Inventor of Books, the Feathered Serpent of the Aztecs, the Kukulkan
and Gukumatz of the Maya, the morning star Tlahuizcalpantecuhtli, the
Spiral God. Conceived when the virgin Chimalman swallowed an emerald,
or was visited by Onteol in a dream, or was hit by an arrow in the womb, or
born by Coatlicue, the mother of the stars of the Milky Way; around his neck
he wears the ehecailacocozcatl, the wind jewel, a central symbol of Mesoamer-
ican pre-Columbian culture (Fig. 45.10). The wind jewel is a logarithmic
spiral, a cross section of a snail shell, usually Strombus (de Borhegyi 1966) or
196 The Perfect Shape

Fig. 45.10 Quetzalcoatl with the wind jewel on his chest, from the Codex
Borbonicus. Illustration Eddo, Creative Commons Attribution 3.0 Unported

Fasciolaria (Tozzer and Allen 1910). Such cross-cut snails are found at
archeological sites (de Borhegyi 1966) and were probably worn by priests.
The crosiers of medieval Europe are similar to the bow of the Oseberg
ship, often in the form of a serpent (Fig. 45.11). But although they are
beautiful, they are usually less elegant, with a stiff-looking join between
spiral and staff. Across the join, the position is of course continuous, and
45 Pagan Coils 197

Fig. 45.11 Crosier with serpent eating a flower, France, 1200–1220 AD. Metropol-
itan Art Museum

the direction nearly so, but the curvature is clearly not, with the zero-
curvature staff turning almost directly into a high-curvature spiral. Conti-
nuity of curvature is essential for aesthetics. The mathematical curves
known as splines, which are used everywhere in computer graphics (includ-
ing the generation of elegant letter fonts), are constructed precisely to
ensure continuity of curvature. The Vikings got it right, partly because
they constructed their curves by bending wood, which leads to minimal
curvature automatically.
198 The Perfect Shape

Don’t Drive Along a Crosier


Imagine sitting on a train driving along this Bishop’s crosier. As you enter the
spiral, the curvature increases suddenly, and you will feel an unpleasant jolt
because of the change in centripetal force. Clearly, it is important when
designing a railway track to ensure continuity of curvature. There is a partic-
ular spiral, called Euler’s spiral, where curvature increases linearly inwards as a
function of path length. If you drive a car and turn the steering wheel at a
constant rate, you will describe Euler’s spiral. This spiral is a little tricky,
because there is no simple formula for it—it cannot be expressed in terms of
the usual mathematical functions. Instead, its x and y coordinates as a function
of a parameter t is defined by the so-called Fresnel integrals:
ðt
xðtÞ ¼ cos s2 ds
ð0t
yðtÞ ¼ sin s2 ds
0

These integrals must be evaluated numerically, or by some clever approxima-


tions. If we plot this for positive t, we get a nice-looking spiral (Fig. 45.12).
At the outer end of this spiral, for t ¼ 0, the curvature is zero. Now a smart
trick for constructing a curve on a railway track is to use a segment of an Euler

Fig. 45.12 Euler’s spiral for positive t. Program code in Appendix B


45 Pagan Coils 199

Circular
curve

Straight track Euler easement

Fig. 45.13 Easement of railroad track into a curve using the Euler spiral (black)

Fig. 45.14 Segment of Euler’s spiral with positive and negative branches, rotated

spiral to ease (that is the technical term) from the straight track into the circular
curve with a given radius (Fig. 45.13).
The procedure would be used in reverse for easing from the curve back into
a straight track. This is not just esoteric mathematics—Euler spirals are
actually used for constructing railway easement. Think about it next time
you sit on a train and your coffee cup doesn’t spill in a curve. It may be thanks
to those Fresnel integrals.
200 The Perfect Shape

And do compare Euler’s spiral with the bow of the Oseberg or Begby ship.
The similarity is uncanny. I’m not saying that the ancients knew about Euler’s
spiral, but they certainly knew how to draw a curve with the smoothest
possible change of curvature (Fig. 45.14).
46
A Note on Toilet Paper

Some people do not know the path length of an Archimedes spiral or the
involute of a circle. It is difficult to understand how they can manage in daily
life. How can they know how many rolls to buy if they need a kilometer of
toilet paper? How can they know how large to make the sheet of dough to roll
up into cinnamon buns? The length of the groove on an old LP record?
The equation for the path length of an Archimedes spiral r ¼ aφ is found
using the general equation for ds which is derived in Appendix A.2:
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
 2ffi qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi pffiffiffiffiffiffiffiffiffiffiffiffiffiffi
dr
ds ¼ r 2 þ dφ ¼ ðaφÞ2 þ a2 dφ ¼ a 1 þ φ2 dφ:

Integrating this simple expression is surprisingly difficult—I suggest we cheat


and consult a table of integrals. The resulting path length as a function of total
angle is
 qffiffiffiffiffiffiffiffiffiffiffiffiffiffi  qffiffiffiffiffiffiffiffiffiffiffiffiffiffi
a
sðϕÞ ¼ ϕ 1 þ ϕ2 þ ln ϕ þ 1 þ ϕ2 :
2

Let us consider a toilet paper roll with no core (it is easy to extend the argument
to the case with a core), with an R ¼ 50 mm radius. Assume the paper is
h ¼ 0.1 mm thick. How long is the paper? The number of turns of the spiral
is R/h ¼ 500, so the full rotation angle is φ ¼ 2π  500, or 1000π radians. We
also need the parameter a in the equation for the Archimedes spiral—it is a ¼ h/2π.
Using the equation above, we get s(1000 π) ¼ 78,539.890 mm, or 78.539890 m.

© Springer International Publishing AG 2016 201


Ø. Hammer, The Perfect Shape, DOI 10.1007/978-3-319-47373-4_46
202 The Perfect Shape

We have an answer, but it is not quite correct. The reason is, as we have
discussed in an earlier chapter, that the true curve describing a spiral with a
constant distance between whorls, measured normal to the curve, is not the
Archimedes spiral but the (very similar) involute of a circle. The equation for
the path length of an involute circle is surprisingly simple:
b
sðϕÞ ¼ ϕ2 :
2

Here, b is the radius of the generating circle. For each round of unwinding the
string defining the involute of the circle, the radius of the involute increases by
2πb. This increase should correspond to the thickness of the paper, so b ¼ h/
2π. The length of the toilet paper is calculated as the length of the involute of
the circle: s ¼ 78,539.816 mm. The correction of 74 μm from the Archimedes
solution is not to be sneezed at, and the calculation was easier, too!
We can also express the paper length as a function of whorl distance h and
number of turns n:
b h=2π
s ¼ ϕ2 ¼ ð2πnÞ2 ¼ πhn2 :
2 2

Unfortunately for us spiral enthusiasts there is a very elegant way to achieve the
same result with no knowledge of either Archimedes spirals or involutes. Con-
sider the toilet roll as seen from the side (Fig. 46.1). Although it is not exactly a
circle because the end of the paper makes a small step, we can approximate the
area as A ¼ πR2, at least if the paper is thin. This area must be same as the area of
the extremely long and thin rectangle formed by the rolled-out paper viewed
edge-on: A ¼ sh. Equating these two areas gives s ¼ πR2/h ¼ 78,539.816 mm.
It’s such a nice little trick.

h
s
h
R

Fig. 46.1 Left: Toilet roll from the side, area approximately πR2. Right: Unrolled
paper viewed edge-on, with area sh
47
A Delightful Nuclear Missile Disaster

In 2009, the Russians had made 11 test launches of the Bulava in 5 years. The
new generation of intercontinental ballistic missiles was a matter of national
pride, an enormously costly and advanced development program that simply
had to succeed. But things were not going well. Five of the launches had failed
spectacularly. Heads were rolling. There must have been a feeling of desper-
ation in the cold White Sea air, as the 175 m long nuclear submarine Dmitry
Donskoy was preparing for another test launch. It was early morning,
December 9, 2009.
After the ignition of the first stage, the deadly device shot through the
atmosphere with a thundering noise. First stage burnout and separation.
Second stage ignition, burnout and separation. By now, the missile had
escaped from the atmosphere, and there may have been some subdued cheers.
But the third stage was up to no good, destined to give the engineers another
devastating blow. Shortly after ignition of the third stage engine the nozzle
must have been damaged, sending the exhaust to the side. The missile started
to tumble.
In northern Norway, skies were clear and people were getting to work
around 7.50 am local time. What they saw in the eastern sky was nothing
less than apocalyptic. The giant, expanding vortex was all the more scary
because it was totally unexpected and incomprehensible. Yet in the village of
Skjervøy, photographer Jan Petter Jørgensen had the presence of mind to set
up his camera properly and obtain a perfect photograph (Fig. 47.1).
Literally spiraling out of control, the Bulava had turned from weapon of
mass destruction to a piece of art.

© Springer International Publishing AG 2016 203


Ø. Hammer, The Perfect Shape, DOI 10.1007/978-3-319-47373-4_47
204 The Perfect Shape

Fig. 47.1 The Bulava spiral over Skjervøy, Norway, December 2009. Photo Jan
Petter Jørgensen

The eerie images were soon all over the world, in newspapers, on TV, and
not least on the Internet. The spiral shape struck a nerve among mystics and
conspiracy theorists. It was a wormhole to another dimension, aliens, a “Tesla
death ray” or some other top-secret experiment. Slightly cooler-headed com-
mentators suggested a rotating comet or meteor. The correct explanation was
soon given by rocket engineers and confirmed by the Russian military.
As a first approximation, we can think of the Bulava spiral as an Archimedes
spiral. Recall Archimedes’ definition, where we imagine sitting on a straight
line rotating around a center while moving out along the line with constant
speed. This is analogous to the exhaust moving out from the missile at a
constant velocity, in a plane normal to the line of vision, while the missile
rotates at a constant angular rate.
Because the missile itself was moving at great speed, because the expanding
spiral was probably slowed down a little by tenuous air, and other complica-
tions, the resulting spiral was not quite perfect. But it delighted the whole
world, perhaps except that poor crew in the White Sea.
The Bulava spiral has a ghostly parallel in the night sky: the so-called
pre-planetary nebula IRAS 23166 þ 1655 in the constellation of Lyra
(Fig. 47.2). There, a dying carbon star called AFGL 3068 is ejecting dust
and gas in the phase leading up to the birth of a planetary nebula, such as the
47 A Delightful Nuclear Missile Disaster 205

Fig. 47.2 Hubble telescope image of pre-planetary nebula IRAS 23166 þ 1655.
Other stars in the field are not physically associated with the nebula. Credit
ESA/NASA and R. Sahai. Creative Commons Attribution 3 licence

well-known Ring Nebula in the same constellation. AFGL 3068 has a close
stellar companion, and the binary pair is rotating around the common center
of gravity. Complex interactions between the two stars modify the stellar wind,
and shape the nebula into an Archimedes-like spiral when seen normal to the
orbital plane (Mastrodemos and Morris 1999).
Or think of the solar wind in our own solar system. Let us say that the Sun
has radius r0, and rotates with angular velocity ω in radians/day (Fig. 47.3).
We also assume that there is at the present time (t ¼ 0) a coronal mass ejection
(CME) on the Sun, at latitude 0 and longitude 0 in a fixed reference frame in
the solar equatorial plane. Thus, the particles emitted from the surface at
present have polar coordinates (0, r0) in the solar equatorial plane. The solar
wind has a radial velocity v. This means that the particles emitted t days ago,
when the CME was located at longitude ωt, have now reached a position
(ωt, r0 + vt).
206 The Perfect Shape

r0 (0, r0)

(-ωt, r0+vt)

Fig. 47.3 A coronal mass ejection with velocity v on the surface of the rotating Sun
gives rise to a solar wind in the shape of an Archimedes spiral (top view)

To find the shape of the whole trail of particles we make a simple change of
coordinates to get rid of time: φ ¼ ωt. This gives an Archimedes spiral with
polar coordinates
v
r ¼ r0  φ:
ω

This so-called Parker spiral fits quite well with observations. To get an idea
about the dimensions of the spiral, we can use r0 ¼ 7.0  105 km,
ω ¼ 0.257 rad/day for a 24.5 days solar rotation period, and v ¼ 400 km/s
or 3.5  107 km/day for the slow component of the solar wind. The spiral then
has the equation

r ¼ 7:0  105  1:3  108 φ km;

which we plot for negative φ. The distance between successive whorls of this
spiral is about 8  108 km, or roughly the radius of the orbit of Jupiter.
If the source of the CME is not on the equator, the velocity in the radial
direction will be smaller but there will also be a velocity component parallel
with the rotational axis of the Sun. This will result in a helical spiral similar to
the tower crane spiral we saw earlier.
48
Shaligram-Shilas and the Hands of Vishnu

Tulasi means Matchless. The incomparably beautiful princess Tulasi dreamed


of Vishnu himself as her husband. Brahma promised Tulasi the eternal love of
Vishnu, but first, she would have to go through a penance. She would marry
the daemon Sankhachuda, who had himself lived a life of devotion. Brahma
also granted Sankhachuda the gift of invincibility in battle, for as long as Tulasi
remained true to him. At first sight this seemed a grave mistake, for
Sankhachuda went berserk, terrorizing humans and demigods. Nothing, not
even Shiva, could end his havoc for as long as Tulasi was chaste. Vishnu had to
make a hard choice, suspending morality for the common good: He disguised
himself as Sankhachuda and sneaked into Tulasi’s bedchamber. Now
Brahma’s plan played out. Tulasi’s chastity was broken, and Sankhachuda
was immediately killed on the battlefield. “You have deceived me, and killed
my husband!” Tulasi shouted angrily. “Your heart must be made of stone, and
thus I curse you to remain on Earth and become one”! Vishnu gave a beautiful
answer. “Now you receive the fruit of your penance. Leaving your body and
becoming Divine, you will always be with me. Your body will become the river
Gandaki, and your wonderful hair will become millions of small trees that will
be called Tulasi. And I, accepting your curse, will become many stones on the
banks of Gandaki”.
Since that day, the Hindu devotees of Vishnu, the vaishnavas, have collected
sacred stones, shaligram-shilas, on the banks of Gandaki River in Nepal, and
worshipped them as avatars of Vishnu. They are found in shrines all over
India. A shaligram is a peculiar-looking, shiny, jet-black, Jurassic limestone,
making sacred the surrounding area within 24 miles. The shaligrams are

© Springer International Publishing AG 2016 207


Ø. Hammer, The Perfect Shape, DOI 10.1007/978-3-319-47373-4_48
208 The Perfect Shape

named according to several elaborate classification schemes, one of them given


in the Purana (holy book) of Brahma-vaivarta. Here, a Ranarama is “round
and medium-sized, with two circular marks and a quiver with arrows”. These
“circular marks” are stunningly beautiful, spiral shapes with ribs and spines,
found on many or most of the shaligrams. To paleontologists, they are
ammonites. To the vaishnavas, they are incarnations of the Sudarshana Chakra,
the wheel-like magical weapon of Vishnu, held in his upper right hand
(Fig. 48.1). With 108 (or 9  12) teeth, it looks a bit like an ammonite with
a serrated keel.
In his upper left hand, Vishnu carries another sacred object of interest to us:
Shankha. The Shankha is a conch of Turbinella pyrum, a large marine snail
with a shell up to 22 cm long. Traditionally used as a musical instrument in
India, the Shankha is most sacred to the vaishnavas, and is mentioned many
times in the ancient Hindu texts. Blowing it produces Om, the primeval sound
of Creation, the breath of Vishnu. A distinction is made between the common,
dextral form, known as Vamavarta Shanka, associated with Shiva, and the very
rare, sinistral form, or Dakshinavarta Shanka, connected with Vishnu and his
consort Lakshmi. I am not sure how entrenched this distinction is among the
Hindus, for images of Vishnu often show him with a Vamavarta.
The Shankha as a trumpet of war is evoked in the very beginning of the
Bhagavad-Gita (ca. 200 BC), sounding so beautiful in Sanskrit:

Pancayanyam hrsikeso devadattam dhananjayah


paundram dadhmau maha-shankham bhima-kharma vrkodarah.
Lord Krishna blew his conch Pancajanya; Arjuna blew his conch Devadatta;
Bhima, the inspirer of incredible bravery, blew the great conchshell Paundra.

The Indian classical dance, Bharathanatyam, uses more than 50 mudras, or


hand gestures. The Shankha Mudra is a nice one. Grab your left thumb with
your right hand. Then keep the remaining four fingers on your left hand
together and straight, and touch your left index finger with your right thumb
to form a tent-shaped figure. The side facing away from you (Fig. 48.2) now
forms a fairly convincing snail shell with the opening on the right—a dextral
shell, Vamavarta Shankha, the conch of Shiva. Grabbing your right thumb
with your left hand, you make the Dakshinavarta Shanka, the conch of
Vishnu. If you ever go to a conchologist party, you should try that trick. It
cures throat problems too, according to some.
48 Shaligram-Shilas and the Hands of Vishnu 209

Fig. 48.1 Vishnu, with the Chakra in his upper right hand and the Shankha in his
upper left. Illustration Ramanarayanadatta astri, public domain via Wikimedia
Commons
210 The Perfect Shape

Fig. 48.2 The dextral version of the Shankha Mudra


49
The Quest for the Sublime Spiral

One of the strangest spiral stories is that of the Ionic volute. Among the most
elegant creations of the ancient Greeks, this spiral has caused endless confu-
sion, debate and frustration. The commotion started with the Romans, and
exploded in the Renaissance. It has only got worse since then, and now the
issue is as impenetrable as ever. There is no better illustration of our fascination
with spirals (Fig. 49.1).
The Greeks needed to hold their roofs up, so they made columns. And of
course, being Greeks, they had to decorate them, so they put so-called capitals
on top. Now most Greek monumental architecture was built in one of a small
number of styles, or “orders”. The orders were like pre-made packages of
stylistic elements, somewhat like the “themes” of some word processing and
presentation software. What concerns us here is the Ionic order, whose capitals
had fancy whorls or volutes on the sides, hanging under the top plate called the
abacus.
Naturally, the Romans tried to copy the Greek Ionic volutes, and
there are countless examples of them in classical Roman architecture. As
luck would have it, around 25 BC the Roman architect Vitruvius
produced a magnificent treatise on architecture, called De Architectura. In
the Renaissance, when architects were again desperately imitating every
detail of antiquity, the work by Vitruvius was treasured beyond limit. Now
you can buy it in paperback. The third book of De Architectura contains a
recipe for making a Ionic volute with a diameter of 8 units, and it goes as
follows:

© Springer International Publishing AG 2016 211


Ø. Hammer, The Perfect Shape, DOI 10.1007/978-3-319-47373-4_49
212 The Perfect Shape

Fig. 49.1 Ionic capital from the Erechtheion, Acropolis, ca. 415 BC, British Museum
(photo Steven Zucker, CC:BY-NC-SA 2.0)

Tunc ab linea, quae secundum abaci extremam partem demissa erit, in


interiorem partem alia recedat unius et dimidiatae partis latitudine. deinde hae
lineae dividantur ita, ut quattuor partes et dimidia sub abaco relinquantur. tunc
in eo loco, qui locus dividit quattuor et dimidiam et tres et dimidiam partem,
centrum oculi conlocetur signeturque ex eo centro rotunda circinatio tam magna
in diametro, quam una pars ex octo partibus est. ea erit oculi magnitudine, et in
ea catheto respondens diametros agatur. tunc ab summo sub abaco inceptum
schema volutae in singulis tetrantorum actionibus dimidiatum oculi spatium
minuatur, doneque in eundem tetrantem, qui subest, conveniat.

In translation (Morris Hicky Morgan, 1914):

Then let another line be drawn, beginning at a point situated at a distance of one
and a half parts toward the inside from the line previously let fall down along the
edge of the abacus. Next, let these lines be divided in such a way as to leave four
and a half parts under the abacus; then, at the point which forms the division
between the four and a half parts and the remaining three and a half, fix the
center of the eye, and from that center describe a circle with a diameter equal to
one of the eight parts. This will be the size of the eye, and in it draw a diameter on
the line of the “cathetus.” Then, in describing the quadrants, let the size of each
be successively less, by half the diameter of the eye, than that which begins under
the abacus, and proceed from the eye until that same quadrant under the abacus
is reached.
49 The Quest for the Sublime Spiral 213

Abacus

4.5

3.5 4.0

Fig. 49.2 Following Vitruvius’ description to the letter

At first sight, this seems a relatively clear description. You start the spiral
from the top, and the first quarter whorl has a radius of 4.5 units. Then for
each quarter whorl, you reduce the radius by 0.5 units, until you are under the
abacus again. In the middle, there should be an eye of diameter 1 unit. Let us
try (Fig. 49.2):
That does not look very good. The Renaissance architects were
dumbfounded. Could Vitruvius have been wrong? Never! But maybe he was
a bit . . . imprecise? At this point, in the early 1500s, all hell broke loose, and
the commotion continued for centuries. Countless scholars pondered the
problem, and each had his own ingenious solution.
There is, in fact, nothing in Vitruvius’ text that forces us to use circular
segments. His “describing the quadrants” could well be some kind of freehand
drawing where the reduction in size was continuous and somewhat
unspecified. In addition, the text could be interpreted to mean that the
construction stops after one whorl (under the abacus) rather than continuing
for one more whorl to the eye. If so, the design of the inner whorls is left to the
workman. However, early Renaissance interpreters assumed that the segments
were circular and that the rule continued to the eye, not unreasonably, because
this would allow an exact construction using a compass. An easy way out
would simply be to move the centers of the quarter circles around to make the
spiral continuous—after all Vitruvius does not specify the centers of the
segments (Fig. 49.3):
This is an acceptable engineering solution, but the spiral is approximately of
the Archimedes type, and not particularly pretty. To be precise, it is the
involute of a square—remember the puppy tied to its house?
One of the earliest Renaissance construction rules for volutes is that
of Leon Battista Alberti in his Ten Books on Architecture (1452). It is an
214 The Perfect Shape

Abacus

Fig. 49.3 Centers of quarter circles moved to make the segments continuous

Fig. 49.4 Alberti’s volute (1452)—the involute of a straight line segment

involute of a straight line segment, producing a sequence of joined semicircles


with centers alternating between the ends of the straight line. It looks awful
(Fig. 49.4).
In contrast, thorough studies of classical ruins convinced the Renaissance
architects that the Greeks diminished the distance between the whorls as they
approached the eye, much like a snail shell. However, the discovery of the
logarithmic spiral was still in the future. Instead, scholars started inventing
more and more complex algorithms, maintaining the basic proportions of
Vitruvius but diminishing the radii by decreasing amounts and moving the
centers of the circular segments about in complicated ways while approaching
the eye. Some of the methods, such as the ones by Sebastiano Serlio (1537)
and Giuseppe Salviati (1552) were quite successful, producing fairly elegant
snail-like volutes. A clear and stringent description of these early attempts is
49 The Quest for the Sublime Spiral 215

given by Andrey and Galli (2004). New methods continued to be proposed


through the Baroque (Fig. 49.5).
The modern literature about the development of the Ionic volute is
confusing. One “just-so” story, often repeated, is that the true, Greek
volutes were actually logarithmic (equiangular) spirals. One book (Hersey
2000) even claims that the Vitruvian volutes illustrated above are equiangular
spirals. But as far as we know, the Greeks did not know about the
logarithmic spiral or how to construct one. In fact, the Greek volutes seem
to have all kinds of shapes, sometimes even Archimedes spirals. It is true that
most of them look vaguely logarithmic, but this may have been due to
workmanship and a sense of beauty and proportion rather than mathematics.
A related, persistent story is that of the “whelk-on-a-string”. This ingenious
idea was seemingly first conceived by the architect Sir Banister F. Fletcher
sometime around 1900, and is illustrated in his very popular “A history of
architecture”. Wind a piece of string around the shell of a whelk (or other
trochospiral gastropod) from the aperture to the apex; tie the apical end
of the string to the tip of a pencil, and put the shell apex down on a
piece of paper. Holding the shell fixed, unwind the string while holding it
tight. The pencil will now draw a nice volute (in theory that is; you might
find it is not so easy in practice). In fact, we are constructing the involute
of the logarithmic spiral shell, which will itself be a logarithmic spiral, as

Fig. 49.5 Left: Ionic volute, neoclassicist, Østre Fredrikstad Church, Norway, 1874.
Right: Detail from clock, Copenhagen, 1770
216 The Perfect Shape

we saw in an earlier chapter. Because of Fletcher’s influential book, this


powerful meme is quite entrenched, and it is accepted as fact that the
Greeks actually produced their Ionic volutes using such a trick. But alas we
have no evidence for this either.
The horrible truth may be that the Greeks just drew their volutes freehand.
50
A Very Funny Fish

East of the Urals, there is a town called Karpinsk. On the Moon, there is a
crater called Karpinsky. They are both named after the first president of the
Russian Academy of Sciences, Alexander Karpinsky (1847–1936). It was
Karpinsky who first described the bizarre fossil Helicoprion.
As a paleontologist, I am sometimes asked: Why were the dinosaurs so big? I
have two slightly conflicting answers, which rarely satisfy the questioner.
Firstly, there are some very big animals living now as well, the blue whale
for example. The dinosaurs were not really that much bigger. Secondly, if size
simply varies randomly through time, we would expect there to be a time in
the past when animals were bigger than today—there is no particular reason
why living animals should be the biggest of all time. We could ask a similar
question about strangeness: Why were some extinct animals so strange? Well
firstly, there are some very strange living animals as well, humans for example.
Secondly, if strangeness varies randomly through time, we would expect
stranger animals in the past—there is no particular reason why living animals
should be the strangest of all time.
In any case, Helicoprion is high on my list of strange extinct animals.
Usually, the only part that is found of this animal is a logarithmic spiral of
teeth, up to more than 20 cm in diameter. The teeth are definitely from a shark
or related fish, experts say, and it lived in the Permian, some 270 million years
ago. But there the agreement ends. Where in the shark was this extraordinary
structure placed? In the mouth, presumably, but it could have been at the front
of the jaw or in the throat. One might think that the youngest teeth formed in
the center of the spiral and were pushed outwards while growing, before they

© Springer International Publishing AG 2016 217


Ø. Hammer, The Perfect Shape, DOI 10.1007/978-3-319-47373-4_50
218 The Perfect Shape

were used and shed, but experts seem to agree that the process was quite the
opposite: New, large teeth were formed at the outermost end, pushing older,
smaller teeth back and into the spiral (Fig. 50.1).
CT scans of an unusually well preserved specimen with partial jaws indicate
that the spiral filled most of the lower jaw (Tapanila et al. 2013). Moreover,
these scans reveal that Helicoprion was probably not a true shark, but belonged
to the closely related holocephalians, or rat fishes. The rat fishes are still living,
and they look weird, to say the least. I do like the idea of a 7 m long rat fish
with a logarithmic spiral mouth. Now that is about as strange as it gets.

Fig. 50.1 Tooth whorl of the shark Helicoprion


51
Spirals of the Mind

A modern science ethics committee would not have been amused by Heinrich
Klüver, an ex-patriot German psychologist looking exactly like you would
expect from an ex-patriot German psychologist, with Freudian glasses and
overdone eyebrows. In the 1920s he administered mescaline, the hallucino-
genic substance in the peyote cactus, both to himself and to a large number of
volunteers. Peyote was traditionally used by American indigenous people in
shamanistic practice, and was made famous by the book “The Doors of
Perception” by Aldous Huxley.
One of the most interesting effects carefully documented by Klüver was that
a small set of particular geometric patterns was experienced consistently across
experiments and across subjects (Klüver 1966 [1928]). He classified these
patterns into four main types, which he called form constants: Tunnels, cob-
webs, lattices and . . . spirals. Later experiments emphasized that even the lattice
patterns usually had a spiral aspect. Other researchers found that many sources
of altered states of mind could conjure up the same form constants: Other
drugs, exhaustion, rhythmic body movement over a long time, or illness. People
with near-death experiences sometimes reported a “tunnel of light”.
In an influential paper from 1988, Lewis-Williams and Dowson developed
an idea that had already occurred to Klüver. Considering the widespread
shamanistic practices in early cultures, could it be that the trippy form
constants had been important elements of prehistoric religion? Maybe the
networks and spirals so common in Paleolithic, Neolithic and Bronze Age rock
art depicted the mystic visions obtained in trance? It is certainly an appealing
theory.

© Springer International Publishing AG 2016 219


Ø. Hammer, The Perfect Shape, DOI 10.1007/978-3-319-47373-4_51
220 The Perfect Shape

Fig. 51.1 Crude models of two of the patterns that can self-organize in the visual
cortex

There is another fascinating twist to this story. Neurologists studying the


processing of vision in the brain think they have a good idea why the form
constants appear (Bressloff et al. 2001). The theory is very elegant. First of all,
there is an area of the cortex in the back of the brain where neurons are
activated in a pattern that reflects the image received by the retina in the eye.
For simplicity, we can consider this area, called V1, as a two-dimensional
sheet. Researchers have made models of the interconnections between neurons
in V1, and calculated that some patterns can self-organize spontaneously in
response to noise in the system. These textures of stripes and dots (Fig. 51.1)
are well known from other pattern formation systems studied in physics and
biology (Turing patterns), so they are not really surprising.
But these are not exactly the patterns that are seen when taking mescaline or
LSD. The reason is that the image on the retina is not mapped onto V1
directly, but distorted through a transformation that can be modelled with a
simple mathematical equation. A point on the retina given in polar coordinates
(r, φ), is mapped to a point on V1 given in rectangular coordinates as x ¼ ln r,
and y ¼ φ. This is, strangely, also the definition of the complex logarithm—if
we consider the retina and V1 as complex planes, a point z ¼ a + bi on the
retina is simply mapped to a point ln z on V1. Because the logarithmic
function flattens out for large r, the mapping allocates a larger area on V1 to
the central field of vision than to the periphery.
When your brain interprets a certain pattern on V1, it must, in effect, carry
out the inverse of the mapping from retina to V1, so that we experience the
“real world” with its original geometry. This inverse function is from (x, y) in
rectangular coordinates on V1, to r ¼ ex, and φ ¼ y, in “real-world” polar
coordinates (Fig. 51.2).
51 Spirals of the Mind 221

Retina Visual cortex V1


Retinal - V1 mapping

(r, φ)
r
φ

(x, y)

V1 - retinal mapping

Fig. 51.2 Geometrical transformations between retina and visual cortex

Fig. 51.3 Perceived pattern (tunnel) resulting from horizontal lines in the virtual
cortex

Now we are in position for a little fun. Let us say there is a horizontal,
straight line of excited neurons on V1, at a constant vertical position y ¼ k.
What will you see? We use the mapping from V1 to retina, so we get r ¼ ex,
φ ¼ k. In other words, moving along the line in V1 (increasing x), we will
move out along a radial, straight line with angle k in retinal coordinates, at an
exponentially increasing rate. Consider then the self-organized pattern on V1
shown above, with a set of parallel, horizontal lines at positions y ¼ k, y ¼ 2k,
y ¼ 3k, etc. What will you see? If the lines have some thickness, you will see
this (Fig. 51.3):
222 The Perfect Shape

It is the Tunnel of Light, one of the four form constants. Even the details of
this figure fit well with observations from drug experiments.
Vertical lines on V1 give rise to a different experienced image. Now, on V1
we have x ¼ k. In polar retinal coordinates we get r ¼ ek, φ ¼ y. Moving along
the line on V1 (increasing y), we describe a circle in retinal coordinates. A set of
constantly spaced vertical lines on V1 will therefore be seen as a set of
concentric circles with exponentially increasing radii, which will again look
like a tunnel. Another form constant, the cobweb, will result from another self-
organized pattern on V1—the rectangular grid. The visual impression will be a
spider’s web composed by a superposition of radial lines (from the horizontal
lines on V1) and circles (from the vertical lines on V1).
Now consider the second self-organized pattern shown above, the hexago-
nally packed dots. The horizontal rows of dots will be mapped to radial rows of
dots in retinal coordinates, like so (Fig. 51.4):
It is a lattice, the third form constant. It is also a circular tessellation, with its
logarithmic spirals as seen in a previous chapter. If the dot pattern on V1 were
rotated a little, the spirals would become much more apparent.
Finally, consider a diagonal line on V1. In rectangular coordinates, we have
x ¼ ky for some slope constant k deciding the angle of the line. In retinal, polar

Fig. 51.4 Perceived pattern (spiral lattice) resulting from packed dots in the visual
cortex. Program code in Appendix B
51 Spirals of the Mind 223

Fig. 51.5 Left: Diagonal lines in the visual cortex. Right: Mapped to retinal
coordinates

coordinates, this gives r ¼ eky, φ ¼ y, or, in other words, r ¼ ekφ. You should
recognize this equation by now—the logarithmic spiral! Which is the final
form constant (Fig. 51.5).
If all the spirals in this book have made you dizzy, then maybe the retinal-
cortical mapping is to blame. When you see a logarithmic spiral, it is mapped
to a straight line in your cortex. Maybe the simplicity of a straight line gives rise
to a particularly strong neural response, giving you that swirly, swaying, vortex
feeling?
52
The Spider’s Spiral Spin

The orb web of a garden spider is made in a spiral pattern. Many people have
claimed that these spirals are logarithmic. Probably the first, and certainly the
most eloquent, was the French entomologist Jean-Henri Fabre (1823–1915).
In his book “The Life of the Spider”, he writes at length about the alleged
logarithmic spiral shape of the garden spider Epeira (now no longer considered
a valid Latin name) (Fig. 52.1):

The Epeira, therefore, is versed in the geometric secrets of the Ammonite and the
Nautilus pompilus; she uses, in a simpler form, the logarithmic line dear to the
Snail.

Few people have considered it necessary to check this assertion. Unfortu-


nately, it turns out to be totally incorrect. Don’t believe in authority (believe in
me instead).
An orb web contains two rather different spirals. First, an auxiliary or
provisional spiral is weaved outwards, with a rather large spacing. This is
only a scaffold for the true capture web, which has a much smaller spacing,
spun inwards. Vollrath and Mohren (1985) measured orb webs carefully, and
concluded that the capture web is close to an Archimedes spiral (constant
distance between coils), but the auxiliary web shows increasing distance
between coils further from the center. Therefore, they considered the auxiliary
web to be close to logarithmic. In fact, their data rather indicate that the
distance is increasing linearly with angle. This is a property of a spiral where
radius increases as the square of angle:

© Springer International Publishing AG 2016 225


Ø. Hammer, The Perfect Shape, DOI 10.1007/978-3-319-47373-4_52
226 The Perfect Shape

Fig. 52.1 Jean-Henri Fabre, 1823–1915. Wikimedia Commons, public domain

Fig. 52.2 Galilean spiral, r ¼ φ2, approximating to the auxiliary spider web

r ¼ k φ2

This spiral is sometimes called a Galilean spiral (Fig. 52.2).


Fabre had a nice hat, but he got the math all wrong (Fig. 52.3).
52 The Spider’s Spiral Spin 227

Fig. 52.3 Detail of capture web in my garden, approximating to an Archimedes


spiral at least in its more regular parts. Wrought iron pole to the right is a quadruple
helix. Photo Marte H. Jørgensen
53
The Mystery of the Twisted Tree

Why are trees twisted? To a varying degree, the wood fibers in a tree are
arranged in a helical pattern known as spiral grain. Often most clearly visible
when the bark has peeled off, or when the wood is cracked by frost or drought,
such helices can also be visible in the bark itself, especially in older
trees (Fig. 53.1). It is a spectacular phenomenon, and important for the
forestry industry, but why? The literature is vast and contradictory. The oldest
reference in a review paper by Kubler (1991) is from 1854. In his classic book
Curves of Life, Cook (1914) speculates that trees in the northern hemisphere
should show right-handed helices, in contrast with left-handed trees south of
the Equator. Vague theories about the Earth’s rotation and prevailing winds
are offered, only to be seemingly disproved later in the book by the observation
of two chestnut trees from the same locality in England, but with opposite
chirality. Astonishingly, hundreds of scientific papers later, this discussion has
not been put to rest. According to Skatter and Kucera (1998), trees in the
northern hemisphere grow faster on the southern side because of phototro-
pism, making the crown asymmetric. This is supported by statistics, at least at
high northern latitudes (Eklund and Säll 2000). The prevailing westerly winds
at these latitudes would then cause a counterclockwise torsion of the tree as
seen from above, and right-handed spiral grain is supposed to make the tree
more resistant to these forces. Later papers have questioned several aspects of
this idea (e.g., Wing et al. 2014), and data from south of the equator are still
sparse.
Kubler (1991) advocated another old theory: The spiral grain ensures even
distribution of water and nutrients around the stem even if some of the root is

© Springer International Publishing AG 2016 229


Ø. Hammer, The Perfect Shape, DOI 10.1007/978-3-319-47373-4_53
230 The Perfect Shape

Fig. 53.1 Ancient Salix tree with right-handed coiling, Fredrikstad, Norway.
Photo Marte H. Jørgensen

damaged or dry. The reason this would work is that helical conduits from one
position around the root reach different angular positions at the same height,
depending on the radial position of the conduit. Thus, each single source area
in the root serves the entire tree. It’s quite cool, kind of holistic, and supported
by experiments. You can break the stem of a tree almost throughout, only
53 The Mystery of the Twisted Tree 231

leaving a little sliver of wood, and if the water supply is sufficient the whole
crown of the toppled tree can remain green on the ground for years. You may
have seen such heroic cases in the woods.
Other theories abound, including relief of growth stresses, and increase or
reduction in bending stiffness. These ideas are not mutually exclusive, and
perhaps several of them contribute to the selective advantage of spiral grain.
It’s about time humanity resolves this delightful mystery. A good start could
be a “citizen science” project where people around the world note the coiling
direction of their neighborhood trees. I would love to see a map of that.
Afterword

After all these rotations, have we come any closer to the essence of the spiral? At
least I hope to have convinced you of the fundamental role of the Perfect Shape
in nature, art, religion and engineering, and how sad our lives would be
without it. If I have managed to infect you with the spiral madness, to make
you look for spirals wherever you turn, then I have succeeded in my evil
scheme. And like the spiral itself, the spiral stories never end. I have not written
about the mainspring and balance spring in a watch, or why hair curls, or LP
and CD records, or finger prints, or the Lorenz attractor, or the Solomon
R. Guggenheim Museum, or constricting snakes, or the Torah, or the
Argonauta, or the narwhal, or the magnificent feather duster and Christmas
tree worms, or a thousand other things, and I have barely mentioned the
double helix of DNA. But it had to end somewhere, and so it ends here.
I have been wondering what to write about next. Squares? Circles? Conic
sections? It would not be the same. There is no figure like the spiral.

© Springer International Publishing AG 2016 233


Ø. Hammer, The Perfect Shape, DOI 10.1007/978-3-319-47373-4
Appendix A: Mathematical Derivations

A.1: Coiled-Up Cone


We want to show that diameter D is proportional to path length s in a coiled-
up, expanding tube describing a logarithmic spiral. This means that the tube
can be described as a coiled-up cone.
Consider first the equation for the total path length of the logarithmic spiral
out to a particular radius r, as derived in Appendix A.2:
pffiffiffiffiffiffiffiffiffiffiffiffiffi
1 þ k2
r
s¼ :
k

This equation describes a property of logarithmic spirals due to Descartes—


that the ratio of arc length to radius is constant—and moreover it specifies this
constant in terms of k. Solving for r gives
sk
r ¼ pffiffiffiffiffiffiffiffiffiffiffiffiffi :
1 þ k2

The diameter D of the coiled-up expanding tube when its inner edge is at
radius r from the center of the spiral is

© Springer International Publishing AG 2016 235


Ø. Hammer, The Perfect Shape, DOI 10.1007/978-3-319-47373-4
236 Appendix A: Mathematical Derivations

D ¼ rðφ þ 2πÞ  rðφÞ


ðφþ2πÞ
¼ aek  aekφ
kφ 2πk
¼ ae ðe  1Þ
¼ rðe2πk  1Þ:

r(φ+2π)

D
r(φ)

Finally, inserting r from above, we find that diameter D is proportional to


path length s, as required, with a coefficient depending only on k:
 
k e2πk  1
D ¼ pffiffiffiffiffiffiffiffiffiffiffiffiffi s:
1 þ k2

A.2: Path Length of the Logarithmic Spiral


We want to find the path length s of a logarithmic spiral as a function of angle
ϕ.
Considering first a general curve r(φ) in polar coordinates, we look at the
new radius r0 after a certain small angular increment, dφ. Now look at the
small right-angled triangle with sides b, c and d in the figure below.
Appendix A: Mathematical Derivations 237

c
d ds
b

r’

dφ r

Geometry of a segment of a logarithmic spiral (red)

Because dφ is small, we can identify b with a circular segment of radius r,


and d with a segment of the curve, with length ds. From the definition of
angles measured in radians, we have b  rdφ. We also have c ¼ r0  r ¼ dr.
Now we can give the length of ds, using the Pythagorean theorem:
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
 2ffi
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi dr
ds  d ¼ b2 þ c2 ¼ ðrdφÞ2 þ ðdrÞ2 ¼ r2 þ dφ:

Plugging in the equation for the logarithmic spiral, we get


qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi pffiffiffiffiffiffiffiffiffiffiffiffiffi
ds ¼ ðaekφ Þ2 þ ðakekφ Þ2 dφ ¼ aekφ 1 þ k2 dφ:

By adding (integrating) all the little ds along the curve, we can then calculate
the total distance s from the pole of the spiral (φ ¼ 1) to any point, e.g.,
φ ¼ ϕ:
Z ϕ Z ϕ pffiffiffiffiffiffiffiffiffiffiffiffiffi
sðϕÞ ¼ ds ¼ aekφ 1 þ k2 dφ
1 1
pffiffiffiffiffiffiffiffiffiffiffiffiffi
pffiffiffiffiffiffiffiffiffiffiffiffiffiZ ϕ
a 1 þ k2 kϕ
¼ a 1 þ k2 e dφ ¼

e :
1 k
238 Appendix A: Mathematical Derivations

A.3: The Roulette of the Pole of a Logarithmic


Spiral
When rolling a logarithmic spiral over a table, its pole will describe a straight
line with slope equal to the expansion coefficient k.

c
p
r

x s
To show this, we consider that the table is the tangent to the spiral for some
spiral angle φ. The radius is then r(φ) ¼ ekφ. We also need the distance p from
the pole to the table, i.e., the length of the perpendicular from the pole to the
tangent of the curve. For any curve in polar coordinates, this distance is related
to the curve r(φ) by the following relation (e.g., Gow 1960, eq. 21.10):
 
1 1 1 dr 2
¼ þ
p2 r 2 r 4 dφ

Plugging in the equation for the logarithmic spiral, r ¼ ekφ, we obtain

1 1 1 kφ 2 1 1 2 2 1 þ k2
¼ þ ðke Þ ¼ þ k r ¼ :
p2 r 2 r 4 r2 r4 r2

Rearranging, we then get the so-called pedal equation, giving p as a function of


r:
r
p ¼ pffiffiffiffiffiffiffiffiffiffiffiffiffi :
1 þ k2

We also know from earlier that the total path length of the spiral is
pffiffiffiffiffiffiffiffiffiffiffiffiffi
1 þ k2
s¼ r:
k
Appendix A: Mathematical Derivations 239

The horizontal position of the tangent point is just s, i.e., the total distance
traveled by the contact with the table. The horizontal position x of the pole is
then
pffiffiffiffiffiffiffiffiffiffiffiffiffi sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi 1 þ k2 r2 r
x ¼ s  c ¼ s  r 2  p2 ¼ r  r2  2
¼ pffiffiffiffiffiffiffiffiffiffiffiffiffi :
k 1þk k 1 þ k2

Now, imagine that we roll the spiral with a small angle dφ. The slope of the
trajectory of the pole is then the ratio of the vertical distance moved to the
horizontal distance moved:
pffiffiffiffiffiffiffi
1 ffi
dp d p=dr 1þk2
¼ ¼ 1 ¼ k:
dx d x=dr pffiffiffiffiffiffiffiffi
2
k 1þk

The simplicity of this result is astounding, and I suspect that it can be found by
a more straightforward method! Since the slope is constant, the trajectory of
the pole is a straight line. For k ¼ 0 (circle), the slope is zero, as the center of a
wheel moves along a horizontal line.

A.4: The Path of a Pendulum with Friction


A mass m hangs at the end of a pendulum of length r, and swings in the
x direction (Fig. A.1). The angular displacement is θx(t). We assume that the
friction force is proportional to tangential velocity, with a coefficient c. With
some simplifying assumptions, such as the amplitude being small, and sin θ  θ
for small θ, we can use Newton’s second law to set up a second-order
differential equation for θ:

d2 θx dθx mg
m 2
þc  θx ¼ 0:
dt dt r

The general solution to this equation is


240 Appendix A: Mathematical Derivations

θx
r

cv
m

mg

Fig. A.1 Pendulum with mass m and length r, dampened by a friction force
proportional to velocity (F ¼ cv)

 
c 2mvx þ Ax c
θx ðtÞ ¼ e Ax cos ðhtÞ þ
2mt sin ðhtÞ
2mh

where Ax is the initial angular displacement and vx the initial angular velocity.
The frequency h, in swings per second, is
rffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
g c2
h¼  2:
r 4m

This shows that the pendulum swings with frequency h, and with exponen-
tially decaying amplitude. Now let us say that for the movement in the
x direction, we choose vx very carefully so as to cancel the last term in the
general solution, i.e., 2mvx + Axc ¼ 0, or vx ¼ Axc/2m. This gives
c
θx ðtÞ ¼ Ax e2mt cos ðhtÞ:

If the pendulum also swings in the y direction, we have


 
c 2mvy þ Ay c
θy ðtÞ ¼ e2mt Ay cos ðhtÞ þ sin ðhtÞ :
2mh

In this direction, we let the initial displacement be zero, i.e., Ay ¼ 0. Moreover,


we choose vy ¼ Axh. This gives
c
θy ðtÞ ¼ Ax e2mt sin ðhtÞ:
Appendix A: Mathematical Derivations 241

Now dropping the index on A, and changing variables so that φ ¼ ht and


k ¼ c/2hm we get

θx ðtÞ ¼ Aekφ cos ðφÞ


θy ðtÞ ¼ Aekφ sin ðφÞ

Finally, we use sin θ  θ for small angles (we already made this approximation
when setting up the differential equation), and assume r ¼ 1. This allows a
change from angular to linear displacements:

x ¼ Aekφ cos ðφÞ


y ¼ Aekφ sin ðφÞ

So if we put a pen on the pendulum, it will draw a logarithmic spiral! With the
approximations we have made, each revolution of the pendulum will take a
constant amount of time, even as the amplitude is decreasing (this is why a
pendulum clock keeps time almost independent from the amplitude). There-
fore, it will take an infinite amount of time to draw the spiral all the way into
the pole, although the path length is finite.
In the derivation above, we were very particular about the selection of the
initial velocities and displacements in the x and y directions. This was only to
arrive at a particularly clean solution, which we would immediately recognize
as a logarithmic spiral. Other initial conditions give a “squashed” logarithmic
spiral, with each whorl looking more like an ellipse than a circle:
242 Appendix A: Mathematical Derivations

A.5: The Path of an Expanding Circular Segment


We want to find the path of the tip of a finger when opening the hand, given
that the hand should always form a circular segment of constant length L and
spanning an angle θ. Without loss of generality, let L ¼ π, and the base of the
hand is fixed at (0, 0). The radius of the circular segment is r.

(x, y)

L=π
r

θ
(-r, 0) (0, 0)
r
The center of the circular segment will be at (r, 0), and we have
θr ¼ L ¼ π, or θ ¼ π/r. The position (x, y) of the tip of the circular segment
is then
π
x ¼ r þ r cos θ ¼ r cos  r
r
π
y ¼ r sin θ ¼ r sin :
r

We can reparameterize this curve to make it look more like “standard” spiral
equations in polar coordinates. By a change of variable we set r ¼ π/φ and
obtain
π 1
x ¼ cos φ 
φ φ
π
y ¼ sin φ
φ
Appendix B: Program Code

Program code for reproducing some of the plots in this book is given below.
The scripts can be executed in the free statistical software Past, written by the
author. You can download Past for Windows and Mac at http://folk.uio.no/
ohammer/past.

B.1: Vogel’s Sunflower Equation


The program produces the first 250 florets of the sunflower, according to
Vogel’s (1979) equation.

cleargraphic;
for n:¼1 to 250 do begin
phi:¼ 2*pi*n*(1-(sqrt(5)-1)/2);
r:¼sqrt(n);
drawpoints(r*cos(phi), r*sin(phi), black);
end;

B.2: Hair Whorl


Produces a grid of short line segments, all oriented with a constant angle to the
radius vector.

© Springer International Publishing AG 2016 243


Ø. Hammer, The Perfect Shape, DOI 10.1007/978-3-319-47373-4
244 Appendix B: Program Code

cleargraphic;
for i:¼1 to 21 do for j:¼1 to 21 do begin
x:¼i-11; y:¼j-11; phi:¼arctan2(y, x)+pi/3;
if sqrt(x*x+y*y)<10 then
drawline(x, y, x+1.5*cos(phi), y+1.5*sin(phi), black);
end;
drawpoints(0, 0, black);

B.3: The Vertigo Harmonograph


Carefully selected parameters of a rotating, stretched logarithmic spiral
reproducing the Vertigo pattern.

cleargraphic;
vx:¼vector(5000); vy:¼vector(5000);
for i:¼1 to 5000 do begin
// First compute the spiral with exponential decay
x1:¼cos(i/10)*exp(-i/2500); y1:¼1.4*sin(i/10)*exp(-i/2500);

// Then rotate the coordinates with a rotation matrix


d:¼450; // Rotation period
vx[i]:¼cos(i/d)*x1-sin(i/d)*y1;
vy[i]:¼sin(i/d)*x1+cos(i/d)*y1;
end;
drawpolyline(vx, vy, black);

B.4: The Bug Problem


Simulates four bugs moving towards their neighbours.

cleargraphic;
N:¼4; // Number of bugs
x:¼vector(N); y:¼vector(N);
newx:¼vector(N); newy:¼vector(N);

// Starting positions
x[1]:¼0; y[1]:¼0; x[2]:¼1; y[2]:¼0;
x[3]:¼1; y[3]:¼1; x[4]:¼0; y[4]:¼1;

for i:¼1 to 500 do begin // Time steps


for j:¼1 to N do begin
Appendix B: Program Code 245

next:¼j+1; if next¼N+1 then next:¼1;


d:¼sqrt(sqr(x[next]-x[j])+sqr(y[next]-y[j])); // For normalizing
newx[j]:¼x[j]+0.002*(x[next]-x[j])/d;
newy[j]:¼y[j]+0.002*(y[next]-y[j])/d;
drawline(x[j], y[j], newx[j], newy[j], black);

// Draw red square every 100 steps


if (i-1) mod 100 ¼ 0 then begin
drawline(x[j], y[j], x[next], y[next], red);
end;
end;

// Copy new positions to old


for j:¼1 to N do begin
x[j]:¼newx[j]; y[j]:¼newy[j];
end;
end;

B.5: Julia Set


This script takes a long time to run in Past, up to 10 min. There are more
efficient algorithms for computing Julia sets, but the direct method is the
easiest to understand.

cleargraphic;
cr:¼0.285; ci:¼0.01;
ju:¼array(400,400);
for i:¼1 to 400 do begin
for j:¼1 to 400 do begin
// Complex start value for z. Adjust to zoom.
zi:¼-(i-200)/150; zr:¼(j-200)/150;
k:¼0;
repeat
// Complex multiplication and addition
zrsave:¼zr;
zr:¼zr*zr-zi*zi+cr;
zi:¼2*zrsave*zi+ci;
k:¼k+1;
until (k¼200) or (zi*zi+zr*zr>3); // The value 200 must be
// increased when zooming
if k¼200 then ju[i,j]:¼6
else ju[i,j]:¼ln(k);
246 Appendix B: Program Code

end;
end;
drawmatrix(ju, true);

B.6: Hyperbolic Cosecant Spiral


The hyperbolic cosecant spiral is one of Cote’s spirals.

cleargraphic;
vx:¼vector(1000); vy:¼vector(1000);
for i:¼1 to 1000 do begin
phi:¼(i+40)/30;
r:¼1/((exp(0.1*phi)-exp(-0.1*phi))/2);
vx[i]:¼r*cos(phi);
vy[i]:¼r*sin(phi);
end;
drawpolyline(vx, vy, black);

B.7: Archimedes Spiral Versus Evolute of circle


Plots an Archimedes spiral in black, and the evolute of the unit circle in red.

cleargraphic;
vx:¼vector(600); vy:¼vector(600);

// Archimedes spiral with a¼1


for i:¼1 to 600 do begin
phi:¼i/45;
vx[i]:¼phi*sin(phi);
vy[i]:¼-phi*cos(phi); // Flip up-down to coincide with evolute
end;
drawpolyline(vx, vy, black);

// Evolute of circle with radius 1


for i:¼1 to 600 do begin
phi:¼i/45;
vx[i]:¼cos(phi)+phi*sin(phi);
vy[i]:¼sin(phi)-phi*cos(phi);
end;
drawpolyline(vx, vy, red);
Appendix B: Program Code 247

B.8: Euler’s Spiral


Plots both branches of Euler’s spiral. Past has a built-in function fresnel
for approximating the two Fresnel integrals, returned as a two-vector.

cleargraphic;
vx:¼vector(1200); vy:¼vector(1200); v:¼vector(2);
for i:¼1 to 1200 do begin
t:¼(i-600)/100; // t goes from -6 to +6
v:¼fresnel(phi);
vx[i]:¼v[2]; vy[i]:¼v[1];
end;
drawpolyline(vx, vy, black);

B.9: Entoptic Pattern


The plan here is to plot the pattern experienced when a self-organized
hexagonal closest packing of dots is produced in the visual cortex (V1). It
might then seem natural to traverse the pixels of the dot pattern, mapping
them with the V1-retina transfer function onto the retinal pattern. However,
this would not populate every pixel of the retinal image. Instead, we traverse all
the pixels of the retinal image, and use the retina-V1 mapping to pick pixels
from V1.

cleargraphic;
// Set up an array x with a staggered dot pattern
x:¼array(200,200);
for i:¼1 to 200 do for j:¼1 to 200 do x[i,j]:¼1;
for i:¼1 to 12 do begin
for j:¼1 to 24 do begin
x[i*16, j*8]:¼0;
x[i*16-8, j*8+4]:¼0;
end;
end;

// Blur it four times to make it smooth


xb:¼array(200,200);
for i:¼1 to 200 do for j:¼1 to 200 do xb[i,j]:¼1;
for k:¼1 to 4 do begin
for i:¼2 to 199 do for j:¼2 to 199 do begin
xb[i,j]:¼(1.2*x[i,j]+x[i-1,j]+x[i+1,j]+x[i,j1]+x[i,j+1])/5.2;
248 Appendix B: Program Code

end;
x:¼xb;
end;

// drawmatrix(xb, true); // Plot the input matrix

// For every cell in the output array y, pick the


// corresponding cell in the input array xb
y:¼array(200,200);
for i:¼1 to 200 do begin
y1:¼10*2*(i-100.5)/101;
for j:¼1 to 200 do begin
x1:¼10*2*(j-100.5)/101;
x2:¼ln(sqrt(x1*x1+y1*y1)); // -5 to ln sqrt(200) ¼ 2.65
if x2>-1 then begin
y2:¼arctan2(y1,x1);
y[i,j]:¼xb[y2*20+100,x2*20+40];
end;
end;
end;

drawmatrix(y, true);
Literature

Recommended Books

Bertin, G., & Lin, C.-C. (1996). Spiral structure in galaxies: A density wave theory.
Cambridge, MA: MIT Press.
Cook, T. A. (1914). The curves of life (1979 ed.). USA: Dover Publications.
Davis, P. J. (1993). Spirals, from Theodorus to Chaos. Wellesley, MA: A K Peters.
Hargittai, I., & Pickover, C. (Eds.). (1992). Spiral symmetry. Singapore: World
Scientific (A nice collection of authoritative essays on spirals).
Israel, N. (2015). Spirals—The whorled image in twentieth-century literature and art.
New York, NY: Columbia University Press.
Klüver, H. (1966 [1928]). Mescal, and mechanisms of hallucination. Chicago: Univer-
sity of Chicago Press.
McManus, C. (2003). Right hand, left hand. London: Phoenix (Thoroughly
researched, critical and readable, on handedness).
Purce, J. (1974). The mystic spiral: Journey of the soul. London: Thames & Hudson
(As the title indicates, this book is hardly outstanding in scientific rigor, but the
large number of beautiful illustrations of spirals in ancient art makes it well worth
the money).
Scales, H. (2015). Spirals in time: The secret life and curious afterlife of seashells.
London: Bloomsbury Sigma (Scientifically rigorous and beautifully written).
Thomas, J. (1999). Understanding the Neolithic. London: Routledge.
Ward, G. (2006). Spirals: The pattern of existence. England: Green Magic (This little
book on spirals in culture and nature contains interesting information, but tends to
diverge into mysticism).

© Springer International Publishing AG 2016 249


Ø. Hammer, The Perfect Shape, DOI 10.1007/978-3-319-47373-4
250 Literature

Other References

Anderhub, J. H. (1941). Joco-Seria—Aus den Papieren eines reisendes Kaufmanns.


Frankfurt am Main: Hauserpresse Hans Schaefer.
Andrey, D., & Galli, M. (2004). Geometric methods of the 1500s for laying out the
Ionic volute. Nexus Network Journal, 6, 31–48.
Bernoulli, J. (1691). Specimen alterum calculi differentialis. Acta Eruditorum, 2,
282–289.
Bourne, D. W., & Heezen, B. C. (1965). A wandering Enteropneusta from the
Abyssal Pacific, and the distribution of “spiral” tracks on the sea floor. Science,
New series, 150(3692), 60–63.
Bressloff, P. C., Cowan, J. D., Golubitsky, M., Thomas, P. J., & Wiener, M. C.
(2001). Geometrical visual hallucinations, Euclidean symmetry, and the functional
architecture of the striate cortex. Philosophical Transactions of the Royal Society of
London B, 356, 299–330.
Buryanov, A., & Kotiuk, V. (2010). Proportions of hand segments. International
Journal of Morphology, 28, 755–758.
Calter, P. (2000). How to construct a logarithmic rosette (without even knowing it).
Nexus Network Journal, 2, 25–31.
Cobley, M. J., Rayfield, E. J., & Barrett, P. M. (2013). Inter-vertebral flexibility of the
ostrich neck: Implications for estimating sauropod neck flexibility. PLoS One, 8(8),
e72187.
Dalley, S., & Oleson, J. P. (2003). Sennacherib, Archimedes and the water screw:
The context of invention in the ancient world. Technology and Culture, 44, 1–26.
Darwin, C. (1875). On the movements and habits of climbing plants. London:
John Murray.
Davis Philip, A. G. (1992). The evolution of a three-armed spiral in the Julia set, and
higher order spirals. In I. Hargittai & C. Pickover (Eds.), Spiral symmetry
(pp. 135–163). Singapore: World Scientific.
de Borhegyi, S. F. (1966). The wind god’s breastplate. Expedition Magazine, 8(4),
13–15.
Douady, S., & Couder, Y. (1996). Phyllotaxis as a dynamical self organizing process.
Parts 1–3. Journal of Theoretical Biology, 178, 255–312.
Earl of Rosse. (1850). Observations on the nebulae. Philosophical Transactions of the
Philosophical Society of London, 140, 499–514.
Eklund, L., & Säll, H. (2000). The influence of wind on spiral grain formation in
conifer trees. Trees, 14, 324–328.
Euler, L. (1743). De summis serierum reciprocarum ex potestatibus numerorum
naturalium ortarum dissertatio altera. Miscellanea Berolinensia, 7, 172–192.
Francis, C., & Anderson, E. (2009). Galactic spiral structure. Proceedings of the
Royal Society of London A, 465, 3425–3423.
Fuchs, B., Möllenhoff, C., & Heidt, J. (1998). Decomposition of the rotation curves
of distant field galaxies. Astronomy and Astrophysics, 336, 878–882.
Literature 251

Gardner, M. (1965). Mathematical games. Scientific American, 213(1), 100–104.


Gerbode, S. J., Puzey, J. R., McCormick, A. G., & Mahadevan, L. (2012). How the
cucumber tendril grows and overwinds. Science, 337, 1087–1091.
Ghyka, M. (1946). The geometry of art and life. New York, NY: Sheed and Ward.
Gow, M. M. (1960). A course in pure mathematics. Burlington, MA: Elsevier.
Guy, N. (2011). The rise of the anticlockwise newel stair. The Castle Studies Group
Journal, 25, 113–174.
Guyon, E., Hulin, J.-P., Petit, L., & Mitescu, C. D. (2001). Physical hydrodynamics.
Oxford: Oxford University Press.
Hammer, Ø., & Bucher, H. (2005). Models for the morphogenesis of the molluscan
shell. Lethaia, 38, 1–12.
Hammer, Ø., Hryniewicz, K., Hurum, J. H., Høyberget, M., Knutsen, E. M., &
Nakrem, H. A. (2013). Large onychites (cephalopod hooks) from the Upper Jurassic
of the Boreal Realm. Acta Palaeontologica Polonica, 58, 827–835.
Heezen, B. C., & Hollister, C. D. (1971). The face of the deep. Oxford: Oxford University
Press.
Hersey, G. L. (2000). Architecture and geometry in the age of the Baroque. Chicago:
University of Chicago Press.
Holland, N. D., Clague, D. A., Gordon, D. P., Gebruk, A., Pawson, D. L., &
Vecchione, M. (2005). ‘Lophenteropneust’ hypothesis refuted by collection and
photos of new deep-sea hemichordates. Nature, 434, 374–376.
Hoso, M., Kameda, Y., Wu, S.-P., Asami, T., Kato, M., & Hori, M. (2010). A
speciation gene for left-right reversal in snails results in anti-predator adaptation.
Nature Communications, 1, 133.
Imafuku, M. (1994). Response of hermit crabs to sinistral shells. Journal of Ethology,
12, 107–114.
Kamper, D. G., Cruz, E. G., & Siegel, M. P. (2003). Stereotypical fingertip trajec-
tories during grasp. Journal of Neurophysiology, 90, 3702–3710.
Kauffman, E. G., Harries, P. J., Meyer, C., Villamil, T., Arango, C., & Jaecks,
G. (2007). Paleoecology of giant Inoceramidae (Platyceramus) on a Santonian
(Cretaceous) seafloor in Colorado. Journal of Paleontology, 81, 64–81.
Kenyon, F. C. (1895). In the region of the new fossil, Dæmonelix. American Naturalist,
29, 213–227.
Klar, A. J. S. (2003). Human handedness and scalp hair-whorl direction develop from
a common genetic mechanism. Genetics, 165, 269–276.
Klar, A. J. S. (2004). Excess of counterclockwise scalp hair-whorl rotation in homo-
sexual men. Journal of Genetics, 170, 2027–2030.
Klar, A. J. S. (2009). Scalp hair-whorl orientation of Japanese individuals is random;
Hence, the trait’s distribution is not genetically determined. Seminars in Cell and
Developmental Biology, 20, 510–513.
Klüver, H. (1966 [1928]). Mescal, and mechanisms of hallucination. Chicago: Univer-
sity of Chicago Press.
252 Literature

Kosuge, T., & Imafuku, M. (1997). Records of hermit crabs that live in sinistral shells.
Crustaceana, 70, 380–384.
Krumbein, W. C. (1944). Shore processes and beach characteristics. US Corps of
Engineers, Beach Erosion Board, Technical Memoir no. 3.
Kubler, H. (1991). Function of spiral grain in trees. Trees, 5, 125–135.
LeBlond, P. H. (1979). An explanation of the logarithmic spiral plan shape of
headland bay beaches. Journal of Sediment Petrology, 49, 1093–1100.
Leung, K.-T., Józsa, L., Ravasz, M., & Néda, Z. (2001). Pattern formation:
Spiral cracks without twisting. Nature, 410, 166.
Lewis-Williams, J. D., & Dowson, T. A. (1988). The signs of all times—Entoptic
phenomena in Upper Palaeolithic art. Current Anthropology, 29, 201–245.
Littler, J. W. (1973). On the adaptability of man’s hand (with reference to the
equiangular curve). The Hand, 5, 187–191.
Longo, M. J. (2011). Detection of a dipole in the handedness of spiral galaxies with
redshifts z  0.04. Physical Letters B, 699, 224–229.
Lucas, E. (1877). Problème des trois chiens. Nouvelle Correspondance Mathématique,
3, 175–176.
Martin, L. D., & Bennett, D. K. (1977). The burrows of the Miocene beaver
Palaeocastor, Western Nebraska, U.S.A. Palaeogeography, Palaeoclimatology, Palaeo-
ecology, 22, 173–193.
Mastrodemos, N., & Morris, M. (1999). Bipolar pre-planetary nebulae: Hydrodynamics
of dusty winds in binary systems. II. Morphology of the circumstellar envelopes.
The Astrophysical Journal, 523, 357–380.
Mattheck, C., & Reuss, S. (1991). The claw of the tiger: An assessment of its mecha-
nical shape optimization. Journal of Theoretical Biology, 150, 323–328.
McDonald, J. H. (2011). Myths of human genetics (pp. 40–45). Baltimore, MD:
Sparky House Publishing.
McDowell, J. (1973). Plato; Theaetetus. Oxford: Clarendon Press.
McKinney, F. K., Listokin, M. R. A., & Phifer, C. D. (1986). Flow and polypide
distribution in the cheilostome bryozoan Bugula and their inference in Archimedes.
Lethaia, 19, 81–93.
Mercer, H. C. (1929). Ancient carpenters’ tools (Reprinted 5th ed.). Dover Publi-
cations, 2000.
Nahin, P. J. (2012). Chases and escapes: The mathematics of pursuit and evation.
Princeton, NJ: Princeton University Press.
Noone, C. J., Torrilhon, J., & Mitsos, A. (2012). Heliostat field optimization:
A new computationally efficient model and biomimetic layout. Solar Energy, 86,
792–803.
Raup, D. M. (1962). Computer as aid in describing form in gastropod shells.
Science, 138(3537), 150–152.
Raup, D. M., & Michelson, A. (1965). Theoretical morphology of the coiled shell.
Science, 147(3663), 1294–1295.
Literature 253

Reyssat, E., & Mahadevan, L. (2009). Hygromorphs: From pine cones to biomimetic
layers. Journal of the Royal Society, Interface. doi:10.1098/rsif.2009.0184.
Ryan, A. J., & Christensen, P. R. (2012). Coils and polygonal crust in the Athabasca
Valles Region, Mars, as evidence for a volcanic history. Science, 336, 449–452.
Sassi, M., & Vernoux, T. (2013). Auxin and self-organization at the shoot apical meri-
stem. Journal of Experimental Botany, 64, 2579–2592.
Schilthuizen, M., Craze, P. G., Cabanban, A. S., Davison, A., Stone, J., Gittenberger,
E., & Scott, B. J. (2007). Sexual selection maintains whole-body chiral dimor-
phism in snails. Journal of Evolutionary Biology, 20, 1941–1949.
Shamir, L. (2012). Handedness asymmetry of spiral galaxies with z < 0.03 shows
cosmic parity violation and a dipole axis. Physics Letters B, 715, 25–29.
Skatter, S., & Kucera, B. (1998). The cause of the prevalent directions of the
spiral grain patterns in conifers. Trees, 12, 265–273.
Smith, R. M. H. (1987). Helical burrow casts of therapsid origin from the Beaufort
Group (Permian) of South Africa. Palaeogeography, Palaeoclimatology, Palaeoecology,
60, 155–170.
Stevens, P. (1974). Patterns in nature. Boston, Toronto: Little, Brown and Company.
Tapanila, L., Pruitt, J., Pradel, A., Wilga, C. D., Ramsay, J. B., Schlader, R., &
Didier, D. A. (2013). Jaws for a spiral-tooth whorl: CT images reveal novel
adaptation and phylogeny in fossil Helicoprion. Biology Letters, 9, 20130057.
Teichert, C., & Kummel, B. (1960). Size of endoceroid cephalopods. Breviora Museum
of Comparative Zoology, 128, 1–7.
Torgersen, J. (1950). Situs inversus, asymmetry and twinning. American Journal of
Human Genetics, 2, 361–370.
Torricelli, E. (1644). De infinitis spirabilus. (Republished by E. Carruccio, Domus
Galilaena, Pisa, 1955).
Tozzer, A. M., & Allen, G. M. (1910). Animal figures in the Maya codices. Papers of
the Peabody Museum of Archaeology and Ethnography, Harvard University, 5,
283–372.
Tucker, V. A. (2000). The deep fovea, sideways vision and spiral flight paths in
raptors. Journal of Experimental Biology, 203, 3745–3754.
Tucker, V. A., Tucker, A. E., Akers, K., & Enderson, J. H. (2000). Curved flight
paths and sideways vision in peregrine falcons (Falco peregrinus). Journal of Experi-
mental Biology, 203, 3755–3763.
Vogel, H. (1979). A better way to construct the sunflower head. Mathematical Bio-
sciences, 44, 179–189.
Vollrath, F., & Mohren, W. (1985). Spiral geometry in the garden spider’s orb web.
Naturwissenschaften, 72, 666–667.
Wentworth Thompson, D. (1917). On growth and form. Cambridge: Cambridge Uni-
versity Press.
Wing, M. R., Knowles, A. J., Melbostad, S. R., & Jones, A. K. (2014). Spiral grain in
bristlecone pines (Pinus longaeva) exhibits no correlation with environmental factors.
Trees, 28, 487–491.
Index

A Baird, John Logie, 95


Acanthus scroll, 80, 81 Beaches, 57, 93, 94
Acorn worm, 1, 2 Begby ship, 200
Airships, 46, 47 Begonia, 81, 82
Alberti’s volute, 214 Belemnoids, 166
Allometry, 27 Bernoulli, Jakob, 35, 36, 119, 185
Ammonites, 12, 24, 57, 58, 123, 148, 208 Bernoulli, Johann, 119
Amphidromus, 31, 32 Bhagavad-Gita, 208
Analog computer, 128, 129 Bilayer, 109, 110
Anchors, 166 Bivalves, 130, 147, 148
Antikythera mechanism, 184 Brachiopods, 123, 147, 148, 177
Arabesque, 81 Bronze age, v, 55, 102, 115, 189, 192,
Archimedes 194, 219
screw, 179, 180, 182 Bryozoan, 54
spiral, 5, 8, 9, 13, 20, 96, 125, 126, Bubble chamber, 161–163
145, 170, 173
Aristotle, 7, 73
Asbestos, 111 C
Auger, 180, 181 Cam, 9
Cameroceras, 24
Cave of Swallows, 65
B Cephalopods, 57, 166
Babbage, Charles, 73–78 Chapman, Fredrik Henrik af, 166, 167
Bacteria, 182 Chrysotile, 110, 111
Baculitids, 24 Circular tessellation, 41–52, 99, 222

© Springer International Publishing AG 2016 255


Ø. Hammer, The Perfect Shape, DOI 10.1007/978-3-319-47373-4
256 Index

Circumnutation, 80 F
Cloud chamber, 161, 162 Fabre, Jean-Henri, 225, 226
Cochlea, v, 26 Falcon, 39–40
Coffee, 61, 169–172, 199 Fermat spiral, 5
Coltsfoot, 81, 82 Ferns, 79, 82
Columella, 130, 131 Feynman, Richard, 30
Complex numbers, 106, 107 Fibonacci sequence, 15
Corbusier, 17 Fibonacci spiral, 17, 142
Coriolis Effect, 152 Fiddlehead, 82
Coronal mass ejection, 205, 206 Finger path, 90
Cosmological Principle, 61, 62 Florets, 48–52
Cotes, Roger, 121, 133, 134 Foraminifera, 149, 159, 160
Crosier, 196–200 Form constants, 219, 220, 222, 223
Cucumber, 79, 82, 83 Foucault, Léon, 86
Curvature, 109, 110, 166, 197, Fractal, 58, 82, 107, 108
198, 200 Fresnel integrals, 198, 199

D G
Daemonelix, 139, 140 Galaxy, 60–62, 85–88, 170, 171
Dallas Airport, 142, 143 Galilean spiral, 226
Darwin, Charles, 79 Gears, 184, 186
Descartes, René, 33 Gemasolar power plant, 44, 45
Dextral coiling, 29, 131, 195 Geometric series, 175
Difference Engine, 77, 78 Ghyka, Matila, 38
Diictidon, 140 Gnomon, 23–25, 73–78, 142, 147
Dinosaurs, 2, 57, 140, 144, 177, 217 Golden
Diogenes, 175 section, 15–17
DNA, v, 31, 53, 56, 233 spiral, 15–18, 142
Drill bit, 160, 180 Graptolites, 126
Dürer, Albrecht, 33, 173 Great Mosque, 126, 127
Greek key, 154, 189, 190
Gyrostasis, 142
E
Epicycles, 92
Equiangular spiral. See Logarithmic spiral H
Escher, Maurits Cornelis, 41, 43, 70 Hair whorl, 71, 72
Escherichia coli, 182 Haliotis, 17, 18
Euclid, 7, 15, 73, 75, 137 Harmonograph, 66, 85–87
Euler, Leonhard, 106, 184 Harp, 81
Euler’s spiral, 198–200 Helicoid, 54, 179, 180, 182
Evolute, 36, 124 Helicoprion, 217, 218
Index 257

Helix, v, 8, 30, 41, 53–56, 66, 79, 80, M


102, 126, 127, 131, 158, 179, Maelstrom, 151–158
180, 182, 233 Mars, 119, 156
Hermit crabs, 30 Mersenne, Marin, 33
Hyperbolic spiral, 5, 68, 92, 120, 121, 171 Mescaline, 219, 220
Mirror, 30, 31, 44, 59, 117, 123, 142
Mistletoe, 141–146
I Monotonic functions, 3–5
Involute, 124, 183–187, 192, 201, 202, Moseley, Henry, 124, 125, 132
213–215 Moth, 39–40
Ionic volute, 211, 215, 216 Mouflon, 27

J N
Julia set, 107, 108 Nautilus, 11–13, 17, 23, 25, 26, 37, 48,
57, 73, 147, 159, 225
Newgrange, 190, 191
K Newton, Isaac, 35, 119
Karpinsky, Alexander, 217 Nipkow disk, 95, 96, 98
Kelvin–Helmholtz instability, 151–158
Kepler, Johannes, 170 O
Klüver, Heinrich, 219 Operculum, 131, 132

L P
Labyrinth, v, 115–117 Palaeocastor, 140
Lava coils, 156 Parastichies, 48, 49, 51
Leonardo da Vinci, 17, 157, 180, 181 Parker spiral, 206
Leviathan of Parsonstown, 59, 60 Parsons, William, 59
Lissajous pattern, 85 Peacock, 44–46
Lituus, 5, 82, 133, 134 Pedal, 36
Logarithmic rosette. See Circular Pegea, 183, 184, 186
tessellation Pendulum, 85–88
Logarithmic spiral, 5, 13, 17, 18, 23–27, Phaistos disk, 190, 191
33–41, 44, 46, 48, 66, 69, 71, 73, Phyllotaxis, 49, 51
82, 85–87, 89–92, 94, 99, 100, Pine cone, 48, 109
107, 114, 119–121, 123, 129, Plasmid, 56
132, 141–144, 147–149, 153, Plato, 145
159, 165, 166, 168, 170, 173, Pliny the Elder, 12
174, 185, 187, 195, 214, 215, Poe, Edgar Allan, 151
217, 218, 222, 223, 225 Polar coordinates, 3–5, 9, 13, 36, 38, 50,
Lorentz’ law, 162 153, 205, 206, 220, 222
Loudspeaker, 26 Polygonal spiral, 141–145
Loxodrome, 69–71 Propeller, 101, 180, 182
258 Index

Q T
Quetzalcoatl, 195, 196 Tarxien temple, 190
Tatlin, Vladimir, 102
Teeth, 27, 140, 149, 165, 184, 186, 208,
R
217, 218
Rat fish, 218
Televisor, 95–98
Raup, David, 127
Thaetetus, 145
Rhumb line, 69
Toroidal helix, 55, 56
Rifle, 53–56, 67
Torricelli, Evangelista, 34
Right hand rule, 162
Tower cranes, 125, 206
Rock carvings, 194
Trees, 31, 159, 207, 229, 231
Rope, v, 53–56
Triskele, 190, 191
Roulette, 37
Trochospiral shells, 53
Turbines, 19–21, 44, 58, 208
S Turbo, 131, 132
Salps, 183, 184, 186 Turing, Alan, 220
Samarra, 126, 127 Twining plant, 79
Saturn, 152
Shaligram-shila, 207–208
Shankha, 208–210 V
Sinistral coiling, 29 Verne, Jules, 29
Situs inversus, 31 Viking ships, 192
Smithson, Robert, 101, 102, 142 Vishnu, 207–208
Solar Vitruvius, 179, 211, 213, 214
power, 45, 50 Vogel’s sunflower model, 50, 51
wind, 205, 206 Vortex flow, 152, 153, 157,
Spiders, 222, 225, 226 203, 223
Spiral
casing, 19–21
grain, 229, 231 W
Jetty, 101–103, 142 Wallis, John, 25
Spiral of Theodorus, 145, 146 Wentworth Thompson,
Spirograptus, 126 D’Arcy, 27, 38
Spirorbis, 25 Whirl, 98, 99, 151
Spirorhaphe, 1, 2 Whirlpool Galaxy, 60
Staircases, 53, 65, 140 Whorl expansion rate, 13, 185
Stellarator, 56 Wilson, Charles T.R., 161
Sun, vi, 2, 39, 44, 57, 89, 101–103, 115, Woodhenge, 115, 116
189, 192, 194, 195, 205, 206 Wren, Christopher, 25
Sunflower, v, 48–52, 80
Sun ship, 156
Symmetry, 31, 36, 62, 66, 80, 123, 130, Z
155, 162 Zeno, 175

You might also like