Entwined
ry
of Light
and Mind
"A small gem
of a book”
— James Gleick,
The Washington Post
CATC HING
THE
Light
The Entwined History of
Light and Mind
Arthur Zajonc
OXFORD UNIVERSITY PRESS
New York Oxford
Oxford University Press
Oxford New York
Athens Auckland Bangkok Bombay
Calcutta Cape Town Dares Salaam Delhi
Florence Hong Kong Istanbul Karachi
Kuala Lumpur Madras Madrid Melbourne
Mexico City Nairobi Paris Singapore
Taipei Tokyo Toronto
and associated companies in
Berlin Ibadan
Copyright © 1993 by Arthur Zajonc
First published in 1993 by Bantam Books, a division of
Bantam Doubleday Dell Publishing Group, Inc.
666 Fifth Avenue, New York, New York 10103
First issued as an Oxford University Press paperback, 1995
Oxford is a registered trademark of Oxford University Press
AU rights reserved. No part of this publication may be reproduced,
stored in a retneval system, or transmitted, in any form or by any means,
electronic, mechanical, photocopying, recording, or otherwise,
without the prior permission of Oxford University Press.
Library of Congress Cataloging-in-Publication Data
Zajonc, Arthur.
Catching the light: the entwined history of light and mind I
Arthur Zajonc.
p. cm.
Includes index.
ISBN-13 978-0-19-509575-3 (Pbk.)
1. Light—History. 2. Light—Philosophy. I. Title.
QC352.Z35 1993
535'.09—dc20 92-20204
Printed in the United States of America
For my wife, Heide
Contents
Chapter 1
Entwined Lights: The Lights of Nature and of Mind
1
Chapter 2
The Gift of Light
10
Chapter 3
Light Divided: Divine Light and Optical Science
38
Chapter 4
The Anatomy of Light
58
Chapter 5
The Singing Flame: Light as Ethereal Wave
97
Chapter 6
Radiant Fields: Seeing by the Light of Electricity
124
vii
Contents
Chapter 7
Door of the Rainbow
161
Chapter 8
Seeing Light—Ensouling Science: Goethe and Steiner
188
Chapter 9
Quantum Theory by Candlelight
225
Chapter 10
Of Relativity and the Beautiful
253
Chapter 11
Least Light: A Contemporary View
292
Chapter 12
Seeing Light
330
Notes
345
Acknowledgments
371
Index
373
viii
I’ll tell you how the sun rose a ribbon at a time.
Emily Dickinson.
I am the one who openeth his eyes, and there is light;
When his eyes close, darkness falleth.
the Egyptian god Ra, 1300 B.C.
If the light rises in the Sky of the heart... and, in the
utterly pure inner man attains the brightness of the sun
or of many suns... then his heart is nothing but light, his
subtle body is light, his material covering is light, his
hearing, his sight, his hand, his exterior, his interior, are
nothing but light.
Najm Razi, 1256
All the fifty years of conscious brooding have brought me
no closer to the answer to the question, “What are light
quanta?” Of course today every rascal thinks he knows
the answer, but he is deluding himself.
Albert Einstein, 1951
CHAPTER I
Entwined Lights:
The Lights of Nature
and of Mind
Use the light that is within you to regain your
natural clearness of sight.1
Lao-tzu
In 1910, the surgeons Moreau and LePrince wrote about their
successful operation on an eight-year-old boy who had been
blind since birth because of cataracts.2 Following the operation,
they were anxious to discover how well the child could see.
When the boy’s eyes were healed, they removed the bandages.
Waving a hand in front of the child’s physically perfect eyes,
they asked him what he saw. He replied weakly, “I don’t know.”
“Don’t you see it moving?” they asked. “I don’t know” was his
only reply. The boy’s eyes were clearly not following the slowly
moving hand. What he saw was only a varying brightness in
front of him. He was then allowed to touch the hand as it began
to move; he cried out in a voice of triumph: “It’s moving!” He
could feel it move, and even, as he said, “hear it move,” but
1
CATCHING THE LIGHT
he still needed laboriously to leam to see it move. Light and
eyes were not enough to grant him sight. Passing through the
now-clear black pupil of the child’s eye, that first light called
forth no echoing image from within. The child’s sight began as
a hollow, silent, dark, and frightening kind of seeing. The light
of day beckoned, but no light of mind replied within the boy’s
anxious, open eyes.
The lights of nature and of mind entwine within the eye and
call forth vision. Yet separately, each light is mysterious and
dark. Even the brightest light can escape our sight.
As PART OF what I call “Project Eureka,” a friend and I have
designed and constructed a science exhibit in which one views
a region of space filled with light. It is a simple but startling
demonstration that uses only a carefully fabricated box and a
powerful projector whose light shines directly into it. We have
taken special care to ensure that light does not illuminate any
interior objects or surfaces in the box. Within the box, there is
only pure light, and lots of it. The question is: What does one
see? How does light look when left entirely to itself?
Approaching the exhibit, I turn on the projector, whose bulb
and lenses can be seen through a Plexiglas panel. The projector
sends a brilliant light through optical elements into the box
beside it. Moving over to a view port, I look into the box and
y at the light within. What do I see? Absolute darkness! I see
• nothing but the blackness of empty space.
On the outside of the box is a handle connected to a wand
that can move into and out of the box’s interior. Pulling the
handle, the wand flashes through the dark space before me and
I see the wand brilliantly lit on one side. The space clearly is
not empty but filled with light. Yet without an object on which
the light can fall, one sees only darkness. Light itself is always
invisible. We see only things, only objects, not light.
2
Entwined Lights: The Lights of Nature and of Mind
The exhibit reminds me of a conversation I had over dinner
with the Apollo astronaut Rusty Schweickart. I asked him about
his space walk, specifically about what he saw when looking
out into the sunlit emptiness of outer space. He replied that
although it was difficult to keep the brightly lit spacecraft and
other hardware out of view, if you could do so, then you saw
only the dark depths of deep space studded with the light of
countless stars. The sun’s light, although present everywhere,
(1 fell on nothing and so nothing was seen. Only darkness.
Darkness Within
Two lights brighten our world. One is provided by the sun, but
another answers to it—the light of the eye. Only through their
entwining do we see; lacking either, we are blind.
Arguably the best-studied case of recovery from congenital
blindness is the case of S.B., researched by the psychologists
Gregory and Wallace.3 On December 9, 1958, and January 1,
1959, a blind British male, fifty years old, received cornea
transplants. For the first time since he was ten months old, he
had complete functional use of his eyes. What did he see?
The guardians of S.B. had enrolled him at age nine in the
Birmingham Blind School, where he learned the trade of boot
repairing. Earning his living by that means, he lived a life that
was unusually independent for a blind adult, for example going
for long bicycle rides by holding on to the shoulder of a friend.
He enjoyed gardening and especially any kind of work with his
hands, and was a confident, cheerful, and clearly intelligent
man.
Examining him about a month after his operations, Gregory
and Wallace asked him about his first visual experience follow
ing the operation. S.B. responded by saying that he had heard
a voice, the voice of his surgeon, coming from in front of him
3
CATCHING THE LIGHT
and to one side. Turning toward the sound, he saw a “blur.”
S.B. was unsure what the blur was but reasoned that since he
had heard the voice of his physician, and knowing that voices
come from faces, the blur in front of him must be his physician’s
face. Faces, even long after the operation, were “never easy,”
S.B. reported. Nor were his struggles with seeing confined to
faces. Gregory and Wallace’s research with S.B. (and similar
research before and since) has made it clear that learning to
see as an adult is not easy at all.
After his release from the hospital, Gregory and Wallace took
S.B. to a museum of technology and science. S.B. had a long
standing interest in tools and was clearly excited at the prospect
of seeing what he had until now only handled or heard described.
They took him to a fine screw-cutting lathe and asked him to
tell them what stood before him. Obviously upset, S.B. could
say nothing. He complained that he could not see the metal
being worked. Then he was brought closer and allowed to touch
the lathe. He ran his hands eagerly over the lathe with his eyes
shut tight. Then he stood back a little and, opening his eyes,
— declared, “Now that I’ve felt it I can see it.”
In the case of S. B., the slow process of learning to see con
tinued for the next two years, until his death. Its slow pace and
limited success were sources of deep disappointment to him, as
they invariably are to all such patients. In many situations S.B.,
like others, came to neglect sight entirely, for example by leaving
the lights off at home and navigating in his accustomed way as
a blind man. In most instances the effort to see was simply too
great. Those newly given sight may give up completely, some
times even tragically ending their struggle to see by taking their
own life.
In his systematic study of sixty-six case histories of the re
covery of sight in those bom blind, M. von Senden concluded
that innumerable and extraordinary difficulties need to be over
come in learning to see. The world does not appear to the patient
4
Entwined Lights: The Lights of Nature and of Mind
as filled with the gifts of intelligible light, color, and shape upon
awakening from surgery. The project of learning to see inevitably
leads to a psychological crisis in the life of the patient, one that
can end with the rejection of sight. New impressions threaten
the security of a world previously built upon the sensations of
touch and hearing. Some decide it is better to be blind in their
own world than sighted in an alien one.4
During the last few decades, studies of recovery from con
genital blindness have been corroborated and extended through
vision research on animals. It is now clear, for example, that if
a cat is unable to see forms during the critical period between
the fourth week and the fourth month, even if its environment
is still light, then the cat will be blind forever. The optically
healthy organ of the eye alone is insufficient for sight. During
the first months of life, patterns are elaborated in the eye or •]
brain of the kitten by the act of seeing. Without the nourishment
of seeing in the first months, these structures decay or are never
developed. After the fourth month, the damage is irremediable.5
The natural development of human vision is very similar.
During a critical window in the first years of life, visual as well
as many other sensory and motor skills such as speech and
walking are formed. If this opportunity is missed, trying to make
up for it at a later time is enormously difficult and mostly un
successful.
In the case of Dr. Moreau’s eight-year-old patient, after the
surgeon had worked with the child for some months, the parents
forced him to give the child over to a welfare establishment. By
the following year all that the child had learned to see under
Dr. Moreau’s care was lost. Moreau’s account contains a note
of exhaustion and regret that even with his full attention, he had
accomplished little of permanence. The sober truth remains that
vision requires far more than a functioning physical organ. With
out an inner light, without a formative visual imagination, we
are blind. Moreau writes:
5
CATCHING THE LIGHT
It would be an error to suppose that a patient whose sight
has been restored to him by surgical intervention can there
after see the external world. The eyes have certainly obtained
the power to see, but the employment of this power, which
as a whole constitutes the act of seeing, still has to be
acquired from the very beginning. The operation itself has
no more value than that of preparing the eyes to see; edu
cation is the most important factor.... To give back sight
to a congenitally blind person is more the work of an educator
than of a surgeon.6
Moreau’s child held to those modes of knowing that were
familiar and reassuring to him: touch, hearing, smell. To do
otherwise, to see, would have required a superhuman effort. In
many ways we act like Moreau’s child. The cognitive capacities
we now possess define our world, give it substance and meaning.
The prospect of growth is as much a prospect of loss, and threat
to security, as a bounty. One must die in order to become.
Newly won capacities place us in a tumult of new psychic phe
nomena, and we become like Odysseus shipwrecked in a stormy
sea. Like him we cling tenaciously to the shattered keel of the
ship we originally set out upon, our only and last connection to
a familiar reality. Why give it up? Do we have the strength to
leave, to change? Perhaps the voices encouraging us to venture
out on our own belong only to the cruel Sirens? So we close our
eyes, and hold to what we know.
H Besides an outer light and eye, sight requires an “inner light,”
one whose luminance complements the familiar outer light and
transforms raw sensation into meaningful perception. The light
of the mind must flow into and marry with the light of nature to
bring forth a world. This urges on us a second inquiry. Having
introduced the light of the mind, what, in fact, is the light of
nature?
6
Entwined Lights: The Lights of Nature and of Mind
The Darkness That Is Light
My “box of light” exhibit incites in the viewer the puzzling
question: what is the nature of this invisible thing called light
whose presence calls everything into view—excepting itself?
Over time we as a civilization have given many answers. We
have called it by the names of gods, or made it an action or
attribute of the divine. Even when Western science gave it a
more substantial nature, it always reflected our wonder and our
capacity to imagine. Early in the seventeenth century, Francis
Bacon marveled that the “form and origin of light” had been so
little investigated.7 Why had the specific nature of something
so important as light not been discovered? Almost four hundred
years later, we, like Bacon, are still naturally curious about
what light is made of, how big it is, how it moves, and so on.
In other words, we wish to know its physical nature.
In my own professional life, I first sought to understand light
by means of laboratory research in quantum optics. In laser
experiments performed at institutes in Boulder, Amherst, Paris,
Hanover, and Munich, I studied light and the way it touches
matter. The more I learned of the quantum theory of light,
theoretically and experimentally, the more wonderful light
seemed. Even armed with such sophisticated theories, I have
no sense of closure regarding our knowledge of light. Far from
it, light remains as fundamentally mysterious as ever. In fact,
quantum theory has taken the simplistic, mechanistic concep
tions of light provided by early science and, on the firm basis
of experiment, shown them all to be impossible. In their place,
it has framed a new theory of light that every great modern
physicist from Albert Einstein to Richard Feynman has struggled
to understand—unsuccessfully, as they realized themselves.
Once I understood that for all the power, precision, and beauty
of quantum optics, we still do not know what light is, I got
excited. The old scientific idols of light, like outdated effigies,
7
CATCHING THE LIGHT
have been destroyed; and every attempt to fashion new ones has
failed. Our technical mastery of light has opened all the doors
once closed by a hasty scientific arrogance. I could not resist
entering all the passageways, old and new, within the many
roomed mansion of light. This book is the story of what I found
there.
The first thing I discovered was that light has gathered around
it innumerable artistic and religious associations of extraordinary
beauty. It has been treated scientifically by physicists, sym
bolically by religious thinkers, and practically by artists and
technicians. Each gives voice to a part of our experience of
light. When heard together, all speak of one thing whose nature
and meaning has been the object of human attention and ven
eration for millennia. During the last three centuries, the artistic
and religious dimensions of light have been kept severely apart
from its scientific study. I feel the time has come to welcome
them back, and to craft a fuller image of light than any one
discipline can offer.
Light touches all aspects of our being, revealing a part of
itself in each encounter. A history of those meetings can lead
us around and into the nature of light. Long before it was the
object of scientific study, light, and especially the sources of
light, were venerated as divine—an image of godly nature.
Mythologies of all civilizations are rich with tales of the sun,
moon, and stars, of fire, rainbow, and aurora. These, too, touch
on the being of light, for they are part of the human experience
of it. In what follows I will speak equally of the quantum theory
of light and the Zoroastrian god of light, Ahura Mazda. I will
approach light from many sides, mythic and spiritual as well as
historical and technical. Different ages and different peoples
have been drawn to one or another of light’s many parts. As I
studied, it seemed to me that the characteristics of a culture
are mirrored in the image of light it has crafted. Each, in its
own way, has attempted to uncover light’s nature and meanings,
8
Entwined Lights: The Lights of Nature and of Mind
and so authored a tale of light. In the telling of that story, the
culture reveals as much about itself, about the light of its peo
ple’s minds, as about nature’s light. These twin themes twine
their strands around the central axis of these pages like the
serpents that spiral the healing staff of Hennes, the god of -»•
communication: the changing nature of two lights, the outer light
of nature but also the inner light of the mind. I have grown
convinced that the two are inseparable.
So, as we follow the historical path of light, we will attend
to both the changing ideas of light and the changing human
consciousness that studied it. We have looked into the coun
tenance of natural light for many, many years, wondering what
or who it is. Light has grown old during the millennia of our
looking, its features changing utterly, its tender, childhood face
nearly completely disguised. Light displays now a sterner, more
useful and mathematical countenance, but even today other
visages—artistic, scientific, and spiritual—complement the for
mer. What will light look like tomorrow? Throughout all times
and images, the same sun has warmed the earth and lit the
planet. From the birth of the very idea of light, through its
contemporary maturity, to its final form at the end of time, light
will have sighted whole kingdoms and nourished prairie, tree,
and flower. How have we changed this thing called light through
the lights of our own consciousness? In the mingling of nature
and mind arises an understanding of the life of light. This book,
therefore, is a biography of that invisible companion who ac
companies us inwardly as much as it does outwardly—light.
9
CHAPTER 2
The Gift of Light
Thou sun, of this great world both eye and soul.
Milton
As the beasts of the earth were being created, the Titan Epi-
metheus (whose name means “afterthought”) assumed the task
of giving to each some faculty for their protection and survival.1
To the turtle he gave a hard shell, to the wasp his stinger, to
others speed and cunning. When he finally came to the human
species, all nature’s powers had already been allotted; nothing
at all was left for man. In Plato’s words, man remained “naked,
unshod, unbedded and unarmed.” Despondent, the bungling
Epimetheus turned to his wise brother Prometheus (whose name
means “forethought”). Faced with the helplessness of man, Pro
metheus daringly stole from Zeus the gift of fire, carrying it to
mankind as ancient mariners often carried hot embers, in a giant
fennel stalk. From the light of Prometheus’ gift, man has kindled
his civilizations, his cultures and technologies. The fire and
light of Zeus became the possession of mankind.
For his actions on our behalf, Prometheus was cruelly pun
ished by being chained to the mountains of Caucasus, where
each day his liver, the seat of life, was tom from his side and
10
The Gift of Light
eaten by an eagle sent from Zeus. Nor was mankind left to enjoy
in peace the gift of Prometheus. Zeus, angered and jealous,
commanded the lame craftsman of the gods, Hephaestus, to
fashion a seductive automaton, Pandora, whose infamous box
Epimetheus greedily accepted. Too late he saw its evil contents,
Against his will Pandora lifted the lid and so unleashed on
humanity illness, grief, and pain.2 The gift of fire and all it '
symbolizes is invariably linked with the burden of care. Under
human control, the fire of the gods bums as well as warms,
blinds as well as illumines.
Western CIVILIZATION began with the song of a blind bard three
thousand years ago who, in singing the Iliad and the Odyssey,
gave voice to the Greek imagination, and so begat Western
poetry. Homer’s blindness lent his words purity and power. His
sense world dark, a godly world rose to take its place, and
Homer’s memory seemed to reach back to archetypal deeds and
an eternal heroic age.
Greek vases show the standing bard rocking as he sang,
wrapped in an inner glory, listening, as he spoke, to a voice
that sang within him. Like Homer, the old wandering minstrels
of north Karelia around the Baltic Sea would rock, eyes closed,
sitting upon a log bench, arms locked with those of a peasant,
chanting antiphonally their ancient epic, the Kalevala.
The Bhagavad-Gita, or “song of god,” is the sung response
of the minister and charioteer Sanjaya to the questions of the
blind king Dhritarashtra. The mightiest earthly power, the king,
is blind. He sees through the eyes of another, his charioteer
and counselor, whose spiritual gifts extend his sight. When the
king asks about the happenings on a distant, sacred field where
those he loves prepare to battle one another, Sanjaya is able to
see and hear the intimate conversation between the accom
plished prince Arjuna and the divine Krishna, who like himself
11
CATCHING THE LIGHT
has taken on the form of a charioteer. Here the soul-spiritual
faculty of higher sight is separated out, into the person of the
charioteer. He becomes the bard who sings to a blind and worldly
royalty. The charioteer, like the poet, must see further than
others, and speak and steer according to what he sees.
Is it only coincidence that the most renowned soothsayer of
antiquity, Tiresias, was blind from his seventh year? He lost
ed his sight for seeing the goddess Athena bathing, that is, for
I seeing a god unveiled.
The motif is eternal. The light of day makes way for the light
of night, of blindness, of inner sight. As Plato wrote: “The mind’s
eye begins to see clearly when the outer eyes grow dim.”3 The
Romantic poet Novalis understood fully the efficacy of darkness.
His Hymns to the Night begins with a sublime antithesis: “What
living, sense-endowed being does not love above all the won
drous appearances spread around him, the glorious light....”
And yet Novalis tells us that for all its beauty, he turns away
from the day “to the holy, unspeakable, mysterious night.” Out
of the dark solitude of loss comes the poet’s light and voice. In
the midst of outer darkness, of blindness, an inner light illumines
an imaginal landscape of beauty and reality. The blind bard
sings the world he sees into the hearts of his listeners so that
they, too, might, for an evening, shed the cares of their world
for the beauty of his.
WHAT IS THE source of poetic light that illumines the night of
Homer’s blindness? It is imagination, which is also important
to common sight. The light of imagination will occupy half of
our history, because of its significance for both the ancient world
and poetry and the present world and science. No matter how
brilliant the day, if we lack the formative, artistic power of
imagination, we become blind, both figuratively and literally.
We need a light within as well as daylight without for vision:
poetic or scientific, sublime or common.
12
The Gift, of Light
As we will discover, the mind is subtly and usually uncon
sciously active in sight, constantly forming and re-forming the
world we see. Thus, we participate in sight. The habitual pat
terns by which we see are set down in the first years of life.
Even the simplest, most “objective” acts of cognition require
our involvement. In addition, the nature of that participation is
specific to each culture and historical period. A Coke bottle
dropped from an airplane into a society of bushmen may be seen
as many things, but never as a container for a carbonated bev
erage. Human consciousness has changed over time and differs
between cultures.
In antiquity, our role in seeing, in granting meaning to the
sense world, was felt more keenly than today; the inner light
was closer to consciousness. Unlike the ancient Greeks, we live
habitually in a scientific world view that too often treats our
participatory role in cognition as unessential or illusory. Yet to
see, to hear, to be human requires, even today, our involvement,
our ceaseless participation. An example will intensify the ar
gument: the puzzling phenomenon of Greek color vision.
The Wine-Dark Sea of Antiquity
The sun rose on the flawless brimming sea into a sky all
brazen—all one brightening for gods immortal and for
mortal men on plowlands kind with grain.*
Homer, The Odyssey
The atmosphere and landscape of Homeric Greece appears at
once alike and apart from our own. The sun still rises over
plowlands kind with grain, but few of us waken to a bronze sky -
brightened by gods immortal.
Walking on island shores, a captive of the beautiful nymph
Kalypso, Odysseus gazed longingly on the “wine-dark sea”
yearning to return to his native Ithaka and his beloved wife,
13
CATCHING THE LIGHT
Penelope. Standing today on an island coastline in the Aegean,
I see neither a wine-dark sea nor a brazen sky, but the extraor
dinary blue of water and heavens I so love.
Of the countless epithets attached by Homer to the sky or
sea, none, say linguists, can be understood to mean “blue.”
The sky may be referred to as “iron” or “bronze,” the sea as
black, white, gray, purple, or wine-dark—but never blue. Did
the ancient Greek lack the experience of blue, was he partially
color-blind? Or do we perhaps see here another instance of the
presence of an interior light, of the activity of vision? Since
1810, when Goethe first pointed to the curious lack of blue in
Greek usage, scholars have been puzzled by it and similar ab
sences in the color language of early Greek poetry.5
From the careful analysis of color terms that do exist in ancient
Greek, and our modem knowledge of color blindness, compel
ling arguments have been made against the hypothesis that the
Greeks possessed a physical eye different from our own. But
then we have shown that sight entails more than an operating
physical organ. In considering the following examples of color
vision in Homer’s Greece, bear in mind the significant inner,
psychological pole of sight. If we do so, then we may unravel
the puzzle that has confounded so many.
Some five hundred years after Homer, the great student of
Aristotle, Theophrastus, wrote a treatise on stones in which he
described a stone called kyanos, a stone we now identify with
the precious blue mineral lapis lazuli. When we encounter
kyanos in its adjectival form, it is natural to think of it as referring
to blue (related to our term “cyan”). Although the association
seems natural enough, occurrences in Homer defy that inter
pretation.
In fury and grief at the loss of his friend Patroklos, Achilles
slew Hektor, pierced the heels of this noble son of Priam, and
attempted to defile his body by dragging it for twelve days on
the plains of Troy. In doing so, “a cloud of dust rose where
14
The Gift of Light
Hektor was dragged, his k-yanos hair was falling about him.”6
Are we to understand that Hektor had a head of blue hair? In
order to stay his senseless debasement of a worthy prince and
warrior, Zeus sent Iris to Achilles’ immortal mother, Thetis, on
the sea’s floor. Iris, “storm-footed,” sprang into the sea and,
finding Thetis, bade her join Zeus. Ashamed to mingle with the
gods, Thetis “took up her kyanos veil, and there is no darker
garment,” and followed Iris to Olympus.7 From these and many
other instances we learn that kyanos meant dark rather than
blue. Yet there was no other word for blue in Homeric Greek.
Homer and other early poets simply lacked a term for blue. To
them blue was not a color in our sense, but the quality of
darkness, whether describing hair, clouds, or earth.
Similarly for chloros, the term that later Greek color theorists
call green, a puzzle confronts us. In the Iliad, honey is chloros;
in the Odyssey, so is the nightingale; in Pindar, the dew is
chloros, and with Euripides, so are tears and blood! From its
use, we can see it means not green but moist and fresh—alive.
We still speak of unseasoned wood or an untrained laborer as
green. For the ancient Greek, these connotations were the pri
mary meanings. So disconnected from the external perception
of color were they that the psychological quality of “freshness”
or “darkness” could become the perceived attribute. They saw
the moist freshness of tears and so saw green. When enraged,
we may remark metaphorically that we “see red.” I would suggest
that we should understand the Homeric world’s use of such color
expressions not metaphorically but literally. Neither the sunlight
nor their eyes were different from ours. Rather, what they
brought as the interpretive light of an antique imagination
changed the way they saw, just as a similar light continues to
shape the way we see today.
A more recent related example is provided by “The Case of
the Colorblind Painter” reported by Oliver Sacks and Robert
Wasserman in 1987.8 Jonathan I. had been a successful painter
15
CATCHING THE LIGHT
until, at age sixty-five, he had a minor car accident. He suffered
a concussion and the trauma normally associated with such an
accident, but took away no lasting physical injury. He did,
however, become completely color-blind, a persistent condition
whose sudden and unaccountable onset was triggered by the
accident. He saw the world, in his own words, as if “viewing a
black and white television screen.” The case is moving and
tragic. An artist whose entire life had been lived through color
now saw none. Ophthalmologists and neurologists including
Sacks and Wasserman subjected Mr. I. to the full panoply of
medical tests to no avail. The cause of his color blindness has
remained a mystery. Sacks and Wasserman summarize their
study by saying, “Patients such as Mr. I. show us that color is
not a given but is only perceived through the grace of an ex
traordinary complex and specific cerebral process.” Moreover,
while physiological computations may well be going on, color
vision “is infinitely more; it is taken to higher and higher levels,
admixed inseparably with all our visual memories, images, de
sires, expectations, until it becomes an integral part of ourselves,
our lifeworld.”
The “lifeworld” of Homer on the shores of Troy was profoundly
different from our own. His memories, associations, desires,
and expectations were unlike those we bring to the battlefield.
The mental organ of sight used by the blind bard was then a
cultural commonplace, but it was also significantly different from
the mind-set with which we see today. We need to soften the
notion of ourselves as equipped with fixed vidiconlike eyes and
static computerlike brains to produce the equivalent of con
sciousness. The blossoms of perception unfold out of a far richer
and self-reflexive union of mental and natural lights.
In the cases of S.B. and Mr. I., we confront a situation where
persons could not see what we would all agree was in fact
“there,” before their eyes. They lacked an interior light and so
remained functionally blind. The opposite situation occurs when
someone sees something we would say is “not there.” We nor-
16
The Gift of Light
mally term such experiences hallucinations. They arise when
an individual’s psychological state is sufficiently strong to fab
ricate an experience akin to that produced by the senses. Did
the ancient Greeks hallucinate the color of their wine-dark sea?
The linguistic evidence suggests otherwise, or we would have
to imagine an entire culture collectively hallucinating. Yet, in
a certain sense, their inner emotions or stage of development
did “color” the world they saw. Study of other language groups,
for example Chinese or American Indian languages, supports
the interpretation that cultures see the world, down to its very
textures and colors, in ways significantly different from our
own.
Over millennia, the two lights of nature and mind have in
teracted to present differing worlds to different ages. Like a
blind bard granted sight, we will have difficulty at first imagining
Poseidon.
CATCHING THE LIGHT
our way into ancient understandings of sunlight and the sighted
eye. They will appear, initially, unfamiliar and even absurd.
Yet the strangeness may be largely a reflection of the modem
imagination we bring to ancient experiences. At every stage, we
will need to reimagine the universe, to participate in it em-
pathetically in order to hear the epic song of light.
The magnificent bronze statue of Poseidon, lifted from the
Aegean to grace the National Museum of Athens, has only
dark hollows where eyes once were. They were not empty in
450 B.C., when special inlaid gems filled those now sightless
pools. Here was the sacred seat of a special glory that brought
cognitive life to the supple and powerful figure. When toppled
from his pedestal and cast into the sea, the gems—Poseidon’s
eyes—were stolen. The god who once ruled the sea, now blind,
was deposed. Our story of light begins with the ancient and
sacred understanding of the eye, so close to light. Empedocles,
physician and demi-god, will reset the gemstones in Poseidon’s
face. Later, others will remove them.
The Lantern and the Eye
w The light of the body is the eye: if
therefore thine eye be single, thy whole body Lt? 6“^
shall be full of light. But if thine eye be
evil, thy whole body shall be full of
darkness. If therefore the light that is in
thee be darkness, how great is that darkness!
Matthew 6:22-23
In his book on ancient philosophers, Diogenes Laertius tells the
tale of a pestilence that struck the Sicilian city of Selinus in the
middle of the fifth century B.C.10 From the stagnant, sewage-
filled waters of the main river there arose disease and death,
claiming the lives of “both citizens and women.” Hearing of
Selinus’ distress, the noble physician, scientist, statesman, and
18
The Gift of Light
poet Empedocles came from the neighboring town of Acragas
dressed in the purple robe of wealth, girded with a belt of gold.
On his feet he wore bronze slippers, on his head a laurel wreath;
behind him followed a train of young boys to attend his needs.
Discovering the source of the pestilence, Empedocles caused
two nearby rivers to be rechanneled, mixing thereby their sweet,
flowing waters with the rank river water of Selinus, and so
relieving the Selinuntines from their ills.
The story is entirely believable, considering the appalling
state of rivers around urban centers in present times. Paris,
perched upon the He de la Cite, was famous even in the Roman
period for the stench of the Seine; the Baltic Sea, once a mag
nificent natural resource, has become the poisonous dumping
ground for industrial Poland. For Empedocles to have identified
the cause of disease, conceived a plan of action, and then
engineered and excavated at his own expense a canal suited to
dispersing the fetid waters of Selinus is an action deserving of
high commendation. So we are not surprised to learn that when
Empedocles appeared to the Selinuntines afterward, he was
praised and worshiped as a god. What does surprise us is that,
in response to their praises, Empedocles threw himself into a
fire, apparently without harm, to confirm their opinion of his
divinity.
In Empedocles, we find not only an impressive early scientist
whose specific ideas on vision will concern us, but also the last,
belated example of a human type that vanished from Greece at
his remarkable death when he disappeared into the volcanic
crater of Mount Etna, bronze slippers and all. Empedocles was
not only a scientist-physician but also a poet and shaman who
wrote, in addition to his insightful book On Nature, a rather
more puzzling spiritual-religious tract, Purifications .u We pos
sess only fragments of these two works, but from them we can
still form an opinion of the range and character of Empedocles’
person and thought.
In the Purifications, we learn from Empedocles himself of his
19
CATCHING THE LIGHT
divine origins, that he has been condemned to “wander thrice
l ten thousand seasons far from the company of the blessed, being
I bom throughout the period into all kinds of mortal shapes.”12
He is a god sentenced to live as a bird, a mortal, and in countless
other forms for the horrid transgression of eating the human flesh
of sacrifice. As we study his account of sight, remember the
paradoxical blend that is Empedocles, shaman and scientist in
fifth-century Greece, less than one hundred years before Plato.
The world of Homer, of the gods, the pageants of mystery re
ligions, and the guarded rites of initiation are not distant from
the natural science of early Greece. A spiritual cosmos provided
the protecting chambers in which the birth of natural science
took place.
According to Empedocles, the divine Aphrodite, goddess of
love, fashioned our eyes out of the four Greek elements of earth,
water, air, and fire, fitting them together with rivets of love.13
Then “as when a man, thinking to make an excursion through
the night, prepares a lantern,” lighting it at the brightly blazing
hearth fire and fitting it around with glass plates to shield it
from the winds, so did Aphrodite kindle the fire of the eye at
the primal hearth fire of the universe, confining it with tissues
in the sphere of the eyeball.14 Marvelous passages were fitted
. into the eye, permitting it to transmit a fine interior fire through
I the water of the eye and out into the world, thereby giving rise
to sight. Sight proceeded from the eye to the object seen; the
eyes rayed out their own light.
When, a few hundred years after Empedocles, the evangelist
Matthew wrote, “the eye is the light of the body,” he was thinking
not only metaphorically but scientifically. The image of the eye
as a lantern was a cultural and scientific commonplace when
Matthew was composing his gospel.15
In the struggle to unravel the mystery of sight, sunlight played
a lesser role. Empedocles recognized the existence of the sun’s
light; that much is clear from other fragments such as “It is the
20
The Gift of Light
Earth that makes night by coming in the way of the [sun’s] rays,”16
an astute observation for the time. He seems, however, to
have considered sunlight as only part of the whole process, and
recognized that something more was required for vision, some
thing essential provided by man: the light of the body.
Platonic Sight
Plato, like Empedocles, was permitted to study the secret doc
trines of Pythagoras, at least until he (again like Empedocles)
betrayed Pythagoras’ teachings to the uninitiated through his
writings. Plato’s account of vision is, not surprisingly, similar
to, if fuller than, that of Empedocles. When blended together
with the later geometrical tradition of sight begun by Euclid and
the medical tradition codified by Galen, Plato’s treatment would
persist for almost 1,500 years! In this tradition, the light of the
eye played fully as important a role as the light of the sun.
According to Plato, the fire of the eye causes a gentle light ||
to issue from it. This interior light coalesces with the daylight,
like to like, forming thereby a single homogeneous body of light. ■=•
That body, a marriage of inner light and outer, forges a link
between the objects of the world and the soul. It becomes the
bridge along which the subtle motions of an exterior object may
pass, causing the sensation of sight.17
In this view, two lights—an inner and outer—come together
and act as the mediator between man and a dark, cavernous
external world. Once the link of light is formed, the message
may pass, like Iris, Homer’s messenger goddess, from one world
to the other. The eye and the sun display to Plato a deep har
mony, one still appreciated by the German poet Goethe when,
in the introduction to his own Theory of Color (1810), he penned
the poem:
21
CATCHING THE LIGHT
Were the eye not of the sun,
How could we behold the light?
If God’s might and ours were not as one,
How could His work enchant our sight?is
Once again, the mind’s eye is not passive, but plays its own
significant part in the activity of seeing. The image of an interior
ocular fire captured vividly the ancient sense of that action, so
convincingly that it dominated philosophy for 1,500 years.
We come to know the world, in large part, through sight.
Quite naturally, Plato used sight as a metaphor for all knowing,
calling the psyche’s own organ of perception the “eye of the
- soul” or “mind’s eye.”19 Our word “theory” has its origin in the
Greek word theoria, meaning “to behold.” To know is to have
seen, not passively but actively, through the action of the eye’s
fire, which reaches out to grasp, and so to apprehend the world.
Our activity, present in seeing and knowing, is an element
integral to the Platonic understanding of vision. Sight entails
the seer in an essential, formative action of image making or
imagination. To such as Moreau’s child or to S.B., the effort of
that constructive act was a constant and exhausting reminder of
their past blindness. To us who see, the world is instantly and
effortlessly intelligible; at least most of the time.
Consider the figure on page 23. It is but one of many similarly
“ambiguous figures.” Allow yourself time to play with it. At first
only one figure shows itself, an old woman or a young girl.
Without an iota’s change on the “objective” printed page, the
delicate chin of the young girl becomes the lumped nose of an
old hag. Feel the shift from one picture to the other. It takes
place entirely within you. With a little practice you can even
control what you see.
The physical difference between one image and the other is
nil, while the “soul distance” between them is huge. What has
changed? Your own activity; the character of your participation
22
The Gift of Light
Old woman or young maid?
can shape and reshape itself, and you can feel it. With every
act of perception, we participate unawares in making a mean
ingful world. In response to outer light, an interior light flashes,
bringing intelligence with it. It is the light that did not brighten
the newly opened eye of Moreau’s child when turned to see its
first light.
Times of Transition
In the Bhagavad-Gita, in Homer, Empedocles, and Plato, vision
entails an essential human activity of movement out from the
eye into the world. In the centuries following Plato, a shift
gradually took place that only reached its conclusion with Rene
Descartes in the seventeenth century. The concerns of science
changed during this long period. The influence of Plato and then
of Aristotle lingered long into the medieval period. As long as
this was the case, sight was as much or more a soul-spiritual
23
CATCHING THE LIGHT
process as a physical one. By the sixteenth century, however,
a profound shift seems nearly mature. Natural philosophers such
as Kepler and, to a much greater extent, Galileo are less con
cerned with the soul’s translation of external stimulation into
meaningful perception, and more preoccupied with the physics
of the eye viewed as an inanimate, physical instrument. The
change is not universal, swift, or uniform, but a watershed is
crossed nonetheless, first by those few scientists in the oft-
dangerous vanguard of research. In their hands, sight becomes
a question of mechanics rather than a species of soul-spiritual
activity so characteristic of many earlier thinkers.
The shift is characteristic and of central importance. We meet
it first in the evolution of man’s experience of seeing. We will
discover it again when we study light itself. What begins as a
lively, soul-spiritual experience, be it of light or sight, atten
uates, clarifies, and divides into optics and psychology. More
than an interesting historical observation, our changing view of
light is symbolic of a major change in consciousness, an im
portant threshold crossed in the history of the mind.
Like the ambiguous figure, nature presents herself in indef
inite guises. How we see her depends as much on us as on her.
Only together do meaningful worldly images arise. The wa
tershed crossed, therefore, is not the divide between ignorance
and wisdom, but more like the ambiguous shift from young girl
to old woman. Therefore as we read the history of science, we
must be ever conscious of the individuals who enacted it. Their
eyes saw, their hearts yearned for knowledge, and out of their
being ways of seeing the world were bom, flourished, and died.
One way of seeing became for a time the way of many, until a
fresher, more congenial view appeared.
The delicate beginnings of the transition to a mechanical
conception of seeing were evident by 300 B.C. in the optical
studies attributed to the great Alexandrian mathematician Eu
clid. In his book Optics he provided a brilliant geometrical
24
The Gift, of Light
treatment of sight. Euclid continued to believe that a visual ray •
was primary to the whole process of vision, and advanced several
very sensible arguments in favor of the position.
For example, we often do not see things even when looking
at them. Drop a needle on the ground, Euclid suggests, and
then wonder as you search for it why you don’t see it immediately.
Your field of view certainly encompasses the needle. In modem
terms, the needle is certainly imaged on the retina, but remains
unseen. Then suddenly, in a flash, you see it. If sight depends
only on light from outside falling on objects, and then traveling
into the eye, one would see it immediately. Obviously light was
being reflected from the needle and into the eye throughout the
search, so, reasoned Euclid, sight cannot in the first place
depend on external light. The puzzle is solved, however, if we
adopt the doctrine of the visual ray. In searching for the needle,
the eye’s own visual ray reaches out and passes back and forth
across the ground. Only when it strikes the needle do we see
it!
The visual ray of Euclid is different in important ways, how
ever, from the luminous and ethereal emanation of Plato and
Empedocles. In Euclid’s hands, the eye’s fiery emanation has
become a straight line, a visual ray, susceptible to deductive
logic and geometric proof. His extensive mathematical studies
yielded many fruits and became the basis for later Arab inves
tigations and for laying the foundation for the discovery of linear
perspective by Brunelleschi, Alberti, and Diirer centuries later.
But mathematization came at a price. It distanced man from the
earlier and more immediate experience reflected in the Platonic
understanding of vision.
The significance of mathematization should not be underes
timated. Without abstraction, science as we know it cannot exist.
Yet in order to analyze one must stop experiencing and go on
to represent the object of study with thoughts of crystalline
clarity, for example, with mathematical concepts. Euclid did
25
CATCHING THE LIGHT
just this. Plato’s somewhat elusive, immaterial bridge of light
between object and eye, became through Euclid a geometry of
visual rays, cones, and angular measurement. Everything
needed for the study of geometrical optics was developed, but
in the process one can detect an important distancing from the
subjective human experience of seeing. Euclid’s meticulous
mathematical style of argumentation has replaced the more po
etic treatment of Empedocles or Plato. As every physicist knows,
the elegant forms of mathematics can easily outshine the dull
stirrings of experience, and eventually come to replace the phe
nomena they originally were invented to describe. Euclid’s han
dling of light foreshadows the growing separation of sight as
lived experience from sight as a formal object of investigation.
The history of light has turned a comer, and with it the mystery
of sight entered a new phase, one that blossomed first in Arab
lands, to culminate finally in the work of another great geometer
and mathematician, Rene Descartes.
The Arab Connection
Toward the end of the Roman empire the stage was set for further
developments in the history of the mind. The closing of the
_ Platonic Academy in A.D. 529 by Justinian was the final death
knell of Greek philosophy in the West and the dawn of the Dark
Ages. For many centuries, the Academy had been a sanctuary
in which the ideas of Plato and his followers flourished. With
the rise of Christianity, however, pagan thought was in danger
of being eradicated. In A.D. 389, the great library in Alexandria
with its half-million scrolls was destroyed by rioting Christians.
Under a state that sanctioned the Roman Church, the Platon-
ists, who still revered the pagan gods, were persecuted and
hounded as dangerous heretics. When Justinian’s soldiers swept
into the Platonic Academy, the last disciples of Plato had to
26
The Gift of Light
flee Athens. The seven great sages of the Academy departed
with their precious books bound for Persia where the emperor
Khurso I received them graciously at his magnificent summer
palace in Jundishapur (near what is today Dizful, Iran).20
In the court of Khurso I, and at the illustrious Academy of
Jundishapur, literature, the arts, science, and philosophy flour
ished. The Athenian refugees found here a cosmopolitan at
mosphere of remarkable tolerance. The indigenous religions of
Zoroastrianism and Manichaeism mingled with eastern religious
thought, as well as pagan, Christian, and Jewish influences.
Jundishapur was founded as a prisoners’ camp following the de
feat of the Roman emperor Valerian in A.D. 260 by Shapur I.
By the sixth century it had become the greatest center of learning
in the world, boasting an outstanding astronomical observatory,
medical school, and the world’s first hospital. Jundishapur was
known then, and for centuries thereafter, for its physicians and
wise counselors. The rise of Islam blunted the impact of Jun
dishapur, but the leaders of Jundishapur’s Academy were the
nucleus around which the scholarship and learning of Islam
formed.
With the rise of Islam in the seventh century, a cultural
revolution of unprecedented scope took place on the Arabian
peninsula. Following the establishment of the new religion by
Mohammed and a system of governance for the vast empire won
through holy wars, Islamic scholars became tremendously active
in collecting and translating Greek manuscripts. Baghdad, dur
ing the ninth century under the guidance of the scholar and ^>©5
translator Hunayn ibn Ishaq, became a great center of learning,
and Arab science and scientists rose quickly in importance.
While thinkers in the West forsook the concerns of Hellenism
for religious questions, especially the matter of salvation, phi
losophers and physicians of the Islamic Near East, under the
influence of Jundishapur, were busy mastering, commenting on,
and furthering the knowledge of antiquity.
27
CATCHING THE LIGHT
The famous philosopher, mathematician, astronomer, and op
tician Ibn al-Haytham figured prominently in these develop
ments.21 In his hands, the history of sight took another significant
step away from earlier and more spiritual or psychological views
and toward a mathematical and physical theory of vision.
Born in Basra (Iraq) in A.D. 965, Ibn akHaytham, or Alhazen
as he came to be known in the West, became the greatest optical
scientist of his age. As a child and young man, Alhazen had
attempted to attain knowledge of truth through the Islamic re
ligious sciences of his day. Dismayed by the elusiveness of this
goal and the rancor he saw between competing religious sects,
he resolved to concern himself with a “doctrine whose matter
was sensible and whose form was rational.”22 Truth was one, he
felt, and throughout the following decades he maintained his
initial resolve to avoid the vagaries of the spiritual sciences.
Instead, he produced dozens of treatises on mathematical and
scientific subjects, the most influential of which was his Optics.
One hundred and fifty years after his death in 1040, the Optics
was translated into Latin and subsequently became the foun
dation for future optical research. Two aspects of his work will
be of special concern to us: his replacement of the Platonic
theory of vision with his own quite different theory, and his
study of the camera obscura. Both reflect Alhazen’s reimagin
ation of light.
The PROMINENT Greek accounts of vision had given full weight
to the inner activity of the seer. As we have seen, this came to
be embodied in their view that a pure fire, essential to sight,
resided within the eye and rayed out, sunlike, to illuminate the
world. This view was taught in various forms in the West until
the twelfth century, for example by the great teacher William
of Conches at the cathedral schools of Chartres and Paris. A
profound student of Plato, Conches also drew from Galen the
28
The Gift of Light
view that food was transformed from matter into a spiritual light
in a series of stages. Its first transformation occurred in the liver,
where it became “natural virtue.” Passing then on through the
heart, it became “spiritual virtue,” moving finally into the brain,
where it was refined into a luminous wind that animated the
organs of sense and provided the interior ray of the eye.23
Another important Greek school of thinking held that vision
occurred by the transmission of husks or forms (called eidola
or simulacra) from the object to the eye. The atomists of the
Greek world believed that films or images peeled off objects, or
were impressed onto the air by them, and streamed to the ob
server, where they entered the eye. The tiny reflected image of
the world that is visible when we look at the dark pupil of our
neighbor’s eye was taken by them as evidence of these husks.
Obvious problems existed with this theory. How, for example,
did a husk the size of a mountain become small enough to enter
the eye?24 This view likewise found its reflection in the Middle
Ages, but setting it aside for the moment, we can return to the
Arab connection.
In the Near East, a view of vision was being developed that
was complementary to Plato’s view as taught at Chartres. It
emphasized outer light, and received a special impetus in the
Arab world. Alhazen marshaled a series of logical arguments in
support of the view that sight proceeded not partly, but entirely
by means of light entering the eye from objects around us. He
considered the following situation. One cannot long look at the
sun without great pain. If one maintained a theory in which the
flow was away from the eye, then how could there be pain? If,
however, there were some kind of transmission from the sun to
the eye, its overwhelming action on the eye could account for
the discomfort. Another of his arguments concerned afterimages.
Stare at a bright light or a window for thirty seconds and then
close your eyes. A clear sense impression floats in view with
the same contours as the original but usually showing colors
29
CATCHING THE LIGHT
complementary to those of the lights seen. Again Alhazen took
this as evidence that something affects the eyes from without,
impressing itself on the eye so strongly that an effect persists
even after the light is extinguished. These, and many other
phenomena, were joined with carefully reasoned arguments to
reject the visual theories of Plato and others. Alhazen conceded
that mathematicians may still find it useful to draw “visual rays”
from the eye to the object for the geometrical study of light, but
in doing so they “use nothing in their demonstrations except
V* imaginary lines... and the belief of those who suppose that
something [really] issues from the eye is false.”25
Thus have the rays of Empedocles’ interior fire been
quenched. In their place Alhazen offers a carefully elaborated
theory of exterior, physical rays that can be combined with
Euclid’s precise language of mathematics to provide a compel
ling scientific account of vision. The eye, once the site of a
sunlike, divine fire, fast became a darkened chamber, awaiting
an external force to enlighten it.
Vision in a Dark Chamber
At this time, a device known as the camera obscura, which
literally means “darkened chamber,” affected the scientific
imagination so greatly that by the seventeenth century it had
become the model for the eye. While antecedents can be found
earlier, its first clear description appears in the writings of Al
hazen.26
On a bright day, stand within a darkened room. Punch a small
hole the size of this letter “o” in an opaque curtain over a window
in the room. Outside is a bright world, inside a dark room; they
are connected only by the light filtering through a single small
aperture. On the wall of the darkened chamber opposite the
hole, a wonderful inverted image of the outside scene appears
30
The Gift of Light
in full detail. In his study of the camera obscura, Alhazen ar
ranged several candles in a row on one side, their flickering
images then appearing in a similar but now inverted row on the
screen. If one held a theory of “husks,” they all must pass
through the same small aperture without interfering with one
another. Somehow the light from each candle passes simulta
neously through the same single point without obscuring the
image. Amazingly, an entire landscape, rich in color and detail,
can make its way undisturbed into a camera obscura through a
single tiny hole. In fact, make the hole too large and the image,
though brighter, becomes blurred.
The specific connection of this experiment with vision had to
wait four hundred years, until the Renaissance genius Leonardo
da Vinci made the extraordinary suggestion that the eye itself
| is a camera obscura. The eye, too, according to da Vinci, is a
dark chamber into which an image of the world is projected. , 5
In the first years of the seventeenth century, the mathema
tician and astronomer Johannes Kepler developed a complete
geometrical explanation for the camera obscura, and went on to
give a detailed and successful explanation of the optics of the
eye and vision in terms of it. As in the camera obscura, the
outer world was projected onto an inner screen in the eye. Kepler
declared that “vision occurs when the image of the whole hem
isphere of the world that is before the eye... is fixed on the
reddish white concave surface of the retina.”27 Yet Kepler, like
so many scientists before him, was deeply disturbed by one fact:
the image on the screen of a camera obscura js inverted! How
can the image on the retina be upside down when we see the
world right side up? Innumerable fantastic schemes had been
invented to rectify the image, but Kepler’s geometrical argu
ments were so tightly reasoned that, even in the absence of
direct observational evidence, there was simply no escaping the
conclusion: the retinal image must be inverted. Kepler accepted
this and left to others to explain how the image could be righted.
31
CATCHING THE LIGHT
Today we locate the solution to this problem in the mind, as
sociating it with psychology of vision. In Kepler’s words,
How the image or picture is composed by the visual spirits
that reside in the retina and the optic nerve, and whether
it is made to appear before the soul or the tribunal of the
visual faculty by a spirit within the hollows of the brain, or
whether by the visual faculty... all this I leave to be dis
puted by the physicists [philosophers]. For the armament of
opticians does not take them beyond this first opaque wall
encountered within the eye.28
At this point optics ends and the light of the body, that is
the soul’s activities, must be engaged in order for us to see the
world right side up.
Descartes
Experimental verification of Kepler’s deductions was finally
achieved by Rene Descartes. Descartes’s optical studies contain
a revealing illustration of the visual system that unwittingly
depicts not only Descartes’s understanding of the anatomy of
the eye and of visual optics, but also his philosophy of percep
tion.29
Above and at some distance from an enormous eye are ar
ranged three geometrical objects: a circle, a diamond, and a
triangle. Rays are drawn from them through the lens of the eye
below, and are focused on the retina. The rear membranes of
the eye have been removed in order that the philosopher (Des
cartes himself?) can view the three images projected onto the
rear surface. The external world depicted in the upper part of
the picture is shown in light; the lower part surrounding the
observer is dark. As with Alhazen, for Descartes the world is
bright, and the eye dark.
32
The Gift of Light
Optical analysis by Descartes. The philosopher viewed the world through
an ox’s eye whose back had been scraped to make it transparent. The
image he saw was inverted.
33
CATCHING THE LIGHT
The illuminating interior ray or fire had vanished. Yet Des
cartes still holds a two-stage theory of vision. In the first, light
(which he conceived of as material and mechanical) is conveyed
through the physical organ of sight to a common sensorium in
the body. The mechanical stimuli are then, in Descartes’s view,
“perceived” by a spiritual principle within man. For Descartes
the world of extension, of substance—res extensa—reached all
the way into the body but could not of itself complete the process
of vision. A spiritual principle, the mind or soul—res cogitans—
was still required. Like the philosopher in the illustration ob
serving the flickering retinal images from his dark vantage point,
the immaterial mind observed the mechanical proddings of the
world in the sensorium.
Although the light of the eye that reached out and granted
meaning to raw sensation had retreated from the body, it re
mained in Descartes’s dualist position as a disembodied spirit,
a vestige of the past. Yet even this faint echo of a Greek heritage
was destined for at least temporary extinction.
Modern Sense Physiology
We shall, sooner or later, arrive at a mechanical equivalent
of consciousness.
Thomas Huxley
The final development in the evolutionary drama of sight (and
I must leave much out) occurred in our century. By the mid-
19005, the neurophysiology and psychology of vision had ad
vanced to an extraordinary degree. The detailed knowledge we
now possess of brain structure and function, of the neural anat
omy of the eye and visual pathways, is truly staggering. In the
flush of excitement that naturally accompanies a century of dis
covery, many feel that they hold now in their hands “the me-
34
The Gift of Light
chanical equivalent of consciousness,” as Thomas Huxley called
it.
The Harvard biologist and Nobel laureate David Hubei speaks
for many scientists when he states that the brain is a machine
“that does tasks in a way that is consonant with the laws of
physics, an object that we can ultimately understand in the same
way we understand a printing press.”30 Moreover, contrary to
Descartes, we have no need to appeal “to mystical life forces—
or to the mind” to account for perception, thought, or emotion.
They are purely and simply states of the physical brain.
Hubei rightly recognizes the profound implications of this
view for everything we do. Our image of the mind sets the agenda
for everything from education to love relationships. According
to Hubei, once we understand that mind is an illusion, and that
brain is the sole reality, then we can restructure our systems of
education and social institutions to serve the brain, not an an
tiquated notion of “spiritual man.”
In traditional language, the substitution of a purely material
and sensual image for a spiritual reality is idolatry. In his in
sightful little book Saving the Appearances, Owen Barfield sug
gests a connection between the biblical injunction against
idolatry and the veneration of models so common to modern
scientific practice.31 Scientific models certainly have their right
ful place. But when does a model become an idol, that is, when
is it taken for something other than a model, becoming “reality”?
The model of an atom as a miniature planetary system is helpful
only as long as it is not taken literally. Quantum physicists
discovered long ago the dangers of idolatry. Neurophysiologists
have yet to leam the lesson. For many of them, the brain has
become an idol; it has become quintessential man.
The dangers associated with this kind of adulation of the brain
are innumerable. The image we have of ourselves is a powerful
thing; it shapes our actions, and so also the world we fashion
for us and for our children. It is important, therefore, patiently
and carefully to distinguish between idol and fact.
35
CATCHING THE LIGHT
I am not suggesting a simplistic Romantic return to the past.
There is no turning back. Yet are Hubei and the legions of
scientists who think like him right to reduce our humanity to
brain function? The answer is, quite simply, no. The brain as
described by Hubei is a carefully crafted and dazzling image
fashioned from the fruits of scientific research, one full of in
sights, but which is ultimately mistaken for something it is not.
Is it possible to embrace the results of science without falling
into such idolatry? Yes, but this, perhaps more than any other,
is the challenge we confront in our times. Our success or failure
in fashioning a nonidolatrous science will determine much for
our future.
Rekindling the Fire of the Eye
The movements we have traced are like a contrapuntal harmony
where one melody sounds against the other. As the light of the
eye dims, that of the world brightens. As the beacon of the eye
gradually retreats, the power of sunlight projects itself deeper
and deeper into the human being until finally the ethereal em
anations of Plato, and even the Cartesian spectator, vanish from
the Western scientific sense of self. Yet some data and scientific
developments indicate the possibility of a “postmodern” view
of self and vision that has room in it for the light of the eye. In
them, the interior ray may once a|igain find a place, even if under
another guise.
We have learned that our consciousness is not immutable.
Our habits of thought become perceptions, and while powerful
and pervasive, these are not universal or “true.” We should
learn to take responsibility for them. Do they accord with our
deepest intentions and the good of our society and planet? Or
do we need to “reimagine” ourselves and our world? In this way
Matthew’s remarks make real sense: “If thine eye be evil, thy
36
The Gift of Light
whole body shall be full of darkness. If therefore the light that
is in thee be darkness, how great is that darkness!” Our light,
a light of meaning, fashions a world, forms it from the light of
day. If our light be darkness—be evil—then we bring darkness
and evil into our whole body, personal and social. If it be light—
be good—then health flows into us, and into the world.
Plato’s light of the eye was a light of interpretation, of “in- =-
tentionality,” as modem phenomenologists would say; a light
that grants meaning. Cognition entails two actions: the world
presents itself, but we must “re-present” it. We bring ourselves,
with all our faculties and limitations, to the world’s presentation
in order to give form, figure, and meaning to that content. The
beautiful and productive images we craft on the basis of expe
rience are images only—fruits of the imagination. They are no
less true for being so. If in our enthusiasm we forget this, then
images become idols demanding propitiation at their high altars.
Nor should these reflections become grounds for abandoning the
path of knowledge, because it is a philosophy of growth and
development. The organs of insight we bring are neither fixed
nor limited, but malleable and expansive. Thus the importance
of integrating the insights into light gained by artistic and spir
itual disciplines as well as scientific ones.
(X
<\va °
37
CHAPTER 3
Light Divided:
Divine Light and
Optical Science
Our original question—what is the nature of light?—has been
answered differently by different peoples. To the Egyptians it
asked after man’s relationship to the god Ra. They sought first
a moral or spiritual answer, not a mechanistic one. By contrast,
we search to explain the nature of light by tracing light rays
through intricate optical systems. We seek light’s mathematical
and physical lawfulness. The sequence of words—what is
light?—does not have a unique meaning. The Egyptian answer
is utterly different from that of quantum optics, but are they
necessarily at odds? Or does the Egyptian yearn to know a
different part of light’s expansive being?
We should be open to the possibility that the essential ques
tions posed about nature in past or future ages may be quite
different from those posed by us today. As C. S. Lewis wrote
in his lovely book The Discarded Image, it is not that our present-
day understanding is unsubstantiated, but we should realize that
“nature gives most of her evidence in answer to the questions
we ask her.” The questions we ask, as well as the answers we
are willing to accept, reflect our temper of mind. The images
of one age may be discarded by another less because of new
discoveries than because of new priorities and new questions,
all of which reflect a changing psyche.
38
Light Divided: Divine Light and Optical Science
By remaining open to conflicting proposals regarding the na
ture of light, we allow the wide-ranging drama of the past its
place, and so read the full biography of light rather than only
the fragment we have written ourselves. We can also wonder
afresh at our present understandings of light and see the future
as being still undetermined. As we approach the ancients’ un
derstanding of light, we should leave our own hard-won, con
temporary images at the threshold to see as others have seen.
The Lost Eye
I am the one who openeth his eyes, and there is light; When
his eyes close, darkness falleth.
Ra speaking; from the
Turin papyrus, 1300 B.C.
Two eyes looked down on the civilization of the Nile, the “two
eyes of Horus,” the sun and moon. No more significant symbol
existed in ancient Egypt than the eye of the sun-god Ra. His
eye—the sun—was creative, his vision was life itself. It was
said that mankind arose from the tears of his eye. In Egyptian,
the very words for tears and men sounded similar.
The nature of light was clear to the Egyptians. As the priest
scribe wrote in the above fragment 3,300 years ago, when Ra
“openeth his eyes. . . there is light; When his eyes close, dark
ness falleth.” The gaze of Ra was the light of day. For men and
women of that civilization, to stand within daylight was to stand
in the sight of their sun-god. The power of vision to illuminate
the world was universalized, projected onto the grandest scale,
becoming the brightness of day. The gaze of God was light. Light
was God seeing.
We are reminded of the Greeks, who felt the force of their
vision, the “light” of their own eye, and developed a theory of
39
CATCHING THE LIGHT
The solar eye of Egypt, taken from a funerary
papyrus, 1000 B.C.
vision based, in part, on that experience. From the mythology
of the Egyptian world with its multitude of stories about the eyes
of Horus or Ra, we come to realize that prior to the individualized
light of Greek visual theory, sunlight itself was felt to be an
— emanation of an eye, that of the sun-god Ra. In neither case
was light a substance or thing, but rather was felt to be the
power of seeing. To see was to illumine. For Empedocles, the
human eye was like a lantern lit at the hearth of creation. When
open, it rayed forth into the world and man saw. For the Egyptian
priest, the sun itself was an eye, which, when open, brought
the day and, when closed, the night. The kinship between the
eye and the sun was felt deeply for many centuries, from ancient
Egypt to the medieval mystics. In Persian and Greek mythology
the identical image reappears—the sun and moon are the eyes
of gods placed into the heavens.2
The earliest answer given to our question—what is the nature
of light?—must be: it is the sight of God. Mankind, formed
0 from the tears of Ra or by means of his sight alone, shares
somewhat in his nature, like debased gods.3 By the time of the
Greek philosophers, we, like the gods, illuminate the world with
our sight. The sight of Ra lights up the cosmos; the sight of man
lights up our personal world.
40
Light Divided: Divine Light and Optical Science
The change from universal illumination and the eye of god
to the human is beautifully told in the Egyptian story of the “lost
eye” of Horus or Ra that appears in many variations within
Egyptian mythology.4 It seems that the Eye of the supreme god
of Egypt wandered away from its appointed course as the Sun,
becoming lost in the watery depths of this world and living as
a lioness in the eastern mountains of the sunrise. Ra sent Shu
OK?
and Tefnut in search of it, but by the time the Eye was found P™'11
and returned to the face of Ra, another eye had been fashioned
to take its place. The original Eye was outraged, but the god
Thoth pacified and healed it. Ra went still further and made a
place for it within the enclosed serpent form of the uraeus, and
placed it in the middle of his forehead “where it could rule the
whole world.” In the depiction of the pharaohs, this same em
blem rested on their heads. The Eye of Ra, the Sun, was no
longer free to rove unfettered, but was now forever encompassed
by the asp-serpent. Delimited and individualized by the power
of the serpent, it became the ruler of this world. The all-mighty
pharaoh, therefore, is depicted as crowned with the uraeus. The
eye of Ra becomes the eye of man; the light of god the light of
man.
Light was, and has remained, an aspect of God. Imaged in
countless ways as sight or angel or one of a thousand other
things, it has been inseparable from man’s groping to represent
the spirit. From Egypt we move to ancient Persia, where light,
darkness, and the divine unite to form a glorious religious uni
verse.
Light in a Dark World
Legend has it that at the age of thirty, Zoroaster, whose life had
been spent in careful attention to the path of righteousness,
stood in the river Daiti in order to draw water for ritual libations.
41
CATCHING THE LIGHT
i Drawing the water, he looked up and saw a man fair, bright,
| and radiant, whose silken gown was woven of angelic light. The
divine emissary led him to the great god of creation, the god of
light, Ahura Mazda, for whom Zoroaster had longed constantly.
So began the visions and ministry of the Persian prophet Zo
roaster in the second millennium B.C.5 The Egyptian imagina
tion had elucidated the kinship between light and eye in god
and man. The Persians, by contrast, explored the origin and
signification of light—and its negation, darkness—through a
majestic cosmic myth of creation.
In the dualistic religion of Zoroastrianism, the antagonism
between light and darkness was embodied in the form of warring
spiritual powers. Ahura Mazda, the uncreated creator of all that
is good on heaven and Earth, was opposed by an adversary,
also uncreated, but wholly malign, the “Hostile Spirit,” Angra
Mainyu, or Ahriman.6 The ultimate polarity of good and evil,
of light and darkness, is expressed in the struggle between these
two beings and their company of created spirits. Ahura Mazda
fashioned the world first in a purely spiritual, disembodied state
beyond the reach of Angra Mainyu. Following this came a second
I creation of the world, but now in a material existence, which
Angra Mainyu together with his Daevas (demons) immediately
attacked, mingling his own darkness with the static, luminous
perfection of the original creation. As a consequence, the seas
became salty, fire was tainted by smoke, deserts appeared, and
I all physical existence became a mixture of good and evil, light
I and darkness.
The earliest earthly source of light, fire, had always played
an important role in the practice of worship of Indo-Iranian
civilization. Under the influence of Zoroastrianism, it gradually
grew to assume ever greater significance in temple and home
rituals, so that by the sixth century B.C. there were great numbers
of fire temples. Their holy fires were ritually fed in a threefold
manner with dry fuel, incense, and animal fat by priests who
42
u;
Light Divided: Divine Light and Optical Science
never allowed them to fail. Perhaps worshipers saw in the fires
a symbol for the redemption of the earth. The wood, itself a
mixture of light and darkness, was transmuted by the alchemical
action of fire, as if by the fiery might of Ahura Mazda, to become
radiant light once again. The smoke and residual ash were the
irredeemable darkness woven into creation by Angra Mainyu
and now cast out by the fire of god.
The mission of human beings, as a creation of Ahura Mazda,
was also symbolized in the fire rituals. Ahura Mazda and his
Yazatas (beneficent spirits) shared the task of redeeming the
world by separating good from evil, the light from the darkness,
with humans.
According to Zoroaster, we live in a time of suffering, sorrow,
illness, and death, in a time when the worlds of light and dark
ness are mingled. Yet he insisted that we are not to flee the
suffering of the world but should assume the responsibilities
entrusted to us by Ahura Mazda. A third age will come when
the pristine first creation is restored, after the ultimate defeat
of Angra Mainyu and his legions by Ahura Mazda, the Yazatasi " ?
and man. Then good will be separated from evil, the light from 'iojid's
the darkness, and redemption accomplished. We now live in
the middle of the “Three Times” that frame Zoroastrian cos
mogony—Creation, Mixture, Separation. The role Zoroaster as
signs to man in world evolution is one of enormous dignity and
responsibility for the future, and one that places him squarely
in a cultic imagination whose root metaphors are light and dark
ness.
Zoroaster’s spiritual universe was peopled with innumerable
divine beings whose primeval actions had given rise to all ex
istence. Light and darkness surround us as images of those
ancient deeds. As creations of light, human moral actions now
will determine the future forms of existence for better or worse.
Our spiritual actions, like those of the gods, will one day image
themselves in light or darkness. What is inner will become outer.
43
CATCHING THE LIGHT
Angelic Light—Human Light
How you have fallen from heaven, bright son of the morn
ing, felled to the earth . . .
Isaiah 14:12-15
On the first day God created light; on the fourth He “made two
great lights, the greater to govern the day and the lesser to
govern the night; and with them he made the stars.” Placed in
the vault of heaven at the beginning of time, the sun, moon,
and stars continue to give light to the earth, govern day and
night, and separate the light from the darkness.7 So reads the
Old Testament.
The creation account in Genesis is revolutionary. We find in
0G it no tale of battle among the gods; no supernatural warfare of
the kind that appears in Babylonian, Greek, or Norse mythology.
The innumerable gods found in the myths of every culture are
reduced by the Jews to a single being: Jahve. God, in a solitary
action spanning seven days, created heaven and earth, man and
woman. While his original state was paradisaical, the acquisition
of divine wisdom was, once again, joined with labor and loss.
Eating the forbidden fruit of the tree of knowledge of good and
evil meant for Adam and Eve the loss of Paradise and the
beginning of want. To woman Jahve said, “in sorrow thou shalt
bring forth children,” and to man, “in the sweat of thy face shalt
thou eat bread.”8
In the Christian imagination, the Fall of Man was prompted
by the temptation of Lucifer; humanity’s ultimate redemption to
come through Christ. In one apocryphal tradition, these two
beings, Lucifer and Christ, were originally brothers, both high
\
angels of light. A medieval maxim declared Christus verus Lu
cifer, Christ is the true Lucifer, which makes all the more sense
as Lucifer’s very name means “bearer of light.” This being is
no Ahriman, the dark spirit of denial from Zoroastrianism, but
44
Light Divided: Divine Light and Optical Science
a being of radiant beauty who fell from the heights through the
sin of pride. In Isaiah’s words, he is “the bright son of the -
morning” who because he thought, “I will scale the heavens; I
will set my throne high above the stars of God,” was cast out
of heaven. The story is told in the Jewish Books of Adam and
Eve.9
God made man in his own image and blew the breath of life
into him. Then came the archangel Michael and he brought
Adam before the host of angels and bade them worship man in
the sight of God as the Lord desired. Lucifer balked, responding:
“I will not worship an inferior and younger being than I. I am
his senior in Creation; before he was made was I already made.
It is his duty to worship me.”
For defying the warnings of Michael and the dictates of God,
Lucifer was cast down onto the earth. There he saw Adam and
Eve, who were still in Paradise. Miserable at his separation from
God and envious of the happiness of the first pair, he beguiled
Eve and through her Adam, so that they ate of the forbidden
tree.
The Fall of Man was inseparable from the fall of the bearer
of light, Lucifer. As angelic light was cast down onto the earth,
man and woman ate the fruit of the tree of knowledge, and so
were cast out to labor in the lands east of Eden.
The expulsion and fall of Lucifer, and the accompanying Fall
of Man, finds its counterpoise in the incarnation of Christ—a
luminous being of the sun. According to the gospel of John,
Christ is “the light of the world,” a light that shines into the
darkness. John says it more bluntly still in his first letter. “This
is the message we have heard from him and proclaim to you,
that God is light and in him is no darkness at all... if we walk l"\A
‘ !
in the light, as he is the light, we have fellowship with one
another.”10
John the Baptist “came as a witness to testify to the light. . . .
He was not himself the light; he came to bear witness to the
45
CATCHING THE LIGHT
light. The real light which enlightens every man was even then
coming into the world.”11 Christ was to be understood as the
“real light”—the true spiritual, consummately angelic light of
the world. Without the burden of sin, that is, without expulsion
from heaven, this high angel of light was nevertheless incarnated
in the flesh of this earth. Resisting Lucifer’s temptations to pride,
he prepared for the shame and degradation of the Passion. Those
who scorned him asked how could this be the Messiah, if the
Lord God allowed him to be scourged, crowned with thorns, and
crucified between two criminals? Yet in the Christian imagi
nation, without these trials of sacrifice and love, performed in
utmost humility, the light of resurrection could not shine. And
might not the salvation of fallen man also entail the redemption
of Lucifer—of fallen Light?
PROMETHEUS, TOO, was finally unbound from his agonies on
the Caucasus by the hero Heracles. Having accomplished
eleven of his twelve labors, which included an initiatory de
scent into hell and a rising again, Heracles journeyed to Mount
Atlas in search of the golden apples of Hesperides. Near the
end of that final task, he crossed the Caucasus and found and
freed Prometheus from the agonies of his daily torture. Thus,
the greatest hero of Greek myth, having completed his own
\ soul’s transformation through twelve mighty labors, could free
the first benefactor of mankind. Himself free, Heracles could
Jree others,,
The shared, cultural imagination of man as sojourner in a
dark, fallen world into which light has been mingled offered
each soul the Heraclean task of world redemption, whether as
a follower of Zoroaster or Christ. This imagination attained its
most radical expression in the teaching of the third-century
Iranian gnostic Mani.
46
Light Divided: Divine Light and Optical Science
Manichaeism—Religion of Light
Rich Friend of the beings of Light! In mercy grant me
strength and succour me with every gift.'2
Manichaean Hymn
In the main hall of a rustic, hillside shrine, thirty miles south
of Ch’uan-chou on the eastern coast of China, is a stone statue
of a Buddha-like figure.13 Its various members have been skill
fully carved from stones of differing hues so that the entire statue
appears luminous. Many of its features are familiar to us from
Buddhist statuary elsewhere; he sits in peace, cross-legged on
a lotus, dressed in a kasaya, and enclosed in the sacred radiance
of an embracing halo. Looking closer, however, we notice that
this is no ordinary Buddha. In place of downcast eyes, curly
hair, and the clean chin characteristic of a Buddha, we see long
straight hair draping his shoulders, framing a bearded face. His
penetrating eyes stare straight at the approaching pilgrim. Look
ing to the nearby inscription, we read: “Mani the Buddha of
Light/’ ~ ’
The origins of Manichaeism were far away from that unique
Chinese shrine in time as well as space. When Mani was but
twenty-four years old, “there flew down and appeared before me
that most beautiful and greatest mirror-image of myself. . . my
divine Twin.”14 From that day in mid-April A.D. 240 in the
ancient city of Ctesiphon (near present-day Baghdad), there
arose a religious movement that spanned the Roman empire,
Persia, and the Far East. Mani, “the Envoy of Light,” taught
to his followers what his Twin had revealed to him, traveling
widely, healing and teaching as he went. His was a religion of
light whose central creation myth and ascetic practices acted
so powerfully on the human imagination that Manichaeism
I spread without benefit of warfare, forced conversions, or active
royal support. Perceived as a threat by every orthodoxy, Mani
47
CATCHING THE LIGHT
died a martyr’s death, but not before he released what was the
most widely held heretical vision ever to capture the imagination
of man. Its last beautiful refrains were still being sung in the
thirteenth century by the devoted followers of various Christian
“Manichaean” heresies in southern Europe, until they, too, were
ruthlessly destroyed.
It is not surprising that we find Manichaeism first arising in
the countryside that fostered Zoroastrianism, for in many things
they were similar. According to Mani, Light and Darkness were
equal powers originally strictly separated from one another, but
through repeated invasions they became mingled. When given
the chance during the first conflict of the creation drama, the
<- Prince of Darkness, Ahrmen, and his hosts seized upon and
conquered the First Man, Ohrmizd, of the Realm of Light. By
this means Light and Darkness were mixed and the world brought
into existence. So, too, was the stage set for all future attempts
at redemption. In order to liberate the First Man, the Father of
Greatness evoked the Friend of Lights, who in turn issued the
Living Spirit. Together with the Elements of Light, the Living
Spirit entered the Realm of Darkness and liberated Ohrmizd,
but in his flight Ohrmizd had to leave behind his luminous
, amour, the Five Elements. The purest particles of light from
<- the Elements came to form the sun and moon, those slightly
L corrupted formed the stars, and a third of the Light was left
behind, profoundly corrupted, to form the earth. In an attempt
to thwart the redemptive efforts of the Beings of Light, the Prince
^of Darkness (not Jahve) created Adam and Eve in the image of
| the third divine messenger, the not-yet-incamate Christ. (All
this was taken as a fantastic heresy by the young orthodox
Church.) Through Adam and Eve and their descendants, Light
was to be trapped forever in the dark realm of Ahrmen. Mani’s
' mission was its release. Light, imprisoned in the matrix of matter
since the battles of Ahrmen, should make its way home again.
Within Manichaeism, the Elect (i.e., priests) had an espe-
48
Light Divided: Divine Light and Optical Science
cially important role to play in the project of redemption. For
through their earthly conduct they could release Light and carry
it within them through the gate of death. Even the food they ate
contained enchanted Light. When rightly prepared by Hearers
(those who heard the teaching but were not themselves Elect)
and eaten by the Elect, the Light within it would be released
into the priests’ care. At their death the Elect journeyed in spirit
to the Moon with their harvest of Light. The Moon waxed with
the bounty brought by them until, when full, a Column of Light \
would rise from the Moon and streamed on for a final return to \
the Sun, its proper celestial home.
During his life, Mani had traveled widely, preaching his mes
sage of Light and Darkness, of knowledge (gnosis) and salvation.
He envisioned a universal religion that brought together the
heritages of Buddha, Jesus, and Zoroaster. Manichaeism’s in- (_
fluence was enormous. One of the greatest figures of early Chris-
tianity, St. Augustine himself, was a Manichaean from 373 to
382, and although later he vigorously attacked its doctrines, his
thinking was never really free of its effects, and indirectly
through his writings a metaphysics of light, bom of Mani’s vision,
made its way into orthodox Christian thinking.
As Christ, Light incarnate, was delivered by jealous Jewish
priests to Pilate for execution, so was Mani, the Envoy or Buddha
of Light, given to the Persian king Vahram I for execution at
the instigation of the zealous Zoroastrian priest Kirdir, who
tolerated no rivals. Mani, however, was confident that his teach
ings would live on long after his death (in A.D. 277), and so
they did.
Throughout the history of the early Church, Mani’s cosmogony
of light, and tenets of many kindred heresies, constantly chal
lenged orthodox doctrines. Their elaborate imagination of our
universe as caught in the struggle between spiritual light and
ponderous darkness informed the lives of countless followers.
In addition, the contemporary Dutch theologian G. Van Gro-
49
CATCHING THE LIGHT
ningen sees a less personal but equally effective influence of
gnostic thinking. In the gnosticism of Mani and similar sects,
Van Groningen detects an earnest concern with the scientific
and philosophical issues that take their point of departure from
man.15 These sects, not the early Church, were deeply com
mitted to knowledge. They aspired to a knowledge or gnosis of
the spiritual cosmos not in order to master the earth, but to free
I the soul that it might return to its native realm of light. Still,
gnostic aspirations, Van Groningen suggests, foreshadow sci
entific ones. The early Church saw in such striving the luciferic
sin of pride. Salvation, after all, was given by grace alone, not
-» by knowledge. Through the Middle Ages the western Church
carefully restrained those in its fold whose heterodox studies
(and there certainly were such) might take them too far. Others
outside the fold were treated more harshly. Yet, as he had
predicted, Mani’s impulse lived on. The last dramatic scene in
that history was the flowering and destruction of the Cathars, or
“Pure Ones,” of thirteenth-century southern France.
Between the villages, chateaus, cities, castles, and through
out the countryside of Languedoc, lean, pious men journeyed,
two by two, long-haired and dressed in black. Gaunt from hard
fasting, they spoke quietly, without anger or worldliness, of
Christ and a path of light, the path of salvation. Many were their
acts of charity, and constant was their attention to God and a
pure life. To the young and simple they told fairy tales of ex-
I traordinary beauty;16 to others they gave religious instruction
according to their place within the community. In devout sim
plicity, they celebrated together the rituals and sacraments of
their church. To the people they were known as bons hommes,
or “good men”; within their religion they were perfecti. Passing
an elderly perfectus who was ill-clad and crippled, the Count of
Toulouse said: “I would rather be this man than a king or
emperor.”17
While, in fact, distinct from Manichaeism, the Cathars (and
related heresies of the period) shared the Manichaean view that
50
Light Divided: Divine Light and Optical Science
the world was the creation of Satan, and that through their
righteous conduct, the divine Light within them could be saved.
Failing that, they would return to life in the form of another
sentient being (metempsychosis) to try again.
The established Church of Rome first ignored and then fought,
with more and greater force, the incursions of the Cathars into
its dominion. The Church sent representatives from its own
newly established mendicant orders, most notably St. Bernard j
in 1145 and St. Dominic in 1205. Trudging barefoot under the
hot sun and dust, St. Dominic begged for alms and an audience.
Tormented wherever he went, ridiculed and mocked, he
preached the doctrines of the Catholic Church not only in his
words, but also in his actions of devotion and poverty. While
these certainly made an impression on his hearers, neither he
nor St. Bernard had much success in converting Cathars. When
they failed, repression was accomplished through violence.
While St. Francis, in Assisi, was writing his beautiful Canticle
to the Sun, a glorious paean to all existence, the Cathars were
being ruthlessly destroyed by sword and fire in southern
France.18
The final blow came in 1238 with the fall of the Cathar High
Place, the mountaintop fortress at Montsggur. Two hundred per-
fecti sheltered there were burned without trial. While it took
another century of systematic repression, execution, and bru
tality to extinguish the ideal of “purity” espoused by the Cathars,
by 1330 the Cathar church in France was no more. St. Bernard
of Clairvaux, who fought the Cathars with all his strength, said
of them, “No sermons are more thoroughly Christian than theirs,
and their morals are pure.”19 Their austere heresy took with
consummate seriousness the reality of angelic light in a darkened
world. They were the last in the West to live fully within an
imagination of spiritual light. With the eradication of the Man
ichaean heresy, a mighty imaginative stream ended, effectively
disappearing from the stage of external history.
Yet the metaphysics of light espoused by Manichaeism found
51
CATCHING THE LIGHT
a more orthodox echo in the religious and scientific convictions
of the thirteenth century’s most significant figure in the biography
of light.
Robert Grosseteste’s
Cosmogony of Light
In 1229, just three years after the death of their beloved St.
Francis of Assisi, the English Franciscans invited the elderly
churchman and scholar the Archdeacon of Leicester, Robert
Grosseteste, to teach the young friars at their newly built Oxford
school.20 Six years later, in 1235, he was consecrated Bishop
of Lincoln, the largest diocese in England. As teacher and then
as bishop, Grosseteste was famous for his upright and moral life
in an age when clergy were often more interested in temporal
delights than the final rewards of the hereafter. Grosseteste had
a special regard for the poor Franciscans who shunned the riches
of the Church for those of the soul, and they returned his af
fections in kind. Combined with his upright character was a
love of music, a sharp wit, a devotion to learning, and a single-
minded preoccupation with the nature of light.
Among Grosseteste’s books, one is a gemstone that outshines
all the rest. It is the slender volume entitled De Luce, or “On
Light.” In order to appreciate it we need to remind ourselves
of Plato’s influence on medieval scientific thought.
Near the end of his life, Plato wrote a lengthy dialogue Ti
maeus, which begins by recounting a conversation between an
aged Egyptian priest and the Greek poet Solon in the holy city
of Sais. The priest, having access to documents of great antiq
uity, described to Solon the most ancient history of Greece,
going back long before the deluge. Timaeus, the astronomer in
Socrates’ group, continued the tale by going back even further,
giving an account of the generation of the universe. The “likely
52
Light Divided: Divine Light and Optical Science
story” told by Timaeus is of the creation of the world by God
through his subordinate Demiurge. Yet its character is less a
religious than a scientific presentation, relying on reason, evi
dence, and speculation, and not on divine inspiration. The world
is formed according to number, ratio, and geometry, wrote Plato.
The five elements of earth, water, air, fire, and ether, or quin
tessence, are composed of triangles that join to form beautiful
solids: the five regular “Platonic solids.” His cosmology was
thoroughly mathematical, and in Greek times that meant pri
marily geometrical. Throughout the Middle Ages the Timaeus,
perhaps more than any other document, shaped the scientific
imagination of the West, and together with Genesis challenged
medieval thinkers to fashion their own “likely story” of the
world’s beginnings. The Timaeus was one of the few texts of
Greek thought that had made its way, at least in partial trans
lation, into the early intellectual life of Western Europe.
Grosseteste’s De Luce is the only comparable work of scientific
cosmogony between the time of Plato’s Timaeus and the eigh
teenth century. It is a watershed document. On the one side,
receding into a haze of unclarity, rise the magnificent heights
of prescientific Greek and Christian thought; on the other are
the as-yet-unborn achievements of modern science and tech
nology. Some years ago the eminent French historian Alexander
Koyre advanced the striking opinion that De Luce was the first
step toward the establishment of a modem mathematical science
of nature.21 On opening the book, we therefore would expect to
see pages dense with mathematics, calculations, and geomet
rical figures. Instead, we find a story of creation, a Neoplatonic
genesis in which the entire universe comes into being through
the expansion and modification of light.
Light, according to Grosseteste, was the first form of cor- o
poreity, and from it all else followed. Multiplying itself from a
single point infinitely and equally on all sides, light formed a
sphere and together with this action arose matter. Following the
53
CATCHING THE LIGHT
expansion of light to its ultimate extent came a phase of differ
entiation caused by condensation and rarefaction, which led in
turn to the separation of heavens and earth and to the origin of
| the thirteen spheres, nine heavenly and four mundane. We will
not follow Grosseteste’s arguments in detail, but I would em I
phasize the twin themes of light and measure (or geometry) that
pervade the work.22 Like Plato, he embedded within his cos I
mogony his vision of God as Geometer and Numerator, who
“ordered all things by number, weight, and measure.”23 The
medium chosen by God for his creation was light. Yes, De Luce
offers us the tender beginnings of mathematical science, but
everything is implicit, contained in its structure, style, and
Grosseteste’s manner of inquiry.
All of material creation, then, is condensed light. The modem
architect Louis Kahn was unwittingly paraphrasing Grosseteste
when he said to an interviewer for Time magazine, “We are
actually bom out of light, you might say. I believe light is the
maker of all material. Material is spent light.”24 Yet according
to Grosseteste, the condensation of light is only one pole of
God’s creative action, the pole leading to sensual creation.
Grosseteste also believed deeply in the supersensible pole of
creation, in angelic light. According to his view, we stand upon
the earth surrounded by the four visible kingdoms of nature
=. (mineral, plant, animal, and man), but abovei man rise the nine
hierarchies of angelic beings: Angels, Archangels, Archaei, Ex-
L usiai, Kyriotetes, Dynamis, Thrones, Cherubim, and Seraphim.
Together with the recognizably scientific aspects of his cos
mogony of light, Grosseteste embraced an explicitly spiritual
metaphysics of light. To Grosseteste, God’s declaration “Let
there be light” possessed two aspects. One aspect would ulti
mately become the light of our physical existence condensing
even to the form of matter, but the other aspect was a light of
intelligence embodied in the purely spiritual, angelic creations
of God. In his commentary on the opening chapters of Genesis,
54
Light Divided: Divine Light and Optical Science
and especially in his detailed studies of the four major writings
of Pseudo-Dionysus, Grosseteste revealed his profound interest
in the spiritual ordering of our cosmos, and once again he saw
it as best symbolized as a cosmos of light, but now angelic light.
In Grosseteste, we also find the first advocate for mathematics
as a pivotal element of the new science. He wrote, “nothing
magnificent in [the sciences] can be known without mathemat-
ics.”25 Grosseteste has been called the first expositor of exper
imental science. He provided, at least in theory, the twin pillars
essential to the development of modem science: mathematics
and experimentation. A. C. Crombie declared him “the first to
.set out a systematic and coherent theory of experimental inves
tigation and rational explanation by which the Greek geometrical
method was turned into modem experimental science.”26 Others
followed who would actually enact the program Grosseteste de
scribed, but he articulated with forceful clarity the tenets that
would become basic to the new science over the following cen
turies. Both the metaphysics and physics of light played through
his mind with equal ease and honesty. It was, after all, God’s
created universe, material and spiritual, and the place of man
as a citizen within both called for knowledge of both realms
whether through revelation, reason, or experimentation.
Grosseteste’s thinking had been influenced not only by Pla
tonism and Christian theology, but also by the first great phi
losopher of the Islamic world, al-Kindi. During the ninth
century, al-Kindi had not only vigorously argued for an improved
version of Euclid’s extramission theory of sight, but also sought
to integrate it with his spiritual philosophy of nature. The eye
is not alone in sending forth rays, but “everything in this world
produces rays in its own manner like a star. . . . Everything that
has actual existence in the world of the elements emits rays in
every direction, which fill the whole world.”27 The universe of
al-Kindi was woven together by a vast web of lightlike rays
linking stars to earth, magnets, fire, sound, and so on. Optics
55
CATCHING THE LIGHT
thus became an archetypal image of a universal process: the
radiation of power. Echoing the cosmology of al-Kindi, Grosse
teste saw the sun and moon in the eyes of the human face; “for
the head is borne towards the heavens and has two lights, as it
were sun and moon.”
Robert Grosseteste is for us a figure who stands, Janus
headed, facing in two directions. One face gazes upward and
back, evoking a metaphysics of light that embraces hosts of
angelic beings and emanations of light that are active in God’s
creation as intermediaries and enactors of His will. His other
face gazes earthward to a future when the physics of light will
develop to its fullest extent as modern optics, and most espe
cially as a mathematical physics grounded in experimental ob
servation. Both faces, both of Grosseteste’s enthusiasms, are
rightfully a part of our study of light.
Light WAS ONCE the sight of God. As the gaze of Ra spanned
space light stretched from one comer of the universe to the
other. In the night sky, planets and stars once played host to
gods and angels, who in turn passed on their gift of light to man
through mishap or cunning. In Christianity, the greatest cosmic
being of light, Christ, ancient brother to Lucifer, came down to
earth. It was as if the Persian sun god Ahura Mazda were to
don vestments of clay. The cultic imagination of light reached
its highest form in the gnostic-Manichaean sects in which the
redemption of the world was the release of light through human
actions. The final grand inflorescence of this religious imagi
nation perished in the fires that consumed the Cathar perfecti.
Henceforth light would grow to be understood scientifically.
Arising from Greek philosophy, moving through Persian and
Arab lands to the desk of Grosseteste and his contemporaries,
man’s imagination of light profoundly changed, becoming ulti
mately the conceptual basis for the optical sciences of modern
56
Light Divided: Divine Light and Optical Science
physics. Grosseteste still felt the presence of a glorious spiritual
tradition that could nourish his inner life. He sought to mate
the first buds of natural science with his religious thought in a
metaphysics of light. With the passing of the Cathars and of
Grosseteste, the religious tradition of angelic light faded. Over
time, science pruned away the trappings of spirit to fashion a
material and mathematical imagination of light. In doing so, it
similarly reshaped its image of man and cosmos.
57
CHAPTER 4
The Anatomy of Light
Hail holy Light, offspring of Heaven first
born!
Or of th’ Eternal coetemal beam
May I express thee unblamed? Since God is
light,
And never but in unapproached light
Dwelt from eternity, dwelt then in thee,
Bright effluence of bright essence increate!
Milton, Paradise Lost
On the Threshold of Scientific Sight
Outside the sublime cathedral of Santa Maria del Fiore in Flor
ence, the day was lovely and a light wind pushed billowing white
clouds across an azure sky. It was exactly the kind of day Filippo
Brunelleschi had been waiting for, so he let it be known among
his friends that a long-promised demonstration would take place
in the doorway of the cathedral.1
Standing on the threshold, Brunelleschi looked across the
cobbled piazza to the beautiful Baptistery of San Giovanni, al-
58
The Anatomy of Light
ready three hundred years old. Within its dim interior, glorious
mosaics hovered overhead on its domed vault, including the
huge figure of Christ at the Last Judgment. Dante Alighieri had
been one of the innumerable Florentines brought into the com
munity of the faithful through baptism there. Just now, however,
standing on the threshold of the cathedral, Brunelleschi did not
see the sacred history of the baptistery, but rather a network of
lines receding to an infinite horizon. He looked on the piazza
with the eyes of the great inquirers who would come later: Ga
lileo, Descartes, and Newton. For today he was not only architect
and sculptor, but also geometer and optical scientist.
His friends gathered around him, the slight youthful artisan,
and listened intently to his animated explanation of the dem
onstration he would perform before them. A genius, he would
go on to design a great dome for the cathedral behind him—an
enormous task that would last sixteen years, employing scores
of craftsmen and artists and consuming vast sums of money.
His task today, however, was far more modest in budget and
scale. It required only a wooden panel twelve inches square, a
mirror, the brushes of a miniaturist, paints, and a single man.
Yet in that afternoon, Brunelleschi performed magic, a trick
whose consequences would reach further by far than those of
his dome, or indeed anything else he would ever create. Quite
simply, he was changing the way we see.
On the threshold of the cathedral, Brunelleschi made the first
( drawing in linear perspective. Scholars place the date between
I 1412 and 1425.2 For the first time in history, man was creating
a scene so realistic that when it was viewed from the right place,
“the spectator felt he saw the actual scene” before him, as
Brunelleschi’s contemporary Manetti tells us.3
In order to prove the power of perspective illusion, Brunelles
chi contrived an ingenious demonstration. Those present that day
viewed his perspective rendering of the Florentine baptistery in a
bizarre but highly dramatic way. (See the figure on page 60.)
59
CATCHING THE LIGHT
Brunelleschi’s illusion of perfect perspective. A mirror in his right hand
and the painting in his left, Brunelleschi looked through a peephole toward
the baptistery. With or without the mirror, the scene was the same.
They stood in the door of the cathedral, facing the baptistery.
We should imagine Brunelleschi’s friends holding the painting,
one by one, with its painted surface away from them, and peeping
through a hole drilled through its center. Through the hole they
saw the piazza and baptistery. Into the line of view they now
inserted a mirror whose size and shape were such that, when
held at arm’s length, the reflected image of the painting just
[ filled the mirror. But the painted scene viewed in reflection was
exactly that seen directly. In other words, the viewer saw the
same thing with and without the mirror: a perfect illusion. To
P enhance the illusion further, Brunelleschi made the painting’s
? sky out of burnished silver, so highly polished that its mirror
* surface reflected the windswept clouds as they drifted overhead.
The illusion was masterful and complete.
The effect was stunning, and its consequences were destined
to dominate all of painting until the twentieth century. How had
he done it? Why hadn’t it been done before? Euclid’s theory of
60
The Anatomy of Light
vision was completely adequate to the task and had been known
for centuries. Yet something had been lacking, not some new
fact but a new spirit, a new scientific way of seeing.
A Spiritual Geometry
For the mind, only that can be visible which has some
definite form; but every form of existence has its source in
some peculiar way of seeing, some intellectual formulation
and intuition of meaning.4
Ernst Cassirer
In a small room of the Frick mansion in New York hangs a
painted panel not much larger than the one by Brunelleschi in
the doorway of the Duomo. It is a work by Duccio painted in
1310 depicting the temptation of Christ by Satan. Two enormous
figures, Christ rejecting his dark adversary, stand on a stylized
mountain amid a landscape barren but for a few surrounding
clusters of buildings. From the standpoint of linear perspective
or the way a camera would have caught the scene, nothing is
correct in the painting; but from the standpoint of spiritual
relationships, the arrangement is perfect. The sky, made of gold »
leaf, suggests that we are not looking into a physical but rather
a moral or spiritual space. Everything—the colors, the gestures,
the relative scale and positions of each figure, building, and
mountain—reveals a sacred order that may well be at variance
with the laws of vision, but is in true harmony with the values
of fourteenth-century Italy. In just this way did the painter, and
the painting’s viewers in Siena, see their world.
What mattered most for them was not the physical but rather
the spiritual geometry of space. This had been true of Egyptian
art and it remained true of Duccio’s. Pharaoh and the Madonna
had to be larger than those who served them because their
spiritual stature required it. Painting depicted reality, and in
61
CATCHING THE LIGHT
- i’|i
F>
ii ■'
r
f ■i
T
rTI
The Temptation of Christ on the Mountain by Duccio.
doing so it created a language of expression in profound accord
with human experience, not the laws of optics. Only when “what
was” changed, when Brunelleschi and his associates began to
experience their world from the standpoint of Euclidian analysis,
did the means arise to represent their vision: linear perspective.
It is difficult, almost inconceivable, for us to imagine a culture
that does not see pictorial images as we do. Perhaps an example
from cross-cultural studies will help. Twenty years ago, two
acquaintances of mine worked in remote villages in Africa with
62
The Anatomy of Light
indigenous peoples. Their project was to improve infant care in
the hopes of lowering the staggering infant mortality rate. One
friend was a professional photographer and so developed a slide
show on infant care. Once it was perfected to his satisfaction,
he packed his high-tech presentation into his truck and headed
out to the bush to join the woman who is now his wife. After
some efforts they devised a way to show the slides to women of
the village. At the end he asked them how they liked the pre
sentation. Very nice, the women said. The colors were beautiful,
and the shapes were very interesting. After a few conversations,
it dawned on him that during the evening’s performance much
or all of the slide imagery was simply not seen as he had seen
it. The colors were great, but they didn’t see the images as
representative of infants, objects, or people. The context and
scale of the images were so remote from their experience as to
make them nearly meaningless. This single instance could be
multiplied endlessly with startling and confusing anecdotes from
early missionaries, anthropologists, and explorers.
In a more systematic cross-cultural study of pictorial percep
tion, Jan Deregowski studied the perceptual abilities of Zambian
children and adults.5 One aspect he studied was the assignment
of spatial arrangements and depth to pictorial images. From his
work and others, it is clear that many visual clues are ignored
or misinterpreted (from our viewpoint) in other cultures. Con
sider the phenomenon of “split type” drawing. When African
children and adults were presented with a perspective drawing
of an elephant from above, and another showing the elephant
unnaturally split, also seen from above, they almost universally
preferred the split-type drawing.
The “properly” drawn perspective view of the elephant was,
according to them, “jumping about dangerously.” The prefer
ence for split-type drawing is remarkably widespread, according
to anthropological studies. When representing an animal or face,
the African artist, like the child or pre-Renaissance artist, could
63
CATCHING THE LIGHT
The “split” elephant was preferred by African children
and adults to the top-view, perspective drawing.
not leave out the “hidden” side. Our sense that linear perspec
tive renders a scene truly is purely cultural. It can only appear
when we, as a culture, have sufficiently deadened the active
light of interpretation. Then the eye or legs that are hidden but
that we know are there can be left out; the relative size of pharaoh
can be the same as or even smaller than that of his servant,
even though we know he is the greater being. One can almost
feel the withdrawal of interpretive light with the birth of linear
perspective. Yet even here, our mind is still active, otherwise
we would struggle, as Moreau’s child did, to see the world.
From these and our previous considerations of sight, we re-
64
The Anatomy of Light
alize that every age and culture has crafted its own sensory
reality, forged its own view. We may apply the same lesson to
the discovery of linear perspective.6
The pre-Renaissance history of perspective is complex and
controversial. As the eminent art historian Erwin Panofsky has
argued, “each period in western civilization had its own ‘per
spective,’ a particular symbolic form (as Ernst Cassirer called
them) reflecting a particular Weltanschauung or world view.”7
Elements of perspective can be found dating back to the Greeks,
but not until 1425 did a coherent, clear understanding of linear
perspective appear. Brunelleschi and the Renaissance longed
to see and represent the world in accord with “visual truth,” as
Lorenzo Ghiberti called it.8 They “endeavored to imitate na
ture,” to school their sight on the mathematical lawfulness of k< CCLUJSC
nature. They incorporated into themselves the geometric spirit
that breathed through the age and thereby helped lay the foun
dation for the style of scientific investigation to follow.
By 1525 the spirit of artistic-scientific perception had so
developed that Albrecht Diirer could publish his unique Paint
er’s Manual. Opening it, one does not find a single section on
color or painting as such. Rather, it begins with instructions
concerning the geometrical construction of spirals and conic
sections, then goes on to polygons, the Pythagorean theorem, -
towers, sundials, and regular and irregular solids. There is even
a detailed section on the “Construction of the Alphabet” from
arcs, circles, and lines. Nowhere is the spirit of the manual,
and so of the period, more evident than in the final section on
linear perspective.
The principles of linear perspective derive from Euclid’s con
cepts of the visual ray and the visual cone as elaborated by
Ptolemy and al-Kindi. Rays from the eye extend to every point
of the seen object, forming thereby an irregular cone with its
apex at the eye. Imagine a plane surface (the painter’s canvas
for example) somewhere between the object and the eye. Each
65
CATCHING THE LIGHT
Goddess and craftsman. A perspective drawing machine by Albrecht
Diirer.
ray will pass through the canvas at a particular point. Suitable
markings of each point on the canvas will create an accurate
representation of the object in perfect perspective.
In his Painter’s Manual, Diirer includes descriptions and
drawings of various devices that permit the draftsman to apply
Euclid’s theory and so to make completely accurate perspective
drawings. The most revealing device is his last, wonderfully
illustrated by Diirer himself. On the left of the drawing is a
woman, unclothed but for a carefully placed drapery, reclining
as a goddess might on a sea of pillows. Perhaps it is a study of
Aphrodite, the beautiful and wanton goddess of love who, curi
ously, was wife to fiery Hephaestus, the lame god of craft. Yet
something is amiss. She lies not on a couch but on an optical
bench outfitted for “projection and section,” as it would be called
today. On the right is the draftsman whose eye is positioned
steadily behind a pointer. From it his line of sight (Euclid’s
cone of visual rays) is directed to the object (that is, the “god
dess”) through a frame crisscrossed with stout black thread. The
grid so formed acts as a coordinate system superimposed on the
sensuous figure of the reclining nude. Point by point, the drafts
man, with utmost visual acuity, transfers her divine figure to
the cross-ruled paper on the table before him. Recall that young
Tiresias had been blinded by a sight such as this. To the right,
66
The Anatomy of Light
however, we have no youth but a descendant of Aphrodite’s
husband, Hephaestus the divine craftsman.
The juxtaposition is remarkable. To the left is a figure whose
pose evokes erotic emotions or sentiments of sublime sensuality
that had long been part of the artistic tradition. To the right is
the dispassionate, objective eye of the draftsman recording in
precise detail every feature of the body before him. It is the
meeting of two streams of time, one flowing from out of the past,
a stream of beauty both sensual and spiritual, and a second
stream flowing from out of the future, one that aspires to clarity
and control. They are wed in the picture plane of the artist’s
canvas as Aphrodite was wed to Hephaestus.
Nor is Albrecht Diirer alone in his enthusiasms. The most
powerful image I know that reflects the shift in consciousness
then under way is Leonardo da Vinci’s anatomical study of a
man and woman in the act of love, of procreation, dating from
the first years of the sixteenth century.
Human dissection had long been under religious and popular
prohibitions, only gaining legality in the century before Leo
nardo. The story of Michelangelo’s secretive anatomical studies
is well known. Under cover of night he would steal into the
morgue in order to dissect the unclaimed bodies of beggars by
the light of a flickering candle. Until Leonardo’s time, little new,
attentive work had been done in human anatomy. The medical
texts of antiquity were given prominence over original anatomical
study and illustration. Leonardo was certainly the most brilliant
anatomist-artist of the period.
In his anatomical study of human sexual intercourse, Leo
nardo takes the consummate act of intimate human affection
and displays it in anatomical cross section; the bodies of the
two lovers are cleaved from head to groin. The inquisitive eye
of the artist-anatomist has fixed on this action as it might on
any other and submitted it to the clear, calm gaze of the new
spirit. The drawings are still suffused with an artistic tenderness
not to be found in modem anatomical texts. We should not
67
CATCHING THE LIGHT
' ■ >
I '
h* ‘v*
...... , f 4*
B
k
,ua
K £if
i ani^Ss
-
Bib
^5$
■ W1
-^1
1
<
The anatomy of love, from the anatomical drawings of Leonardo
da Vinci.
68
The Anatomy of Light
imagine Leonardo, Diirer, and Brunelleschi as revolutionaries
bent on the overthrow of a godly universe. They viewed their ef
forts as a service to God who Himself had created the world ac
cording to number and measure. Leonardo longed to see deeper
than the surface appearance of things, and hoped “that it might
please our Creator that I were able to reveal the nature of man
and his customs even as I describe his figure.”9 The Gothic ca
thedral had also been the expression of a conviction that God in
his creation worked geometrically.10 Renaissance painters saw
perspective painting as a similar imitation of His work.
The PHILOSOPHICAL origins of the new attitude to nature can
be found in the remarkable thought of cardinal and mathema
tician Nicolaus of Cusa in the fifteenth century." For hundreds
of years nature had been alternately condemned as the product
of the Fall, man’s expulsion from the Garden of Eden, and
mystically revered as the work of an almighty Creator. By the
thirteenth century, the tide shifted toward a view of the earth
as a glorious illustration of God’s wisdom and power. The Lord
had given man two books, it was said: the Bible, which was
written by the “light of revelation,” and Nature, written by a
second light, the “light of experience.” Both were expressions
of the same hand. The study of scripture and the philosophical
investigation of nature might be understood as two roads leading
back to the same ultimate source.
Nicolaus of Cusa took matters a significant step further by
likening God to the then-novel mathematical concept of infinity. oo
Man is a finite creature, and since the multiplication of a finite
number by any other is always finite as well, God (as infinity)
is forever unattainable. The ratio of any finite number to infinity
is always infinite, no matter how great the finite number. Up to
this time the cosmos of the medieval mind had been a continuous
chain of beings (Angels, Archangels . . . ) stretching from man
upon the earth to the heights of heaven and God. Grosseteste
69
CATCHING THE LIGHT
had certainly still embraced such a vision, but with Nicolaus
of Cusa the chain was broken. As if in a second Fall, man and
woman were separated further still from the divine. Cusa ob
served that man, finite being that he is, can never attain to the
Godhead, no matter how many rungs he climbs on the ladder
leading to the spirit. Mathematical logic had proven that our
slightest imperfection places us infinitely far from God. Better
then to study the earth, the finite creation of God, than to
presume to moral perfectibility and proximity to the Godhead.
For centuries, humankind, eyes fixed on the sun and stars, had
aspired to become gods. Then, quite suddenly, the last hope
was taken away for a direct ascent, and we had to be content
to study the reflected luster of God in earthly knowledge.
The light of nature, therefore, became an ever-more-
trustworthy source of human understanding, and Renaissance
artists such as Leonardo da Vinci, Brunelleschi, and Albrecht
Diirer stood ready to see by its light. Bookish scholarship was
spurned: “If indeed I have no power to quote from authors,”
wrote Leonardo, “it is a far bigger and more worthy thing to
read by the light of experience which is the instructress of their
masters.”12 Moreover, following Nicolaus of Cusa, mathematics
brought with it the surety now felt to be lacking in theological
or philological analysis, and once again Leonardo was in the
, lead: “There is no certainty where one can not apply any of the
mathematical sciences.”13
The enormity of this transition should not be missed. We
witness here other faces of the paradigm shift described by
Thomas Kuhn in his Structure of Scientific Revolution. As Al
brecht Diirer looked through the Cartesian latticework of threads
at a reclining Aphrodite, and as Leonardo da Vinci sketched
the act of love in anatomical cross section, their way of seeing
passed from that characterizing one era to that of another. They
crossed over the threshold to scientific sight. In the two hundred
years from Duccio to Diirer, Western man moved from a religious
70
The Anatomy of Light
perspective to a scientific one, from a moral to a material uni
verse. Brunelleschi had strode from the sacred, interior space
of the Duomo in Florence to its doorway, to the threshold of a
new vision. His successors passed from the doorway into the
lively, secular space of the piazza. Initially there lingered the
memory, the sounds and smells of the rituals performed inside,
but the life of the piazza held its own fascination. Its clamorous
voices did not sing psalms in church Latin, but called out in
the Italian used by Leonardo and Galileo. Its offering was not
the body of Christ, but wares suited to worldly needs. What
seemed beyond reach in one domain could be acquired in an
other, if one but had the currency of the marketplace.
CONTINUING ACROSS the piazza of Santa Maria del Fiore, one
comes to the east door of the baptistery, the “Door of Paradise”
as Michelangelo called it. Commissioned of Lorenzo Ghiberti
in 1425, about the time of Brunelleschi’s demonstration, it in
cludes the exquisite panel of the wise King Solomon receiving
the beautiful Queen of Sheba in his resplendent temple, boldly
executed in the new linear perspective. Solomon’s temple had
always been the biblical archetype for church architecture, the
inspiration of architect and mason. The temple’s master builder,
Hiram of the line of Tubal-Cain, is not shown in the company
of the priestly king and queen. He remains, like so many other
master craftsmen, unseen.14 Yet the times are changing. The
marriage taking place in fifteenth-century Florence and else
where is not between secular and religious powers, but between
techne and sophia, between craft and wisdom.
Until now, learning required leisure, and leisure required
slaves or the support of the Church. Scholars and priests spurned
the manual arts as beneath the dignity of their noble aspirations.
Yet the mood of the time was shifting. After all, God, too, had
71
CATCHING THE LIGHT
• t
----------- -4
j
The marriage of the Queen of Sheba to Solomon the Wise,
'N1® ‘u rendered with exquisite linear perspective in bronze by
Ghiberti.
labored for six days in creating this world, and Jesus was a
carpenter. Human labor might be more than penance, it might
even be the imitation of God.'0 Monastic life changed during
the Middle Ages, coming to value work as prayer, and thereby
helping to form a climate of thought where Hiram could appear
at the flank of the Queen of Sheba. I imagine her as an inter
mediary between the priestly wisdom of Solomon and the earthly
forces, of which Hiram was supreme master.
From the union of scholar and craftsman, there flowed a new
stream. Its current carved the canyons we now inhabit. Galileo
was one of the early merchants to ply its waters, and one of the
things he sought was the body of light.
72
J
The Anatomy of Light
The Perfect Argument
SAGREDO: 1 cannot without great astonishment hear it
attributed as a prime perfection and nobility of the bodies
of the universe that they are invariant, immutable, and
inalterable. For my part I consider the earth very noble and
admirable precisely because of the diverse alterations,
changes, generations, etc., that occur in it incessantly.
Galileo, 1632
In his three monumental compositions, Paradiso, Purgatorio,
and Inferno, Dante had given detailed information on the dis
position of heaven, the anguish of souls in purgatory, and the
eternal damnation of the wicked. In view of Galileo Galilei’s
ultimate condemnation by the Inquisition, it is curiously ironic
that in 1588 the Florentine Academy invited him, then a gifted
young scholar and mathematician of but twenty-four years, to
address its august body on the location, size, and arrangement
of hell as described in the Inferno.11 In retrospect it seems an
unlikely subject for the young man who was later to lecture the
world on other topics, such as the mathematics of projectile
motion, the construction of the telescope, and his revolutionary
observations made with it. Perhaps the later condemnation of
an aged and blind Galileo by the Holy See was intended to speed
him on his way to eternal rehabilitation in a place whose di
mensions and accommodations would be well known to him from
his youthful research. His own position was that “the holy Bible
and the phenomena of nature proceed alike from the divine
Word, the former as the dictate of the Holy Ghost and the latter
as the observant executrix of God’s commands.”18 Thus no con
tradiction could exist between the two books (scripture and na
ture) if both were read correctly, and no damnable violation of
God’s trust was performed in attending to and experimenting
with the book of nature.
What then was the argument between Galileo and his im-
73
CATCHING THE LIGHT
placable opponents? In large part it was over the location of
perfection. Aristotle in his On the Heavens had stated the matter
clearly. The earth, immobile and at the center of the universe,
was surrounded onionlike by sheaths first of the four elements
(earth, water, air, and fire, in that order), and then by the seven
planets: moon, Venus, Mercury, sun, Mars, Jupiter, and Saturn.
From the moon out, the universe was made of an incorruptible
“quintessence" or fifth substance, whose sole movement was
perfect circularity. All heavenly objects, quintessential in na
ture, should also, therefore, have a perfect and unblemished
shape—the sphere. Moon, sun, planets, and stars were con
ceived of and “seen” as polished, transcendent, spherical ob
jects unsullied by the dirt, chasms, and mountains of the
“sublunar” realm we inhabit. Perfection was up, rank upon rank
above man, as Dante, Grosseteste, Dionysius, and myriad others
had imagined, in a great angelic chain of being, now stiffly
reflected in the incorruptible, crystalline spheres of the planets.
It seemed axiomatic that spiritual perfection was reflected in
physical perfection. Odysseus, washed up naked and briny on
the Phaiakian shore of the island of Skeria, “looks like one of
heaven’s people” after a quick bath.19 His uncertain origins were
recognized as noble by his skill in games, casting his discus
many times farther than all the Phaiakians.20 The spiritual per
fection of sun, moon, and stars should, likewise, find full re
flection in their physical attributes: unblemished, spherical, and
eternal.
In 1609 Galileo heard that “perspective glasses,” what we
would call telescopes, had been made by the Dutch spectacle
makers of Middelburg. Galileo investigated the properties of
various lens combinations and constructed several improved
telescopes for himself. Pointing his novel device at the heavens,
he invaded sacred space with a secular and scientific mind.
Unlike his contemporaries, Galileo was not content with leaving
the sanctuary of the Church for the bustle of the piazza, but
74
The Anatomy of Light
insisted on bringing the money changers into the temple. Gazing
“heavenward,” he saw, instead of angels and perfection, craters
and mountains on what had once been the mirror surface of the
moon. New planets (moons) circled distant Jupiter, and even
the sun, that source of God’s pure light, was marred by hideous
blemishes—sunspots. Perfection lost. No longer could one look
up to it on a starry night, no longer did the gods or God illumine
the universe as Helios. The sun was indeed a “fiery rock,” as
Anaxagoras had prophetically declared two thousand years ear
lier. The analytic and earthen eye—Brunelleschi’s, Leonardo’s,
Diirer’s, and Galileo’s—had reached out and touched the heav
enly spheres. The skies’ crystalline purity turned to common
stone, and the ground shook under both the university and the
Church, always the great bastions of conservative thought. In
the light of Galileo’s discoveries, where could one now hope to
find perfection?
Galileo could answer that he had not destroyed perfection,
but only displaced it to its rightful location. In exploding
the myth of incorruptible heavens, he moved perfection to the
mind and the insubstantial universe of pure mathematics. The
earth and its manifold creatures had always been seen as
transient and corrupt; now the entire universe shared the fate
of man—imperfection. Pure thought alone, mathematics,
remained inviolate. By unbelievable coincidence, this imperfect
world had hidden beneath its scaly skin the pristine forms of
mathematics as laws of nature. As Galileo’s contemporary Jo
hannes Kepler declared, “Geometrical reasons are co-eternal
with God.”21 From the time of the ancient Pythagoreans, and
before them the astronomer-priests of Babylon, men had seen
number in the world. Freed from religious constraints, that math
ematical view erupts in the sixteenth century to become one of
the two main currents in scientific thought.
While perfection was becoming disembodied a second and
equally essential stream developed in Galileo’s mind as if to
75
CATCHING THE LIGHT
balance the first. No longer was Aristotelian physics tolerable
with its four elements, doctrines of matter, form, potentiality,
and actuality, etc. In its place should stand clear causal ac
counts, understood to be material and mechanical in character,
for every phenomenon.
Views such as these had been advanced in antiquity by Greek
atomists, but they were abjured as atheistical and dangerous.
Plato in the Laws of his ideal city reserved a specific punishment
for those who held and promulgated such views.22 They were to
be taken to the desert, placed in solitary confinement on bread
and water, and taught the error of their ways. If unrepentant,
they were to be executed. All this from a philosopher whose
own beloved teacher, Socrates, had been sentenced to death for
corrupting the youth of Athens. Galileo was not sentenced to
death for displacing perfection; he had powerful supporters, and
was politic enough to dissemble to the Inquisition while laboring
furtively to publish his heretical views.
If Aristotle was to be set aside, what could act as the basis
for secure knowledge? If one wished to begin with sense impres
sions, how could one move beyond their often misleading report?
In earlier times, one saw through them the indirect messages
of the gods. What stood behind them now? Especially in his
early writings, Galileo’s position was that sense impressions were
purely subjective, that “tastes, odors, colors, and so on were
no more than mere names so far as the object in which we place
them is concerned, and they reside only in consciousness.”23
For example, something we experience as hot is really only “a
multitude of minute particles having certain shapes and moving
with certain velocities.” Such is heat. Although Galileo appeared
cautious in his speculations about an ultimate materialistic,
corpuscular basis for the world, later philosophers and scientists
felt no such scruples. If a divine world existed, it was severely
separate from the material world of the senses, and little or
nothing stood between them.
76
The Anatomy of Light
PLATO HAD ARGUED in his cosmological dialogue the Timaeus
that between two numbers a middle term could always be found
that mediated and made more beautiful the mathematical re
lationship they expressed. His Demiurge created the world ac
cording to such a threefold relationship. In the graduated
r-
cosmos, intermediaries had connected earth and heaven, man
and God. With Cusa and Galileo the middle term disappeared,
and a yawning abyss was felt to separate the material from the
mathematical, the real from the ideal, and man from his Maker.
A Body Called Light
The Grimms’ tale of the “Bremen Town Musicians” ends with
a group of frightened thieves cautiously reentering their forest
hideout in the black of night. They have already been driven
out once by four intrepid animal musicians. The lead thief tiptoes
across the dark room toward the fireplace, where he sees two
glowing orbs, which he takes to be hot embers. They are, in
fact, the radiant eyes of the cat who shrieks and claws as he
reaches for her, at which point donkey, dog, and rooster join
in the fray to drive the intruders away.
The glow of a cat’s eye, and her ability to navigate at night,
convinced early opticians of the reality of the visual fire within
the eye.* Yet the question still remained, if the antique mind
thought a source of light was in the eye, why can’t you see at
night? Various answers were given, but Aristotle’s was influ
ential and straightforward: dark air was opaque. Light the lamp
*We now understand the phenomenon as the reflection of dim light in
the spherical eye of the cat, and make use of the effect by putting tiny
glass beads in the lettering of traffic signs to reflect the light of oncoming
cars.
77
CATCHING THE LIGHT
and it becomes transparent. Light for Aristotle was “the ac-
tualization of the potentially transparent.”
To explain: modern calculators or portable computers often
have LCDs or “liquid crystal displays.” They work by making
segments of the display selectively opaque. It is done by sand
wiching a special “liquid crystal” between transparent elec
trodes and applying small voltages to those regions of the
electrodes that should be changed from clear to black, from
transparent to opaque.24 Here the action of electricity changes
the state of the liquid from opaque to transparent, or as Aristotle
might say, from potentially transparent to actually transparent.
In just the same way, fire can change the state of air from dark
to “light,” from potentially transparent to actually transparent.
Then you, with your active eyes, see through the room. For
■ Aristotle, light was not a thing but a condition or state of a
I medium. Aristotle’s was a disembodied understanding of light,
a subtle conception that easily accounted for the invisibility of
light. Light had no substance or structure, but by the seventeenth
century matters would change.
In 1611, Galileo had brought with him to Rome not only his
famous telescope, but also a little box containing remarkable
stone fragments of a mineral called by its Bolognese discoverer
spongia soils, or solar sponge.2a Placed in a dark room, these
cold stony fragments would continue to glow if they had recently
been exposed to light. Cold light! According to Aristotle, fire
rendered air transparent, not cold stones. Perhaps the philos
opher was wrong, and light, like heat, was better understood as
a corpuscular and mechanical action. Galileo, therefore, sug
gested that when substance is reduced to “truly indivisible at
oms” then “light is created.”26
Light may be a body like all others but only smaller, or
perhaps even the smallest. Perfect, incorruptible light, the sign,
78
The Anatomy of Light
symbol, and very nature of God or Ahura Mazda, is really only
another earthly body. Is such a suggestion any less revolutionary
than the displacement of perfection from a spiritual universe to
a mental abstraction? From the Milky Way (newly explained by
Galileo) to the tiniest atoms of light, everything is uniformly
material and, as a consequence, equally amenable to the new
methods of scientific investigation being developed. Reason and
experiment are now adequate to encompass the two infinities of
the macro- and microscopic.
Galileo hesitated to give an unequivocal opinion on the nature
of light, and even declared emphatically that he had “always
been in the dark” on the question of “the essence of light” but
a scant two years before his death. After reading an attack on
his materialistic theory of light, Galileo rejoined with a story.
He was willing to be imprisoned in a pitch-dark cell, he said,
surviving on bread and water, if only he would be guaranteed
on his release into light that he would know what it really was.
If his sentiments were genuine, then he died unrequited, longing
still to know the true nature of light.
Galileo had, however, voiced an opinion, one that gained
intellectual momentum and conceptual detail. Light was not God
but a body. If a body, then it would have an anatomy open to
dissection and scientific investigation like other bodies. None
achieved more in this vein than Sir Isaac Newton.
Newton: Standing on the
Shoulders of Giants
JFe are like dwarfs sitting on the shoulders of giants; hence
we can see more and further than they, yet not by reason
of the keenness of our vision, but because we have been
raised aloft and are being carried by men of huge stature.
Bernard of Chartres, 1115
79
CATCHING THE LIGHT
Isaac Newton’s fame began like Galileo’s with the construction
of an instrument, the same kind of instrument, a telescope in
> 1669.27 As holder of the Lucasian chair in mathematics at Cam
bridge University, Newton, then in his late twenties, could do
more or less what he had a mind to. And his mind chose to
struggle with an array of questions that ranged from alchemy
and theology to mathematics and optics. Out of his reflections,
readings, and experimentation with optical phenomena, came
the idea for a novel kind of telescope, one that did not rely on
glass lenses* but on a curved mirror for gathering and concen
trating light from distant objects. Laboring in his own workshop,
the Lucasian professor ground a suitable mirror surface and
made the tube and mount for what would become the prototype
for all major astronomical telescopes the world over to this day.
He took pride in the accomplishment of his own hands, and
must have relished the initial failure of professional instrument
makers who were later charged with replicating his design.
Except for a few colleagues at Cambridge, Newton’s remark
able talents and diverse accomplishments were unknown. Ru
mors of his device eventually reached the recently formed “Royal
Society for the Improvement of Natural Knowledge,” who nat
urally requested the pleasure of examining it. Newton obliged
by sending them a twenty-five-inch-long version of his reflecting
telescope. After Society members studied it, they were unani
mous and generous in their praise, longing to know more of
Newton and to have him as one of their own. In response to
their accolades, Newton sent his famous letter of February 6,
1672, to Oldenburg, then secretary of the Royal Society. In it
he laid out the results of his experiments on the nature of light
and color, and put down the foundation for the understanding
that dominates popular discussions of light to this day.
*Refracting telescopes of the period suffered from “chromatic aberration,”
that is, images of differing color focused in different planes.
80
The Anatomy of Light
The origins of Newton’s conception of light began early, with
his boyhood love of tinkering and his natural inclination toward
a mechanical philosophy of nature, but we will pick it up in
1665 when it “pleased Almighty God in his just severity to visit
the town of Cambridge with the plague of pestilence.”28 During
the two years of plague in Cambridge, Newton lived quietly at
home with his mother in Woolsthorpe. These two years are
legendary in the history of science, sometimes called the anni
mirabiles, because of the sheer number and significance of the
scientific ideas that came to Newton during his months of retreat.
Certainly his previous years of energetic study had created the
requisite fertile bed, but the scale of originality he demonstrated
during these two years of meditation was unprecedented. Among
the accomplishments Newton credited to those years (and there
were others) were the invention of calculus, his theory of gravity
and planetary dynamics, and his theory of light and color. Any
one of these would have guaranteed his name to posterity.
What connected such disparate subjects as these, and so
informed his and now our understanding of light, was a fun
damental principle of which Newton was supreme master—
analysis. It was not original with him, but he applied it with an
unprecedented practicality and brilliance. As an example of the
power of his analysis, let’s look at an important derivation from
his gravitational theory.
All material objects attract one another, but exactly how?
How strongly and in what way does attraction depend on distance
or the sizes of the objects? Newton solved the problem twice,
once for himself and once for us. In the year 1666 while at
home, Newton was musing in a garden on the problems of mo
tion. As we all know, an apple fell. Yet Newton saw in its motion
not what he and others had seen before. In the descent of the
apple he also saw the travels of the moon. They were the same!
If gravity could extend to the moon (a new thought), then the
moon, too, should be falling. If, in addition, it was moving
81
CATCHING THE LIGHT
laterally as it fell, it might just go around the earth instead of
hitting its surface. “Whereupon,” we are told in early accounts,
“he fell a-calculating what would be the effect of that suppo
sition.”29 His calculations showed that his supposition might in
fact work.
Twenty years later, in his Principia, Newton needed to con
vince us, his astonished readers, that he had not been hallu
cinating when he saw the apple and the moon as one, and for
this he needed analysis. Ironically, Newton recognized that al
though a derivation of the universal law of gravitation for large
bodies such as the earth could be done easily using his newly
< invented calculus, he chose to use the ancient and trusted meth-
] ods of geometry, but adapted to modem requirements. These
requirements were none other than those implied by the calculus
where the ratios used may well be between quantities that ap
proach zero. As Newton’s major biographer Richard S. Westfall
writes, “Euclid would not have recognized his offspring.”30 With
Newton, the principle of analysis became formally embodied in
mathematics and so, too, in his thinking. In that moment a
fundamentally new kind of thinking, and so, too, of seeing,
entered the West, one that challenged the human imagination
to attain new heights of abstraction.
Until the seventeenth century, mathematics had been domi
nated by plane geometry. Even though the elements of geometry,
such as lines, triangles, etc., are purely ideal and nonphysical
in character, they remained visualizable. In neoplatonic
thought, geometry dwelt midway between the material sense
world and the pure unextended, formless world of Ideas. Indian
philosophy made a similar distinction. Above the physical level
was the rupa or “form” realm; beyond it was the formless, arupa
realm. With the development of calculus, mathematics crossed
to the realm of arupa, to the arena of hitherto unimaginable
quantities.31
As mathematicians such as Newton and Leibniz were con-
82
The Anatomy of Light
ceiving of infinitesimal quantities, they and others were simul
taneously groping toward a material atomism of the kind hinted
at by Galileo. The earth is an enormous mass, but it can be
thought of as the nearly infinite sum of tiny masses. This is the
key idea behind several of Newton’s demonstrations in the Prin-
cipia. The earth as a whole will attract the moon and the apple
in identical ways because every atom of earth has its own unique
gravitational relationship to the moon and apple. The total effect
is gotten by simple addition (or integration in the continuous
case).
Can analysis, so successful in physical dynamics, be applied
to light? What are the least parts of which light is made? Can
they be added together and separated not only in the mind but
in the laboratory as well? Newton again provided the answer
and showed others the means.32
In LATER YEARS Newton was painted again and again as standing
in a darkened room with a narrow beam of light passing from a
shuttered window to a prism held before his calm but penetrating
eyes. Exiting the prism was a rainbow of colors that brightened
the air, falling on the far wall of the room. Here was the analysis
of light, the extraordinary moment seen by Newton as the break
ing up of light into its “least Parts.”
In keeping with the artist’s image of him, Newton opens his
Opticks with a definition of the fundamental unit of light out of
which the entirety of his optics would systematically be built.
“Definition I: By Rays of Light I understand its least Parts, and
those as well Successive in the same Lines, as Contemporary
in several Lines.”33 The “ray of light” is the fundamental unit
or conceptual atom of his theory. The origin of this definition
is certainly Newton’s early and abiding corpuscular conception
of light.34 However, the analysis of light occurs on two levels,
one mental and the other physical, and when under attack,
83
CATCHING THE LIGHT
Newton made good use of the distinction. In such times he
invariably responded that his “least Parts” or rays were purely
formal, theoretical constructions that did not commit him to a
particular physical model of light.35
Light rays, according to Newton, are created in the sun and
travel through space to us unchanged by reflection, scattering,
’■I or refraction. Moreover, each kind of ray produces in the eye a
different sensation: red, green, blue, and so on. Natural sunlight
is the sum of many such rays and so appears white. A prism
acting on white light is the analyzing instrument that separates
its constituent rays into their original classes. If the first prism
is followed by a second, the “colorific rays” can be brought
i together again and so re-create white light. In Newton’s hands
* white light, like the earth, was analyzed into its “least Parts,”
the various color-producing rays, and the prism was the device
that performed the physical analysis.
Although reasonably careful to hide his particular model of
light behind adroit philosophical language, Newton could not
resist formulating his views on the nature of light, at least in
the form of questions.36 For example, in query 29 of the Opticks
he asks, “Are not the rays of light very small bodies emitted
from shining substances?” The smallest such bodies he thought
would evoke violet and blue impressions with larger and larger
corpuscles correspondingly responsible for green, yellow, or
ange, and red. Our sensations of color, therefore, were to be
understood as our subjective response to the objective reality
of corpuscular size.
In addition to suggesting that lights of various colors might
i ultimately be small bodies of various sizes, Newton explained
' important optical phenomena by calculating the change in tra
jectory of such bodies, for example, as they passed from air to
water and were refracted. These ideas were drawn directly from
his study of mechanics. The Principia, his book on mechanics,
even contained a derivation of the law of refraction according
to the corpuscular model of light.
84
The Anatomy of Light
Thus, not only was light a body, but its laws of motion were
none other than the laws Newton had already discovered as
those governing the motions of planets and apples. Forces of
attraction and repulsion pulled and pushed projectiles of light
through the world. Left undisturbed, they would travel in straight
lines according to the law of inertia like all other material ob
jects. The dynamics of light were identical to the dynamics of
the planets.
The cosmos was unified, from phenomena on the grandest
scale to those on the most minute, from the stars to the particles
of light they emitted. It was a single vision. Gone was the
multiplicity of being in which a unique intelligence was asso
ciated with each heavenly sphere. All was reduced to matter
moving in obedience to the laws Newton unearthed. It was an
impressive, even a compelling vision.
Newton, conscious of his accomplishments, feigned modesty
by reaching back to a medieval image. In Chartres cathedral
above the south portal rise beautiful, large stained-glass lancet
windows. They show the New Testament Evangelists seated on
the shoulders of Old Testament prophets, symbolizing thereby
the foundation on which they stood as they continued God’s
work. Their vision rested on the past vision of the prophets.
Newton’s accomplishments likewise rested on changes in con
sciousness wrought by his predecessors. Correctly, he wrote: “If
I have seen further, it is by standing on the shoulders of
Giants,”37 adding to their views his own powerful vision of the
world.
ASIDE FROM THE objections of a few lone scientists, theologians,
and artists, the response to Newton’s achievement was euphoria.
Scientists and philosophers of the day were dazzled, and con
temporary poets and artists likewise joined in lavish praise of
Newton.38 For many years, some had felt the traditional sources
of poetic vision were failing; they were dead or dying. The Muses
85
CATCHING THE LIGHT
no longer sang as they once did. The “twilight of the gods” had
come, and like the fairy beings of legends, the gods had retreated
before the crass onslaught to more congenial climes. If a poet
would speak truth, where could he turn for inspiration? Finally
an answer could be given—to Newton. John Hughes envisioned
him this way:
The great Columbus of the skies I know!
’Tis Newton’s soul, that daily travels here
In search of knowledge for mankind below.
0 stay, thou happy Spirit, stay,
And lead me on thro’ all the unbeaten Wilds of Day.39
Newton became the inspiring spirit, the Muse of poet and phi-
losopher alike, and Tnany were the songs sung in his praise or
verses written to make poetry of his scientific tracts.40
It would be a century before poets would turn on Newton and
the despotism of his botanizing eye. They would then lament
the dismembering of the world into parts, so that true wholes
were nevermore seen. But for now most were content.
In the years following the 1704 publication of the Opticks,
Newton’s theory of optics, shorn of its subtleties, was broad
cast to the widest audiences. Three different means were
used. At the universities, his theory was usually accepted
uncritically and without philosophical sophistication, as the
basis for instruction. Since the truth had been discovered,
there seemed little need for original optical research, and
so none was undertaken. The competing wave theory of the
Dutch physicist Christian Huygens was lost in the blaze of
enthusiasm for Newton’s optics, and for Newton himself. Not
\ surprisingly, therefore, little criticism arose from university
quarters.
For the public, another disseminator of Newton’s thought was
the popular traveling lecture-demonstrator who visited science
\
86
The Anatomy of Light
clubs for both the upper and middle classes, demonstrating and
explaining the latest seemingly magical, scientific develop
ments. The third means was through popular accounts written
by literary figures. Perhaps the most illustrious to undertake
this task was the brilliant wit of the French Enlightenment,
Voltaire. Tutored by his equally brilliant mistress, the “immortal
Emily,” whose love of mathematics and mathematicians was
legendary, Voltaire sought “to remove the thorns from Newton’s
writings without loading them with flowers which do not suit
them.” Following the publication of his Elements of Newton’s
Philosophy, even his Jesuit adversaries had to admit that “all
Paris resounds with Newton, all Paris stammers Newton, all
Paris studies and learns Newton.”41
THE FORCE of these many efforts shaped a new way of seeing
in the widest literate circles. The view of light gradually became
sharper and more popular. What had at first been the province
of a few mathematically minded scientists was taken up again
and again by artist and author and academic. What their ren
derings sacrificed in philosophical subtlety was replaced by
materialistic simplicity. Gone were Newton’s or Galileo’s hesi
tations about the “true” nature of light. Light was a body, and
its movements were like those of all other bodies. Such were
the facts.
In France, a parallel, if distinctive, line of inquiry had been
launched by Rend Descartes some decades earlier. Like New
ton’s understanding of light, it was founded on a rational and
predominantly material conception of the universe. Cartesian
and Newtonian physics vied for dominance for more than a
century. Ironically, the origins of Descartes’s universal math
ematical science undercut his rationalist agenda, for Descartes
received his mandate in the classical manner of all biblical
figures, through visitation in a dream.
87
CATCHING THE LIGHT
Descartes’s Dream
I frankly confess that in respect to corporeal things I know
of no other matter than that which the geometers entitle
quantity.
Descartes, 1644
While at Cambridge, the twenty-two-year-old Newton purchased
a copy of Rent: Descartes’s book Geometry. He read the first
two or three pages and found he could understand no more. He
returned to the beginning and started again, managing to com
prehend another three or four pages, and so it continued until
he had satisfied himself concerning the mathematics of this
French master. Here was a book and a thinker worth troubling
over.
Forty-six years Newton’s senior, Descartes had advanced a
philosophy of nature in which the universe was a mechanism,
from the simplest atom to the most complex aspect of the living
human anatomy. “The rules of nature are the rules of me
chanics,” he declared.42 This dictum was taken up and pro
moted in the most powerful ways by his friend and advocate,
the commanding French monk Marin Mersenne. The “mechan
ical philosophy” of the seventeenth century was largely forged
by these two Jesuit-trained giants, rising in Descartes as the
fruits of his solitary meditations, and then promulgated by Mer
senne through his vast correspondence and highly influential
connections. At their hands, as well as under the swords of the
Duke of Bavaria’s Catholic armies during the Thirty Years War,
the magical philosophy of Renaissance animism, Hermeticism,
Cabalism, and all their attendant trappings, were destroyed. In
their place would emerge Descartes’s “universal and admirable
science.” How deliciously ironic that the conception of Des
cartes’s consummately rational science occurred in a dream.
In November of 1619 Descartes, a twenty-three-year-old sol-
88
The Anatomy of Light
dier-philosopher, retired for the winter to a modest house near
| Ulm.43 During the previous twenty months, he had traveled
through Germany and Holland, corresponding with mathema
ticians and philosophers whose range of interests included not
only the orthodox problems of science, but also the connections
between science and spirituality. Some were certainly involved
in Rosicrucian spiritual pursuits, which in these years were
experiencing an enormous surge in public attention.44 Two
short tracts concerning Christian Rosenkreutz and his brother
hood had just been published, the Famafraternitatis (1614) and -i
the Confessio (1615), causing the so-called “Rosicrucian Furor.” J
The Confessio put forward the tenets of the fraternity of Christian
Rosenkreutz, which included the conviction that on the basis
of God’s revelation and human invention, observation, and un
derstanding, “if all books should perish, and by God’s almighty
sufferance, all writings and all learning should be lost, yet
posterity will be able only thereby to lay a new foundation, and
bring truth to light again.” This was to Descartes’s liking. He
searched for a member of the “invisible brotherhood” but main
tained he had found none. Through his contacts with other
seekers, however, the impressionable Frenchman was infused
with the holy significance of his task: the creation of a new
science of nature.
In the middle of these influences, and following a period of
intense solitary meditations, Descartes tells us that on November
10 he had a vision and a three-part dream that revealed to him
his vocation and the foundations of the science he was to create.
He held fervently that this single episode was the most important
occurrence in his entire life, and in gratitude for it he vowed
to make a pilgrimage of thanksgiving to the Virgin at Lorette, »
a promise he kept five years later when he traveled there on
foot from Venice. Descartes’s own detailed account of his vision
and dream has unfortunately been lost, but we have a faithful,
if incomplete and pale, summary of it from his early biographer,
89 \A^.
CATCHING THE LIGHT
Baillet, who studied it. To our unholy eyes it may seem unre
markable, but Descartes knew it to be his personal Pentecostal
visitation, his life’s epiphany.
In the first part of his dream, Descartes finds himself strug-
. gling against a tempestuous wind in order to reach the Church
j of the College of La Fleche (where he and Mersenne had been
educated) to say his prayers. Descartes turns to show courtesy
to a man he has neglected to greet, and in that moment is blown
violently against the church. He is presently informed that some
one has something for him—a melon. Descartes awakens to the
experience of pain, turns over on his right side, and prays for
protection. Falling asleep once more, he has another dream.
Concerning it we only know that it fills him with terror; he is
awakened by a noise like the sound of thunder and sees thou
sands of shimmering sparks in his room. In the final part of his
dream, Descartes sees a dictionary and a Corpus poetarum, on
his table, open at a passage of Ausonius: quod vitae sectabo
iter?, which means “what path shall I follow in life?” An un-
i known man appears and hands him a bit of verse on which
| Descartes reads the Latin phrase, est et non, “yes and no.”
We possess, through Baillet, but a few meager bits of Des
cartes’s own interpretation of the momentous night. Signifi
cantly, the lightning was taken by Descartes to be “the Spirit
—r of Truth that descended upon him and took possession of him.”
The dictionary symbolized all the various sciences, and the
Corpus poetarum “marks particularly and in very distinct man-
L ner, Philosophy and Wisdom linked together.”
Thus, we find the youthful Descartes in his solitary reflections
in Germany, searching fora means to truth. He has made contact
with the forbidden sciences and is deep in meditations on his
future life. Descartes tells us that “the genius that heightened
in him the enthusiasm which had been burning within him for
several days past, had forecast these dreams to him before he
had retired to his bed.” Thus Descartes knew that the night of
90
The Anatomy of Light
November 10 would bring the answer to his passionate quest.
What was the answer he heard? It was, as Jacques Maritain
calls it, the Pentecost of Reason.
In the months following, Descartes would write his most im
portant foundational work, Discourse on Method, whose original
title was to have been Project of a Universal Science Destined to
Raise Our Nature to Its Highest Degree of Perfection. As the
Confessio had said, all the prior works of scholars needed to be
set aside. Until now we have “been governed by our appetites
and our preceptors. As a result of these reflections, it was borne
in upon me [Descartes] that as far as all the opinions I had
received thus far were concerned, I could not do better than to
undertake once and for all to get rid of them.” The Spirit of
Truth had descended on him; it was his solitary task to adjust
everything to “the level of reason.” His admirable and universal
science was not the work of a brotherhood, occult or public,
but the creation of a single mind, his own. Descartes’s was to
be a godly science that rises with certainty to the simple intuited
nature of things as God knows them. Maritain has called this
science “the mythology of modern times that promised every
thing and denied everything, that raised above all things the
absolute independence, the divine aseity of the human mind.”
It was the child of a mischievous genius conceived in a philos
opher’s brain—the Dream of Descartes. Within this dream, what
is light?
SPACE, Descartes ARGUES, is inconceivable apart from matter.
Therefore, where there is space or “extension” there must also
be matter, always. In an analogy that befits his nationality,
Descartes likened our universe of space to the winemaker’s vat
just after the harvest. The vessel is filled with half-crushed
grapes, which are completely surrounded, even permeated, with
the juices they have given up. Likewise, the vast reaches of
91
CATCHING THE LIGHT
space carry planets, stars, and moons like grapes in their own
juices. Space for Descartes is filled with an atomistically con-
ceived plenum, a material fluid that occupies all voids and drives
the planets in their courses like bits of grass caught in the
whirlpools of a stream. Descartes could thus command, “Give
me motion and extension and I will construct the universe!”45
And light?
In Descartes’s view, between the eye and every object is a
column of plenum along which an action can travel. Light is
neither a projectile nor a fluid flow, but a “tendency towards
motion” in the plenum that propagates at infinite velocity along
the column. Sight, he writes, is like the blind man’s stick. As
he walks he feels around him with the stick. An object that
strikes one end instantly causes a push at the other. In the same
way, an object affects the plenum around it, causing a concussion
at the eye, and so we see. Sight and light, according to Descartes,
were to be understood as pure mechanism. This is a watershed
analysis that offered an imagination of nature and a basis for
scientific inquiry that would dominate science for three hundred
years. As the distinguished Harvard historian of science A. I.
Sabra writes, mechanical analogies had been used long before
Descartes to help explain certain optical phenomena, “but the
Cartesian theory was the first clearly to assert that light itself
was nothing but a mechanical property of the luminous object
and of the transmitting medium. It is for this reason that we
may regard Descartes’s theory of light as the legitimate starting
point of modem physical optics.”46
While the impact of his conception of nature and science
proved seminal, the specifics of Descartes’s theory of light were
relatively short-lived. Across the waters of the channel, Newton
drew out certain of the absurdities to which such a view of light
and cosmos led. By contrast, Newton’s own dynamical formu
lation of optics along mechanical lines gained widespread ac
ceptance, but it, too, suffered from serious deficiencies. In the
92
The Anatomy of Light
midst of the triumphant march of Newton’s corpuscular optics,
persistent critical voices could be heard. Many spoke of specific
failings of the theory.47 For example, an analysis of two people
staring at one another in the projectile view would seem to
require that particles of light pass simultaneously along the same
path in opposite directions. Also, if the sun was giving off
copious numbers of particles, why does it not waste away over
time? Proponents replied that the corpuscles are extremely
small, so that the sun loses but a drop or two of substance in
the course of a day. Or what of the innumerable corpuscles that
must pass through the pinhole of a camera obscura without
affecting the image? Again smallness came to the rescue. Yet
if they are so small, then how can they be so effective? When
investigating the impact of intense, focused light, the experi
mentalist John Michell had been troubled by the melting of his
copper-plate detector at the focus of a two-foot reflector. Light
can yield dramatic effects. If the bodies of light are small, they
must also be exceedingly numerous in order to account for such
power.
Nieuwentijdt in his book of 1718, The Religious Philosopher,
calculated that 4.1866 X 1044 particles are emitted each second «•
from a candle flame.48 This approaches the number of protons
in the entire earth. Nieuwentijdt took the calculation as evidence
for the constant attention of God to his creation, the “Direction
of an Omnipresent Power, extending its Care over all things,
even the smallest of Bodies.” If light corpuscles were immutable
as Newton claimed, why does the solar sponge seem to emit
light of a different color from that which originally illuminated
it? Such were the unresolved issues.
Descartes’s theory avoided such difficulties by declaring light
to be only a “tendency towards motion,” not entailing actual
motion. In this way Descartes tried to avoid the absurdity of
having an object (corpuscles of light) move simultaneously in
two opposite directions. The particles of the plenum do not
93
CATCHING THE LIGHT
actually move, but only “tend” to move. Rays of light are,
therefore, “nothing else but the line along which this action
tends.”49 Light is not, according to Descartes, the flight of a
projectile, but the propagation of that action.
Hidden within Descartes’s obscure conception of light is a
strength. As puzzling as his view appears, it holds the seed of
an enormously fruitful conception of light, the so-called wave
theory, that would not only overthrow Newton’s understanding
of light, but would ultimately work against Descartes’s own
dream of a mechanical universe.
While there were public and poetic accolades for the cor
puscular theory, scientists of stature like Huygens questioned
it along the above lines. They put forward alternative conceptions
of light. Some spoke of light as a material fluid of fire that flowed
from the fountain of the sun and back again. Others filled the
universe with a substance far more tenuous than air but of
unbelievable rigidity whose vibrations were light. All shared a
material conception of light, but the most successful theories
also possessed elements that would, in the end, undermine that
very position.
''X
SWIM OUT INTO the ocean and notice a few things. The swells of
. the open sea slowly increase in size as they approach the shore.
As they speed toward you there is a moment, especially for the
inexperienced, in which you worry that you will be carried along
and dashed to pieces on the beach. As the wave and the feeling
pass, you notice that you and the water around you did not travel
with the wave toward the beach, but only bobbed up and down.
No water went by, it only moved up and down. What did pass?
You saw it coming, and see it still heading for the beach. What
is it, this wave? It is a form, a shape, a specific action of the
sea.
Could light, like sound, be a shape, a vibration of a universal
94
The Anatomy of Light
ether? Nothing moves, or moves only slightly, except the form
and figure of light. Light then would not be substance, but rather
form! With these thoughts, a small but serious competitor arose
to the projectile theory of light, one that would gradually gain
momentum and find a surprising ally in the new investigations
then under way in electricity.
WHAT had BEGUN as a god, and in Greek hands had become a
luminous inner fire whose ethereal effluence brought sight, be
came in the Middle Ages “the first corporeal form” of Grosse
teste. His was still a metaphysics of light, but within that first
corporeal form lay the seeds of light’s final corporeal form: New
ton’s corpuscular optics. Along the way, perfection was deposed
from its ancient haunts, and in its place reigned abstract math
ematics in brilliant formality. Abstract perfection and substantial
reality, these replaced the moral perfection and power of im
mortal gods. This was far more than an exchanging of ideas; it
entailed a profound transformation in the West’s very way of
perceiving. A material and mechanical eye replaced the moral
and spiritual one of earlier times.
With the scientific revolution of the sixteenth and seventeenth
centuries, mankind entered a new age. It saw itself as entering
manhood and so strove to put away childish things. Light was
shorn of its metaphysical dress, its body was displayed naked
like one of Leonardo’s cadavers. Light’s hard skeleton was laid
bare by the incisive minds of Descartes and Newton, its com
position and principles of motion exposed to enlightened sci
entific sight. As Bernard de Fontenelle, secretary to the Paris
Academy of Sciences, wrote in 1686, the world was but a staged
performance, an opera. The attentions of science were not on
the dramatic action no matter how lavishly set or passionately
enacted, but rather with the calculating view held by one usually
unnoticed during the production—the stage engineer.
95
CATCHING THE LIGHT
The opera-goer does not care how the stage machinery works.
By contrast the engineer in the pit hears every cue, sees the garish
makeup of the actors, operates the lighting, raises the curtains,
and sets in motion the mechanisms of the stage. The performers
and technicians conspire to create an illusion, one that entertains
a contented audience, but the engineer alone knows the full un
garnished truth.
According to Fontenelle, nature conspires in a similar manner
to present a beautiful pageant to the human senses. We can be
entertained by a sunset, a bird’s song, or stirred by human
passions to love and war, but behind these external happenings
lies nature’s secret machinery, the means by which she contrives
to present her illusions to an unwitting audience. Unlike the
common man, the scientist is not satisfied with the show, with
seeing only the surface phenomena. Rather he relentlessly
searches for the real mechanical operations of nature. As Fon
tenelle wrote, “He that would see nature as she truly is must
stand behind the scenes of the opera.”
Fontenelle, like Voltaire, wrote in order to enlighten the cu
rious public about the mechanisms scientists were imagining as
operative behind the outer display of nature. From his pen, as
well as that of Mersenne, ran the ink that would enamor the
public of a mechanical cosmos. The dream of Descartes would
become the dream of an age. The enthusiasms of youth, however,
are often contradicted by the lessons and puzzles of a long life.
Such was the case also with the history of light and the me
chanical universe. Light kept true to its own inviolable char
acter, always ready to cast a new shell of curious design upon
the shore of human imagination.
96
CHAPTER 5
yah- c$
The Singing Flame:
Light as Ethereal Wave
We shape the clay into a pot, but it is the
emptiness inside that holds whatever we
want.
Lao-tzu,
Throughout history, space has been like the emptiness inside
Lao-tzu’s clay pot. It’s not the pot, but the emptiness inside that
holds whatever we want. Ever since we created the concept,
space has held whatever we put into it. We have imagined space
to be many things, and that act of imagination has had impli
cations for our image of light. Endow space with divinity and
light is godlike; discover its shape and light is geometrical; fill
it with matter and light is substantial. From Moses to Einstein,
the history of light is also the history of space.
When on the first day of creation the Lord said, “Let there
be light,” the early Christian fathers understood the first light
to be that noble, spiritual reality they termed lux, which was
the soul of space. They, and medieval scholars after them,
labored long and hard to distinguish lux from its emanation or
97
CATCHING THE LIGHT
bodily counterpart, called lumen. The distinction for us is dif
ficult to grasp, yet was essential to their view of the world.1
Lux was God-given, essential light, the being of light, and
as such a reflection of its Maker. Augustine viewed it as the
simplest, noblest, most mobile and diverse of all corporeal
being. Lumen, by contrast, was the material means by which
our perception of the being of light (as lux) arose. When we
sense the brilliance of the sun, we are perceiving its lux, but
we do so by means of the unseen lumen that connect it to us.
Between the time of Augustine and Galileo, the being of light
that ensouled space (lux) retreated, leaving behind its hard
material vestige (lumen) as a fossil record for the curious natural
philosopher.
Light has fascinated and continues to fascinate, for, as Leo
nardo wrote, “Among the studies of natural causes and laws, it
is light that most delights its students.”2 From the radiant eye
of ancient Egypt to the quantum-field theories of today, light
has sculpted space to suit its demands.
In Grosseteste’s imagination, the unfolding of light from its
primordial hearth gave birth to space as light multiplied itself,
generation after generation, until, spent, it died at the shoreline
of the universe it had created.
The space of Euclid and Brunelleschi was pure geometry.
Light and vision proceeded as rays, a pure geometry of sentient
lines linking soul to world.
For Descartes, space was dimensioned, extended, and there
fore had to be substantial. He could not conceive of an extended
space apart from substance; where there was one, there must
be the other. Looking up by night or by day, one sees “a sky
that is made of liquid matter,” and the planets swirl in its
whirlpools like grass in the eddies of a stream.3 Light, whatever
it is, must transit this material medium. By the end of the
eighteenth century, this material medium was the ether whose
motions became the lumen that caused vision, and lux was no
98
The Singing Flame: Light as Ethereal Wave
longer an attribute of God but merely a subjective phantom of
the mind.
Our understandings of light are thus entwined with our con
ceptions of space. They have coevolved: moral space and spir
itual light, perspective space and geometrical light, material
space and substantial light. Each age emphasized one face of
light and so revealed its own predilections. The corpuscular
conception of light emphasized its substantial nature, and yet
it suffered from serious deficiencies. Perhaps, as some sug
gested, rather than consider light to be a material substance or
fluid, it is a pure form, a figure dancing. After all, the mysterious
nature of sound has finally succumbed and shown itself to be
the vibration of air. Might not light profit from a similar treat
ment? It seems appropriate that a mathematician, whose life’s
work is constant contemplation of the insubstantial, should first
advocate light as a figure dancing through the ether.
At fifty-three years of age, and nearly blind, Leonhard Euler,
the greatest mathematician of the eighteenth century, the author
of innumerable mathematical and physical treatises, and the
most prolific member of both the Prussian and Russian acade
mies of science, honored the petition of a curious young German
princess from Anhalt-Dessau to hear his thoughts on science.
His letters to the princess between 1760 and 1762 covered every
aspect of science and quickly captured the attention of Europe,
with thirty-six editions in nine languages. Having written the
princess concerning sunlight, he anticipated her question, and
ours. “Having spoken of the rays of the sun, which are the focus
of all the heat and light that we enjoy, you will undoubtedly
ask, What are these rays? This is, beyond question, one of the
most important inquiries in physics.”4
His response, coming as it did in an age reverberating with
the ideas of Newton, was a piece of heresy. The rays of sunlight
are, he rejoined, “with respect to the ether, what sound is with
respect to air,” and the sun is but a bell ringing out light.5 With
99
CATCHING THE LIGHT
these words the first significant challenge to the Newtonian con
ception of light was sounded. The princess’s reply: “Yes, but
what, sir, is sound, and what the ether?” If we would understand
light by analogy with sound, then we must first have some cer
tainty about the nature of the latter.
Knowledge concerning sound did not come quickly. Its spir
itual origins were enveloped in the nimbus of the Word, its
investigators included the masters of all times. Yet haltingly,
certainty arose concerning it, and by Euler’s time a full response
could be given to the princess. We, like he, must deviate from
our tale of light to explore for a while the history of sound. Once
we understand sound, we will appreciate the graceful analogy
Euler suggested for the nature of light.
Of Harmonies and the Vacuum
From the magnificent north spire of Chartres cathedral, the “Sun
Tower,” bells call the faithful to worship. Three hundred feet
below, in the archivolts of the Royal Portal, are other, silent
bells of stone. They appear as part of the twelfth century’s
portrayal of a learned Christian life, a scholarly and meditative
life that moved through the mastery of seven “liberal arts.” These
seven disciplines formed what William of Conches called “the
proper and only instrument of all philosophy.” Four of them
were devoted to the sciences. These four, the Quadrivium, “en
lightened the mind,” and one was the science of music.6 The
medieval patron of music shown in the sculptural depiction at
Chartres is not an early Christian composer but the pagan phi
losopher Pythagoras, who sits working while listening to the
ringing of bells and the tones of a lyre, instruments by which
he first discovered the presence of number in all things. It may
seem that he is listening to earthly tones alone, but the Chartres
masters knew well that he listened also to the inaudible concord
100
The Singing Flame: Light as Ethereal Wave
Ik .
Pythagoras at Chartres cathedral listening to the harmony
of the spheres.
101
CATCHING THE LIGHT
of heavenly music, whose divine harmonies echoed throughout
science from Pythagoras to Johannes Kepler. Sound, like light,
has a spiritual as well as a secular history, one whose trajectory
parallels that of its visual companion. According to the Finnish
folk epic, the Kalevala, the world was sung into existence.
Before sound could become the referent for a mechanical image
of light, as it was for Euler, it had to empty itself of its own
spiritual nature, it had to become a body and not the eternal
reverberation of the Word.
The transition to an entirely mechanical image of sound took
place in the seventeenth century. A characteristic figure in the
story was the learned Jesuit Athanasius Kircher, who combined
a religious temperament with a modem bent toward experimental
inquiry and debunking pagan superstitions. For example, on
the one hand, he understood Job’s line “When the morning stars
sang together.. .” as a reference to the Pythagorean harmony
of the heavenly spheres, a subject on which he expounded at
some length in many books.7 Yet while longing to hear the Lord’s
pure song, Kircher tinkered in his laboratory and museum of
sound at the Roman College devising novel acoustical instru
ments that he often used to disabuse the commoners of their
superstitions. For instance, on Pentecost, that religious feast set
aside to commemorate the miracle of the Holy Spirit in which
tongues of flame descended on the first Christian disciples,
Kircher and his companions dragged enormous speaking trum
pets to the top of St. Eustachius mountain. Acting the part of
an angelic choir, they sang litanies that could be heard in vil
lages up to five miles away. Over two thousand villagers followed
the call from above, bearing gifts to these brazen angels.
Yet Kircher was also an earnest investigator whose most im
portant scientific experiment was the first attempt to study the
propagation of sound through a vacuum. The very idea of the
vacuum, of space devoid of matter, was a hotly contested phil
osophical point in Kircher’s day. Aristotle had long before de-
102
The Singing Flame: Light as Ethereal Wave
dared that “Nature abhors a vacuum,” and his words carried
enormous weight. Notwithstanding the controversy, in 1650
Kircher inverted a column of mercury (as is commonly done in
mercury barometers) and placed a small bell in the empty space
above the mercury. Waving a lodestone—a magnet—by the
enclosed iron clapper, he caused the bell to ring, and hearing
it do so through the vacuum, he quite sensibly concluded that
air was not necessary for the transmission of sound. Others in
the Accademia del Cimento of Florence repeated the experiment
with similar results and conclusions. All this seemed to support
the view, held by the French philosopher and atomist Gassendi,
that sound was caused by the emission of a fine stream of in
visible particles that made their way from the source to the ear.
Not until ten years after Kircher’s experiment did Robert Boyle,
using a much-improved, von Guericke vacuum pump, succeed
in showing that the ringing of a bell could not, in fact, be heard
if carefully suspended in an evacuated glass container. Air was
necessary for the transmission of sound.
Interestingly, one could still see through the vacuum; it was
not dark, meaning that light, by contrast, did not require air for
its transport. Although light, unlike sound, did not require air
in order to travel, it remained entirely possible that a far more
subtle substance remained in the region, a substance that could
not be eliminated by Boyle’s vacuum pump.
The discovery that sound requires a material medium for its
travel, and allied discoveries such as the precise measurement
and prediction of the speed of sound, all succeeded in tying
sound to the earth. In the scientific imagination, sound became
a mechanical and material phenomenon. But the specific nature
of sound, as Bacon wrote, remained “but superficially ob
served,” and was “one of the subtlest pieces of nature.”8 How
is sound physically produced, and what is its specific nature
during its flight through the air?
Sing a few notes, and as you do so touch your larynx or put
103
CATCHING THE LIGHT
your fingers firmly into your ears. The vibrations you feel hint
strongly at the mechanism associated with the production of
sound. Look closely at a bowed violin string, or the end of a
mouth harp. Its blurred motion naturally suggests the association
of vibration with the generation of sound, an association already
made by Aristotle.9 Yet how do such vibrations correspond to
sound in detail?
The mathematical relationship between pitch and the tension
or the length of a string goes back to Pythagoras, but the as
sociation of it with a specific frequency of vibration originates
with Galileo at the end of the “First Day” of his Discourses
Concerning Two New Sciences of 1638. Should you happen to
scratch your fingernail on a blackboard, the squeal produced
not only sends a shiver up your spine, but you also feel a
vibration at your fingertip. Try it with a piece of hard chalk,
and you will notice that the vibrating chalk leaves a trail of
marks behind that are closer or farther apart depending on
whether the pitch is high or low. This observation was first
casually noted by Galileo when he ran a hard iron chisel on a
brass plate. Then he repetitively ran his chisel over the brass
plate under varying conditions. Although the screech must have
been ear-shattering, the results were exhilarating. With each
vibration the chisel left a mark in the plate. By counting the
marks per inch, Galileo could correlate the pitch produced with
the vibrations per second. He was even able to confirm the
ancient Pythagorean ratio for the musical interval of the fifth by
sounding his chisel-plate instrument, comparing it to a well-
tuned clavichord. Pitch and vibrational frequency were thereby
yoked together. Sound became a vibration that passed through
the material medium of air.
Remember the image of the ocean wave moving toward the
shore. The wave, rushing by, raises and lowers the water but
does not drive it toward the beach. All wave motion is similar
in this regard. A small, sometimes very small motion is enough
104
The Singing Flame: Light as Ethereal Wave
to send a “shape” rippling quickly away from its source. The
medium that carries that shape merely shakes in the tiniest way,
yet a thunderclap travels eleven miles in a single minute and
can be heard twenty miles away.
TWO HUNDRED YEARS after Galileo, during the performance of
the grand trios of Beethoven, and long after it was well estab
lished, the vibrational basis of sound received unexpected visual
confirmation. While listening to the string trio, the attention of
John Leconte, M.D., wandered to two fishtail gas burners near
the piano. The flame of one was seen to be pulsating in exact
synchrony with the music.10 Even the trills of the cello rippled
beautifully across the sheet of flame so that “A deaf man might
have seen the harmony.” The hidden vibrations of sound were
made visible by a “sensitive flame.”
The cold, dim glow of his solar sponge had suggested to
Galileo that light might be a body, the smallest in existence.
Perhaps he was wrong, perhaps one could replace his dull ma
terial conception of light with a view of it as tiny waves rippling
through an ether like the trills of a cello rippling across an open
flame. Perhaps light is a singing flame, a delicate vibration of
the luminiferous ether.
The Two Faces of Knowledge
Sound, shorn of its ancient associations with the creative Word
of God, presented itself to the clear-eyed intellectuals of the
eighteenth century as a wave of alternating compression and
rarefaction in the invisible medium of air. With this view of
sound established, it was natural to consider light to be of a
similar nature. The imagination, scientific or otherwise, quite
705
CATCHING THE LIGHT
naturally sees the unknown in familiar idioms: God in the image
of man, light in the image of sound.
Before going further, we should pause in order to realize the
incompleteness of the picture we have produced of sound. Pure,
unvarying vibration carries no meaning; this is a theorem in
mathematical physics. Bow a B-flat on your violin forever and
nothing has been conveyed, not music, not voice; no “signal”
has been transmitted. In fact your sense of hearing knows this
perfectly well and so will tend not to hear an unvarying back
ground sound whether it be a waterfall or the hum of a badly
ballasted fluorescent light. In order to speak we must modulate
the pure tones we produce, forming them into words. Meaning
arises as much through silence as through sound. Count Maurice
Maeterlinck knew this when he wrote, “Souls are weighed in
silence, as gold and silver are weighed in pure water, and the
words which we pronounce have no meaning except through the
silence in which they are bathed.”11
Vibration of the air is like the unformed lump of clay before
the hands of the sculptor have shaped it into the powerful cre
ation that can so move us. The human larynx acts like hands,
shaping the monotone of the physicist’s audio oscillator into the
diction of meaningful speech. Could the same be true of light?
Adelard of Bath in the twelfth century wrote that sight was due
to a “visible breath.”12 He thought that we “inhale” external
light (lumen) and “exhale” the light of meaning (lux), which
reminds one of the reciprocal action of outer and inner light
written of by the Greeks.
To see sound or light as vibration only is to reduce Michel
angelo’s David to marble dust. One may in a sense be correct
to do so, but in the process we lose the truth the statue embodies.
The untouched marble is all potential. Like the god Proteus, it
can assume any shape. Left to itself, it conveys nothing. To use
the language of light, it is as if lux must set upon lumen so that
speech, music, or a bird’s gentle song can all touch us, phys-
706
The Singing Flame: Light as Ethereal Wave
ically and psychically. If vibration is the body of articulate tone,
its spirit is reflected in its infinitely nuanced form. Sight like
hearing requires a modulated and crafted form of light for mean
ing. Stabilize images perfectly on the retina and they disappear.
This is a fact of sense psychology. We see only change, move
ment, life.
The SPECIFICS OF speech and hearing were separated off into
another discipline (the beginnings of sense physiology and psy
chology) during the eighteenth century, leaving the “bodily as
pects” to physical scientists. Among them Leonhard Euler
provided the first carefully supported vibrational theory of light
in his New Theory ofLight and Color of 1746. Luminous objects
“vibrated,” he wrote, and the ether carried those vibrations to
the eye as air carried sound to the ear. To advance his own
vibrational theory, he first needed to discredit Newton’s cor
puscular view, and this he did more systematically than anyone
before him, using many of the objections I raised at the close
of the previous chapter. Still, even Euler and his vibrational
theory could not explain all of light’s many manifestations, and
primary among them were the phenomena of diffraction. The
word “diffraction” was coined by the Jesuit Father Francesco
Maria Grimaldi of Bologna, Italy, following his detailed inves
tigation of light phenomena in 1665. Diffraction effects are sub
tle presences in our lives of which we are usually unaware.
Take from your wallet a credit card. It very likely possesses
a mirrored patch in which an image hovers. Occasional magazine
covers, promotional gimmicks, and every science museum dis
play similar three-dimensional, “holographic” images. All are
diffraction phenomena.
On a rainy night, look through your umbrella at a streetlight.
The multiple, colorfully transformed image of the light you see
is a diffraction phenomenon.
107
CATCHING THE LIGHT
The diffraction of light shows up _es en
circling the hand and coin. Here laser light is
used.
With your fingers very close together but not quite touching,
hold them up to your eye and look through the small gap toward
a source of light. The pattern of dark lines and shapes you see
is a diffraction effect.
Of the countless arrangements that give diffraction, the sim
plest is surely light passing by an opaque edge. At an unrelenting
boundary, free white light abruptly meets and succumbs to dark
ness. The offspring of this meeting are subtle but unmistakable.
Where before one noted only light and dark, variegated, parallel
bands of color rhythmically intrude into each region. Unnoticed
until the seventeenth century, the phenomenon enriches the
108
The Singing Flame: Light as Ethereal Wave
metaphor of light. Once again light struggles with darkness and
the meeting gives rise to color.
None of these phenomena could be explained with a corpus
cular view of light. Nor was Euler’s wave theory adequate to the
task, because it was still missing a key concept. If light was
like sound, then such diffraction effects should appear in sound
as well. At that time none was known. As it turns out, sound
does indeed show such effects but not so readily. For example,
sing a note in the bathroom. Slowly vary the pitch and notice
the change in intensity. When the frequency of vibration is
resonant with the room, the sound waves interact to produce a
clear increase in volume. Even this simple acoustic phenomenon
could not yet be understood. Something important was missing
from the understanding of waves whether acoustical or luminous.
The concept lacking was provided by the English scientist Young
and masterfully applied by his contemporary Fresnel in France.
THOMAS Young was a prodigy and polymath of amazing reach.13
Already reading by age two, in his early years Young was largely
self-taught in mathematics through the calculus, and the natural
sciences (including the making of telescopes and microscopes).
He displayed an early and abiding fascination with languages,
studying first Latin, Greek, French, Italian, and then going on
to Hebrew, Chaldean, Syriac, Samaritan, Arabic, Persian, Turk
ish, and Ethiopic. Young, in fact, successfully deciphered parts
of the famous Rosetta Stone independently of Champollion and
so helped to unravel the hieroglyphic script of ancient Egypt.
In London, Edinburgh, and Gottingen, he studied medicine,
receiving his M.D. in 1796. His interest in vision and his study
of the eye led him to the seminal suggestion in 1801 that color
vision occurs through the retina’s sensitivity to three principal
colors: red, yellow, and blue. Maxwell and Helmholtz later
modified and extended Young’s views on tricolor vision to the
109
CATCHING THE LIGHT
form now accepted today. Precocious, a lover of dancing and
dressage, Young was an elegant and brilliant intellect whose
importance for us lies in his revolutionary “principle of inter
ference.”
Deeply influenced by Euler, Young subscribed to a vibrational
theory of light and a universal luminiferous ether. In addition,
however, he advanced a principle that he maintained could make
diffraction phenomena understandable, which was a daring
claim. Just as intersecting water waves may enhance or cancel
each other, Young suggested that the undulations of the ether
can be strengthened or weakened to the point of extinction
through similar interference. We now are habituated to such
ideas in high-school science classes, but we should not under
estimate the apparent insanity of Young’s suggestion. According
to this principle, parts of a uniformly illuminated screen can be
made dark by the addition of more light. Light plus light equals
darkness? This is just what Young was proposing.
Henry Brougham spoke for many contemporaries when he
declared Young’s principle of interference to be “one of the most
incomprehensible suppositions that we remember to have met
with in the history of human hypotheses.”14 Together with their
rejection of the principle of interference, critics also rejected
the hypothesis that light is the sensory effect of vibrations in a
luminiferous ether. About the ether Lord Brougham wrote:
“From such a dull invention nothing can be expected.” Young,
misunderstood time and again, came to see himself as a modern
Cassandra who spoke nothing but truth but whom no one could
understand. Still, the specters of diffraction and polarization left
little real comfort for those who imagined light to be a stream
of tiny projectiles, and the time would not be long in coming
for the overthrow of their treasured hypothesis, and the vindi
cation of Young’s incomprehensible principle of superposition
(an incomprehensibility, by the way, that becomes all the greater
in quantum mechanics).
no
The Singing Flame: Light as Ethereal Wave
The TREATMENTS OF light by Newton, Descartes, Huygens,
Young, and Euler differed in particulars but shared an important
feature. They all stressed an analogical understanding of light.
That is, light was viewed as like something else: a bit of matter,
or the waves on a pond’s surface. However, Euler’s work also
developed a parallel approach, namely the formal and mathe
matical description of nature.
Euler modeled light on sound, and so participated in the well-
established tradition of proceeding by analogies with better
understood phenomena. His mathematical temperament, how
ever, led him to pursue an alternate description of phenomena
as well, one that created a more abstract imagination of nature
through the application of advanced mathematics including the
newly developed methods of calculus.
Euler was certainly not the first to apply mathematics to na
ture. In fact, like the history of light and sound, the application
of mathematics to nature has its own fascinating history that
mirrors the evolution of human consciousness. Going back in
time to ancient Babylonia, priest-astronomers stood atop their
ziggurats or stepped pyramids to observe the movements of sun,
moon, planets, and stars. On the basis of their observations,
they fashioned a purely arithmetic astronomy of amazing pre
cision.15 Significantly, they completely lacked any picture of the
cosmos other than that provided by religious mythology. In the
intervening centuries, the consciousness of the astronomer
changed profoundly. The planets and stars were no longer the
dwelling places of the gods, but distant masses arranged geo
metrically around the earth or sun. Mathematics also changed.
Its objects were not limited to the numbers and operations of
arithmetic, nor to the elements and proofs of Euclidian geometry,
but by the eighteenth century included radically new and seem
ingly unimaginable entities such as the infinitesimals of the
111
CATCHING THE LIGHT
calculus and, by the century’s close, the first approach to non-
Euclidean geometries. The perfection that Galileo drew down
from the heavens and placed into mathematics quietly unfolded
in sometimes disquieting ways.
Euler and his eighteenth-century contemporaries made great
strides in the application of modem mathematical analysis to
nature. Like the dance of sun and shadow about an object, Euler
and every physicist after him would use two languages to discuss
the nature of our world. One appeals to our sensual imagination,
the other to abstract reasoning. From his day to ours, the being
of light has been pursued using not only the nets of mechanical
analogy, but also the far more subtle and insubstantial web of
mathematics. Many raced to join Fontenelle, who watched na
ture’s performance from the wings of the opera’s stage where
pulleys and props, makeup and lighting, could be scrutinized
at close range. For a handful of others, however, the machinery
of nature was interesting, yes, but not nearly so beautiful as the
view of her forms and patterns to be seen through the lens of
mathematics.
Euler addressed not only, nor even primarily, a German prin
cess, but rather a different and more intimate audience in his
hundreds of papers and books. He wrote these works for a tiny
and elite community, one whose precarious existence depended
on the whim and generosity of despots such as Catherine II or
Frederick the Great. In the very same year in which his im
mensely popular Letters to a German Princess appeared, Euler
published a learned volume entitled Rigid Mechanics, which
sold all of twelve copies in its first two years. Amazingly, the
spiritual progeny of those twelve readers, among whom would
number Lagrange, Laplace, Poisson, Fourier, and Gauss, were
the first mathematical physicists. They would, in the end, shape
all of scientific culture. They, and not the German princess,
labored hard to nurture the tender seedling of modern science.
In the process, the very language of scientific communication,
112
The Singing Flame: Light as Ethereal Wave
especially with the invention of the calculus, became less and
less the common property of educated people. The widening
abyss between the mathematician’s intimidating discourse and
popular wit is probably nowhere better illustrated than in the
encounter between Euler and Denis Diderot, the brilliant ar
chitect of the great French Encyclopedia, at the court of Empress
Catherine II of Russia.
In 1773, the irreverent philosophe Diderot went to St. Pe
tersburg as the first librarian of Catherine’s newly formed library.
His boldness, eloquence, and atheism promised so to corrupt
the younger members of the court that the older courtiers im
posed upon the empress to still the tongue of the libelous French
man. Catherine reluctantly agreed, but wished it to be done
without her. It was, therefore, arranged that Diderot should meet
“a Russian philosopher, learned mathematician and distin
guished member of the Russian Academy, who was prepared to
prove to him the existence of God, algebraically, and before the
whole court.”16 (A project, by the way, continued by the math
ematical genius, Kurt Godel.) The occasion arrived, the whole
court was present, and the Russian philosopher (Euler!) ad
vanced gravely toward Diderot, and spoke in an earnest voice,
full of conviction, “Monsieur, (a + b") / z = x, therefore God ex
ists: answer that!” Diderot, though unmatched in his own in
tellectual domain, was not Euler’s equal in mathematics, and
so withdrew from the contest. Shortly thereafter Diderot thought
better of his position at the court of Catherine and returned to
France. Even the encyclopedist Diderot joined the ranks of the
German princess when confronted by the pronouncements of a
modem mathematical physicist.
Insubstantial mathematics and materialist world conceptions
that longed for mechanical models of unseen realities: these
were the features of the mental world of the eighteenth century.
Earlier light had inhabited other and very different imaginal
landscapes, but now its nature had of necessity to take up
113
CATCHING THE LIGHT
residence in the psychological space offered to it by the best
minds of the period. Is it ever otherwise? Does not our tenor of
mind shape the world we see?
Yet the human spirit is restless and nature forever compliant,
willing to answer as yet undreamed questions, capable of open
ing up vast new vistas, revealing still undisclosed parts of her
being. Puzzling features, unexplained phenomena of light re
mained and encouraged continued investigation. Under their
influence, a Frenchman devised a theory so successful that it
overthrew Newton’s image of light and established his own wave
theory as the supreme imagination of light.
Light at the Roadside
Diffraction was a recalcitrant fact that pointed incessantly toward
a wave theory. However, the puzzle of “polarized light” also
resisted comprehension and seemed initially to favor a corpus
cular view. And for advocates of the wave picture, above all
else, what in fact was the all-pervading ether? These issues
drove the research efforts of wave theorists to modify Euler’s
views such that the wave model of light became a formidable
and finally triumphant adversary of the corpuscular view.
Young’s principle of superposition, as bizarre as it might have
seemed to some, contained the essential missing ingredient. It
still waited for a gifted, mathematically minded spirit to cast
the principle into an elegant and powerful formalism. Ironically,
that figure was to be found at the roadside supervising the design
and construction of roads and bridges—Augustin Fresnel of the
Corps des Ponts el Chaussees.
IF ONE NEEDED to point to a single event that marked the crossing
to a new mathematical imagination of light, it would have to be
114
The Singing Flame: Light as Ethereal Wave
the March 1819 sitting of the Paris Academy of the Sciences.
Among the members of that committee were the greatest math
ematical physicists of the century, most of whom were staunch
corpuscularists. They were to judge the best scientific treatment
of the intractable problem of diffraction. Only two submissions
were received, one absurd and the other a treatment of such
power, scope, and mathematical sophistication that its author,
a relatively obscure provincial French engineer, and his theory
burst into the forefront of research into the nature of light.17
Alone and unaware of the work of Thomas Young, the civil
engineer Augustin Fresnel had busily pursued beautifully de
signed experiments on diffraction (using the local blacksmith
as his fine instrument maker) and developed the mathematical
foundations of a wave theory of light that could account for what
he saw. All the great mathematical physicists of his day, such
as Poisson, Biot, and Laplace, were supporters of a Newtonian
conception of light and therefore strongly antagonistic to the
wave theory of light. Fresnel had to submit his prize treatise to
these men to be judged, and over the years it would be against
them that the unknown engineer from the provinces leveled his
attacks in one scientific paper after another. By his sophisticated
use of the principle of interference, together with his masterful
application of the calculus, Fresnel made new and marvelous
predictions—many of which were confirmed. Yet his analysis
was not always complete.
In his contest paper, Fresnel included formal solutions to
diffraction problems so difficult that he could not solve them for
individual cases. In other words, he only analyzed the diffraction
problem abstractly but could make no concrete experimental
predictions. His brilliant adversary Poisson proceeded to solve
one of Fresnel’s intractable equations and then tried to use it
to show up an apparent absurdity in Fresnel’s theory. Poisson
pointed out that Fresnel’s own theory clearly predicted a spot
of light directly behind a small opaque obstacle. Shine light, for
115
CATCHING THE LIGHT
example, on a BB and directly behind it, in the middle of its
shadow, there should be a spot of light as bright as if no BB
were there! Absurd, declared Poisson. The experimentalist Ara
go, who was a friend and supporter of Fresnel, did the exper
iment and found a spot of just the kind predicted by Poisson
using Fresnel’s theory. Fresnel’s wave theory was vindicated.
Poisson inadvertently had driven home the last nail into the
coffin that carried the corpuscular theory of Newton.
Where in common life do such diffraction effects show up?
Although studied by Fresnel using special light sources and
instruments, a completely analogous phenomenon can be seen
many nights around the full moon! As gossamer clouds drift over
the moon, close around it appear rings of color: bluish near the
moon, becoming white further out, and ending in a reddish band.
Called the aureole or corona, this lovely sight is caused by the
diffraction of moonlight around water droplets or ice crystals in
just the same way Poisson’s spot is formed, but now repeated
for each of the cloud’s many millions of droplets.18 Instead of
only obscuring light, the droplets and crystals of the cloud can
put light where none should be if one imagines light as geo
metrical or corpuscular rays only. The aureole should remind
us of the impossible feat performed by light, its uncanny ability
to appear in the heart of the darkest shadow.
Fresnel and Young managed to explain another stubborn bit
of experimental data, namely the phenomena of polarization.
The most common instance of polarization is associated with
glare. Whether driving or boating, the reflection of sunlight from
surfaces can be annoying and even hazardous. Such glare can
be magically reduced by wearing so-called Polaroid sunglasses.
The secret of this innovation relies on the understanding of
polarization developed largely by Fresnel.
Pass light through a clear crystal of Icelandic spar and it
seems to become in some sense “oriented.” This feature can be
evidenced by passing the oriented light (or polarized light)
116
The Singing Flame: Light as Ethereal Wave
through an identical second crystal. It, too, is completely clear,
and yet for certain relative orientations of the two crystals, no
light passes through the second one. Corpuscular theorists had
suggested that particles of light, in passing through the first
crystal, were selected according to their shape. If the second
crystal was aligned in the same way, then light passed through,
otherwise not. It would be like trying to put square pegs into
square holes; they only fit if oriented certain ways. Polarization
phenomena stumped wave theorists because sound does not show
any polarization effects. If light was a wave just like sound, then
they should show identical effects. Fresnel suggested a solution.
If the vibration in the ether, which he viewed as light, were
“transverse” to the direction of propagation, then one could
accommodate an orientation.
Sound is a wave of compression and rarefaction in air. Speak
into one end of a pipe and the wave propagates through it at
the speed of sound in air. The wave motion is along the direction
of compression. If a domino chain could be equipped with
springs to reset each domino as it fell, the wave front of falling
and reviving dominoes would be a good image of sound.
Light is more like waves on a string or on water than sound
waves. A string can vibrate up and down, or left and right.
These correspond to two different linear polarizations of light
vibrations and so account for the “orientation” of light.19 The
elimination of glare, which is largely light polarized horizontally,
can be accomplished by orienting polarization filters to pass only
vertically polarized light.
One of the most surprising and beautiful polarization phe
nomena is that of polarization colors. To see them at work, simply
crush an old-fashioned piece of cellophane or a transparent
candy wrapper between two Polaroid filters (if you have an old
pair of Polaroid sunglasses, you can just pop out the lenses). A
glorious set of colored swatches suddenly springs into view and
their colors change as you rotate the Polaroid lenses! Close
117
CATCHING THE LIGHT
Rarefaction .
Compression -] | j 1=0
I
1 I r
1 i n can
lass
Depiction of a sound wave in air. Sound, moving to the right, is a series
of compressions and rarefactions, as shown above.
observation shows that with every change in the thickness of
the cellophane, a different color appears.
SEEING THESE COLORS, I cannot help being reminded of the
“Synchromy” paintings of the American artist Morgan Russell.
In his paintings, color itself becomes a language divorced from
figure. Under one, called Synchromy in Blue-Violet, the catalog
inscription reads, “Then God said, ‘Let there be light!’ and there
was light. ...” At the painting’s center and dominating it is a
yellow that Russell spoke of as symbolizing that light and the
first eye that saw it. To Mrs. Whitney he wrote, the “bursting
of the central spectrum in my picture. . . has surely a vague
analogy with what must have happened to the first visual organ.”
In the teens of this century, Russell was also occupied with the
construction of what he called “light boxes.” These were small
wooden boxes whose two longer sides were left open. Lamps
could be mounted within the boxes, and painted transparent
tissues were then mounted on the open sides. When lit from
within, Russell’s art literally glowed, radiating color into space
like the rays of sunlight through panels of stained glass. Dis
content within the confines of a box, light moves out and, where
possible, through its surround, calling into life the otherwise
mute, dark colors of tissue or glass.
On May 16, 1832, the great English astronomer William
118
The Singing Flame: Light as Ethereal Wave
Herschel sent to his physicist friend Whewell what he, too,
called “a box full of light.” It lacked the gaily painted sides
and artistic pretensions of Russell’s light boxes, but, Herschel
explained, the carton was “not an instantaneous light box but
one that works slowly.” It contained the scientific papers of
Augustin Fresnel.
Between 1820 and 1835, British scientists studied and ap
plied Fresnel’s ideas. As they did, it began to appear as if the
essential mathematical relationships that characterize light had
finally been captured and set down in Fresnel’s papers, all of
which could be contained in a modest-sized cardboard carton.
Galileo would gladly have stolen into the dark box of Fresnel,
perhaps with a candle, to discover in that solitary cell the math
ematical truth about light. And yet for all the magisterial suc
cesses of Fresnel’s theory, light was not at home or at peace in
Herschel’s box of light. It wanted out both scientifically and
spiritually. The electrical experiments of Faraday and the dy
namical theory of Maxwell tore open the box and scattered the
papers. Is light a wave? Maxwell and Faraday would answer
yes, but of what is it made, what kind of wave? And contemporary
with Young and Fresnel were the thinkers, poets, and artists of
Romanticism and American transcendentalism for whom the
essence of light could never be caught in equations or put into
a box. Their revision of light was far more angry and radical
than that of the scientists, and it, too, will form an integral part
of our history of light.
The Death of the Material Ether
Discontent not to see, we strain every instrument and intellectual
muscle to make out the nature of an invisible, everyday thing—
light. Not seeing it, we speculate. Perhaps, like sound, it is a
fleeting figure whose form speeds through space on a quivering
119
CATCHING THE LIGHT
medium. As sound is carried on the air, and ocean waves on
water, perhaps light is a wave borne by a supposed ether. Re
sisting direct observation, light was imagined by most physicists
throughout the nineteenth century to be the vibration of an
ethereal matter, but there were significant problems with such
a view.
Air and water we know, but what is the character of this
material ether? It seems hardly satisfactory to explain one in
visible thing, light, by another, the ether. What are its density,
texture, consistency, and other physical properties? From the
fact that we and the earth are hurtling through space, and so
through the ether, without apparent effect, it must be extremely
subtle. Yet the enormous velocity of light, 186,000 miles per
second, requires of the ether other, seemingly contradictory
properties. Again we can proceed by analogy, this time with
waves on a string or rope.
Stretch a long rope between two posts. Tap one end and you
will notice that the disturbance moves swiftly to the opposite
end, where it reflects back and returns to its origin, again to
reflect and strike out again. . . . Draw the rope more taut and
repeat the experiment. The form moves faster now. Slacken the
rope and it moves slower. Clearly the speed with which the
disturbance travels depends on the tension in the rope. We
sense the rightness of this relationship from our own experience.
Snap your fingers. If you snap them with more force, your fingers
move more quickly. Similarly the rope’s tension is that force
which acts to restore the rope to its original position. The larger
the restoring force, the quicker the rope snaps back into place,
and the faster the disturbance scoots along.
Speed increases with tension. However, a second feature of
the rope also affects speed, and that is its mass. It is harder to
snap back a heavy rope than a light one under the same tension.
Not surprisingly, therefore, as the mass goes up (actually the
mass per unit length) the speed goes down. On a piano, guitar,
or violin, instrument makers take advantage of this relationship
120
The Singing Flame: Light as Ethereal Wave
using heavy strings for the bass notes and thin ones for the high
notes. Careful experimentation and theoretical analysis show
that the speed of the disturbance is given by the formula
v = ''•/(!'Im). The speed v is equal to the square root of the
tension divided by the mass. Increase the tension and the dis
turbance speeds up, increase the mass and it slows down, in
exact accord with the above equation.
By analogy we can ask of the ether, if light is a wave within
this elusive medium, then it, too, must have a source of tension
and a mass density. What are they? The answer puts flesh onto
an otherwise vague notion, providing us with a concrete material
image of the ether and so, too, of light. Everywhere light reaches,
the ether exists also, sustaining the vibrations that are light. It
is an elastic solid so resilient that a wave traveling in it can
circle the earth over seven times in a single second. Yet that
same earth must be capable of passing through the ether at
mosphere of the universe unimpeded.
Already in 1746 Euler had made explicit estimates of the
physical properties of the ether based on the comparison of
the propagation velocities of sound and light. On these grounds
he argued that the density of the ether must be at least a hundred
million times less than that of air, and its elasticity a thousand
times greater, to account for the extreme rapidity of light. Thus
the ether was to have a resiliency greater than steel, but was
also millions of times more subtle than air.
A century later, Sir George Stokes suggested a model of the
ether at the close of a technical paper.20 Glue and water together
form a stiff jelly that can on the one hand act like a solid for
rapid vibrations, yet will allow the easy passage of a slowly
moving body. Perhaps the ether is of a similar substance. The
exceedingly rapid, small vibrations of light race through the
medium at 186,000 miles per second, while the ponderous mo
tions of the planets creep along their orbital paths at a mere
10,000 miles per hour, plowing the ether aside as they pass.
Following Fresnel’s lead, able French and English mathe-
121
CATCHING THE LIGHT
maticians created dynamical models of the ether like Stokes’s,
that is to say, they elaborated detailed pictures of the motions
and interactions of “ether molecules” required by the known
properties of light. From light’s incredible speed, requirements
had been placed on the ether’s elasticity; from the motions of
the planets and comets, limitations were put on the ether’s
density; from the facts of polarization, the structure of ether
interactions could be suggested. Following on many experimen
tal results, the work of these great mathematical scientists ap
peared to be converging on an elaborate material and mechanical
model of the luminiferous ether that could account for every
single experimental phenomenon.
Their picture of light possessed only one absolutely funda
mental failing, namely that the ether was material. Most sci
entists of the nineteenth century were locked into a mode of
imagination that was rigorously materialistic. No matter how
subtle and unusual its properties, if the ether was something,
then that thing must possess a substantial nature of some kind.
If it did not, then it simply could not exist. For them the opposite
of matter was spirit, and to admit the immateriality of light was
to open the floodgates of speculative natural theology.
In the middle of the eighteenth century, Bishop Berkeley and
Chevalier Ramsay had each advanced views of the ether as
essentially spiritual, reaching back to the prisca sapientia (pri
mal knowledge) tradition of Egypt, Greece, Persia, and the
Hermetic writings already familiar to us from the previous chap
ter. Ramsay called the ether “the body of the Great Oromazes
[i.e., Ahura Mazda], whose soul is truth. ... He diffuses himself
everywhere.”21 Trinitarians made the analogy between, or sug
gested the identification of, the Holy Ghost and the Universal
Ether. After 1875, such sentiments received renewed consid
eration when a “crisis of faith” struck many late-Victorian sci
entists, who then often sought to reconcile science and religion
by means of spiritualism.22
122
The Singing Flame: Light as Ethereal Wave
To deny a material nature to light or the ether was, therefore,
to court a reversion to prescientific, spiritual imaginations of
light, a view that the reigning scientific minds of the age
abhorred. We should not underestimate the significance of re
ligious inclinations (or disinclinations) in the conduct of science,
especially with regard to the advocacy of poorly supported hy
potheses. In this context we can understand the insistence of
most scientists of the period that light and the ether must at
root be material. The detailed dynamical models of the period
offered them, literally, a concrete understanding of light that
satisfied their metaphysical bias, and moreover could be used
to make predictions as to how the ether would show itself, even
if only in extremely subtle ways.
The ether was accordingly systematically searched for in nu
merous laboratories and observatories both in Europe and among
the fledgling American scientific community. Every suggested
experiment was done and redone, and even today new experi
ments in search of a material ether continue to be performed
with magisterial precision. By 1900, the indications were grow
ing clear; by 1990, they are undeniable. The material ether does
not exist. It was a hypothetical fiction bom of a materialistic
imagination.
Light is not a luminous ripple on the material substrate of
the ether. Still, although innumerable experiments deny the
ether, an equal number seem to affirm the wavelike character
of light. If we take both seriously and suppose light to be, in
some sense, a wave, then what is it that is waving? In the cases
of water waves, sound waves, vibrating strings . . . something is
always waving. The figure of sound is borne by the air. What
bears the fleeting figure we call light? One thing has become
certain, whatever it is, it is not material!
123
CHAPTER 6
Radiant Fields:
Seeing by the Light
of Electricity
For the invisible things of Him from the cre
ation of the world are clearly seen, being
understood by the things that are made, even
his eternal power and Godhead.
St. Paul, Romans 1:20
Well before the bankruptcy of material imaginations of light
became evident, another and more productive view dawned in
unexpected quarters.1 Far from the ancient and learned halls of
Oxford and Cambridge, and therefore also far from the pompous
tyranny of academic tradition and rivalry, a failing blacksmith
and his gentle wife had a son, the third of four children. The
year was 1791, the surroundings were poor and inauspicious—
a London slum. The ailing father eked out a spare existence as
a smith, his family living above the coach house. The daily fare
was modest. In 1801, when food prices were high, the young
124
Radiant Fields: Seeing by the Light of Electricity
son would be given a single loaf of bread to supply his meals
for the week. Mean outer circumstances, however, can some
times belie remarkable destinies.
From these unlikely beginnings, Michael Faraday, the great
est experimental scientist of all time, arose. By the time of his
death at age seventy-six, he had been elected an honorary fellow
to the scientific societies in Paris, Brussels, St. Petersburg,
Florence, Copenhagen, Stockholm, Berlin, Munich, Vienna,
etc., and been offered knighthood as well as the presidency of
both the Royal Society and the Royal Institution of London,
honors that he steadfastly declined. When his much-loved col
league pressed the aged Faraday to accept them, he replied,
“Tyndall, I must remain plain Michael Faraday to the last.”2
From this “plain” soul would come a revolutionary conception
of light, one not fettered to the materialistic worldview of his
time. Historians have traced to this simple son of a blacksmith
the seminal imaginative ideas of modern field theory. I see in
him an uncommon figure who cared too deeply about truth to
be beguiled by the fashionable models of the day, and who knew
that nature was the constant arbiter of ideas. How did it come
about that plain Faraday saw so deeply into the heart of nature?
What was his character, his education, and the special gift that
made him the pivotal individual of the nineteenth century for
our biography of light?
To judge from his childhood and adolescence, Faraday’s suc
cessful career of scientific discovery seemed improbable in the
extreme. He was a Cinderella figure rising from the ashes of his
father’s smithy to become a prince of natural philosophy. Mi
chael Faraday’s formal education was halted after he had learned
the bare rudiments of writing, reading, and arithmetic. He was
withdrawn from school by his tenderhearted mother after being
beaten for mispronouncing the letter “R.” Henceforth he grew
up at home and on the streets of London.
Although the family was poor in things material, they led a
125
CATCHING THE LIGHT
rich religious life. The latter proved to be an especially formative
and lasting influence on Faraday. From those early years he
maintained a quiet and unswerving religious faith as practiced
within the small Christian sect into which he was bom, the
Sandemanian Church. Throughout his life, here was Faraday’s
spiritual home. Faraday saw no disagreement between his work
and that of his God, for, paraphrasing a favorite Sandemanian
passage from St. Paul, Faraday wrote, “even in earthly matters
I believe that the invisible things of Him from the creation of
the world are clearly seen.”3 The passions of objective scientific
research and intimate religious sentiment alike moved through
the breast of Michael Faraday. In subtle yet significant ways
each played into the other, so that his agnostic colleague Tyndall
could write of his mentor: “The contemplation of Nature, and
his own relationship to her, produced in Faraday a kind of
spiritual exaltation which makes itself manifest here. His reli
gious feeling and his philosophy [science] could not be kept
apart; there was an habitual overflow of one into the other.”4
One of the “invisible things” Faraday investigated was light.
Following St. Paul’s precept, he sought knowledge of the invis
ible by understanding “the things that are made,” that is, by
constant, open-minded observation and experimentation. Noth
ing was so important to him as a truly revealing phenomenon.
Faraday labored with unflagging energy and genius toward the
archetypal phenomena of electromagnetism. Unknown to any
before Faraday, once found, these phenomena suggested a to
tally new imagination of light.
Electric Waves
At the age of thirteen, Michael Faraday launched upon his career
as an errand boy for Mr. G. Riebau of 2 Blandford Street, a
French immigrant bookseller and bookbinder. Among the books
126
Radiant Fields: Seeing by the Light of Electricity
of the shop and under the kind encouraging eye of Mr. Riebau,
Faraday’s curiosity flourished. Within a year Faraday had signed
on as an apprentice bookbinder, a trade he mastered well over
the next seven years. But his education at the shop went far
beyond matters of bookbinding. While his hands received their
training in dexterity, for which he was admired throughout his
life, his spare moments were spent in reading. “There were
plenty of books there, and I read them,” wrote Faraday simply.
Riebau also recollected how Faraday would copy drawings of
electrical machines and the like into his own volumes, and how
many mornings he would take an early walk, “Visiting always
some Works of Art or searching for some Mineral or Vegetable
curiosity—Holloway water Works, Highgate Archway. ..”
Faraday’s energies received a more specific focus when, in
1810, at age nineteen, he joined the City Philosophical Society,
which had been founded but two years earlier under the guiding
spirit of Mr. John Tatum, whose house, library, scientific
equipment, and abilities as a lecturer supported the Society.
Every Wednesday evening Tatum or another club member
would speak to a scientific theme that concerned him. Faraday’s
practice of writing out and binding a meticulous summary of
the lectures, including drawings, in the end stood him in good
stead.
Mr. Riebau occasionally showed Faraday’s four bound quarto
volumes of notes to friends and customers, including a Mr.
Dance. On seeing them, the latter arranged for the young Far
aday to gain admission to the lectures of Sir Humphrey Davy
of the Royal Institution, the most famous scientist in England
and a charismatic speaker. The effect of the lectures was to
convert Faraday into a disciple. He proceeded to write up Davy’s
four lectures with the same care he had Tatum’s.
Shortly thereafter, in the autumn of 1812, Faraday’s appren
ticeship with Monsieur Riebau was at an end, and before long
Faraday was in despair. He longed to be a scientist but saw no
127
CATCHING THE LIGHT
way of fulfilling his aspirations. In desperation, he wrote to Sir
Joseph Banks, president of the Royal Society, to apply for a
scientific situation, no matter how menial. Banks did not even
reply to Faraday’s repeated inquiries. Yet fate worked in re
markable ways. Humphrey Davy suffered injury to his eyes
through a laboratory explosion, and probably through the rec
ommendation of Mr. Dance, Faraday came to serve as Davy’s
amanuensis for a few days. In December, Faraday wrote to beg
a position of Davy, including with his entreaty his bound volume
of meticulous notes of Davy’s lectures. Though flattered, Davy
(who was also of low birth) could offer him nothing. But again
circumstances chanced to move in Faraday’s favor. Davy’s as
sistant became involved in a brawl and was dismissed. That
very evening, as Faraday was undressing in his bedroom, a
thundering knock startled him. A footman descended from a
carriage with a note from Sir Humphrey Davy requesting Faraday
to call on him at the Royal Institution the next morning. At the
interview, Davy offered him the menial position of assistant, a
guinea a week, two garret rooms atop the Institution, fuel, and
candles. Faraday asked only that he be provided with aprons
and, most important, be granted liberty to use the apparatus of
the Institution. Davy and the Institution agreed and Faraday’s
apprenticeship in science with England’s greatest chemist was
begun March 1, 1813.
Faraday’s dexterity and energy quickly showed, and his ser
vices were sought on many fronts in the research and lecturing
activities of the Institution, where Faraday played the role of
assistant to all. During his first three years, he accompanied
Davy on his travels to European laboratories as his valet and
assistant, meeting in the process the great scientists of France
and Italy. It was not long before Faraday’s own mettle began to
show through a series of modest chemical publications, and then
the discovery of benzene. But our specific concern with Faraday
is his central role in providing a new imaginal form for the
128
Radiant Fields: Seeing by the Light of Electricity
One of the many beautiful Chladni plate patterns. Sand accumulates
on those parts of the vibrating plate that are still.
nature of light, one that ultimately broke free from the materi
alistic constraints of his contemporaries.
Perhaps not coincidentally, the period of Faraday’s interest
in the nature of light occurred simultaneously with his research
into the nature of sound, music, and musical instruments (1828—
30), an interest that dated back to his days in Riebau’s shop,
where singing was a favorite pastime. Of special interest to
Faraday were the so-called Chladni figures.
Some years earlier, in 1785, Chladni had discovered that
beautiful patterns could be made to appear on thin metal plates
when sand was sprinkled onto their surface, and then the plate
set to sounding by bowing an edge like a stringed instrument.
Faraday studied these lovely phenomena and demonstrated them
to his audiences at the Royal Institution.
In just these years Fresnel’s important articles on the un
129
CATCHING THE LIGHT
dulatory theory of light appeared in English in a popular form
whose mathematical level was sufficiently low that Faraday could
read the articles with understanding and pleasure, admiring
Fresnel’s clarity and precision. In a similar vein, Herschel, in
his 1830 treatise on natural philosophy, emphasized again and
again the similarities between sound and light, making full use
of the Chladni figures as an aid to the imagination. The analogous
character of light and sound initially argued by Euler and others
in the previous century was now very much in the air.
Faraday’s religious convictions led him to believe deeply and
steadfastly in the unity of nature, in the idea that what appeared
on the surface to be disparate was at its core one. Perhaps
vibration was such a unifying idea under which not only sound
and light could be joined, but electrical effects also. To this
end he energetically pursued a series of investigations in search
of an undulating electrical wave of some kind. His discovery of
just such an effect, one of the most important in his life, goes
under the name of “electromagnetic induction.”
As a name only, it is unfamiliar to most people outside physics
and electrotechnology, but in its applied form it is universally
familiar. Within almost every appliance and atop millions of
telephone poles are transformers whose lineage goes back to
that moment in August of 1831 when Faraday discovered the
principle behind these devices. The impact of Faraday’s dis
covery both practically and for our understanding of light is so
great that we must pause to examine it. For in attempting to
understand what appears to be a purely electrical effect, the
foundations were inadvertently set for a new understanding of
light. The effect appears in two important ways.
For his first experiment, Faraday wound two separate coils
of wire around a torus (or doughnut) of iron. He connected one
to a sensitive meter whose needle deflections would show when
a small current was passing through the coil. To the second coil
he connected a battery through a switch. With the switch closed,
130
Radiant Fields: Seeing by the Light of Electricity
i liihunomrlrr
Batten-
Faraday’s archetypal experiment concerning electromagnetic induction.
When the switch on the left is closed, a current surges through the left
coil, which induces a current to flow in the circuit on the right, which
deflects the galvanometer needle.
current flowed through the left-hand coil: with it open, there
was no current. While watching the meter, Faraday opened and
closed the switch. Neither with the switch open nor with the
switch closed did the meter show a deflection. That is, there
was no current in either case in the coil on the right. However,
at just the moments of closure and opening, in that instant, there
was a sharp deflection of the needle. Moreover, the needle
deflected in opposite directions depending on whether the switch
was being opened or closed. This showed that a brief surge of
current appeared in the right coil when one started or stopped
the current flow in the coil on the left. In other words, current
was induced in the right coil only when there was a changing
current in the coil on the left. Change in one circuit induced
change in another. By contrast, when the current was unchang
ing, there was no effect on nearby coils.
One final caveat concerning the nature of the connection
between coils. While the above “induction” phenomenon is en
hanced by the iron torus core linking the two coils, the torus is
131
CATCHING THE LIGHT
by no means essential to the experiment. No electrical current
passes through it. Remove the iron core and a small but still
detectable induction effect is seen. A change in the electrical
state in one coil induces a corresponding change in the other.
To understand electromagnetic induction, Faraday suggested
that a “wave of electricity” is caused by sudden changes in the
current through the first or “primary” circuit. This electrical
wave travels through space and induces a similar disturbance
in the nearby “secondary” coil of wire. Hence the deflection of
the needle.
Clap your hands and the disturbance echoes through space,
strike a match and darkness progressively gives way to flickering
light. Likewise, throw a switch on a powerful electrical source
and the surge of current that follows causes an invisible electrical
disturbance to propagate through space that can be caught in a
web of circuitry some distance away. When pushed to its full
potential, Faraday’s first experiment metamorphoses into the
radio communication of a space probe sending its pictures back
from Jupiter, Saturn, and Uranus hundreds of millions of miles
away. Create an electrical disturbance in the on-board antenna
of the spacecraft, and three hours later a tiny but exactly cor
responding disturbance arises in a sensitive earth-based an
tenna. The investigations and ideas forged by this blacksmith’s
son continue to propagate new technologies and new knowledge.
Yet, just what is this “electrical wave” that it can connect
distant circuits without a visible material connection of any kind?
Michael Faraday’s efforts to answer this question by further
experimentation and cautious speculation would occupy him for
nearly thirty years. The suggestion he finally put forward ulti
mately transformed our scientific image of light completely.
The second and related experiment, performed hot on the
heels of the first, replaced one coil of wire (the one connected
to the battery) by a magnet. Faraday discovered that by moving
a magnet into and out of a coil of wire, a small current could
132
Radiant Fields: Seeing by the Light of Electricity
be set to flow. If the magnet was stationary relative to the coil,
no current; when it moved, current arose. This is one of the
great archetypal phenomena of electromagnetism. Its practical
and theoretical significance are hard to exaggerate. The pro
duction of electrical power at every generating station is nothing
more than this action, in one form or another. Not only did
industry make good use of Faraday’s discovery, but so would
pure science down to Albert Einstein, who used Faraday’s ar
chetype at the turn of the twentieth century when putting forward
his “principle of relativity.”
Natural Truth and the Shadow
of Speculation
The view I am so bold as to put forth considers, therefore,
radiation as a high species of vibration in the lines of
force. . . .
Michael Faraday
The worldview held by most early-nineteenth-century scientists
was straightforward. The universe was filled with material ob
jects, between which stretched several elusive but material
ethers whose motions conveyed the forces of gravity, light, heat,
electricity, and magnetism from one object to another. Every
where there was substance: ponderable, massive matter. Never
before or since has materialism attained a more contented and
comprehensive grip on the world. The radiant eyes of the Egyp
tian god Ra that once lit a civilization were tightly shut.
Into this arena stepped the slight figure of Michael Faraday
with views that would overturn the fundamental conceptions of
his own beloved scientific community. Gentle, deferential, and
religious man that he was, Faraday hardly cut the figure of a
revolutionary. He was, however, a persistent experimenter and
133
CATCHING THE LIGHT
thinker. To resist the tide of current scientific opinion, which
favored Fresnel’s vibrational ether theory of light, would be
heresy. Yet in two addresses to audiences at the Royal Insti
tution, the first in 1844 and the second in 1846, Faraday dared
to speak, and so carried science across a divide in the Western
imagination of light.
From his earliest researches in science, Faraday possessed
a love for truths gotten through direct experience, and conse
quently he held an abiding distrust of speculative theories, such
as the molecular ether theories popular in his day. Ever aware
of the danger of such speculation in his own research, Faraday
penned the following entry in his diary for December 19, 1833,
shortly after his discovery of electromagnetic induction: “I must
keep my researches really Experimental and not let them deserve
any where the character of hypothetical imaginations."3 The key
word here is “hypothetical.” Time and again new ideas and
imaginations must be grounded in honest experimental fact,
otherwise fantasy takes the place of cautious creative thinking.
In his discourses of 1844 and 1846, Faraday very cautiously
advanced his own views on the ultimate nature of matter, elec
tricity, and light. His first attack was on the naive corpuscular
view of atomic matter then current.6 In place of atoms thought
of as small “blobs” of impenetrable matter, Faraday suggested
that atoms were purely “centers of forces.” Since we know of
objects by their attributes, and these are conveyed by forces
only, we should not add the unnecessary concept of a material
source to the forces.
Buddhist philosophers might put it this way. Can one think
of an “attribute-bearer” apart from all attributes, such as size,
shape, position, etc.? As an example, consider a penny. It is
hard, round, thin, copper-colored, nearly an inch across, and
carries embossed images on both sides. One by one, eliminate
these attributes of the penny. Try it. First erase the images and
the penny becomes a slug. Go on to imagine it as having no
134
Radiant Fields: Seeing by the Light of Electricity
definite color, size, or shape. Can you do it? I cannot. If one
cannot think of an object without its attributes, then why retain
the notion of attribute-bearer at all? Similarly, if we only know
of the world only through various forces, why posit the existence
of force-bearers?
Since all properties such as hardness, color, and so on were
understood to be the result of forces, then atoms (which Faraday
did see as necessary) are simply the geometrical foci or centers
of those forces. Substantial atoms, thought of as tiny, dense bits
of matter, disappear completely, but as they do so, the atmo
sphere of forces that was thought to surround them becomes all-
important. Force, not substance, is the true being of the world,
and it, not the ether, reaches from one end of the universe to
the other. Forces may constellate themselves in myriad ways
and patterns to create chemical species, a visual, tactile, and
fully sensual, “corporeal” world, but at root everything is still
force, not ponderable substance. Faraday’s ontology was very
different from his colleagues’.
Recall Descartes’s image of a cosmos filled with matter, a
plenum that was like a whirling stream of water in which planets
moved like floating bits of chaff, twigs, and leaves. Faraday
suggested that Descartes’s stream was actually a sea of pure
forces. Points of matter—atoms—were only the starlike inter
sections of myriad raying lines of force that spread out from
these centers to weave their way through the universe.
In his second address two years later, Faraday, apparently
inadvertently, took his ideas an important step further.7 Ac-
cording to the usual account, on April 10, 1846, Faraday and
his collaborator Wheatstone were outside the Royal Institution’s
lecture hall waiting for the clock to signal the beginning of their
lecture. It was Faraday’s understanding that Wheatstone would
be giving the lecture, but just moments before the evening’s
performance was to begin, the intensely shy Wheatstone bolted
down the stairs. Caught unprepared, Faraday nevertheless
135
CATCHING THE LIGHT
jumped into the breach, speaking first on the scheduled topic
of Wheatstone’s “Electro-magnetic Chronoscope,” but then fol
lowing it impromptu with his famous “Thoughts on Ray-
Vibrations.”8 Short of material on which to speak, Faraday
“threw out as matter for speculation, the vague impressions of
my mind.”
His vague speculations were, while new to others, not new
to him. They had been maturing during the previous fifteen
years of his investigations. Already in the earliest of his reports
on Experimental Researches in Electricity from November of
1831, Faraday had used the concept of “lines of magnetic force,”
the idea so central to his 1844 and 1846 discourses. If you
sprinkle iron filings on and around a magnet, they align them
selves in definite patterns. Faraday developed this picture into
a powerful tool of the imagination. From his dogged investiga
tions, and by careful reasoning, it seemed possible to him that
one could conceive of the entire universe as threaded through
with an infinity of such “lines of force.” Reading his papers,
one can watch the birth of a new scientific concept of the most
extraordinary significance as it emerges from Faraday’s exper
iments and imagination, that of the “field.”
How is the earth bound to the sun? Perform the following
thought experiment, Faraday suggests. Imagine a space with the
sun alone. It resides solitary in its universe. Now, in an instant,
place the earth ninety-three million miles away. Is it reasonable,
he asks, to suppose that the distant presence of the earth causes
there to rise up suddenly in the sun a power of attraction? Where
would such a power come from? Would not such a conception
violate our sense that forces must, in some sense, be conserved?
Far better that the isolated sun, even before the earth’s ap
pearance, should have spread its influence as it does its light,
even if no objects are there to feel it. Does the sun only shine
when eyes are there to see? So, too, gravitational lines of force
thread their way through space even before the earth appears.
136
Radiant Fields: Seeing by the Light of Electricity
Visualization of magnetic field lines with iron
filings.
The earth, then, reacts to a local, not a distant force, that is,
to the force of the field at its own location.
In an era when all space was permeated by ethers, Faraday’s
suggestion could be interpreted to mean that in the ether were
stresses and strains similar to those in the trusses of a bridge.
The field idea, in this view, was but a superficial account of
what was still, at root, a mechanical and material interaction.
Faraday, however, chose to reach beyond the materialist imag
inations of his contemporaries. Originally, he had considered
137
CATCHING THE LIGHT
the lines of force (his term for “fields”) as nothing more than
useful fictions, but the more he thought about them, the greater
their reality appeared, until they became for him more real than
the atoms of matter that were thought to be their sources, or the
ether that supported them.9 The results of his investigations into
electromagnetic induction, moreover, convinced him of the ex
istence of electric waves, that is, of the vibrations of such lines
of force. He was, in fact, well prepared for Wheatstone’s flight.
The revolutionary step taken in his 1846 lecture to Wheat
stone’s audience was his suggestion that the vibrations called
light, so elegantly described by Fresnel and others, were not
vibrations in an ether at all, but rather the vibrations of physical
lines of force. Faraday’s theory “endeavors to dismiss the aether,
but not the vibrations.”10
Vibrations were essential, the ether was not. Not content with
transforming atoms into centers of force alone, Faraday also
radically altered our conception of space and light. Lao-tzu’s
pot was emptied of its ethereal substances and filled with force,
a force whose movements were light. Descartes would have
stormed the Royal Institution if he could have gotten passage
across Lethe and the English Channel, and even Faraday’s most
sympathetic reviewers thought he had, for once, gone too far.
To make something as insubstantial as lines of force the onto
logical basis of the world seemed preposterous to the materi
alistic imagination, and yet herein lay the pregnant seeds of
field theory as developed by physicists less than a century later.
In the clear, if self-effacing, voice of the blacksmith’s son, we
hear the first sounds of a vast new vision of light.11
He knew his thoughts reached beyond the simple facts of the
science he loved so much, and yet just those facts seemed to
call out for an immaterial interpretation of light. Ever the gra
cious and careful philosopher, he advanced his view circum
spectly, even apologetically, closing his remarks to those in
attendance with the words “I think it likely that I have made
138
Radiant Fields: Seeing by the Light of Electricity
many mistakes in the preceding pages, for even to myself, my
ideas on this point appear only as the shadow of speculation. . . .
He who labours in experimental inquiries knows how numerous
these are, and how often their apparent fitness and beauty vanish
before the progress and development of real natural truth.”12
His did not vanish.
Specters of Philosophy
But the splendour that supplies
Strength and vigour to the skies,
And the universe controls,
Shunneth dark and ruined souls,
He who once hath seen this light
Will not call the sunbeam bright.
Boethius,
from “The True Light”13
In the year A.D. 524, a lone philosopher-sage of Theodoric’s
court in Ravenna awaited his torture and death at the whim of
his suspicious lord. He had used his years of study and medi
tation on the great Greek philosophers, as well as on the fruits
of the new Christian faith, in service of the temperamental Theo-
doric. The solace of his studies now deserted him. Lost in self-
pity, he was startled by a vision of a woman. Her stature was
at once huge and common. Through tear-veiled eyes he saw her
clear eye, quick sight, and grave countenance. Her gown, woven
by her own hand, bore images of wisdom and the universe, as
well as the signs of wounds made by those lusting for a snatch
of knowledge. The divine Sophia, Philosophy, had come into
the dark and solitary chamber to console a condemned man,
the philosopher Boethius.
Inquiring, she discovered the real cause of Boethius’ grief,
139
CATCHING THE LIGHT
his forgetfulness of his true self. Her balm, then, was to “dissolve
this cloud with gentle fomentations that thou mayest behold the
splendour of true light.” In these, his final days, Boethius wrote
his greatest work, The Consolation of Philosophy, so influential
on the centuries to follow. When the book was completed, Bo
ethius was hanged by a rope around his forehead and beaten to
death with clubs. The last utterance of Philosophy in the Con
solation reads: “Great is the necessity of righteousness laid upon
you if ye will not hide it from yourselves, seeing that all your
actions are done before the eyes of a Judge who seeth all things.”
These words were prophetic, for legend has it that after the
execution of Boethius and his supporter Symmachus, a great
fish’s head was served to Theodoric. In his guilt-ridden mind,
it took on the features of those whose death he had caused. He
wept and anguished, inconsolable until his own death shortly
thereafter.
Six hundred years later, the Chartres master Alan of Lille,
in a period of mental anguish, was similarly visited.14 The com
forter called herself Natura, the being of Nature. Her every
feature and attribute reflected the universe, her every word
gauged to reawaken in Alan the knowledge of her that he and
all the world seemed to have lost. Alan called the song she sang
the Plaint ofNature, a song of violations unheard by humankind.
Six centuries later, still at the hands of Goethe, Faust in his
high-vaulted Gothic study was lost in despair.15 Having mastered
every discipline through a long life of study, he felt himself a
fool no wiser than before. In angry desolation, he called out to
unseen forces: “Spirits that hover near to me, give me an answer
if you hear my voice!” His eye fell on the signs of the Macrocosm
and Earth Spirit.
Faust then invoked, even commanded, the Earth Spirit to
appear. Unbidden, Natura had comforted and taught Alan of
Lille, and Sophia before her had consoled the condemned Bo
ethius. Faust, by contrast, bent the Spirit of the Earth to his
140
Radiant Fields: Seeing by the Light of Electricity
Faust in His Study by Rembrandt, 1652.
own needs and will. “Obey! Obey,” he cried, “although my life
should be the price!” When the Earth Spirit did appear, Faust
reeled before its awesome shape of flame, but even so dared to
declare himself a spirit kindred to this mightiest spirit of nature.
Faust’s arrogance was bluntly rejected by the Earth Spirit with
the words, “Your peer is the spirit you can comprehend; mine
you are not!” The spirit Faust could comprehend then entered
his chambers; the plodding pedant Wagner, dressed in night-
141
CATCHING THE LIGHT
gown, nightcap, and holding a flickering candle, a sorry coun
terpoise to the flame-formed Earth Spirit of a moment before.
The Earth Spirit seems to have grown ever more remote, more
difficult of access in these later days. The only way of ap
proaching her, that of scholarship a la Wagner, most often leads
not to her but only to a shadowy, even demonic reflection of her
true self. Michael Faraday’s great successor, James Clerk Max
well, fought a Faustian battle as a young man, and confronted
a fearful counterimage of divine Sophia. The son of a distin
guished family in Edinburgh, Scotland, and a very promising
young mathematician, Maxwell reported his own vision in Gothic
chambers when a third-year student at Trinity College, Cam
bridge, in 1852.16
The midnight bells had struck and Maxwell left his “con
founded hydrostatics” to prepare for bed, asking himself doubt
fully, “with voice unsteady, If of all the stuff I read, I / Ever
made the slightest use.” He looks at his future, a life likely to
be full of outer accomplishment and lofty station, and knows it
will be but the fruit of reason put under “the control of worldly
pride.” Across his papers, so littered with mathematical prob
lems, spectral shapes flicker, becoming little marching crea
tures, the “glorious ranks” of professors past and present, “who
scrutinised the trembling lines.” As if this were not enough, an
awful form then arose before the despondent undergraduate, a
creature intent on defending the timeworn practices of academia.
Angular in form and feature,
Unlike any earthly creature,
She had properties to meet your
Eye whatever you might view.
Hair of pen and skin of paper;
Breath, not breath but chemical vapour;
Dress,—such dress as College Draper
Fashions with precision due.
142
Radiant Fields: Seeing by the Light of Electricity
Eyes of glass, with optic axes
Twisting rays of light as flax is
Twisted, while the Parallax is
Made to show the real size.
Primary and secondary
Focal lines in planes contrary,
Sum up all that’s known to vary
In those dull, unmeaning eyes.
As this wretched mathematical “Hag” proceeded to exorcise
all feeling for beauty and poetry from the heart of Maxwell by
demonic prayer and invocation, “Suddenly, my head inclining I
I beheld a light form shining; I And the withered beldam, whin
ing, / Saw the same and slunk away.” In place of the hideous
shape and voice of the “artificial spectre,” he caught glimpses
of “the being whom she aped.” Maxwell does not call her by
name but characterizes her as Boethius did his goddess Philos
ophy, and Alan of Lille his being Natura. Her modest, tranquil
radiance far outshone the ill-fitting guise of pedantry, for no
matter how pompously dressed, pedantry always veils, never
reveals.
Yet, Maxwell declared, creation can withstand both reason
and calculation, as long as one accompanies and adds to them
the action of worship.
Worship? Yes, what worship better
Than when free’d from every fetter
That the uninforming letter
Rivets on the tortured mind,
Man, with silent admiration
Sees the glories of Creation,
And, in holy contemplation,
Leaves the learned crowd behind!
The young Maxwell longed to leave the learned crowd of
Cambridge pedants behind, just as Faust, for all his learning,
143
CATCHING THE LIGHT
longed to leave Wagner to his dusty volumes in order that he
himself might find truth. Such proved more difficult than the
young Maxwell imagined.
Just a half year after his “vision,” Maxwell fell seriously ill,
and landed at the home of the Reverend C. B. Tayler, rector
of Otley, near Ipswich, suffering from a “brain fever.” During
this time, Maxwell studied Sir Thomas Browne’s classic Religio
Medici (Religion of a Physician), in which the devout Anglican
tread his way judiciously between heresies to a philosophic faith
acceptable to the Church of England. Maxwell’s ceaseless strug
gle to reconcile his genuine religious feelings with the rigorous
demands of natural philosophy was sincere. The closing of his
“Student’s Evening Hymn” dating from this period reflects his
pious sentiments.
Teach me so Thy works to read
That my faith,—new strength accruing,—
May from world to world proceed,
Wisdom’s fruitful search pursuing;
Till, thy truth my mind imbuing,
I proclaim the Eternal Creed,
Oft the glorious theme renewing
God our Lord is God indeed.
In these lines we recognize Browne’s two books of divinity:
“Thus there are two bookes from whence I collect my Divinity;
besides that written one of God, another of his servant Nature,
that universall and publik Manuscript, that lies expans’d unto
the eyes of all; those that never saw him in the one, have
discovered him in the other.”17
Although his youthful anguish over the schism between God
and nature must surely have echoed in later years, the gap
between faith and science became clearer and greater. Maxwell
quickly came to hold that any attempt to harmonize the two
“ought not to be regarded as having any significance except to
144
Radiant Fields: Seeing by the Light of Electricity
the man himself.”18 It was a safe and conventional position given
sanction by Luther and Calvin, and in our own day by the
neoorthodoxy of the great German Protestant theologian Karl
Barth.
The divine being of nature, as also the being of light, faded
from view in the centuries and millennia that separated the
luminous Egyptian eye from the ether whose vibrations com
ported themselves to the differential equations that Maxwell was
destined to discover. More often than not, Natura gave way to
the mathematical Hag. The centuries separating Alan of Lille
from Maxwell marked the twilight of the gods. They, the gods,
saw fit to remove themselves to regions less hostile than ours.
As the German romantic poet Holderlin put it: “Of course the
gods still live, but right up there in another world!”19
Having segregated his spiritual from his physical interests,
Maxwell was ready to mathematize Faraday.
The Electromagnetic Universe
The greatest alteration in the axiomatic basis of physics—
in our conception of the structure of reality—since the foun
dation of theoretical physics by Newton, originated in the
researches ofFaraday and Maxivell on electromagnetic phe
nomena .
Albert Einstein20
Einstein’s words prepare us for what is about to happen. The
understanding of reality held by science, and increasingly by
everyone since Newton, was changing profoundly. We are ap
proaching the stunning culmination of that revolution with Max
well’s mature treatment of light. Until now, a mechanical
universe of matter in motion had captivated the scientific imag
ination, but the researches of Faraday and Maxwell would ul
timately entail a transformation of that worldview down to its
145
CATCHING THE LIGHT
very foundations. New powers, electric and magnetic, had been
explored in the previous hundred years, and now our language,
imagery, and mathematical sophistication were ready for their
greatest achievement, the electromagnetic imagination of light.
Maxwell accomplished the revolution without philosophical fan
fare, but through the instrument of mathematical physics.
STATIONARY BICYCLES fall over. For that reason, we provide
children with training wheels when they are first learning to ride
one. They are not essential for learning the skill, and they make
fast maneuvering awkward, but they do mollify fears. The ether
was like a pair of training wheels for nineteenth-century phys
icists. It seemed obvious to everyone that wave motion was
carried by some kind of medium, so one was cooked up. Maxwell
was as sure about it as the rest. In his 1878 article on the
“Ether” for the Encyclopedia Britannica, Maxwell wrote: “What
ever difficulties we may have in forming a consistent idea of the
constitution of the aether, there can be no doubt that the inter
planetary and interstellar spaces are occupied by a material
substance or body ...” otherwise how could light travel through
those regions? Within thirty years, however, the ether would be
abandoned. Once the electromagnetic theory of light got fully
under way, the training wheels came off, and the new machine
rode beautifully without them.
The most significant development to this end was Maxwell’s
mathematization of Faraday’s scientific imagination. Time and
again Maxwell spoke of his youthful resolve to understand all
the known phenomena of electricity and magnetism first by read
ing all of Faraday’s Experimental Researches, and only subse
quently permitting himself to read the prevailing theory of his
day. In Faraday’s language, Maxwell wished first to read the
book of “natural truth” before studying “shadowy speculations.”
Reading Faraday, Maxwell the mathematician was surprised to
146
Radiant Fields: Seeing by the Light of Electricity
find a kindred soul, someone who thought mathematically even
though he expressed himself in pictures. Faraday’s invention
and the use he made “of his idea of lines of force in coordinating
the phenomena of magneto-electric induction shews him to have
been in reality a mathematician of a very high order,” wrote
Maxwell.
Upon finishing his first early reading of Faraday, Maxwell
resolved to translate Faraday’s thinking into the language of
mathematics. In this project he was singularly successful. His
paper “A Dynamical Theory of the Electromagnetic Field,” fin
ished in 1864, is a landmark in the history of science. In it,
Maxwell synthesized all of the disparate knowledge of electricity
and magnetism into a single set of four equations, now called
Maxwell’s equations. Every electrical, magnetic, or optical ex
periment that has been done, or will be done (barring certain
quantum effects), finds its formal theoretical explanation in
terms of these four equations. Through his “translation” Maxwell
demonstrated the extraordinary power and mathematical ele
gance of Faraday’s ideas. About Maxwell’s achievement, the
Nobel laureate Richard Feynman once said, “Ten thousand
years from now there can be little doubt that the most significant
event of the nineteenth century will be judged as Maxwell’s
discovery of the laws of electrodynamics. The American Civil
War will pale into provincial insignificance in comparison.”21
For our purposes, however, Maxwell’s great synthesis of elec
tric and magnetic knowledge was important because it carried
along with it an entirely unexpected stowaway. From the earliest
beginnings of serious investigation into electricity and magne
tism until the time of Maxwell, the studies of light and elec
tromagnetism were entirely separate. Certain materials, like
amber, attracted bits of paper after being rubbed. This principle
was already known to the ancient Greeks, who called it the
elektron effect, whence the name for the electron was derived.
Benjamin Franklin had demonstrated that lightning was an elec-
147
CATCHING THE LIGHT
trical effect, and many were drawing lightning down into lab
oratories and even social parlors, playing with the newly
discovered phenomena of electricity (a life-threatening pursuit,
as some unfortunates found out). Countless amateurs and profes
sionals alike were busy with the fascinating study of electrical
effects. Other investigators were grinding lenses for telescopes
and microscopes, and studying the intricate phenomena of light
and color. What could these two realms of nature have to do
with one another? Everything! was Maxwell’s (and Faraday’s)
answer. This unifying vision more than anything else is what
caused Einstein to call the researches of Faraday and Maxwell
the origin of “the greatest alteration in our conception of the
structure of reality since Newton.”
In reading his 1864 paper, we can witness the moment of
epiphany when the new structure of reality dawned on Maxwell.
In it Maxwell develops an imagination in which every body is
bathed not by a material ether, but by the electromagnetic field.
Throughout space, surrounding and even penetrating all objects,
are Faraday’s lines of force. All electrical and magnetic effects
(attraction, repulsion, induction, etc.) can be elegantly and ex
actly explained by Maxwell’s theory within this view. Although
brilliant, the synthesis of all electrical and magnetic effects is
not the revolution of which Einstein spoke, but then Maxwell
suggests the totally unexpected.
Toward the end of his paper, Maxwell turned his attentions
away from electrical and magnetic effects to light, away from
amber and magnetite, and toward the candle. The complex equa
tions that he manipulated so skillfully, equations that completely
baffled the older Faraday, suggested a bold hypothesis. For in
his analysis Maxwell derived an equation for the electromagnetic
field that was exactly analogous to the equation Euler had de
rived for the propagation of sound waves. Moreover, from that
equation a prediction for the velocity of light could be made,
and it agreed well with the best measurements then available.
148
Radiant Fields: Seeing by the Light of Electricity
In a moment of classic understatement, Maxwell concluded,
“The agreement of the results seems to shew that light and
magnetism are affections of the same substance, and that light
is an electromagnetic disturbance propagating through the field
according to electromagnetic laws.” In this single sentence, Max
well proposed a profound change in our image of light, one in
which light, electricity, and magnetism would now, and forever
after, be entwined. Two arenas of physics, which to all outward
appearances have nothing in common, were to be united. Far
aday’s ray vibrations had come to maturity in the mathematical
imagination of Maxwell. Light was declared an electromagnetic
wave whose vibrations rippled through space.
Notice that Maxwell’s thinking retained the ideas of the past,
when he wrote that electricity, magnetism, and light were “af
fections of the same substance.” That substance was the ether.
Maxwell and everyone else still thought in terms of the ether,
but his mathematical analysis did not require it at all. For the
first time, the wave conception of light was given a full theoretical
treatment that was, in fact, free of the ether. It would take four
decades before the scientific community would fully realize this,
but the ether was dead. Like a pair of training wheels, the ether
had served a useful purpose, but could now be removed. The
electromagnetic view of light rode all the better without them.
One who urged their removal was Heinrich Hertz. In 1887,
he had experimentally verified that the laws of reflection and
refraction, known so long for light, were also obeyed by the
invisible electrical disturbances produced by sparks. Yes, light
seemed to be but a special kind of electromagnetic disturbance.
Hertz admired the mathematical formulation of light given by
Maxwell, and warned against mistaking the “gay garments [i.e.,
the ether] we have used to clothe it” for the “homely figure” the
real theory presents.
As the field concept gained acceptance Maxwell and others
came to see the electromagnetic field itself as a repository of
149
CATCHING THE LIGHT
energy. Maxwell was totally clear about his conception of the
energy associated with the electromagnetic field. “I want to be
understood literally. ... On our theory the energy resides in the
electromagnetic field, in the space surrounding the electrified
and magnetic bodies, as well as in those bodies themselves.”
The field possessed energy, according to Maxwell’s theory. From
a theorem due to Poynting, one was able to follow the flow of
electromagnetic field energy explicitly. The results are very sur
prising. Why does a light-bulb filament get hot and glow? Not
from the flow of electricity through the filament, but rather be
cause as current flows, field energy streams into the filament
from the space surrounding it. About this the mathematics
seemed clear. Energy flows around and into conductors, not
through them.
Could it be that at root everything is a form of field energy
organized into various forms? Such was the question of Maxwell
and other “energeticists.” Impressive efforts were made in this
direction, but floundered.22 Only since Einstein and the devel
opment of quantum theory have we learned in what sense energy
and matter are related.
FARADAY had believed in the eloquence of phenomena. He was
a master inquirer, a rapt observer and ingenious experimenter
who trusted what he saw. Besides the language of phenomena,
two others were spoken: one imagination of nature was mathe
matical and the other mechanical. Faraday could not understand
the former and distrusted the latter. Faraday entreated Maxwell
to translate his treatment “out of their hieroglyphics, that I also
might work upon them by experiment.” His request was not
answered. Faraday’s distrust of mechanical models in favor of
phenomena has also fared poorly in the intervening years.
The figure who, perhaps more than any other, labored longest
and hardest to combine a mechanical picture of the world with
150
Radiant Fields: Seeing by the Light of Electricity
the rigors of mathematics was William Thomson, later Lord
Kelvin. With his efforts, the material ether and its associated
imagination of light finally exhausted itself. In its wake, the
natural truths of experiment encouraged fresh spirits to unheard-
of conceptions of the being of light that reached beyond even
those of Faraday and Maxwell.
The Final Failure
One word characterizes the most strenuous efforts . . . that
I have made perseveringly during fifty-five years: that word
is FAILURE.
William Thomson
In 1841, at the tender age of eighteen, William Thomson, al
ready an enthusiastic reader of Faraday, entered the scientific
community by publishing a sophisticated mathematical com
parison of Fourier’s theory of heat and electrical action at a
distance. It was a paper that Faraday, thirty years his senior,
could not have read with much comprehension.23 Over the next
decade Thomson’s theoretical studies of heat established his
reputation, and also convinced him that all physical phenomena,
including light and electromagnetic forces, would be understood
in terms of matter moving according to mechanical laws, if only
one could discover the nature of the ether. It was a conviction
held throughout his career, one he labored mightily to realize.
If he was right, then substance must be everywhere. “We
now look on space as full,” wrote Thomson, full of a fluid whose
sole attributes he considered to be extension, incompressibility,
and inertia. All colors, heat, electromagnetic effects, etc., are
but the reflection of the more fundamental, lawful motions of
this fluid. Based on the success of Fresnel’s wave theory of light,
151
CATCHING THE LIGHT
Thomson declared of the ether, “its existence is a fact that cannot
be questioned,” and went on to calculate its probable density
from recent data on solar energy. Faraday’s lines of force were
wonderful, yes, but one needed to understand them in terms of
the mechanical properties of the ether if one was to have any
understanding of electric, magnetic, or gravitational forces.
Thomson, therefore, launched a grand attempt to create a uni
versal physics unified around the principles of simple, inertial
matter moving in accordance with mechanical laws.
Light, as usual, proved the most recalcitrant feature of the
world for Thomson. Maxwell, although sympathetic to Thomson’s
program, had been forced to abandon, at least temporarily, a
theory involving the ether. Instead, he had retreated to a math
ematical treatment for which there was no readily visualizable
model. Thomson, by contrast, refused to give in. Adopting and
transforming the conceptions of brilliant mathematical col
leagues such as Stokes and Green, he fought his whole life long
to find tenable mechanical hypotheses for the ether.
If you travel to Glasgow University, where Thomson lectured
for over five decades, you will find there a demonstration begun
by Thomson a century ago and still running. Its intention is to
elucidate his view of the ether. Midway in a container of water,
a slab of wax is lodged. Bullets are placed above the wax and
small corks are located below it. After a year or so, the bullets
slowly make their way down through the wax, while the buoyant
corks migrate upward through it.
All of space was likewise to be understood as filled by a
highly elastic substance that can vibrate at the exceedingly high
frequencies of light required by Fresnel’s theory, and which
can, at the same time, allow entire planets and stars to make
their slow and ponderous way about the universe. Stimulated
by the work of his German colleague Helmholtz on fluid motions,
Thomson proposed that space was filled with “vortex atoms” that
move like smoke-ring structures in the ether. It was, in fact, a
152
Radiant Fields: Seeing by the Light of Electricity
brilliantly conceived hypothesis, able to account for reflection,
refraction, dispersion, and polarization phenomena. Yet, as the
American physicist Willard Gibbs wrote at the time, for all its
successes, and the audacious genius of its author, Thomson’s
theory “should not blind us to the actual state of the question.
It may still be said for Maxwell’s electrical theory, that it is not
obliged to invent hypotheses but only to apply the laws furnished
by the science of electricity.”24 Thomson had constructed but
another idol to account for light, one no truer than those that
had gone before. His hypothetical vortex atoms were short-lived
and of little long-term value for the development of the science
of light. Better, Gibbs maintained, not to invent such images
but rather to stick to the pure abstract mathematical theory of
Maxwell. Like the Muslims with their prohibition of images
of God or the Prophet, Gibbs was concerned that the images of
vortex atoms blind the scientist to the homely mathematical
truths of light.
Gibbs was an ocean away from Thomson, a safe distance.
For Thomson, now Lord Kelvin, lobbied hard for his views,
and his powers of intimidation were legendary. Over the years,
he had become the reigning patriarch of British physics. Even
as an eighty-year-old man, Lord Kelvin remained a formidable
presence, as is seen in a story told by Ernest Rutherford (dis
coverer of the nuclear atom) of his Royal Institution lecture on
radium.2'’
I came into the room, which was half dark, and presently
spotted Lord Kelvin in the audience and realized that I was
in for trouble at the last part of my speech dealing with the
age of the earth, where my views conflicted with his. To my
relief, Kelvin fell fast asleep, but as I came to the important
point, I saw the old bird sit up, open an eye and cock a
baleful glance at me! Then a sudden inspiration came, and
I said “Lord Kelvin had limited the age of the earth, provided
no new source [of heat] was discovered. That prophetic ut-
153
CATCHING THE LIGHT
terance refers to what we are now considering tonight, radi
um!” Behold! the old boy beamed upon me.
Rutherford realized that the new heat source for the earth’s
remarkable warmth, even after millions of years, was the energy
given off by the decay of radium. For eighteen years, Kelvin
had fought geologists by “proving” from the principles of ther
modynamics that the earth was young. He had, of course, un
derstandably neglected to account for the undiscovered heat
provided by radioactive decay in the earth’s core, but one still
had to be circumspect to avoid Kelvin’s wrath.
For all his power, many of Kelvin’s most passionately held
convictions were to die in the face of simple experimental facts,
the “natural truths” so loved by Faraday. One such fact that he
lived to see was the careful experimental search for the ether
performed by Michelson and Morley in 1887. The experiment
showed unambiguously that there was no ether. From his own
analysis, Kelvin could see nothing wrong with the experiment,
and in his famous “Two Clouds” address of 1900, he called this
experiment a very dense cloud hanging over the ether theory of
light he favored. Only five years later the cloud was removed
by Einstein’s special theory of relativity, but with it there also
disappeared the ether theory of light. The second cloud Kelvin
saw concerned the spectral colors of light associated with in
candescence. The understanding of heat (thermodynamics) and
electromagnetism before 1900 was inadequate to account for the
colors given off when so-called black-bodies were heated. This
discrepancy would only be removed by the advent of quantum
mechanics.
Lord Kelvin was prescient. From these two clouds over nine
teenth-century physics would arise the twentieth century’s ex
traordinary transformation of physics and a new image of light.
A third cloud, however, also existed, one that would make its
own significant contribution to modern physics.
154
Radiant Fields: Seeing by the Light of Electricity
0
410 420 440
I
.1
iih r
Ca
Part of the solar spectrum.
RECALL THAT IN 1612 Galileo had pointed his telescope at the
sun and seen dark blemishes passing over its surface, to the
disbelief of contemporary astronomers. The self-educated, Ba
varian instrument maker Joseph Fraunhofer (1787—1826) took
Galileo’s spots on the sun one step further. In 1814 Fraunhofer
attached a telescope to a prism and examined the spectral
colors of sunlight more care!dully than Newton or anyone had
done. The perfect continuity of colors seen by Newton, one
blending imperceptibly into the other, was in fact broken by
black lines. Whereas Galileo had seen dark blemishes on the
bright surface of the sun, Fraunhofer discovered dark blemishes
in the otherwise glorious phenomenon of the spectrum. What
were these dark lines that marred the rainbow of sunlight?
Not until 1859 did two Heidelberg professors, Gustav Kirch
hoff and R. W. Bunsen, establish the unambiguous connection
between the patterns of dark spectral lines seen in sunlight, and
the similar bright sequences of spectral analysis of light pro
duced when vaporized materials were put into the flame of a
Bunsen burner. Every element possesses a unique light signa
ture, a unique spectrum of discrete lines.
With this discovery astronomers could for the first time turn
their telescopes to the sun and stars, and, from the spectra they
155
CATCHING THE LIGHT
saw, determine the kinds and amounts of elements present at
those distant locations. Terrestrial physics had, through the
spectral analysis of light, made a leap into the cosmos. The sun
and stars were seen to possess spectra just like those that could
be produced in Heidelberg, leading to the simple conclusion
that the stars differed little (aside from average temperature)
from our familiar world.
The spectral analysis of light must join the Michelson-Morley
and “black-body” experiments as a third cloud on Kelvin’s oth
erwise triumphant, clear view of the horizon of physics. Nine
teenth-century physics had succeeded brilliantly in nearly every
domain. Yet unexplained phenomena of light persisted. It took
courage to look at them carefully, but those who dared ultimately
were rewarded beyond their greatest expectations.
In THE FACE of these issues Lord Kelvin could, for all his bra
vado, be candid about his own accomplishments. During the
banquet celebrating the fiftieth jubilee of his professorship at
Glasgow, he responded to a toast of high praise with these words:
“One word characterizes the most strenuous efforts for the ad
vancement of science that I have made perseveringly during
fifty-five years: that word is FAILURE. I know no more of electric
and magnetic force, of the relation between ether, electricity
and ponderable matter than I knew and tried to teach to my
students of natural philosophy fifty years ago in my first session
as Professor.”26
Kelvin had steadfastly resisted the dawning of a new imagi
nation of light. If one insisted on understanding light as material,
then one was certain of failure. Kelvin, and those who thought
like him, were forced to step down. The threads first spun by
Faraday and woven by Maxwell needed still to be picked up by
fresh minds, and fashioned in ways less fettered to conceptions
of the past. Led by experiment and the prompting of genius, a
156
Radiant Fields: Seeing by the Light of Electricity
new century of light dawned in 1900, one whose implications
we are still struggling to comprehend.
We stand poised to cross the threshold of the twentieth cen
tury, yet I would have us draw back for a moment to consider
more closely what has become of light, and the phenomena it
presents to us.
Distairbainices of the Heart
When the lamp is shattered
The light in the dust lies dead—
When the cloud is scattered
The rainbow’s glory is shed.
When the lute is broken,
Sweet tones are remembered not;
When the lips have spoken
Loved accents are soon forgot.
Percy Bysshe Shelley
In the last years of his life, Faraday’s mental powers gradually
slipped away, sending him into a gentle senility. When his lone
disciple, Tyndall, who self-consciously longed to play Schiller
to Faraday’s Goethe, made his last visit to the old natural phi
losopher (Faraday detested the new designation of “scientist”),
he found that “the deep radiance, which in his time of strength
flashed with such extraordinary power from his countenance,
had subdued to a calm and kindly light.” Tyndall remembered
how he had “knelt beside him on the carpet and placed my hand
upon his knee; he stroked it affectionately, smiled, and mur
mured in a low soft voice.”27
Faraday’s last speechless pleasures were the sunsets he could
see from the gardens of his retirement house. In earlier years
the sight of a rainbow, a storm, or a sunset would set his eyes
157
CATCHING THE LIGHT
aglow. Here were “natural truths” on a grand scale, truths worthy
of our wonder and awe.
Yet throughout the nineteenth century, many asked, how did
such phenomena look through the lens of the electromagnetic
theory: Was the rainbow but a web of electromagnetic equations,
the sunset but the effect of differential light scattering, was
lightning only countless millions of electrons? Had Maxwell’s
mathematical Hag in the end gained the upper hand? Was it
enough to be a devout Anglican while she pillaged the temples
of nature, smashing gods and goddesses, only to raise up in
their place the idols of the scientific imagination?
John Stuart Mill maintained staunchly that the love of natural
beauty was no obstacle to scientific knowledge: “the intensest
feeling of the beauty of a cloud lighted by the setting sun, is
no hindrance to my knowing that the cloud is vapour of water,
subject to all the laws of vapours in a state of suspension.”28
But is the inverse also true? Does not theoretical, scientific
knowledge obscure and even slay the sentiments we have on
seeing the sublime? Concerning another sunset, we recall Henry
David Thoreau’s words for Christmas 1851: “I, standing twenty
miles off, see a crimson cloud in the horizon. You tell me it is
a mass of vapor which absorbs all other rays and reflects the
red, but that is nothing to the purpose. . . . What sort of science
is that which enriches the understanding, but robs the imagi
nation? If we knew all things thus mechanically merely, should
we know anything really?”29
To Thoreau the brilliance of worldly light seemed to threaten
the very existence of the spiritual light of imagination. Halfway
around the world the Japanese philosopher Keiji Nishitani
turned to the fourteenth-century Buddhist priest Muso Kokushi,
who expresses himself similarly in his book Muchu mondo (Ques
tions and Answers in a Dream). Muso recalls the ancients who
maintain that every sentient being possesses a spiritual light
drawn from “the samadhi of the Storehouse of the Great Light.”
158
Radiant Fields: Seeing by the Light of Electricity
The “miracle-light” of all the Buddhas is drawn from this same
source. In our own lives, every act of insight, even in such
ordinary matters as distinguishing east from west, black from
white, “is the marvelous work of that spiritual light. But fools
forget this original light and turn to the outside in search of a
worldly light.”30
During Faraday’s lifetime, the struggle was alive between
those content with worldly light and others who remembered,
even if only dimly, an original light. Goethe and Novalis in
Germany, Coleridge in England, and Emerson and Thoreau in
America; these poetic souls, far more than contemporary sci
entific minds, sought to understand science differently. Some
romantic thinkers saw no way out but the rejection of science
outright. Far more interesting were those others who followed
Goethe’s example and reexamined the essential nature of science
to determine if there was really no room in it for man.
The “unweaving of the rainbow,” as Keats would call it, had
taken many centuries. Jahve had placed the first rainbow in the
heavens when the sun dawned on Noah after the flood. As long
as the rainbow stood in the sky, the covenant between God and
mankind stood firm: “I set my bow in the cloud, and it shall be
for a token of a covenant between me and the earth . . . that the
water shall no more become a flood to destroy all flesh.”31 The
rainbow was a promise of protection. Over the centuries, every
optical discovery, and new conception of light, found application
to the rainbow. To many poetic hearts, this was the “unweaving”
of the rainbow at the hands of the mathematical Hag, an act
that threatened to remove Iris from the heavens, and so to void
the covenant between God and humankind.
Instead of merely mourning Iris’s loss, could one not under
stand science in a way that left nature her soul, and human life
its meaning? Together with the scientific breakthroughs of the
early twentieth century—quantum mechanics and relativity—
we will, therefore, listen to those who sought to reconceive
159
CATCHING THE LIGHT
science and light along more spiritual lines, and in a way that
welcomed the being Natura even if she appeared in new dress.
First, however, we can draw together many centuries of light’s
biography by considering one of light’s most beautiful effects:
the rainbow.
760
CHAPTER 7
Door of the Rainbow
Then as we walked, there was a
heaped up cloud ahead that changed
into a tepee, and a rainbow was the
open door of it.
Black Elk'
In the spring of 1988 I saw an ancient rainbow. I had seen
modern rainbows often in life, but this was my first experience
of a rainbow that stretched back to the beginning of things.
In an annual rite whose origins are beyond recorded memory,
the citizens of Gubbio, Italy, race to the monastery of St. Ubald
atop the mountain on whose flank they live. In the early morning,
three enormous, twenty-foot, phallic “candles,” or ceri, crafted
of wood, are shouldered by squadrons of men and boys and
ceremoniously marched through the various quarters of the city.
Atop each candle stands the flamelike form of a revered saint,
sanctifying this pagan ritual.
From morning to afternoon, the preparations go on with feast
ing and merriment. Then with the sun low in the western sky,
the ceri are taken to the town gate whose road leads to St. Ubald’s
basilica atop the summit. From my hilltop position in front of
the church, I could see the ceri and, all along the winding road,
161
CATCHING THE LIGHT
the thousands of spectators and uniformed racers, the latter
ready to shoulder the huge “baton” when their leg of the primal
relay race would come. With a roar it began, and with amazing
swiftness the tall, erect ceri wound their way uphill in a bizarre
procession of grace and speed.
As they rounded the final curve all near me surged and
yelled. The ceri passed by and we turned to watch them speed
through the open church doors of St. Ubald to claim victory.
Then I saw the rainbow, old and magnificent, arched in the
sky above them and the church. In an instant I understood,
The race could not end in the dark nave of a church, but
would continue until a generation of racers passed beneath
the shimmering arch of colors suspended overhead. The flick
ering saints were pagan gods aglow in the reddish light of
dusk, and the rainbow was their aureola. The race was not
to St. Ubald but to a timeless sanctuary, “and a rainbow was
the open door of it.”
It was an ancient rite and an ancient rainbow. They do not
often occur any longer in our modern world. Once, long ago,
they were more common.
By A DEEP water hole in the desert land of Australia, an abo
riginal bushman crouches with a smoking firebrand. Thirst has
drawn him to the water’s edge, but not without hazard. Lacking
the powerful magic of fire and smoke, the bushman knows that
an enormous many-colored serpent would rise from the pool’s
depths to drown and devour him. Afraid of the fire in his hand,
the watery beast will leave the thirsty visitor unharmed. Through
out Australia the poolside creature bears the beguiling name
“rainbow serpent” after its beautiful, multicolored body that
occasionally raises itself over the earth to stand arched among
the clouds. An ancient, powerful, and magical being, it has
occupied the imagination of aborigines since their origins.
762
Door of the Rainbow
The Rainbow Serpent painted on bark by Namirrgi of the Dangbon Group
on the Arnhem Land Plateau.
WHEN WANDERING Navajo gods came to a canyon chasm or
river gorge, a rainbow bridge would overarch the abyss that they
might pass easily through their lands. At their heels the sacred
coyote would play, delighted by the newborn, rainbow trail.
Some of the rainbow bridges solidified, becoming the natural
rock bridges that today span ravines in the desert southwest.
“To THE Trojans there came as a messenger wind-footed Iris,
in her speed, with the dark message from Zeus of the aegis.”2
Dark Zeus, master of the storm and thunderbolt, often called
swift Iris to carry his messages to fellow immortals and to men.
Iris, goddess of the rainbow, bridged the gulf of sky between
the heights of Olympus and the battlefields before the walls of
Troy. “Running on the rainy wind,” as Homer speaks of her, I
imagine Iris, in the form of a glorious rainbow, arching peace
fully from stormy sky to a tumultuous plain carrying the thoughts
163
CATCHING THE LIGHT
of Zeus to Akhaian or Argive. The rainbow, in the archaic Greek
mind, was a goddess and messenger—swift and changeable.
Rainbow—Daughter of Wonder
With the dawn of a sentient humanity, the spectral arc of the
heavens we call the rainbow began its enchantment of the human
mind. It has been a source of wonder, myth, and superstition,
and has also been a phenomenon pregnant with implications for
the developing scientific study of light.
Among the ancient Semitic peoples of the Near East, the
rainbow marked the transition from a remote age of growing
iniquity to our own, a theme that enters our own tradition through
the book of Genesis. The rainbow’s appearance to Noah sealed
the covenant established between Jahve and “all flesh that is
upon the earth.” Never again would there be a destruction of
the world by flood, never again would a sinful mankind be
inundated and destroyed.3 Seeing the rainbow in our half-lit sky
can act as a comforting reminder of that pledge of God to man
as we live precariously in a seriously threatened world.
To Homer the rainbow was a manifestation of the goddess
Iris; from him we know her genealogy. Thaumas, the god of
wonder, “chose Electra for his spouse I of the deep-flowing
Oceanus child; I she bore swift Iris, fair-haired harpies too.
In the Greek imagination Iris, the rainbow and messenger of
the gods, united the deep-flowing outer sea called Oceanus that
encircled the ancient world with the god of wonder, Thaumas.
Thaumas was the Greek word meaning miracle. In the story of
Iris’s origins, the Greeks embodied their own sense of awe at
the circling, colored rainbow; it was a miracle, a thaumas, one
wed to water, to Oceanus. The rainbow is the offspring of water
and wonder; Iris is the daughter of Oceanus and Thaumas. As
164
Door of the Rainbow
Plato wrote: “He was a good genealogist who made Iris the
daughter of Thaumas.”
In the early history of the rainbow, the arc of colors that
stretched from heaven to earth became quite naturally a bridge
that connected two worlds. As the Greek playwright Aristo
phanes wrote, “And Iris, says Homer, shoots straight through
the skies with the ease of a terrified dove,” bearing messages
between gods and men. The motif was common in cultures other
than the Greek as well. For North American Indians, Polyne
sians, and others, it was a pathway for souls to the upper world,
in Japan the “floating bridge of heaven.” In the Edda of Iceland,
the god Odin in the form of Gangleri asked the way from heaven
to earth and “Harr answered and laughed aloud: Now, that is
not wisely asked; has it not been told thee that the gods made
a bridge from earth to heaven, called Bifrbst? Thou must have
seen it; it may be that ye call it rainbow. It is of three colors,
and very strong, and made with cunning and with more art than
other works of craftsmanship.” Closely guarded by Heemdel,
the gods will make the last of their daily crossings over it at the
time of their departure from our world, the “twilight of the gods.”
Then will the rainbow be destroyed.
The rainbow’s mystery continues to enchant. Approach it and
it recedes from you such that you can never walk beneath its
vault, or reach its legendary end. What is the rainbow, woven
of light and rain and eye? Thaumas, wonder, accompanies it
throughout the centuries; Iris is as beautiful and miraculous now
as she was to the ancient Greek. As a youth, Gerard Manley
Hopkins captured the puzzle.
It was a hard thing to undo this knot.
The rainbow shines, but only in the thought
Of him that looks. Yet not in that alone,
For who makes rainbows by invention?
And many standing round a waterfall
165
CATCHING THE LIGHT
See one bow each, yet not the same to all,
But each a hand’s breadth further than the next.
The sun on falling water writes the text
Which yet is in the eye or in the thought.
It was a hard thing to undo this knot.
The wonder felt by Hopkins or by Homer, which they shaped
into poetry, is the root of philosophy, and the basis of science.
Returning to Plato on Iris: “This sense of wonder is the mark
of the philosopher. Philosophy indeed has no other origin, and
he was a good genealogist who made Iris the daughter of Thau-
mas.”4 The genealogy of the Western mind reaches back to Iris,
to man’s wonder at the rainbow, and reaches forward to the
twilight of the gods, and the destruction of the rainbow, the end
of the covenant between Jahve and mankind.
The Phenomenon
It is not noon—the Sunbow’s rays still arch
The torrent with the many hues of heaven. . . .
Lord Byron
Where, when, and under what circumstances do rainbows ap
pear? Inseparable from an understanding of a phenomenon is
clarity about the details of its occurrence. We have all seen
rainbows, but have you noticed in which direction they occur,
at what time during the day, what the exact shape is, and how
high they are in the sky? What are the colors of the rainbow
and in what sequence do they occur? The answer to each of
these questions is a hint about the cause of the rainbow.
While 1 was a research physicist at Boulder, Colorado, my
wife and I were often pleasantly interrupted by rainbows at
dinnertime. Small evening cloudbursts would quietly appear.
766
Door of the Rainbow
Coming off the mountains to the west, they would head out to
the eastern plains that meet the Rocky Mountains at Boulder.
The clouds and rain would pass overhead, obscuring for a mo
ment the bright sunlight raying over the mountains at dusk.
From our apartment we could not see the rainbow, but we knew
it was there. We kept an umbrella by the door for rainbows.
Going out back to the field behind the apartments, we looked
out over the plains at the magnificent bows of colors hanging in
the half-dark sky.
Perhaps my description has awakened a memory of a rainbow
you have seen, and seen clearly. Now answer the questions with
which I began. Where was my rainbow? Low on the eastern
horizon. When was my rainbow? At evening. Already with these
two observations we have the start of what I will call the “figure”
of the rainbow, that is, its geometrical and temporal form. Once
we get clear the figure of the rainbow, we can see the rainbow
with understanding, as well as with our eyes.
Another feature of the rainbow’s figure is the place of the sun.
Where is it at the time of a rainbow? Virgil knew, and tells us.
As when the rainbow, opposite the sun
A thousand intermingled colors throws
With saffron wings then dewy Iris flies
Through heaven’s expanse, a thousand varied dyes
Extracting from the sun, opposed in place.
The sun stands opposite the rainbow. To find a rainbow, turn
your back to the sun, and a thousand intermingled colors mark
the path of dewy Iris’s flight.
What are the number, kind, and order of the colors that
comprise the figure of a rainbow? Virgil has told us a thousand
of varied dyes. Aristotle asserted there are three: “the first and
largest is red,” with green and purple following as one moves
inward. Xenophanes agreed: “This which they call Iris is also
167
CATCHING THE LIGHT
a cloud, purple, and red and yellow-green [c/iZoroiw] to behold.”5
The Icelandic prose Edda also held to three.
The invariant sequence of colors in the primary bow (from
the inside out) are: violet, blue, green, yellow, orange, red.
Their precise number is variable as they blend one into another,
although often one sees three prominent hues.
What about the shape of a rainbow? Shelley responds:
From cape to cape, with a bridge-like shape,
Over a torrent sea,
Sunbeam-proof, I hang like a roof,
The mountains its columns be.
The triumphal arch through which I march,
With hurricane, fire, and snow,
When the powers of the air are chained to my chair,
Is the million-coloured bow.. . .
Aristotle correctly identified the shape of the rainbow as a
segment of a circle that is never greater than a semicircle, at
least under ordinary circumstances. Occasionally, when the po
sition of sun, observer, and spray are just right, more of the
rainbow may show itself. In his journal kept during the voyage
of the Beagle, Charles Darwin recorded the sighting of a re
markable rainbow.
The successive mountain ranges [of southern Chile] ap
peared like dim shadows, and the setting sun cast on the
woodland a yellow gleam, much like that produced by the
flame of spirits of wine. The water was white with the flying
spray [of the Beagle], and the wind lulled and roared again
through the rigging: it was an ominous, sublime scene. Dur
ing a few minutes there was a bright rainbow, and it was
curious to observe the effects of the spray, which, being
carried along the surface of the water, changed the ordinary
semi-circle into a circle—a band of prismatic colours being
continued from both feet of the common arch across the bay,
168
Door of the Rainbow
From
sun
42”
Antisolar
point
The geometry of the rainbow.
close to the vessel’s side: thus forming a distorted, but very
nearly entire ring.6
If its shape is usually semicircular, where is the center of a
rainbow? It can be found simply as follows. Draw a line from
the sun through the observing eye, continuing it to the ground
under the rainbow. There you will notice a shadow, the shadow
of your own head. The line connecting sun, eye, and shadow is
the axis around which the rainbow circles.
A second line can be drawn from the eye to the rainbow itself.
No matter where on the rainbow, and no matter whether in a
spring shower or the mist of a garden hose, the angle formed
between the first line and the second is always forty-two de
grees—the rainbow angle.
A careful observer will also notice several other aspects of
the rainbow. First, the rainbow appears like a boundary between
a light-filled, interior space and a dark exterior band—Alex
ander’s dark band. The colors of the rainbow thus arise at the
769
CATCHING THE LIGHT
meeting of light and dark. I called the dark exterior a band
because at its outer edge a fainter, “secondary” rainbow often
appears, again at a precise angle, now fifty-one degrees. The
colors of the secondary bow are inverted, with red on the interior
and blue outside. This perversion of the color sequence gave
rise to the widespread belief that the secondary bow was the
handiwork of Satan, who created it as a parody of the Lord’s
rainbow, which had marked his covenant with Noah. In Germany
and Arabia, it was called the “devil’s rainbow.” Thus, beside
the signs of God, there always stood the reminders of his ad
versary.
One final aspect of the rainbow phenomenon is the so-called
supernumerary arcs. These are ephemeral arcs usually of al
ternating pink and green that sometimes appear just below the
primary bow. Look for them in the next rainbow you see; they
are often missed, but their graceful beauty rewards the attentive
observer.
In Boulder, my wife and I would stand next to each other
under the protection of our umbrella, each watching Iris run.
Behind us the sun was setting, before us on the ground were
the shadows of our two heads. A puzzle—reflect on it with
Hopkins: “The rainbow shines, but only in the thought / Of him
that looks.” Two heads make two shadows, and so also two axes
pass from sun through two pairs of eyes to the shadows of our
heads. Thus do two rainbows circle the axes we each define.
We “See one bow each, yet not the same to all, / But each a
hand’s breadth further than the next.” Her bow and mine stood
apart, as did we two. I would literally have had to see through
her eyes to see her rainbow, and she had better close one eye
so as to remove the final ambiguity. This was Hopkins’s puzzle
when writing on the rainbow.
All these features belong to the figure of the rainbow. In seeing
the arc of colors before us, we should gather into the phenomenon
the sun at our back, our eyes, the mist ahead, its colors and
170
Door of the Rainbow
accessory bows. The geometry and timing of the rainbow together
with these are all part of the godly figure of Iris.
Until the seventeenth century in France, the rainbow was
known primarily as iris in honor of the Greek messenger goddess.
Rene Descartes replaced that ancient name with a more prosaic
and, until then, seldom-used phrase: arc-en-ciel, “arc-in-the-
sky.” This is one more instance where the transition from the
mythic to the scientific reflects itself in the evolution of language.
But evolution continues, and I wonder if the grandeur of the
rainbow will not one day call for a name again filled with imag
ination.
Although displaced from her heavenly haunts, Iris still reigns
in one part of our universe. Consider once more the line passing
from the sun through the pupil of the observing eye. When that
line is extended out to sunlit mists, a rainbow appears to en
circle, like Oceanus, that axis of sight. But another far more
modest ring of color encircles the same axis. It rests upon the
aqueous organ of vision itself, between a black interior and a
white exterior, at the threshold to an inner world, and “is the
open door of it.” That slight ring of color carries still the name
of Iris—messenger of the gods.
Unweaving Rainbows
Let us now explain the nature and cause of halo, rainbow,
mock suns, and rods, since the same account applies to
them all.
Aristotle7
The history of the theory of the rainbow runs from Aristotle
through Grosseteste, Descartes, Newton, and the wave theo
rists.8 New concepts of optics either arose from careful study of
the rainbow or were quickly applied to it in order to explain its
171
CATCHING THE LIGHT
mysterious features. The rainbow thus provides an ideal occasion
for observing in a single natural phenomenon the changing eye
mankind has brought to optical effects. It will reflect the evo
lution of light as we have followed it from remote antiquity to
the end of the nineteenth century.
Our view of the rainbow began to change from the mythic
to the scientific during the Hellenic period. Already we find
an impressive treatment by Aristotle in his Meteorology, where
it is considered a sublunar “meteor” of light together with
halos, the aurora, comets, and meteorites. His observational
account is remarkable for its clarity, and is the first to con
sider carefully the geometrical and physical basis for the rain
bow. Aristotle submitted Iris to the same rigorous analysis by
which Euclid had geometrized sight. As regards the physical
origin of its colors, Aristotle treated clouds as made up of
mist that acted as innumerable tiny mirrors, so small that they
reflected the sight of the observer without maintaining the
image of the sun (remember the visual ray of the Greek eye).
Thus, from the eye a visual ray ran to the rainbow (i.e., to
the mirrorlike droplets of a cloud), and from there to the sun.
Aristotle’s theory of colors viewed all colors as intermediates
between white and black. Applying his theory to the rainbow,
three colors arose: “When sight is relatively strong the change
is to red; the next stage is green, and a further degree of
weakness gives violet.” These were the colors of Aristotle’s
rainbow.
The next great step in the understanding of the rainbow was
given by Robert Grosseteste in his little book On the Rainbow
and the Mirror? He opens with thoughts regarding visual rays,
and his metaphysics of light. He writes,
We must not think that emanation of visual rays is just an
imaginary idea without reality, as those persons profess who
consider the part and not the whole. But we ought to know
172
Door of the Rainbow
that a visible species is a substance of like nature with the
sun, which lights and radiates. The visible species, when
conjoined with the radiation of an external illuminated body,
completes perception.10
Grosseteste here summarizes beautifully the Platonic under
standing of light common to the Middle Ages. The two species
of emanation, from eye and sun, are like one another, and when
they are conjoined, perception occurs.
From thoughtful considerations regarding reflection and the
characteristics of rainbows, Grosseteste goes on to conclude
that Aristotle’s explanation of the rainbow as due to reflection
of light from clouds is impossible. He suggests refraction
instead. “It is therefore necessary that the rainbow be made
by the refraction of the sun’s rays in the moisture of a convex
cloud.”11
The phenomena associated with refraction are manifold and
familiar. Whenever light (or for the Greeks, sight) enters or exits
a new medium at an angle, it appears deflected at the interface.
Grosseteste thought that perhaps the moisture of a cloud, when
specially shaped into a convex form, might refract the light of
the sun into a rainbow. All details of how this would occur are
missing, but Grosseteste had advanced a seminal idea that would
prove correct.
Approximately a century after Grosseteste’s little treatise,
Theodoric of Freiberg (not Boethius’ murderous lord) took mat
ters a step further. Sometime after 1304, Theodoric performed
a study of the rainbow that was revolutionary in method as well
as results. His predecessors had considered the watery medium
of the cloud as a whole, neglecting the role of each tiny raindrop.
By contrast Theodoric considered the cloud to be made of in
dividual drops each of whose interaction with light was impor
tant. If a single drop could be studied with sufficient precision,
then by compounding the effects of many similar drops, one
173
CATCHING THE LIGHT
could build up a rainbow. This is just the principle of analysis
that was later used so effectively by Newton in the calculus and
the problem of gravitation.
The small size of a water drop made its study difficult, so
Theodoric enlarged it to a spherical glass vial of water placed
in the sun. From his investigations, he concluded that as light
enters the raindrop it first refracts, is then reflected at the con
cave inner surface at the rear of the drop, and then exits, once
again with a refraction. “Such radiation, I say, serves to explain
the production of the rainbow,” wrote Theodoric.12
As if this were not enough, Theodoric went on to consider
the production of the fainter secondary bow. He rightly recog
nized that an exactly similar process takes place but now with
an additional reflection inside the droplets.
Finally, in putting the droplets back into the sky, Theodoric
realized that each drop can offer only a single color to the
eye depending on the precise geometrical relationship between
eye and drop and sun. Thus a rainbow, seen all at once, must
be the product of many such drops. “Consequently, if all of
the colors are seen at the same time, as happens in the rain
bow, this must necessarily result from different drops which
have different positions with respect to the eye and the eye
to them.”13
Look at dew in the morning sun. Each droplet sparkles
like a gemstone. Find one that shines brightly, and slowly
move your head up and down. The glistening colors you see
always pass through the same ordered sequence of the rainbow:
red, yellow, green, violet. Place these dewdrops in the sky
as rain, and they continue to glisten, each with a single color.
Move your head again and new droplets shine red while the
old ones turn yellow. Geometry is everything in the figure of
light. Colorless in itself, each droplet sparkles, a diamond
of colored light standing between the eye of Horus and the
eye of man.
174
Door of the Rainbow
Optical THEORY GRADUALLY encircled the rainbow, but two final
contributions were still needed. The first was provided by Rene
Descartes and the second by Isaac Newton.
Descartes’s contribution to the rainbow was subtle but pro
found, uncovering the reason for the rainbow angle and also
the presence of Alexander’s dark band. By way of introduction,
consider for a moment the trajectory of water from your garden
hose. Slowly vary the angle of the nozzle from straight forward
to directly up, and note the changing distance the water travels.
Initially, as the angle increases, so does the distance. However,
when the nozzle is at a forty-five-degree angle, the water
reaches its farthest; increasing the angle still more only causes
the water to retreat. As Galileo first noted for artillery, max
imum range is achieved at forty-five degrees, with shorter
distances being traveled for angles both greater and less than
that.
In studying the rainbow, Descartes deduced a similar behav
ior. As a codiscoverer with Snell of the first proper mathematical
law of refraction, Descartes had all the theoretical tools (but
one) for a modern account of the rainbow.
We can follow Theodoric’s good example, as Descartes also
did, and use a large spherical vial of water to simulate a raindrop.
Send into it a single narrow beam of white light. Allow the beam
first to enter the sphere at its center.
Slowly and steadily slide the entering beam across the water
sphere toward its edge. The beam emerging from the vial cor
respondingly moves farther and farther to one side until (as with
the hose) it slows to a halt and doubles back. At the moment it
halts, the refracted and reflected beam bursts into color—the
colors of the rainbow. The angle between the incoming and
outgoing beams is then just forty-two degrees—the rainbow an
gle. “Doubling back” is the key to the rainbow. Without it, no
175
CATCHING THE LIGHT
Light rays entering a raindrop (above) and leaving (be
low). Notice how they crowd around ray seven. Here
is where the rainbow will appear.
amount of refraction would be sufficient to produce a rainbow
in the sky.
As an additional bonus, the dark band receives an expla
nation. Just as no water falls beyond the maximum range of the
hose no matter at what angle you hold the nozzle, so also no
light is refracted beyond the rainbow angle no matter how light
hits the droplet of water. This absence of light implies a dark
region beyond forty-two degrees—Alexander’s dark band.
The secondary bow just outside the dark band finds a similar
explanation. The additional interior reflection present in the
latter has the consequence that the emergent beam approaches
176
Door of the Rainbow
a new fifty-one-degree rainbow angle but nowi from the opposite
side. The refracted and reflected rays halt as before
I and retreat
back to larger angles yielding colors in inverted order. Alex
ander’s dark band is preserved in the space between forty-two
and fifty-one degrees, where neither can reach, a dark abyss
between the bow of Jahve and that of Satan.
Descartes had presented the complete geometry of the rain
bow, but it was entirely monochrome. He lacked a reasonable
explanation for the essence of the bow: its colors.
OVER TWO THOUSAND years after Homer, in 1672, Newton, then
still a solitary genius at Cambridge University, sent a letter to
the Royal Society of London. The content of that letter was an
outline of a new theory of light and color that, within a century,
would complete not only our scientific understanding of the
rainbow, but change our imagination of it as well. The rainbow,
he suggested, was composed of “rays diversely refrangible,”
each one of which induced the sensation of a specific color.
Recall Newton’s corpuscular theory of light. White light, ac
cording to him, is not at all what it appears to be. In fact it is
made up of rays or atoms of light of several kinds, each kind
capable of causing a different color sensation. Newton made the
significant experimental observation that lights of different colors
are refracted by slightly different amounts. The import of this
for the rainbow is great. Although Descartes could account for
the rainbow angle as the extreme angle prior to return, there
remained still the question of the colors of the rainbow. Des
cartes understood the geometry of the rainbow in black and
white only. Newton’s theory granted it color. The rainbow angle
is slightly different for every one of the rays that together com
pose the white light of the sun. That is, the extreme angle prior
to return changes systematically from red through green to violet.
This is called dispersion. Regardless of the correctness or falsity
177
CATCHING THE LIGHT
of Newton’s corpuscular theory, the fact of dispersion was the
last missing piece. The mechanism of dispersion, its physical
cause, is not what Newton imagined it to be. However, his
account, together with its implied corpuscular dynamics, carried
the day nonetheless.
The eighteenth century rejoiced. At long last the rainbow had
given up her secret, the goddess Iris was finally despoiled by
Sir Isaac Newton, a bizarre irony for a man who had so little
use for women. The poets sang Newton’s praises, among them
James Thomson:
Even Light itself, which every thing displays,
Shone undiscovered, till his brighter mind
Untwisted all the shining robe of day;
And, from the whitening undistinguished blaze,
Collecting every ray into his kind,
To the charmed eye educed the gorgeous train
Of parent colours.
Light itself was now illuminated by the even brighter light of
Newton’s mind, and stood revealed as a train of parent colors
normally twisted into whiteness. In prism and rainbow the “yel
low tresses” of sunlight are disentangled and displayed as rays
diversely refrangible.
With the rainbow now explained, only a few minor curiosities
remained, such as the supernumerary bows of pink and green
just below the primary bow. The modem account understands
these to be due to interference effects of the kind first suggested
by Thomas Young. Thus do we find in the rainbow a chronicle
of every feature of optical theory, old and new.
Around it also has swirled the emotions of Romantic hearts
who saw in the triumph of optics over the rainbow the death of
poetic sensibilities. Mark Twain wrote of his beloved Mississippi
that he could see her two ways: as writer or as riverboat pilot.
Through the eyes of the writer, every swirl, fallen bough, or
178
Door of the Rainbow
lapping wave was the occasion for prose redolent with allusions
and feelings. To the calculating eyes of the pilot, these very
same signs were the indicators of shifting sandbars, the telltale
effects of a recent storm, or some other natural occurrence that
could threaten the safe passage of his vessel. Are these two
views not completely incompatible? While one reigns the other
sleeps, like day replacing night and night day.
Those who lived under the inspiration of night sent up a lament
for the loss of the rainbow.
Awful Rainbows
. . . Do not all charms fly
At the mere touch of cold philosophy?
There was an awful rainbow once in heaven:
We know her woof her texture; she is given
In the dull catalogue of common things.
Philosophy will clip an angel’s wings,
Conquer all mysteries by rule and line,
Empty the haunted air, the gnomed mine—
Unweave a rainbow. . . .
John Keats, Lamia (1820)
Some, with Alexander Pope, would eulogize Newton, declaring,
“Nature and nature’s laws lay hid in night, / God said, ‘Let
Newton be!’ and all was light.” To them, the development of
scientific consciousness was progress, and its accomplishments
distinguished our time from that of earlier barbarism and ig
norance.
Others, with Keats, lamented the loss of Iris—swift and beau
tiful—from the heavens. They saw the advent of science as
hastening the twilight of the gods, the destruction of the Bifrbst
bridge. The rich, inner soul-world of the rainbow was plundered,
and in its place streamed abstract atoms of color. No matter
179
CATCHING THE LIGHT
how much finery the poet might hang on Newton’s text, its lifeless
bones shone through. A past golden age seemed lost forever.
In response they often turned their backs to the new scientific
imagination of nature.
The painter Benjamin Robert Haydon included in his diary
for December 28, 1817, a description of an “immortal dinner
party”1’ at which Keats, Wordsworth, and Lamb were present.
Keats, in a highly humorous way, abused Haydon for including
Newton’s head in his painting: “a fellow who believed nothing
unless it was as clear as the three sides of a triangle.” In his
account of the party, Haydon told of how Lamb and Keats
“agreed that he [Newton] had destroyed all the poetry of the
rainbow by reducing it to its prismatic colours. It was impos
sible to resist Keats,” and they all drank “Newton’s health, and
confusion to mathematics.”10 Their attitude was captured by
Thomas Campbell in his poem on the rainbow.17
Triumphal arch, that fill’st the sky
When storms prepare to part,
I ask not proud Philosophy
To teach me what thou art.
Can all that optics teach, unfold
Thy form to please me so,
4s when 1 dream of gems and gold
Hid in thy radiant bow?
When Science from Creation’s face
Enchantment’s veil withdraws,
What lovely visions yield their place
To cold material laws!
From a consciousness that saw in the rainbow a covenant with
God—or a goddess herself—we have come to see the rainbow
as an ephemeral phenomenon of mundane light manufactured
by refraction and rain. We have exchanged an ancient view or
imagination of the world for the one of contemporary science,
180
Door of the Rainbow
and in doing so we have, in Keats’s words, “Unwoven the rain-
bow.”
The shift in view is a simple fact. This may be seen as progress
by some and a fall from grace by others, but we simply do
inhabit a different world from that of the Australian aborigine
or ancient Sumerian. It differs less because of technical ad
vances in our external world than because of a revolution in our
way of thinking and seeing, a revolution located within us. We
are the inheritors not only of an external culture but of an inter
nal one as well. In fact these two are inextricably entwined,
one with the other. External history is inseparable from the his
tory of our interior landscape, and so the history of the rainbow
will be a history of the mind seen as if through one facet of a
many-sided gem.
The history of the rainbow from the age of myth to contem
porary optics is an exemplar wrought in miniature of our un
folding penetration of and relation to natural phenomena. It is
a powerful story that moves from Iris, messenger goddess of
Homer’s Greece, through the philosophers of antiquity, to Des
cartes, Newton, and Young, who laid the foundation fora modem
understanding of the rainbow.
Yet while engaging in itself, the history of the rainbow hides
within it another story far more significant than an external
history of science. For the changing images of the rainbow reflect
to us momentous changes in the fabric of consciousness itself.
The history of light, the rainbow, and more generally the history
of science continue to act as a text in which we read the psy
chogenesis of the mind.
Using Your Owu Light
Seeing into darkness is clarity.
Knowing how to yield is strength.
181
CATCHING THE LIGHT
Use your own light
and return to the source of light.
This is called practicing eternity.
Lao-tzu18
Owen Barfield begins his brilliant little book Saving the Ap
pearances by reflecting on the rainbow.19 His question is simply:
7s it really there?
Since you cannot approach a rainbow, cannot touch it, smell
it, get under or inside it, is it really there, or is it only a special
scattering of light from tiny fluid globes? One consideration in
favor of its reality: We have all seen rainbows. As a civilization
we share the experience, and so whatever rainbows really are,
they exist as a collective phenomenon.
But then, Barfield asks, consider a tree. Accord to it the same
treatment you did the rainbow. It certainly differs from a rainbow
in that you can approach it, touch it, perhaps even smell it.
You still cannot get inside or easily under it, but its reality
seems hardly an issue. Yet ever since Galileo, and especially
in our own century, physics has insisted that the tree, like the
rainbow, is made up of small globes of matter called atoms that
scatter light according to particular laws. Some small fraction
of that light makes its way to your eye and you “represent” the
tree to yourself by an unknown and miraculous process.
Quivering aspen leaves show us their silvery undersides as
dark green pines gently rock in the same breeze, creaking their
higher branches. Perhaps an evening shower has passed over
head and a rainbow hovers above the treetops and before a
distant hill. Physics may say that this is all a complex electro
magnetic interaction among neighboring cellulose molecules,
water, their constituent atoms, and light, but we see neither
light waves nor atoms; we see, hear, smell, and feel trees and
rainbows. They and we are the familiar objects of consciousness.
182
Door of the Rainbow
We care for such things, and to such common things we are
willing to give ourselves.
Children, trees, and rainbows exist. About this neither Bar-
field nor I wish to argue. What they are in themselves, in their
essence, we may not fully know, but we do “represent” them
to ourselves in robust colors and character. And strangely, our
entire contemporary culture represents them roughly as we do.
If we take to seeing rainbows and trees when others do not, our
hallucinations may land us in a psychiatrist’s office. Thus the
peoples of every time and place shared representations. Barfield
calls them “collective representations.”
Now comes the critical two-part question: How do such rep
resentations arise, and can they differ from one place or time
to another? Think back to our discussion of vision. Recall the
puzzle of the interior ray, the light that spreads from the lamp
of the body, our eyes. From the case of S.B. and others born
blind, we learned that seeing entails far more than the possession
of an operational sense organ. Into raw sensations flow things
such as memory, imagination, mental habits, feelings, and even
our will (in as much as we attend to something). Without the
light that we bring to sensations, the world is meaningless and
dark. If the physical eye is a camera obscura, a “dark chamber,”
then sight requires, in addition, a spiritual eye for which Em
pedocles’ image of the lantern is a more fitting metaphor.
Modem evolutionary biology quite naturally sees the present-
day physical eye as the end product of a long evolutionary
development. Eyeless fish swim the dark, lightless waters of
Echo River in the Mammoth Caves of Kentucky. Lacking light,
the fish evolved no eyes. By contrast, light has bathed our planet
for millions of years. Under its persistent influence, physical
organs of perception have arisen suited to it. Is it not also
possible that over the tens of thousands of years that intelligent
humankind has existed, a similar evolution might have taken
place in the development of the spiritual capacity that grants
183
CATCHING THE LIGHT
meaning to the proddings of our senses? Perhaps as Goethe says:
“The eye owes its existence to the light. Out of indifferent animal
organs the light produces an organ to correspond to itself; and
so the eye is formed by the light, for the light so that the inner
light may meet the outer. ... If the eye were not sunlike, how
could we perceive the light?”20
If, over millennia, the creative outer light of the sun has
called forth our physical eyes, what were the forces that crafted,
kindled, and colored our inner light? As the English chemist
and philosopher Michael Polanyi puts it, our language, tools,
and action create faculties: “. . . we interiorize these things and
make ourselves dwell in them.”2' By dwelling in them, new organs
of cognition arise. Musing on the history of light and the rainbow,
I think we can begin to sense those changeable psychic forces
in which man has lived, and that have granted meaning to
experience.
In our study of light, we have been concerned not with the
“right” explanations of contemporary optics, but with our chang
ing imaginations of light. Accord to each its rightful power and
glory, for they reigned long and well in their time. Nor did the
electromagnetic view of Maxwell have a longer tenure; we have
not yet gotten to the quantum theory of light.
It would be better if we asked not, were prior views of light
true, but rather, what is the significance of the view? What lived
within the soul of Egyptian or Greek that was content with the
eye of Horus and swift-winged Iris? As John Stuart Mill wrote,
paraphrasing Coleridge, “The long duration of a belief is at least
proof of an adaption in it to some portion or other of the human
mind; and if, on digging down to the root, we do not find, as is
generally the case, some truth, we shall find some natural want
or requirement of human nature which the doctrine in question
is fitted to satisfy.”22
Our conscious awareness of light, the rainbow, and, indeed,
all phenomena has changed vividly during the thousands of years
184
Door of the Rainbow
that we as a species have viewed them. As humans have evolved
biologically, so, too, has human consciousness evolved. More
over, that evolution continues in our own day, and its future
remains open. What will we make of it? I wonder.
Some, rejecting the evolution of mind entirely, will hold fast
to a nineteenth-century materialistic imagination of light in an
age whose science cannot support them. They may speak of
“photons” but are actually only revisiting Newton’s corpuscular
theory of light. Others, better informed, will adopt a radical
relativism. All knowledge is purely conventional, they will say.
To ask after the nature of light is but a socially sanctioned game;
the concept of light itself is merely a convention. Their decon-
structive enterprise is useful in that it undermines the compla
cency of science, but in the end it offers nothing in its place.
The end result of deconstruction is a pile of ruins: nihilism.
Evolutionary change is accepted by practitioners of this view,
but it (evolution) is ultimately meaningless.
The third avenue is one seldom taken because it demands
the most. It requires that we imagine light, ourselves, and the
world as richer and deeper than the mechanical imagination can
capture. Goethe, among others, advocated such an understand
ing. With care, but not fear, they sought a way that the physical
might open out on the spiritual. The future consciousness nec
essary to accommodate their science could not be mechanical
merely, but always ready to dwell lovingly in new phenomena,
actions, and expressions in order that fresh faculties of cognition
be created. The essence of light would not be a physical thing,
an idol, but a spiritual reality or agency.
As we enter the twentieth century in the pages ahead, all
three avenues are taken. The first two, materialistic realism and
various forms of relativism, are easily treated because many
scientists and philosophers have worked these fields already.
By contrast, the modem spiritual history of light is less familiar,
and seems, at first, antithetical to its scientific treatment. Yet
185
CATCHING THE LIGHT
I think this view mistaken. The torch carried by Goethe and
others such as Coleridge and Emerson, was picked up by the
philosopher and “spiritual scientist” Rudolf Steiner. Therefore,
as we watch the quantum theory of light unfold through the
struggles of Planck, Einstein, and Bohr, we will also follow the
simultaneous creation of a modern spiritual imagination of light
by Rudolf Steiner.
THE BIOGRAPHY of light is like the rainbow; a complex harmony
of many features that weaves together the rigorous figures of
natural law and the changing human soul to create a transient
appearance of colors. The sun low at one’s back can write into
the mist before us the text of two majestic bows—one the Lord’s
and the second Satan’s—separated by a chasm of darkness.
Graceful pink arcs work their way into the light-filled space
beneath the brighter primary bow. The unpretentious intricacy
of the rainbow easily becomes a metaphor for human life. The
mythic and scientific mingle still within us.
Goethe’s Faust, when broken and dispirited, was gently awak
ened by the sounds of Aeolian harps and the song of the spirit
Ariel. In the watery vapors of a nearby cataract, he saw hovering
a glorious rainbow. His words capture all the tumult of the
waterfalls and his anguished soul, his dark thoughts, and the
tranquil radiance of sun and human spirit. Between these two
flowers the rainbow whose many-hued reflections mirror life’s
aims and actions.
Nay, then, the sun shall bide behind my shoulders!
The cataract, that through the gorge doth thunder
I’ll watch with growing rapture, ’mid the boulders
From plunge to plunge down-rolling, rent asunder
In thousand thousand streams, aloft shower
Foam upon hissing foam, the depths from under.
Yet blossoms from this storm a radiant flower;
186
Door of the Rainbow
The painted rainbow bends its changeful being,
Now lost in air, now limned with clearest power,
Shedding this fragrant coolness round us fleeing.
Its rays an image of mans efforts render;
Think, and more clearly wilt thou grasp it, seeing
Life in the many-hued, reflected splendour.23
187
CHAPTER 8
Seeing Light—
Ensouling Science:
Goethe and Steiner
The purest and most thoughtful minds are
those which love color the most.
John Ruskin
Imagine that your laboratory was on the grounds, and amid the
struggles and joys, of a hospice run by Mother Teresa. In that
context, what would be the important questions concerning light
and color? My introduction to color science occurred in such a
setting, under the tutelage of an unusual and wonderful man—
Michael Wilson.
The year was 1974; Wilson and I were driving from Exeter
University to his color lab in Clent, near Birmingham, for a
week of work together. As we wound along the Atlantic coast
of south England, he pointed out the misty beauties of heather
and sea, but the mountains of Wales, which he hiked at every
opportunity, were his real love. There, sky and stone, pasture,
sheep and shepherd, mingled on blue-green hills. Forty years
188
Seeing Light—Ensouling Science: Goethe and Steiner
my senior in science, Wilson was a master of color; I was a
young American graduate student more accustomed to electron
impact excitation of helium than the rainbow’s hues. I had come
on the advice of a trusted professor, knowing Wilson had been
chair of the Color Group of England and author of several im
portant scientific papers on color. I knew less about his private
life, but in the days that followed, and during the years of
friendship that ensued, I would learn just how closely his life
was connected to light and color.
From the moment we drove onto Wilson’s estate, I knew my
introduction to color was to be strikingly different from university
physics instruction in Ann Arbor, Michigan. Leaving the VW
van by an old farmhouse, we walked a path up to Sunfield
Children’s Home. Some thirty years before, Wilson and a few
dedicated friends had founded a home for mentally handicapped
children that had since grown into a large residential community.
Climbing the hill that overlooked Sunfield’s property, we looked
out over fields, trees, and groups of obviously disabled children
with their attentive staff. I suddenly realized that I would study
color here, within this community of handicapped children, and
that this same community had surrounded and supported the
color work of Michael Wilson. I had become accustomed to
“centers of excellence,” staffed entirely with brilliant, high-
strung scientists and technicians. At Sunfield I found another
tempo and style of investigation, one guided by compassion
rather than grant deadlines or scientific rivalry. The setting of
Sunfield encouraged me to adopt a different attitude to my sci
entific work, one I have struggled to maintain ever since.
On our way back we stopped to visit with staff and children
in the main building. It was then I began to discover the full
range of Wilson’s engagement with color, for not only had he
worked scientifically with the straight formalism of modem color
theory, but for years he had also sensitively experimented with
color as a therapy for the handicapped children of Sunfield.
189
CATCHING THE LIGHT
Clearly, here was a man of independent mind and warm heart,
one who commanded the respect not only of scientists on both
sides of the Atlantic, but who also received, and returned in
kind, the love of the Down’s syndrome, spastic, and autistic
children of Sunfield. Wilson’s love of color had not only purified
his mind, as Ruskin had predicted, but it had also bridged
worlds I had always seen kept severely apart. I longed to know,
what were light and color to this man?
From Land to Goethe
In November of 1957, Edwin Land (the inventor of instant pho
tography) lectured on color, with demonstrations, to the National
Academy of Sciences and to the Rockefeller Institute for Medical
Research.1 His presentations—widely reported in the press—
had startled the scientific community. In them Land challenged
the very foundations of contemporary color theory. Six months
later, Land lectured to the Royal Photographic Society in Lon
don, and shortly thereafter Wilson began his well-known study
of Edwin Land’s radical revision of color science.2
During my visit, Wilson led me through Land’s demonstra
tions, one by one, in his well-equipped color laboratory (located
in a converted set of farm sheds). Land’s demonstrations were
truly astonishing. Nothing I had learned at the university could
explain what I was seeing. The standard basis for understanding
color had been laid down first by Newton. With the subsequent
advent of the wave theory of light, the connection between color
and wavelength then became commonplace. Together, these
formed the orthodox framework for the understanding of color.
Although sufficient for the rainbow, with them alone one simply
could not make sense of what I was seeing. Land’s experiments
seemed to challenge the scientific notions of color with a greater
force than any previous experiments.
190
Seeing Light—Ensouling Science: Goethe and Steiner
NEWTON had shown that if one extracted, say, yellow light
from the spectrum produced by a prism, and mixed it with
orange light similarly produced, then a color intermediate
between the two—a yellow orange—appeared. Its particular
hue depended on which color dominated the mixture, orange
or yellow. Land performed the same experiment but with a
single important modification. He projected the yellow and
orange light beams through black-and-white photographic
transparencies. The transparencies depicted an identical still-
life scene but photographed through different-colored filters.
With only the yellow image projected, one saw a purely mon
ochrome-yellow still life on the screen. None of the original
colors of the scene were present, only shades of yellow. The
same was true when the second image alone was projected
through the orange filter. Now, however, the still life was
entirely in shades of orange.
With Newton in mind, what would you expect to see if both
images were projected on top of one another? Hues somewhere
between yellow and orange as before? That is what I expected,
and most of the members of the National Academy of Sciences
expected the same. However, you do not see yellow oranges,
far from it! Reenacting Land’s demonstrations with Wilson, I
saw what appeared to be a full range of colors, including reds,
blues, and greens. But these were colors I “knew” simply could
not be there! My eyes told me one story, my training as a
physicist told me another. What was going on?
Michael Wilson answered my questions in his book-crammed
study, our feet wanning near one another at an electric hearth.
In doing so, he took me back to the dawn of the nineteenth
century, and to the German poet Goethe. Ironically, the basis
for understanding what had shocked the National Academy of
Sciences was to be found in the color studies of Germany’s great
191
CATCHING THE LIGHT
literary genius. Now I understood the name of Wilson’s institute:
the Goethean Science Foundation.
We can join Goethe at the moment of his entry into the story
of light. In doing so, we witness the entry of a new vein into
light’s history, one that carries us away from the mechanical
and electrical imaginations of light and toward a renewed spir
itual conception of it.
Colors of the Eye
We reserve the right to marvel at color's occurrences and
meanings, to admire and, if possible, uncover color’s se
crets 3
Goethe
In January of 1790, Germany’s great literary genius, Johann
Wolfgang von Goethe, lifted a prism to his eyes hoping to glimpse
the secret of color. The prism had been borrowed long before,
along with other optical equipment, from privy councillor Hofrat
Biittner of Jena, but now a messenger stood impatiently at the
door insisting upon its immediate return. The owner had un
derstandably despaired of ever getting back the equipment, all
of which had languished untouched for months in Goethe’s
closet. Goethe realized that Biittner’s importuning could be put
off no longer. The box of optical equipment was retrieved, but
before handing it over to the messenger, Goethe could not resist
pulling out a prism. He wished to see, if ever so briefly, Newton’s
“celebrated phenomenon of colors” known to him from child
hood.
The prism before him, and Newton’s theory of light firmly in
mind, Goethe looked at the white walls of the room expecting
them to be dressed in the colors of the rainbow. Instead he saw
only white! Amazed, he turned to the window, whose dark cross
frame stood out sharply against the light gray sky behind. Here,
192
Seeing Light—Ensouling Science: Goethe and Steiner
at the edge of frame and sky, where light and darkness met,
colors sprang into view in the most lively way. Like a shot
sounding in the mountains, Goethe realized and spoke aloud,
“Newton is wrong!” Returning the prisms was now unthinkable,
so Biittner’s servant was, yet again, sent away empty-handed.
Goethe’s study of color had finally begun.4
For the next forty years, Goethe experimented with light and
color, seeking not only color’s secrets, but also a method of
investigation more congenial to his temperament, one that was
at once objective and yet devoted to nature, simultaneously
science and art.
At the end of his long life, Goethe looked back and declared
at least modest success. In his own estimation, his scientific
studies ranked above all else, greater than Faust, his poetry,
plays, and novels. All these, he said, would dim in the luster
of other poets. Less perishable would be his contributions to
science, and foremost among them would be his massive study
of light, darkness, and color. To his secretary Eckermann,
Goethe often remarked in his last years that “I do not pride
myself at all on the things I have done as a poet. There have
been excellent poets during my lifetime; still more excellent
ones lived before me, and after me there will be others. But I
am proud that I am the only one in my century who knows the
truth about the difficult science of color.”5 This was Goethe’s
ripe judgment spoken after fifty years of effort in botany, color,
zoology, geology, meteorology, and many other areas of scientific
endeavor. His words have troubled a great many of his admirers.
His singular stature in the history of world literature is un
questioned, but his science has remained an enigma. Was the
great man duped, or have his efforts been consistently misun
derstood?
In ORDER TO probe the character of Goethe’s genius, we must
learn to see the world partly through his eyes. For his proposal
193
CATCHING THE LIGHT
is not given in the idiom of conventional science and so is easily
misread. Goethe does not advance a competing theory of light
in the usual way, but rather offers a wholesale reinterpretation
of the scientific enterprise itself. He does so not abstractly as
a philosopher, but concretely as an artist-scientist, or as Emer
son would later speak of him, as the paradigmatic “poet-savant.”
In Goethe, imagination joins with experiment in a way that brings
us suddenly much closer to light.
Biittner’s impatient messenger had nudged Goethe into a mud
dled reenactment of Newton’s prime experiment, and so into
color science. Yet the peculiar manner of his own entry into
color science was, Goethe felt, inappropriate for others. Char
acteristically, he chose to introduce his readers to color with
precisely those phenomena long dismissed as inconsequential
or outright misleading—optical illusions.6
One EVENING WHILE at an inn, Goethe noted the entry of an
especially beautiful young woman. Her white face was radiant,
contrasting strongly with her jet-black hair. A scarlet bodice
shaped her ample figure to advantage; Goethe was fascinated
by her. From across the room, he watched intently as she stood
in the light. After some moments, she stepped away, but as she
did so a striking twin appeared on the white wall behind, exactly
where she had been. Now, however, a black face appeared,
surrounded by a bright nimbus of light. Instead of a scarlet
bodice, the dark beauty was dressed in magnificent sea green.
The specter was even more lovely than the visitor.
We have already met and puzzled on similar perceptual co
nundrums, but in Goethe’s three-part treatment of light and
color, Towards a Theory of Color, they take center stage. Instead
of dismissing them as mischievous phantoms of the mind, Goethe
informs us at the outset that these very phenomena will be the
basis for his entire theory. Indeed, for him there is no such
thing as an optical illusion: “Optical illusion is optical truth!”
194
Seeing Light—Ensouling Science: Goethe and Steiner
he defiantly declares.7 In them is evidenced the living inter
action of our inner nature with outer nature. Of especial value
are pathological instances of color experience, for through them
the truths of color and cognition are most clearly rendered. No
wonder Goethe anticipated Dalton in the study of color blind
ness.
It was a brilliant stroke. At the outset Goethe distinguished
himself by advocating a radically different approach to color
and light, which, true to his poetic genius, was literally founded
on the imagination. What had been mere illusions became sign
posts on the path to truth. Phantoms became facts, and through
them the subtleties of color vision could slowly be unraveled.
In Goethe’s wake followed a stream of notable scientists. They
self-consciously followed his lead by studying illusions and path
ologies of perception. Edwin Land’s presentation to the National
Academy of Sciences, as well as the fascinating studies of Oliver
Sacks, are in the same tradition and ultimately rest on the
opening paragraph of Goethe’s Theory of Color.
AMONG Goethe’s papers is preserved a simple portrait of a
woman, but in reversed colors; perhaps the same striking figure
he saw that evening in the inn. Staring at it for a minute, and
then slipping it to the side, one sees the original fair-skinned
visitor reappear, hovering phantomlike over the ground behind
it. From experiences such as these, Goethe fashioned his ap
proach to light, one that starts with the eye’s own world of colors
before going on to investigate those of outer nature.
Rekindling the Empedoclean Fire
How do we see physically? No differently than we do in
our consciousness—by means of the productive power of
195
CATCHING THE LIGHT
imagination.. Consciousness is the eye and ear, the sense
far inner and outer meaning.
Novalis
In mundane imitation of the tavern scene, place a small brightly
colored object before you on a white sheet of paper. Stare fixedly
at it for thirty seconds or so. Now slide the object away, or avert
your eyes to a neutral-colored surface, and relax. Before you
hovers an image whose shape is identical with that of the viewed
object, but whose color is different from, in fact opposite to,
that of the original. It may wander and fade, but can be renewed
for a time by blinking. Its edges change, passing through a color
sequence, growing inward, and then fading from sight altogether.
If you have done the exercise, then you have seen a negative
afterimage.
Since Aristotle first described them in On Dreams, afterimages
have drifted in and out of the literature concerning sight. Bright
objects induce dark afterimages, dark objects induce light ones;
a red object provokes a green image, while a green one calls
forward a red afterimage. In these phenomena we sense a lawful
pattern, a truth about sight. Goethe called it the “law of required
change.”8 The eye “is compelled to a form of opposition: setting
extreme against extreme. . . quickly merging opposites and
striving to achieve a whole.” The pattern of appearance can be
elegantly represented by the color circle.
When stimulated by a color taken from one part of the color
circle, the eye seeks “to complete the circle of colors within
itself” by providing its opposite.9 This law takes on all the more
significance when the opposite color appears not only subsequent
to the initial impression, but simultaneous with it, as is the case
in the phenomenon of colored shadows, also carefully studied
by Goethe. While somewhat difficult to follow in words, the
effect is striking, and I encourage you to try it for yourself.10
First noticed by Otto von Guericke (the inventor of the vacuum
196
Seeing Light—Ensouling Science: Goethe and Steiner
Illuminated by red light alone
Red lamp Screen
*
Red
shadow
Green
shadow
White lamp
Illuminated by white light alone
Colored shadows appear when a region is lit by two differently colored
sources. How can the green shadow arise without green light?
pump) in 1672, colored shadows appear whenever two differently
colored sources of light illuminate a single region, as shown in
the diagram above. Clearly, two shadows can form, one caused
by blocking the light of each source. Allow one source to be
uncolored, for example, and the second to be red. Then a little
thought shows that one shadow is illumined only by red light
and the other only by white light. There are, therefore, three
regions on the screen: red alone, white alone, and red plus
white. What do you expect to see? Red alone gives a bright red
region, which makes sense. Red and white together give a pale
red, which also meets our expectations. But much to everyone’s
surprise, white alone is seen as green! Neither source of light
itself is green; no green anywhere, and yet one sees an unam
biguous green.
Here once again is the “law of required change.” When the
197
CATCHING THE LIGHT
eye (understood now as the entire visual system) is under the
dominant impression of red, as it is in the above, then it responds
by “seeing” white as green, which is the color complementary
to red. If one changes the red source to blue, then the white
region is seen as yellow orange, the complementary of blue, and
so on.
These phenomena of “chromatic adaptation,” as they are now
called, demonstrate to us the important truth that sight is active,
and possesses its own laws. Edwin Land’s color investigations
can only be understood if we include this active dimension to
sight. Prior to Goethe, such effects had been dismissed as dis
tractions. For Goethe, by contrast, such illusions were the proper
starting point and basis of a scientific study of color. Especially
through illusions, we gain a glimpse of the character of sight,
and of the “ideal,” imaginative power active within it.
A student of Goethe, as well as of modem color science,
Wilson was especially qualified to analyze Land’s experiments.
Wilson demonstrated that the blues seen in the Land experi
ments were caused by the same process responsible for colored
shadows, so carefully studied since Goethe’s time. The human
visual organization accommodates to the dominant color, yellow
orange in Land’s case. The lack of this color—like a colored
shadow—will be seen as blue violet, the complementary color
to yellow orange. The context, as well as the object itself,
strongly influences the color we see. All the inexplicable colors
in Land’s presentations derive from our unconscious adaptation
to the color context.
In the work of Goethe and Land, the mobility and intelligence
active in sight is drawn sharply to the fore. One cannot neglect
the power of the eye, its interior light, as Empedocles would
have said, if we would understand our experience of color. The
color “illusions” of Land and Goethe provide us with a rare
glimpse into the ubiquitous presence of mind in vision. Our
every perception is literally colored by context, prior experience,
198
Seeing Light—Ensouling Science: Goethe and Steiner
indeed, by every aspect of our inner world. These are all active
in producing color. As Oliver Sacks remarked after his study
of the color-blind painter, color is an integral part of ourselves,
of our lifeworld. As such, we affect it, but it also affects us, far
more deeply than we realize.
The CHILDREN had just left. Their watercolor paintings, rich
with colors, were still glistening in the afternoon light. The most
able children of Sunfield had finished a session with their art
teacher. She had worked according to a therapeutic plan de
veloped in collaboration with the home’s doctor, the class
teacher, and Michael Wilson. A pale yellow crescent moon hung
in the center of each painting surrounded by a dark blue sky.11
The children were drawn into the radiant experience of light,
and the embracing mood of the night sky.
At Sunfield, all those who might benefit from it were given
color therapy. Children unable to paint were treated in a color
projection room or color pool. My own session with Ursula Gral,
the color-room therapist, convinced me of the subtle but powerful
effects of color.
Sitting me in a small room, Ursula gradually and silently led
me on a journey into color space. The room lights changed color,
dimming to a deep blue. When they were at their dimmest, a
curtain opened before me and colors (projected from behind the
wall) delicately filled the large screen, the room, and me. Ursula
played the controls of Wilson’s instrument like a virtuoso, slowly
changing the colors, moving them through a metamorphic se
quence of quiet power. I was calmed and then awakened through
the agency of pure color.
Only at the end did I notice the small footprint on the screen.
During a therapy session for one of the children, Ursula ex
plained, the child stood up and walked toward the field of color
as one who has long been lost and suddenly sees the trail leading
199
CATCHING THE LIGHT
home. She caught his arm as his foot hit the screen. In order
to grant the child’s wish to live in color, Wilson designed and
constructed an entirely novel therapeutic environment—a color
therapy pool.
In a warm, dimly lit room, a tile wading pool is filled with water
that shimmers and sparkles like a live jewel. As a therapist and
child slowly make their way into the pool, I notice how brightly
their bodies and limbs are illuminated. Wilson has installed
strong colored lights around the perimeter of the pool just below
the waterline. Very little of the light could escape because of total
internal reflection, and as with my light box, without something
off of which to reflect, light, as always, remains invisible. Only
when the child enters is it clear that one is bathed in color as well
as water. Wilson explained to me that for some children, it is only
in the pool that they become aware of their own bodies. Swathed
in color, they notice themselves for the first time.
Sunfield explored the therapeutic use of color with modesty
and courage. Yet to use it effectively, they needed to know the
inner character of color intimately and truly. Electromagnetic
theory might help Wilson technically, but another kind of knowl
edge was required for its actual use. Here, Goethe’s color studies
were a natural point of departure for Wilson and the staff of
Sunfield. For in them the inner aesthetic dimensions of color
are as important as its outer aspects.
Color as the Character of Light
And so my nature studies rest on the pure base of living
experience.'2
Goethe
Goethe’s scientific interest in light was bom on a hillside outside
Rome under the luminous blue dome of the Mediterranean sky.13
200
Seeing Light—Ensouling Science: Goethe and Steiner
In conversations with landscape painters while he was in Italy,
Goethe felt a significant disquieting dissatisfaction—they
seemed to lack any real basis for the aesthetic use of color.
What was it, he asked, that determined the artistic application
of color? Was it only “caprice determined by taste which in tum
was fixed by custom, prejudice, artistic convention”?14 Con
versely, were colors to be chosen in mimicry of external ap
pearance alone, or was the artist guided by the cognoscenti
whose own tastes defined the norm of the period’s art? Goethe’s
temperament could not rest until he had seen through to the
unitary principles of color, principles that would embrace also
the “moral” dimensions of color. Once they were found, Goethe
felt sure that these insights would have aesthetic implications.
Significantly, Goethe’s entry into color drew its initial impulse
not from science, but from art.
After returning to his home in Weimar, Goethe opened a
compendium of science to refresh his vague memory of the
scientific explanation of color. He found there the conventional
discussion of Newton’s corpuscular theory, but it seemed of no
value for his own ends. On the verge of abandoning the question,
he reflected that he should be able to uncover the truth he sought
about color purely through his own observations and experi
mentation. The realization led to his fateful request of optical
equipment from Hofrat Biittner, and his first hallway experiment.
As Goethe lowered Biittner’s prism from his eyes and turned
away from the window, he also turned firmly away from the
corpuscular theory of light advanced by Newton. Goethe was
convinced that colors were not somehow secretly present in white
light only to be sorted out by the prism. What Goethe saw
convinced him this was an error. Darkness as well as light
appeared to be equally necessary for the production of prismatic
colors. Yet what Goethe really sought was not an alternative
mechanical explanation for color production, but another way
of explaining altogether. He rejected mechanical models of the
201
CATCHING THE LIGHT
kind common to his predecessors and contemporaries, but what
were his alternatives?
Within Germany at this time, there flourished a philosophical
movement that attempted to connect natural science to a lofty
idealism. Friedrich Schelling, Lorenz Oken, and G. W. F. He
gel were but three of a much larger group of “nature philoso
phers” who would have been only too glad to have the giant
Goethe join forces with them. Here, too, however, Goethe re
mained true to his own approach to light. The speculative phil
osophical thought systems of these countrymen were pale and
unhealthy to Goethe’s artistic eye. For example, Hegel, whom
Goethe admired, and who was one of Goethe’s few supporters
in his battle with Newtonianism, offered up the following defi
nition of light: “As the abstract self of matter, light is absolute
levity, and as matter, it is infinite self-extemality. It is this as
pure manifestation and material ideality, however, in the self-
extemality of which it is simple and indivisible.”15
If such were the fruits of Hegel’s dialectical method, Goethe
wanted nothing to do with them. Goethe felt compelled to make
his own way toward light. Trusting to his artistic sensibilities,
he sought out light where it lived, in the sensual phenomena of
color. Here, not in Hegel’s “oyster-like, grey, or quite black
Absolute” (as Hegel himself called it in a letter to Goethe),
could one read the life history of that precious resource called
light.
Goethe declined to define light either mechanically or ab
stractly, saying simply that “In reality, any attempt to express
the inner nature of a thing is fruitless. What we perceive are
effects, and a complete record of these effects ought to encom
pass this inner nature. We labor in vain to describe a person’s
character, but when we draw together his actions, his deeds, a
picture of his character will emerge.”16
Attempt to define a human being in terms of psychological
theories, and you inevitably fail to reveal his or her essential
202
Seeing Light—Ensouling Science: Goethe and Steiner
being, but as a novelist, describe how she walks, places her
hand at her side, moves her head and mouth.. . and imme
diately the inner nature of a person is revealed. Goethe knew
this method well. Therefore, he did not seek causes but a history
of appearances in which light could show its manifold nature
through color. These would combine to become a vivid and lively
biography of light, whose text would lead one to an intimate and
gentle knowledge of the being of light, spiritual as well as
physical. As with humans, so, too, with our investigation of
light, suggests Goethe. If we would know the nature of light,
then we should look to its actions and gestures, which are
colors.17
Colors, then, can act as the entryway to light’s inner nature.
He called his method a “gentle empiricism.” Learning from
Francis Bacon, but reaching beyond the narrow constraints of
induction, Goethe practiced a style of investigation which, he
wrote, “makes itself in the most intimate way identical with its
object and thereby becomes actual theory. This heightening of
the spiritual powers belongs, however, to a highly cultivated
1,18
age.
If we would anticipate that future, highly cultivated age and
practice Goethe’s “gentle empiricism” with light, then we must
explore the full range of color phenomena offered by light, be
coming “in the most intimate way identical with” them. By doing
so, we move through color, becoming not only more intimate
with the nature of light, but simultaneously initiating a trans
formation of self that leads to new faculties of insight.
We TOUCH HERE on two aspects critical to Goethe’s understand
ing of science. First, in common with all scientists, Goethe
searched for patterns, the hidden lawfulness within the welter
of color phenomena, but for him they were to be exalted per
ceptual experiences, not abstract substitutes for nature’s glory.
203
CATCHING THE LIGHT
Second, Goethe fully emphasized the significance of Bildung or
“self-transformation” in his scientific methodology.
Fashioning Organs for Seeing
Goethe saw the human being as constantly engaged in a process
of self-formation. We have learned that even natural organs such
as the eye require imagination to see. If the blind cannot be
given sight by physical means alone, then how much more true
must this lesson be for those organs of cognition that allow us
to “see” natural laws. The seeing of lawful patterns within the
multiplicity of phenomena requires suitable inner organs. They
are not given at birth, but develop in life. Nor should we confuse
these capacities with analytic facility or logic, as valuable as
these may be in their own right. In addition to analytical rea
soning, all scientists (and we too) rely on a kind of seeing, a
capacity for insight, that has been schooled through thoughtful
experience. With it one sees that which others, staring at the
same phenomena, may never see. In just this way, scientists
make their observations and discoveries.
It has happened dozens of times. Every walk I take with an
experienced naturalist convinces me that Goethe’s attention to
Bildung and the “organs” of insight was apt. Standing with a
geologist before an outcropping of rock, he sees more than I
who stand next to him. I make a few distinctions, he a hundred,
and each one tells a story to him of which I know nothing:
glaciation, a lake bed, or volcanic lava flow; he finds the fossil
under my foot. I feel not only illiterate but blind. Not only does
the geologist interpret the phenomena more fully, he sees things
I miss utterly. I cannot even see the text, much less read it. He
is Holmes, I poor Watson with all my bookish knowledge. As
Emerson wrote, “We animate what we see, we only see what
we animate.”19
204
Seeing Light—Ensouling Science: Goethe and Steiner
Such seeing can become refined, raised to a high art. As
though through a powerful telescope, these cognitive organs
bring the distant coastlines of uncharted scientific regions into
focus. Through them, and not by analytical reasoning alone, are
nature’s essential patterns discerned and scientific discoveries
made.
Remembering the “light of the body” can help us to under
stand Goethe’s point. The sighted eye requires more than the
input of natural light; it also requires Empedocles’ inner, ocular
light of intelligence. If we neglect the animating light of coherent
intelligence that illumines and flows through all our senses, then
the glory of the world stands mute before our inquiring spirit.
Goethe emphasized the importance of a light that is within. In
his words, “if the eye were not sun-like, how could we perceive
the light?”
How does one kindle the Empedoclean fire of the eye?
Goethe’s reply was simply, by active engagement with the world.
Our every thoughtful interaction is formative and profoundly
educative: “Every object, well-contemplated, creates an organ
of perception in us.”20 He considered the instrument of sight as
a paradigm of organ development and formation. He understood
it as crafted by light itself as it worked on the human organism.
“From among the lesser ancillary organs of the animals, light
has called forth one organ to become its like, and thus the eye
is formed by the light and for the light so that the inner light
may emerge to meet the outer light.”21
Goethe’s words harken back to ancient ideas, yet we can
place them into a much more modem evolutionary context. Light
is formative. Under its influence plants grow, but also the eye
was formed. Similarly under the influence of mountains, stones,
and streams, the geologist develops organs of cognition, kindles
an Empedoclean light that elucidates his beloved kingdom.
Goethe’s is a participatory science in which the sight of an
idea, the epiphanous moment, is supreme. How can one come
205
CATCHING THE LIGHT
to see an idea within the phenomena, to see through color, the
character of light? Only by building up the organs needed. The
chemist laboring over her laboratory bench, the mathematician
working at the blackboard, are working as much on themselves,
educating capacities for insight, as they are on the chemicals
in the flask or the equations on the board. Our every engagement
and action is pedagogic, educating new organs for discovery,
artistic or scientific.
The luminous eye slowly darkened during its long transit from
Egypt to Alhazen and Descartes. In Goethe’s hands it once again
comes to light and life. Can we tum this new life back on light
itself, hoping to penetrate to its essential nature? Can we develop
the “eye” that will see light invisible?
The HISTORY OF light is, in large part, a history of idolatry. One
image of light after another has been offered in place of light.
Goethe, like all the greatest scientists (Galileo and Newton in
cluded), knew the difference between the ingenious productions
of the human mind and the laws of nature. Yet Goethe’s position
differed from theirs in his unswerving commitment to the con
crete phenomenal dimensions of nature. His was an artistic, not
a mathematical spirit, one that relished the smell of pigments
and the sound of a well-turned phrase. If Goethe, like Faraday,
would see into nature’s eternal workings, then the facts or phe
nomena themselves would have to be raised to a high theoretical
level so he could “see” in them the pattern nature was weaving.
Like a portraitist working his pigments to reveal the character
of the human being before him, Goethe worked with natural
phenomena until they revealed to his attentive eye what he called
nature’s “open secret.” His method was that of a consummate
artist; but has a scientific discovery ever been made in any other
way? Emerson rightly observed, “never did any science origi
nate, but by a poetic perception.” Goethe forcefully holds our
206
Seeing Light—Ensouling Science: Goethe and Steiner
attention to the epiphanous moments in science, to the poetry
that is the heart of science.
Seeing Ideas
My perception is itself a thinking, and my thinking
a perceiving.22
Goethe
Like Goethe reaching down to grab a prism from Biittner’s case
of optical equipment, we can begin to cultivate new eyes for
light by putting a prism to our eyes to see the celebrated phe
nomenon of colors. Taking a prism into one’s hand leads to a
wealth of confused expectations. The unspoken question is, “I
wonder what I’ll see?” With our first look, wonder rises. Then,
after a minute of smiles, a slightly intent, bemused expression
of interest appears. It seems to ask, “How can I make sense of
this lovely jumble?” In that moment, we step away from the
pure pleasures of naive “empirical phenomena,” as Goethe
called them, and toward the arena of “scientific phenomena.”23
We move from wonder, through interest, toward insight, and so
approach the sight of the ideal.
It happens slowly, with the experimenter beginning to change
the conditions under which the phenomenon occurs, distin
guishing those that are essential to the effect from others that
are not. One is quickly led to simplification. Both light and dark
are needed to produce prismatic colors, and the simplest ar
rangement is a single straight boundary between light and dark
regions.
The illustration on page 208 shows the two significant configu
rations, light above, dark below; or the opposite, light below and
dark above. When viewed through a prism, blues and violets ap
pear along one edge (which edge depends on how you look
207
CATCHING THE LIGHT
Violet Yellow
Indigo <- Orange
BlueX X Red
When viewed through a prism, edge colors appear where light meets dark.
through the prism); along the other edge are yellows and reds.
These two arrangements of light and dark are complements of
each other. When seen through a prism, they yield the “cool”
colors for one arrangement and the “warm” colors for the other.
To Goethe’s eye, the very simplicity of the experiment and the
character of its results supported its importance. Recall also the
aesthetic question, still in Goethe’s mind from his Italian jour
ney. The polarity of warm and cool in color was a well-known
characterization among artists. Here it occurs again quite natu
rally, but in a scientific setting. One arrangement of light and
dark yields the warm colors, its opposite the cool. Color’s artistic
polarity seems to have an objective basis in experiments as well.
Only green and magenta are missing if we wish to complete
the full color circle. These magically appear by modifying the
edge arrangements to form thin bands of light and dark, green
appearing in the middle of the white band, and magenta in the
black band.
In order to understand the foregoing phenomena, Goethe did
not seek an abstract theory, but sought special examples that
best represented those relationships of light to darkness which
208
Seeing Light—Ensouling Science: Goethe and Steiner
Newtonian Goethean
spectrum spectrum
Black White
Boundary *" Boundary
colors White ;.<■ "lack , ■ colors
Black White
(a) (b)
Green appears as the white region narrows (a). By contrast, as the black
region narrows, magenta appears (b).
are responsible for the appearance of warm and cool colors.
These he called “f/r-phenomena,” that is, primal or archetypal
phenomena. In the case of warm colors, the yellows and reds
of dawn and dusk offer us the instance we seek. By learning to
see them, we come to comprehend one pole of the mystery of
color.
As the sun sets, its light passes through more and more of
the atmosphere en route to the eye. Thus, in its journey from
sun to us, light passes through the darkening or “turbid” medium
of air, as Goethe called it. In the process, all the warm colors
arise. This is the archetypal relationship between light and dark
ness that yields red, orange, and yellow—light through dark
ness. The stronger the darkening, the redder the color.
The blue vault of the daytime sky offers us the archetypal
instance of the other pole of color. Here light does not pass
through darkness, but just the opposite—darkness passes
209
CATCHING THE LIGHT
through light. Looking up, we gaze into the dark depths of space,
but once again the atmosphere intervenes. Now, however, it
plays a different role. In this case, the air catches the light. We
look, therefore, through the light-filled medium of the atmo
sphere into the darkness. Or, if we follow Goethe and conceive
of darkness as equally, if oppositely, active to light, darkness
shines through the light-filled air, and the cool colors arise.
Once you learn to see the law of color in the colors of the
heavens, you will see examples of it everywhere, from the blue
haze over a smoky pool table, to the use of “atmospheric per
spective” by a painter (that which is distant appears blue be
cause of the intervening turbidity of air). Prismatic colors, as
in the boundary colors discussed above, are more complex but
can also be understood in this way. In all cases, light and
darkness meet in a turbid medium to create color.
The account of color production given by the most recent
theories of physics offers a similar, if far more exact and math
ematical account. In them, colors arise through the “scattering”
of light. The turbid medium provides innumerable scattering
centers, be they molecules in air or a glass prism. From them
light is scattered according to strictly mathematical laws, and
in the process colors are produced. Even the rainbow, set be
tween Alexander’s dark band and a luminous interior region,
can be understood in an analogous way. Where light meets
darkness, colors flash into existence. Colors are, therefore, the
offspring of the greatest polarity our universe can offer. In
the mythic language of Zarathustra, colors are a reflection of
the mighty battle relentlessly waged between the god of light,
Ahura Mazda, and the dark hosts of Ahriman. In Goethe’s lan
guage, “Colors are the deeds and sufferings of light,” the deeds
and sufferings of light with darkness.24
If we follow Goethe’s pathway into color, we are not led to
models of light in terms of waves or particles, but to a perception
of those relationships between light and darkness that give rise
210
Seeing Light—Ensouling Science: Goethe and Steiner
to color. Seen aright, the phenomena of the blue sky and sunset
are the theory, and true to its Greek root, theoria, theory really
is a “beholding.” In Goethe’s words, “The highest thing would
be to comprehend that everything factual is already theory. The
blue of the heavens reveals to us the fundamental law of chro-
matics. One should only not seei anything further behind the
phenomena: they themselves are the theory.
theory. ”25
Like the geologist reading rocks, or Newton seeing the apple
fall, or Archimedes crying “Eureka!” we can grow to perceive
the laws of chromatics in the blue of the heavens and the first
light of dawn. Through studying the action of light in darkness,
and darkness in light, we come to sense the “deeds and suf
fering” that are color. Once we have kindled an Empedoclean
light within, fashioned the requisite organs of insight, the ar
chetypal phenomena appear, and in them we see an idea.
Philosophy since Plato has divided knowing into two insular
realms: ideas and experience. Naively, but tenaciously, Goethe
ceaselessly sought a way to experience ideas, to bridge the
chasm others thought unbridgeable. The union of idea and ex
perience may seem impossible, but as Goethe says, “nothing
forbids us from seeking a loving approach to that which lies
beyond our reach.”26
Goethe’s method gradually converts facts to theory, seen real
ity to ideal reality, and is Goethe’s response to philosophical
dualism. One cannot take truth by force, but perhaps indirectly,
through phenomena, sign, and symbol we may approach her.
“The True, which is identical with the divine, does not allow
itself to be recognized by us directly. Rather we discern it only
in reflection, in instance, symbol, in particular and kindred
appearances. We become aware of it as incomprehensible life
and yet cannot renounce the wish to comprehend it.”
Goethe’s method requires a reciprocal enhancement of both
211
CATCHING THE LIGHT
natural phenomena and the observing mind. It all begins with
wonder, as Plato rightly said, but then passes on to interest and
so to active inquiry. In the process, new organs of perception
are fashioned that are suited to seeing the essential aspects of
the phenomena before us. As we enhance our cognitive capac
ities we simultaneously enhance the world we see until, ulti
mately, we behold the ideal within the real as archetypal
phenomenon. To them one rises, as Goethe described the pro
cess, and from them one can descend in order to understand
specific phenomena. They are the ultimate experience, and the
limit beyond which one cannot legitimately go. Most of us,
however, do not recognize this and so go on, putting, for ex
ample, a model or idol in place of the archetype. “The sight of
an archetypal phenomenon is generally not enough for people;
they think they must go still further; and are thus like children
who after peeping into a mirror turn it round directly to see what
is on the other side.”27
GOETHE’S sense of scientific understanding is grounded in in
sight, not model building, and so is true to the heart of both
science and art.
Every scientific discovery from Galileo to Einstein can trace
its origin to the eureka experience in which a phenomenon
becomes transparent to the ideal, and an idea is seen. From
this exhilarating moment, the scientist works to translate his or
her insight into words and symbols. In the process, the eureka
experience is often lost while its technical power is retained.
Goethe was more interested in the former, seeking constantly
for means that would permit everyone to have their own epiphany
into nature’s ways, to see ideas.
GOETHE performed FOR philosophy a common piece of parlor
magic. The magician stands with two disconnected solid metal
212
Seeing Light—Ensouling Science: Goethe and Steiner
rings before the audience, one in his left hand, the other in his
right. He taps them together to prove the impossibility of their
union. Then, before their very eyes, he clangs the two and they
are linked. Each ring passes through the middle of the other.
In an instant the topology has changed utterly. Now all such
distinctions as inside-outside simply lose their meaning. Like
the two rings, Goethe considered the realms of thinking and
perceiving as interpenetrating. Perceiving is at once outside and
at the center of thinking, and thinking likewise passes through
the heart of seeing, and surrounds it.
Two worlds, kept so long apart, are united in our perception
of archetypal phenomena. To see them we must fashion new
organs of cognition, for they cannot be gained by logic alone.
Once known, they represent the highest we can hope to attain.
To the artist, one final and all-significant aspect of this method
is that in perceiving the archetypal phenomenon, one does not
denude or degrade nature but exalt her. The sunset is still
gloriously red, not reduced to differential absorption and scat
tering. The perception of a scientific idea does not require the
death of the beautiful.
The last chapter of Goethe’s Theory of Color undertakes a
preliminary treatment of the “sensory-moral” effects of color.
In these pages, Goethe describes his inner response to color,
connecting it back to earlier sections of the book. The eye’s
tendency to complete the color circle is related to the principles
of color harmony; the polarity of warm and cool colors takes on
new meaning in light of his prism experiments. The inner aspects
are as much a part of the experience of color as the redness of
red. The archetypal phenomena include the moral with the sen
sual.
Remember Goethe’s motivating question framed on a hillside
outside Rome regarding the use of color by artists. Remember,
too, the therapeutic use of color at Sunfield Children’s Home.
Textbook physics cannot answer the real questions of Goethe or
Wilson. Such answers come only by working with the phenomena
213
CATCHING THE LIGHT
themselves, because then we fashion organs for a science in
which the beautiful as well as the useful, the human as well as
the physical, can be experienced. Thus, Goethe performed a
second piece of conjuring. As Gaston Bachelard once wrote,
and as Goethe amply demonstrated, “The phenomena of the
world, as soon as they acquire a little consistency and unity,
turn into the human truths.”28
More Light!
Throughout his life Goethe was a lover of nature and especially
of light. When he was young, that love was passionate; when
he was old, it became quiet but intense. As a child, Goethe
revered God through his works, his creation of minerals, plants,
animals, and heavens. Above all these ranked the sun. To this
god, the child Goethe once constructed an altar after the fashion
of Old Testament prophets. On his father’s ornate, red, four
sided music stand he arranged his most precious specimens:
crystals, ores, shells, and plants. Yet something especially fine
was needed for the summit—a flame with gently rising smoke,
perhaps. In a small porcelain saucer the young priest placed a
tablet of incense, completing the altar. Lighting the tablet was
all that remained in order to consummate the ceremony.
The secret service occurred at dawn with only Goethe in
attendance. As a brilliant sun rose above the apartments to the
east, Goethe used a magnifying glass to focus the sunlight onto
the incense. Like a Zoroastrian priest, the child connected the
sacred fire atop his altar to the sun. The liturgy was complete.
Through the power of the sun, and with a little technical assis
tance from the lens, the mystery was enacted. Johann was con
tent.
Goethe lived long enough to see his color science ignored or
rejected by scientific contemporaries in whom mechanical con-
214
Seeing Light—Ensouling Science: Goethe and Steiner
ceptions of light were unalterably rooted. They preferred a math
ematical language to that of color experience, physical models
to archetypal phenomena. Yet this greatest of Germany’s many
geniuses never swerved from his own judgment concerning the
high value of what he had accomplished.
The light of the ferryman’s lamp in Goethe’s fairy tale The
Lily and the Green Snake turned all it touched to gold. Similarly,
colors are the precious ripples that sparkle in light’s wake. For
decades Goethe studied them, saying at his life’s end, “I have
known light in its purity and truth, and I consider it my duty
to strive after it.”29
In his final conflict with death, light was still Goethe’s last
request. A half hour before his passing, Goethe commanded
that the window shutters be opened so that more light might
stream into the room where he lay. His earthly striving at an
end, how fitting that Goethe’s last words are said to have been:
“More light!”30 Ruskin was right, those who love color are pure.
Surely to them, if to anyone, will be granted nature’s open
secret—light.
kt SUNFIELD I came to appreciate Goethe’s approach to color,
and his understanding of the essence of scientific inquiry as a
seeing of ideas, but his treatment of the “sensory-moral” aspects
of color, so important to Wilson’s color therapy, still puzzled
me. Orthodox physics had no room in it for moral dimensions
of color. I asked Wilson how he understood them. In response,
he told me a story. When he was a young man, training on the
violin and for orchestral conducting, he worked with an opera
company. During this time, he became fascinated by stage light
ing and, through that, by color. His mother gave him, therefore,
what was at first a highly puzzling book on color by a thinker
she admired. After reading half of it, Wilson threw the volume
across the room in despair, only to pick it up again later. In
215
CATCHING THE LIGHT
the intervening years, he had become a careful student of this
same philosopher and suggested that if I wished to know more
about the moral and spiritual dimensions of color as Goethe
conceived them, then I had best read him, too.
Rudolf Sterner’s Metaphysics
of Light
Only a weak light glimmers, like a tiny point in an enor
mous circle of blackness. This weak light is no more than
an intimation which the soul scarcely has the courage to
perceive, doubtful whether the light might not itself be a
dream, and the circle of blackness, reality.
Vasily Kandinsky
While in Berlin during 1908, the painters Vasily Kandinsky
and Gabriele Miinter met old friends from their Munich days,
Alexander and Maria Strakosch. Alexander Strakosch recalled
how they, together with other dear friends from Munich, would
meet, spending wonderful evenings together in Kandinsky’s Ber
lin atelier. It was a time of profound seeking, personally and
artistically, for them all. What they sought could not be provided
by the mechanical imagination of nature and self offered by
nineteenth-century science. Many felt the need for a more ex
pansive and ensouled imagination of both. Strakosch wrote,
“Many friends, knowing of our searches, drew our attention to
the weekly lectures at the Architektenhaus as essential to hear.
These were the lectures of Rudolf Steiner, the same scholar and
philosopher who, in the early 1890s, had been in Weimar as
editor of Goethe’s scientific writings. He had left the stodgy
atmosphere of the Goethe-Schiller archives for the fervor of
Berlin, where, in 1897, he assumed editorship of the avant-
garde Magazine for Literature, a position that placed him
squarely in the lively literary circles of Germany’s greatest city.
216
-
Seeing Light—Ensouling Science: Goethe and Steiner
The audience at the Architektenhaus, however, had not gath
ered for literary ends, but rather to hear Steiner speak about
his spiritual philosophy of man and universe. By the time Kan
dinsky and friends heard him in 1908, Steiner’s activities in
Berlin had changed fundamentally from those of a Berlin intel
lectual to a spiritual philosopher. He had already written and
lectured extensively on meditation, Christianity, the spiritual
history of mankind, and the spiritual dimensions of our universe,
mostly to Theosophical audiences (he would found his own An-
throposophical Society in 1913).
On March 26, 1908, Kandinsky, the Strakosches, and their
company attended Rudolf Steiner’s Architektenhaus lecture on
“Sun, Moon and Stars.” The content of the lecture clearly placed
Steiner with those who understood light in terms of spirit, that
is, with figures such as Grosseteste, Mani, and Zoroaster. He,
like Grosseteste, stood with a foot in two worlds, one in the
world of spirit and the other planted firmly in the world of
science. Here was the essential and eternal tension out of which
Steiner sought to create harmony instead of discord.
The lecture must have struck a special chord in Kandinsky,
for directly after it he painted the Ariel scene from Goethe’s
Faust, lines of which Steiner had read at the close of the lecture.
Kandinsky gave the painting to Maria Strakosch. In it a healing
rainbow hovers above a waterfall. Suspended mist calmly rises
from a watery tumult, catching the light, bending it into the
panoply of colors that bathe a broken Faust. Life is like the
rainbow. We have heard the lines before:
The rainbow mirrors human aims and action.
Think, and more clearly wilt thou grasp it, seeing
Life is but light in many-hued reflection.
Light is connected by Goethe to life; the rainbow reflects not
only sunlight, but human aims and action. Its colors are an
217
CATCHING THE LIGHT
image of our aspirations and deeds. Here was the seed for an
imagination of light that did not rob artistic imagination as it
enriched scientific understanding. Kandinsky and Steiner
understood Goethe’s metaphor and each gave it expression in
his own medium.
The SON of a minor railway official, Rudolf Steiner had been
educated at the Vienna Technical University, the MIT of Aus
tria, completing an undergraduate degree in mathematics, phys
ics, and chemistry. In addition to the usual diet of technical
subjects, he closely studied the works of Kant, Fichte, Hegel,
Nietzsche, and Darwin, later earning his Ph.D. in philosophy
while editing Goethe’s scientific works in Weimar.32 Steiner
knew science and academic philosophy, but from his personal
experience of spiritual phenomena, he felt that a broader con
ception of the world was required than that offered by nineteenth
century thought. Early on, the treatment of light became a
paradigmatic instance of the deep conflict he felt between his
scientific training and personal spiritual experiences.
As a university student, Steiner, like others in his day, grew
critical of the wave theory’s ether hypothesis.33 Yet Steiner went
much further. He felt that while light could manifest as color
in sense phenomena, light itself was essentially extrasensory.
In it one met the purely spiritual within the sense-perceptible.
As he wrote, “light is already spiritual. Within the sense-
perceptible the spiritual holds sway.”34
Not only light, but much of existence needed to be understood
spiritually, as well as physically, in Steiner’s view. Orthodox
nineteenth-century science was ill equipped and undisposed to
take up this task, so on the basis of his philosophical studies
and Goethe’s scientific investigations into color, botany, and
biology, Steiner proposed a “science of the supersensible.” His
Architektenhaus lectures were but a single instance of that am
bitious project.
218
Seeing Light—Ensouling Science: Goethe and Steiner
As a complement to the justifiable physical investigations of
light, Steiner offered a spiritual imagination of light. As we have
seen, ancient civilizations invariably viewed light and its heav
enly sources as divine. Steiner sought a modern Christian meta
physics in which the cosmos celebrated by Egyptian priests,
Greek philosophers, and Robert Grosseteste could find new life
on a philosophical footing at peace with science. As long as the
observations, experiments, and thinking of natural science re
mained strictly objective, Steiner felt that no conflict could arise
with spiritual science. Often, however, natural science extends
its reach beyond its observations, refusing a place for the spirit.
Such a position was, to Steiner, philosophically untenable and
morally bankrupt. He felt the time was ripe for a spiritual science
that could move from sense phenomena to spiritual phenomena
without lapsing into vague mystical utterings. He rejected the
then-fashionable approach of spiritualism and psychic research
as simply a displaced materialism of the spirit, and sought rather
to follow Goethe’s scientific methodology, which implied the
development of organs of cognition suited to every domain of
experience, including domains of the spirit.
Spirit Light
The universe seen from within is light; seen from without,
by spiritual perception, it is thought.'5
Rudolf Steiner
Like so many before him, Steiner drew on the opening pass
age of the Gospel of St. John when seeking the spiritual di
mensions of light. John speaks there of the Logos—or Word—
as the Light of the World. Everything we see, from sun to stream,
cloud, animal, and man is, says Steiner, an embodiment or
image of that divine-spiritual reality—the Logos.
279
CATCHING THE LIGHT
The purest appearance of the outer physical body of the
Logos is the light of the Sun. Sunlight is not only material
light. To spiritual perception it is also the garment of the
Logos. In sunlight, spirit streams to the earth, the spirit of
love. . . . Together with physical sunlight streams the warm
love of the Godhead for the earth.36
Light is the pure body of the Word, of the Logos. The very
same light that illumines our world is the most perfect image of
God’s creative song. Grosseteste had called light the first cor
poreal form; Steiner agreed. The electromagnetic theory of light
was, in Steiner’s view, but a pale and abstract reflection of light’s
much greater nature, one of which we should remain forever
mindful.
In LECTURES AT the close of 1920, given to an audience well
acquainted with his spiritual world conception, Steiner offered
a remarkable vista onto the nature and biography of light. He
performed a kind of spiritual archaeology of the ancient origins
of light, providing an account that was mythic in its proportions
and language, yet one he intended to be an exact imagination
of the genesis and being of light.37 Once again Goethe provides
the entry point for our considerations. “Colors are,” Goethe said,
“the deeds and sufferings of light.” The metaphor is exact; yet
what can act and suffer but beings?
The spiritual hierarchies spoken of in early Christianity (An
gels, Archangels, and so on), and which were so much a part
of Grosseteste’s world, were and still are real according to Stei
ner. To each of us is given a personal guardian Angel, over
groups there hovers an inspiring Archangel, and a Zeitgeist or
Archai rules every age. Steiner’s many descriptions of the nature
and evolution of the hierarchies were detailed and of a scope
unknown since the Gothic cathedral builders. It seems that the
220
Seeing Light—Ensouling Science: Goethe and Steiner
angelic orders have evolved over enormous periods of time in
ways somewhat similar to man’s own spiritual evolution. In par
ticular, in aeons past, the angelic hierarchies passed through a
“human” stage. During that period they possessed an inner or
moral world analogous to ours, one full of life and struggle, of
deeds, noble and ignoble. Then, in a remarkable passage, he
suggests that what the angelic host once carried within them
selves as moral realities have since become “world-thoughts,”
which we in turn experience as the light of the world. The actual
physical light that surrounds us is the fossillike remains of an
ancient moral world lived by angelic beings. What was inner
becomes outer.
With disarming logic, Steiner goes on to say that the moral
world we are now nurturing within our souls will likewise one
day become the light, or darkness, of a future evolutionary phase
of the cosmos. “We see about us today a world of light; millions
of years ago it was a moral world. We bear within us a moral
world, which, millions of years hence, will be a world of
light. . . . And a great feeling of responsibility toward the world
to-be wells up in us, because our moral impulses will later
become shining worlds.”38
Within such a view, it can never be a question of separating
the moral from the physical. In Emerson’s words, “Every natural
fact is a symbol of some spiritual fact. . . . The world is em
blematic. The laws of moral nature answer to those of matter as
face to face in a glass.”3’ Thus, every physical reality is but
the obverse of a moral reality. Steiner’s vision extended Emer
son’s to the greatest limits of human and cosmic history. In his
view, the physical world is the fruit of the moral world. Pure
hearts truly will illumine future worlds. Or, if we harbor darkness
within, then a dark world will be its lawful consequence in the
distant future. We are cocreators of the world not only through
the deeds of our hands but, in even greater measure, through
the spiritual impulses we foster inwardly. In Steiner’s words,
221
CATCHING THE LIGHT
“The physical and the moral do not exist side by side. No, they
are only different aspects of something that is in itself one. The
moral world order reveals itself out of the natural world order.”
The connection of physical and moral, of sensual and spir
itual, was, and remains, largely heretical within religious as
well as scientific circles. Powerful forces within Protestant the
ology have insisted on dividing religion from science utterly.
Likewise, within the scientific community, religion has been
viewed as treating an aspect of life entirely distinct from those
investigated by science. Max Planck spoke for many when he
said, “There can never be any real opposition between religion
and science; because the one is the complement of the other.”
For scientist and theologian alike, there have been two worlds,
and so two truths: one scientific and the other religious.
Goethe, Emerson, and Steiner, by contrast, acknowledge dis
tinctions within life, but refuse to fragment it. Truth might be
many-sided, but at root it is one. To Steiner, Planck’s views are
naive. Simple faith, as Planck envisions it, cannot withstand
the impressive and ever-growing force of natural science. If
modem habits of visualizing nature as a deterministic mecha
nism prevail, “then there is no possible way of saving the moral
realm . . . there is simply no evidence anywhere in this moral
realm of a power to prevail against the realm of the natural
order.”40 By implication, if the moral realm falls, so, too, does
the light of future aeons. Thus, Steiner sees the segregation of
scientific and spiritual knowing not only as a hindrance to a
comprehensive understanding of light, but also as a threat to
our future. To Steiner, the natural world around us grows out
of the moral world within us just as the butterfly evolves from
the caterpillar. Its shape will depend on the worm we place
inside the chrysalis.
Goethe’s COLOR THEORY had been either neglected or dis
missed. Steiner’s conception of light, together with his entire
222
Seeing Light—Ensouling Science: Goethe and Steiner
philosophy, met with more violent reactions. Attacks on him
mounted during the early twenties from all points of the compass:
clergy, academics, scientists, politicians, and proto-Nazi
groups, to name a few. Most occurred in the press, but others
were more brutal. During a 1922 public lecture in Munich,
Steiner escaped serious harm only through the quick action of
a cadre of youth who placed themselves on the stage between
a group of assailants and the speaker. The touring agency of
Sachs and Wolff informed Steiner that they could no longer
guarantee his safety, and therefore would not arrange further
public lectures for him.
On New Year’s Eve, 1922—23, following a lecture by Steiner
on the history of natural science, the headquarters of his An-
throposophical Society located in Domach, Switzerland, was set
afire by an arsonist. The structure, which seated over a thousand,
had been named the Goetheanum after Goethe. Lovingly crafted,
carved, and artistically painted by hundreds of volunteers over
the previous ten years, it was the impressive flower of Steiner’s
Anthroposophy. In that single night, Rudolf Steiner and his
collaborators watched as their years of labor were lost in flames,
in the moral darkness of persecution.
Broken but not defeated, Steiner redoubled his efforts during
the final two years of his life. He formed a University for Spiritual
Science and designed a new organic, concrete structure to be
come the second Goetheanum. It became the home for the im
pulse to which he had given voice, one that affirmed the union
of religion, science, and the arts, and which saw the moral world
of man as the true progenitor of light.
WITH GOETHE AND Steiner, a major contrapuntal theme is in
troduced into our biography of light. Until now the story line
has flown smoothly from ancient mythic to modem scientific
imaginations of light. These two figures, however, mark a re
naissance of the mythic. Both were citizens of the modem age,
223
CATCHING THE LIGHT
both understood science, and yet they sought to transform it so
as to include within it light’s spiritual dimensions.
To my eye, our world needs the enrichment of myth and an
ethic bom of compassion. Every laboratory should be sited as
Wilson’s was, on a sunny field, where scientists remain cognizant
of the questions suffering poses. In the figures of Goethe and
Steiner, I sense the germ of a new mythology and of a science
of compassion.
Novalis wrote wisely when he said that within the flame, all
natural forces are active. The truth of his words rests on the
fact that the flame, and the light it sheds, are as much moral
and spiritual forces as natural ones. Light is not two things, but
one. Its efficacy lies in the unity. The scientific study of light
never need diminish its full stature. But we, like Navajo gods,
walk a fragmented landscape riven by canyons and gorges. Un
like them, we are timid, holding fast to the firm familiar soil
beneath our feet, unwilling to throw a rainbow bridge across the
dark chasms of our world. Not lunacy, but courage is needed
to see our world whole, to know that love must become concentric
with insight.
224
CHAPTER. 9
Quantum Theory by
Candlelight
Some candle clear bums bright somewhere
I come by.
I muse at how its being puts blissful back
With yellowy moisture mild night’s blear-all
black,
Or lo-fro tender trambeams truckle at the
eye.
Gerard Manley Hopkins,
“The Candle Indoors"
On the evening of October 21, 1929, America relived one of
the great moments in the biography of light. That night, Thomas
Alva Edison, Henry Ford, and President Hoover strode together
down the main street of Ford’s new living history park, Green
field Village, to Edison’s old laboratory (moved there by Ford
from Menlo Park, New Jersey). At his old lab bench, frail with
225
CATCHING THE LIGHT
age and overcome with emotion, Edison sat ready to reenact the
final moments in the creation of his incandescent electric light
of fifty years before. Broadcast over hundreds of new radio sta
tions to millions of listeners, Ford asked everyone to turn off
all their electric lights in honor of the occasion. That evening,
all of Greenfield Village, and much of America, awaited Edison’s
invention by candlelight. When the glorious moment came, Edi
son threw the switch that fed electrical power to his bulb’s fragile
carbonized filament. As its feeble glow lit his aged face, in
unison, America threw the switches that fed power through the
circuits of Greenfield Village and darkened homes across the
continent. For a few minutes, the light of all past ages had been
called back to life so that it might be ritually extinguished in a
triumphant moment of technical bravado.
Now, even more than in 1929, the candle is an anachronism,
but perhaps as a consequence it has taken on a sacred conno
tation. We, like many before us, place its small fire upon our
altars, by our bedside, or, if we have the courage, on the boughs
of our Christmas trees. We look, and as Gerard Manley Hopkins
put it, we “muse at how its being puts blissful back / With
yellowy moisture mild night’s blear-all black.” Yet the candle
flame seems to give off more than light. In a time of utility, it
is more a symbol than a household technology. And like the
rainbow, it, too, has been a doorway into the understanding of
light.
EACH Christmas for thirty-five years Michael Faraday gave a
series of lectures “before a juvenile auditory,” which quickly
became immensely popular. During the holidays of Christmas
1860-61, he gave his last series of juvenile lectures in the great
theater of the Royal Institution on The Chemical History of a
Candle,1 lectures that have since become classics in the history
of science. Although seventy years old, he spoke with the voice
226
Quantum Theory by Candlelight
of youth. “I claim the privilege of speaking to juveniles as a
juvenile myself,” he said; and so he did, with a joy and zest
that belied the fifty years of service he had rendered the Insti
tution. The apprentice bookbinder had become England’s most
distinguished and beloved scientist. Faraday invited his young
listeners to join him in his investigations, and as he told them,
“There is no better, there is no more open door by which you
can enter the study of natural philosophy than by considering
the physical phenomena of a candle.”
Light a candle and notice first the perfect cup formed below
the flame to carry the melted wax. The flame, reaching down
the wick, melts the wax at the candle’s center, while a current
of air rising around the candle keeps the rim cool and high, and
so creates a vessel perfectly suited to hold its molten contents.
The liquid within is drawn up the wick by the same forces that
draw sap up a tree or plant: capillary action. Instead of feeding
leaves and flowers, however, the liquid wax vaporizes in the
dark inner region of the flame closest to the wick, mingles there
with the air, and feeds the flame. If this were all, as it is for
some flames, a candle would shed little light. The bright yellow
cone that spreads its gentle radiance, however, is due to tiny
glowing embers of unburned carbon, the same that turn up as
soot when the wick is too long. Cold, it is the blackest of
substances, but when hot, soot becomes beautifully luminous.
To the poetic eye of Gaston Bachelard, the candle flame is a
model phenomenon. In it “The most vulgar material of all pro
duces light. It purifies itself in the very act of giving off light.
Evil is the nourisher of good. In the flame, the philosopher
encounters a model-phenomenon, a cosmic phenomenon, a
model of humanization.”2 As model phenomenon, the candle
flame is symbolic, incorporating into its nature a moral as well
as a physical aspect. Coarse matter is purified in the flame
becoming light. Looking at a flame, Paul Claudel wonders at
the transformation it effects: “Whence matter takes flight to wing
227
CATCHING THE LIGHT
itself into the category of the divine.”3 To the poet, it offers a
model of humanization, to the scientist, an unsolved riddle;
either way, the candle flame draws us toward it like moths.
In his Christmas lectures on the candle, Faraday neglected
to show his audience of children one feature which, beyond all
others, held within it the germ that revolutionized physics and
ushered in a fundamentally new conception of our physical
world. That neglected feature was the color of candlelight.
Look at a candle flame. It flickers, waves, and pulses, a lu
minous, yellow, insubstantial form. View it now through a prism
and there appears the familiar sight of the spectrum as a beau
tiful, flame-shaped array of rainbow colors. Within the delicate
colorful form lies hidden an entirely new understanding of na
ture. “In the flame of a lamp all natural forces are active,”4
wrote Novalis. Its peaceful flame offers, as Faraday has told us,
the surest doorway to knowledge. We must pass through it to
make our way to the quantum theory of light.
Light from Heat
Light is the genius of the fire process. Light makes fire.
Novalis
Any heated solid or liquid, whether it be clay in a potter’s kiln
or the filament of an incandescent lamp, glows in a way entirely
similar to soot in a candle flame. It is a universal law that is
oblivious to the substance being heated. Hot matter always gives
off light, and always in the same way.
Every potter, glassblower, and metalworker knows this law.
The colors of the heated, radiant substances with which they
work change predictably with temperature. At low temperatures
the blacksmith’s iron is the same dull red as the potter’s clay.
The hotter either becomes, the brighter and more orange and
228
Quantum Theory by Candlelight
then yellow the glow of the material. Every temperature has a
unique color (the so-called color temperature). This is the basis
for optical methods of temperature measurement often used in
glass and metal shops. In astronomy it allows scientists to de
termine the temperature of the sun’s surface to be eleven thou
sand degrees Fahrenheit. Heat and the color of light are linked.
Anyone INTERESTED IN the light of a candle or that of the sun
might naturally send its light into a prism in order to examine
the spectrum produced. Just such experiments were being done
in England and Germany in the nineteenth century with startling
results. While the English astronomer Herschel was examining
a solar spectrum, he caused it to fall on a row of thermometers.
He was not the first to do so, but he noticed something that
others had not. The thermometers that accidentally extended
beyond the visible red edge of the spectrum registered a sig
nificant rise in temperature. Where nothing was to be seen, a
heat effect was still produced. Similar observations were being
made by others about the same time but at the other end of the
spectrum, beyond the violet edge. Now, however, the effects
produced were not heat but chemical in nature. Minerals
changed color or glowed when placed beyond the violet end of
the spectrum. These investigators had discovered what we now
call infrared and ultraviolet radiation. The radiant heat one feels
from a cast-iron stove is an example of invisible infrared radia
tion. At the other end of the spectrum, not only minerals change
color in ultraviolet radiation, but so do we. Our suntan depends
in large part on ultraviolet light, as does photosynthesis in
plants.
These forms of radiation were found to behave like visible
light in all respects except that they produced no response in
the human eye. Like dogs that can hear pitches outside the
range of human hearing, creatures do exist that evidently can
229
CATCHING THE LIGHT
Photochemistry
4
ffFff
Colors — z uv : ray
X
Telecommunication Sight Medical
The electromagnetic spectrum.
see in these parts of the spectrum. We, on the other hand, must
rely on less direct means to demonstrate their existence. The
discovery of “invisible forms of light” broadened enormously
the conception of light held by physics. Nowadays we conceive
of it as running the entire gamut from radio waves to gamma
rays. The spectrum of candlelight is far broader than meets the
eye.
Many experimental studies of light from glowing hot bodies
were being done in Berlin in the last years of the nineteenth
century. The complexity of the candle flame with its varied
burning conditions and temperature regions led the Berlin sci
entists Rubens, Lummer, and Pringsheim to invent a light source
for which the temperature-color relationship could be studied
more minutely. As a consequence, theirs became the nineteenth
century’s most accurate spectral analysis of light emitted from
hot bodies. For technical reasons, it was called the study of
“black-body radiation.”5 Conceptually the experiment simply
quantified what was obvious to the eye when it viewed a glowing
body through a prism, but took into careful account the invisible,
infrared part of the spectrum.
At a given temperature a very particular, continuous spectrum
of colors is seen. For example, when, to the eye alone, hot metal
looks orange, then the brightest color in the spectrum is orange
with colors to either side (red to one side, and yellow-green-
230
Quantum Theory by Candlelight
blue to the other) fading off gradually in intensity. The specific
form of the intensity distribution, color by color, was the function
carefully measured by Rubens and others in Berlin. The question
was, how to understand it, that is, how to understand the colors
of candlelight?
As we have seen, the nineteenth-century physics community
was confident that between the mechanics of Newton, the elec
tromagnetic theory of Maxwell, and the pristine science of ther
modynamics, it could account for everything. One needed only
to increase the computational power available and all unresolved
problems could be solved. Certainly the modest light of Fara
day’s pretty candle was child’s play, and should be easy game
for the collective scientific wisdom of the previous three cen
turies. And so the best theoretical physicists of the period in
both Germany and England set out to calculate from first prin
ciples the spectrum of a candle flame. Yet they failed. In 1899,
physics could not account for the colors of candlelight!
An Upright Mam
The study of light has resulted in achievements of insight,
imagination and ingenuity unsurpassed in anyfield of men
tal activity; it illustrates, too, better than any other branch
of physics, the Vicissitudes of theories.
Sir J. J. Thomson, 1925
During the closing years of the nineteenth century, an indus
trious theoretical physicist had joined the faculty at the Uni
versity of Berlin. His name, now inseparable from the birth of
quantum mechanics, was Max Planck.6 The descendant of a
line of pastors, scholars, and jurists, Planck took from them the
conservative and upright bearing that he maintained throughout
his life. His ancestors had been exemplars of the traditional
231
CATCHING THE LIGHT
Enlightenment values of rationalism and tolerance, joined with
strong ecumenical Protestant convictions. Max Planck followed
their good example, standing in the eyes of his contemporaries
as a paragon of classical moral precepts, as well as master of
all the best fruits of classical physics. Yet life is complex. His
Prussian loyalties would later demand unsavory compromises
with Nazi science, but in 1900 his tenacious intelligence was
forcing him reluctantly to cross the border to a new imagination
of light.
At the instigation of a friend, Planck embarked upon his
theoretical analysis of light, fully expecting that classical phys
ics would be adequate to the task. Over the course of time,
however, try as he might, none of his efforts bore satisfactory
fruit. As did others, Planck began by assuming that luminous
bodies could be modeled as a collection of atomic oscillators
vibrating at distinct frequencies. With egalitarian impartiality,
the laws of thermodynamics distributed the available heat
energy evenly among them. All calculations based on this
model, including Planck’s, were in catastrophic disagreement
with the data. In 1899, after four long years of faltering,
Planck made a seemingly small but very significant concession.
In his theory of black-body radiation, he condescended to an
apparently unjustified mathematical artifice whose full import
he dimly sensed but did not then, or for several years there
after, fully appreciate.
Consider the swinging mass of a pendulum clock. Its fre
quency of oscillation is determined, as Galileo first discovered,
by the length of the pendulum. Its amplitude (that is, how far
it swings) is determined only by how it is set into motion, by
the energy imparted to the pendulum. Strange as it seems,
Planck was forced into making the extraordinary assumption
that at the quantum level, this description of the pendulum’s
motion is inadequate. In truth, not all amplitudes are allowed.
That is, we cannot pull the pendulum back to any height we
232
Quantum Theory by Candlelight
wish, but only to certain initial heights. Thus, the energy of the
pendulum cannot take on all values. Quite to the contrary, only
a discrete set of such amplitudes and energies is possible. They
are specified by Planck’s famous relationship: E = nhv. That is,
the energy of an oscillator, such as a pendulum, can only take
on values that are integer multiples of h (Planck’s constant, a
very small number) times the frequency v (the number of ticks
per second).
This single assumption changes everything. Energy can be
given to or taken from the pendulum only in specific increments.
Like walking up stairs, you cannot take half a step at a time.
And for high-frequency oscillators (corresponding to blue light),
the steps are bigger than for low frequencies (red light). In
addition, at low temperatures not enough energy is available to
excite the high-frequency blue oscillators or, put another way,
to take the large steps needed. But other oscillators are around
with smaller “red” steps that can be more easily climbed.
Planck’s analysis, therefore, led to the prediction that at low
temperatures, the red oscillators alone would be active in ab
sorbing and emitting energy. Consequently, the spectrum of
candlelight (low temperature) would be predominantly red, and
the missing blue part of the spectrum would appear only for
higher temperature phenomena.
At root, the energy given to or emitted by the oscillators is
electromagnetic, that is, light. Thus, Planck’s assumption, if
taken literally, implied that light itself might be quantized,
existing only in discrete units. Planck initially thought his as
sumption was merely a mathematical trick that expedited his
calculation, and would one day disappear. The English physicist
James Jeans agreed, suggesting that h be sent to zero at some
point in the calculation, and all would be well again with the
world. Yet Planck’s constant refused to be dismissed. Pandora’s
box of quantum mechanics had been opened, and all the atten
dant ills that flowed from Planck’s analysis could not be stuffed
233
CATCHING THE LIGHT
back into the tidy enclosure of nineteenth-century physics. Her
schel’s “box of light” was in apparent disarray. Out of this
dilemma was fashioned the new quantum of light. The colossal
implications of Planck’s modest assumption were apparent only
to a few of his peers. Henceforth, the energy that is light would
be quantized.
With this one assumption, Planck was able to derive a math
ematical formula that fit the data he had at hand quite well. To
test his theory further, Planck invited Rubens to tea at his home
outside Berlin. Rubens brought with him the results of his most
recent measurements. Planck reciprocated with his new formula
for the distribution of black-body radiation. For the first time,
theory agreed with experiment completely. The success had all
stemmed from Planck’s bizarre assumption about the discrete
motion of a pendulum.
Planck must have intuited the profound significance of this
moment, because in the flush of this success he took his beloved
son Erwin on a long and memorable walk in the Griinewald
forest outside Berlin. Erwin, only seven at the time, knew of
his father’s long preoccupation with the analysis of light. As
they walked the father confided to his child that he felt the
discovery he had made was destined to be of the same signifi
cance as those of Copernicus or Newton. It was the stuff of
revolutions. These were bold, prophetic words from an upright,
cautious man, a heresy uttered to no one other than his son,
and yet they were spoken truly. The paradoxical demands that
Planck’s theory of light has placed on the human imagination
continue to reverberate a century later in the form of wave
particle duality.
Planck himself was reluctant to accept the quantum of light
he had invented, struggling time and again to avoid the need
for it. One of the very few to take up the implications of Planck’s
analysis of candlelight was the still-unknown physicist Albert
Einstein.
234
Quantum Theory by Candlelight
The Reckless Quantum
According to the assumption contemplated here, when a
light ray spreads from a point, the energy is not distributed
uniformly, but consists of a finite number of energy quanta
that are localized in space, move without dividing, and can
be absorbed or generated as a whole .’
Albert Einstein, 1905
Einstein’s greatest accomplishments date to the remarkable pe
riod from 1902 to 1908 when he held the unpretentious position
of patent-office technician. One extraordinary scientific paper
after another was written by this unknown giant in the few hours
between work and domestic duties. During this period, Einstein,
always a bold and prescient thinker, daringly leaped beyond
Planck to suggest in 1905 that light be considered a collection
of independent particles of energy.8 With this hypothesis, Ein
stein went on to explain other recent experiments that had also
posed insurmountable difficulties for the classical wave theory
of light. The immediate reaction to his suggestion, supporting
calculations, and predictions was silence. Particles of light? The
idea was appalling. It flew in the face of a century-long vindi
cation of the wave theory. Huygens, Euler, Fresnel, Faraday,
Maxwell—had not their efforts finally culminated in the enor
mously successful view of light as an electromagnetic wave? Not
only did experiments on the interference of visible light support
it, but now also others at much longer wavelengths, including
the newly discovered infrared and radio waves. Undeterred,
Einstein was, in his own words, “ceaselessly preoccupied with
the incredibly important and difficult” question of the consti
tution of light. His research convinced him that “the next phase
of the development of theoretical physics will bring us a theory
of light that can be interpreted as a kind of fusion of the wave
and emission [particle] theories.”9
235
CATCHING THE LIGHT
-f L. a
r
—- - ■
V
’H 4
\
' 3
Albert Einstein receiving the Planck Medal from
Max Planck in 1929.
236
Quantum Theory by Candlelight
In these early years, one of America’s greatest experimen
talists, Robert A. Millikan, had been studying the emission of
electrons from metal surfaces when illuminated by light. The
details of this phenomenon formed a strong if not completely
convincing argument in support of Einstein’s particle view. Still,
the particle view of light was profoundly objectionable to Mil
likan, so objectionable that even in the face of his own data,
he called it a “bold, not to say reckless, hypothesis.”10 That
such a hypothesis could explain his data was an embarrassment,
especially as no one took it seriously. Even Planck himself, who
after 1905 had come to know and respect Einstein for his other
scientific contributions, was sharply critical of the notion of light
being advanced by Einstein.
An example of the furor that attended the introduction of light
quanta took place in 1909 on the occasion of an address by Ein
stein to the eighty-first meeting of German scientists and physi
cians in Salzburg. Planck, Rubens, Stark, and other physicists
were in the audience. Einstein gave a masterful overview of the
situation regarding light, the supposed ether, and the necessity
for light quanta. On the basis of his penetrating analysis of key ex
periments, Einstein maintained that “our current foundations of
the radiation theory must be abandoned.” He went on to ask the
question that was on everyone’s mind. Would it not be possible to
derive Planck’s result, and explain the key experiments without
the “horrendous-looking hypothesis of light quanta,” as he called
it? Could we not retain at least the classical beauty of Maxwell’s
electromagnetic fields for the propagation of light through free
space, and treat only the material processes of emission and ab
sorption differently? Einstein’s response was a firm no. He went
on to reason, starting from Planck’s own result, that to escape by
such a route was impossible. The detailed results from experi
ments on black-body radiation constrained one in the most tena
cious way, forcing one to assume light quanta, he said.
The first to respond to the lecture was Planck himself. Planck
237
CATCHING THE LIGHT
was willing to admit that there was a quantum discontinuity, but
he simply felt convinced that it could and must be located entirely
in the atomic nature of matter, and not ascribed to light itself.
How could particles of light give the enormous variety of interfer
ence effects that wave optics could explain so readily? How could
a particle interfere with itself? It was a logical absurdity!
Einstein responded by suggesting that the light quanta need
not interfere with themselves, but could interfere with other
quanta as they traveled. A brilliant response; he did not know
then that this way out was also blocked! If quanta existed, they
were even more shocking in their behavior than he realized. In
fact, Einstein was forcing the entire physics community into
accepting a quantum theory of light, a theory he firmly rejected
later in life. Photons, the light quanta first introduced by Ein
stein, do interfere with themselves, as we will see later on. The
stakes were far higher than even the boldest thinker of the period
understood. Even Einstein’s view of light as energy quanta would
prove inadequate. Planck’s caution was well grounded. As he
had prophesied to his son, it was a revolution in thought the
likes of which had not been felt since Copernicus or Newton,
and the implications were frightening.
The BATTLE BETWEEN Einstein and Planck received reinforce
ments from another side. Parallel with the passionate investi
gation into the nature of light was an entirely similar, and equally
significant, investigation into the nature of matter. One of the
key figures in this research was the charismatic Danish physicist
Niels Bohr. His penetrating study of matter drew him at first
indirectly, and later directly, into the debate on light. The the
oretical controversies in which Bohr participated will seem re
mote from our common experience of light, but in actuality they
are not. In this instance, the connection is not by way of can
dlelight but through lightning and the aurora borealis.
238
Quantum Theory by Candlelight
Night Sky and Neon Signs
Lightning is the father of light."
Franz von Baader
To the nineteenth-century nature philosopher Franz von Baader,
light is bom from a sheath of darkness, just as an infant departs
from the mother’s womb. In light there dwells God’s realm, the
creator’s love; and in darkness resides the creator’s anger. The
force of God that liberates light from darkness is the force of fire,
and the expansion of that light is the Holy Spirit. One and the
same divine being eternally manifests both in regions above as
beneficent light, and in regions below in the terrifying form of
lightning.12
The release of light is always swift, whether from the spark
of steel against flint, or a candle wick catching fire. Light is
bom suddenly, and in danger, out of darkness. To von Baader,
lightning was, therefore, emblematic of the genesis of light; it
was the father of light. Nothing stands in greater contrast to a
flash of lightning as it rends the air than quiescent candlelight.
If, as von Baader says, lightning is the father of light, then the
unlit candle must be its mother. Yet lightning presents not only
symbolic truths but scientific ones as well.
We can begin, as we did with the candle flame, by studying the
spectrum of lightning. This is especially easy because no slit is
needed before the prism; the flash itself already confines the light
to a narrow if irregular route. The dispersed image of lightning
looks very different from that of a candle flame. In place of a
smoothly varying band of colors, sharp lines of color appear, one
separated from the next in an apparently irregular fashion. These
are the “spectral emission lines” of the elements that make up our
atmosphere. As the powerful surge of electricity passes swiftly
along the lightning’s channel, each element—nitrogen, oxygen,
239
CATCHING THE LIGHT
etc.—is caused to emit light of a kind that is completely charac
teristic of the element. The universal black-body spectrum,
which was oblivious to the material heated, is gone. In its place
one sees a kind of light uniquely tied to the individual elements
that make up our atmosphere. The spectrum is not that character
istic of a heated body, but that produced by the flow of electricity
through gas. Phenomena such as these—and their laboratory
counterparts—captured the attention of Bohr in Copenhagen.
They were responsible for launching a second attack on the em
battled electromagnetic conception of light.
Lightning is not the only natural phenomenon that shows a
line spectrum. Another and much quieter phenomenon associ
ated with the polar night also shows its spectral colors, and it
also has a history rich in emblematic meanings.
WHEN VIEWED FROM several thousand miles away in space, the
polar regions of the earth show an extraordinary oval of light
many hundreds or even thousands of miles in diameter. In the
north the oval is centered over the northwest tip of Greenland.
Over time the light ring grows and contracts, changing its color
and shape as well as its size. To someone on the ground, the oval
of light appears like veils or draperies of gentle, mobile, deli
cately radiant color that can last for several hours and stretch
across the entire night sky. Growing, fading, and rippling, these
lights possess a morphology entirely of their own and have evoked
wonder and stimulated all manner of speculation over the millen
nia. 13 Since Greek times, they have been called the aurora bo
realis, which means “northern dawn.” (There is also an aurora
australis or southern dawn in the southern hemisphere.)
While the possibilities are infinite, a typical aurora might de
velop in something like the following way: The night is moonless,
dark, cold, and clear. At first, one sees only a simple, broad white
arc in the northern sky with a smooth lower edge, its ends never
240
Quantum Theory by Candlelight
touching the earth. The lower border then develops folds and
kinks, and colors now show themselves, turning the silent glow
from white to yellow, green, or red. The once-tranquil arc begins
to pulse and becomes irregular in shape. With luck, the observer,
by now completely enchanted, might see the magnificent “cor
ona,” which appears like streams of rippling light spreading out
from a common center high overhead.
In antiquity, Aristotle reported the appearance of such
“chasms, trenches and blood red colors” in the night sky. Later,
Plutarch repeated the report of an earlier fifth-century B.C. aurora
that was visible in Greece for seventy-five nights. Sightings like
these are rare in the Mediterranean, but in some far northern (or
southern) regions an auroral display can be seen every clear dark
night. Especially in these regions, the lore concerning the aurora
is rich and magical, reminding us of the rainbow’s power of en
chantment.
Among the Eskimo, the aurora is associated with the souls
of the dead. Those who suffered a voluntary or violent death
cross a narrow and dangerous bridge and enter heaven through
a hole. Once arrived, they play ball, and their game is seen
below as the aurora. The Eskimo word for the aurora is aksanirg,
meaning “ballplayer.” The Eskimo of northern Canada called
the aurora the “dance of the dead,” while in Greenland stillborn
or premature children are said to play ball in the heavens fol
lowing their untimely death, and so give rise to the aurora.
In Northern Europe, the Finns tell the tale of Repu, a fox
whose tail flashes fire but not heat—the aurora. The modem
Finnish word for the aurora belies its origins in the tale of Repu;
revontuli means “fox fire.”
As a prognosticator, unusually dramatic auroral displays have
been almost universally taken as omens of war and disaster.
Even in the twentieth century, the interpretation of the Miracle
of Our Lady of Fatima, when the Virgin appeared to three chil
dren on May 13, 1917, is connected with the aurora. The Virgin
241
CATCHING THE LIGHT
The aurora borealis.
242
Quantum Theory by Candlelight
told the children “when you will see a night illumined by an
unknown light, know that the chastisement of the world is at
hand.” On January 25, 1938, an unusually brilliant auroral
display appeared that was widely taken to foretell Hitler’s in
vasion of Austria three months later.
WHEN the Finns held that Repu’s tail flashed bright fire but
not heat, they were not far off the mark. Unlike the light from
a candle flame or other incandescent sources, the light of the
aurora is not due to the glow of a heated body, but is a “cold”
light much more akin to the light given off by a neon sign or
Galileo’s solar sponge, than by a candle. Like lightning, auroral
light can be sent through a slit and prism to produce spectra.
It, too, shows discrete luminous lines similar to those seen in
lightning spectra, and both are like those investigated by Kirch
hoff and Bunsen.
The physics behind the brilliant red light of a neon sign can
help us understand the light of lightning and aurora. In both
cases their light is produced by electricity passing through a
rarified gas atmosphere. If the gas is neon, the light is red.
Other gases give other colors when electricity passes through
them, and as Kirchhoff and Bunsen verified, the spectrum of
each is unique. The source of electricity for a neon sign is
obvious, but if the aurora and lightning are similar, where are
the dynamos that power them?
If, as he watched for sunspots, Galileo had simultaneously
studied the night sky in Norway, he would have noticed a cor
relation. The intensity of the auroral display corresponds to
sunspot activity with a two-day delay between them. The blem
ishes of the sun are huge eruptions that send an energetic stream
of charged particles racing toward the earth.
That stream is channeled by the magnetic field of the earth
to the poles, and the aurora shines out when that dark solar
243
CATCHING THE LIGHT
Sunspots, the unsightly imperfections discovered by Galileo.
stream strikes the earth’s upper atmosphere. The earth’s invis
ible magnetic forces draw down a dark solar wind, brightening
the long polar night with fox fire.
Planck’s magisterial analysis of light alone, even with the
hypothesis of the quantization of light energy, is insufficient to
account for lightning, or auroral light. Something more is
needed, namely a new theory of matter. The aurora poses new
questions concerning the structure of the atom and its production
of light. For this we turn to Niels Bohr.
Born IN Denmark in 1885, Bohr left Copenhagen for England
in 1911, immediately after receiving his doctorate. He first
244
Quantum Theory by Candlelight
joined Sir J. J. Thomson (discoverer of the electron) at the Cav
endish Laboratory in Cambridge, England, and then went on to
work with Sir Ernest Rutherford in his laboratory at Manchester.
Rutherford was busy at the time with his famous alpha-particle
experiments that would lead him to a picture of the “nuclear”
atom in which most of the atom’s mass is concentrated at its
center, in a positively charged nucleus. Fresh from collaboration
with these men, Bohr returned home in 1913, and shortly af
terward turned his attention away from nuclear structure and
toward the problem of light.
Study of the spectral lines of the elements had, by 1913,
developed into an active area called spectroscopy. Thousands
of lines crowded spectroscopic tables without any apparent or
der. One day in casual conversation with a colleague, Bohr
learned that the physicist Rydberg had found a formula that
predicted many of the prominent spectral lines, but it completely
lacked any theoretical base. He had discovered a pattern within
the jumble, but still no reasons existed for it. Bohr quickly
recognized that by combining Rutherford’s nuclear atom with
the Planck hypothesis of discontinuity, he could derive Ryd
berg’s formula and so explain the detailed appearance of line
spectra, such as appear in the aurora or lightning. But as with
Planck, the derivation came only at a very considerable cost to
the presumptions of nineteenth-century physics. Bohr’s atom
was to be understood as like a planetary system with electrons
circling a central nucleus in concentric orbits. This much could
be thought of along traditional lines. The transitions, however,
from one orbit to another needed to take place discontinuously
as “quantum jumps.” With each quantum jump, a quantum of
light or photon was given off. Its color or frequency could not
be given by a classical analysis of the atom’s motion in any
sense, but by Planck’s famous relationship: E = hv.
Physicists in England and Germany responded to Bohr’s sug
gestions incredulously, but as James Jeans remarked to col-
245
CATCHING THE LIGHT
leagues, although it might have nothing else in its favor, it did
possess the “very weighty justification of success.” Planck’s
quantum was irrepressible. Here was an entirely new domain
of application for the idea. With Bohr’s theory much could be
explained, even if no detailed mathematical theory yet existed.
When the young Wemer Heisenberg from Munich later suc
ceeded in developing such a theory, Bohr’s planetlike, electron
orbits disappeared from the atom, which became a stranger
object still. Very little of the traditional, mechanistic world
conception seemed left standing. Atoms were, after all, the
fundamental building blocks of all matter, and yet nearly every
recognizable feature of them disappeared before careful scru
tiny. The electron has no orbit, no real position in the atom;
light is emitted during discontinuous quantum jumps; how was
one to think such a world? It seemed impossible!
When Einstein first heard Heisenberg present his new theory
of quantum mechanics in the famous Berlin colloquium series
hosted by Planck, he was shaken.14 Afterward, the great man
invited Heisenberg home to talk. Immediately, Einstein com
plained that in Heisenberg’s formulation the concept of the
“electron path” had no place! Heisenberg, who had read Ein
stein’s work on relativity carefully, quoted Einstein back to him,
pointing out that since we cannot, in fact, ever observe such a
path, it made no sense to introduce the concept into the theory.
(Einstein had made similar arguments when advancing parts of
his relativity theory.) Much to Heisenberg’s amazement, Ein
stein replied, “Perhaps I did use such a philosophy earlier, and
also wrote it, but it is nonsense all the same.” If the loss of a
path for the electron was bad, the notion of a quantum jump
was, if anything, even less satisfactory to Einstein. There seemed
to him no physical basis for understanding Heisenberg’s quan
tum theory. It provided a formalism and predictive power, but
no explanatory value in the usual sense. Against this he rebelled
for the remainder of his life. The boldest thinker of the decade
246
Quantum Theory by Candlelight
balked once the full implications of quantum theory, which he
himself had helped create, were known. Bohr, by contrast, began
cautiously but gradually to assume a more and more radical
position as experiments suggested it.
Bohr WAS ENTIRELY Einstein’s peer, and a man of enormous
charm and personal force. When Einstein finally met him in
1920, he felt the power of Bohr’s personality: “Not often in my life
has a human being caused me such joy by his sheer presence. . .
[and to the physicist Ehrenfest] 1 am as much in love with him
as you are. He is like an extremely sensitive child who moves
around the world in a sort of trance.”15 Bohr, like Planck, was
reluctant to give up the great beauty of classical field theory for
an atomistic conception of light. For twenty years, Bohr resisted
the concept of particulate light with every weapon in his arsenal,
giving up even the sacred notions of causality, and the conser
vation of energy and momentum in individual events, if only he
could avoid the conception of light quanta. He, again like
Planck, argued that the locus of the problem was not light, but
matter. The magnificent edifice of electromagnetic theory should
not be modified, but rather our conception of substance. It
should be reconceived along lines advocated in his famous paper
of 1913.
What one detects as a granularity of light is, he argued, not
a feature of light but only the signature of its origins. An analogy
can make this clear. A wonderful yet ridiculous photograph of
Einstein exists, taken during a visit he made to the American
southwest. It shows him in a broad-brimmed sombrero and pon
cho, standing in front of an adobe hut surrounded by native
southwest Indians. He is grinning impishly, his unkempt white
hair sprouting out from under his hat, and his eyes twinkling.
He was in the southwest, but no matter how you dressed him,
or in what company you put him, he was still Einstein. Similarly,
247
CATCHING THE LIGHT
when light is emitted from matter, it carries away with it the
costume of the source. Spectral lines, for example, clearly re
veal, in the arrangement of their colors, the particular element
from which they were emitted. One should not, however, confuse
the present dress for the essence of free light. The being of light
may bear the imprint or signature of the source, but light left
to itself is a continuous wave, the argument went.
Bohr’s view, one which has persisted with many successes
until this day,10 conceived of matter as atomic and quantized,
but insisted that light as it traveled between emitter and detector,
that is when left to itself, was a pure electromagnetic wave.
The drama heightened when truly convincing experimental
evidence for light quanta appeared in 1922. A. H. Compton
scattered very high energy light in the form of X rays from free
electrons. The effect was immediately intelligible if one thought
of the collision between the light and the electron as if it were
a collision between two billiard balls, one being the electron
and the other a quantum of light. The wave theory seemed hard-
pressed to give any account of the effect at all. The tide was
clearly turning, and an atomistic conception of light was gaining
ground. As the superb Munich physicist Arnold Sommerfeld put
it, Compton’s discovery “sounded the death knell of the wave
theory of radiation.”
Still, Bohr resisted. Together with Kramers and the young Har
vard graduate Slater, a theory of the Compton effect was devised
that saved the view of light as a wave. However, in order to retain
the wave theory, Bohr made an enormous concession. He gave up
strict causality, as well as the conservation of energy and momen
tum for individual quantum events. Conservation would occur
only on average, statistically, and causality only overall. This was
an extraordinary compromise consciously made, one Einstein
protested adamantly when he heard of it. Causality had been sa
cred for centuries, how could it be thrown out in any way whatso
ever? So, too, with the conservation of energy.
248
Quantum Theory by Candlelight
The theory of Bohr, Kramers, and Slater could be tested by
a refinement of Compton’s experiment in which X rays are scat
tered off electrons—it failed. Although his theory failed, Bohr
had dared to challenge the bedrock classical conceptions of
causality and energy conservation. Once sounded, the challenge
refused to die. Was the world a causal chain from beginning to
end, or did quantum theory require a looser conception of ne
cessity? In the first decades of the century, these were questions
debated in the widest circles, far beyond the confines of physics
laboratories.
Living Philosophy
Berlin at the turn of the century was a heady place. While Max
Planck was occupied at the Berlin University with the invention
of quantum theory, not far away at the old Cafe des Westens,
prominent writers, artists, and intellectuals gathered every eve
ning to discuss Van Gogh, Nietzsche, Freud, and to sow the
seeds of revolution. The poet Emst Blass experienced life in
these years as “a spirited battle against the soullessness, the
deadness, laziness, and meanness of the philistine world.”17
Nineteenth-century materialism was under attack on every front.
Quantum theory and relativity would undermine mechanistic
explanations, artists were discontent with the constraints of ac
ademic art, and intellectuals were announcing a countercultural
revolution in ideas.
New attitudes rejected the arid mechanisms of the previous
age, and longed to replace them with a more vital Lebensphi-
losophie, or living philosophy of the world. Physics, never per
formed in cultural isolation, shared in the spirit of the times.
One eminent professor of physics said in a 1925 address, “It is
interesting to observe that even physics, a discipline rigorously
bound to the results of experiment, is led into paths which run
249
CATCHING THE LIGHT
perfectly parallel to the paths of the intellectual movements in
other areas of modem life.”18
Together with the revolutions in physics occurred a revolution
in sensibilities, but the winds of change did not affect everyone
equally. The first quarter of the twentieth century was often a
pitched battle between the voices of tradition and those of a new
life. Our chronicle of light reflects these struggles. Planck, Ein
stein, and Bohr will fight to protect the traditional underpinnings
of science while simultaneously advancing revolutionary un
derstandings of light. Artists and spiritual thinkers of the period,
while sometimes appreciative of science’s grand accomplish
ments, nonetheless criticized its one-sidedness and longed to
incarnate Emerson’s dream, that is, to unite severe science with
a poetic and spiritual vision.
In Paris, Robert Delaunay painted directly out of color and
light, juxtaposing color forms with only a hint of realism. In
1912 Klee translated Delaunay’s essay “On Light” for the
expressionist art magazine Der Sturm. Delaunay’s words in that
essay could have been Klee’s or Kandinsky’s: “So long as art
is subservient to objects, it remains description, literature. ...”
Rather should light itself be “treated as an independent means
of representation.”19 These artists turned their trade away from
depiction of sense objects and toward what had, until then,
remained unrepresented and extrasensory—the lights of nature
and mind.
Thus, with the dawn of the twentieth century, artistic and
spiritual currents were astir that reanimated and recast the an
cient and sacred lineage of light. Artists courageously explored
the new light’s first feeble glimmerings, while spiritual philos
ophers such as Steiner pursued light along pathways of medi
tation. The range of views alive during the first twenty-five years
of the century was incredible, and the relationships between
them were sometimes explosive. We have inherited all it at
tempted, both cultural and countercultural.
250
Quantum Theory by Candlelight
LIKE A RHIZOME whose tendrils stretch out from invisible im
mortal roots, the metaphysics of light had lain buried beneath
the nineteenth century’s scientific achievements; yet it had gone
on working still. When the soil overhead grew thin, an unsus
pected stem reached upward, toward the light, and blossomed
at the dawn of our century in the arts, literature, and in visionary
philosophy. W. B. Yeats’s words could well have been written
to describe the renewed inspiration offered by a metaphysics of
light among artists, poets, and philosophers of the time.
The shapes of beauty haunting our moments of inspiration
[are] a people older than the world, citizens of eternity,
appearing and reappearing in the minds of artists and or
poets.. . and because beings, none the less symbols, blos
soms, as it were, growing from invisible immortal roots,
hands, as it were, pointing the way into some divine laby
rinth.
Until the First World War, the two worlds of light—science
and spirituality—develop in relative isolation from each other.
Following the war’s devastation, however, they confront one
another with renewed intensity. Some will plead for union, others
for segregation, and the figure of light will bear the impress of
those controversies.
Crippled by the devastation of the First World War, and the
harsh terms of the Versailles treaty that ended it, Germany’s
economy was in a state of collapse. Many of the most talented men
of the new movement in science, art, and religion had perished in
the trenches. Among the ruins of Europe, living philosophy
gained momentum. The flames of discontent were, however, also
being fanned by early factions of Hitler’s fascist corps, who fed on
the popular feelings of resentment. Reactionary sentiments grew
and Hitler managed to divert some of the countercultural move-
251
CATCHING THE LIGHT
merit’s impetus away from its original intentions and used it for
his ends. Prominent individuals unwittingly added their own rea
soned voices to the growing tide of repression.
The smoke of many fires darkened the sky. During the years of
Hitler’s rise to power, Einstein’s theory of relativity was labeled a
“Jewish theory” and Einstein resigned his position at Planck’s
Kaiser Wilhelm Institute in Berlin. Planck, von Laue, and Hei
senberg maintained a precarious existence within the German
scientific community during the Third Reich, struggling to retain
a shred of humanity in an increasingly insane world. During the
war years, Planck, now very old, took up lecturing tirelessly on
his understanding of the relationship and division between sci
ence and religion. Still, the furnaces of hatred belched forth their
poisonous vapors. Planck watched helplessly as his own Kaiser
Wilhelm Institute was “cleansed” of Jewish scientists, and then
used to advance Hitler’s dream of a pure Aryan world order. The
abstract ideal of science disconnected from the values of its soci
ety fell apart. Good science is good not because it is value free,
but because it cherishes noble values.
On February 15, 1944, a massive Allied bombing raid on
Berlin obliterated the outlying town of Griinewald and with it
Planck’s home. All his meticulously preserved letters and doc
uments were destroyed in the conflagration. Finally, later that
same year, his beloved son Erwin, with whom he had walked
alone in the Griinewald forest at his first expansive moment of
discovery, was arrested by the Gestapo for his role in the as
sassination attempt on Adolf Hitler. His father moved heaven
and earth to save Erwin’s life, to no avail. The pain of his loss
nearly killed him. It was a very, very dark time. Perhaps, as
von Baader felt, out of such darkness, coming swiftly like a bolt
of lightning, light can rise again.
252
CHAPTER 10
Of Relativity and
the Beautiful
It teems that the human mind has first to
construct forms independently before we can
find them in things.
Albert Einstein.
Unlike Planck or Bohr, Albert Einstein began his study of light
not with a candle flame or the aurora, nor with any outer phe
nomenon, but with a Gedanken experiment—an experiment he
could perform purely in thought. Einstein would conceive many
such thought experiments, trusting in each to his inner sense
for the truth they contained. In contrast to many skeptics, he
dared to believe “that pure thought can grasp reality, as the
ancients dreamed.”1 In his first thought experiment, Einstein
locked horns with light. So elusive to laboratory experiments,
perhaps pure thought would have better luck in grasping the
reality of light.
Einstein himself tells us that probably in early 1896, as a
sixteen-year-old, he had the impossible thought; if one runs
after a light wave with a velocity equal to that of light, then the
253
CATCHING THE LIGHT
light wave should stop. Traveling with a friend, you do not notice
the speed of the plane carrying you both. By running with light,
likewise, it would seem to stop. “However,” noted Einstein,
“something like that does not seem to exist!”2 He had arrived
at a marvelous paradox, one he could think, but to which he
saw no solution. One can imagine running at any speed, and
so accompanying any moving object. Yet light was different; its
travel seemed somehow special. Unmoving light? Such a thing
was impossible! Yet by thinking the impossible, a revolution
was bom.
The paradox of running with light would simply not let go of
Einstein’s imagination. Throughout his years as an undergrad
uate student, he turned it over and over in his mind, longing
to unravel the mystery of light, and together with it, the nature
of the supposed ether that carried light. For nearly ten years he
harbored the puzzle, vainly suggesting experiments to his in
structors, grappling with his paradox. During those years he
matured scientifically, and like the grain of sand in the growing
oyster, his question concerning the nature of light was the ir
ritating nucleus around which grew the pearl of relativity. Fi
nally, in 1905, during a year of amazing productivity, the being
of light finally revealed itself to him, at least partly. Under his
steady scrutiny a new thought form was shaped, one capable of
addressing his early question; it has become known as Einstein’s
special theory of relativity. The paradox of running with light
found an answer. Together with that answer came innumerable
predictions, many of which startled Einstein’s audience, and
continue to startle us still.
Planck had been reluctant to suggest that light might be
quantized into what later would be called photons. Unlike his
Prussian counterpart, Albert Einstein’s thinking was more dar
ing. He felt unconstrained by either traditional family values or
the habits of thinking characteristic of classical physics. When
the celebrated French mathematician Henri Poincard wrote a
254
Of Relativity and the Beautiful
letter of recommendation for Einstein in 1911, he emphasized
his uncanny openness to new ideas, and his extraordinary ability
to see all they imply. “One can, above all, admire in him the
facility with which he adapts himself to new concepts and draws
all the consequences from them... he is not one attached to
the principles of classical physics.”3 Indeed, as we have seen,
Einstein was one of the first to take Planck’s work seriously and
quickly to develop its implications.
Others at the close of the nineteenth century had been groping
toward a new view of light, space, and time, but in every case,
some vestige of the classical conception held them back. They
could not free themselves from a mechanical image of light, nor
from the framework of absolute space and absolute time. Far
aday’s ontology of force had been “solidified,” by accommo
dating it to ether theories. Throughout the nineteenth century,
science viewed light as a vibration in the universal, material
ether. It was a view that was failing, but still proved amazingly
persistent. Early on, Einstein recognized the problem and wrote,
“the incorporation of wave-optics into the mechanical picture
of the world was bound to arouse serious misgivings.”4 One such
misgiving must certainly have arisen from Lorentz’s proof at the
end of the nineteenth century that the rigidity of the ether would
not merely need to be large, but infinite, in order to support the
vibrations of light. As Einstein remarked, “The laws [of Maxwell]
were clear and simple, their mechanical interpretations clumsy
and contradictory.”5
As a consequence of such considerations, physicists on the
continent began to dematerialize the ether, stripping it of un
necessary physical attributes. Important steps were taken by the
period’s recognized master, Lorentz. He eliminated every me
chanical feature of the ether, save one. Gone were its mass and
resilience, but it still defined an absolute frame of reference. It
was immobile. All motion was to be measured relative to it.
From this frame, one had a “God’s eye view” in which all things
255
CATCHING THE LIGHT
could be seen as they really were. Einstein, but twenty-six years
old, questioned even this. He worked inwardly, theoretically,
forging new ways of thinking about old things such as light, the
ether, space, time, and causality. Once the interior forms were
perfected, they could be compared to outer phenomena to see
if they could withstand the stringent demands of experiment.
Of the many rich veins we might follow into Einstein’s theory
of relativity, none glitters more brightly than light. As we follow
it, two revolutionary features of the theory will quite naturally
occupy us: the complete collapse of the hypothetical ether, and
the unique significance of the speed of light. They reflect the
two postulates, as Einstein called them, on which the special
theory of relativity was founded.
Relativity and Faraday’s Archetype
For the rest of my life I will reflect on what light is!
Albert Einstein, c. 1917
In the opening paragraphs to his famous 1905 paper, Albert
Einstein moved dramatically beyond the views of British phys
icists, and went further even than his continental colleagues,
when he wrote, “The introduction of a ‘luminiferous ether’ will
prove to be superfluous. . . .”6 Not only did Einstein suggest
that previous scientists were wrong about their substantial ether,
but even the totally ephemeral ether of the Dutch physicist
Lorentz was unnecessary. In that remarkable paper, Einstein
successfully developed a framework for physics that did not
possess a unique frame of reference against which to measure
motion. All motion was to be judged only with reference to other
objects; no viewpoint was privileged with the designation of
absolute rest. Moreover, within each frame of reference, all of
the laws ofphysics applied equally well. This was Einstein’s first
256
Of Relativity and the Beautiful
postulate, which he called “the principle of relativity.” To some
one at home in one frame of reference, everything was explicable
according to the very same laws used in every other inertial
frame. It seemed an innocent assumption, yet it led to radical
changes in our understanding of the world, including the ether.
Since all frames were perfectly equivalent, there was no need
for a single absolute frame of reference provided by the ether.
In order to introduce his pivotal principle of relativity, Ein
stein turned to the archetypal phenomenon of electromagnetic
induction discovered by that fertile experimenter Michael Far
aday. This same experiment had first implied a possible con
nection between light and electric disturbances for Faraday,
already a revolutionary association. Now it was pressed into
service by Einstein for a second and equally significant end.
Recall that when a magnet is moved relative to a coil of wire,
a current arises in the coil. But, asked Einstein, what is “really”
moving in such an experiment, the magnet or the coil?
To make the situation clearer, let us perform a thought ex
periment. Imagine a doughnut-shaped space station floating
freely in space. Its metal skin is a good conductor of electricity.
Toward it speeds a spaceship, ordinary in all things except that
it has been made into a huge, cylindrical magnet. One end of
it is a north magnetic pole, the other a south pole. The magnetic
spaceship approaches the station and flashes through the central
hole of the space station at great speed. On the station a large
current surge is detected in the metal skin. Why? According to
Faraday’s law, the surge is due to an electricfield induced around
the circular station by the moving magnet. Calculations are made
by space-station scientists, and yes, experiment agrees with
theory. But now shift to the viewpoint of those aboard the space
ship. They have not been moving at all, and toward them rushes
the space station. As it flashes by they also correctly predict a
current surge in the station of exactly the right kind, but their
explanation of its origins differs completely from that of scientists
257
CATCHING THE LIGHT
on the space station. To those in the ship, the current is due
to the movement of charges (electrons) in the skin of the station
acted on not by an electric field (as those on the station maintain)
but purely due to magnetic forces. Who is correct? Did the
current arise by the action of electric or magnetic forces?
Clearly, there is no way to tell. Nor does it matter for practical
purposes which view we take, because in either case a current
arises. There is, however, an enormous difference in what we
say causes the current. From one perspective, Maxwell’s elec
tromagnetic theory credits the current to the action of a magnetic
field, while in the other case, it arises from the action of an
electrical field. That is, a physicist sitting in the space station
(which plays the role of the coil in Faraday’s induction exper
iment of chapter 6) will give one explanation for the current
(calling it an induced electric field effect), while a second riding
on the magnetic spaceship will give a very different account
(saying it is due to the motion of charges in a magnetic field).
They will both be using the same theory of electromagnetism,
but the explanations they offer will be fundamentally different.
How can this be?
Faraday’s coil/magnet experiment provides us with a beau
tiful, indeed, an archetypal instance of relativity. What one
observer judges to be caused by electric fields, another observer
in motion relative to the first will judge as due to magnetic fields,
or some combination of both. Faraday’s simple experiment is,
in the hands of Einstein, the basis for another revolution in our
understanding of light and causality. First, he will use it to
dismiss absolute motion and with it, the ether. Second, it re
quires us also to view electric and magnetic fields as related
relativistic quantities. They are not “out there” like some kind
of surrogate ether, but transform into one another according to
Einstein’s equations. Accounts will, therefore, differ for physical
effects such as current production, but a good theory (like Max
well’s) can be used in any reference frame with equal success.
258
Of Relativity and the Beautiful
Just this universal applicability makes the ether unnecessary.
The ether had always provided the absolute frame of reference
against which all motion was to be judged and the “true” theory
applied. Now one no longer needed to say what was “really”
moving, the magnet or the coil; it just doesn’t matter, so long
as one is willing to give up the goal of a single true account for
physical events. Each observer can give his own perfectly con
sistent causal account of why things happen. However, what
each says “causes” an event (for example, the current in the
coil), or even the timing of the events, these will now depend
on their relative states of motion.7 There is no “God’s eye view.”
Accept this, and the accession is a major one, and all preferred
frames vanish, including that offered by the ether. That per
sistent artifact of the materialistic imagination, the ether, is
finally gone!
Wait! Einstein might have no need for the ether to explain
electrical and magnetic effects, but what then is light? All wave
theories supposed it. Maxwell and those following him took light
to be an electromagnetic vibration in the luminiferous ether,
propagating according to his four equations. If light is a wave,
but there is no ether, then what is waving? Try to imagine sound
without the motions of air. Yet this is just the avenue Einstein
was offering for light. Einstein’s suggestion received experi
mental support during these same years. The consequences for
our scientific understanding of light are enormous. Light is to
be an electromagnetic wave, but without a material medium to
support its motion. Light’s last vestige of materiality, its physical
body so to speak, had been taken from it.
With the ether gone, what remains? Perhaps the electric and
magnetic fields themselves are real even if the ether is not. No,
Faraday’s archetypal experiment and the principle of relativity
clearly demonstrate that it is wrong to think of the electric and
magnetic field at a particular time and place as having a single
true value. Observers in motion with respect to one another
259
CATCHING THE LIGHT
measure very different values for the fields at the same point.8
Faraday’s field concept gets us closer to light, but in using his
idea we cannot separate it from ourselves. Our account will
reflect our bias that we are stationary and the world moves around
us. Ptolemy took his viewpoint on the universe to be the earth,
Copernicus chose the sun; the accounts they gave of the motions
of the heavenly bodies were equally accurate but very different.
Likewise, relativistically we are always implicated in every mea
surement. Physicists must always specify what frame of refer
ence they are assuming for their calculations. It does not matter
in principle what frame they choose (although in practice one
frame is usually much easier to work in than another), but the
story they tell will be colored throughout by the choice. If one
forgets this lesson from relativity, then one runs the danger of
slipping back into a naive realism that takes one view and
assumes it to be true universally.
Alternatively, we might turn hopefully to quantum theory for
consolation. If light waves are now problematic, then return to
a corpuscular view. But while quantum theory speaks of a par
ticlelike nature to light, the photon remains massless and pos
sesses, in addition, its own set of unusual properties. It
implicates the observer even more deeply through Heisenberg’s
uncertainty relations. No, quantum theory can deepen the mys
tery, but it cannot relieve it. Light is not substantial, at least
not in the form of a massive particle a la Newton, nor even in
the form of undulations through a material medium A la Euler
or Maxwell. If one conceives of the universe as matter or its
movement, light is the exception that shatters that prejudice.
The nature of light cannot be reduced to matter or its motions;
it is its own thing.
Pause to consider the history of light once again, now in
cluding the work of Einstein, for we have come far. First ex
perienced as the gaze of God, light was a spirit reality, eternal
and omnipresent. From the time of the Greeks to those of New-
260
Of Relativity and the Beautiful
ton, light became first geometrical and then material. With Far
aday and Maxwell the era of electromagnetic light dawned, but
light still carried along with it the trappings of its recent past,
a material medium—the ether. Aside from Faraday, scientists
simply could not imagine light—a real thing—apart from ma
teriality. Their thinking did not allow it; matter was synonymous
with reality. Then in 1905, Einstein suggested an alternative.
The cost was high: it granted no absolute meaning to motion,
declared the ether superfluous, allowed no single privileged
causal account for physical phenomena, implicated us in all
observations, and predicted a host of attendant effects such as
length contraction and time dilation. Yet for all that was lost,
Einstein provided much in return. If there were now many causal
accounts, he gave physicists the means for translating one to
another, accomplished by what is called a Lorentz transforma
tion. Later, he also offered the basis for seeing a deep connection
between matter and energy (E = me2). Although not material,
light could be thought of as a form of energy whose behavior in
this world is imaged in Maxwell’s electromagnetic field equa
tions. Rationality remained, simple or subtle matter did not.
COPERNICUS had FORCED human consciousness from its ancient
haunts to the sun. The earth was no longer the still point around
which the cosmos moved; the earth itself moved around the sun,
and it was but an insignificant star in the galaxy, which in turn
was but one among countless others. No single material place
was the center and focal point of God’s creation, but all were
created equal. Einstein went further. Now there was not even
a place of rest, everything everywhere was equally in motion.
With Copernicus, location had become relative; with Einstein,
motion. Together they emancipated human consciousness, al
though many longed for the security of the earlier view. One by
one the traditional, outer props of science, religion, and society
261
CATCHING THE LIGHT
were being taken away. Humanity now stands alone and alien
ated in the limitless universe. Bereft of parents and far from
home, each of us has now to become a center unto himself or
herself, and find the spiritual strength to hang unsupported in
the void, drawing support from within not from without.
Amid this extraordinary flux of new ideas, the loss of absolute
space, time, and the ether, there remained but one constant of
unique significance, one “truth” that remained independent of
all reference frames—the speed of light! Around it swirled all
other velocities, positions, and times, but light was, as Einstein
wrote in 1905, “always propagated in empty space with a definite
velocity c which is independent of the state of motion of the
emitting body.” This was Einstein’s second postulate and pillar
for his theory of relativity.
The Speed of Sight
No corporeal substance can be so subtle and swift as this.
William of Conches,
12th century
The night sky of ancient Greece did not suffer from the obscuring
elements of city lights and pollution. In town and field, the stars
were a nightly presence whose beauty and order were the spring
board of science. The earliest measurements of the speed of
light were made by observers who, like so many before and after
them, turned their eyes to the stars and wondered.
If, as Empedocles, Plato, and many others held, during the
act of sight something passes from the eye to the object seen,
then in seeing the distant stars, something must pass through
the greatest imaginable distance, and do so in an instant. After
all, on opening our eyes, the entire world from horizon to horizon,
and out to the most remote star or planet, immediately springs
262
Of Relativity and the Beautiful
into view. Whatever travels there must do so at the highest
speed. The measurement implicit in this argument, a common
one in antiquity, was not actually a measurement of the speed
of light, but really of the “speed of sight.” The concern of ancient
authors with the speed of perception continued for two thousand
years. John Pecham, the most widely read medieval optician,
wrote only of it, and not the speed of light, in his important
work Perspectiva Communis. Still the speed of light itself was
treated implicitly or explicitly from Plato on.9
Aristotle berated Empedocles and Plato for their views of
sight.10 In his view nothing streamed from either eye or object.
Light was, as we have seen, a state or quality of the medium,
and so just as water may freeze all at once, the transformation
of the “potentially transparent” to the state of actual transpar
ency, which is light, can occur everywhere simultaneously. It
is simply wrong, according to Aristotle, to think of light as
propagating at all. If his position is difficult to understand,
consider a student who is, as are many, confused by Aristotle’s
claim that light does not propagate. Call the student’s state of
confusion darkness. As the instructor, firelike in his enthusiastic
explanation of the puzzle, goes on, the student’s face suddenly
“lights up.” The student has changed from a state of confusion
to one of comprehension, and has done so instantly. Similarly,
when the world “lights up,” propagation is not a logical neces
sity. Light may well be a universal change of state.
The general view that light travels at infinite speed was taken
up, for example, and repeated by Augustine: a ray of light
“manifestly passes through those great and immense spaces
immediately, in a pulse beat,” he wrote. Or elsewhere, “our
visual ray does not reach near objects sooner and distant objects
later, but passes through both intervals with the same swift
ness.”11 Bear in mind Augustine’s close association of God and
Christ with light. It made eminent theological sense that God’s
presence, near and far, be simultaneous.
263
CATCHING THE LIGHT
By the twelfth century, Plato’s influence had made its way to
the Chartres masters, influencing their understanding of light
and vision. According to William of Conches, the visual fire
flowing from the eyes is subtle and swift: “no corporeal substance
can be so subtle and swift as this. Thus it must be instantly
here and instantly there [among the stars].” Still, William rea
soned that because the visual fire was, in his view, a material
substance, its movement required time. What is divine can travel
at infinite speed, and so be all places at once, but what is
material must travel through space and in time. Once again
spiritual and physical considerations intertwine.
While light was taken by most to travel instantly throughout
space, dissidents like William of Conches felt that although
light might travel swiftly, it still required time to do so. Alhazen
and Roger Bacon were two of the earliest to hold this opinion.
By the time of Rend Descartes, the controversy was well estab
lished, with Descartes taking the side of instantaneous propa
gation in the strongest possible terms.
Recall Descartes’s plenum and his model of vision in which
sight is explained by analogy to a blind man with his cane. If
all space is filled with a material medium, then, as with the
blind man’s rigid stick, a shock or movement at one end will
instantaneously appear as a shock or movement at the other end.
Descartes felt so strongly about this point that he was willing to
hang the truth or falsity of his entire philosophical system on
the prediction. To an unknown friend he wrote in 1634 that
light “reaches our eyes from the luminous object in an instant;
and I would even add that for me this is so certain, that if it
could be proved false, I should be ready to confess that I know
absolutely nothing in philosophy.”12 In Paris, forty-one years
later the Danish astronomer Ole Rpmer demonstrated that Des
cartes was wrong. The completion of the syllogism is left to the
reader.
By careful observation of the moons of Jupiter, and using
Cassini’s new measurement of the planetary distances from the
264
Of Relativity and the Beautiful
sun, Rpmer was able to show that light travels at a finite, if
extraordinarily high speed. We have already seen the signifi
cance of that speed for Maxwell, leading him to suggest that
light was an electromagnetic wave. Its significance for the special
theory of relativity was greater still. In the words of the French
physicist Marie-Antoinette Tonnelat: “The second postulate of
Einstein profoundly modifies the status of the speed of light:
what was a kinematic and essentially relative entity now becomes
a phenomenon describable by an invariant law.”13
The goal of ever more precise measurement of this remarkable
quantity, the speed of light, has captured the imagination of
great experimentalists for the last three hundred years. The
American physicist Michelson dedicated the last twenty years
of his life to ever more exact measurements of light’s speed of
travel. His efforts culminated in a huge experiment involving a
three-foot-diameter steel tube one mile long. Sealed and evac
uated, light flashed back and forth through its airless corridor.
The results—a new value for the speed of light—reached the
impatient Michelson on his deathbed on May 7, 1931. He could
die, two days later, content.
Since then ever more precise methods have been devised to
measure the speed of light. The last used highly stabilized lasers
and special techniques for measuring ultrahigh optical frequen
cies. Measurement finally became so exact that the main in
accuracy in them was due to uncertainties in the world’s standard
unit of length: the meter. In 1983 it was therefore decided to
end the three-hundred-and-eight-year history of the measure
ment of the speed of light forever. Instead of defining a unit of
length, as had always been the case before, the speed of light
would be defined! Its value was taken to be the then-current
best-measured value of 299,792,458 meters per second. From
1983 on, the speed of light was no longer a quantity that could
be measured; rather it was defined as a matter of pure convention
to be the above value.
Having given up the definition of the meter for that of the
265
CATCHING THE LIGHT
speed of light, physicists had to go back and adjust the unit of
length so it would be in agreement with the new standard. The
meter, therefore, is now defined as the distance traveled by light
in vacuum during a time interval of 1/299 792 458 of a second.
The carpenter’s measuring tape is really so many light-seconds
long. Every ruler can now trace its pedigree back to the speed
of light. As Grosseteste imagined, light, as it spreads, gives rise
to space.
The ANCIENT QUESTION of the speed of light has been answered.
It is finite, and we have decided to call that finitude 299,792,458
meters per second. Never again will a researcher measure its
value. What had begun as the infinite speed of sight has become
the finite speed of light. Yet while experimentalists resolved one
issue, relativity theory opened another, taking the great but finite
speed of light, and declared it unique.
The Uniquely Universal Speed
IFe shall find in what follows that the velocity of light in
our theory plays the role, physically, of an infinitely great
velocity.'*
Einstein, 1905
Drop a small stone into still water and watch the ripples expand.
When the waters have quieted, skip a rock on the water’s sur
face, and each time the stone touches, an expanding ring of
waves grows outward. The speed at which the rings expand over
the water’s surface is exactly the same in both cases. The speed
of water waves does not depend on the speed of the stone,
dropped or tossed, but is determined only by the properties of
water. Similarly with sound; the air alone, not the source speed,
determines the speed of sound. All waves propagate at speeds
266
Of Relativity and the Beautiful
set solely by the media that support them. Therefore, when
Einstein put down his second postulate—“the speed of light is
independent of the speed of its source”—it might seem that he
was saying nothing new. In 1905 (almost) everyone thought light
was a wave, so to them his second postulate was obviously true.
But as we have seen, in that same paper Einstein had dismissed
the ether; the very basis for the unique determination of the
speed of light was therefore gone! Until 1905 wave velocities
could be independent of source velocity only because they were
completely fixed by the media in which they moved (water, air,
ether, etc.). With no medium left for light waves, how could
light possibly “know” how fast to travel? Einstein’s reply was
startling. He entirely ignored the question, and asserted as a
postulate that the velocity of light was independent of source
velocity. Period. When this is taken together with his first pos
tulate, the principle of relativity, the consequences of his theory
for our conception of time and space were truly revolutionary.15
Think back to Einstein’s first thought experiment—running
with light. What the second and first postulates imply is that
not only can you never catch up to a light wave, but you can’t
even get close. Give your friend a flashlight, turn it on, and
measure the speed of light: 299,792,458 meters per second.
Ask him to keep measuring, just to make sure the speed of light
does not change. Now start running. For argument’s sake, let
us assume that relative to your friend with the flashlight, you
are running at ninety-nine percent the speed of light. Not know
ing relativity, he will think you have almost caught up to the
light wave. While running, you reach out and measure the speed
of light yourself: 299,792,458 meters per second! How can this
be? You are running at nearly the speed of light, and you
measure the light as traveling at the same speed as before. You
reason, it must “really” be going at twice the speed of light.
Your friend with the flashlight, however, keeps getting the same
value you do. What is going on?
Nothing else travels the way light does. All observers will
267
CATCHING THE LIGHT
measure its speed in vacuum to be 299,792,458 meters per
second. To make this possible we must change our conception
of space and time, but if we do so, everything hangs beautifully,
if strangely together. Is light really like this? Perhaps Einstein’s
theory is brilliant but wrong; theories have been disproved be
fore. In Einstein’s day, essentially no evidence one way or the
other existed to support or refute his theory. Since 1905, how
ever, extraordinarily precise experiments have been performed
that confirm it.
I remember well the 1977 experiment of Alain Brillet and
Jan Hall performed at the Joint Institute for Laboratory Astro
physics in Boulder, Colorado, where I was also working. Down
in the lowest basement of the laboratory, they suspended a
granite table, atop which was one of Jan Hall’s stabilized helium
neon lasers. They rotated the granite slab and laser slowly,
searching, as had Michelson and Morley, for an “ether wind”
as the earth plowed its way through the hypothetical ether. The
earth moves around the sun at about thirty thousand meters per
second. Therefore, if an ether existed, one might expect to detect
an ether wind of up to this speed. In the Brillet and Hall ex
periment, such a wind would have shown up as a small shift of
the laser light’s frequency of oscillation. Michelson and Morley’s
1887 experiment had found no ether wind down to 4,700 meters
per second. If it was smaller than this, their experiment could
not have detected it. Brillet and Hall also saw no evidence for
the ether, and their upper limit on the ether wind was fifteen
meters per second, which is still the best Michelson-Morley
style measurement to date.16 Other kinds of measurements have
lowered the highest possible value for the ether wind still further,
down to 0.05 meter per second! As Einstein’s first postulate
implies, no trace of a luminiferous ether has yet been found.
Similar developments have taken place in tests of the second
postulate—which says that the speed of light is independent of
source motion. All manner of atomic light sources have been
268
Of Relativity and the Beautiful
moved by experimentalists at speeds ranging from the very slow
to ninety-two percent the speed of light. In every case the mea
sured value of the speed of light was found to be the same.17
Referenced to nothing (that is, without an ether), the speed of
light is unvarying. If something comparable existed in spatial
terms, it would have the property that no matter where you went,
it was always the same distance away from you. You could get
neither closer to it, nor farther away from it; the separation
would be fixed. This is what infinity is like. No matter how far
toward it or away from it one travels, it remains the same infinite
distance away. As Einstein said, the speed of light plays the
role of an infinitely great velocity. Light has no place, but it
does have a speed and we are always separated from it by
299,792,458 meters per second.
Special relativity has been tested many times, and it has met
all challenges. Einstein’s run with light has truly transformed
our understanding of it. Einstein had dared to ask what light
would look like if one traveled beside it at the speed of light.
The implications of the answer have been vast. We might well
tum the question around and ask, what would the world itself
look like while running with light? What is the world like from
a “light’s eye view”?
As every schoolchild now knows, nothing but light can reach
the speed of light. No massive object, no matter how small its
mass, can attain this ultimate speed. One would have to de
materialize to go as fast as light. Like young Einstein, imagine
that you are light. Leaving your physical body behind, take
refuge in your light-body and wing through space. What would
you see?
It has become common knowledge since Einstein that as one’s
speed approaches that of the speed of light, space and time
change in significant and surprising ways.18 Lengths along the
direction of motion shrink and moving clocks run slow. Exactly
how these would change the appearance of things has been
269
CATCHING THE LIGHT
studied quite carefully in recent years, but I have never seen a
serious treatment of what would happen if one traveled at the
ultimate speed—the speed of light.
We do know that as one speeds up, the clocks we see flashing
past us tick slower and slower, and the shapes of objects distort.
Distances traveled seem to shorten—so-called length contrac
tion. What happens when the impossible finally is achieved and
we become light? Does time ultimately slow to a stop? To nat
uralists in 1911, Einstein remarked that what to us might be
centuries would, to a living organism, be “a mere instant,”
provided the organism travels with nearly the speed of light. ”
What about space? Does distance completely disappear? In his
seminal 1905 paper, Einstein declared that, “For v = c all mov
ing bodies—viewed from the ‘resting’ frame—shrivel up into
plane figures.” Everything is here and now, forever! In running
with light, are we led full circle back to light eternal and om
nipresent, outside of space and time? Or maybe Aristotle was
right, and light is a universal and instantaneous change of state,
at least when seen from the vantage point of light itself. By
musing steadfastly on puzzles like these, perhaps we, like Ein
stein, will uncover yet another image of light.
Einstein’s relativity, Planck’s quantum of light, Steiner’s an
gelic light, and Delaunay’s artistic light all arose contempora
neously. Why? Remembering C. S. Lewis, we can ask: what
were the questions most on the hearts of those living through
the years from 1900 to 1925, and what do they reveal about the
underlying tenor of their minds? During the previous three
hundred years, every outer source of security had been removed.
The guiding institutions of the past, whether sacred or secular,
had been called into question or, especially in the course of
war, destroyed. Simple faith no longer sufficed, existential ques
tions loomed large, and few adequate answers were forthcoming.
270
Of Relativity and the Beautiful
Relativity theory, from one angle, can be seen as completing
what was begun at the dawn of modem science by Copernicus
and Galileo. In leaving the safety of sacred and classical tra
ditions, we left much behind, but took with us still our body
(space) and heartbeat (time). Though alone, we could cling fast
to these. But no, according to relativity, even these were to be
offered up. Space and time are not absolutely given. A more
fluid structure and malleable rhythm is required. Thus every
sure material prop, even the abstract comforts of absolute space
and time, must be given over to the fire. We have faithfully
followed the lineage of light to the twentieth century, and our
place in the universe seems more completely meaningless than
ever before. Many might deny it, but every external underpin
ning of classical science has been withdrawn. What, in the end,
remains? Emerson’s refrain, penned decades earlier, sounds
clear and true: “Nothing is at last sacred but the integrity of
our own mind.”
The human spirit had, until the modem era, been cared for
and sheltered from such disconcerting knowledge by church or
state. Like the innocent youth Parzival, we had been raised in
idyllic ignorance of such harsh conceptions of the world. Only
when surprised by knights in glittering armor, whom he mistook
for angels, did Parzival leave home dressed as a fool, without
so much as a look back at his distraught mother. She died at
his departure, and he had much to learn. Year after year, Par
zival discovered his failures and stupidities, but never relin
quished his aspiration. Finally, utterly alone in the dark and
cold of the forest wilderness, a tattered figure, he dropped the
reins on his horse’s neck and, profoundly humbled, found his
way back to the Grail castle as its new king. Like Parzival, light
had first lived in an enchanted countryside as a favored and
divine child. Then, donning the armor of theory, it rode forth
to conquer all the phenomena of light. Early success bred ov
erconfidence, but not all phenomena fell before its lance. Can-
271
CATCHING THE LIGHT
dielight, the aurora, and lightning, to name but a few, stood
firm. The old armor was stripped off and another far more subtle
took its place, made of mercury instead of steel. The ramparts
of the Grail castle vanished from Mont Salvat; they reappeared
as the close-fitted plates of the human skull, the earthly sanc
tuary of the mind.
There, within that infinite interior space, Einstein created a
theory possessing deep inner consistency, that urges a new self-
reliance. With two postulates as anchor points, a brilliant
thought structure arose that violated common sense, but fostered
our courage to think devotedly. Einstein seemed endowed with
a kind of inner gyroscope that allowed him to keep his bearings
in every storm. I think his poise was but a reflection of his
faithfulness to thinking, and to a view that held nature to be an
expression of Intelligence. What Lebensphilosophie rejected was
not rationality, but the lifeless form it took during the late nine
teenth century. I would suggest that we see relativity theory as
a reflection of an evolutionary movement of the human psyche
toward true autonomy. The inflexible geometry of past science
reflected the restricted structure of its imaginative base. By
contrast the dynamic imagination of modem science liberates,
but it also endangers. The stability we once found without, we
must now find within. One of the most beautiful examples of
this change occurred around the world’s geometry. Not since
Brunelleschi invented linear perspective had the geometry of
human experience undergone such a radical shift.
In the decades leading up to the period of relativity theory,
the very architecture of space was revolutionized. Until then the
mathematical imagination, and with it all of scientific thinking,
had been dominated by a single book. No text, other than the
Bible, had done more to shape the thinking of the West than
Euclid’s Elements. Since the invention of printing alone, more
than one thousand editions have appeared. Yet the mathematical
framework the Elements espoused grants an unfounded privilege
272
Of Relativity and the Beautiful
to one view, excluding the very idea of non-Euclidean geome
tries. The roots of a more flexible attitude to geometry reach
back to the Renaissance creators of linear perspective, but the
development of their first insights into the modem discipline of
“projective geometry” had to await the work of great mathe
maticians such as Poncelet (1788-1867), Cayley (1821-95),
and Klein (1849-1925). By the time of Einstein, non-Euclidean
geometries and the even more comprehensive theory of projec
tive geometry had broken the grip of Euclid on mathematical
and spacial thinking, and a new imagination of space could be
bom.
To show something of the flavor offered by the new geometry,
consider the circle and cube of Euclidean geometry. Once given,
these forms are fixed. The curvature of a circle is absolutely
uniform and determined by its particular radius; similarly, the
comer angles of a cube are always ninety degrees and all sides
are of equal length. In projective geometries these invariants
disappear. The circle and cube can undergo an infinity of meta
morphic changes. The figure below exemplifies but one of the
possibilities; you must imagine the multitude of others not
drawn.
Olive Whicher’s lovely drawing allows us to see geometric
forms as crystallizations bom from light. Rays, like beacons,
stream through space in an orderly fashion; their intersections
defining and lengths embracing familiar geometrical figures. In
addition, these rays must be imagined in motion, the figures
ever-changing, so that all stasis disappears. The space of forms
is entirely mobile, in flux, a geometry of streaming metamorphic
life, and mathematics is alive with the vital force of the modem
imagination.
Now freed of Euclidean constraints, we have to support our
selves on the gossamer wings of the geometry that our own hands
have crafted. Seen in its light, the contours of continents we
overfly may change strangely from one moment to the next, but
273
CATCHING THE LIGHT
A projective geometric construction.
as Blake once wrote, “No bird soars too high if he soars with
his own wings.”
Light and the Architecture of
Space-Time
Structure is the giver of light.
Louis Kahn
The new architecture of space presupposes architects to imagine
it, and builders to build it. Like a modem gnostic, the architect
Louis Kahn was constantly attentive to light.20 Where silence
and light met, in the space formed by their union, he discovered
inspiration. Here was the Sanctuary of Art, the Treasury of the
Shadows. Within this sanctuary, he wrote, “the artist offers his
work to his art.”
274
Of Relativity and the Beautiful
Silence to Light
Light to Silence
The threshold of their crossing
is the Singularity
is Inspiration
(where the desire to express meets the possible)
is the Sanctuary of Art
is the Treasury of the Shadows
(Material casts shadows, shadows belong to light).
Silence broods. In it resides the restless, “unmeasurable de
sire to be.” When the desires of Silence meet Light, “the mea
surable, giver of all presence,” worlds arise. From these two
originates all that is made.
.. . all material in nature, the mountains and the streams
and the air and we, are made of Light which has been spent,
and this crumpled mass called material casts a shadow, and
the shadow belongs to Light.
So light is really the source of all being. And I said to
myself, when the world was an ooze without any kind of
shape or direction, the ooze was completely infiltrated with
the desire to express, which was a great congealment of
Joy, and desire was a solid front to make sight possible.
To make sight possible. Around and within that original con
gealment of Joy rayed Light, the source of all being. Spending
itself, it built up not only mountains, streams, and air, but also
the sighted creatures who see it. As Richard Wilbur writes, “It
was the sun that bored these two blue holes,” our eyes. Kahn’s
cosmogony was one of light, the world its offspring.
Architecture and the new geometry of space-time faced each
other in the first light of the new century, at once intellectually
liberated and spiritually inspired. Like figures shaken from a
textbook of projective geometry, buildings of glass, steel, and
concrete grew in the landscape. The organic architecture of
275
CATCHING THE LIGHT
Gaudi in Barcelona, Steiner in Switzerland, Le Corbusier in
France, caught in concrete the dynamic morphology of organic
forms. Bruno Taut’s “Glass House” allowed light to magically
penetrate even its structural walls through glass bricks. To invert
a thought of Kahn’s, light seems the giver of space.
The INTERPLAY OF light and space displayed yet another facet
in the hands of Einstein when he extended his special theory
of relativity to embrace the phenomenon of gravitation. His
general theory originated with what Einstein called the happiest
thought of his life.21
In NOVEMBER OF 1907, while sitting in a chair in the Beme
patent office where he worked, it suddenly occurred to Einstein,
“If a person falls freely he will not feel his own weight.” A
simple observation, but within it lay the gravitational analogue
of Faraday’s archetypal experiment of electromagnetic induc
tion. The relativity of electric and magnetic fields that had led
Einstein to his special theory here possessed an exact parallel
in gravitation, one that would lead him ultimately to his general
theory of relativity. Einstein’s “happiest thought” has since be
come enshrined as the “equivalence principle” of relativity the
ory. It states that gravitational attraction is experimentally
indistinguishable from steady acceleration. Confined to a room,
the pressure on your feet could as well be due to it accelerating
upward (as in an elevator or spaceship at blast-off), as to the
gravitational attraction of the earth.
Einstein was able to use the equivalence principle to predict
that the path followed by light would be bent as it passed a
massive body such as the sun. Arthur Eddington’s 1919 English
expedition to Principe Island off the coast of Spanish Guinea
succeeded in capturing in photographs just this effect. The
deflection was tiny, but in essential agreement with Einstein’s
276
t“
Planet
(a) (b)
Einstein’s principle of equivalence. To the astronauts inside, the expe
rience of acceleration is exactly like being at rest on a planet.
i
I
l
The bending of starlight by the mass of the sun, as predicted by Einstein.
The apparent change in a star’s position was first demonstrated by Ed
dington’s expedition.
277
CATCHING THE LIGHT
prediction of seven years earlier. Why should weightless light
bend as it travels through empty space simply under the action
of a massive body? If light had mass, then fine, but it does
not. Then why should it bend? Einstein had argued that since
light would appear to follow a curved path in an accelerated
room it would do the same in the equivalent gravitational
situation; what is true for one must be true for the other.
Therefore, the bending of light by planet, suns, and all massive
objects should occur.
No longer should we imagine light piercing its way through
space in straight lines, but rather as weaving through the
curved structure of “spacetime.” Just as electricity and mag
netism had been wedded together by the special theory of
relativity, space and time likewise became inseparably linked
in the reality of Einstein’s universe. The pathways of light
held a special place in the new theory, for while they curved,
they still mapped out the shortest distances or “null geodesics”
connecting one space-time point to another. If the sun and
starlight that fills all space could be made visible, we would
see the glistening, mobile, sculptural architecture of the cos
mos. Einstein even said that this immaterial architecture of
space-time could be understood as a kind of rehabilitated
ether. It differed utterly from earlier conceptions but provided
the malleable framework through which light sped and space
time was defined. He said, “According to the general theory
of relativity, space without ether is unthinkable; for in such
space there not only would be no propagation of light, but
also... no basis for space-time intervals in the physical sense.
But this ether may not be thought of as endowed with the qualities
of ponderable media. . . .”22
Dismissed at the opening of the century, a new understanding
of the ether has been suggested not only by Einstein, but also
by prominent physicists since. Their proposals are based not
only on relativity theory but also on their attempts to under-
278
Of Relativity and the Beautiful
stand quantum theory. The material ether of Kelvin is dead, but
for some the ether lives anew in a far more subtle, immaterial
guise.
DURING the FIRST decades of the twentieth century, Einstein’s
meditations on light never flagged. Nearly everyone who visited
him in later years seems to tell the same story. He had done
much to father the strange quantum of light, but he was con
vinced that quantum theory could not be the entire story. Many
were the revolutions of the first decades, but it seemed to him
that the demands of causality required a nonprobabilistic theory.
As he often said, “God does not play dice.” To Einstein the
quantum physics of the microworld was still incomplete, a frag
ment of the truth. We might think we have reached the end,
but we are only fooling ourselves. Even the concept Einstein
had introduced, the quantum of light or photon, could not be
understood properly in quantum theory. In 1951 he declared,
“All the fifty years of conscious brooding have brought me no
closer to the answer to the question, ‘What are light quanta?’
Of course today every rascal thinks he knows the answer, but
he is deluding himself.”23
It is forty years since Einstein warned against the arrogance
of scientific surety regarding light. Efforts to understand light
have not abated since his death, and yet the essence of light
remains an enigma.
We STILL LACK a picture for the photon. A remarkable feature
of relativity is that its predictions, including those it makes
about light, do not rely in any way on a model for light.
Relativity does not need to know whether light is a wave or
particle, it does not need to answer the question “What are
light quanta?” in order to predict the bending of light around
279
CATCHING THE LIGHT
the sun. Whatever light is, it must conform to the principle
of relativity. Much of relativity’s beauty and universality stem
from this remarkable feature of the theory, one which Einstein
appreciated deeply. How far can one go without a picture of
light, how many predictions can one make without knowing
what light really is?
In recent years, this avenue of thinking has been pursued,
especially by the remarkable late physicist Richard Feynman.
Following the path of beauty and not that of image, Feynman
stalked the phenomena of physics with an ancient and honorable
instrument, the confidence that perfection lay at the root of
existence. The dictates of beauty have commanded many over
the centuries, guiding the imagination of inquirers down to the
present. From the shapely attractions of space-time we turn to
the abstract beauty of perfection.
Following the Beautiful
It isn’t that a particle takes the path of least action but that
it smells all the paths in the neighborhood and chooses the
one that has the least action.
Richard Feynman24
Something of a legend in his own lifetime, Richard Feynman
was a brilliant young prankster at the secret Los Alamos labo
ratories where J. R. Oppenheimer guided his scientific team to
produce the first nuclear bomb. As a provocative member of the
congressional committee to investigate the Challenger disaster,
Feynman’s irrepressible genius revealed the tragic flaw of
booster design responsible for the mishap.
For someone of such brilliance and originality as Feynman,
what was it about physics that excited him and drew him into
the subject originally, then sustained his fascination with it for
280
Of Relativity and the Beautiful
decades? When all the hard things come easy, as they did for
Feynman, to what does one hold fast? The answer lay in a remark
made to him by his perceptive high-school physics teacher, Mr.
Bader. Noticing the boredom of his exceptional student, he gave
him something lovely to think about. In the vocabulary of phys
ics, Mr. Bader told Feynman that light, like all things, always
follows the path of the beautiful.
Bending Light, Again
While fishing, every child has noticed that when you dip a fishing
pole into the water, it seems to break. Put a spoon into a glass
of water, and when seen from the top, the spoon shows a kink
in it where none appeared before. We are not surprised that
when removed from the water, both the fishing pole and spoon
regain their pristine forms. We are accustomed to such behavior
in spoons and fishing rods. Perhaps we even know something
about refraction; that light (or the visual ray) bends when it
passes from one medium to another: from air into water, for
instance.
Why does light do this? Why does it bend when entering
water from the air? Or, to enlarge the question, why does it
travel just as it does in general, sometimes bending just so
much, other times reflecting at exactly a certain angle? Sim
ilarly, why does a stone or an electron, or anything, travel in
exactly the way it does? What stands behind the behavior of
all physical systems, behind all the laws of physics? It seems
hugely presumptuous even to ask such a grand question, and
yet at a very powerful level, there is a single fascinating
answer. It was about this that Mr. Bader spoke to the young
Feynman that day, and I believe it was this that held Feynman
for a lifetime.
The actions of light can once again lead us to the heart of
281
CATCHING THE LIGHT
things. By puzzling with others over the simple bending and
reflecting of light, we will be led in the end to Bader, Feynman,
and, surprisingly, to the contemporary challenge of quantum
mechanics. The tale begins in the exotic location of second-
century Egypt.
The FIRST SYSTEMATIC observations of refraction seem to have
been made by the Alexandrian astronomer Claudius Ptolemy in
the second century A.D. From his important treatise on optics
there has survived what appears to be a table of data giving the
angles of light in air and water.
ANGLE IN AIR ANGLE IN WATER
10 8
20 15>/2
30 2214
40 29
50 35
60 4014
70 4514
80 50
At first blush, the numbers all seem reasonable. The angles
in water are all less than the angles in air as they should be;
the overall pattern seems right. One becomes suspicious, how
ever, when one considers the reported accuracy of the measured
angles in water. To measure to half a degree of arc is pretty
remarkable for the second century A.D. Look more closely at
these angles, and subtract adjacent numbers
ibers from one another.
If you do so, the following results are obtained.
282
Of Relativity and the Beautiful
Angle
in air
AIR
WATER
Angle \
in water
Refraction at an air-water surface.
ANGLE IN AIR ANGLE IN WATER DIFFERENCES
10 8 7>/2
20 1516 7
30 22 16 616
40 29 6
50 35 516
60 4016 5
70 4516 416
80 50
The column of differences is clearly an arithmetic progression:
416, 5, 516. .. and so on, adding a half each time. Careful
comparison with angles in water as measured today shows that
the actual angles are not as given by Ptolemy at all. In other
283
CATCHING THE LIGHT
words, the column of water angles was not created by measure
ments but by theory. Ptolemy forced his “measurements” to fit
his theory. They are too “good” to be true.
The point here is not that Ptolemy was dishonest, but that a
certain motive existed for his falsification of the data, a very
high sort of motive. Namely, he wanted the data to be beautiful,
and beauty for the Hellenistic world was geometric, the perfec
tion of mathematics. The angles in water just had to be in an
arithmetic progression. It is such a beautiful sequence, and
measurements are so inexact, that it just must be correct.
As was also the case in his magisterial treatment of astronomy,
Ptolemy started his study of optics with certain metaphysical
assumptions about the way the world is ordered. Metaphysics
literally means “beyond physics.” One of his metaphysical val
ues was that the universe is formed according to number; it is
truly a “cosmos,” which in Greek means “order.” Nor should
we imagine that such values are outdated. Einstein, among many
others, tenaciously upheld the values of beauty and order as
essential features important to the theoretician. Nor was Ptolemy
alone in his own day.
Although he disagreed with Ptolemy about the specifics of
refraction, the fourth-century philosopher Damianos displayed
a similar spirit when discussing refraction in his own treatise
on optics. He derived the law of refraction on the basis of the
following logic: “if Nature does not wish to permit our visual
ray to wander about fruitlessly, she will let it break at equal
angles.” Thus, according to Damianos, there is a single “best”
way for nature to refract light, otherwise fruitless visual chaos
would result. The best way is to arrange for the angle of refraction
always to be exactly half of the incident angle.
Eight hundred years later, Grosseteste agreed with Damianos,
making even clearer the metaphysical basis for the derivation.
“And it is shown to us by this principle of natural philosophy,
that every operation of nature is by the most finite, most ordered,
284
Of Relativity and the Beautiful
AIR
WATER
Grosseteste’s early and erroneous law of refraction,
based on his sense of the beautiful.
shortest and best means possible.”25 Unfortunately, the facts of
refraction do not agree with the best possible world as imagined
by Damianos and Grosseteste; shortest may not be best.
The gut scientific reaction to all this is to call such meta
physical argumentation nonsense. Yet Mr. Bader’s conversation
with Feynman sounded not unlike the musings of Damianos and
Grosseteste, with an essential difference; Mr. Bader’s definition
of “best” was right! It was, in Plato’s technical terminology,
divine.
Causes: Necessary and Divine
Plato and, until the seventeenth century, all philosophers after
him, held that causes were of at least two types; one he called
285
CATCHING THE LIGHT
“divine,” while all others were merely “necessary.” Divine
causes are “endowed with mind and are the workers of things
good and fair.”26 By contrast, necessary causes proceed by
brute force and, left to their own devices, would lead only to
chance, chaotic results if they were not directed by divine
causes.
Carpenters with tools and materials at the ready, like nec
essary causes, still require an idea, usually expressed by ar
chitect and owner, before a house—the right house—can be
built. As with house construction, so also with the world. The
forces of nature in themselves are unintelligent, brute, accessory
causes called forward to “execute the best as far as possible”
by higher divine causes, says Plato.
Aristotle termed Plato’s divine cause causa finalis, or final
cause, distinguishing it from three other accessory causes:
material cause (the lumber of the house), efficient cause (the
builders with their tools), and the formal cause (the house
plan of the architect). The final cause of the house is its use
as a shelter. The end to which it is put, its purpose, is the
causa finalis.
In what possible sense can we make use of final causes in
optics? When puzzling over the phenomena of refraction and
reflection, one quite naturally asks after the forces at work, the
pushes and pulls that act on the light (whether thought of as
particles or waves). We do not ask what is the “best” path for
light to follow. Yet there is a tradition in physical science—of
which Damianos, Grosseteste, and Feynman are a part—that
does ask just this question.
That tradition goes back at least as far as Hero of Alexandria
(c. 125 B.C.), who proved that the path actually taken by the
visual ray (or by light) in reflection from a mirror is the shortest
possible path connecting the eye to the source of light with a
reflection off the mirror. All other paths are longer. Light follows
the best path, if by best we understand shortest.
286
Of Relativity and the Beautiful
Both Damianos and Grosseteste knew of Hero’s derivation,
and sought to adapt its argument to the case of refraction. The
difficulty is that for refraction the distance traveled is patently
not the shortest physical distance connecting light source and
eye. So they argued for an alternative notion of best, namely
equal angles. Where they failed was not in their use of meta
physical criteria, but rather in their choice of metaphysics. They
needed one closer to Hero’s criterion. They needed the notion
of shortest optical distance, which usually differs from a straight-
line distance.
Instead of the simple criterion of shortest distance, Fermat
advanced his “principle of least time.” Light always travels the
path that minimizes not the distance, but the time it takes to go
from one place to another. Like a savvy commuter trying to get
to work during rush hour, light may well choose to go a longer
physical distance if its travel time is shorter. The bending of
light as it passes from air into water, refraction, is a case of
light following a path that is longer in length but actually shorter
in time. With Fermat’s principle of least time, and no more
geometrical knowledge than Hero possessed, it is possible to
derive the true law of refraction (Snell’s law), as well as of
reflection. Light does follow the best path, if only we know what
is meant by best.
Traveling All Roads
If the story ended here, it would already be wonderful, but
Fermat is only the first modem to apply a much broader and
more powerful, unifying principle that is much loved in con
temporary physics. What Mr. Bader told young Feynman was
the twentieth-century version of Fermat’s principle. Generalized
by the French mathematician Maupertius in 1744, it is known
today as the “principle of least action.” The specifics are not
287
CATCHING THE LIGHT
important, but suffice it to say that for an astonishingly wide
range of physical phenomena, be it optics, mechanics, hydro
dynamics, electrodynamics, or quantum theory, one can define
a quantity called “action,” and go on to derive the laws of physics
for the domain of interest by finding the path along which the
action is smallest.27 Thus its name: the principle of least action.
The father of the modem quantum theory of radiation, Max
Planck, saw in this principle a very high truth. Unlike the
“differential” laws of physics that specify the moment-by-
moment forces acting on a particle as it moves, the principle of
least action seems to work with the whole path, undivided.
Formally, one can always pass from the integral to the differ
ential formulation, but each carries with it a particular attitude
toward the world. In one—the differential view—we are the
general contractor overseeing a construction crew and the ma
terials used; we are concerned only with necessary causes. But
we can also step back from our manual labors to admire the
work even before it has reached its physical completion, seeing
it whole in our mind’s eye. In doing so, we are free to imagine
other designs, other paths chosen, and then to select the best,
to conceive a cause divine. Both ways of seeing the task are
valid, each has its rightful domain. In physics, the judgment is
nature’s, she has chosen her “divine cause” and it seems to be
the principle of least action.
What excited the young Feynman was, I believe, the expe
rience of seeing the beauty and wholeness of this universe, of
seeing into the choices that nature had made. That first seeing,
which was an instance of scientific inspiration, bore splendid
fruits when later, as a brilliant graduate student at Princeton,
Feynman created his “path integral formulation” of quantum
theory on just this same principle of least action. It has been
called by many the most beautiful of all formulations of quantum
theory. In its view, all roads are taken, all paths are traveled
by a quantum particle, be it light or matter. When taken together
288
Of Relativity and the Beautiful
the countless paths distill to one whose action is least. Whatever
light is, here is where we will find it.
TWO AVENUES OF research lead us into light. One is directed
toward the universally true, and operates with mighty principles
such as relativity and least action. The other tends toward the
infinitely small, the building blocks of the world. Although we
call them by other names, even today we allow Plato’s distinction
between “divine” and “necessary” causes. Einstein termed them
principle theories and constructive theories.28 With the former
we flatter ourselves that we have detected the rationality of the
universe, its underlying intelligence; with the latter we are con
cerned with the nitty-gritty engineering of how nature realized
her objectives. The play being written, we are curious to peer
behind the wings at the stage machinery. Goethe and most of
twentieth-century philosophy warn us not to make too much of
the “reality” we discover there. The discoveries we make, the
“realities” we see, may well reflect more about us than the object
of our investigation, and this has proven doubly true of modem
physics. Yet one is loath to forgo the pleasure of drawing pictures
from our theories. We wish to know not only how light behaves,
but what it is. Unwilling to accept the answers of the past, we
forge our own understanding, cast our own idols.
Neither the theory of relativity nor the principle of least action
speaks about the nature of light. Starting from apparently in
nocent beginnings with intuitively sensible assumptions, how
ever, these theories do make extraordinary deductions that lead
to universal laws of beauty and power. Yet for some, laws alone
are not enough. Another avenue must also be followed, and it
is the path of dissection. We know it from previous chapters.
Ever since Leucippus and Democritus inaugurated atomist phi
losophy, science has sought for the elementary constituents of
nature. Light will of necessity be reduced through this analysis
289
CATCHING THE LIGHT
to its least parts. And yet here, too, a thousand surprises await,
because the least part of light quite literally conceals a whole.
We live poised between the part and the whole, the differential
and the integral, the fragmentation of a darkly ahrimanic spirit
and the enticements of a luminous luciferic one, between, as
Schiller called them, the tendency to substance and the tendency
toward form.29 Between them both we find our way, giving to
each its due. Schiller likened this kind of engagement with the
world to true play, for only in play is one free.
LOUIS Kahn once told of a task he imposed upon himself.
I gave myself an assignment: to draw a picture that dem
onstrates light. Now if you give yourself such an assignment,
the first thing you do is escape somewhere, because it is
impossible to do. You say that the white piece of paper is
the illustration; what else is there to do? But when I put a
stroke of ink on the paper, I realized that the black was
where the light was not, and then I could really make a
drawing, because I could be discerning as to where the light
was not, which was where I put the black. Then the picture
became absolutely luminous.
Even the artist must invoke darkness in order to illustrate
light. Louis Kahn knew this truth more clearly than most. I am
no different. The finite expressions of language compose my
palette, and my limited conceptual world defines the dim mark
ings I can make. Like Kahn, all I am able to do is put my words
where light is not. But perhaps by my doing so, this book, too,
will become absolutely luminous.
290
Of Relativity and the Beautiful
U'lkx <Q MbCtkiuUL,
* ■
£1?
v ■ t
pru*' *vo
- 3/wv^xn O|
i
Ctk/ U1ACA1A M be Iz XxfUM 1l» (THZ.
3 .
* \ \ ' / /
£ \ ' S'
- - / g-
J
v___
s__
— \ <
LU
FHJ
/ I \ wl
/^{?Uq thP/^l ^*Ukvr»vn)' H0W
Silence and Light by Louis Kahn.
291
CHAPTER I I
Least Light:
A Contemporary View
Ever splitting the light! How often do they
strive to divide that which, despite every
thing, would always remain single and whole.
Goethe
Our discussions have reached a turning point. The changing
understandings of the past must give way to the tentative grop-
ings of the present. The stories told are now our stories. In our
best judgment, formed on the basis of everything we know, what
do we regard as the essential nature of light? To rainbow, candle,
prism, and mirror are now joined deeply puzzling quantum phe
nomena in which the character of light seems paradoxical in the
extreme.
Rather than describe this “modern” light abstractly, we will
search among the countless experiments of quantum optics for
those concrete instances that most clearly express the photon’s
essential behaviors. These experiments will dramatize the
unique features of light quanta, and become what Goethe would
gladly have called the archetypal experiments of the quantum
292
Least Light: A Contemporary View
domain. By patiently working with them, we will, in Goethe’s
view, fashion new organs of cognition better suited to our subject
than the traditional faculties we have inherited from classical
physics. The challenge of quantum phenomena is, then, a chal
lenge to our complacency of mind, an incentive to reach beyond
ourselves. Like fish swimming the black waters of Mammoth
Caves, we have become attuned to the darkness. Only after
centuries of effort have we moved into light-filled waters. Blinded
still by the habits of the cave, we are reluctant to explore and
fully contemplate the open sunlit terrain. Only by self
consciously doing so can we ever hope to approach an under
standing of light.
The phenomena of light, even when restricted to quantum
effects, are innumerable. Yet certain of them teach us more
than others. Like a great sculpture or work of art, in them the
essential is brought to the surface, and every extraneous feature
is chiseled away. Such phenomena carry a special significance
and, in each field, are few in number. Goethe called them
“archetypal phenomena,” and I will adopt his term here as well.
In them we encounter self-evident manifestations of natural law.
They are the experiential windows that look into the eternal
principle by which nature shapes the flow of phenomena. If we
have the eyes really to see an archetypal phenomenon, then we
have the wit to understand it. What are the archetypal phenom
ena of light at the quantum level?
By turning away from the common experiences of light to the
quantum, we follow the modem trend toward a study of the least
parts of light. Contemporary quantum experiments investigate
light when reduced to its weakest possible level. As we saw,
Max Planck and Albert Einstein were the first to posit the ex
istence of an elementary quantum of light, suggesting light was
not infinitely divisible, and so possessed a least part. In 1926
the American chemist G. N. Lewis gave the quantum of light
its current name: the photon. Research physicists have probed
293
CATCHING THE LIGHT
the photon and quantum optics is the specific subfield that takes
upon itself the study of light as a quantum object. As I have
grown to appreciate firsthand, it is a marvelous study full of
dramatic moments that ultimately offer a fresh way of thinking
about light.
Can we find archetypal experiments in which the photon
clearly, if paradoxically, shows some aspect of itself? Then, like
Einstein steadfastly running after light, we can attend single-
mindedly to the puzzling features of the photon, hoping that by
doing so we might finally come to understand the mystery it
presents.
If we wish to study the least part of light, the isolated photon,
the first task clearly is to construct a source of them. At first
blush it would seem trivial. Simply take any light source, a
candle even, and dim it sufficiently so it gives off a slow but
steady stream of single photons. Just this was done for decades
until the fundamental logical flaw in doing so was realized. No
one ever checked experimentally to see if these sources were
really producing single photons. One assumed without experi
mental evidence that dim light is a single-photon stream. What
evidence would convince us that a source is a single-photon
source? Obviously, a suitable means of testing light sources is
required.
Perhaps the most elegant single-photon test has been realized
by the French group of Alain Aspect, Philippe Grangier, and
G. Roger.1 Although technically somewhat difficult, it is con
ceptually supremely simple. One passes the suspected photon
into an optical instrument that divides light, sending half one
way and half another way. If at some point the light is no longer
divisible, then we have reached the level of the photon, or
“atom” (from the Greek, meaning indivisible) of light. The op
tical device that divides light in half is a half-silvered mirror,
or “beamsplitter.” When light falls on it, half the light is trans
mitted and half reflected. Imagine a single, atomlike photon
294
Least Light: A Contemporary View
COINCIDENCE
COUNTER
Light
Half-silvered
mirror Detector
Light as a particle. In this experiment a single photon falling on the half
silvered mirror goes to one detector or the other, never both.
striking the beamsplitter. What will happen? If it is truly in
divisible, then it will go one way or the other but not both. This
is called “anticorrelation.” If, on the other hand, the light can
be divided, both branches will have light, half will go one way
and half the other. This is the test.
In contrast to an atomic theory of light, the wave theory
conceives of light as infinitely divisible; there is no lower limit
to how weak the light’s intensity can be. Therefore the beam
splitter will always divide the light, transmitting half one way
and reflecting the other half. The beamsplitter is, therefore, a
litmus test for light: wave or particle. Performing it on a variety
of light sources has yielded startling results. All traditional
sources such as candles, incandescent lamps, gas discharge or
fluorescent lights, and even lasers show no anticonelation, what
ever their intensity. No matter how weak their light, the data
show that the light they emit splits at the beamsplitter in a way
that can be well described by a wave theory of light. All the
295
CATCHING THE LIGHT
usual sources of light thus fail the simplest criterion for single
photon emission. Single-photon sources are not natural.
All is not lost. Humans are ingenious. If conventional light
sources do not suffice, we can contrive new ones that serve much
better. In recent years, two such sources have been developed,
one relying on atomic cascade and the other on what is called
in the trade “two-photon, parametric, down-conversion.” The
specifics of these sources need not concern us, but in both cases,
two closely related photons are generated, one of which signals
the presence of the other to the experimenter. When properly
used, these sources have been shown to pass the single-photon
litmus test. The light they produce displays the anticorrelation
looked for in single-photon sources.
Thus, only in recent years have physicists had good sources
of nonclassical, quantum-mechanical light, and we have been
enjoying ourselves ever since. Dozens of elegant experiments,
often yielding strange results, have been done in recent years
using single-photon sources. I will relate a few that I think most
clearly reveal the subtle nature of light. Within them we may
find the archetypal experiments we are seeking. Now that we
have the single photon, let us study its behavior. It may be
hopeless to express the nature of light abstractly, but through
its actions we can catch a glimpse of its character.
Archetypal Instances
Behind the tireless efforts of the investigator there lurks a
stronger, more mysterious drive: it is existence and reality
that one wishes to comprehend.2
Einstein, 1934
Early on, Einstein drew attention to the perverse nature of the
photon. As usual, he suggested an insightful thought experi-
296
Least Light: A Contemporary View
Mirror
F
I
L
M
Half-silvered mirror
Light as a wave. Interference requires that light travel both paths to the
film.
ment. Use a single-photon source, he once said to Bohr, to
generate an isolated photon. Pass the photon through a beam
splitter. We know that the photon has gone one way or the other;
this was after all the very test we used to declare the source fit
to be called a single-photon source. Check as often as you like;
the anticorrelation is always there. Now, said Einstein, replace
the two detectors with mirrors that redirect the photon so that
whatever way it travels, it ultimately lands on the same section
of a photographic film. What do you see?
With the arrival of the first photon, a single, tiny black spot
appears roughly where one expects. The following few photons
likewise show up as spots scattered around the first. So far
everything is fine. But as time passes and the spots accumulate
one by one, something altogether unexpected appears. The spots
align themselves into dark and light bands, into clear interfer
ence fringes. The experimental facts are unambiguous. When
we set up to see interference fringes, we see them, even with
single photons. When we set up to see which single path the
photon has taken, we find out. The problem is not with the
phenomena, but with the inadequate thoughts we bring to them.
297
CATCHING THE LIGHT
0,02s
10s
■ *,»£*
Interference fringes build up slowly, one photon at a time.
In the first case, we think waves; in the second, particles. In
the first, we think both paths; in the second, one path. How can
both be true? What kind of a thing could possibly behave this
way? To quote Einstein at sixteen, “something like that does
not seem to exist”! Hold on to this thought for a moment.
Ever since Thomas Young, interference fringes have been
understood as due to light traveling by two paths. But the source
of light used in this experiment was from a single-photon source,
which has the demonstrated property that it sends photons along
only one path, not two. The appearance of interference fringes,
however, requires that in some sense the single, indivisible
quantum of light, the lone photon, takes, or at least is affected
by, both paths! Bohr paraphrased Einstein’s summary of the
situation as follows: “In any attempt of a pictorial representation
of the behavior of the photon we would thus meet with the
difficulty: to be obliged to say, on the one hand, that the photon
298
Least Light: A Contemporary View
always chooses one of the two ways and, on the other hand, that
it behaves as if it had passed both ways.”3
Here is an archetypal instance of wave-particle duality. For
several decades, Einstein’s thought experiment lacked a proper
laboratory realization. Finally, in 1986, Aspect, Grangier, and
Roger successfully performed a lovely version of the experiment
with unambiguous results. The solitary photon does indeed in
terfere with itself. One thing—the photon—seems somehow
simultaneously related to two distinct paths. The implications
of the experiment are fundamental. The root structure of quan
tum mechanics, and our idea of light, must accommodate this
experimental fact. We have found an archetypal phenomenon
but lack the ideas with which to see it rightly.
Goethe was right. Try though we may to split light into fun
damental atomic pieces, it remains whole to the end. Our very
notion of what it means to be elementary is challenged. Until
now we have equated smallest with most fundamental. Perhaps
for light, at least, the most fundamental feature is not to be
found in smallness, but rather in wholeness, its incorrigible
capacity to be one and many, particle and wave, a single thing
with the universe inside.
In recent talks and articles, the distinguished physicist John
Archibald Wheeler has reminded the physics community that
the quantum ambiguity demonstrated by single-photon interfer
ence experiments is even more serious than it first appears. He
did so by proposing an interesting thought experiment, one first
hinted at by Einstein and Carl Frederich von Weizsacker.' In
it the previous two dichotomous experiments (anticorrelation and
interference) coexist until the bitter end, and the wave-particle
ambiguity is sustained throughout the photon’s flight.
Imagine that a single photon passes through a beamsplitter
as before, but the choice is delayed as to whether detectors or
mirrors are installed. In principle, photons can be “split” at
huge distances from the detectors, and the delay made as long
299
CATCHING THE LIGHT
as one pleases, years even. The puzzling nature of light sinks
in while we wait. During that entire time, what can we say about
the path followed by the photon? If we think of it as a particle,
then, when the paths split long ago, the photon took one of them.
However, we may well decide to run an interference experiment
on that photon, and make our decision to do so well after the
photon has passed the fateful decision point. Nevertheless, we
will see interference. If, on the other hand, we think of the
photon as a wave going both paths, then again, long after the
branch point has been reached, the experimenter can switch to
a measurement of path, again with success. Prior to the final
moment of measurement, it seems that we cannot say the photon
travels both paths, but nor can we say it travels one path; a
situation that is certainly very awkward! However, after we com
plete the measurement, we can apparently declare whether the
photon had traveled to us by a single path or by both. In Whee
ler’s words, by our delayed choice we seem to “have an una
voidable effect on what we have a right to say about the already
past history of that photon.” Delaying the choice of what
experiment we run has dramatized the essential ambiguity of
single-photon physics.
Something very strange confronts us here. Conventional, well-
defined objects have a clear, unique history. They originate
somewhere, travel a path, and in the end arrive somewhere else.
Single photons (indeed, all quanta) are also well defined, at least
mathematically, but are different. The photon’s history of motion
shows an irreducible and paradoxical, wave-particle character
that is unlike anything we know from classical physics. What
is the source of the ambiguity?
At the Max Planck Institute for Quantum Optics outside Mu
nich, I recently collaborated in a demonstration of Wheeler’s
proposed delayed-choice experiment.5 The results were as quan
tum theory predicted, convincingly showing us that a single
photon is ambiguous in the quantum-mechanical sense all the
300
Least Light: A Contemporary View
way until the detection. The single photon makes no “decision”
at the beamsplitter, either for particle or for wave. It is neither,
or both. The ambiguity is essential to light, one exceedingly
difficult to grasp but all the more important for its elusiveness.
If the photon exists between sightings, then our description of
that existence must allow for the facts of single-photon inter
ference experiments.
ATTITUDES toward THE delayed-choice experiment, and other
experiments like it, vary. The majority of physicists by far simply
do not concern themselves with the meaning of their quantum
calculations. Nor do they trouble themselves about the impli
cations of archetypal quantum experiments. Science is not, they
say, concerned with truth or meaning, but only with prediction
and control; it is an instrument. Nineteenth-century scientific
arrogance here changes to twentieth-century cynicism. Power
we have and that is enough; true knowledge we relinquish for
ever. According to this view, modem science is like the as
tronomy of ancient Babylonia three thousand years ago. Without
any physical notions about our solar system, astronomer-priests
could predict the course of stars, sun, moon, and planets with
amazing accuracy by applying purely arithmetic procedures to
data from the past. Likewise, we can apply a quantum algorithm
to make predictions about quantum phenomena without any
notion of what a photon, an electron, etc., really is. Yet there
is a difference. Babylonian astronomer-priests sought relent
lessly for the meaning of their observations, but it was a spiritual
meaning wedded to rituals, practical life, and their participation
in the life of their gods and goddesses. With this they were
content. We have shed every spiritual framework for science.
The instrumentalist view would also abdicate hope for a true
physical understanding.
As the gods disappeared the Greeks fashioned a rational,
301
CATCHING THE LIGHT
geometric cosmos to replace the dying spiritual one. In the
sixteenth century, matter joined reason, and the universe grad
ually became a clockwork. Twentieth-century physics and phi
losophy called much of this view into question. It asked whether
any true knowledge of matter, light, or by extension, all of nature
was possible. The deconstruction of scientific knowledge has
been liberating, but freed from its tyranny, we now verge on a
demoralizing nihilism. If scientific knowledge is only instru
mental, is there meaningful knowledge of any kind? In their
hearts, no serious, practicing scientists believe in instrumen
talism; no teacher glows warm with enthusiasm in presenting a
meaningless calculational procedure to his or her students. To
what, then, does physics lead?
When we are confronted with the black void of pure instru
mentalism, the temptations of reactionary idolatry are very near.
Having lost the gods, we fall in love with the beautiful idols we
can raise in their places. Atoms, quarks, tiny black holes...
they are reified, garlanded, and dragged forward to assume a
place in the temple. Calling them real, we animate them with
the false life of fear, our fear of the unknown being of nature.
We feel caught between the sophisticated emptiness of decon
structionism and the gaudy creations of our own hands. To what
can one hold? Goethe’s answer would have been that given to
Hegel and the readers of his Theory of Color; hold to the phe
nomena. They can be trusted. Rightly seen, they will become
the theory.
To see we must represent objects to ourselves, we must apply to
them concepts that fit their nature. Failing this, we fail to see. We
are like the once-blind S. B. who remained mostly sightless even
when surgery repaired his eyes. He possessed the raw material for
sight, but did not see. The experiments of quantum optics have
given us the raw material of light, but we still lack the concept
suited to its nature. Once we fully possess it, the paradox of wave
particle duality will disappear while “wave-particle-ness” re-
302
Least Light: A Contemporary View
mains. When that intelligence is brought to archetypal quantum
phenomena, we will see them with understanding and not confu
sion. In Emerson’s language, we will have “named” them, and
once named, a phenomenon becomes a theory. We can then rec
ognize the presence of “wave-particle-ness” in many places, hav
ing seen it clearly in one. Like tenderness discovered in
childhood, we will know it when we meet it.
LIKE ADOLESCENTS READING Shakespeare, we feel the language
of quantum phenomena to be at once strange and wonderful.
The meanings are obscure, some words unknown. All the more
reason to pause over it, to reread the text and muse over its
possible meaning. Understanding it to be a play, we can me
chanically act out the parts, but much is left untouched thereby.
If only the script could be penetrated, the meanings discerned,
then how much richer the production would be. Surely nature,
like a found dramatic script, deserves to be read for meaning
and not merely acted out for profit. As we hear the Sirens of
the past, we should know that everything depends on the courage
and freshness of the reader. The phenomena of light refute a
nineteenth-century reading of the world, even if the modem
staging remains fragmentary and hard.
In recent years growing dissatisfaction with instrumentalism
has led many scientists to adopt other attitudes toward quantum
phenomena. Some espouse “critical realist” perspectives, in
cluding realist interpretations of quantum mechanics. Regarding
quantum phenomena, there are, broadly speaking, two such
positions. The larger group holds that we must fundamentally
change our understanding of the photon and all other elementary
particles. They are real but profoundly unfamiliar. For example,
one cannot speak about the “path” of the photon. The photon
does not have a trajectory as such, a set of particular positions
that change with time. In order to describe the photon’s history
303
CATCHING THE LIGHT
we must rise above the particulars to a more abstract level. One
adds together the “quantum amplitudes” for all possible paths
the photon can travel, and the “superposition state” so con
structed is the proper quantum mechanical description of the
photon. The evolution of the superposition state replaces the
concept of trajectory familiar to us from classical physics. If
one takes a realist view, then the superposition state, which is
a kind of metaphysical mixture of possible paths, is the photon
as it travels from beamsplitter to detector. Wheeler calls the
superposition state the “Great Smoky Dragon.” Although the
head and tail of the dragon show up, its huge body is obscured
by smoke. So too with the photon. We may know all about its
emission and detection, but the path it follows in between is
ultimately ambiguous.
Wheeler goes further, reaching beyond the views held by
most physicists when he uses delayed choice to show how
strongly our present measurements connect us with the past. In
the above experiments, only the final act of detection lifts the
quantum ambiguity and dispels the smoke to reveal a classical
world in which even the photon’s path can be talked about. The
final detection apparently determines the already-past history of
the photon. In an interview, Wheeler remarked, “To the extent
that it [the Great Smoky Dragon or photon] forms a part of what
we call reality, we have to say that we ourselves have an un
deniable part in shaping what we have always called the past.”6
A second and much smaller group interprets the single
photon, delayed-choice, interference experiment very differ
ently. David Bohm and Basil Hiley are the foremost advocates
of a view that pictures photons and all quantum objects in more
traditional ways. Quanta have positions, trajectories, and real
path histories. Of course Bohm and Hiley must provide in their
theory for the paradoxical phenomena of quantum mechanics.
They do so by introducing something unknown until now. Phys
ics has traditionally divided the world into particles (electrons,
304
Least Light: A Contemporary View
quarks...) and fields (gravitation, electromagnetic, and nu
clear). They suggest the addition of a third entity, the so-called
quantum potential. It acts like an immaterial ether whose pur
pose is to guide the photon in something like the way a remote-
controlled airplane is guided by radio signals. All quantum
effects are carried by this information field.
The quantum potential has unusual features. First, the quan
tum potential does not “push” on objects, but informs their
motion, structuring reality through its own form. Since it does
not exert a force, the quantum potential is not directly detectable
by physical means; one can only gauge its presence indirectly.
Second, the quantum potential is instantly sensitive to all
changes in the experimental arrangements, even very distant
ones. The second feature is called nonlocality, and is a central
feature of all present discussions of quantum theory, whether
by Bohm or the quantum realists. We will encounter the concept
again, and treat it more fully then. The Bohm and Hiley theory
offers us the attractive possibility of imagining particle motion
as we always have, but it comes at a cost. The cost is the
introduction of a kind of “ghost” field, or immaterial ether—
the nonlocal, quantum potential.
One may understand the photon to be a nonclassical, quantum
object and so give up all meaningful statements about its history.
Or with Bohm and Hiley one may populate space once again
with a new quantum-mechanical, nonlocal ether. One view lo
cates the quantum character of reality in a hidden environment
(Bohm/Hiley), while the other integrates it into the photon itself.
The former is a kind of dualism, the latter a monism. Yet both
views insist that our world is not what it was once thought to
be. One view gives up history in favor of a “quantum reality,”7
the other advances a new “implicate order,” as Bohm calls it,
of which our reality is but a partial projection.8
Quantum phenomena show the world, at least on the atomic
scale, to be utterly unlike the living-room world of sofas and
305
CATCHING THE LIGHT
armchairs. Its structure is other, its order unfamiliar; yet for all
that, it is no less real. All conceptual schemes, no matter how
they differ one from the other, must accommodate the data of
archetypal quantum phenomena. To my mind, therefore, it is
of far less importance which one theory we take to be true than
to see what they all have in common. Each points from a different
direction to a common core.
Walking around Michelangelo’s David, we see it from differ
ent sides. Reading about Michelangelo’s life, studying his con
temporaries and his other sculpture, we enlarge our orbit around
the David. Rather than fixate on one “view,” should we not (like
Einstein) hold on to our self while learning to see through the
eyes of others? Similarly with regard to quantum phenomena.
We should move among every equivalent theory, each interpre
tation, learning to see through its eyes. Each will highlight a
different blemish of the old order and offer an individual remedy
that reveals as much about its author as about the quantum. Yet
for having circled the phenomenon in this way, something al
together wonderful happens. We change. We see David or the
quantum phenomenon anew. Theories become aids to reflection,
as Coleridge would have called them, not the codification of
truth. Too quickly do theories become idols and so stand in the
way of insight instead of promoting it. Moving among competing
views, one is freed from the tyranny of monocular vision, and
like the Indian god Varupa, encompasses the world with a thou
sand eyes.
What, then, are the essential insights into light offered us by
quantum phenomena? We have begun, but to answer this ques
tion fully, we need to move farther around the statue. Already
the outline is becoming clear. Notions such as history, place,
and identity will need reworking, but in order to make these
clearer we need to study other quantum phenomena. So far we
have examined only single-photon, quantum effects. It is now
time to go further to two-photon and many-particle phenomena,
306
Least Light: A Contemporary View
to enlarge our orbit as we continue to fashion new capacities for
insight.
Entangled Light
The aim remains: to understand the world.
John Bell
Once again we will follow Einstein’s lead. Deeply troubled by
the implications of quantum mechanics, Einstein sought a way
to demonstrate its incompleteness convincingly. In his view,
quantum mechanics was only a partial picture of a far more
subtle and complex reality, much of which remained hidden.
If this were the case, then the peculiarities of quantum theory
would be no surprise to anyone. Quantum theory never pre
dicts individual events, but only the probabilities of them
occurring. But incomplete knowledge always shows up as un
certainty in predictions, whether at Las Vegas gaming tables
or weather predictions. Ignorance leads to chance. Yet is the
converse also true—do uncertainties in experiment results
always point up our ignorance? In the nineteenth century, a
confident physics would have answered yes, uncertain results
always reveal partial knowledge. Einstein insisted that ration
ality requires an unambiguous theory, and quantum mechanics
is ambiguous and therefore not rational. It is not a proper
theory at all.
To make his point, in 1935 Einstein, together with B. Po
dolsky and N. Rosen, once again resorted to a thought exper
iment, a brilliant one that has proved of unparalleled significance
for the foundations of quantum mechanics, especially when, in
1964, its consequences were elucidated by a modest but brave
young physicist from Ireland, John Bell.
307
CATCHING THE LIGHT
PHYSICISTS would agree that no single person has done more
important work in the foundations of quantum mechanics than
John Bell. His unique stature was due, I think, to his rare
combination of deep modesty, profound courage, and unmis
takable brilliance. Shortly before his untimely death in 1990,
I had the privilege of organizing a workshop on the foundations
of quantum theory together with my Amherst College colleague
George Greenstein, but it was Bell’s willingness to attend that
brought people together. For a solid week he acted as an in
sightful and radically honest critic of quantum theory during our
intensive conversations on its foundations. It proved a wonderful
gathering, filled with humor and music, as well as helpful treat
ments of the major unresolved questions in the field.
As Bell emphasized again and again, for all its power, and
few knew it better than he, orthodox quantum mechanics is
simply not good enough. It will be superseded. Bell once wrote
that the fate of quantum mechanics is apparent simply by ex
amining its own internal structure. Look carefully and you will
see its harsh future: “It carries in itself the seeds of its own
destruction.”9 Einstein and Bell possessed the uncanny ability
to see clearly the problematic malaise of quantum theory, and
they both held tenaciously to the aspiration that a better and
truer understanding of nature was possible. On the basis of Bell’s
theorems, some of the deepest issues in quantum mechanics
became open to experimental examination for the first time.
Those investigations are now reasonably complete, and their
significance for our contemporary reimagination of light and
substance is huge.
In THE YEARS since 1975 when Bell proved his famous theorems,
physicists have successfully performed Einstein’s 1935 thought
308
Least Light: A Contemporary View
experiment, since dubbed the EPR experiment after its authors
Einstein, Podolsky, and Rosen. It is an archetypal experiment
which, if understood fully a la Bell, carries us far into the new
order of things. It asks after reality, seeking for the “be-ables”
of a theory, as Bell termed them. What are the real properties
that inhere in things? Especially for light, the answer leads to
a far more subtle and, I think, beautiful understanding than
ever before. We begin very simply.
Objects of this world bear attributes: my pen has color, shape,
mass, and so on. Every object must have some well-defined
attributes by which we know it. Thus, my pen cannot be red
and green at the same time, without me suspecting my sanity.
Without sensible attributes, an object loses its identity; in a
sense, it disappears. Einstein asked, what are the real attributes
of quantum objects? By these we shall know them, or we will
not know them at all.
Since quantum objects are delicate things, one needs a means
of testing for an attribute that is noninvasive, to use medical
terminology. After all, if one is not careful, the object may be
disturbed so violently that it changes in one’s hands. Then, as
if wearing rose-colored glasses, I will see not the object’s color,
but the color of my glasses. In order to avoid this problem,
Einstein suggested the following. Let the objects of study be
produced in pairs, like identical twins bom together and indis
tinguishable except for one attribute. Staying with the analogy
of the twins, let the one distinctive attribute be a birthmark,
under the left arm, say. By this mark alone can we tell who is
the prince and who the pauper.
The twins rise each morning, bid each other farewell, and
head in opposite directions to laboratories at opposite ends of
town. One twin (we don’t know which) reaches his destination
slightly before the other. A scientist, uncertain as to who he
has before him, checks for the hidden birthmark. In that instant,
his uncertainty disappears as he recognizes the lad. Not only
309
CATCHING THE LIGHT
does he know with certainty who he has before him, but can
also predict with absolute confidence, without observing the
distant twin, who will show up in his colleague’s lab a few
minutes later. In a sense, observation on one twin is sufficient
to make true statements about both. The scientists keep track
of their results, X’s for pauper (birthmark), O’s for prince (no
birthmark).
SCIENTIST A SCIENTIST B
Feb. 1 X Feb. 1 o
Feb. 2 0 Feb. 2 X
Feb. 3 X Feb. 3 0
Feb. 4 X Feb. 4 0
Feb. 5 0 Feb. 5 X
Notice the perfect correlation; whenever one is X, the other
is 0. This is precisely how we expect the world to behave. Each
twin wends his way to one laboratory; one has a birthmark and
the other does not, one is the pauper while the other is the
prince. All is well with the world.
Einstein proposed a quantum version of this experiment, the
EPR experiment. Forty years later, Bell proved that the pre
dictions of quantum mechanics and those of any, alternative,
“local, realistic” theory would differ in specific measurable
ways. The important words are “local” and “realistic.” What is
a local, realistic theory? They are just the theories that give
explanations about which we would feel good. They are com
monsense theories. In addition, Bell’s theorem is very general.
It does not advance one competitor to quantum mechanics, but
all commonsense competitors in one fell swoop! The stakes are
very high. On the basis of Bell’s theorem, one good experiment
can eliminate every commonsense competitor to quantum me
chanics.
310
Least Light: A Contemporary View
The experiments have been done, most beautifully by Aspect
and his collaborators in France. They undeniably show the quan
tum-mechanical predictions to be correct, not those of any Ein-
steinian, local, realistic, commonsense theories. Experiments
necessitate, therefore, that something in Einstein’s view of ra
tionality must be given up. The question remains, what? The
EPR/Bell archetypal experiment urges on us the possibility of
a more flexible form for rationality than that of traditional sci
ence. We need not give up rationality, but rather must broaden
its meaning. Our conceptions of rationality were unreasonably
limited by our bias toward a mechanical universe. After all,
shall we declare the mathematics of quantum theory “irrational”?
Mathematics can perform with complete calm the most extraor
dinary things. Let us tum to Einstein’s quantum thought ex
periment, and see in what ways rationality needs enhancement.
Briefly stated: In laboratory EPR experiments, two photons
take the place of the twins. They are produced simultaneously
and also carry an attribute or “birthmark” called polarization.
Moreover, they are produced in a special way such that if one
is X-polarized the other is O-polarized, or vice versa. If we do
not check their polarization, then, as for the twins, the photons
are indistinguishable. We allow the photons to fly off in opposite
directions to distant detectors that are able to measure polari
zation. One arrives and its polarization is checked. If it proves
to be X-polarized, then the scientist there knows the other is
O-polarized. Repeat the experiment many times. As before, the
data from the two detectors are found to correlate exactly; for
every X there is a corresponding 0, for every 0 an X. Einstein
declared any attribute for which such sure predictions are pos
sible to be a real attribute of light; that is, it existed even before
the measurement was made. After all, how can distant actions
on one twin or photon immediately affect the properties of the
second? Effects act locally, not at a distance. In order to affect
a distant object, a signal must travel the entire distance, and
311
CATCHING THE LIGHT
that takes time. Again, this is common sense, an obvious feature
of any good theory, according to Einstein. He termed it “lo
cality.” Commonsense theories are realistic and local.
The correlations above are much like those for the twins.
Quite naturally we would initially interpret all this to mean that
each photon is really polarized from the outset, just as each twin
really is marked or not at birth. All this is completely reasonable;
it is exactly the way a rational world should be arranged. Un
fortunately, for correlated quantum particles, matters are oth
erwise! Quantum attributes are not so neatly separable, our
actions are not so nicely localized. Bell and the recent EPR
experiments drive this point home unforgivingly. In general,
light correlations do not behave like twin correlations under all
circumstances. Bell’s theorem pointed out those special situa
tions for which quantum theory and local, realistic theories
differ, and subsequent measurements have come down clearly
in favor of quantum predictions. Common sense is thus violated
in an unavoidable way, but what is the significance of the vi
olation?
Einstein’s view was one in which objects possess real, en
during attributes such as color, polarization, and a path of
travel. EPR experiments, especially the most recent ones by
Aspect and collaborators in France,10 measure the attribute of
photon polarizations for various detector orientations. Their
results, when coupled with Bell’s theorem, convincingly dem
onstrate that there is no local, realistic way of understanding
polarization correlations! The experiments and logic that lead
to this conclusion are compelling. They demand of us an
extremely deep reformulation of what we take light to be. To
make clear the grand implications of these pivotal experiments,
we can examine their implications for theoretical interpretation.
Again two prominent tacks are taken. Both are important aids
to reflections as we ponder the archetypal quantum phenomena
of light.
312
Least Light: A Contemporary View
The FIRST AND most common response is that paired photons
cannot be understood as individually polarized (i.e., marked)
from birth at all! In this understanding of quantum mechanics,
the attribute of polarization is no longer granted to each photon
of the pair separately, but is somehow only a shared or holistic
property of a new kind of object. Here is an instance in which
the whole is rigorously not the simple sum of its parts. I will
call this view “quantum realism.” According to it, one cannot
conceive meaningfully of two separate photons traveling through
space each with its own unique and enduring polarization. Quan
tum reality is different; separability is lost, and the relationship
of whole to part is not that given by Kant to the inorganic world.
Once two objects have interacted quantum-mechanically, they
join to become a new single, “entangled” entity, as Schrodinger
called it. The new entangled state of light does not travel in the
usual sense, but evolves in a more holistic way, retaining its
ambiguous nature throughout.
At this point an enormous problem arises, one seen early on
in the development of quantum mechanics. If quantum reality
is entangled in surprising and subtle ways, how is it that we
see sense reality as separate and disentangled? This is the “mea
surement problem.” Quantum reality is whole, all of a piece
until a measurement is made. Then somehow what was one
becomes two; what was an entangled whole becomes disentan
gled parts, and does so instantly. In the case of paired photons
in the EPR experiments, an exact polarization is found at one
detector, and in addition, a distant paired photon simultaneously
shows up with an exactly correlated polarization! The ambigu
ous, entangled state is somehow reduced, and the reduction can
take place over arbitrarily large distances, instantly. Clearly,
Einstein’s condition of locality is violated. How does orthodox
quantum theory account for this nonlocal reduction of entangled
313
CATCHING THE LIGHT
light, that is, for the simple fact of measurement? It cannot.
Many have been the proposals, past and present, but none has
yet succeeded in addressing the measurement problem con
vincingly. If quantum reality is, then the passage from it to sense
reality is miraculous. At this point several physicists and phi
losophers, including the Nobel laureate Eugene Wigner, intro
duce the active agency of the mind. The miracle of reduction,
they say, takes place by the action of mind in the moment of
cognition. Only when one fully includes its role, they argue,
can one account for knowledge.
Turning aside from the measurement problem, at least for
the moment, we can search for a surer footing with regard to
the nature of light. As proved by experiments with photons, the
attribute of polarization does not have a unique, enduring “local
reality” in the Einstein sense. Perhaps the EPR identity problem
is only associated with the attribute of polarization. Light is
specified by characteristics other than polarization, which (we
hope) may not suffer the same ambiguity. Perhaps these other
attributes will be firmer, sharper, and less illusive. For massive
particles, for example, the attribute of electric charge is un
ambiguous (a so-called super-selection rule requires charge
never to appear in an entangled state). But the photon has no
mass, or charge. In fact only four attributes are used to define
light formally; they are polarization, wavelength, direction, and
intensity. We have already seen that the history of a photon’s
path is ambiguous, so direction is no better than polarization.
Experiments show that wavelength (i.e., color) and intensity
also suffer exactly the same fate as polarization.11 All can be
come entangled, one color with another, and so on. The con
clusion: There is no truly unambiguous attribute of light! Today’s
quantum-optics experiments challenge the root conceptions we
have about the separable atomistic structure of the world. More
over, much the same can be said to hold for material particles.
For example, even the mass of objects can be put into the same
kind of entangled superposition state that light evinces.
314
Least Light: A Contemporary View
How bad is this? Very bad. If true, and it is true, then it
implies there are parts of reality (assuming there is a reality)
where attributes do not map onto things simply. It would be as
if you found an object that was no particular color, shape, size,
mass, etc. It was nothing in particular, but still remained a very
specific thing. What are the attributes of light? It seems a simple
question, and quantum theory gives what is at first a simple
answer: polarization, wavelength, direction, and intensity. More
careful thought, however, shows that quantum reality handles
the attributes of light far differently than does sense reality.
Sense objects must possess sharply defined attributes. Light,
quantum mechanically considered, need not. Its attributes are
more holistic; in general they exist in inseparable or entangled
combinations, at least until the moment of measurement, what
ever that is.
Ever since Galileo, Descartes, and Newton, science has
sought for the “primary qualities” of things, that is for the un
ambiguous and irreducible attributes of reality. The senses pro
vide only secondary experiences, but behind them, they argued,
lay the primary qualities of extension, or mass, or solidity, etc.
What are the primary qualities of light that vouchsafe its un
ambiguous existence? The extraordinary response given by
quantum realism is that there are none. Light, as an enduring,
well-defined, local entity vanishes. In its place a subtle, entan
gled object evolves, holding all four of its quantum qualities
suspended within itself, until the fatal act of measurement.
What are THE responses of physicists to these metaphysical
implications? Foremost is what Bell has called FAPP, “for all
practical purposes.” We have a practical way to proceed, they
say, and would do better to give up all hope of an objective
picture of reality, a suggestion strongly advanced by Bohr. The
procedures laid out by quantum theory are adequate “for all
practical purposes.” The troubling areas are not large and mostly
315
CATCHING THE LIGHT
philosophical, it is said, and can often be made nearly to vanish
(although they never disappear entirely); so FAPP physics is
good enough. Schrodinger protested in the strongest terms.
A widely accepted school of thought [Bohr’s] maintains that
an objective picture of reality—in any traditional meaning
of that term—cannot exist at all. Only the optimists among
us (and I consider myself one of them) look upon this view
as a philosophical extravagance born of despair in the face
of a grave crisis. We hope that the fluctuations of concepts
and opinions only indicate a violent process of transformation
which in the end will lead to something better than the mess
of formulas that today surrounds our subject.12
The grave crisis will not go away by ignoring it. Here are the
seeds of self-destruction referred to by John Bell as latent in
quantum theory itself. Besides, there are many interesting av
enues to explore, exciting implications to play out, even if they
are uncomfortable for traditional commonsense realism. The
world may very well prove to be rational in the end, but its
structures are far more varied than those immediately revealed
to our senses. We should learn from Bell, Einstein, and so many
others to be awake to what confronts us, to face it clearly and
carefully, and to be ready to transform boldly our most habitual
ways of seeing.
Light falling ON the eye provokes sight. Until that moment,
light lives in a universe of its own, all its commonsense prop
erties caught up into itself. As we saw in the last chapter, even
space and time lose their meaning when one imagines traveling
at the speed of light. Intervals of time and distance disappear.
As light touches every atom it enfolds atomic history into its
own, entangling itself ever more widely. Ironically, as it does
so the drama of the EPR results in one way declines. Quantum
316
Least Light: A Contemporary View
theory usually predicts the largest divergences from common
sense for the simplest entanglements of light. Quantum effects
are generally diluted as entanglement grows, disappearing for-
all-practical-purposes (FAPP). Yet even this is not always true,
and when it is not, the drama only increases.
Every single-photon interference experiment, every two-
photon EPR experiment, and many I have not touched on have
also been performed with matter, that is with electrons, or other
atomic particles. For example, neutrons, the uncharged con
stituent of the nucleus, have been used recently in magnificent
demonstrations of wave-particle duality in just the way I have
used photons.13 Most recently, several experimental research
groups have even used atoms, many times more massive than
neutrons, and placed them into superposition states that show
the same interference effects as massless photons.14
In addition, a few many-particle experiments have now been
performed that convincingly show that the stunning effects due
to entangled states do not always diminish with increased num
bers of particles. The experiments at IBM’s Watson Research
Center by Webb, Tesche, and Washbum have woven billions
of electrons into collective, entangled states to produce im
pressive quantum-mechanical effects.15 High-temperature su
perconductors also promise to bring such collective quantum
states ever closer to everyday life. The remoteness of quantum
paradoxes may not be long-lived. Then what will we make of
our world?
Quantum realism holds that the world really is the way or
thodox quantum mechanics says it is. Properties of photons such
as path and polarization do not exist properly before measure
ment. Ambiguity disappears in that instant by mechanisms not
yet known to us. Many are the problems with such a view, but
what are the alternatives? There are several, but to my mind
the most enduring has been, and likely will be, that of David
Bohm’s mentioned before.
317
CATCHING THE LIGHT
For every experiment performed to date, Bohm has a consis
tent theoretical account. In every case, a “real” photon travels
a real path, with a real polarization and wavelength. The en
tanglement of quantum realism is not, in his view, an entan
glement of quantum particles, but of his new quantum potential.
The new physics of quantum mechanics is carried by it, in
cluding the mysteries of nonlocality. The electron, neutron, or
photon is guided by this ghost field. If one grants Bohm the
existence of his quantum potential, then he can explain every
experiment with the additional benefit that even measurement
is accounted for! The world of sense reality, of particles and
fields, is very like it has always been, but it is understood by
Bohm to be a projection of a far more subtle “implicate order”;
the holism of EPR and entanglement are essential aspects of its
nature.
Bohm’s is an attractive view, but requires adherents to accept
the profoundly new implications it entails. Among these are the
quantum potential, and the existence of a source of energy within
all elementary particles. Bohm argues that the quantum potential
does not “push” or “pull” objects the way other forces do (for
example, gravity), but rather the quantum potential “informs”
their motion as the movements of a remote-controlled model
airplane can be directed from the ground. Very little energy is
needed to send the signal to the airplane, but the consequences
of its reception can be dramatic. The behavior of quanta are
like those of the plane, which are guided by the quantum po
tential.
At this point, several problems arise. No one has ever found
direct evidence of the quantum potential. The tiny electromag
netic signals that guide model airplanes and space probes can
be detected with sensitive equipment, as they are on the crafts
themselves. Despite experimental searches, why have we never
found the “ghost field” of the quantum potential? If it is real,
then it must transmit at least a small amount of energy to the
318
Least Light: A Contemporary View
receiving particle. That energy should be detectable, but has
never been seen. In addition, the model airplane has an engine
and servomotors internal to itself. These are directed by the
information carried by the radio signal. Bohm’s view implies,
as he himself says, that every elementary particle possesses an
internal “engine.” Particles appear inert at one level, but at the
level of the quantum they are “self-movers.” Yet all experimental
evidence to date, even at the highest energies and smallest
dimensions, has turned up no evidence of structure in the elec
tron or of “engines” within elementary particles. It is nearly
impossible for most physicists to imagine an engine within the
photon or electron, or that the ghostly quantum potential can
be so all-pervasive and yet so elusive. To them, FAPP and
quantum realism, even with the measurement problem, are eas
ier views to adopt. Still, to my mind, Bohm offers a serious
alternative, even if few, far too few, have joined him in his
research into implicate-order physics.16
In EITHER VIEW, that of Bohm or quantum realism, a monumental
shift has taken place in our conception of things. It goes by the
humble name of “nonlocality,” but within it is concealed a
revolution in thinking. No new predictions will arise from it, no
new high-tech gadgets, but if taken seriously, it urges us to do
what Goethe, Thoreau, Whitman, and a thousand others have
petitioned for a century: to conceive of things whole. As Francis
Thompson phrased it,
All things . . .linked are,
That thou canst not stir a flower,
Without troubling of a star.17
If one day we awoke and truly saw the world this way, the
ramifications would shake our psyche to its foundations. As-
319
CATCHING THE LIGHT
suming we remained sane, the relationship of I to thou, of
individual to planet, of my actions to yours, would be revolu
tionized. Edward Lorenz’s theory of nonlinear systems delivered
the “butterfly effect.” The flight of a butterfly in Rio de Janeiro
can, in fact, change the weather of Japan. If chaos dynamics
has shown us the extreme delicacy of our world, quantum physics
reveals its deep intimacy.
Every imagination of light has been held within a larger cul
tural imagination of man and world. Are we now at the watershed
to a new one, and could it possibly support a true ecology of
human, animal, plant, and mineral communities? In recent de
cades, light has taken on a new and subtle figure; one can only
hope it is but a symptom of a larger evolutionary change in the
structure of our imagination that supports an ecological con
sciousness.
WHEN I TRY to imagine light without a particular color, direction
of propagation, etc., I understand the struggles of medieval
theologians or artists as they sought to picture God, and appre
ciate their choice of light as an attribute of the divine. God’s
existence was not at issue, but his features were. Anything
definite one said about them, any image that depicted Him, was
of necessity partial and potentially misleading. Even such a
simple thing as the location of God, where God is, was full of
danger. Amazingly enough, light, too, suffers from a special
difficulty with regard to location. We have seen its other attri
butes entwined in subtle ways, but what of the simple sense of
place? As the next section shows, the very concept of place
loses meaning for light as for no other object.
320
Least Light: A Contemporary View
The Place of Light
Hast thou perceived the breadth of the Earth? declare if
thou knowst it all. Where is the way where light dwelleth?
and as for darkness, where is the place thereof?
Book of Job
In his story of world creation, before anything else could happen,
Hesiod had first to bring forth a brooding space filled with
possibility, called Chaos. In Greek, its root meaning was “gap.”
Here, in the place provided by Chaos, rose the “wide-bosomed
Earth,” as Hesiod called it, “a sure, eternal dwelling-place for
all the deathless gods who rule Olympus’s snowy peaks.”18 The
place of earth was created before the earth itself.
Everything must have location, a place where it is. Aristotle
declared that “A natural scientist must inform himself not only
on the infinite, but also on place.”19 What then is the place of
light? One would suppose that with the photon concept of light,
a straightforward response would be forthcoming; but quantum
theory and experiment again conspire to make the place of light
completely elusive.
In AN IMPORTANT paper of 1949, Eugene Wigner and T. D.
Newton, then at Princeton, looked for the place of elementary
quantum particles: electrons, protons, mesons, and photons.2
They were interested in “localized states,” that is to say, they
sought a clear way to define position for particles within the
formalism of quantum mechanics. Everything began well
enough. Elementary particles that possess mass, like the elec
tron and neutron, presented no special problems. However,
when they turned their attention to light, a sudden hitch ap
peared. They commented that light was different; they could
find no mathematical object within quantum theory that properly
321
CATCHING THE LIGHT
corresponded to the concept of place or position as we know it.
That observation, their trouble in finding the place of light,
persists to this day.
In their analysis of the laser, Marian Scully, Murray Sargent,
and Willis Lamb carefully describe how light is reflected back
and forth between the two mirrors of a laser cavity. If light could
be conceived of as corpuscular, we would naturally picture to
ourselves a kind of tennis match with balls of light bouncing
back and forth in the laser. This is not the case! In their words:
“Photons are not localized at any particular position and time
within the cavity like fuzzy balls; rather, they are spread out
over the entire cavity. In fact no satisfactory quantum theory of
photons as particles has ever been given.”21
Many times over the last sixty years, the quantum theory of
light has been searched for a way to give a position to light,
and it has consistently denied what it otherwise provides so
readily for massive particles. Why is light so recalcitrant about
declaring its position? The answer seems connected to the trans
verse nature of the electromagnetic field. Remember Fresnel’s
discovery that polarization could only be explained by consid
ering light to be a transverse wave. This innocent observation,
when carried over into the quantum theory, makes the definition
of position impossible.
It does not mean that there is no commonsense meaning to
the place of light at all, but it is always limited in a critical
way. There have been several experimental tours de force in
recent years that attempt to locate light’s position in one way or
another. One of the most dramatic and easiest to understand is
what is called “light-in-flight” or LIF holography. First dem
onstrated in 1978 by Nils Abramson in Sweden, the technique
has since been developed so far that it can produce “stop-action”
images of traveling light pulses no thicker than the breadth of
a hair.22 Yet what one sees here is a wave front, not a particle
of light, that is to say, only one spacial coordinate is well defined.
322
Least Light: A Contemporary View
By contrast, a massive particle has sharp values for all three
spacial coordinates. Not so light.
Interesting recent developments have connected an EPR-type
experiment to the question of location, not only for light, but
for all elementary quantum systems. In 1989 J. D. Franson of
Johns Hopkins University suggested a variation of the EPR
experiment in which the ambiguous attribute is not photon po
larization, but its time of emission. In terms of our twins analogy,
the ambiguity would now be in their time of departure. It is
important to realize that one is not simply ignorant of the de
parture time, but far more radically, that departure time does
not exist as an unambiguous feature of the twins. The ambiguity
in the time of photon or particle creation results in a parallel
ambiguity in its position. Franson’s argument applies to massive
particles, such as electrons and neutrons, as well as to light,
as long as they are in an entangled state. But in addition to the
fundamental ambiguity entailed by his considerations for all
particles, light has the added problem that even for disentangled
states, the concept of position is invalid.
By now it should be evident that light possesses a nature
unique to itself. Every natural assumption we make about it,
assumptions common to us from daily life, leads to errors. En
tering light, we cross into another domain, and must learn to
leave behind what we hold dear from the past, and cleave only
to the archetypal phenomena of light at every level, down to the
quantum. Particles, waves, location... all should be left, like
soiled sandals, at the threshold of the temple. The light within
is of a different order than the objects without; it inspires us to
subtle reflections not common to the marketplace. We stand like
Brunelleschi, in the portal between sanctuary and piazza. He
looked out, intent on the geometry of sight; we are turned in,
absorbed in the morphology of light. From the union of our open
imaginations with the firm facts of light will rise the offspring
of insight into it.
323
CATCHING THE LIGHT
The CONTEMPORARY ARTIST Janies Turrell has said, “Light is not
so much something that reveals, as it is itself the revelation.”
Holding to light, not to the objects it illumines, is the point.
We need time to work with it, live with it, think about it, and
see into it.
Turrell has rightly been called a light smith, forging his sculp
tural art out of pure light. His installations are architectural
spaces made bare of all but the essentials so that light itself can
become the object. Everything leads to light. Turrell speaks of
his working with light as an attempt “to create an experience
of wordless thought,” an engagement with immaterial reality. “I
have an interest in the invisible light, the light perceptible only
in the mind. A light which seems to be undimmed by the entering
of the senses. I want to address the light that we see in
dreams. . . .”23 I have a similar aspiration, that each of us be-
come a light smith.
Turrell senses the mystery of light’s insubstantial power and
fashions artistic forms with that energy. Light, he says, “has a
quality seemingly intangible, yet it is physically felt. Often
people reach out to try to touch it. My works are about light in
the sense that light is present and there; the work is made of
light. It’s not about light or a record of it, but it is light. Light
is not so much something that reveals, as it is itself the reve
lation.” In this and preceding chapters, I have made an anal
ogous attempt, drawing into view the selfless, ever-present, but
elusive actions of light.
I have written of light’s life, it remains for us to form a picture
of its character. We cannot do so abstractly but only by drawing
together the many sides it has shown us, by moving through one
aspect to the next, always asking, what is it that can show itself
in so many ways? We have watched the transformation of light
from a divine presence to a material object, we have seen the
324
Least Light: A Contemporary View
body of light become ever more subtle, immaterial, and para
doxical. Every attribute of light can and does become entwined,
entangled, and nonlocal. Color, place, polarization, intensity
... everything by which light is classically named loses its com
mon significance; a new name is needed. Circling through the
life of light, we can hold its essence only gently, as we might
hold a fledgling bird ready to fly. Fitted for air and space, it
touches the earth awkwardly. Following it in flight, running with
light, we seem caught up into its mysterious nature, at once
here and there, a connective tissue weaving together all of ex
istence, a whole whose parts are wholes themselves, a thing for
which time and space disappear. I cannot describe it, my imag
ination can only just touch its hem, but I do know that at its
core there seems to live an original or “first light” within which
wisdom dwells, a wisdom warmed by love and activated by life.
Around it a many-roomed mansion has risen, and our wanderings
there have not exhausted its riches.
As WE LEAVE light’s expansive dominions, the heavens dim and
darkness quietly falls. Within that darkness there is a silent
murmur, a still voice that whispers of yet another and unsus
pected part to light, for even utter darkness shimmers with its
force.
Dark Light
Yet mystery and manifestations arise from the same source.
This source is called darkness.... Darkness within dark
ness, the gateway to all understanding.
Lao-tzu
On May 18, 1885, the French poet Victor Hugo at age eighty-
three had a stroke. Four days later, during his death struggles,
325
CATCHING THE LIGHT
he, like Goethe, spoke of light, saying, “Here is the battle of
day against night.” Hugo’s last words continued what in life he
had always done: searched the darkest recesses of human nature
for its brightest treasures. As he died he whispered, “I see black
light.”
Is there a light in darkness? Is the night empty, void, and
dead, or does it, too, offer more than appearances suggest? If
we trusted the poets enough to ask, their reply would be unan
imous. Whether it be Novalis, Goethe, Hugo, or Nerval, the
night has always invited the poet to descend into its fathomless
seas. Beneath dark waves every stoke can strike a thousand
sparks, like the bioluminescent glow of a sailboat’s wake. Dark
seas shine in poetic imagination, and in journeying to the sub
lime, one enters through the doorway of darkness. The French
poet Gerard de Nerval wrote of it this way: “In seeking the eye
of God, I have seen a vast sphere black and fathomless, whence
the night that inhabits it shines forth on the world and continually
deepens.”24 The night shines forth with its own dark light, en
treating the bard to read by its radiance, the eye of God.
Distrusting the poet’s perceptions, we turn to the physicist
for levelheaded counsel and are perhaps surprised to learn that
they, too, speak of light in utter darkness. I am gratified, because
I would have all physicists be poets. Yet what is the physicist’s
dark light, and how do they sense it?
DURING A 1948 walk with Niels Bohr, the Dutch physicist
H.B.G. Casimir told Bohr of a difficult calculation he had just
performed. It had to do with the possibility of a force of attraction
existing between two uncharged metal plates. The calculation
had been long and arduous, but surprisingly, the result was
wonderfully simple in its final form. The plates should attract
each other inversely as the fourth power of their separation. How
326
Least Light: A Contemporary View
could something so simple arise from such a complex analysis?
Casimir was suspicious. Bohr agreed and pointed Casimir in a
totally unexpected direction. Casimir’s calculation had exam
ined the detailed mechanisms by which atomic motions in the
metal plates might induce unexpected electromagnetic fields and
thus attract a nearby plate.25 He had used the still relatively
new quantum theory of Schrodinger and the relativity of Einstein,
but Bohr urged Casimir to turn his attention away from the plates
and to the empty space around them. It would have been sense
less advice twenty-five years earlier, but in the intervening de
cades a full quantum theory of light (quantum electrodynamics)
had been developed, and one of its features was a new under
standing of the vacuum, of emptiness.
Where before the vacuum had been understood as pure emp
tiness—no matter, no light, no heat—now there was a residual
hidden energy. Take away everything, cool to absolute zero in
temperature, and still the vacuum remains, and it is shimmering
with a special kind of light. Called the “zero-point energy of
the vacuum,” it seems an essential part of quantum-field theory.
Bohr suggested that Casimir look to it, to the vacuum, for the
force between two metal plates. Casimir followed Bohr’s pre
scient advice and in a two-page derivation, he arrived by an
elegant path at what had caused him so much trouble before.26
Since Casimir’s calculation, experiments have shown the force
to occur with exactly the form he predicted.
Without launching into the details of quantum electrody
namics (or QED), I can still draw attention to a few features of
the calculation that are suggestive of a new way of understanding
emptiness. According to QED, after one has removed all matter
and all light from space, an infinite energy still remains. Since
there is no way to extract this energy out of the vacuum, theorists
dismissed it as a curious artifact of the theory, of little real
significance, at least until Casimir made his calculation. By
inserting two parallel conducting plates into the vacuum, Casi-
327
CATCHING THE LIGHT
mir showed that the structure of the vacuum could be changed.
Move one of the plates and it changes again.
One can calculate the zero-point energy of the volume between
the plates for both plate separations. It is infinity both times!
However, Casimir realized that by subtracting the one infinity
from the other, carefully, he could obtain a finite result. Infinity
minus infinity can, in some cases, give a reasonable, finite
result. This calculation proved significant not only for the spe
cific predictions it made, but also because it showed physicists
how to eliminate the infinities popping up in their QED cal
culations.27
For our purposes, Casimir’s problem suggests another way of
understanding darkness. Darkness may be a far richer, more
subtly structured fullness than we have thought until now. Re
cent theory and experiments only encourage these musings fur
ther. Even in the deepest shadow, we can seek, and perhaps
find, a hidden light.
VICTOR Hugo saw darkness as a twofold being, divided between
the fallen Islamic angel and tempter of man, Iblis, and the
heavenly human being he called Christ. Both belonged to the
night: to one belonged the emptiness of despair, to the other
the true comfort of compassion. The key to each is within us.
At seventy-five, Hugo wrote of darkness with the key in his
hand.28
0 darkness, the sky is a gloomy precinct
Whose door you close, and whose key the soul owns;
And night divides itself in half, being diabolical and holy,
Between Iblis, the black angel, and Christ, the starry Hu
man Being.
It is empty, and yet within its gloomy precinct lives the holy,
starry Human Being.
328
Least Light: A Contemporary View
We HAVE TRAVERSED an enormous span in our natural history
of light, both scientific and sacred. Like a jeweler studying a
faceted diamond whose fleeting colors change with the smallest
turn, we have held light always before us, rotating it slowly
before our eyes. It has taken on a thousand forms, changing in
our hands like the Greek god Proteus. Yet for all that, it is one
being still. Paraphrasing Herder, we can say, “That which is
called light in creation is, in all its forms and in every being,
one and the same spirit, a flame unique.”29
Across from that flame unique, however, has always been a
changing face. First the countenance was that of an ancient
Egyptian, then a Greek, a Manichaean, a Cathar, a Scholastic
bishop... a physicist and ours. Each culture brought its own
question, each the wealth and limitations of its imagination, and
the flame unique lit each according to its own nature. The history
of light is not a steady convergence to truth, nor does it end in
a relativism that makes inquiry meaningless. Insights into light
belong to every culture; truth has many homes. The history of
light mirrors the history of mind, and continues to offer fresh
aspects to new awakenings of the soul.
Again and ever again, it is ourselves whom we study in study
ing light. Thus we uncover the evolution of that knowing con
sciousness we prize so highly, but whose life and growth we too
often neglect. After we accept the history of its changing char
acter, where does mind now tend, what are the possibilities for
its conscious cultivation, and what will be the new fruits of
tomorrow’s awakenings to light?
329
CHAPTER 12
Seeing Light
Our whole business in this life is to restore
to health the eye of the heart whereby God
may be seen.1
Augustine
“The divine light penetrates the universe according to its dig
nity,” wrote Dante in his Paradiso.2 To the medieval imagina
tion, as light penetrated the realms of stone and glass, it dignified
town and village by calling into being the Gothic cathedral.
Light and geometry were the intertwined themes of these sacred
structures. For the first time, an architecture had been discov
ered that could free walls of their structural loads so they could
become the jeweled membranes of a Heavenly City. Sacred
geometry ordered the shape, but light, like the breath God
breathed into Adam, gave the cathedral life, swelling it just as
the Virgin’s womb swelled with the Light of Mankind.
As St. Bernard conceived it, the passing of light into and
through stained-glass windows was a literal reenactment of the
conception and birth of the infant Jesus by the Virgin mother.
In both instances the material kingdom was penetrated but not
violated. “As a pure ray enters a glass window and emerges
unspoiled, but has acquired the color of the glass.. . the Son
330
Seeing Light
of God, who entered the most chaste womb of the Virgin,
emerged pure, but took on the color of the Virgin, that is, the
nature of a man and a comeliness of human form, and he clothed
himself in it.”3
Two worlds, the physical and the spiritual, were one within
the sanctuary of the cathedral; and light played a leading role
in that unitary imagination of God in man’s domain. The great
art historian of the Gothic, Otto von Simson, wrote,
With its sublime theology of light it [the Mass] must have
conveyed to those who listened, a vision of the eucharistic
sacrament as divine light transfigured through the darkness
of matter. In the physical light that illuminated the sanc
tuary, that mystical reality seemed to become palpable to
the senses. The distinction between physical nature and
theological significance was bridged by the notion of cor
poreal light as an “analogy” to the divine light.4
There are levels to light, and more than one may be active
at any one moment. Some, from the past, may be alien to us,
and their union with others that are more familiar may be dis
concerting. Yet, while we may be deeply confused, the culture
we study seems oblivious to our dilemma. One feature of sci
entific progress has been the segregation of levels once inte
grated, for example, the isolation of values from scientific
knowledge, the photon from the Incarnation. This separation
has had its costs as well as its benefits.
Traditional cultures, as well as the early history of our own,
were, to use the language of the Cambridge anthropologist Ernest
Gellner, “multi-stranded.”5 When a member of the Nilotic tribe
of the Nuer looks at a cucumber and, in complete seriousness,
identifies it as a bull, he entangles himself in no logical absurdity
because he lives in a multistranded mental world. The strand
331
CATCHING THE LIGHT
of cucumber as a vegetable food is woven into, but not confused
with, the strand of cucumber as totem. The history of the West
is a history of the growing segregation of the strands of con
sciousness, the separation of the moral and spiritual from the
sensual and physical, the loss of the unity felt by the Nilotic
clansman.
The felt unity of the spiritual and physical that still pervaded
the consciousness of the thirteenth century reflects itself in early
understandings of history. The outer, historical events of the
Old and New Testaments were simultaneously a revelation of
the spirit. Entering a Gothic cathedral, one passed through a
portal whose intricate stone statuary rendered visible the history
of God’s actions on earth. In Chartres, the window depicting
the genealogy and birth of Jesus, the beginnings of things, stands
above the entryway. Moving down the nave aisle, one passes
between images drawn from the history of the Jews and the
miracles of Christ’s life. Where transept crosses nave, the sac
rifice of Christ crosses the linear axis of history through which
one has just passed. The altar railing delimits the spaces of
present and past, of the Shepherd and sheep, the priest and his
flock. Often, the iconography of the apse rising above and behind
the altar is that of the Last Judgment and new Jerusalem, our
eschatological future. The journey is through time, conceived
of as at once secular and sacred. The dismembering of this unity
is a relatively modem event.
Every early culture has understood its genesis and history as
one woven of threads both divine and mundane to form a mul
tidimensional mythology of creation, destruction, and migration.
The organization of the drama is temporal. In Hindu cosmolo
gies, world time is ordered into “yugas,” the Maya have their
“katuns,” the Aztecs their many ages and five “sun” periods.
Among the Greeks, Hesiod denominated the ages of man as
Golden, Silver, Bronze, Heroic, and Iron, the last of which is
our own. Even our word “world” stems from the Old Germanic
compound wer-aldh meaning “the life or age of man.
332
Seeing Light
To each age was associated not only external events, but a
moral order as well. The Golden Age, writes Hesiod, was a time
when men “lived like gods, carefree in their hearts, shielded
from pain and misery.. . . They knew no constraint and lived in
peace and abundance as lords of their lands.”6 After this blessed
age, a second childish and unnatural race arose, followed by a
third brazen race of mortals, who were a harsh, violent people,
full of destruction. Then came the “divine race of heroes,”
among whose number were Achilles and Odysseus. The fifth
race of men, our own, is a tragic mingling of justice and injustice,
a time of disorder, when all holy alliances between families and
friends are violated, and the hallowed patterns of life are lost
completely. Hesiod’s was a mythopoeic, multistranded history
of human nature, not a chronology of purely mundane events.
The rich, psychospiritual story of our origins has only recently
given way to another imagination. Since the rise of science in
the sixteenth and seventeenth centuries, the physical origins of
the world have gradually disentangled themselves from its spir
itual origins. Divine cosmogony found a rival in the upstart
physical cosmogony of astronomy and physics. With the ap
pearance in 1859 of Darwin’s On the Origin of the Species by
Means of Natural Selection, the battle moved closer to home,
from the realm of cold matter, planets, and stars, to plants,
animals, and Homo sapiens.
In purging our history of the sacred, we have also unwittingly
lost sight of our cognitive role in history. In our struggle for an
accurate chronology of events and a description of influences,
we have neglected the minds, or mentalities, of the figures who
enacted the history. We have neglected the “soul of history.”
At best we give a chronology of ideas, an intellectual history,
but we leave the history of their thinking (as opposed to their
thoughts) untouched.
Implicit in life in every culture is a social and symbolic
system, a tacit knowledge that, at least in part, is used to
construct reality. Over two hundred and fifty years ago, the great
333
CATCHING THE LIGHT
Italian philosopher Giovanni Battista Vico recognized the en
tanglement of mind and society when he wrote that “the world
of civil society has certainly been made by men, and that its
principles are therefore to be found with the modifications of
our own human mind.”7
Light has lived through all ages, yugas, and societies. Its
transmutations exemplify a dramatic evolution in consciousness.
What has been true for peoples is also true within the course
of an individual life.
Psychogenesis
Then again, do you not see the year assumingfour aspects,
in imitation of our own lifetime? For in early spring it is
tender and full offresh life, just like a little child. . . .
Ovid
In both Utica, New York, and Washington, D.C., hangs an
allegorical series of four paintings by the American artist Thomas
Cole, in which the seasons of life from infancy to old age are
imaged. In the early springtime of life, an infant appears in a
small boat whose shape has been sculpted into the figure of
Hours. At the helm stands a radiant angelic being who pilots
the craft and her young charge out of a dark cavern into luxuriant
growth bathed in a misty dawn light.
When, in the next panel, the infant has become a youth, the
landscape opens up into a vast, exotic, and exciting prospect.
The youth, now fully animated, yearning for the multitude of
dreams outspread before him, takes the helm, while his unnot
iced spirit guide gestures farewell from the bank. In the third
canvas, Manhood, the boat is poised at the brink of a barren
and dangerous cataract. The helm is broken, the sky is dark,
the light is menacing, and the now-mature soul seems utterly
lost. Only from the heights at the upper left of the painting does
334
Seeing Light
The Voyage of Life: Youth by Thomas Cole.
335
CATCHING THE LIGHT
there stream a faint light of hope in which we can make out the
delicate shape of his angelic companion. The final season of
life, Old Age, is rendered as a vast and lifeless expanse of rock
and water; “the stream of life has now reached the ocean to
which all life is tending,” as Cole himself characterized it.8 The
figure of Hours is broken and gone. Seated, the timeworn and
bearded figure certainly recognizes that his journey through life
is at an end. For the first time, he sees the spirit who has
accompanied him throughout. It gestures toward a brilliant field
of light, his luminous future.
In these four canvases, time has become space. As if at a
family reunion, or perhaps in a traditional peasant village, all
the ages of man are simultaneously present. By simply shifting
one’s gaze, decades can lapse, moving from infant to grand
mother, from worried father to youthful daughter, all with the
turn of an eye. The full gamut of human experience is drawn
into the lines of every face, and into the creases of every hand.
These are the images of hope and care, of perennial aspirations
and the harsh knowledge of mortality, set side by side.
Cole’s cycle made a deep impression on the American public.
Shortly after they were hung, a man of middling years was found
viewing them alone. He had been looking a long time before
remarking, with a melancholy air, “Sir, I am a stranger in this
city, and in great trouble of mind. But the sight of these pictures
has done me great good. They have given me comfort. I go away
from this place quieted, and much strengthened to do my duty.”
He had found his own life in Cole’s archetypal, artistic depiction,
and felt recognized. Someone had seen, and that helped.
Also, according to Hindu teachings, the human life is divided
into four stages. In the first, the sole responsibility of youth is
to learn; in the second, one becomes a worldly householder
active in the affairs of life; in the third, one retires to the forest
for a period of meditation; and finally, in the fourth phase, one
becomes a wandering mendicant sage.
Six centuries before Thomas Cole, but long after the rise of
336
Seeing Light
Hindu philosophy, an unknown artist painted the crypt ceiling
of the Cathedral of Anagni in central Italy, also showing four
ages to life’s span.9 Wonderfully geometrical in its composition,
a series of concentric circles spreads like ripples in a pond,
linking a peripheral cosmos with HOMO, the human being.
Shown nude and upright, the human form is ringed around with
the words mikrocosmos, id est minor mundus, “microcosm, that
is to say, a little world.” Man is a little world. The rings are
segmented neatly into four quadrants, and in each is the painted
bust of an age of man: child, youth, prime, old age. With each
is also written the corresponding psychological temperament:
sanguine, choleric, melancholic, and phlegmatic, respectively.
In every nuance of the painting, an ordering of both human life
and the natural universe announces itself.
Absent in Anagni is Cole’s powerful, dynamic play of light
and color; the figures are motionless models from whom it is
difficult to imagine much comfort could be drawn for a troubled
life. Each frame is complete unto itself. These are static stages
of life, not moments in a passage through it. But whether in
Anagni, in India, or Utica, New York, the stations of life’s
journey are objects of meditation, and these are, most notably,
inner, developmental stages whose character implicitly shapes
the character of our knowing, not an arid outer chronology. I
have tried to paint the history of light in similar colors, from
the inside as well as the outside.
Human history and biography are inevitably suffused with the
operations of the psyche. Only more recently have we grown
certain of a similar entanglement of outer nature and the human
mind. The transformations of cultures over time have had pro
found effects on the insights humanity has had into nature. We
have seen the character of successive ages reflected in the im
ages they have made of light. These form a sequence, not of
disjointed fragments, but a whole that unfolds in time, a series
of awakenings that bespeaks an inner evolutionary development.
Heraclitus was right: “It is in changing that things find repose.”
337
CATCHING THE LIGHT
What seems eternal, must be seen anew: “The sun is new each
day.” Scientific and prescientific knowledge of light have re
flected the continual metamorphosis of our inward instruments
of knowing. The very existence of that transformation suggests
the possibility of further evolution, individually and culturally,
and the possibility of relinking the moral and sensual, the phys
ical and spiritual, in a fresh, unitary imagination.
Past change occurred with little self-consciousness. Mistakes
could be left behind. The time of unconscious change is over,
as environmental and nuclear hazards daily bring home to us.
We now inhabit the entire planet, and have learned the potency
of our accomplishments. Future evolution must be shaped self
consciously. It is a new and dangerous technology.
What should be the nature of future knowledge; how will we see
light tomorrow? From all that has gone before, a fruitful avenue
seems clear. First, it will require the modest recognition that we
all are only partly sighted creatures, and so know only a part of
nature. As Novalis—who was both a poet and a mining engi
neer—wrote: “But it is vain to attempt to teach and preach Na
ture. One born blind does not learn to see though we tell him
forever about colors, lights and distant forms. Just so no one will
understand Nature who has not the necessary organ, the inward
instrument, the specific creating instrument.. . . ”10
If we lack the “necessary organ, the inward instrument” as
Novalis called it, we will need to cultivate it.
Clearsight
Finally, I must tell you that as a painter 1 am becoming
more clearsighted before nature.
Paul Cezanne1'
Cezanne often painted the same scene over and over again.
338
Seeing Light
Standing on the bank of a river, he told his son that the motifs
he saw so multiplied themselves “that I think I could occupy
myself for months without changing place.” What did Cezanne
hope to accomplish, or to see, by tirelessly reworking the same
vista? As if by way of answering, Cezanne wrote to Emile Ber
nard: “Get to the heart of what is before you.... In order to
make progress, there is only nature, and the eye is trained
through contact with her. It becomes concentric through looking
and working.”12
The eye becomes concentric, aligned with nature, through
the artist’s ceaseless action of looking and working, of struggling
to see clearly a single gesture of nature’s infinitely varied rep
ertoire, and then to paint it. In guiding the hand across a canvas,
one fashions and refines fresh senses, new capacities of mind
suited to seeing that which until then had eluded the eye. In
the end, one “gets to the heart of what is before us.” Like the
alchemist, whose outer actions were but an image of an inner
transformation, the artist, in creating outwardly, simultaneously
accomplishes an equally precious inner work—clearsight.
Sitting quietly beneath the shade of a tree, a Buddhist monk
meditates on one of the forty traditional subjects for meditation
that form the basis of Buddhist practice. At knee level before
him is the earth kasina, a disk of earth alike in all outward
respects to a carefully crafted mud pie, and the object of his
unswerving contemplation. As he meditates on the earth kasina
it gradually disappears. In its place hovers an “afterimage” that
initially may last only a few minutes, but that in the end becomes
as firm in the mind as the mud pie is in the shade of the tree.
Following the earth kasina further leads the meditant through a
series of disappearances and new appearances that, for the
Buddhist adept, demarcate transitions to realms beyond every-
339
CATCHING THE LIGHT
day consciousness. Over the course of months and years, if
necessary, the monk will follow the images of earth from one
realm to another until he, like Cezanne, has become sufficiently
clearsighted to get to the heart or essence of earth. Then the
sign of earth becomes, as Buddhaghosa described it, “like the
disc of a mirror, a well-burnished conch-vessel, the round moon
issuing from the clouds, white cranes against a rain cloud, and
makes its appearance as though bursting the grasped sign, than
which it is a hundred times, a thousand times more purified.”13
Among the nine other kasinas or devices (which include water,
air, heat, blue, yellow, red, white, and space) is also the kasina
of light. Of it, Buddhaghosa writes: “he who grasps the Light
device grasps the sign in light entering through a wall-crevice,
key-hole, or window-space.”14 That is, every manifestation of
light is potentially the occasion for the true grasping of light,
be it the dappled disks of light beneath the shade of a tree, or
the moonbeam that furtively makes its way through a chink in
the wall. Each instance offers an occasion for enlightenment,
for seeing light.
William Blake boldly put it this way: “If the doors of per
ception were cleansed, everything would appear to man as it is,
infinite.”
The PHILOSOPHER Schopenhauer once recorded a remarkable
conversation between Goethe and himself concerning light.
Schopenhauer sensibly suggested that light is a purely subjec
tive, psychological phenomenon, and that without sight, light
could not be said to exist. Goethe responded vehemently, as
Schopenhauer describes it: “ ‘What,’ he [Goethe] once said to
me, staring at me with his Jupiter-like eyes. ‘Light should only
exist in as much as it is seen? No! You would not exist if the
light did not see you!’ ”15
Goethe as botanist well knew the life-giving powers of light.
340
Seeing Light
In addition, he felt that light brought forth not only life but also
could, through its ceaseless action, create the very organ suited
to perceive it. Evolution has occurred in the context of light,
and over time the body responded with the organ of sight. Goethe
phrased it this way: “The eye owes its existence to the light.
Out of indifferent animal organs the light produces an organ to
correspond to itself; and so the eye is formed by the light for
the light so that the inner light may meet the outer.”16 Light,
ever active, created the eye. It sculpted an organ suited to itself,
like the streaming water shaping the stones over and through
which it flows. Had light not “seen” man, we should never have
seen the light.
What Goethe pointed to at the concrete level of human phys
iology is likewise true for more subtle organs of soul. Cezanne’s
looking and working, the Buddhist in his meditations, Goethe
through his ceaseless artistic and scientific undertakings, each
of these aimed at the birthing of the “necessary organ, the inward
instrument” needed to see more deeply into nature. And the
opportunity for this practice is ever-present, for, as Goethe says,
“Every object, well contemplated, opens up a new organ within
us.”17
The artist and the monk both know that through a disciplined
practice they can internalize nature so that they can realize new
capacities of mind. Personal growth is not only a matter of
memorizing sacred texts (which the monk has certainly done),
or of academic artistic analysis (which Cezanne has also done),
but it requires praxis, daily labor, to fashion fresh, hard-
won, soul faculties. Every action of the hand and the eye sculpts
the soul. Piaget called the process accommodation, the devel
opment of new cognitive structures. Goethe, Novalis, Emerson,
Steiner. . . spoke of new senses that would reveal fresh aspects
of nature’s infinite being.
The artist and monk are distinct from us not because of what
takes place in them, but because they apply themselves self-
341
CATCHING THE LIGHT
consciously to transformation. They educate themselves for an
end they have chosen. By contrast, most of us are educated by
others and for ends we have not chosen. Traditionally, artists,
philosophers, and religious figures have formed that small, self-
conscious segment of society that engages in the important if
often painful business of introspection and prophetic critique.
They sense the dangers of an unthinking, habitual mode of
seeing and know the need for tireless renewal.
If EVERY CULTURE and period have thought so variously about
light, and if quantum theory has stripped light of its naively
imagined attributes, then what of certainty is left to the nature
of light at the end of this evolutionary drama? Everything. The
true artist, monk, and scientist are not searching to grasp knowl
edge as object, but rather as event. The moment of critical
importance is the moment pointed to by Goethe, the moment of
aperqu, or insight. For millennia one can see the sun rise and
never notice the rotation of the earth. One can throw a thousand
rocks and never see their parabolic flight. We can wake each
morning for sixty years to the glow of the dawn and never see
light. Why? Because one rushes past the immediate offerings
of the senses to what we suppose to be the hidden, enduring,
primary objects of reality. The habits of our culture, the dogmas
of our education constrain our sight. Atoms become immortal
gods, photons their stem messengers.
Knowledge is not an object to be traded like chattel, but an
epiphanous moment to be cherished. Too often we bypass epi
phanies for a currency more ready to hand, for an abstract notion,
an old insight in new dress, an equation whose characters could
reveal, but which, in fact, we allow to go unread. These are the
stones of knowing instead of the bread, they are idols instead
of gods. Epiphanous knowing presupposes organs for insight,
inward instruments; and new knowing requires new instruments.
342
Seeing Light
We each possess the rudiments of every organ, but we deny
them the nurture they need, we neglect the praxis in whose light
they might grow and flourish.
The photon is not an object to be hefted in the hand like a
clod of earth. Its very elusiveness draws us to the essentials of
knowing as seeing, to knowledge as event. Understood in this
way, insights into light’s multifaceted nature are not the exclu
sive property of twentieth-century physics. Our present insights
do not exhaust the wealth of light, but complement the genuine
insights of the past. The future also becomes clearer. The ar
chetypal experiments of quantum optics and the meditations of
the contemporary poet offer us equally the opportunity for fresh
aper<?u, but only if we have the patience to educate ourselves
in light, to make ourselves concentric with its nature, as Cezanne
and Buddhaghosa recommend.
OVER MILLENNIA, CULTURES have embraced and discarded count
less images of light. Within a single lifetime, likewise, we have
lived within and shed successive understandings of light.
Through research, artistic praxis, and quiet contemplation,
light’s elusive being constantly re-creates itself in our mind’s
eye, offering fresh epiphanies to every generation. When seen
with a thousand eyes, light will finally rest with us in the haven
we have made.
Seeing light is a metaphor for seeing the invisible in the
visible, for detecting the fragile imaginal garment that holds our
planet and all existence together. Once we have learned to see
light, surely everything else will follow.
343
Notes
Chapter 1 I Entwined Lights:
The Lights of Nature and of Mind
1. Lao-tzu, Tao Te Ching, quoted in Hastings, Encyclopedia of
Religion (New York: Charles Scribner’s Sons, 1908—26), vol.
8, p. 51.
2. Quoted in M. von Senden, Space and Sight: The Perception
ofSpace and Shape in the Congenitally Blind before and after
Operation, trans. Peter Heath (Glencoe, IL: The Free Press,
1960), p. 40.
3. R. L. Gregory and J. G. Wallace, “Recovery from Early
Blindness: A Case Study,” in Perception, ed. Paul Tibbetts
(New York: Quadrangle/New York Times Book Co., 1969).
4. M. von Senden, Space and Sight, p. 20.
5. David H. Hubei, Eye, Brain and Vision (New York: Scientific
American Library, distributed by W. H. Freeman, 1988),
chap. 9.
6. Quoted in M. von Senden, Space and Sight, p. 160.
7. Marjorie Hope Nicolson, Newton Demands the Muse (Prince
ton, NJ: Princeton University Press, 1966), p. 75.
345
Notes
Chapter 2 I The Gift of Light
1. Plato, “Protagoras,” sections 320d—32le. The Collected
Dialogues of Plato, ed. Edith Hamilton and Huntington
Caims (Princeton, NJ: Princeton University Press, 1969).
2. Hesiod, Works and Days, trans. Apostolos N. Athanassakis
(Baltimore: Johns Hopkins University Press, 1983), lines
50-100.
3. Plato, “Symposium,” 219a.
4. Homer, The Odyssey, trans. Robert Fitzgerald (Garden City,
NY: Doubleday, 1961), bk. 3, lines 1—4.
5. Eleanor Irwin, Color Terms in Greek Poetry (Toronto: Hak-
kert, 1974).
6. Homer, The Iliad, trans. Robert Fitzgerald (Garden City,
NY: Doubleday, 1974), bk. 24, lines 401—03.
7. Homer, Iliad, bk. 24, lines 93-94.
8. Oliver Sacks and Robert Wasserman, “The Case of the
Colorblind Painter,” The New York Review of Books, vol.
34 (November 19, 1987), pp. 25—34.
9. Benjamin Whorf, Language, Thought and Reality (Cam
bridge: MIT Press, 1964).
10. Diogenes Laertius, Lives of the Philosophers, ed. and trans.
A. Robert Caponigri (Chicago: Henry Regnery Co., 1969),
chap. 8.
11. See E. R. Dodds, The Greeks and the Irrational (Berkeley,
CA: University of California Press, 1951), pp. 145—46.
12. Empedocles, Katharmoi (Purifications) in Kathleen Free
man, Ancilla to the Pre-Socratic Philosophers (Cambridge:
Harvard University Press, 1983), p. 65, frag. 115.
13. Empedocles, On Nature, in Freeman, Ancilla, p. 61, frags.
85-87.
14. Freeman, Ancilla, pp. 60-61, frag. 84.
15. The Revised Standard Version of the Bible translates Mat
thew 6:22—23 as “The eye is the lamp of the body. So if
346
Notes
your eye is sound, your whole body will be full of light;
but if your eye is not sound, your whole body will be full
of darkness. If then the light in you is darkness, how great
is the darkness.”
16. Freeman, Ancilla, p. 58, frag. 48.
17. Plato, Timaeus, 45b; David C. Lindberg, Theories of Vision
from al-Kindi to Kepler (Chicago: University of Chicago
Press, 1976), chap. 1.
18. This is a paraphrase of a Neo-platonic text. See J. W. von
Goethe, Theory of Color in Scientific Studies, ed. and trans.
Douglas Miller, vol. 12 of Goethe: Collected Works in En
glish (New York: Suhrkamp, 1988), p. 164.
19. Paul Friedlaender, Plato, an Introduction, trans. Hans
Meyerhoff (Princeton, NJ: Bollingen, 1973), p. 13.
20. De Lacy Evans O’Leary, How Greek Science Passed to the
Arabs (London: Routledge & Kegan Paul, 1964).
21. On Ibn al-Haytham or Alhazen see the article by A. I.
Sabra in The Dictionary ofScientific Biography, ed. Charles
Gillispie (New York: Charles Scribner’s Sons, 1972), vol.
6, pp. 189-210; and Lindberg, Theories of Vision, pp. 60-
86.
22. Quoted in Dictionary ofScientific Biography, vol. 6, p. 190.
23. Lindberg, Theories of Vision, p. 91.
24. A. C. Crombie, “Early Concepts of the Senses and the
Mind,” Scientific American, vol. 210, no. 5 (May 1964),
pp. 108—16; see also Lindberg’s objections to Crombie in
Theories of Vision, p. 207.
25. Lindberg, Theories of Vision, p. 66.
26. The camera obscura was probably known in some form in
antiquity. References in Euclid and the pseudo-Aristotle
Problemata indicate some familiarity with it, but Alhazen
appears to be the first to describe it fully. See John Ham
mond, The Camera Obscura: A Chronicle (Bristol, CT: Hil-
ger, 1987), pp. 1—7.
347
Notes
27. Johannes Kepler, quoted in Lindberg, Theories of Vision,
p. 203. Kepler combined his work with Felix Plater’s rec
ognition of the retina (instead of the lens) as the sensitive
tissue of the eye.
28. Lindberg, Theories of Vision, p. 203.
29. Lindberg, Theories of Vision, p. 201.
30. Hubei, Eye, Brain, and Vision, p. 222.
31. Owen Barfield, Saving the Appearances (New York: Har
court, Brace & World, 1965).
Chapter 3 I Light Divided:
Divine Light and Optical Science
1. C. S. Lewis, The Discarded Image (Cambridge: Cambridge
University Press, 1964), pp. 222-23.
2. Ahura Mazda in Persian mythology has the sun for his eye.
See M. N. Dhalla, History of Zoroastrianism (New York:
Oxford University Press, 1938), p. 213.
3. W. Max Mueller, Egyptian Mythology in The Mythology of
All Races, ed. Louis Herbert Gray (Boston: Marshall Jones
Co., 1943), vol. 12, p. 70.
4. Veronica Ions, Egyptian Mythology (New York: Paul Ham-
lyn, 1975), p. 41.
5. Mary Boyce, ed. and trans., Textual Sources for the Study
of Zoroastrianism (Totowa, NJ: Barnes and Noble Books,
1984), p. 75. Some authors, ancient and modem, put Zo
roaster back as far as 6000 B.C.
6. Mary Boyce, Zoroastrians: Their Religious Beliefs and Prac
tices (London: Routledge & Kegan Paul, 1979); and Boyce,
Textual Sources for the Study of Zoroastrianism.
7. RSV, Genesis 1:14-19.
8. RSV, Genesis 3:16—19.
9. Jeffery Burton Russell, The Devil: Perceptions of Evil from
348
Notes
Antiquity to Primitive Christianity (Ithaca, NY: Cornell Uni
versity Press, 1978), pp. 207—09.
10. RSV, 1 John 1:5.
11. RSV, John 1:67-69.
12. Jes P. Asmussen, Manichaean Literature (Delmar, NY:
Scholars’ Facsimiles & Reprints, 1975), p. 88.
13. Samuel N. C. Lieu, Manichaeism in the Later Roman Em
pire and Medieval China (Manchester, U.K.: Manchester
University Press, 1985), pp. 210—13.
14. The Cologne Mani Codex, “Concerning the Origin of his
Body,” ed. and trans. Ron Cameron and Arthur J. Dewey
(Missoula, MT: Scholars Press, 1979), p. 19.
15. G. Van Groningen, First Century Gnosticism: Its Origins
and Motifs (Leiden, Neth.: E. J. Brill, 1967), pp. 23ff.
16. “Two Cathars Tales,” Journal for Anthroposophy, trans.
Christopher Bamford, no. 36 (Autumn 1982).
17. Zoe Oldenbourg, Massacre at Montsegur, trans, by Peter
Green (London: Wiedenfeld & Nicolson, 1961), p. 50.
18. Steven Runciman, The Medieval Manichee (Cambridge:
Cambridge University Press, 1982).
19. Denis de Rougemont, Love in the Western World (Princeton,
NJ: Princeton University Press, 1983), p. 82.
20. James McEvoy, The Philosophy of Robert Grosseteste (Ox
ford: Clarendon Press, 1982).
21. Alexandre Koyr6, Diogenes 4 (1957), pp. 421—48; and
McEvoy, The Philosophy of Robert Grosseteste, p. 210.
22. For the importance of light and measure in Gothic archi
tecture see Otto von Simson, The Gothic Cathedral (Prince
ton, NJ: Princeton University Press, 1988), chapter 2.
23. The Bible, Wisdom 11:21; this was a favorite text of Grosse
teste’s. On Grosseteste’s view of God as mathematician, see
McEvoy, pp. 167—80.
24. Louis Kahn, interviewed in Time, January 15, 1973.
25. Robert Grosseteste, quoted in A. C. Crombie, Robert
349
Notes
Grosseteste and the Origins of Experimental Science, 100-
1700 (Oxford: Clarendon Press, 1953), p. 143.
26. Crombie, Experimental Science, pp. 10—11.
27. See Lindberg, Theories of Vision, chap. 2.
28. McEvoy, Philosophy of Robert Grosseteste, p. 372.
Chapter 4 / The Anatomy of Light
1. Samuel Edgerton, Jr., The Renaissance Rediscovery of Lin
ear Perspective (New York: Harper & Row, 1976), espe
cially chaps. 1 and 10. Also see John White, The Birth
and Rebirth of Pictorial Space (London: Faber & Faber,
1967).
2. Martin Kemp, The Science of Art (New Haven: Yale Uni
versity Press, 1990), p. 9.
3. Antonio di Tuccio Manetti, The Life of Brunelleschi, trans.
Catherine Enggass (University Park, PA: Pennsylvania
State University Press, 1970), pp. 42—46.
4. Ernst Cassirer, Language and Myth, trans. Suzanne K.
Langer (New York: Harper & Brothers, 1946), p. 8.
5. Jan Deregowski, “Pictorial Perception and Culture,” re
printed in Image, Object and Illusion (San Francisco: W. H.
Freeman & Company, 1974), chap. 8.
6. Suzi Gablik, Progress in Art (New York: Rizzoli, 1976),
chaps. 6-7.
7. Edgerton, The Renaissance Rediscovery of Linear Perspec
tive, chap. 10, pp. 157—58, summarizing the article by
Erwin Panofsky, “Die Perspective als ‘symbolische
Form,’ ” Vortraege der Bibliotek Warburg 1924—25 (Leip
zig, Ger.: 1927), pp. 258—331.
8. Quoted by Richard Krautheimer and Trude Krautheimer-
Hess, Lorenzo Ghiberti (Princeton, NJ: Princeton University
Press, 1956), p. 14.
350
Notes
9. Leonardo da Vinci, The Notebooks of Leonardo da Vinci,
ed. and trans. Edward MacCurdy (New York: George Bra-
ziller, 1955), p. 93.
10. Otto von Simson, The Gothic Cathedral (Princeton, NJ:
Princeton University Press, 1988), chap. 2.
11. Ernst Cassirer, The Individual and the Cosmos in Renais
sance Philosophy, trans, by Mario Domandi (Philadelphia:
University of Pennsylvania Press, 1963), chaps. 1 and 2.
12. Leonardo, Notebooks, p. 57.
13. Leonardo, Notebooks, p. 612.
14. See the Masonic Temple Legend for interesting background
in Charles William Heckethom, Secret Societies (New Hyde
Park, NY: University Books, 1965), vol. 1, bk. 8, chap.
1.
15. George Ovitt, Jr., The Restoration ofPerfection (New Bruns
wick, NJ: Rutgers University Press, 1987); and Lynn
White, Jr., Medieval Technology and Social Change (New
York: Oxford Galaxy Book, 1966).
16. Galileo Galilei, Dialogue, trans. Stillman Drake (Berkeley,
CA: University of California Press, 1953).
17. Stillman Drake, Galileo (New York: Hill & Wang, 1980),
p. 24.
18. Galileo Galilei, “Letter to the Grand Duchess Christina,”
Discoveries and Opinions of Galileo, trans. Stillman Drake
(Garden City, NY: Doubleday & Co., 1957), p. 182.
19. Homer, Odyssey, bk. 6, line 243.
20. Homer, Odyssey, bk. 8. Odysseus, modem soul that he
was, warned that looks can be deceiving: “In looks a man
may be a shade, a specter, and yet be master of speech
so crowned with beauty that people gaze at him with
pleasure.”
21. Johannes Kepler, Epitome of Copernican Astronomy, bk.
IV, I, 3.
22. Plato, Laws, bk. VII; see also Francis Cornford, Before and
351
Notes
After Socrates (Cambridge: Cambridge University Press,
1972).
23. Galileo Galilei, The Assayer in Discoveries and Opinions,
p. 274.
24. The technology is slightly more complicated, relying on a
change in the polarization of the light as it passes through
the liquid crystal.
25. We would call it barium sulfide.
26. Galileo Galilei, The Assayer in Discoveries and Opinions,
p. 278.
27. Richard S. Westfall, Never at Rest: A Biography of Isaac
Newton (Cambridge: Cambridge University Press, 1980),
pp. 232ff.
28. Westfall, Never at Rest, p. 141.
29. Westfall, Never at Rest, p. 154.
30. Westfall, Never at Rest, p. 426.
31. Rudolf Steiner, Philosophic und Anthroposophie, “Mathe-
matik und Okultismus” (Domach, Switz.: Verlag der Rudolf
Steiner-Nachlassverwaltung, 1965).
32. Alan E. Shapiro, “Newton’s Definition of a Light Ray,” Isis,
vol. 66 (1975), pp. 194-210.
33. Isaac Newton, Opticks, 4th edition (New York: Dover,
1952), p. 1.
34. A. I. Sabra, Theories of Light from Descartes to Newton
(New York: Cambridge University Press, 1981), chap. 11,
for example, argues a dogmatic atomism of light.
35. Shapiro, “Newton’s Definition of a Light Ray.”
36. G. N. Cantor, Optics After Newton (Dover, NH: Manchester
University Press, 1983), chap. 2.
37. Newton to Robert Hooke, letter of February 5, 1676, quoted
in Westfall, Never at Rest, p. 274.
38. Marjorie Hope Nicolson, Newton Demands the Muse
(Princeton, NJ: Princeton University Press, 1946).
39. John Hughes, The Ecstasy (1735), quoted in Nicolson, New
ton Demands the Muse, p. 11.
352
Notes
40. See, for example, “To the Memory of Sir Isaac Newton,”
in The Complete Poetical Works ofJames Thomson, ed. J. L.
Robertson (London: 1908); or Nicolson, Newton Demands
the Muse, p. 12.
41. From Journal de Trevoux (Paris: Chez E. Ganeau, 1701).
42. Quoted in Stephen Mason, A History of the Sciences (New
York: Collier Books, 1962), p. 169.
43. Jacques Maritain, The Dream ofDescartes (New York: Phil
osophical Library, 1944). See chap. 1 for an engaging ac
count of the dream.
44. Francis A. Yates, The Rosicrucian Enlightenment (Boulder,
CO: Shambhala, 1978).
45. Quoted in Mason, History of the Sciences, p. 169.
46. A. I. Sabra, Theories of Light, p. 48.
47. Cantor, Optics After Newton, chap. 3.
48. Cantor, Optics After Newton, p. 33.
49. Sabra, quoting Descartes; Theories of Light, p. 60.
50. Bernard de Fontenelle, Conversations on the Plurality of
Worlds, trans. H. A. Hargreaves (Berkeley, CA: University
of California Press, 1990), “The First Evening.”
Chapter 5 / The Singing Flame:
Light as Ethereal Wave
1. Edward Grant, ed., A Source Book in Medieval Science
(Cambridge: Harvard University Press, 1974). See Bartho
lomew the Englishman “Concerning the Properties of
Things,” trans. Bruce Eastwood, p. 383; also Vasco Ron-
chi, The Nature of Light, trans. V. Varocas (Cambridge:
Harvard University Press, 1970), p. 62; and Lindberg,
Theories of Vision, p. 134.
2. Leonardo da Vinci is here quoting John Pecham.
3. Rene Descartes, in Sambursky’s An Anthology of Physical
Thought (New York: Pica Press, 1975), p. 244.
353
Notes
4. Leonhard Euler, Letters of Euler on Natural Philosophy
Addressed to a German Princess, ed. David Brewster (New
York: J. & J. Harper, 1833), p. 77.
5. Euler, Letters to a German Princess, p. 85. The notion of
light as vibration existed before Euler, most especially in
the thought of Christian Huygens, but it did not successfully
challenge Newton’s corpuscular view.
6. Adolf Katzenellenbogen, The Sculptural Programs of
Chartres Cathedral (Baltimore: Johns Hopkins, The Uni
versity Press, 1959), pp. 15—22.
7. John Hawkins, General History of the Science and Practice
of Music (New York: Dover, 1963), vol. II, chap. 133.
Joscelyn Godwin, Harmonies of Heaven and Earth (Roch
ester, VT: Inner Traditions, 1987).
8. Francis Bacon, quoted in Dayton C. Miller, Anecdotal His
tory of the Science of Sound (New York: Macmillan, 1935),
p. vi.
9. Aristotle, Sound and Hearing, quoted in Miller, Anecdotal
History of Sound, p. 3.
10. John Leconte, “On the Influence of Musical Sounds on the
Flame of a Jet of Coal-Gas,” Philosophical Magazine, 4th
series, vol. 15 (1858), p. 235.
11. Maeterlinck quoted in Arthur Symons, The Symbolist Move
ment in Literature (New York: E. P. Dutton & Co., 1919),
p. 92.
12. Adelard of Bath, “Natural Questions,” in A Source Book in
Medieval Science, ed. Edward Grant.
13. Alexander Wood, Thomas Young, Natural Philosopher
(1773-1829) (Cambridge: Cambridge University Press,
1954); Dictionary ofScientific Biography, vol. 14, pp. 562-
72.
14. Cantor, Optics After Newton, p. 146.
15. Otto Neugebauer, The Exact Sciences in Antiquity (Provi
dence, RI: Brown University Press, 1957).
354
Notes
16. R. J. Gillings, “The So-called Euler-Diderot Incident,”
American Mathematical Monthly, vol. 61 (1954), pp. 77-
80.
17. Francois Arago, Biographies of Distinguished Scientific
Men, trans. Smyth, Powell, and Grant (London: Longman,
1857), pp. 399-471.
18. Robert Greenler, Rainbows, Halos, and Glories (New York:
Cambridge University Press, 1980), chap. 6.
19. Sound waves are “longitudinal waves,” which means that
the vibration is in the same direction as the direction of
propagation and not transverse to it.
20. Sir George Stokes, “On the Constitution of the Ether,” May
1848, in Nineteenth-Century Aether Theories, ed. Kenneth
F. Schaffner (New York: Pergamon Press, 1972).
21. G. N. Cantor, “The Theological Significance of Ethers,”
Conceptions of Ether, eds. G. N. Cantor and M. J.S. Hodge
(Cambridge: Cambridge University Press, 1981), p. 149.
22. Spiritualism sought for physically demonstrable effects of
the spirit, and so materialized psychical research and ex
perimentation. It clearly remained true to the materialistic
imagination of the nineteenth century, even when treating
the immaterial.
Chapter 6 / Radiant Fields:
Seeing by the Light of Electricity
1. Joseph Agassi, Faraday as a Natural Philosopher (Chicago:
University of Chicago Press, 1971); and L. Pearce Wil
liams, Michael Faraday (New York: Basic Books, 1965).
I have relied heavily on L. Pearce Williams’s biography of
Faraday throughout this chapter.
2. John Tyndall, Faraday as a Discoverer (New York: D. Ap-
pieton and Co., 1873).
3. Quoted by G. N. Cantor, “Reading the Book of Nature: The
355
Notes
Relation Between Faraday’s Religion and His Science,” in
Faraday Rediscovered (New York: Stockton Press, 1985),
eds. David Gooding and Frank James, p. 71.
4. Tyndall (1894), quoted by Cantor, “Reading the Book of
Nature,” p. 74.
5. Cantor, “Reading the Book of Nature,” p. 74.
6. Michael Faraday, Experimental Researches in Electricity
(London: Richard Taylor & William Francis, 1855),
vol. 2, pp. 284ff.
7. See for example the account of Frank A.J.L. James, ‘“The
Optical Mode of Investigation’: Light and Matter in Fara
day’s Natural Philosophy,” in Faraday Rediscovered,
p. 149.
8. Faraday, Experimental Researches, vol. 3, pp. 447ff.
9. See for example Faraday’s June 1852 paper, “On the Phys
ical Lines of Magnetic Force,” in Experimental Researches,
vol. 3, pp. 438ff.
10. Faraday, Experimental Researches, vol. 2, p. 451.
11. Some scholars argue that Faraday advanced in turn two
types of field theories. One operated through empty space
by the very properties of the field or space itself (his 1846
view), and later, perhaps under the influence of W. Thomp
son (Kelvin), Faraday changed his view completely to that
of a continuous, nonparticulate ether. (See Barbara Giusti
Doran, “Origins and Consolidation of Field Theory in Nine
teenth-Century Britain,” in Historical Studies in the Phys
ical Sciences, [Princeton, NJ: Princeton University Press,
1975], vol. 6, pp. 133-260.) I disagree. From reading the
sources, I suspect that the second view was never really
adopted by Faraday. When he does mention the ether after
1846, he usually surrounds the reference with so many
qualifiers that he effectively keeps it at arm’s length, while
accommodating the enthusiasms of colleagues such as Kel
vin.
356
Notes
12. Faraday, Experimental Researches, vol. 2, p. 452.
13. Boethius, The Consolation of Philosophy, trans. V. E.
Watts (Baltimore: Penguin, 1969).
14. Alan of Lille, Plaint of Nature, trans. James J. Sheridan
(Toronto: Pontifical Institute, 1980).
15. See the “Night” scene in Johann Wolfgang von Goethe,
Faust, ed. and trans. Stuart Atkins (Boston: Suhrkamp/
Insel, 1984).
16. See “A Vision” in the poems of James Clerk Maxwell
printed at the end of The Life of James Clerk Maxwell,
Lewis Campbell and William Garnet (London: Macmillan
and Co., 1882).
17. Sir Thomas Browne, Religio Medici and Other Works, ed.
L. C. Martin (Oxford: Clarendon Press, 1964).
18. Quoted in Ivan Tolstoy, James Clerk Maxwell (Edinburgh:
Canongate, 1981), p. 59.
19. Friedrich Holderlin, Brot und Wein in Poems and Frag
ments, trans. Michael Hamburger (Cambridge: Cambridge
University Press, 1980).
20. Albert Einstein, “Maxwell’s Influence on the Development
of the Conception of Physical Reality,” in Sir J. J. Thom
son, James Clerk Maxwell (New York: Macmillan, 1931),
pp. 66-67.
21. Richard Feynman, Lectures on Physics (Reading, MA:
Addison-Wesley, 1965), vol. 2.
22. Jed Z. Buchwald, From Maxwell to Microphysics: Aspects
of Electromagnetic Theory in the Last Quarter of the Nine
teenth Century (Chicago: University of Chicago Press,
1985).
23. On William Thompson, Lord Kelvin, see Harold Issadore
Sharlin, Lord Kelvin: The Dynamic Victorian (University
Park, PA: Pennsylvania State University Press, 1979), and
David B. Wilson, Kelvin and Stokes (Bristol, CT: Adam
Hilger, 1987).
357
Notes
24. Willard Gibbs, quoted in Dictionary ofScientific Biography,
vol. 13, p. 386.
25. Sharlin, Lord Kelvin, p. 237.
26. Quoted in Sharlin, Lord Kelvin, p. 226.
27. Tyndall, Faraday as a Discoverer, p. xvii.
28. John Stuart Mill, Autobiography, 1924, p. 129. Quoted in
“The ‘Spectre’ of Science,” by C. J. Wright, Journal of the
Warburg and Courtauld Institutes (1980), vol. 43, p. 187.
29. Henry David Thoreau, The Journals of Henry D. Thoreau,
ed. Francis H. Allen and Bradford Torrey (Boston: Hough
ton Mifflin Co., 1906), vol. Ill, pp. 155-56.
30. Quoted in Keiji Nishitani, Religion and Nothingness, trans.
Jan Van Bragt (Berkeley, CA: University of California
Press, 1982), p. 167.
31. RSV, Genesis 9:13.
1
Chapter 7 / Door of the Rainbow
1. Black Elk, Black Elk Speaks (Lincoln, NE: University of
Nebraska Press, 1988), p. 25.
2. Homer, Iliad, bk. 2.
3. Flood myths are extraordinarily widespread, ranging from
Plato’s story of the lost continent of Atlantis to South Amer
ican Indian legends. For further discussion see The Flood
Myth, ed. Alan Dundes (Berkeley, CA: University of Cal
ifornia Press, 1988).
4. Plato, Theatetus, 155d.
5. Xenophanes, frag. 28D.
6. December 10, 1834, off Chonos Archipelago. Charles Dar
win, Journal of Researches during the Voyage of HM.S.
Beagle (New York: Hafner Publishing Co., 1952), p. 269.
7. Aristotle, Meteorologica, trans. H. D. P. Lee (Cambridge:
Harvard University Press, 1952), bk. II, 371b, line 19.
8. Carl B. Boyer, The Rainbow: From Myth to Mathematics
358
Notes
(Princeton, NJ: Princeton University Press, 1989); Robert
Greenler, Rainbows, Halos, and Glories.
9. Boyer, The Rainbow, pp. 89ff.; a translation of De hide is
contained in Bruce S. Eastwood, “Robert Grosseteste on
Refraction Phenomena,” American Journal of Physics, vol.
38, pp. 196ff (1970).
10. Grosseteste, in Eastwood, American Journal of Physics,
p. 196.
11. Eastwood, American Journal of Physics, p. 198.
12. Theodoric of Freiberg, quoted in Boyer, The Rainbow,
p. 114.
13. Boyer, The Rainbow, p. 116.
14. James Thomson, The Complete Poetical Works of James
Thomson, ed. J. L. Robertson (New York: 1908), pp. 436-
42.
15. See “Newton’s Rainbow and the Poet’s” in M. H. Abrams,
The Mirror and the Lamp (Oxford: Oxford University Press,
1953).
16. The Autobiography and Memoirs of Benjamin Haydon, ed
ited from his Journal by Tom Taylor, a new edition with
an introduction by Aldous Huxley (New York: Harcourt,
Brace, 1926), vol. 1, p. 269.
17. Thomas Campbell, “The Rainbow” (1820).
18. Lao-tzu, Tao Te Ching, trans. Stephen Mitchell (New York:
Harper & Row, 1988), p. 52.
19. Owen Barfield, Saving the Appearances; A Study in Idolatry
(New York: Harcourt, Brace & World, 1965).
20. J. W. von Goethe, Zur Farbenlehre, in Goethes Werke,
Hamburger Ausgabe, ed. Erich Trunz (Munich, Ger.:
C. H. Beck, 1982), vol. 13, my translation. See also J. W.
von Goethe, “Introduction to Theory of Color" in Scientific
Studies, p. 164.
21. Michael Polanyi, Knowing and Being (Chicago: University
of Chicago Press, 1969), p. 148.
359
Notes
22. John Stuart Mill, Mill on Bentham and Coleridge, ed. F. R.
Leavis (London: Chatto & Windus, 1950), pp. 99—100.
23. J. W. von Goethe, Faust, trans. Albert G. Latham (New
York: E. P. Dutton, 1908), pt. II, p. 15.
Chapter 8 / Seeing Light—Ensouling Science:
Goethe and Steiner
1. Edwin Land, Proceeding of the National Academy of Sci
ences, vol. 45 (1959), pp. 115—29, 636—44; and Scientific
American, vol. 200 (May 1959), pp. 84—99.
2. Michael Wilson and R. W. Brocklebank, The Journal of
Photographic Science, vol. 8 (1960), pp. 141—50, and Con
temporary Physics, vol. 3 (1961), pp. 91-111.
3. J. W. von Goethe, from a letter to the artist Josef Carl
Stieler, in Goethe’s Color Theory, arr. and ed. by Rupprecht
Matthaei, trans, by Herb Aach (New York: Van Nostrand
Reinhold, 1971), p. 202.
4. This episode is recounted in “Confessions of the Author,
included in Goethe’s Color Theory, p. 199.
5. J. P. Eckermann, Conversations with Goethe, trans. Gisela
C. O’Brien (New York: Frederick Ungar, 1964), Thursday,
February 19, 1829, p. 149.
6. Goethe, Scientific Studies, pp. 168ff.
7. Discussed in Rudolf Magnus, Goethe as Scientist, trans.
Heinz Norden (New York: Schuman, 1949).
8. In recording a conversation with Goethe on February 1,
1827, Eckermann states: “This led our conversation to a
great law which pervades all nature, and on which all life
and all joy depend. “This [after-imaging],’ said Goethe, ‘is
the case not only with all our other senses, but also with
our higher spiritual nature; and it is because the eye is so
eminent a sense, that this law of required change is so
striking and so especially clear with respect to colors.’”
360
Notes
J. P. Eckermann, Conversations of Goethe with Eckermann
and Soret, trans. J. Oxenford (London: Bell, 1883), rev.
ed., pp. 216-17.
9. Goethe, Scientific Studies, p. 178.
10. The easiest arrangement is to use the light from two flash
lights (or candles), one colored, for example, with a red
cellophane and the other left uncolored (this light should
be dimmer if possible). Set them apart so two clear shadows
can be formed. One will appear red and the other green. I
sometimes see colored shadows in the evening when the
red of the sunset passes through a window and mingles with
the white light of a room light. Two distinctly colored shad
ows appear.
11. Michael Wilson, manuscript report on “Color in Therapy”
presented to the Color Group of Great Britain, Imperial
College, February 3, 1971.
12. Goethe, quoted in Denis L. Sepper, Goethe Contra Newton
(Cambridge: Cambridge University Press, 1988), p. 28.
13. J. W. von Goethe, “Konfession des Verfassers,” Goethes
Werke, Hamburger Ausgabe, vol. 14, pp. 251-69; and Sep
per, Goethe Contra Newton, chap. 2.
14. Goethe, “Konfession des Verfassers,” p. 254.
15. G. W. F. Hegel, Philosophy ofNature, ed. and trans. M. J.
Petry (London: Allen, 1970), vol. 2, sec. 276, p. 17.
16. Goethe, Scientific Studies, p. 158.
17. Goethe, Scientific Studies, p. 158.
18. Goethe, Scientific Studies, p. 307. This excerpt only re
translated by Frederick Amrine.
19. Quoted by Benjamin DeMott, in Teaching What We Do
(Amherst, MA: Amherst College Press, 1991)
20. J. W. von Goethe, “Significant Help Given by an Ingenious
Turn of Phrase” in Scientific Studies, pp. 39ff.
21. Goethe, Scientific Studies, p. 164.
22. Goethe, Scientific Studies, p. 39.
361
Notes
23. J. W. von Goethe, “Empirical Observation and Science,”
a letter to Schiller, January 15, 1798 in Scientific Studies,
pp. 24-25.
24. Goethe, Scientific Studies, p. 158. This famous line has
been translated innumerable ways.
25. Goethe, Scientific Studies, p. 307.
26. Goethe, Scientific Studies, p. 21.
27. Goethe, in J. P. Eckermann, Conversations with Goethe,
February 18, 1829, p. 147.
28. Gaston Bachelard, The Flame ofa Candle, trans. Joni Cald
well (Dallas, TX: Dallas Institute of Humanities and Cul
ture, 1988), p. 20.
29. Lexikon der Goethe Zitate, ed. Richard Dobel (Zurich: Ar
temis Verlag, 1968), p. 524, no. 18.
30. Friedrich von Muller, in Lexikon der Goethe Zitate, p. 534,
no. 25.
31. Strakosch quoted in Sixten Ringbom, The Sounding Cosmos
(Abo, Fin.: Abo Akademi, 1970), p. 67.
32. Rudolf Steiner, Philosophy ofSpiritual Activity, trans. Wil
liam Lindeman (Hudson, NY: Anthroposophic Press,
1986).
33. Rudolf Steiner, The Course of My Life, trans. Olin Wan-
namaker (Hudson, NY: Anthroposophic Press, 1951),
p. 69.
34. Rudolf Steiner, Riddle of Man, trans. William Lindeman
(Spring Valley, NY: Mercury Press, 1990), p. 130.
35. Rudolf Steiner, Truth-Wrought Words, trans. Arvia
MacKaye Ege (Spring Valley, NY: Anthroposophic Press,
1979), p. 184.
36. Rudolf Steiner, The Gospel of John, trans. Maud Monges
(Hudson, NY: Anthroposophic Press, 1984), lecture of May
20, 1908.
37. Rudolf Steiner, “The Connection of the Natural with the
Moral-Physical. Living in Light and Weight,” in Colour
(London: Rudolf Steiner Press, 1935), pp. 87ff.
362
Notes
38. Rudolf Steiner, Truth-Wrought Words, p. 185.
39. Ralph Waldo Emerson, “Nature,” chap. 4, The Selected
Writings of Ralph Waldo Emerson, ed. Brooks Atkinson
(New York: Modem Library, 1968), pp. 15, 18.
40. Rudolf Steiner, Fruits of Anthroposophy (London: Rudolf
Steiner Press, 1986), pp. 48-49.
Chapter 9 I Quantum Theory by Candlelight
1. Michael Faraday, The Chemical History of a Candle deyr
York: Viking Press, 1960).
2. Gaston Bachelard, The Flame of a Candle, p. 20.
3. Paul Claudel, The Eye Listens, trans. Elsie Pell (New York:
Philosophical Library, 1950), p. 154.
4. Novalis, The Novices of Sais, trans. R. Manheim (New York:
C. Valentine, 1949), p. 95.
5. Thomas Kuhn, Black-Body Theory and the Quantum Dis
continuity 1894-1912 (New York: Oxford University Press.
1978). An excellent reference for much of the material
concerning black-body radiation, Max Planck, and the dis
covery of the quantum of action.
6. J. L. Heilbron, The Dilemmas ofan Upright.Wan (Berkeley.
CA: University of California Press, 1986).
7. A. Einstein, “On a Heuristic Point of View Concerning the
Production and Transformation of Light,” Collected Papers
ofAlbert Einstein, trans. Anna Beck (Princeton. NJ: Prince
ton University Press, 1989), vol. 2, p. 87.
8. Reference to his paper and M. J. Klein, “The First Phase
of the Bohr-Einstein Dialogue,” in Russell MeCormmach.
ed., Historical Studies in the Physical Sciences (Philadel
phia: University of Pennsylvania Press, 1970). vol. 2,
pp. 1-39.
9. A. Einstein, Physikalische Zeitschrift, vol. 10 (1‘X'WL
p. 817.
10. Klein, “Bohr-Einstein Dialogue,” p. 7.
363
Notes
11. Franz von Baader, “Uber den Blitz als Vater des Lichts,”
in Franz von Hoffmann, ed., Samtliche Werke (Aalen, Ger.:
Scientia, 1963), vol. II.
12. Antoine Faivre, “Tenebre, Eclair et Lumiere chez Franz
von Baader,” Lumiere et Cosmos (Paris: Albin Michel,
1981), p. 268.
13. Robert H. Eather, Majestic Lights (Washington, D.C.:
American Geophysical Union, 1980); and Syun-ichi Aka-
sofu, “The Aurora,” in Light from the Sky (San Francisco:
W. H. Freeman, 1980).
14. Werner Heisenberg, Encounters with Einstein (Princeton,
NJ: Princeton University Press, 1989), pp. 112ff.
15. Einstein, quoted in Abraham Pais, Subtle Is the Lord (New
York: Oxford University Press, 1982), pp. 416—17.
16. It is called the semi-classical or neo-classical theory of
quantum mechanics.
17. Ernst Blass, “The Old Cafe des Westens,” in The Era of
Expressionism, ed. Paul Raabe, trans. J. M. Ritchie (Dal
las, TX: Riverrun Press, 1980), p. 29.
18. Gustav Mie, inaugural lecture, January 26, 1925, Univer
sity of Freiburg in Breslau, quoted by Paul Forman, “Wei
mar Culture, Causality and Quantum Theory (1918—27)”
in Historical Studies in the Physical Sciences, ed. Russell
McCormmach (Philadelphia: University of Pennsylvania
Press, 1971), vol. 3, pp. 1-116.”
19. Wemer Haftmann, The Mind and Work of Paul Klee (Lon
don: Faber & Faber, 1954), pp. 57-58.
364
Notes
Chapter 10 I Of Relativity and the Beautiful
1. Albert Einstein, Ideas and Opinions, “On the Method of
Theoretical Physics” (New York: Crown Publishers, 1954),
pp. 270-76.
2. Quoted in Abraham Pais, Subtle Is the Lord, p. 131.
3. Henri Poincare, quoted in Jeremy Bernstein, Einstein (New
York: Viking, 1973), p. 103.
4. A. Einstein, in Paul Arthur Schlipp, ed., Albert Einstein:
Philosopher-Scientist (La Salle, IL: Open Court, 1949),
p. 25.
5. Einstein, “Ether and the Theory of Relativity,” a 1920 ad
dress in Sidelights on Relativity (New York: Dover, 1983).
6. See Arthur Miller, Albert Einstein’s Special Theory of Rel
ativity (Reading, MA: Addison-Wesley Publishing Co.,
1981), p. 166.
7. On the question of causality and relativity, see David Bohm,
The Special Theory of Relativity (New York: W. A. Ben
jamin, 1965), chap. 28.
8. There are special combinations of E and B that are invar
iant. See J. D. Jackson, Classical Electrodynamics (New
York: Wiley, 1975), pp. 517-19.
9. David C. Lindberg, “Medieval Latin Theories of the Speed
of Light,” in Roemer et la Vitesse de la Lumiere (Paris: Vrin,
1978), pp. 45-72.
10. Aristotle, De anima, II, 7;Desensu, VI. See Sabra, Theories
of Light, chap. 2.
11. Lindberg, Theories of Vision, p. 48.
12. Sabra, Theories of Light, p. 48.
13. M.-A. Tonnelat, “Vitesse de la Lumiere et Relativity,” in
Roemer et la Vitesse de la Lumiere, p. 282.
14. Albert Einstein, “On the Electrodynamics of Moving Bod
ies,” trans. Arthur Miller, in Albert Einstein’s Special Theory
of Relativity, p. 401.
365
Notes
15. James H. Smith, Introduction to Special Relativity (Cham
paign, IL: Stipes Publishing, 1965), chap. 2.
16. A. Brillet and J. Hall, “Improved Laser Test of the Isotropy
of Space,” Physical Review Letters, vol. 42(1979), pp. 549-
52. See also O’Hanian’s Classical Electrodynamics (Boston:
Allyn and Bacon, 1988), pp. 157—164.
17. D. Newman, G. W. Ford, A. Rich, and E. Sweetman,
“Precision Experimental Verification of Special Relativity,”
Physical Review Letters, vol. 40 (1978), pp. 1355-58.
18. See J. Terrel, Physical Review, vol. 116 (1959), pp. 1041ff.
19. Goethe, Scientific Studies, p. 263.
20. See John Lobell, Between Silence and Light: Spirit in the
Architecture of Louis I. Kahn (Boulder, CO: Shambhala,
1979); and Louis Kahn, Light Is the Theme, comments on
architecture complied by Nell E. Johnson (Fort Worth, TX:
Kimbell Art Foundation, 1975).
21. Pais, Subtle Is the Lord, chap. 9.
22. Einstein, “Ether and Relativity.”
23. Quoted in Emil Wolf, “Einstein’s Researches on the Nature
of Light,” Optics News, vol. 5, no. 1 (Winter 1979), pp. 24—
39.
24. Richard Feynman, Lectures on Physics (Reading, MA: Ad
dison-Wesley, 1968), vol. II, chap. 19, p. 9.
25. “Metaphysical Derivations of a Law of Refraction: Damianos
and Grosseteste,” Bruce S. Eastwood, Archivefor the History
ofExact Sciences, vol. 6(1970), pp. 224—36; and Journal of
the History ofIdeas, vol. 28(1967), pp. 403—14.
26. Plato, Timaeus, 46e.
27. For a sober and detailed treatment, see Wolfgang Yourgrau
and Stanley Mandelstam, Variational Principles in Dynam
ics and Quantum Theory, 3rd edition (London: Sir Isaac
Pitman & Sons, 1968).
28. Einstein, Ideas and Opinions, p. 228.
29. Friedrich Schiller, The Aesthetic Letters, trans. J. Weiss
(Boston: C. C. Little and J. Brown, 1845).
366
Notes
Chapter 11/ Least Light:
A Contemporary View
1. P. Grangier, G. Roger, and A. Aspect, “Experimental Evi
dence for a Photon Anticorrelation Effect on a Beamsplitter
A New Light on Single Photon Interferences,” in Euro
physics Letters, vol. 1 (January 1986). On the general de
velopment of the modem conception of the photon, see
Richard Kidd, James Ardini, and Anatol Anton, American
Journal of Physics, vol. 56 (1988), pp. 27-35.
2. Quoted in Arthur Fine, The Shaky Game (Chicago: Uni
versity of Chicago Press, 1986), p. 106.
3. N. Bohr, “Discussions with Einstein on Epistemological
Problems in Atomic Physics,” in Albert Einstein: Philoso
pher-Scientist, P. A. Schlipp, ed. (La Salle, IL: Open Court.
1949), pp. 200-41.
4. J. A. Wheeler, in The Ghost and the Atom, eds. P. C. W.
Davies and J. R. Brown (Cambridge: Cambridge University
Press, 1986), p. 64; and Mathematical Foundations of
Quantum Theory, ed. A. R. Marlow (New York: Academic
Press, 1978); and C. F. von Weizsacker, Zeitschrift fur
Physik, vol. 70 (1931), p. 114.
5. T. Hellmuth, H. Walther, A. Zajonc, and W. Schleich,
“Delayed-Choice Experiments in Quantum Interference.”
Physical Review, vol. 35 (1987), pp. 2532—41. See also
John Horgan, “Quantum Philosophy,” Scientific American
(July 1992), pp. 94-104.
6. Wheeler, The Ghost in the Atom, p. 69.
7. Fritz Rohrlich, From Paradox to Reality (Cambridge: Cam
bridge University Press, 1987), p. 22.
8. David Bohm, Wholeness and the Implicate Order (New York:
Ark Paperbacks, 1987).
9. J. S. Bell, Speakable and Unspeakable in Quantum Me
chanics (Cambridge: Cambridge University Press, 1987),
p. 27.
10. A. Aspect, et al., “Experimental realization of Einstein-
367
Notes
Podolsky-Rosen-Bohm Gedankenexperiment: A new niola-
tion of Bell’s inequalities,” Physical Review Letters, vol. 49
(1982), pp. 91—94; and “Experimental test of Bell’s ine
qualities using time-varying analyzers,” pp. 1804—7.
11. These are the so-called quantum beat experiments.
12. Erwin Schrodinger, What Is Life? and Other Scientific Es
says (Garden City, NY: Doubleday, 1956), pp. 161—62.
13. U. Bonse and H. Rauch, eds., Neutron Interferometer (New
York: Oxford University Press, 1979).
14. See “Making Waves with Interfering Atoms,” Science, vol.
252 (May 17, 1991), pp. 921-22.
15. See articles by C. Tesche, S. Washbum, and R. Webb,
for example, in New Techniques and Ideas in Quantum
Measurement Theory, ed. Daniel Greenberger, vol. 480
(1986), Annals of the New York Academy of Sciences, pt.
II.
16. D. Bohm, B. J. Hiley, and P. N. Kaloyerou, “An Onto
logical Basis for the Quantum Theory,” Physics Reports,
vol. 144, no. 6 (1987), pp. 321-75.
17. Francis Thompson, quoted in Alan J. Friedman and Carol
C. Dorley, Einstein as Myth and Muse (Cambridge: Cam
bridge University Press, 1985), p. 44.
18. Hesiod, Theogony.
19. Aristotle, Physics, trans. Richard Hope (Lincoln, NE: Uni
versity of Nebraska Press, 1961), p. 58.
20. T. D. Newton and E. P. Wigner, Reviews of Modern Phys
ics, vol. 21 (1949), pp. 400—06; see also E. R. Pike and
Sarben Sarkar, “Photons and Interference,” in E. R. Pike
and Sarben Sarkar, eds., Frontiers in Quantum Optics (Bos
ton: Adam Hilger, 1986), pp. 282—317.
21. Murray Sargent III, Marian Scully, and Willis E. Lamb,
Jr., Laser Physics (Reading, MA: Addison-Wesley Pub
lishing Co., 1974), p. 228.
22. J. A. Valdmanis and N. H. Abramson, “Holographic Im
aging Captures Light in Flight,” Laser Focus World, Feb
ruary 1991, pp. 111—17.
368
Notes
23. James Turrell, in Occluded Front, ed. Julia Brown (Los
Angeles: The Lapis Press, 1985), p. 46.
24. Gerard de Nerval, Le Christ aux Oliviers in Dictionary of
Foreign Quotations, compiled by Robert and Mary Collison
(New York: Facts on File, 1980), p. 138.
25. H. G. B. Casimir and D. Polder, “The Influence of Retar
dation on the London-van der Waals Forces,” Physical Re
view, vol. 73 (1948), p. 360.
26. H. B. G. Casimir, “On the attraction between two perfectly
conducting plates,” Proceeding of the Koninklijke Neder-
landse Akademie van Wetenschappen, vol. 51 (1948),
p. 793.
27. See I. J. R. Aitchison, “Nothing’s plenty: The vacuum in
modem quantum field theory,” Contemporary Physics, vol.
26 (1985).
28. Victor Hugo, Derniere Gerbe, November 26, 1876.
29. Herder, speaking of life, quoted in Bachelard, The Flame
of a Candle, p. 2.
Chapter 12 I Seeing Light
1. See Margaret Miles, “Vision,” The Journal ofReligion, vol.
63 (1984), pp. 125-42.
2. Dante Alighieri, The Paradiso, trans. John Ciardi (New
York: New American Library, 1970), canto 31, lines 22ff.
3. St. Bernard (or an imitator), quoted in Millard Miess, “Light
as Form and Symbol in Some Fifteenth-Century Paintings,”
Art Bulletin, vol. 27 (1945), pp. 175-81.
4. Otto von Simson, The Gothic Cathedral (Princeton, NJ:
Princeton University Press, 1988), 3rd ed., p. 55; and all
of chap. 2.
5. Ernest Gellner, Plough, Sword and Book: The Structure of
Human History (Chicago: The University of Chicago Press,
1988), chap. 2.
369
Notes
6. Hesiod, Works and Days, pp. 1 lOff.
7. Vico, quoted in Charles M. Radding, A World Made by
Men: Cognition and Society, 400-1200 (Chapel Hill, NC:
University of North Carolina Press, 1985).
8. Louis L. Noble, The Course ofEmpire (New York: Lamport,
Blakeman & Law, 1853), p. 289.
9. See plate 5 of Elizabeth Sears, The Ages of Man: Medieval
Interpretations of the Life Cycle (Princeton, NJ: Princeton
University Press, 1986).
10. Novalis, The Disciples at Sais and Other Fragments, trans.
F.V.M.T. and U.C.B. (London: Methuen and Co., 1903),
pp. 137.
11. Paul Cezanne Letters, ed. John Rewald (New York: Hacker
Art Books, 1976), p. 327, to his son, September 8, 1906.
12. Cezanne, Letters, pp. 303 and 306, July 25, 1904.
13. Buddhaghosa, Path of Purity, trans. Pe Maung Tin (Pali
Text Society, 1975), p. 145.
14. Buddhaghosa, Path of Purity, p. 200.
15. J. W. Goethe, Goethes Gesprache (Leipzig, Ger.: Bieder-
mann, 1901-11), vol. 2, p. 245.
16. Goethe, Goethes Werke, Hamburger Ausgabe, vol. 13,
p. 323.
17. Goethe, Goethes Werke, Hamburger Ausgabe, vol. 13,
p. 38.
370
Acknowledgments
The writing of this book would not have been possible with
out the help of many friends and teachers. My interest in
Goethe, Steiner, and the humanistic and spiritual dimensions
to light were first piqued by Professors Ernst Katz and Alan
Cottrell, two early mentors, to whom I owe a great debt. My
study of light from the vantage point of quantum physics was
stimulated by research visits to the Ecole Normale Superiure
with Marie Anne Bouchiat, the Max Planck Institute for Quan
tum Optics with Herbert Walther and Marian Scully, the Uni
versity of Hanover with Jurgen Mlynek, and the University of
Rochester with Leonard Mandel. Among many local colleagues
whose conversations provoked and clarified my thinking I would
like to give special mention to George Greenstein, Herbert Bern
stein, K. Jaganathan, Larry Hunter, Dudley Towne, and Bob
Krotkov. For literary and artistic themes, I thank Frederick
Amrine, Douglas Patey, Douglas Miller, Christopher Bamford,
and Joel Upton for their interest. I would like to thank William
Irwin Thompson for including me in Lindisfarne gatherings,
which provided the occasion for developing certain threads of
the book, and also his son Evan Thompson for pushing me to
think ever more deeply about vision.
In addition to the many colleagues important to the creation
of this book, I would like to thank Amherst College, especially
its library staff, and Laurence Rockefeller for support during a
371
A cknowledgments
sabbatical leave. Without such support, writing would have been
even more protracted than it already was. Of course, without
the encouragement of my family, and the excitement of my sons,
August and Tristan, the project would have been impossible.
Finally, my thanks to Leslie Meredith for her early enthusiasm
and careful editing, and most especially to Patricia van der Leun
whose constant support, good judgment, and unflagging attention
to the book helped shape it from inception through publication.
To all of these goes much of the credit for what is good in
Catching the Light. I reserve whatever errors and infelicities it
contains to myself.
372
Index
Afterimages, 29-30, 194-95, 196 in electromagnetism, 133, 257
Ages of man, 332-33 light quanta and, 292-93
Alan of Lille, 140 Architecture
Alexander’s dark band, 169-70, concept of space and, 272-74
175, 176-77 Kahn’s use of light in, 274, 275
Alhazen, 28, 29-32, 264 of space-time, 276-80
Ambiguous figures, 22-23 Aristophanes, 165
Anagni, Cathedral of, 337 Aristotle, 74, 75
Analysis. See also Mathematics; aurora borealis and, 241
Spectroscopy. final cause and, 286
of light, 83-85, 155-56, 289-90 rainbow and, 167, 168, 172
of physical phenomena, 81-83 speed of light and, 263, 270
Ancient civilizations. See also understanding of light, 77-78
Human cultures; specific Art
cultures. application of color in, 200
nature of light in, 39-43 artistic dimensions of light and, 8
views of rainbow, 161-65 cultivation of sight and, 339
Angelic light, 44-46, 54-55, 57 cultural relativity of perception
Steiner’s spiritual hierarchies and, 62-64
and, 220-21 perspective drawing in, 59-60
Animals, vision research on, 5 spiritual geometry in, 61-62
Anni mirabiles, 81 at tum of the century, 249-50
Anticorrelation, 295-96, 299 Aspect, Alain, 294, 299, 311, 312
Aperqu (insight), 342-43. See also Atoms. See also Corpuscular theory
Insight. of light; Material view of light.
Arc-en-ciel, 171 as forces, 134-35
Archetypal phenomena (“l/r- quantum mechanics and, 244—46
phenomena”) “vortex," 152-53
color theory and, 208-10 Augustine, Saint, 49, 263, 330
373
Index
Aureole (corona), 116 Bohr, Niels, 238, 326-27
Aurora borealis, 240—43 quantum theory and, 244—49,
cross-cultural lore around, 241, 315-16
243 view of light, 247-49
phenomenon of, 240-41 Boundary colors, 207—11
physical view of, 243-44 Box of light
Australian aborigines, 162-63 exhibit, 2—3
Herschel’s, 119, 234
Baader, Franz von, 239 Boyle, Robert, 103
Babylonian astronomer-priests, Brain, and models of vision, 34—36
111, 301 “Bremen Town Musicians” tale, 77
Bachelard, Gaston, 227 Brillet, Alain, 268
Bacon, Francis, 7 Brougham, Henry, 110
Bacon, Roger, 264 Browne, Thomas, 144
Bader, Mr., 281, 285 Brunelleschi, Filippo, 58-61, 71
Baillet, A., 90 Buddhaghosa, 340
Barfield, Owen, 35, 182-83 Buddhist philosophy
Beamsplitter, 294-96, 299 elimination of attributes in, 134-
Beauty. See also Wonder. 35
love of, and scientific knowledge, meditation in, 339—40
157-60, 179-81, 214-15 Bunsen, R. W., 155
path of, and light, 281, 285, “Butterfly effect,” 320
287-89 Biittner, Hofrat, 192, 194, 201
Bell, John, 307, 308, 316
Calculus, 82, 111, 113
Bernard, Saint, 330-31
Camera obscura, 30—32
Bernard of Chartres, 79
Campbell, Thomas, 180
Bhagavad-Gita, 11—12 Candlelight, 225—26
Bible, 18, 44, 154 color of, 228, 230, 231, 239
Black-body radiation, 154, 156, physical phenomenon of, 227—28
230-31 Planck’s theory of light and,
Einstein’s particle theory and, 232-34
235-38 “The Case of the Colorblind
Planck’s theory of light and, Painter” (Sacks and
235-38 Wasserman), 15-16
spectral emission lines and, 240 Casimir, H.B.G., 326-27
Blake, William, 340 Cassirer, Emst, 61
Blindness. See also Color Cathars, 50—51, 56
blindness. Cathedral of Anagni, 337
interior light and, 16-17 Catherine II of Russia, 113
poetic imagination and, 11—12 Causality, 76
sight following recovery from, 1— Compton effect and, 248—49
2, 3-6 Lorentz transformation and, 261
Boethius, 139-40 necessary and divine types of,
Bohm, David, 304-05, 317-19 285-86, 289
374
Index
Cizanne, Paul, 338, 339 Consciousness. See also Human
Chartres cathedral, 100-02 cultures; Mind.
Chladni figures, 129, 130 ecological, 320
Chloros, 15 evolution of, 181-87, 270-76
Christianity, 26, 48-52. See also mutability of, 13, 36-37
Judeo-Christian tradition. reality of, 182-83
Chromatic adaptation, 198. See also Copernicus, 261
“Law of required change." Corpuscular theory of light. See
Claudel, Paul, 227-28 also Ether; Material view of
Clemens, Samuel L. (Mark Twain), light.
178-79 Faraday’s attack on, 134-35
Cold light, 78 Goethe’s color studies and, 201
Cole, Thomas, The Voyage of Life, Newton and, 83-85, 87, 92-94,
334-36, 337 99, 107, 116, 177
Collective representations, 183 Cusa, Nicolaus of, 69-71, 77
Color blindness, 15—16, 195
Colored shadows, 196-98 Damianos, 285
Colors of light. See also Prism; Dark light, 325-29
Spectroscopy, Darkness and light
candlelight and, 228-31 color production and, 207-11
colors of the rainbow and, 167— in Judeo-Christian tradition, 44-46
68, 172 in Manichaeism, 47-52
in Zoroastrianism, 41-43, 46
Michelson-Morley experiment
Darwin, Charles, 168-69, 333
and, 154, 156
Davy, Humphrey, 127, 128
sensory-moral effects of, 213-14, Deconstruction, 185
215
Delaunay, Robert, 250
solar spectrum and, 155-56 Delayed-choice experiment, 299-
Color temperature, 229 301
Color theory Descartes, Reni, 32, 87, 264
darkness and light and, 207-11 concept of space, 91-92, 98,
Goethe and, 191, 192-95, 200- 135
03, 207-12, 208, 214-15 dream of, 88-91
Land and, 190-91 nature of light and, 91-94
Color therapy, 199-200, 213, 215 rainbow and, 171, 175-77
Color vision “Devil’s rainbow,” 170
“The Case of the Colorblind Diderot, Denis, 113
Painter” and, 15-16 Diffraction phenomena
Homeric Greece and, 13-15, 16 contest on explanation of, 114—
pathological instances of, 195 15
retinal sensitivity and, 109-10 in everyday life, 116
Commonsense theories, vs. wave theory and, 107-09, 110,
quantum theory, 310-11 114-19
Compton, A. H., 248—49 Diogenes Laertius, 18
375
Index
The Discarded Image (Lewis), 38 Maxwell’s electromagnetic field
Dispersion, 177-78 and, 150
Dream of Descartes, 88-91 in Planck’s theory of light, 233-
Duccio di Buoninsegna, 61 34
Diirer, Albrecht, 65-67 in vacuum, 327-28
Entanglement, 313, 316—17
Eddington, Arthur, 276 Epimetheus, myth of, 10-11
Edison, Thomas Alva, 225-26 EPR experiment, 309—12, 313—14,
Egyptian mythology, 39, 40, 41 323
Eidola, 29 Equivalence principle, 276-79
Einstein, Albert, 145, 253. See Eskimo lore, 241
also Relativity theory. Ether
electromagnetism and, 133, 256— decline of material view of, 119—
61 23, 146, 149, 254-55
particle theory of light and, 235— experimental attempts to detect,
37 154, 268
principle of equivalence, 276-79 immaterial view of, 122—23,
principle of relativity, 256-57 256, 278-79, 305
running with light paradox and, material view of, 91—92, 98,
253-54 110, 134, 146, 149, 151-54
thought experiments on quantum (See also Wave theory.)
mechanics, 296-99, 307, theory of relativity and, 256—62
308-10, 311-12 Euclid, 65, 272-73
views on quantum theory, 235- Optics, 24—26
37, 246-47, 279 Euler, Leonhard, 99-100, 107,
Electrical field, 257-58 121
Electromagnetic field mathematical description of
Bohr and, 247-48 nature and, 111—14
discovery of electromagnetic Euripides, 15
induction and, 130-33 Eye. See also Retina.
Einstein’s “principle of as source of light, 77-78
relativity” and, 257-61
Faraday’s speculations on, 136- Fall of Man, 44-45
39 FAPP (“for all practical purposes”)
Maxwell’s theory of light and, physics, 315—16, 317, 319
146-50 Faraday, Michael, 125—26, 133,
Elements, spectral lines of, 155— 157-58
56, 239, 240, 245 electromagnetic induction and,
Elephant, split-type drawing of, 130-33, 257
63-64 introduction of forces by, 134-35
Emerson, Ralph Waldo, 221, 271 lectures for children, 226-27,
Empedocles, 17, 19-21, 198, 205 228
Energy. See also Relativity theory. Maxwell’s mathematization of
conservation of, 248, 249 work of, 146—50
376
Index
theory of light, 133—39 Goethe, Johann Wolfgang von, 14,
“Thoughts on Ray-Vibrations,” 184, 185-86, 196, 289, 299,
136-38 340-41
Faust (Goethe), 140-42, 186-87, color studies by, 191, 192—95,
217 200-03, 207-12, 208, 214-15
Fermat, Pierre de, 287 Faust, 140-42, 186-87, 217
Feynman, Richard, 147, 280—81, nature of light and, 202—03
287-89 participatory understanding of
Field theory, 125, 136, 149-50, science and, 203—06, 210—11,
259-60. See also 212
Electromagnetic field. Theory of Color, 21—22, 194—95,
Finnish folklore, 11, 102, 241, 243 302
Fire, 42-43. See also Candlelight; Gothic cathedral, 330—31
Heat, light from. Gral, Ursula, 199—200
Fontenelle, Bernard de, 95—96 Grangier, Philippe, 294, 299
Frame of reference, 256—57, 258— Gravitation
60 Newton’s theory of, 81—83
Franklin, Benjamin, 147—48 theory of relativity and, 276—79
Fraunhofer, Joseph, 155 Greece, ancient. See also Homer;
Fresnel, Augustin, 114—18, 119, Plato.
130, 134 accounts of vision, 28—29
causality and, 76
Galen, 28 color vision in, 13-17
Galileo Galilei, 73-76, 78-79, elektron effect and, 147
104, 155 linear perspective and, 65
Gellner, Emest, 331—34 myth of gift of fire in, 10—11
Geometry. See also Linear natural science in, 18—21
perspective; Mathematics; nature of light in, 39—40
Measurement. tale of Empedocles in, 18—21
non-Euclidean, 272—74 view of the rainbow in, 163—64
perfection and, 75 Gregory, R. L., 3—4
physical and spiritual worlds Grimaldi, Francesco Maria, 107
and, 82 Grosseteste, Robert, 52—57, 172—
of the rainbow, 177 73, 220
Ghiberti, Lorenzo, 71—72 Guericke, Otto von, 196
Gibbio, Italy, 161-62
Gibbs, Willard, 153 Hall, Jan, 268
Glare, 116-18 Hallucinations, 17
Gnosticism. See Manichaeism. Hanson, J. D., 323
God. See also Physical and spiritual Haydon, Benjamin Robert, 180
worlds; Religious feeling, and Heat, light from, 228—31
scientific thought. Hegel, Georg, 202
concept of infinity and, 69—71 Heisenberg, Werner, 246, 260
light as sight of, 39—41, 56 Heracles, 46
377
Index
Heraclitus, 338 Individuals, developmental stages
Hero of Alexandria, 286, 287 of, 334—38. See also Mind.
Herschel, William, 130 Infinity
Hertz, Heinrich, 149 concept of, 69-71
Hesiod, 321, 333 in quantum electrodynamics,
Hiley, Basil, 304-05 327-28
Hindu philosophy, 336 speed of light and, 269
Hitler, Adolf, 251-52 Infrared radiation, 229
Holography, 107, 322-23 Insight
Homer, 11, 13-17, 163, 164 cultivation of, 204—06, 338-42
Hopkins, Gerard Manley, 165-66, moment of, 212, 342-43
170, 225, 226 Instrumentalism, 301-02, 303
Horus, 39, 40 Interference
Hubei, David, 35 rainbows and, 178
Hughes, John, 86 single-photon studies and, 296-
Hugo, Victor, 325-26, 328 301
Human cultures, 17, 329. See also Young’s principle of, 110, 115,
Ancient civilizations; 178
Consciousness; Mind; specific Interior light. See also Physical and
cultures. spiritual worlds.
evolutionary development of, Greek color vision and, 13—17
249-50 historic accounts of vision and,
meanings of light for, 8-9 20-22, 24-26, 28-30
mythologies of human history in, poetic vision and, 11—12
332-33 postmodern view of vision and,
pictorial perception in, 62-64 36-37
Husks, theory of, 29—31 Iris (goddess the rainbow), 163,
Huxley, Thomas, 34 164-65, 167-68, 171
Huygens, Christian, 86 Islam, 27
Ibn al-Haytham. See Alhazen. Jeans, James, 233
Ibn Ishaq, Hunayn, 27 Judeo-Christian tradition, 44—46.
Icelandic mythology, 165, 168 See also Christianity.
Ideas, catching sight of, 206, 207- Jundishapur, Academy of, 27
14. See also Insight. Justinian I, 26—27
Idolatry, and scientific models, 35,
36, 302, 306 Kahn, Louis, 274-76, 290, 291
Imagination Kalevala (Finnish folk epic), 11,
experimentation and, 194, 195 102
Greek color vision in, 13—15 Kandinsky, Vasily, 216—18
light of, 12-13 Keats, John, 179, 180, 181
poetic, 11—12, 206-07 Kelvin, Lord. See Thomson,
Implicate-order physics, 319 William.
Incandescence, 154 Kepler, Johannes, 31, 32
378
Index
Khurso I (emperor of Persia), 27 quantum phenomena and, 306—
al-Kindi (Islamic philosopher), 55— 07
56 as sight, 39-41
Kircher, Athanasius, 102-05 as source of being, 275
Kirchhoff, Gustav, 155 as source of sight, 62, 184, 205,
Klee, Paul, 250 316, 340-41
Knowledge. See also Rationality; Lightning, 239, 240
Science. Light quanta. See Photon.
causality and, 76 Linear perspective
cultivation of insight and, 338— Brunelleschi and, 58—61, 71
42 cultural influences and, 64-69
as event, 342-43 Ghiberti and, 71—72
future, 36, 338 pre-Renaissance, 65
gnosticism and, 50 spiritual geometry and, 61-62
Liquid crystal display (LCD), 78
Kokushi, Muso, 158—59
Living philosophy
Koyre, Alexander, 53
(Lebensphilosophie), 249-50,
Kuhn, Thomas, 70 272
Kyanos, 14-15
“Local, realistic” theories, vs.
quantum theory, 310-11
Lamb, Willis, 322 Locality condition, 310-12, 313.
Land, Edwin, 190-91, 195 See also Location, of light;
Lao-tzu, 1, 97, 182 Nonlocality, concept of.
Laser, 322 Location
“Law of required change,” 196, of center of rainbow, 169
197-98 of light, 321-25
Least action, principle of, 287-89 of perfection, 74-76
Lebensphilosophie, 249—50, 272 of rainbow, 167
Leconte, John, 105 Lorentz, Hendrik Antoon, 255,
Leonardo da Vinci, 31 256
LePrince, Dr., 1-2 Lorentz transformation, 261
Lewis, C. S., 270 Lorenz, Edward, 320
Lewis, G. N., 293 Lucifer, 44—45
LIF (“light-in-flight”) holography, Lux vs. lumen, 97-98, 106
322-23
Light Maeterlinck, Maurice, 106
Magnetic field, 257-58
attributes of, 314-15, 323, 324— Manichaeism, 47-52
25 Materialistic realism, 185
bending of, 276, 277, 278, 281- Material view of light. See also
82, 287 (See also Refraction.) Corpuscular theory of light;
historical shift in accounts of, Ether. \
23-26 Cartesian vs. Newtonian
physical and conscious levels of, viewpoints and, 87, 92—94
330-31 Galileo and, 76, 78-79
379
Index
Grosseteste and, 53-56, 95 Mind. See also Consciousness,
Newton and, 83-85, 87, 92-94 ambiguous figures and, 22-23
Mathematics. See also Geometry. concept of nonlocality and, 319-
applied to description of nature, 20
111-14 cultivation of capacity for sight
development of calculus and, 82 and, 4-6, 338-43
Grosseteste’s cosmogony of light Descartes’s account of vision
and, 53—55 and, 31—32
Nicolaus of Cusa and, 69-71 Greek accounts of vision and,
perfection and, 75, 284 13-17, 28-29
Mathematization, 25-26 as illusion, 35—36
of Faraday's work, 146-50 inversion of retinal image and,
Matthew (apostle), 36-37 31-32
Maupertius, Pierre Louis Moreau invisible light and, 324
de, 287 mutability of consciousness and,
Maxwell, James Clerk 13, 36-37
“A Dynamical Theory of the organs of insight and, 204-06
Electromagnetic Field,” 147— participation in sight, 12-13,
49 183
religious feeling and scientific Platonic sight and, 21-23
thought in, 142-45 psychogenesis and, 334-37
treatment of light, 146-50, 259 role in history, 333—34, 337—38
Measurement. See also Geometry. theory of relativity and, 270
Grosseteste’s cosmogony of light Miracle of Our Lady of Fatima,
and, 53-55 241, 243
Moon, and diffraction effects, 116
quantum phenomena and, 313-
Moreau, Dr., 1-2, 5-6
14, 317-18
Morley, Edward Williams, 154
Mechanical philosophy, Cartesian,
Motion, and Einstein’s principle of
88, 91-94
relativity, 256-61. See also
Meditation, 339-40
Speed of light.
Mersenne, Marin, 88 Miinter, Gabriele, 216
Metaphysical assumptions, 284-85, Music, 100-02, 129. See also
287 Sound.
Meter, standard for, 265-66
Michelangelo, 67 Nature, as light of experience, 69-
Michell, John, 93 71
Michelson, Albert Abraham, 154, Nature of light
265 analysis and, 289-90
Michelson-Morley experiment, 154, in ancient civilizations, 39-43
156 contemporary view of, 292
Mill, John Stuart, 158, 184 Grosseteste’s cosmogony of light
Millikan, Robert A., 237 and, 52-57
Milton, John, 10, 58 location and, 321—23
380
Index
mystery of, 7-8 Perspective drawing, 59-60, 66
“Nature philosophers,” 202 Philosophy
Navajo mythology, 163 specters of, 139—42
Nerval, Gerard de, 326 at turn of the century, 249-50
Newton, Isaac, 79-87 union of idea and experience in,
color and, 191 211-14
contemporary response to, 85— Photon. See also Quantum
87, 93 phenomena; Single-photon
corpuscular theory of light of, studies.
83-85, 87, 92-94, 99, 107, archetypal experiments and,
116, 177 292-93
gravitational theory of, 81-83 location of light and, 321-25
Optiks, 83-85, 86 name of, 293
rainbows and, 177-78 properties of, 260, 279-80
Newton, T. D., 321-22
sources for, 294—96
Nieuwenlijdt, Bernard, 93
Physical and spiritual worlds. See
Nihilism, 185, 302
also Religious feeling, and
Nishitani, Keiji, 158-59
scientific thought.
Nonlocality, concept of, 218, 305,
Galileo and, 73—76
317, 319—20. See also Locality in Grosseteste’s cosmogony of
condition.
light, 53-56
Novalis, 12, 196, 224, 228, 338
neoplatonic view of geometry
Optical illusion and, 82
Nicholas of Cusa and, 69—71
ambiguous figures and, 22-23
perfection in, 74—76, 95
color science and, 194—95
Optics (Alhazen), 28, 29-32 rainbows and, 179—81
Optics (Euclid), 24-26 sound and, 100—05
Optiks (Newton), 83-85, 86 Steiner’s view of science and,
Ovid, 334 221-22
traditional unity of, 331—33
Painter’s Manual (Diirer), 65-67 Western segregation of, 332, 333
Panofsky, Erwin, 65 Physics
Parzival, story of, 271-72 FAPP (“for all practical
Path of light. See also Refraction, purposes”), 315—16, 317, 319
perfection and, 286-87 Newtonian vs. Cartesian, 87,
photons and, 303-04 92-93
through space-time, 278 turn of the century revolution in,
Pecham, John, 263 249-50
Pedantry, 143 Piaget, Jean, 341
Perfection. See also Beauty, Pictorial perception, 62-64
location of, 74—76 Pindar, 15 I
mathematics and, 75, 284 Planck, Max, 222, 231—35, 252,
Persia, ancient, 41-43 288
381
Index
Planck Institute for Quantum location and, 323
Optics, 299-300 many-particle experiments and,
Planck’s constant, 233 317
Plato, 10, 12, 264, 289 Quantum potential, 305, 318-19
account of vision, 21—22 Quantum realism, 313, 317, 319
Timaeus, 52—53, 77 Quantum theory, 7. See also
Platonic Academy, destruction of, Quantum phenomena.
26-27 Bohr’s application of, 244—49
Plenum, 92, 135 candlelight and, 225-26
Plutarch, 241 colors of light and, 154
Podolsky, B., 307, 309 Compton effect and, 248-49
Poetic imagination, 11-12, 206- Einstein’s early advocacy of,
07, 326 235-38
Poincare, Henri, 254-55 Einstein’s resistance to, 246-47,
Poisson, Simeon Denis, 115—16 279
Polanyi, Michael, 184 “path integral formulation” in,
Polarization, 116-18, 311, 312, 288-89
313, 322 place of light and, 321—22
Polarization colors, 118 Planck and, 231-35, 237-38,
Poseidon, bronze statue of, 17 245-46
Primary qualities, 315 vs. “local, realistic” theories,
Prism 310-11
edge colors and, 207-11
Goethe and, 192—93, 201 Radiation, invisibility of, 229-30
Newton and, 83-85 Rainbow
“Project Eureka,” 2-3 ancient views of, 161—65
Projective geometry, 272-74,275-76 as bridge, 163, 165
Prometheus, 46 center of, 169
Ptolemy, Claudius, 282-84 colors of, 167—68, 172
Pythagorus, 101, 102, 104 evolution of theory of, 171-79
Kandinsky’s painting of, 217-18
Quantum electrodynamics (QED), location of, 167
327-28 phenomenon of, 166—71
Quantum phenomena. See also physical vs. spiritual views of,
Quantum theory. 179-81
Bell’s theorems and, 308 supernumerary arcs in, 170
critical realist perspectives on, Rainbow angle, 169, 175—76, 177
303-05 “Rainbow serpent,” 162—63
Einstein’s thought experiments Rationality
on, 296-98, 298-99, 307, Einstein’s principle of relativity
308-10 and, 260—61
EPR experiment and, 309—12, enhancement of meaning of,
313-14, 323 310-12
insights into light from, 306-07 Lebensphilosopie and, 272
382
Index
Redemption, 46, 49 Sabra, A. I., 92
Reflection Sacks, Oliver, 15-16, 195, 199
cats’ eyes and, 77 Sargent, Murray, 322
of lasers, 322 Saving the Appearances (Barfield),
rainbow and, 174 35, 182-83
Refraction Schiller, Johann Christoph
angle of, 282-85, 287 Friedrich von, 290
law of, 175, 287 Schopenhauer, Arthur, 340
light as body and, 84 Schrodinger, Erwin, 316
Ptolemy’s data on, 282-84 Science. See also Geometry;
rainbow and, 173, 174 Knowledge; Mathematics;
Relativism, 185 Philosophy; Physics.
Relativity theory, 254 emergence of mathematical and
“equivalence principle” of, 276- material thought in, 74—76,
79 78-79, 95-96
evolution of human Goethe’s participatory
consciousness and, 270-76 understanding of, 203-06,
experimental confirmation of, 210-11, 212
268-70 Grosseteste and, 53-55
general theory of relativity, 276 love of beauty and, 157-60
gravitation and, 276-79 models in, as idols, 35-36
special theory of relativity, 154, modem, and meaning, 301-03
268-74 poetic imagination and, 206-07
Religion. See God; Physical and role of religious inclinations in
spiritual worlds; Religious conduct of, 123, 126, 143-45
feeling, and scientific thought; separation of the physical and
specific religious traditions. spiritual and, 95, 221-22,
Religious feeling, and scientific 333
speculative theories in, 134
thought, 123, 126, 142-45
Res extensalres cogitans, 34 Steiner’s spiritual view of, 218-19
Retina subjective experience and, 25-26
Scully, Marian, 322
color vision and, 109-10
Self-consciousness, 338, 340-41
inversion of image on, 31—32 Senden, M. von, 4—5
Riebau, Mr., 127
Shelley, Percy Bysshe, 157, 168
Roemer, Ole, 264—65
Sight
Roger, G., 294, 299 active dimensions of, 198-99
Rosen, N., 307, 309 Alhazen’s account of, 29-30
Rosenkreutz, Christian, 89 cognition and, 204-06
Rosicrucianism, 89 cultivation of capacities for,
Russell, Morgan, “Synchromy” 338-43
paintings, 118 Descartes’ view of, 92
Rutherford, Ernest, 245 development of cognitive
Rydberg, J. R-, 245 capacity for, 4-6
383
Index
Greek accounts of, 28—29 measurements of, 262—65
historical shift in accounts of, as measurement standard, 265-
23-26 66
light as, 39-41 theory of relativity and, 256,
light as source of, 62, 184, 205, 262, 265
316, 340-41 Spirit light. See Angelic light.
recovery from blindness and, 1— Spiritual perspectives on light. See
2, 3-6 also Ancient civilizations;
source of, in Empedocles, 20 Physical and spiritual worlds;
speed of, 262-63 specific religious traditions.
Simulacra, 29 Christian Fall of Man and, 44-46
Single-photon studies Descartes and; 34
attitudes among scientists on, Grosseteste’s cosmogony of light
303-07 and, 53-56
interference fringes, 296-301 Manichaeism and, 47—52
single-photon ambiguity and, Steiner and, 219—24
299-301 Zoroastrianism and, 41—43
sources of light for, 294—96 Stages of life, 334—38
Solar sponge (spongia solis), 78, Stained-glass windows, 330—31
105 Steiner, Rudolf, 186, 216—18
Sommerfeld, Arnold, 248 account of light, 219—24
Sound Stokes, George, 121
diffraction effects and, 109 Strakosch, Alexander, 216, 217
light as analogous to, 130 Strakosch, Maria, 216, 217
meaning and, 106-07 Sunfield Children’s Home, 189-90,
physical and spiritual views of, 199-200, 213, 215
Sunlight, 20-21, 29. See also
100-05
Prism.
polarization phenomena and, 117
Sunspots, 75, 155, 243—44
Space, concept of, 97-100
Supernumerary arcs, 170
Descartes and, 91-92, 98, 135
Superposition
geometry and, 272-76 quantum phenomena and, 303-
Lord Kelvin and, 151-53 04, 317
theory of relativity and, 270, 271 Young’s principle of, 110, 114
Space-time, architecture of, 274-80
Spectral emission lines, 239, 240, Telescope, 74—75, 80
245 Thaumas (Greek god of wonder),
Spectroscopy, 154, 156, 230-31, 164-65
245 Theodoric of Freiberg, 173-74
Speed of light Theodoric (sixth century), 139—40
effect of travel at, 269-70 Theophrastus, 14
as independent of source motion, Thermodynamics, laws of, 232-33.
267-70 See also Heat.
as infinite, 263-64 Thompson, Francis, 319
384
Index
Thomson, J. J., 231, 245 234, 260, 298-301, 302-03,
Thomson, James, 178 317
Thomson, William (Lord Kelvin), “Wave-particle-ness,” 302-03
151-54 Wave theory
“Two Clouds” address, 154 color and, 190
Thoreau, Henry David, 158 Compton effect and, 248-49
Time, concept of, 270, 271. See Descartes’ views and, 94
also Space-time. diffraction phenomena and, 107—
Tonnelat, Marie-Antoinette, 265 09, 110, 114-19
Turrell, James, 324 divisibility of light and, 295-96
Twain, Mark, 178-79 Einstein’s particle theory and,
Tyndall, John, 157 235-38
electricity and, 130, 132
Ultraviolet radiation, 229 Fresnel’s work and, 115-17, 130
Unity Huygens and, 86, 94
of electromagnetic and light mathematical foundations of,
phenomena, 130, 147-48 115-16, 130
of light, 224 ocean waves and, 94-95,104-05
traditional view of physical and sound and, 104-05, 117
spiritual worlds and, 331-33 speed of wave propagation and,
Ur-phenomena. See Archetypal 266-67
phenomena. Young’s work and, 110, 114,
115
Weizacker, Carl Frederick von,
Vacuum, 102-05, 327-28
299
Van Gronigen, C., 49-50
Westfall, Richard S., 82
Vibration. See Electromagnetic
Wheatstone, Charles, 135, 136
induction; Sound; Wave Wheeler, John Archibald, 299, 304
theory.
Whicher, Olive, 273, 274
Virgil, 167
Wigner, Eugene, 321-22
Vision. See Sight.
Wilbur, Richard, 275
Vision research, 5. See also Retina. William of Conches, 28, 264
Visions. See Dream of Descartes; Wilson, Michael, 188-90, 191,
Hallucinations; Philosophy, 198, 215-16
specters of. Wonder, 164-66, 212. See also
Visual cone, 65-66 Beauty.
Visual ray, 65-66
Voltaire, 86 Xenophanes, 167-68
Voyage of Life, The (Cole
paintings), 334-36, 337 Yeats, W. B., 251
Young, Thomas, 109-10, 114, 115
Wallace, J. G., 3-4
Wasserman, Robert, 15—16 Zero-point energy, 327-28
Wave-particle duality, paradox of, Zoroastrianism, 41-43
385
Grateful acknowledgment is made for permission to reprint from the following:
Page 33: The Bodleian Library, University of Oxford, Vet. B3.e.l05., p.
125.
Page 60: Theories of Vision from al-Kindi to Kepler, by David C. Lindberg,
edited by Allen G. Debus, copyright © 1976 The University of Chicago
Press.
Page 62: Duccio, The Temptation of Christ on the Mountain, copyright The
Frick Collection, New York.
Page 64: From “Pictorial Perception and Culture,” by Jan B. Deregowski.
Copyright © 1972 by Scientific American, Inc. All rights reserved.
Page 66: Reprinted from The Painter's Manual, by Albrecht Diirer. Translated
by Walter L. Strauss (Abaris Books, 1977), fig. 67, p. 434.
Page 68: Charles D. O’Malley and J. B. de C. M. Saunders, Leonardo da
Vinci on the Human Body, copyright 1952 by Henry Schuman, Inc., New
York.
Page 72: Alinari/Art Resource, New York.
Page 108: Photo courtesy M. Cagnet, M. Francon, and J. C. Thrierr: /Was
optischer Erscheinungen, Berlin-Heidelberg-New York: Springer, 1962.
Page 118: Fundamentals of Physics, 3d ed., by David Halliday and Robert
Resnick, copyright © 1988 by John Wiley & Sons, reprinted by permission
of John Wiley & Sons, Inc.
Page 131: Figure from Physics: For Scientists and Engineers, Third Edition,
copyright © 1990 by Raymond A. Serway, reprinted by permission of Saun
ders College Publishing.
Page 141: Faust in His Study, etching by Rembrandt, B. 270. Reprinted by
permission of Rijksmuseum-Stichting.
387
Credits
Pages 169 and 176: Robert Greenler, Rainbows, Halos, and Glories, Cam
bridge University Press, 1980.
Pages 197, 208, and 209: American Journal of Physics, vol. 44 (1976).
Page 236: Courtesy of Niels Bohr Library, American Institute of Physics,
New York.
Pages 242 and 244: Robert H. Father, Majestic Lights: The Aurora in Science,
History, and the Arts, AGU, 1980, copyright by the American Geophysical
Union.
Page 274: Projective Geometry: Creative Polarities in Space and Time, by
Olive Whicher, copyright © 1971 Rudolf Steiner Press, London.
Page 277: Modern Physics by Kenneth W. Ford, copyright © 1974 by John
Wiley & Sons, reprinted by permission of John Wiley & Sons, Inc.
Page 291: Copyright 1977, Louis I. Kahn Collection, University of Penn
sylvania and Pennsylvania Historical and Museum Commission.
Page 335: Thomas Cole, The Voyage of Life: Childhood, Ailsa Mellon Bruce
Fund, © 1992 National Gallery of Art, Washington, D.C.; Thomas Cole,
The Voyage ofLife: Youth, Ailsa Mellon Bruce Fund, © 1992 National Gallery
of Art, Washington, D.C.
388
JOF
7
“An amazing synthesis and a joy to read—I have not enjoyed h
for a long time....An extraordinary work” ’ q,. ’ boolc so ntm^
author of The Man Who Mistook fikWifefi^^
From ancient times to the present, from philosophers to quant
nothing has so perplexed, so fascinated, so captivated the mind1""1 ?hyS‘cisits, '
nature of light. In Catching the Light, Arthur Zajonc takes us on^ .elusiiv’ e
ney into history, tracing how humans have endeavored to underst^°''Ur-
nomenon of light. Blending mythology, religion, science, literature d* Phe-
ing, he reveals in poetic detail the human struggle to identify th- ’ •
identifythevitakonn^ Pa‘nt’
tion between the outer light of nature and the inner light of th<
heh-nanspir,
Zajonc demonstrates, for example, the complexity of percej
work of Paul Cezanne, the artist standing on the bank of aoption through the
same scene over and over again, the motifs multiplying before J1^his Wing
ds the
goes on to look at our contemporary world, in which science plays? fi^nd he
part in our theories of lights origin—from scientific perspectives stiS gSreatest
Newton’s “corpuscular theory of light” to such revolutionary'iO ®ir Isaac
^Albert
Einstein’s “theory of relativity” and Niels Bohr’s “quantum jump<$|
With rare clarity and unmatched lyricism, Zajonc illuminates I-
implications of the relationships between the multifaceted srrandW^ ound
experience and scientific endeavor. A fascinating search into , man
Beepest
scientific mystery, Catching the Light is a brilliant synthesis tl» Bl both
entertain and inform. . ■$!
“A small gem of a book, poetic in its style and in its determined col
I
ling of
distant ideas.”
The Washington Post
“Brilliant....A beautifully composed meditation.” Kirkus Reviews
Arthur Zajonc is Professor of Physics at Amherst College and a fellow of both
the Lindisfarne Association and the Fetzer Institute.
Cover photographt: David Parker I Photo Rnrarchm (top);
Dana Spaeth I Photonica (bottom) ISBN 178-0-11-501575-3
Cover design by David Tran
Oxford Paperbacks
Oxford University Press lllllllll
□
9 "780195^095753