The Poetry and Music of Science Comparing Creativity in Science and Art Tom Mcleish Full Chapter PDF

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 69

The Poetry and Music of Science:

Comparing Creativity in Science and


Art Tom Mcleish
Visit to download the full and correct content document:
https://ebookmass.com/product/the-poetry-and-music-of-science-comparing-creativity
-in-science-and-art-tom-mcleish/
More products digital (pdf, epub, mobi) instant
download maybe you interests ...

The Art and Science of Social Research (First Edition)

https://ebookmass.com/product/the-art-and-science-of-social-
research-first-edition/

Digital Archaeology: The Art and Science of Forensics


1st Edition

https://ebookmass.com/product/digital-archaeology-the-art-and-
science-of-forensics-1st-edition/

Sturdevantu2019s Art and Science of Operative Dentistry


7th Edition

https://ebookmass.com/product/sturdevants-art-and-science-of-
operative-dentistry-7th-edition/

The Science of Paediatrics: MRCPCH Mastercourse 1st


Edition Tom Lissauer

https://ebookmass.com/product/the-science-of-paediatrics-mrcpch-
mastercourse-1st-edition-tom-lissauer/
The Evolution of Music Through Culture and Science
Peter Townsend

https://ebookmass.com/product/the-evolution-of-music-through-
culture-and-science-peter-townsend/

Therapeutic Modalities: The Art and Science 2nd


Edition, (Ebook PDF)

https://ebookmass.com/product/therapeutic-modalities-the-art-and-
science-2nd-edition-ebook-pdf/

Music, Text, and Culture in Ancient Greece Tom Phillips

https://ebookmass.com/product/music-text-and-culture-in-ancient-
greece-tom-phillips/

Gonstead Chiropractic Science & Art: The Methodology of


Clarence S .

https://ebookmass.com/product/gonstead-chiropractic-science-art-
the-methodology-of-clarence-s/

Explaining Creativity: The Science of Human Innovation


3rd Edition R. Keith Sawyer

https://ebookmass.com/product/explaining-creativity-the-science-
of-human-innovation-3rd-edition-r-keith-sawyer/
THE POETRY AND MUSIC
OF SCIENCE
PRAISE FOR THE POETRY AND MUSIC OF SCIENCE

‘By carefully engaging with historic and contemporary mathematicians,


scientists, writers, artists, and composers, McLeish is able to illuminate the
commonalities inherent in scientific and artistic endeavours—both the
moments of creative inspiration and the diligent, methodical efforts that fol-
low. The author provides the most sustained, learned, and insightful explor-
ation of the mutual characteristics of music and mathematics that I know of.
McLeish offers us an expansive perspective on the human yearning to dis-
cover and create. In doing so, he also provides a release from narrow views of
the sciences and the arts and an opportunity to reconcile diverse academic
specializations. As a musician and historian, I cherish this book for the
insights it gives into both the compositions I perform and the academic
work I engage in. The book is a delight to read and a profound encourage-
ment for those of us who create—on stage, or in academic forums.’
Elaine Stratton Hild, University of Würzburg
‘In this wide-ranging and engaging meditation on the parallel work of
imagination, inspiration, and intuition found in both art and science,
McLeish closes the illusory gap between the two cultures. McLeish, a physi-
cist, heals the divisions in our modern communities of learning by drawing
on his considerable knowledgeable of history, art, philosophy, music and
literature.’
Bernard Lightman, President of the History of Science
Society and Distinguished Research Professor
of the Humanities at York University
‘Tom McLeish has written a book which shows as well as tells. It reinstates
creativity and imagination as being as much part of the practice of science
as close observation, experiment and rigour. Reading it is a pleasure for all
parts of the brain, and the heart as well as the mind. It offers insights to help
everyone interested in the future conduct of science, and in the creation
and sharing of wisdom and beauty in the 21st century.’
Claire Craig, Chief Science Policy Officer, the Royal Society
‘Tom McLeish writes as a lifelong scientist with a rich appreciation of the
arts and a deep historical knowledge. He makes a compelling case for seeing
the scientific and the artistic imagination as intertwined and interdepend-
ent. McLeish shows us that science and the arts never have been, never
could be, and never should be two divided cultures.’
John Holmes, Professor of Victorian Literature
and Culture, University of Birmingham
‘This kind of book is rarer than it should be, and all the more valuable for
that. It dares to take seriously and probe deeply the interplay of the arts and
the sciences, neither setting one above the other nor looking for glib con-
nections but instead interrogating both as expressions of human creativity and
imagination. In place of the tired and reductive notion of “Two Cultures”,
Tom McLeish reveals—passionately, persuasively, and with great scholar-
ship—that so many meaningful points of contact exist between the sciences
and music, literature and visual art that by the end of the book it would
seem strange to put them back in separate boxes. May this be the beginning
of a new and rich conversation!’
Philip Ball, Science Writer
‘Where do creative ideas come from? There is an answer, and it is the same
for ideas in art as it is for ideas in science. There is a hidden wellspring inside
the human mind from which they arise continuously. Tom McLeish pro-
vides meticulous evidence by interrogating the greatest minds. The result is
a brilliant kaleidoscopic view of the history of imagination.’
Uta Frith FBA FRS, UCL Institute of
Cognitive Neuroscience
‘Within the short compass of this subtle and elegant exposition, McLeish
tackles one of the most disabling narratives of our time. Creativity is nei-
ther a luxury nor a disqualification in a world whose survival requires all
our imaginative resources, and it infuses the arts and sciences in uncannily
similar ways. The author has also created a rare and beautiful thing: few
could embrace such a range of artistic and scientific endeavour with such
an uplift.’
Marilyn Strathern DBE, Professor of Social Anthropology,
Cambridge University
‘Reading this book is like eavesdropping on a series of intimate and detailed
conversations among a set of accomplished scientists and artists on the sub-
ject of imagination, play and creativity, conversations in which McLeish
himself participates. What McLeish’s book reveals is the astounding fact
that these conversations are seamless, involving moments of transform-
ation and insight that are not only familiar but frequent.’
Robert Crease, Chair of the Department of Philosophy,
Stony Brook University
‘An interdisciplinary tour-de-force of astonishing clarity and depth, this
book blows apart the “Two Cultures” argument for the rigid standoff
between the sciences and the rest of the arts and humanities. McLeish is a
21st century Renaissance man, and his words are just the tonic for anyone
who loudly doubts that science involves creativity—as well as for those
who quietly know from lived experience that it certainly does.’
Jennifer Rohn, Principle Research Associate at University
College London and author of Cat Zero
‘Many things have been said and written about the connections between
maths and music but I have never encountered anything remotely as deep
as this. McLeish exhibits mind-boggling inter-disciplinarity with depth as
well as breadth, and his prose is as rich in poetry and music as the subject of
the book.’
Eugenia Cheng, Scientist in Residence at the School
of the Art Institute of Chicago
‘The Poetry and Music of Science will be recognized for years to come as a tri-
umph that bridged the sciences and the humanities in an effort to call out
the creativity and imagination that feeds them both. This book on creativ-
ity is creatively written and a pleasure to read.’
Elaine Howard Ecklund, author of Science vs. Religion
‘Anyone who believes that imagination, inspiration and creativity are the
preserve of the arts should read this beautifully crafted ode to the enter-
prise of scientific discovery.’
Jim Al-Khalili OBE FRS, Professor of Theoretical Physics,
University of Surrey
THE POETRY
AND MUSIC OF
SCIENCE
Comparing Creativity in Science and Art

Tom McLeish

1
1
Great Clarendon Street, Oxford, OX2 6DP,
United Kingdom
Oxford University Press is a department of the University of Oxford.
It furthers the University’s objective of excellence in research, scholarship,
and education by publishing worldwide. Oxford is a registered trade mark of
Oxford University Press in the UK and in certain other countries
© Tom McLeish 2019
The moral rights of the author have been asserted
First Edition published in 2019
Impression: 1
All rights reserved. No part of this publication may be reproduced, stored in
a retrieval system, or transmitted, in any form or by any means, without the
prior permission in writing of Oxford University Press, or as expressly permitted
by law, by licence or under terms agreed with the appropriate reprographics
rights organization. Enquiries concerning reproduction outside the scope of the
above should be sent to the Rights Department, Oxford University Press, at the
address above
You must not circulate this work in any other form
and you must impose this same condition on any acquirer
Published in the United States of America by Oxford University Press
198 Madison Avenue, New York, NY 10016, United States of America
British Library Cataloguing in Publication Data
Data available
Library of Congress Control Number: 2018959076
ISBN 978–0–19–879799–9
Printed and bound by
CPI Group (UK) Ltd, Croydon, CR0 4YY
Links to third party websites are provided by Oxford in good faith and
for information only. Oxford disclaims any responsibility for the materials
contained in any third party website referenced in this work.
To Juanita Kalerghi Rothman,
who exemplified how loving and thinking of science, writing,
music and art bring them together in the human soul.
Preface

This all began on a school visit—I believe in the North Yorkshire


town of Harrogate. I was a visiting university speaker to a group
of bright high-school pupils in support of their ‘General Studies’
activity. I can’t be sure now of the topic—it might have been the
history of science, or the science of plastics, or even the fascinat-
ing story of religion and science—but the message for me came
during the discussion. The UK school system, as many readers
will know, is uniquely specialized after the age of 16, by any inter-
national comparison. This meant that some of the 17 and 18 year-
olds in the group were no longer studying any science or
mathematics topics at all. From the level of their discussion many
of them were clearly extremely bright and could have pursued
any subject they had wished, so I asked them why they had not
chosen to continue with any further studies in science. ‘Because I
saw no room for any imagination, or my personal creativity, in
science’, was the common response. I still remember feeling pain,
and knew then that something had gone very awry in the way we
talk about science in education and in the media. My own experi-
ence as a practicing scientist had been completely different. The
school I attended a generation before had unusually allowed me
to continue studies in French, alongside sciences and mathemat-
ics, before going to university, and there I had been helped to see
that science cannot begin without the first creative step. I began
to perceive that these two fortunate experiences were not unre-
lated. Scientists had failed somehow to communicate the creative
core of science to the young people I was with, and clearly to
many others besides.
The later privilege of working as a University Pro-Vice-
Chancellor for Research (the post would be called Vice-President
for Research in the USA) at Durham University gave me a won-
derful daily insight into the thinking of imaginative academic
x Preface

colleagues from the humanities and social scientists, as well as of


the university’s great scientists. I had, of course, read C. P. Snow’s
Two Cultures, and recognized the failed intercultural communi-
cation that concerned him, but became increasingly convinced
that our disciplines share far more than we recognize. Our
schools and universities are not in general structured in support
of interdisciplinary conversation, but whenever we created
space to talk, some precious spark seemed to be struck. Some of
those sparks took hold and became serious projects that are still
growing.
The first thanks therefore go to the institutions and individ-
uals at Durham who gave their time, energies, and encourage-
ment to an interdisciplinary vision. Professor Veronica Strang,
Executive Director of Durham’s Institute for Advanced Studies is
an interdisciplinary inspiration, and kindly hosted the introduc-
tory seminar that started the project that became this book. Prof.
Giles Gasper is the visionary leader of the extraordinary Ordered
Universe collaboration which brings the complementary perspec-
tives of medieval scholars and scientists together to study the
­science of the thirteenth century. This project shows in a fascinat-
ing way that science possesses a much longer and deeper story
that is usually told. It also illustrates in a tangible way that scien-
tists and humanities scholars have much in common, while
introducing me to the extraordinary thirteenth-century poly-
math Robert Grosseteste, who will also assist us at several points
in this book. David Wilkinson, Principal of St. John’s College
Durham, has my eternal gratitude for his intellectual, faithful,
and personal encouragement to frame theology and science in
fresh ways. I am greatly indebted to Patricia Waugh for pointing
me to the entanglements of science and the writing of novels, to
Julian Horton for a shared love of Schumann and for his expert
analysis of the Konzertstück, to Martin Cann for a long and patient
collaboration between biology and physics.
An invitation from the Notre Dame Institute for Advanced
Study (Indiana) as Visiting Fellow 2017–8 was instrumental, especially
Preface xi

in the research for Chapter 4, but also in support of the book at a


whole. NDIAS Director Brad Gregory inspired me with his vision
for the ‘unity of knowledge’, and I am deeply grateful for the dis-
cussions, readings, and comments of other NDIAS fellows and
ND academic staff, especially Laura Dassow-Walls, Celia Deane-
Drummond, Steve Fallon, Margot Fassler, Patrick Griffin, Chris
Kolda, David Bentley Hart, Elaine Stratton Hild, Henricke Lange,
Bernie McGinn, Richard Taylor, Emily Dumler Winckler, Xinyu
Dong, and my two wonderful undergraduate research assistants
Sofia Carozza and Jeremy Cappello Lee. Without the creative sup-
port of Don Stelluto and Carolyn Sherman, none of that would
have happened.
The third important setting for this book is my new academic
home of the University of York. The imaginative decision to cre-
ate an interdisciplinary Chair of Natural Philosophy, so regener-
ating a tradition dormant in England for about two centuries, has
already proved a joy to work in. In a short time new colleagues
have already nourished the content of this book, including and
especially Mary Garrison, Keith Allen, Kevin Killeen, Jeanne
Nuechterlein, and Elizabeth Tyler.
Listening to artists, art curators, musicians, scientists, and edu-
cationalists talk about their work candidly and honestly has been
an immense privilege and an essential source of wisdom. Thanks
go especially to the Ordered Universe project’s artists in residence
and collaboration Alexandra Carr, Colin Rennie and Cate
Watkinson, and to Vanessa Chamberlin, Janet Graham, Jeremy
Mayall, Graeme Willson, Berry Billingsley, Michael Reiss, Stephen
Blundell, Carl Gombrich, Julie Kornfield, Katie Lewis, Stephen
Blundell, Ard Louis, Wilson Poon, Scott Milner, Christopher
Southgate and Hannah Smithson. I am grateful to Jill Cook of the
British Museum for the inspiration of her Living with the Gods exhib-
ition, and to Jennifer Thompson of the Philadelphia Museum of
Art for access to and discussion of Monet’s Pines at Sunset. Of great
value was the Imagination Institute workshop held in Kings College,
Cambridge in 2016. For this thanks are due to Robert Sternberg,
xii Preface

James Kaufman, Martin Rees, Ashley Zauderer, Michael Berry,


Michael Cates, Melissa Franklin, Philip Ball, Naomi Leonard, Jon
Keating, John Pendry and Herbert Huppert.
I am very grateful to friends and colleagues who have gone the
extra mile by reading and commenting on drafts, often suggest-
ing vital new discussions and sources that I would otherwise have
missed, but at the same time kindly picking up a myriad of small
errors. Any that remain sit very firmly at my door. Especial
thanks go to Carl Gombrich, Mary Garrison, Michael Reiss,
Andrew Steane, and Adrian Sutton for comprehensive over-
hauls, and to Mark Miller, Julie McLeish, Marilyn Strathern,
Victoria Lorrimar, and Iain McGilchrist for particular comments.
Thanks go to Rosie McLeish for very efficient typesetting of the
musical quotations in Chapter 5. Those on this list who share my
surname have also supported the necessary labours with their
patience and love more than they know. Juanita Kalerghi
Rothman has been a lifelong inspiration to connect the scientific
and the artistic, to find aesthetics in engineering and contemplation
in creativity of all kinds. To her this book is gratefully dedicated.
The editorial team at OUP, especially Ania Wronski and Sonke
Adlung, are a perpetual joy to work with, and are most certainly
due thanks for their encouragement but especially their patience.
It was fun.
Tom McLeish
York,
Feast of St. Barnabas, the Son of Encouragement, 2018
Contents

1.  Introduction: Creativity and Constraint 1


2. Creative Inspiration in Science 29
A private moment of discovery 32
Entangling thoughts 39
The strangeness of star-shaped molecules 46
Underground rivers of the mind 52
The Creativity of the New—and doing biology
in a physics lab 58
A conversation about creativity in science 65
3. Seeing the Unseen: Visual Imagination and the
Unconscious 72
The visual metaphor within the scientific imagination 78
Mathematical theory-painting 80
The ancient aesthetic of active seeing 83
The creativity and constraint of a visual project 90
A scientific experience of a visual idea 93
Conversations on creativity with visual artists 99
The great cosmological model and the visual
imagination 103
The visual imagination and astronomy today 113
The excited imagination of the impression 117
An artistic theory of music 124
4. Experimental Science and the Art of the Novel 128
A shared early history 134
The orbits of the early novel and science 137
Newton and Milton—Paradise and procession 140
The art of the probable and the hermeneutic
stance of Robert Royle 143
The arts of fiction and science 150
xiv Contents

Ideation 152
Incubation 154
Illumination 157
Verification and the constraint of form 162
Entanglements of science and literature 168
Humboldt, Emerson, Wordsworth, and the Romantic
scientific aesthetic 170
Émile Zola, Claude Bernard, and the ‘experimental novel’ 178
Creativity and constraint in the novel of the twentieth
century—The Paris Reviews and a Nobel lecture 181
Scientific discovery and the novel 187
5. Music and Mathematics—Creating the Sublime 191
The numerical threads of music 195
Music—the medieval mathematical art 200
Augustine on music 204
Robert Schumann—creative tension, form, and genre 208
A musical close-reading: The Konzertstück for
Four Horns and Orchestra (1849) 219
From music to mathematics 233
A mathematician’s mind 237
A mathematical close-reading and a beautiful
connection—the fluctuation–dissipation
theorem244
An electrical analogy—Johnson–Nyquist Noise 250
A universal truth 253
Music, mathematics, and wordless creation 259
6. Emotion and Reason in Scientific Creation 261
Early modern echoes 274
Scientific testimony to the emotion of ideas 280
Putting out fire with fire: a case study in
creative scientific affect 287
David Bohm on creativity 295
Picasso and Guernica—a documented journey
of aspectus, affectus, and art 297
Contents xv

7. The End of Creation 301


An ur-narrative of creative experience 303
Telling stories of creativity and creation through
theological lenses 312
All the colours of the rainbow 315
The end of creativity 323
Modern minding of the gap 333

Bibliography 341
Index 349
(a) (b)

vorticity maps ω/⟨∣ω∣⟩


(c) (d) 4

–4

Colour Plate A. Experiments (A—microscopy and B—a plot of


velocity contours and vorticity strength in colour scale) and models
(C—a mesoscopic model and D—a continuum theory) of a turbulent
quasi-2D flow of dense swimming bacteria (B. subtilis). Reproduced with
permission from Wensink, H. H. et al. (2012), ‘Mesoscale turbulence in
living fluids’, Proceedings of the National Academy of Sciences, 109, 14308–14313

Colour Plate B. Example of the Ashihda and Kitaoka’s ‘Rotating


Snake Illusion’. Letting one’s eyes wander over the image creates for
most viewers an illusion of rotational motion in the circles. Reproduced
with kind permission of Akiyoshi Kitaoka
(a)

(b)

Colour Plate C. (A) ‘Adoration of the Magi’, 1303–05 (fresco), Giotto


di Bondone (c. 1266–1337)/Scrovegni (Arena) Chapel, Padua, Italy/De
Agostini Picture Library/A. Dagli Orti/Bridgeman Images. (B) ‘Last
Judgement’, 1303–05 (fresco), Giotto di Bondone (c. 1266–1337)/Scrovegni
(Arena) Chapel, Padua, Italy/Mondadori Portfolio/Archivio Antonio
Quattrone/Bridgeman Images
Colour Plate D. ‘L’image Du Monde’ by Gossuin de Metz, With per-
mission, Bibliothèque Nationale de France
Colour Plate E. Claude Monet; ‘Under the Pine Trees at the End of
Day’ (1888), Philadelphia Museum of Art, 125th Anniversary Acquisition.
Gift of F. Otto Haas and partial gift of the reserved life interest of Carole
Haas Gravagno, 1993-151-1 with permission of the Philadelphia Museum
of Art
Colour Plate F. Graeme Willson (1998); ‘Quartet’, Private Collection,
with permission

Colour Plate G. ‘Robert Boyle’ (1627–91), by German painter Johann


Kerseboom (d. 1708), with permission of The Royal Society
Colour Plate H. Pablo Picasso (1937); ‘Guernica’ with permission, Reina Sofía Museum Collection, Madrid
Colour Plate I. The ivory ‘Lion Man’, Schwabia, c. 40 000 bce, with
permission of the Museum of Ulm
(a) (b)

(c)

100

80

60
L*

40

20
100
50
–50 –25 0
0 25 –50 b*
50
a*

(d) (e)

Colour Plate J. (a) Theodoric’s diagram of the formation of the rain-


bow through double-refraction and reflection of sunlight in raindrops
towards an observer; MS Basel F. IV. 30, ff. 33v-34r. (b) a modern illustra-
tion of the dispersion in angles of refracted and reflected light within a
raindrop. (c) A primary and secondary rainbow https://www.pinterest.com.au/
pin/232287293250306356/?autologin=true. (d) and (e) Art installation by
Colin Rennie, Sunderland Art Museum ‘Illuminating Colour’. (b) and
(d) reproduced with permission of Prof. Hannah Smithson
N 1O
Introduction

Creativity and Constraint

Art is limitation; the essence of every picture is the frame.


G. K. Chesterton

Creativity, Inspiration, Passion, Form, Imagination, Composition,


Representation—this powerful list of words leads a reader’s mind
inevitably into the world of the arts. Perhaps it conjures up the
shaping of a block of stone into the form of supple limbs and
torso, or layering darkly tinted oil paints onto canvas to tease the
eye into imagining a moonlit forest at night. Others may think of
a composer scoring a symphony’s climax—she summons the
horns to descend as from a distant mountain peak to meet a har-
monically ascending string bass-line in a satisfying resolution. A
poet at his desk wrestles with meter and rhyme as he filters the
streams of words, metaphors, and allusions that clamour for
place on the page. The double miracle of art is not only that it
allows humans to draw meaning from the world, but also that it
reaches out to its listeners, viewers, and readers so that they may
re-create for themselves something new and personal in response.
Both by words and by images we are changed, troubled, made
more aware as art enriches us in small ways or great. To engage in
art by creation or reception and re-creation is to exercise one of
the capacities that make us human. Indeed, the academic study
of art’s products and process falls under the class of disciplines we
call the ‘humanities’.
2 The Poetry and Music of Science

Experiment, Design, Formulation, Method, Theory, Observation,


Hypothesis, Computation, Trial, Error—another list of words
might lead to a different world of activity. These are more associ-
ated with disciplines we term ‘the sciences’. Their energy seems to
be of a different sort—we are not, perhaps, as emotionally moved
by these terms; they do not suggest as much wild, unpredictable
outcome. Are we encouraged to think, perhaps, of a laboratory
setting—a careful mixing of liquids and a measuring of their
temperature? Is the mental picture one of an observer carefully
preparing a microscope, or calculating by computer the orbit of
a distant planet? If the artistic associations are as likely to disturb
as to excite, are the scientific associations more reassuring (the
French cubist Georges Braque thought that, ‘L’art est fait pour ­troubler,
la science rassure’1)? Or do they disturb in a different way? Very likely
this is a world that is unfamiliar and strange, less accommodating
than the arts and, dare we admit it, less ‘human’ in some way (we
do not class the sciences as ‘humanities’ after all).
But there are other voices that choose the same language to
talk about art and science, and even in the same breath. Philosopher
of science Karl Popper once wrote: ‘A great work of music, like a
great scientific theory, is a cosmos imposed upon chaos – in its
tensions and harmonies inexhaustible even for its creator’2. This
richly layered and dense commentary on music and science will
need some background work to uncover Popper’s meaning—its
allusions immediately fail to intersect with the quite distinct
word-lists that spring from usual talk of art and science. But it
raises suspicions. Is a dualistic division into arts and science really
faithful to our history, our capacities and needs? Does it spring
from a deep understanding of what these twin human projects
attempt to do—is it faithful, dare we ask, to their purpose? And if
not, are we right to ask of our children, ‘are they on the science-
side or the arts-side?’ or to reinforce the well-worn narrative of
1 ‘Art is made to disturb, science to reassure’.
2 Karl Popper (1976 [2002]), Unended Quest: An Intellectual Autobiography. London
and New York: Routledge.
Introduction 3

C. P. Snow that there are ‘Two Cultures’3 at work in our late-


modern world, non-overlapping, mutually incomprehensible,
and doomed to conflict? If we are wrong to categorize culture, let
alone people themselves in this way, then to make exclusive edu-
cational decisions based on such a dualistic assumption will be to
trigger a process of atrophy in one or other aspect of those chil-
dren’s development, and in adult life to have closed off one or
other world of expression, contemplation, creativity, enrich-
ment—of complementary ways of being human.
Doubts intensify about a neat cultural divide if we take the all
too unusual step of listening to an artist, or to a scientist, talk
candidly about their journeys of labour from early ideas to a fin-
ished work. For now, the language-clouds of the arts and the sci-
ences start to collide and overlap. I have an intense memory of
my first lengthy conversation with an artist (also a professor of
fine art at my then home university of Leeds) about our r­ espective
experiences of bringing to light new work in art and in science4.
He spoke of his first experimental attempts to realize an original
conception, of the confrontation of his ideas with the felt con-
straints of material—of paint and photographic print, of the
­necessary reformulation of the original concept, of the repeat of
these frustrated assays not once but many times. I found that I
could tell the story of almost any programme of scientific research
I had experienced in almost precisely the same terms. The discov-
ery was mutual: if I had been surprised by the element of experi-
mentation and trial in his artistic project, then he had not
expected my story of science to speak so vitally of the role of
imagination. Not only that, it became clear to us that not only did
the intellectual and technical histories of our projects map closely
onto each other, but so also our emotional trajectories of excite-
ment, hope, disappointment, rekindling of hope, and resolution
found common expression. The more honest the story-telling in
3 C. P. Snow (1959 [1998]), The Two Cultures. Cambridge: Cambridge University
Press.
4 Professor Ken Hay’s project is discussed in detail in Chapter 3.
4 The Poetry and Music of Science

art or science, the more entangled and related became the experi-
ences of emotion and cognition. Thinking and feeling are closer
under the surface than in our public stories.
Why it is so much less common to discuss the long process of
realization in art than to talk about the final article, composition,
theory, or painting is hard to say. The famous exceptions (such as
the evolution of Picasso’s Guernica5, or the candid reflections of Henry
James’ Art of the Novel6) underline the question. Maybe it has to do
with the tradition of artisan and artist guarding carefully the ‘secrets’
of their trades, thereby to increase by mystique as well as by wonder
the appeal of the finished article. Art has commercial value too—its
finely honed techniques and formulae are secrets worth keeping,
even though they are the vehicles, not the sources, of inspiration. Or
perhaps there is less intention to weave a whimsical web of mystery
around the production of art than a natural reluctance to admit too
much of the false starts, errors, spilt ink, confused ideas, and dead
ends that are the daily experience of any creative activity.
If art is shy about the sweat and tears of working out the form
of an original idea, then science is almost silent about its e­ piphanies
and moments of inspiration. Popper himself, celebrated for the
most detailed modern outworking of a scientific method in his
Logic of Scientific Discovery, wrote at length on how hypotheses may
be refuted, but remained quiet on how they might be imagined in
the first place. While acknowledging the vital necessity of such
imaginative preconception, Popper declared that, as essentially
non-methodological; he had nothing to say about it. There is
some degree of logic and process in the testing and evaluating of
a scientific idea, but there are no such recipes for conceiving
them. Nobel Laureate Sir Peter Medawar lays some of the blame
for our blindness to the role of imagination in science at the feet
of John Stuart Mill’s System of Logic, who writes as if he believed

5 Rudolf Arnheim (1963) The Genesis of a Painting: Picasso’s Guernica Berkeley:


The University of California Press.
6 H. James (1934), The Art of the Novel. Chicago: University of Chicago Press
(2011).
Introduction 5

‘that a scientist would have already before him a neatly ordered


pile of information ready-made—and to these he might quite
often be able to apply his rules.’7 If science gathers to itself a nar-
rative more weighted towards method, and art is more vocal
about creative origins, then these retellings of partial truths will
conspire to drive an illusory distance between them.
Some considerable historical work will be necessary to trace the
origins of buried commonalities of art and science. Even the words
we use to discuss them bear hidden references. A good example
to start with is ‘theory’. Seventeenth-century puritan writer
Thomas Browne was able to say in his 1643 Religio Medici that8
. . . nor can I thinke I have the true theory of death, when I contemplate a skull, . . .

The ancient Greek Orphic priests seem to have used the word
theoria to describe ‘passionate, sympathetic contemplation’9. The
explicitly visual, religious, and contemplative implications at the
root of devising a theory will arise within a close examination of
visual imagination in Chapter 3.
The contrasting traits of silence within the community of
­science on its imaginative energies, and of art on its workaday
reckoning with material reality, is not restricted to our own
times. William Blake, the eighteenth-century poet, artist, and
engraver, famously inveighed against what he perceived was the
destructive dehumanizing of ‘natural philosophy’, the term used
for the quantified and experimental understanding of nature we
would term ‘science’ today. He wrote of his own task:
‘in the grandeur of Inspiration to cast off Rational Demonstration . . . to cast off
Bacon, Locke and Newton’; ‘I will not Reason and Compare – my business is to
Create’.10

7 Peter Medawar (1984), An essay on scians, in The Limits of Science. Oxford: OUP.
8 Kevin Killeen, ed. (2014), Thomas Brown, Selected Writings. Oxford: Oxford
University Press, p. 47.
9 Francis Cornford (1991), From Religion to Philosophy: A Study in the Origins of
Western Speculation. Princeton: Princeton University Press.
10 William Blake, Milton (1804), book 2, pl. 41; Jerusalem, ch 1, pl. 10.
6 The Poetry and Music of Science

For Blake, inspiration has no place in Newton’s work, and reason


none in his own. There is some buried personal dissonance here
given what we know of his own painstaking technical develop-
ments in copper engraving. Yet he was not without cause for
complaint against those early modern philosophers: John Locke,
in his Essays Concerning Human Understanding11 had identified ‘the
imagination’ as the source of false and fantastical ideas, as opposed
to experience, the reliable guide to the true.
I frequently find the same compartmentalized assignments of
inspiration and rationality at work today when talking with high-
school pupils. When visiting schools for ‘general studies’ discus-
sions of science in society or the importance of interdisciplinary
thinking, I like to ask advanced students who have not chosen to
study science subjects (when from their intellectual engagement
with the material it is clear that they could master anything they
chose to) why they made that choice. Among the brightest of
them, I never receive the complaint that the sciences seem too
difficult, but rather that they appear to lack avenues for creativity
and the exercise of imagination. The conversation sometimes also
reflects the expectation of a more playful engagement with the
humanities, contrasting with a notion of seriousness and narrow-
ness in the sciences. I find this painful, doubly sad that these
young people have been offered no insight into the immense
fields for imagination offered by science, and that scientists have
failed in communicating its call on creativity. As Sîan Ede12 writes:
Compared with the cool rationalism of science with its material belief in wholeness,
the theories employed by thinkers in the arts and humanities seem part of a playful
circular game in which the truth is never to be privileged in one direction or another
and is always out of reach.

These echoes of Blake in the words of today’s brightest young


people are painful to hear. They speak to the urgency of a project

11 John Locke (2015), The Clarendon Edition of the Works of John Locke, Oxford:
Oxford University Press, An Essay Concerning Human Understanding Book II.
12 Sîan Ede (2005), Art and Science. London I. B. Taurus.
Introduction 7

that goes beyond the confrontational assumptions of the


‘Two Cultures’ to deeper levels of human motivation, desire,
­experience—one that recognizes the dual qualities of rationality
and inspiration, of seriousness and playfulness, of imagination
and constraint, but challenges their automatic alignment with
the axes of humanities and sciences, exploring instead how they
play out in both.
Admittedly it has never been easy to speak with clarity about
moments of imaginative conception. When inspiration eventu-
ally comes, articulating the experience faithfully is fraught with
difficulty. There is a wordlessness about those moments of vision
that permit and initiate the necessary phase of craft within the
creative process, that plant germs of energy and glimpse distant
impressions of what might be accomplished and perceive the
direction of the road ahead. We know how to desire the moments
that William Whewell called ‘felicitous strokes of inventive tal-
ent’13, but not how to summon, and hardly to describe them.
Here is Shakespeare struggling to do both (and succeeding) in his
100th Sonnet:
Where art thou Muse that thou forget’st so long,
To speak of that which gives thee all thy might?
Spend’st thou thy fury on some worthless song,
Darkening thy power to lend base subjects light?
Return forgetful Muse, and straight redeem,
In gentle numbers time so idly spent;
Sing to the ear that doth thy lays esteem
And gives thy pen both skill and argument.
Rise, resty Muse, my love’s sweet face survey,
If Time have any wrinkle graven there;
If any, be a satire to decay,
And make Time’s spoils despised every where.
Give my love fame faster than Time wastes life,
So thou prevent’st his scythe and crooked knife.

13 William Whewell (1837), History of the Inductive Sciences. London: John


W. Parker.
8 The Poetry and Music of Science

The poet longs for the return of his ‘Muse’, one of the ancient Greek
mythological band of personified inspiration recorded as early as
Hesiod (c. 600 bce) and that Plato invokes to explain acts of creation14.
Shakespeare pleads for his to sing him songs and guide his pen on
the page before Time takes away all opportunity for further art. Yet,
paradoxically he makes this ‘time between inspirations’, this osten-
sibly dry season of complaint—into a sublime sonnet. The actual
song of the Muse is muffled, and the sight of her hidden, by the
humorous complaint of her absence. Unnoticed, the poet’s
­imagined sole source of inspiration, the face of his beloved, is
replaced by his rising outrage that the Muse refuses to come at his
beck and call. In the compressed lines of the sonnet, the first stage of
creativity is conflated with the second; the visitation of inspiration
itself is brought together with the ‘skill and argument’ of the pen.
The long labour of writing must then do battle with Time itself.
Ironically, it is the wasting erosion of time that becomes the topic of
the final work of art. The poet knows that even if inspiration comes,
time is not his friend during the long process of gestation before a
poem lies completed upon the page. No scientist can read this
sonnet without stirring frequent memories of dry days and weeks
when ideas fail to come and fruitfulness all seems elsewhere.
After listening to Shakespeare on inspiration and labour in art,
perhaps we ought then without delay turn to Einstein on creativ-
ity in science (why descend from the summit of Olympus before
you have consulted all of its dwellers?). Here is the master and
motivator of twentieth-century physics on the two components
of creativity15:
The mere formulation of a problem is far more essential than its solution,
which may be merely a matter of mathematical or experimental skills.

14 In most ancient sources there are nine: Calliope (epic poetry), Clio (his-
tory), Euterpe (flutes and lyric poetry), Thalia (comedy and pastoral poetry),
Melpomene (tragedy), Terpsichore (dance), Erato (love poetry), Polyhymnia
(sacred poetry), Urania (astronomy).
15 Albert Einstein and Leopold Infeld (1938), The Evolution of Physics. London:
Cambridge University Press.
Introduction 9

To raise new questions, new possibilities,


to regard old problems from a new angle, requires creative imagination
and marks real advance in science.
~
I am enough of an artist to draw freely upon my imagination.
Imagination is more important than knowledge. Knowledge is limited.
Imagination encircles the world.
Both Shakespeare and Einstein, as they open the door into their
workshops for us —albeit in very different forms, tell of two
phases in the creative process. The first, ‘the visit of the Muse’, the
‘creative imagination’, or ‘the mere formulation’ is the i­ nspiration,
the genesis of an idea. The second is a longer, more directed pro-
cess of developing the germ into its ‘song’ or ‘solution’. It is
intriguing that Einstein chooses to explain his knowledge of the
wellspring of imagination—he ‘draws’ from it—by describing
himself as an artist. He elevates ‘imagination’ as he demotes
‘knowledge’. He wants to make clear that the greater task in sci-
ence is the ‘mere formulation’ of the problem in the first place,
rather than the application of methods to its solution (by ‘mere’
he means ‘fundamental’, ‘elementary’, or ‘constitutive’ rather
than ‘trivial’ of course). The great scientist knows that we find
our way to encircling the Earth, not principally by experiment,
theory, deduction, falsification, or any of those important fea-
tures of scientific method, but by imagination.
Thomas Kuhn famously coined the notion of ‘paradigm shifts’
to denote discontinuous changes in the scientific framework for
understanding nature. They entail revolutions in entire sets of
presuppositions and current mutually supporting scientific ideas.
They typically witness the entry of new ideas not deducible from
prior reasoning16. Classic examples are the Copernican revolution
in cosmology and the shift from classical to quantum physics.
Beyond identifying the growing dissatisfaction with the existing
framework, Kuhn made no suggestions concerning the provenance
16 Thomas Kuhn (1966), The Structure of Scientific Revolutions. Chicago: Chicago
University Press.
10 The Poetry and Music of Science

of the new set of ideas—they are the protoplasm of his revolu-


tions, but the seed of no methodology.
The formulation of the fruitful question, posed in the right
way, constitutes the great imaginative act in science. It requires
a developed sense of the current age of thought, of timing.
Historian of science and chemist Lawrence Principe17 has pointed
out the appropriateness of asking the question about the struc-
ture of the solar system at the turn of the seventeenth century:
when Tycho Brahe’s meticulous observations of planetary motion
and Copernicus’ inspired partial solution to the new paradigm
had opened up a field of potential progress. Kepler’s deductions
together with Thomas Harriot’s and Galileo’s assays in turning
the new telescopes towards the heavens, made asking about
the dynamical consequences of gravity among the sun and the
­planets fruitful in a sense that it had not in any previous century.
If the scientific imagination is fed by the creative and timely ques-
tion, it also needs the nourishment of the discontinuous, of leaps
in thinking that receive their impulse from some other source
than the worthy process of logical deduction. A generation on
from the establishment of the orbits of the moon and planets
within a heliocentric structure of the solar system, Newton’s
great imaginative conception was to contemplate a world in
which the fall of an apple sprung from the same universal field of
force as the monthly procession of the moon.
Einstein and Principe point us to the critical role of the well-
formulated and timely question in science. Yet the silence of any
formulation of ‘scientific method’ on this great creative act tends
to mask its pivotal role, as well as to dull perception of scientific
creativity. Sharpening and highlighting the role of creative
­imagination to a degree faithful to the way science actually works
would deepen the level at which the fundamental motivation for
science is recognized within human culture. An account of sci-
ence that accords appropriate weight to divergent re-imagination
17 Lawrence Principe (2013), The Scientific Revolution: A Very Short Introduction.
Oxford: Oxford University Press.
Introduction 11

of nature as much as to convergent deductive method also finds


itself telling a much longer historical story than the exclusively
post-enlightenment narratives that dominate at present. A liter-
ature-search for the question-form when addressed to the ­natural
world, for example, lengthens the historical line over which we
can map early stirrings of the desire to understand and contem-
plate nature. I have elsewhere written at length on the beautiful
and profound ancient wisdom poem, the ‘Lord’s Answer’ of the
Old Testament Book of Job18. A work of over a hundred verses, it
assumes the unusual poetic form of the repeated question—but
in the light of this discussion appropriately so—since its subject
matter is the human understanding of nature. The origin of light,
the formation of the coastlines and mountains, the provenance
of the hail and lightening, the ability of birds to navigate the Earth
in their migrations—all appear in a grand cosmic sweep of its
author’s enquiry. One stanza from chapter 38 in particular would
have impressed a Newton or an Einstein:
Can you bind the cluster of the Pleiades, or loose Orion’s belt?
Can you bring out Mazzaroth in its season, or guide Aldebaran with its train?
Do you determine the laws of the heaven?
Can you establish its rule upon earth?
The poet looks at the motions of the stars and constellations
across the sky, even noticing that some cluster together while
others, though visually similar in brightness and hue, are
­separated. The lovely Pleiades is one of the very few ‘open star
clusters’ resolvable to the human eye, with up to six or possibly
seven members visible to keen-eyed northern hemisphere
observers during autumn and winter nights. An outstretched
hand-breadth to the south-west of them lie the linear triplet of
bright blue-white stars in ‘Orion’s belt’, far further from each
other than the members of the cluster. The presence of a strange
class of ‘law’, to be obeyed not by humans but by the stars them-
selves, that might contain the statutes binding some stars closely
18 Tom McLeish (2014), Faith and Wisdom in Science. Oxford: Oxford University Press.
12 The Poetry and Music of Science

together while others are far-flung, that oversee their regular and
irregular motions, is an impressive creation of the imagination
even now. The conjecture, buried in such ancient philosophy,
that heavenly and earthly laws might be connected is even more
striking.
The presence of the creatively formulated question in as
ancient a source as the Book of Job19 (undatable other than to
place it within the first half of the first millennium bce) within
the Semitic tradition, carries another salutary message to us late
moderns. Alongside the complex history of ancient Hellenistic
science from 500 bce, it surely erodes any idea that science is in
any way exclusively modern, beginning rootless at the enlighten-
ment and blowing away the cobwebs of centuries of darkness,
magic, superstition and alchemy. Sadly, much popular narrative
of science history has it so, but claiming science as an exclusive
property of the modern world removes the deep and slow cul-
tural development of an imaginative and creative engagement
with nature that develops, at least chronologically, alongside the
story of art in its own multitude of forms20.
The timely question is not the sole province of science. It is
surely not coincidental that the literary genre of the novel arises,
with Daniel Defoe, alongside early modern science, and that the
genre of science fiction underwent an explosion in the last cen-
tury. As Pat Waugh21 has pointed out, the novel is the experimental
medium of artistic creation par excellence. In the safe space of the
novel, inhabited worlds can be summoned into existence and

19 For a magisterial survey of the Book of Job, see the three-volume work by
David Clines, Thomas Nelson pubs. We will encounter an explicit example of
scientific inspiration drawn from reading it in the story of the rainbow, told in
Chapter 7.
20 For a more complete and integrated account of the history of science, see
the now classic work by David Lindberg (2010), The Beginnings of Western Science: The
European Scientific Tradition in Philosophical, Religious, and Institutional Context, Prehistory to
A.D. 1450, Second Edition, Chicago: Chicago University Press.
21 P. Waugh (2015), Beauty writes Literary History, in Corrine Saunders and
Jane MacNaugten, eds., The Recovery of Beauty. London: Palgrave MacMillan.
Introduction 13

their dangers and dark places explored. Questions of the relation-


ship of human beings to time and space, to each other and to the
Earth can be teased out in both internal and external worlds of
their characters. The novelist does not experience unconstrained
freedom, however, but discovers the multiple moral constraints
of the experimental form. Crucially, novelistic writing forces a
more intense outward gaze. For Iris Murdoch, novelistic writing
enables an attention to ‘. . . the inexhaustible detail of the world,
the endlessness of the task of understanding . . ., the connection
of knowledge with love and of spiritual insight with the appre-
hension of the unique’22. Murdoch supplies us with another
glimpse of commonality in the narrative that artists, creative
writers, and scientists adopt when they are trying to articulate
the deepest motivations for what they do. ‘The endless task of
understanding’ and a focus on the ‘inexhaustible detail of the
world’ are the shared delights and common labours of the physi-
cist and biologist also. These novelistic hints are enough to tempt
the much closer look at the relationship between experimental
science and literature in Chapter 4.
Paying close attention to the stories of imagination and work-
manship in the creation of art and science will be the first of this
book’s tasks in the project of reappraising science through the
lens of the humanities. The second task must be a similar study of
the way their creations are received by their respective audiences.
Neither art nor science can exist in a solipsistic vacuum of their
authors. Both must be listened to, observed, received, responded-
to. If the current public narratives of creativity are artificially divided
into the imagination of art and the logic of science, then the
framing of their reception is just as polarized. ‘Science is not with
us an object of contemplation’, complained social thinker Jacques
Barzun23. The impression is that art appeals to the response of
emotion and affect while science connects only to cerebral ­reason.
Such a neat, Kantian, division appeals to a compartmentalized
22 Iris Murdoch (1998), Existentialists and Mystics. New York: Penguin, p. 86.
23 J. Barzun (1964), Science that Glorious Entertainment. London: Harper and Row.
14 The Poetry and Music of Science

and fragmented structural view of culture, but it reinforces the


picture of artificial division of science and art into two realms.
Perhaps this is where they do indeed divide—even if both draw at
depth on a mysterious creative human energy in their produc-
tion, it is conceivable that reading, listening, and hearing science
and art nonetheless pursue distinct mental pathways.
On the other hand, if a distorted impression of creativity arises in
part from selective silences on the part of their practitioners, then
perhaps the same is true of their reception. Comment on the effect
and the enjoyment of art is commonplace. It speaks of a healthy
continuum from artist and performer to receiver and listener. We
may not be able to paint or to sing like the great e­ xponents of art
and oratorio, but we are not silenced as a result from speaking, or
even from critically appraising, paintings or ­performances. There is
understood to be a ‘ladder’ of participation and reception in the
arts. In music, for example, the lower rungs are occupied by those
of us who enjoy concerts, who pick up instruments in the com-
pany of forgiving amateur friends. We would never presume to
perform in public, but nevertheless can confidently express an
opinion on which recording of a s­ ymphony we prefer. The upper
rungs are occupied by the performers on those very recordings.
Here is actor Simon Russell Beale24 talking about his response to
the music of the romantic composer Schubert:
Schubert can make time stand still. In the last, miraculous months of his life, he
expanded his vision of what music could do. His most experimental work is the slow
movement of his B flat Piano Sonata. It is as if he has distilled the process of music-
making. He takes a harmonic progression, explores it, changes a single note,
explores it again; he breaks down a simple melody until only the bones are left and
the music is suspended. The result is a play of pure sound, without external refer-
ence, that gives us a glimpse of eternity.

This is a piece of exceptionally high-quality comment, to be


sure, but it is not unusual for someone who is not a professional

24 Simon Russell Beale (2012), Ferocious, Tender, Sublime, The Guardian, 19th
March.
Introduction 15

musician to write at such serious critical depth about a piece of


music. Beale’s testimony is also an example of the narrative we met
in the case of artistic creators, stories that map rather remarkably
onto the rare but honest stories of science’s own creative process.
Schubert, is, as heard by Beale, experimental, exploratory, even
reductionist and abstractionist. Yet he is also sublime. Beale
describes an example of a deep property of music—its ability to
reconcile us to the passage and structure of time by somehow
suspending us from it. At the level of the abstract, music and
mathematical science have long been compared, albeit inconclu-
sively. Developing a comparison of their creative processes will be
the concern of Chapter 5, but at this point the divergence in their
apparent degree of communicability is stark.
It is harder to find comparable examples of reception and affect
of scientific creation. But this is not because of a lack of inherent
appeal to human desire and need. The ‘ladder of access’ that we
identified in a creative art such as music is not (as observed by
Barzun in different terms) present in our current culture in
­science. This was not always the case—Shelley, Coleridge, and
Wordsworth all thought that science could, and would inspire
poetry (though Shelley foresaw that the inspirational beauty of
science would be a hidden one). So, for articulated contemporary
reception of science, we must usually listen to the scientists
themselves. Here cosmologist Subrahmanyan Chandrasekhar25
describes in remarkable terms an example of the moments of
transport for which science longs:
In my entire scientific life, extending over forty-five years, the most shattering
experience has been the realization that an exact solution of Einstein’s equations of
general relativity, discovered by New Zealand mathematician Roy Kerr, provides
the absolutely exact representation of untold numbers of massive black holes that
populate the universe. This ‘shuddering before the beautiful’, this incredible fact
that a discovery motivated by a search after the beautiful in mathematics should

25 Subrahmanyan Chandrasekhar (1987), Truth and Beauty: Aesthetics and


Motivations in Science. Chicago: University of Chicago Press.
16 The Poetry and Music of Science

find its exact replica in Nature, persuades me to say that beauty is that to which the
human mind responds at its deepest and most profound.

Chandrasekhar’s ‘shuddering before the beautiful’ carries unmis-


takable resonance with Beale’s ‘glimpse of eternity’. It also, perhaps
unwittingly, resonates with accounts of aesthetics that have been
driven beyond the simply beautiful to the ‘sublime’. The cosmolo-
gist is speaking of the extraordinarily simple yet utterly strange idea
of a ‘black hole’. For many years pure conjecture, observational evi-
dence from stellar evolution and highly luminous galactic cores
has pointed increasingly to the inevitable existence of these bizarre
and terrible objects. Black holes are places in the cosmos where
the local presence of matter is so great that gravity generates its
runaway collapse towards a point where density becomes formally
infinite, surrounded by a finite region of space in which the tug of
gravity is so great that no light can escape. Possessing a terrifying
and austere beauty, these objects are as near to instantiated math-
ematics as one could imagine. They can have no other properties
than mass, spin and electric charge. All other attributes that their
original matter once possessed are lost in its irreversible infall. The
normal role of ­mathematics within theoretical physics is to provide
approximate descriptions of natural objects, but in this case the
attribution of a black hole’s triplet of properties is complete.
The experience Chandrasekhar describes is a rarefied and extreme
form of a precious wonder. Einstein put it thus: ‘the most
­inexplicable thing about the universe is that it is explicable’ and
Eugene Wigner pointed towards it in the title of his celebrated
essay The Unreasonable Effectiveness of Mathematics in the Natural Sciences26.
The moment of connection of a constructed pattern of thought,
mathematical, pictorial, or logical, with the deep structure of the
natural world evokes an unparalleled experience of wonder. As
Wigner pointed out, it also appeals through the inherent elegance
of mathematics to the aesthetic sense. More than that, it seems to

26 Eugene Wigner (1960), Communications in Pure and Applied Mathematics, 13, No.
I. New York: John Wiley & Sons, Inc.
Introduction 17

satisfy a need for creative connectedness, the act of understand-


ing, of re-creating an internalized world patterned on the exter-
nal. Such reaching out into the world in abstract thought is
perhaps a flowering of a human response to the ancient questions
of the Book of Job27. Such long perspective suggests that there are
deep and complex commonalities in both the motivation and
imaginative act of creation, as well as in the role of aesthetics, in
the sciences as much as in the arts.
All this is not to claim that there is no tradition of research and
writing on the topic of creativity itself that includes scientific
examples—far from it. A now-celebrated 1926 essay of Graham
Wallas, The Art of Thought28 introduced a four-stage abstract struc-
ture for ‘the creative process’ that we will make critical use of
in the following chapters (finding it serviceable but wanting).
Wallas himself draws deliberately on a genre of The Art of . . . works,
initiated in modern times (it is an echo of Horace’s ancient Ars
Poetica) by writer Henry James in The Art of the Novel and taken up
by William Beveridge in the early twentieth century in his Art of
Scientific Investigation. We will be examining both of these works at
close quarters in Chapter 4. Major edited collections of essays and
research from psychologists29 on creativity have recently spawned
a subfield of cognitive neuroscience. The genre has even pro-
duced a ‘Cambridge Handbook’30. The laboratory-scale essays
into the close-study of the creative process and its context are
informative, although by their nature cannot encompass the
extended and unconstrained creative projects that characterize
substantial works of art and science. At a more selective level,
there are surveys of creative individuals, including those repre-
senting fields of science and art—Howard Gardner’s Creating

27 Tom McLeish, Faith and Wisdom.


28 Graham Wallas (1926), The Art of Thought. Tunbridge Wells: Solis Press (2016).
29 e.g. Robert J. Steinberg, ed. (1988), The Nature of Creativity. Cambridge:
Cambridge University Press.
30 James C. Kaufman and Robert J. Steinberg, eds. (2010), The Cambridge
Handbook of Creativity. Cambridge: Cambridge University Press.
18 The Poetry and Music of Science

Minds31 is an excellent example. Anthony Storr has begun to tease


out conjectural psychological patterns in the minds of very
­creative thinkers32. Other contributions to the creativity litera-
ture are listed for convenience in the bibliography of the present
work. But there is still a clear issue of misunderstanding in the
perception of science within popular culture, as an activity that
calls on minimal imagination. There is little discussion of the way
that imagination plays out in the experience of the thousands of
people engaged in the scientific and artistic work that adds colour
to our communities and national lives. There is also almost total
silence within the educational formation of scientists on the topic
of imagination, of the creative formulation of questions and
hypotheses, or of the experience of scientific ideation. There may
not be a method for this most vital of all scientific processes, but
there are accounts, practices, and a communal experience that
ought to be more widely and openly shared both within and
without the scientific community. Furthermore, as the ­explorations
in this book of visual, textual, and abstract creativity in science
and arts will show, the entanglements of experience between
their artistic and scientific examples are so close that a renewed
interdisciplinary dialogue promises a fruitful dividend for both
cultural communities.
I have suggested elsewhere33 that, because of the ‘missing
rungs’ in scientific ladder of reception, it is lamentably less com-
mon for non-practitioners of science to experience the intensity
of aesthetic response to a new understanding of nature, than for
the scientists whose professional training has taken them to the
ladder’s higher footholds that still exist. But it is not impossible,
and could be as common as the learning of a new tune or appre-
ciating an unfamiliar painting for the first time. In a moving per-
sonal example, a friend told me of the moment when, gazing up
at the moon one evening, he suddenly understood how its phases
31 Howard Gardner (1993), Creating Minds. New York: Basic Books.
32 Anthony Storr (1991), The Dynamics of Creation. London: Penguin Books.
33 Tom McLeish, Faith and Wisdom.
Introduction 19

worked. A life of familiarization with the monthly cycle of cres-


cent, half, full, and gibbous moon was not equivalent to ‘seeing’
how these shapes served as the signature of an illuminated orb.
On that moonlit night shortly after sunset he allowed the two-
dimensional screen of the sky to become a three-dimensional
space in his mind. The moon became a solid sphere, illuminated
by a much more distant sun from different angles on different
days, as seen from the centre of its orbit on the Earth. The celes-
tial geometry and its circling dynamics found a home in his
­imagination—releasing an experience of pure joy. He described
feeling present to the world in a deeper sense than before, and
knowing that this stronger relationship was, once found, not
going to be lost.
The genre of ‘trade book’ science writing has recently pro-
duced examples of exquisite care and intelligent communication
that seem to be painstakingly replacing those removed rungs in
the public latter of science-engagement, giving their readers
some of the experience of personal revelation as my friend’s per-
sonal reconstruction of lunar astronomy. Nobel Prize-winning
physicist Frank Wilczek’s lovingly articulated story of symmetry
in art, philosophy, and physics authentically communicates a
physicist’s delight in the discovered layers of nature’s patterned
structure34. Former editor of the scientific journal Nature, and
now science-writer, Philip Ball, has demonstrated how m ­ olecules
make a story as human as it is material35. Television producer and
writer Simon Singh’s account of the three-century long search
for a proof of Pierre de Fermat’s famous ‘last theorem’ is a thing
of beauty, emotion and an opening to the common gaze of a
world of pure mathematics that engages without patronizing36.
Experiences of such reception in science or in art, achieve at
their most profound such an intensity of emotion and of felt
transformation, that they must draw our exploration to a third
34 Frank Wilczek (2015), A Beautiful Question. London: Allen Lane.
35 Philip Ball (2001), Stories of the Invisible: A Guided Tour of Molecules. Oxford: OUP.
36 Simon Singh (1997), Fermat’s Last Theorem. London: Forth Estate Ltd.
20 The Poetry and Music of Science

level of parallel comparison—that of the human function of


­creative engagement with nature and, if we dare talk of it, of pur-
pose. A nest of questions confronts us here: why do art, and early
science, arise in pre-history? What do they achieve socially and
psychologically today? Where do art and science appear, both
explicitly and hidden, in the complex of cultural narratives? How
do they receive, and provide, value and virtue? The humanities
discipline of theology comes to aid here, for no other reason than
that it is comfortable with the category and narrative of purpose.
Recent writers have attempted to articulate a ‘theology of’ music
(Begbie37), of art (Wolterstorff38), of science (the present writer39),
and found that this trailhead leads to a fruitful landscape within
which these questions of function can be attempted. Critic and
literary scholar George Steiner’s view of art in this regard elicits a
striking reception when read from the perspective of science. In
his moving critique of the humanities in late modernism Real
Presences, he writes40
Only art can go some way towards making accessible, towards waking into some
measure of communicability, the sheer inhuman otherness of matter
speaking at the same time of deep need, and of powerful satisfac-
tion. Writing in despair over a ‘broken contract’ of art with nature
that he sees as vital to reverse, Steiner describes a human condi-
tion ‘out of joint’ with the world in which humans are immersed.
There is a gulf of otherness, of strangeness, with which, for
­reasons we do not understand, we remain uncomfortable. The
divorce is painful. The paradox is heightened when we reflect
that the ‘sheer inhuman otherness of matter’ is the very stuff of
which our bodies are composed. Art may indeed go some way
towards a reconciliation of the human with the material. The
37 Jeremy Begbie (2014), Theology, Music and Time. Cambridge: Cambridge
University Press.
38 Nicholas Wolterstorff (1987), Art in Action Toward a Christian Aesthetic. Grand
Rapids: Eerdmans.
39 Tom McLeish, Faith and Wisdom.
40 George Steiner (1989), Real Presences. London: Faber and Faber.
Another random document with
no related content on Scribd:
As to Porphyry it was doubtless his more practical temperament that
led him to modify the doctrine of Plotinus concerning ecstasy. With
Porphyry the mind does not lose, in that state of exaltation, its
consciousness of personality. He calls it a dream in which the soul,
dead to the world, rises to an activity that partakes of the divine. It is
an elevation above reason, above action, above liberty, and yet no
annihilation, but an ennobling restoration or transformation of the
individual nature.[27]
Gower. One of Porphyry’s notions about the spirits of the air, of
which you told me in our walk yesterday, quite haunted me
afterwards. It contains a germ of poetry.
Kate. By all means let us have it.
Gower. Our philosopher believed in a certain order of evil genii who
took pleasure in hunting wild beasts,—dæmons, whom men
worshipped by the title of Artemis and other names, falsely
attributing their cruelty to the calm and guiltless gods, who can never
delight in blood. Some of these natures hunted another prey. They
were said to chase souls that had escaped from the fetters of a body,
and to force them to re-enter some fleshly prison once more. How I
wish we could see a design of this by David Scott! Imagine the soul
that has just leaped out of the door of that dungeon of ignorance and
pain, the body, as Porphyry would term it, fluttering in its new
freedom in the sunshine among the tree-tops, over wild and town—
all the fields of air its pleasure-ground for an exulting career on its
upward way to join the journeying intelligences in their cars above.
But it sees afar off, high in mid-air, a troop of dark shapes; they seem
to approach, to grow out of the airy recesses of the distance—they
come down the white precipices of the piled clouds, over the long
slant of some vapour promontory—forms invisible to man, and, with
them, spectre-hounds, whose baying spirits alone can hear. As they
approach, the soul recognises its enemies. In a moment it is flying
away, away, and after it they sweep—pursuers and pursued, shapes
so ethereal that the galleries of the ant are not shaken as hunters
and quarry glide into the earth, and not a foam-bell is broken or
brushed from the wave when they emerge upon the sea, and with
many a winding and double mount the air. At last hemmed in, the
soul is forced—spite of that desperate sidelong dart which had all but
eluded them—down into a body, the frame of a beggar’s babe or of a
slave’s; and, like some struggling bird, drawn with beating wings
beneath the water, it sinks into the clay it must animate through
many a miserable year to come.
Willoughby. I wish you would paint it for us yourself. You might
represent, close by that battle of the spirits, a bird singing on a
bough, a labourer looking down, with his foot upon his spade, and
peasants dancing in their ‘sunburnt mirth’ and jollity—wholly
unconscious, interrupted neither in toil nor pleasure by the conflict
close at hand. It might read as a satire on the too common
indifference of men to the spiritual realities which are about them
every hour.
Mrs. Atherton. The picture would be as mysterious as an Emblem
by Albert Durer.
Gower. It is that suggestiveness I so admire in the Germans. For
the sake of it I can often pardon their fantastic extravagances, their
incongruous combinations, their frequent want of grace and
symmetry.
Atherton. So can I, when an author occupies a province in which
such indirectness or irony, such irregularity, confusion, or paradox,
are admissible. Take, as a comprehensive example, Jean Paul. But
in philosophy it is abominable. There, where transparent order
should preside, to find that under the thick and spreading verbiage
meaning is often lacking, and, with all the boastful and fire-new
nomenclature, if found, is old and common,—that the language is
commonly but an array of what one calls

Rich windows that exclude the light,


And passages that lead to nothing;—

This puts me out of all patience.


Gower. The fault you object to reminds me of some Flemish
landscape-pieces I have seen; there are trees, so full of grand life,
they seem with their outstretched arms to menace the clouds, and as
though, if they smote with their many hundred hands, they could
beat away the storm instead of being bowed by it; and underneath
these great ones of the forest, which should shadow nothing less
than a woodland council of Titans or a group of recumbent gods, the
painter places only a rustic with a cow or two, an old horse, a
beggar, or some other most every-day of figures.
Mrs. Atherton. And you mean that the German words are large-
looking as the trees, and the ideas worn and ordinary as the figures?
What will Mr. Willoughby say to that?
Atherton. I think Willoughby will agree with me that it is high time
that we should go back to our theurgic mysticism and Iamblichus.
Here is a letter of his:—

Iamblichus to Agathocles.
I assure you, my friend, that the efforts of Porphyry, of whom you
appear disposed to think so highly, will be altogether in vain. He
is not the true philosopher you imagine. He grows cold and
sceptical with years. He shrinks with a timid incredulity from
reaping in that field of supernatural attainment which theurgy has
first opened, and now continually enlarges and enriches.
Theurgy, be sure of it, is the grand, I may say, the sole path to
the exaltation we covet. It is the heaven-given organum, in the
hands of the wise and holy, for obtaining happiness, knowledge,
power.
The pomp of emperors becomes as nothing in comparison with
the glory that surrounds the hierophant. The priest is a prophet
full of deity. The subordinate powers of the upper world are at his
bidding, for it is not a man, but a god who speaks the words of
power. Such a man lives no longer the life common to other men.
He has exchanged the human life for the divine. His nature is the
instrument and vehicle of Deity, who fills and impels him
(ὄργανον τοῖς ἐπιπνέουσι θεοῖς.) Men of this order do not
employ, in the elevation they experience, the waking senses as
do others (οὔτε κατ᾽ αἴσθησιν ἐνεργοῦσιν οὔτε ἐγρηγόρασι).
They have no purpose of their own, no mastery over themselves.
They speak wisdom they do not understand, and their faculties,
absorbed in a divine power, become the utterance of a superior
will.
Often, at the moment of inspiration, or when the afflatus has
subsided, a fiery Appearance is seen,—the entering or departing
Power. Those who are skilled in this wisdom can tell by the
character of this glory the rank of the divinity who has seized for
the time the reins of the mystic’s soul, and guides it as he will.
Sometimes the body of the man subject to this influence is
violently agitated, sometimes it is rigid and motionless. In some
instances sweet music is heard, in others, discordant and fearful
sounds. The person of the subject has been known to dilate and
tower to a superhuman height; in other cases, it has been lifted
up into the air. Frequently, not merely the ordinary exercise of
reason, but sensation and animal life would appear to have been
suspended; and the subject of the afflatus has not felt the
application of fire, has been pierced with spits, cut with knives,
and been sensible of no pain. Yea, often, the more the body and
the mind have been alike enfeebled by vigil and by fasts, the
more ignorant or mentally imbecile a youth may be who is
brought under this influence, the more freely and unmixedly will
the divine power be made manifest. So clearly are these
wonders the work, not of human skill or wisdom, but of
supernatural agency! Characteristics such as these I have
mentioned, are the marks of the true inspiration.
Now, there are, O Agathocles, four great orders of spiritual
existence,—Gods, Dæmons, Heroes or Demi-gods, and Souls.
You will naturally be desirous to learn how the apparition of a
God or a Dæmon is distinguished from those of Angels,
Principalities, or Souls. Know, then, that their appearance to man
corresponds to their nature, and that they always manifest
themselves to those who invoke them in a manner consonant
with their rank in the hierarchy of spiritual natures. The
appearances of Gods are uniform (μονοειδῆ), those of Dæmons
various (ποικίλα). The Gods shine with a benign aspect. When a
God manifests himself, he frequently appears to hide sun or
moon, and seems as he descends too vast for earth to contain.
Archangels are at once awful and mild; Angels yet more
gracious; Dæmons terrible. Below the four leading classes I have
mentioned are placed the malignant Dæmons, the Anti-gods
(ἀντιθέους).
Each spiritual order has gifts of its own to bestow on the initiated
who evoke them. The Gods confer health of body, power and
purity of mind, and, in short, elevate and restore our natures to
their proper principles. Angels and Archangels have at their
command only subordinate bestowments. Dæmons, however,
are hostile to the aspirant,—afflict both body and mind, and
hinder our escape from the sensuous. Principalities, who govern
the sublunary elements, confer temporal advantages. Those of a
lower rank, who preside over matter (ὑλικά), display their bounty
in material gifts. Souls that are pure are, like Angels, salutary in
their influence. Their appearance encourages the soul in its
upward efforts. Heroes stimulate to great actions. All these
powers depend, in a descending chain, each species on that
immediately above it. Good Dæmons are seen surrounded by
the emblems of blessing, Dæmons who execute judgment
appear with the instruments of punishment.
There is nothing unworthy of belief in what you have been told
concerning the sacred sleep, and divination by dreams. I explain
it thus:—
The soul has a twofold life, a lower and a higher. In sleep that
soul is freed from the constraint of the body, and enters, as one
emancipated, on its divine life of intelligence. Then, as the noble
faculty which beholds the objects that truly are—the objects in
the world of intelligence—stirs within, and awakens to its power,
who can be surprised that the mind, which contains in itself the
principles of all that happens, should, in this its state of liberation,
discern the future in those antecedent principles which will make
that future what it is to be? The nobler part of the soul is thus
united by abstraction to higher natures, and becomes a
participant in the wisdom and foreknowledge of the Gods.
Recorded examples of this are numerous and well authenticated;
instances occur, too, every day. Numbers of sick, by sleeping in
the temple of Æsculapius, have had their cure revealed to them
in dreams vouchsafed by the god. Would not Alexander’s army
have perished but for a dream in which Dionysus pointed out the
means of safety? Was not the siege of Aphutis raised through a
dream sent by Jupiter Ammon to Lysander? The night-time of the
body is the day-time of the soul.
What I have now said—with little method, I confess—sets before
you but a portion of the prerogatives in which the initiated glory.
There is much behind for which words are too poor. I have
written enough, I am sure, to kindle your ambition, to bid you
banish doubt, and persevere in the aspirations which so
possessed you when I saw you last.[28] Farewell.

Gower. That explanation of prophetic dreams and the temple sleep


is very curious and characteristic. No doubt the common phenomena
of mesmerism may have been among the sacred secrets preserved
by the priests of Egypt and of Greece.
Kate. The preference for young and weakly persons, who would
possess an organization more susceptible of such influences, makes
it look very likely.
Atherton. Observe how completely the theurgic element, with
Iamblichus, supersedes the theosophic. In the process of time the
philosophical principles on which the system of Plotinus rested are
virtually surrendered, little by little, while divination and evocations
are practised with increasing credulity, and made the foundation of
the most arrogant pretensions. Plotinus declared the possibility of an
absolute identification of the divine with the human nature. Here was
the broadest basis for mysticism possible. Porphyry retired from this
position, took up narrower ground, and qualified the great mystical
principle of his master. He contended that in the union which takes
place in ecstasy, we still retain the consciousness of personality.
Iamblichus, the most superstitious of all in practice, diminished the
real principle of mysticism still farther in theory. He denied that man
has a faculty inaccessible to passion, and eternally active.[29]
Willoughby. And so the metaphysics and the marvels of mysticism
stand in an inverse ratio to each other. But it is not unnatural that as
the mystic, from one cause or another, gives up those exaggerated
notions of the powers of man and those mistaken views of the
relationship between man and God, which went together to make up
a mystical system of philosophy, he should endeavour to indemnify
himself by the evocations of theurgy, so as to secure, if possible,
through a supernatural channel, what speculation had
unsuccessfully attempted.
Atherton. True; but in this case I should invert the order, and say
that as the promise of theurgy exercised an attraction of growing
strength on an order of mind less fitted for speculation, such
temperaments would readily drop the speculative principle of
mysticism in their eagerness to grasp the illusive prize—apparently
so practical—which a commerce with superior natures held out.
Willoughby. And so the intellectual ambition and the poetical spirit,
so lofty in Plotinus, subside, among the followers of Iamblichus, into
the doggrel of the necromancer’s charm.
Gower. Much such a descent as the glory of Virgil has suffered,
whose tomb at Pausilipo is now regarded by the populace of
degenerate Naples less with the reverence due to the poet than with
the awe which arises from the legendary repute of the mediæval
magician.
Atherton. So the idealism of strong minds becomes superstition in
the weak. In the very shrine where culture paid its homage to art or
science, feebleness and ignorance, in an age of decline, set up the
image-worship of the merely marvellous.
Mrs. Atherton. I think you mentioned only one other of these
worthies.
Atherton. Proclus. He is the last great name among the Neo-
Platonists. He was the most eclectic of them all, perhaps because
the most learned and the most systematic. He elaborated the trinity
of Plotinus into a succession of impalpable Triads, and surpassed
Iamblichus in his devotion to the practice of theurgy. Proclus was
content to develop the school in that direction which Iamblichus—
(successful from his very faults)—had already given it. With Proclus,
theurgy was the art which gives man the magical passwords that
carry him through barrier after barrier, dividing species from species
of the upper existences, till, at the summit of the hierarchy, he arrives
at the highest. According to him, God is the Non-Being who is above
all being. He is apprehended only by negation. When we are raised
out of our weakness, and on a level with God, it seems as though
reason were silenced, for then we are above reason. We become
intoxicated with God, we are inspired as by the nectar of Olympus.
He teaches philosophy as the best preparation for Quietism. For the
scientific enquirer, toiling in his research, Proclus has a God to tell of,
supreme, almighty, the world-maker and governor of Plato. For him
who has passed through this labour, a God known only by ecstasy—
a God who is the repose he gives—a God of whom the more you
deny the more do you affirm.
Willoughby. And this is all! After years of austerity and toil, Proclus
—the scholar, stored with the opinions of the past, surrounded by the
admiration of the present—the astronomer, the geometrician, the
philosopher,—learned in the lore of symbols and of oracles, in the
rapt utterances of Orpheus and of Zoroaster—an adept in the ritual
of invocations among every people in the world—he, at the close,
pronounces Quietism the consummation of the whole, and an
unreasoning contemplation, an ecstasy which casts off as an
incumbrance all the knowledge so painfully acquired, the bourne of
all the journey.
Mrs. Atherton. As though it were the highest glory of man,
forgetting all that his enquiry has achieved, hidden away from the
world,—to gaze at vacancy, inactive and infantine;—to be like some
peasant’s child left in its cradle for a while in the furrow of a field,
shut in by the little mound of earth on either side, and having but the
blue æther above, dazzling and void, at which to look up with smiles
of witless wonder.
Note to page 103.

Iamblichus, de Mysteriis, sect. x. cc. 1, 4, 6; iii. 4, 8, 6, 24; i. 5, 6; ii.


3; iii. 31; ii. 4, 6, 7; iii. 1, 3. These passages, in the order given, will
be found to correspond with the opinions expressed in the letter as
those of Iamblichus.
The genuineness of the treatise De Mysteriis has been called in
question, but its antiquity is undoubted. It differs only in one or two
very trivial statements from the doctrines of Iamblichus as
ascertained from other sources, and is admitted by all to be the
production, if not of Iamblichus himself, of one of his disciples,
probably writing under his direction. Jules Simon, ii. 219.
For the opinions ascribed to Porphyry in this letter, see his Epistola
ad Anebonem, passim. He there proposes a series of difficult
questions, and displays that sceptical disposition, especially
concerning the pretensions of Theurgy, which so much scandalized
Iamblichus. The De Mysteriis is an elaborate reply to that epistle,
under the name of Abammon.
In several passages of the De Mysteriis (ii. 11; v. 1, 2, 3, 7; vi. 6)
Iamblichus displays much anxiety lest his zeal for Theurgy should
lead him to maintain any position inconsistent with the reverence due
to the gods. He was closely pressed on this weak point by the
objections of Porphyry. (Ep. ad Anebon. 5, 6.) His explanation in
reply is, that the deities are not in reality drawn down by the mere
human will of the Theurgist, but that man is raised to a participation
in the power of the gods. The approximation is real, but the apparent
descent of divinity is in fact the ascent of humanity. By his long
course of preparation, by his knowledge of rites and symbols, of
potent hymns, and of the mysterious virtues of certain herbs and
minerals, the Theurgist is supposed to rise at last to the rank of an
associate with celestial powers; their knowledge and their will
become his, and he controls inferior natures with the authority of the
gods themselves.
Iamblichus supposes, moreover, that there is an order of powers in
the world, irrational and undiscerning, who are altogether at the
bidding of man when by threats or conjurations he chooses to
compel them. De Myst. vi. 5.
BOOK THE FOURTH
MYSTICISM IN THE GREEK CHURCH
CHAPTER I.

Questi ordini di su tutti s’ammirano


E di giù vincon si che verso Iddio
Tutti tirati sono e tutti tirano.
E Dionisio con tanto disio
A contemplar questi ordini si mise,
Che li nomò e distinse com’ io.[30]

Dante.

Kate. I have been looking at the pictures in Mrs. Jameson’s Sacred


and Legendary Art, of those strange creatures, the hermit saints—
the Fathers of the desert. Only see this one, what a mane and claws!
The two lions digging the grave there are own brothers to the holy
men themselves.
Atherton. Yet they claimed powers as much above humanity as, to
look at them, you would think them beneath it.
Gower. Religious Nebuchadnezzars.
Willoughby. No shavelings, at any rate, like the smooth-faced
sanctities of the later calendar.
Atherton. You will find among these anchorites almost all the
wonder-working pretensions of mediæval mysticism in full
development, thus early;—the discernment of spirits, gift of
prophecy, miraculous powers of various kinds, ecstasy, exorcism,
&c. &c. I should take St. Antony as a fair specimen of the whole
class.[31]
Mrs. Atherton. Look, here is his picture; there he stands, with
crutch and bell and pig.
Atherton. The bell denotes his power over evil spirits, and the pig
the vanquished dæmon of sensuality. In his life, by Athanasius, there
is a full account of his battle with many dæmons in the shape of
lions, bulls, and bears. He passed twenty years in an old castle
which he found full of serpents. The power of the saint expelled
those unpleasant aborigines. That nose, you see there, was
supposed to possess the faculty of detecting by its miraculous
keenness of scent the proximity of an evil spirit. There is an odour of
iniquity, you must know, as well as an odour of sanctity. This
disposition to literalize metaphors gave currency to the monkish
stories of after times concerning the refreshing fragrance found to
arise from the remains of disinterred saints. In fact, the
materialization of the spiritual, or what passes for such, is the
characteristic principle of the theurgic mysticism within the Roman
Catholic Church. St. Antony, on one occasion, sees his own soul,
separated from the body, carried through the air.
Gower. A striking instance, I should say, of the objectivity of the
subject.
Atherton. One of his visions is not without grandeur. The brethren
had been questioning him one day concerning the state of departed
spirits. The following night he heard a voice saying, ‘Antony, get up;
go out and look!’ He obeyed, and saw a gigantic figure, whose head
was in the clouds, and whose outstretched arms extended far across
the sky. Many souls were fluttering in the air, and endeavouring, as
they found opportunity, to fly upward past this dreadful being.
Numbers of them he seized in the attempt, and dashed back upon
the earth. Some escaped him and exulted above, while he raged at
their success. Thus sorrowing and rejoicing were mingled together,
as some were defeated and others triumphant. This, he was given to
understand, was the rise and fall of souls.
Willoughby. That picture would be really Dantesque, if only a little
more definite. Macarius is another great name, too, among these
Christian ascetics and theurgists—the one who retired to the deserts
of Nitria in the fourth century.
Atherton. He is not only famous for his measure of the
supernatural powers ascribed to his brethren, but his homilies have
been appealed to by modern theopathetic mystics as an authority for
Quietism. He teaches perfectionist doctrine, certainly, but I do not
think his words will bear the construction Poiret and others would
give them. He was at least innocent of the sainte indifférence.[32]
Mrs. Atherton. You said we were to discuss Dionysius the
Areopagite this evening.
Kate. Pray introduce me first. I know nothing about him.
Atherton. No one does know who really wrote the books which
passed under that name. It is generally admitted that the forgery
could not have been committed earlier than the middle of the fifth
century, probably somewhat later. So all I can tell you is, that
somewhere or other (it is not unlikely at Constantinople, but there is
no certainty), about the time when Theodoric was master of Italy—
when the Vandal swarms had not yet been expelled from northern
Africa—while Constantinople was in uproar between the greens and
the blues, and rival ecclesiastics headed city riots with a rabble of
monks, artizans, and bandit soldiery at their heels—while orthodoxy
was grappling with the Monophysite and Eutychian heresies on
either hand, and the religious world was rocking still with the
groundswell that followed those stormy synods in which Palestine
and Alexandria, Asia and Constantinople, from opposite quarters,
gathered their strength against each other—a monk or priest was
busy, in his quiet solitude, with the fabrication of sundry treatises and
letters which were to find their way into the Church under the all-but
apostolic auspices of that convert made by the Apostle of the
Gentiles when he spoke on Mars Hill. The writings would seem to
have been first appealed to as genuine in the year 533. As heretics
cited them, their authority was disputed at the outset; but being
found favourable to the growing claims of the hierarchy, and likely to
be useful, they were soon recognised and employed accordingly.[33]
Willoughby. Proclus could not have been long dead, and his
reputation must have been still at its height, when this anonymous—
let us call him Dionysius at once—was writing his Platonized
theology.
Atherton. With the divines of Byzantium Proclus represented the
grand old world of Greek thought. Even those who wrote against him
as a heathen betray the influence he exercised on their doctrines.
The object of Dionysius evidently was to accommodate the
theosophy of Proclus to Christianity. Another aim, not less
conspicuous, was to strengthen all the pretensions of the priesthood,
and to invest with a new traditionary sanction the ascetic virtues of
the cloister.
CHAPTER II.

They that pretend to these heights call them the secrets of the
kingdom; but they are such which no man can describe; such
which God hath not revealed in the publication of the Gospel;
such for the acquiring of which there are no means prescribed,
and to which no man is obliged, and which are not in any man’s
power to obtain; nor such which it is lawful to pray for or desire;
nor concerning which we shall ever be called to account.—
Jeremy Taylor.

‘I have here,’ said Atherton on the next evening, ‘some notes on the
doctrine of this pretended Areopagite—a short summary; shall I read
it?’
‘By all means.’
So the following abstract was listened to—and with creditable
patience.[34]
(1.) All things have emanated from God, and the end of all is return
to God. Such return—deification, he calls it—is the consummation of
the creature, that God may finally be all in all. A process of evolution,
a centrifugal movement in the Divine Nature, is substituted in reality
for creation. The antithesis of this is the centripetal process, or
movement of involution, which draws all existence towards the point
of the Divine centre. The degree of real existence possessed by any
being is the amount of God in that being—for God is the existence in
all things. Yet He himself cannot be said to exist, for he is above
existence. The more or less of God which the various creatures
possess is determined by the proximity of their order to the centre.
(2.) The chain of being in the upper and invisible world, through
which the Divine Power diffuses itself in successive gradations, he
calls the Celestial Hierarchy. The Ecclesiastical Hierarchy is a
corresponding series in the visible world. The orders of Angelic
natures and of priestly functionaries correspond to each other. The
highest rank of the former receive illumination immediately from God.
The lowest of the heavenly imparts divine light to the highest of the
earthly hierarchy. Each order strives perpetually to approximate to
that immediately above itself, from which it receives the transmitted
influence; so that all, as Dante describes it, draw and are drawn, and
tend in common towards the centre—God.
The three triads of angelic existences, to whom answer the ranks of
the terrestrial hierarchy, betrays the influence of Proclus, whose
hierarchy of ideas corresponds, in a similar manner, to his hierarchy
of hypostases.
Gower. The system reminds one of those old pictures which are
divided into two compartments, the upper occupied by angels and
cherubs on the clouds, and the lower by human beings on the earth,
gazing devoutly upward at their celestial benefactors.
Atherton. The work of Christ is thrown into the background to make
room for the Church. The Saviour answers, with Dionysius, rather to
the Logos of the Platonist than to the Son of God revealed in
Scripture. He is allowed to be, as incarnate, the founder of the
Ecclesiastical Hierarchy; but, as such, he is removed from men by
the long chain of priestly orders, and is less the Redeemer, than
remotely the Illuminator, of the species.
Purification, illumination, perfection,—the three great stages of
ascent to God (which plays so important a part in almost every
succeeding attempt to systematise mysticism) are mystically
represented by the three sacraments,—Baptism, the Eucharist, and
Unction. The Church is the great Mystagogue: its liturgy and offices
a profound and elaborate system of symbolism.
(3.) The Greek theory, with its inadequate conception of the nature of
sin, compels Dionysius virtually to deny the existence of evil.
Everything that exists is good, the more existence the more
goodness, so that evil is a coming short of existence. He hunts sin
boldly from place to place throughout the universe, and drives it at
last into the obscurity of the limbo he contrives for it, where it lies
among things unreal.
All that exists he regards as a symbolical manifestation of the super-
existent. What we call creation is the divine allegory. In nature, in
Scripture, in tradition, God is revealed only in figure. This sacred
imagery should be studied, but in such study we are still far from any
adequate cognizance of the Divine Nature. God is above all negation
and affirmation: in Him such contraries are at once identified and
transcended. But by negation we approach most nearly to a true
apprehension of what He is.
Negation and affirmation, accordingly, constitute the two opposed
and yet simultaneous methods he lays down for the knowledge of
the Infinite. These two paths, the Via Negativa (or Apophatica) and
the Via Affirmativa (or Cataphatica) constitute the foundation of his
mysticism. They are distinguished and elaborated in every part of his
writings. The positive is the descending process. In the path
downward from God, through inferior existences, the Divine Being
may be said to have many names;—the negative method is one of
ascent; in that, God is regarded as nameless, the inscrutable
Anonymous. The symbolical or visible is thus opposed, in the
Platonist style, to the mystical or ideal. To assert anything concerning
a God who is above all affirmation is to speak in figure, to veil him.
The more you deny concerning Him, the more of such veils do you
remove. He compares the negative method of speaking concerning
the Supreme to the operation of the sculptor, who strikes off
fragment after fragment of the marble, and progresses by diminution.
(4.) Our highest knowledge of God, therefore, is said to consist in
mystic ignorance. In omni-nescience we approach Omniscience.
This Path of Negation is the highway of mysticism. It is by refraining
from any exercise of the intellect or of the imagination—by self-
simplification, by withdrawal into the inmost, the divine essence of
our nature—that we surpass the ordinary condition of humanity, and
are united in ecstasy with God. Dionysius does not insist so much on
Union as the later mystics, but he believes, at all events, that the
eminent saint may attain on earth an indescribable condition of soul
—an elevation far transcending the reach of our natural faculties—an
approach towards the beatific vision of those who are supposed to
gaze directly on the Divine Essence in heaven. His disciple is
perpetually exhorted to aspire to this climax of abstraction—above
sight, and thought, and feeling, as to the highest aim of man.
Willoughby. What contradictions are here! With one breath he
extols ineffable ignorance as the only wisdom; with the next he
pretends to elucidate the Trinity, and reads you off a muster-roll of
the heavenly hierarchies.
Gower. And are not these, supplemented by the hierarchy of
ecclesiastics, his real objects of worship? No man could make an
actual God of that super-essential ultimatum, that blank Next-to-
Nothingness which the last Neo-Platonists imagined as their
Supreme. Proclus could not; Dionysius could not. What then? A
reaction comes, which, after refining polytheism to an impalpable
unity, restores men to polytheism once more. Up mounts
speculation, rocket-like: men watch it, a single soaring star with its
train of fire, and, at the height, it breaks into a scattering shower of
many-coloured sparks. From that Abstraction of which nothing can
be predicated, nothing can be expected. The figment above being is
above benignity. So the objects of invocation are gods, demi-gods,
dæmons, heroes; or, when baptized, cherubim, seraphim, thrones,
dominions, powers, archangels, angels, saints; in either case,
whether at Athens or at Constantinople, the excessive subtilisation of
the One contributes toward the worship of the Manifold.
Atherton. The theology of the Neo-Platonists was always in the first
instance a mere matter of logic. It so happened that they confounded
Universals with causes. The miserable consequence is clear. The
Highest becomes with them, as he is with Dionysius, merely the
most comprehensive, the universal idea, which includes the world,
as genus includes species.[35]
Mrs. Atherton. The divinity of this old Father must be a bleak affair
indeed—Christianity frozen out.
Gower. I picture him to myself as entering with his philosophy into
the theological structure of that day, like Winter into the cathedral of
the woods (which an autumn of decline has begun to harm already);
—what life yet lingers, he takes away,—he untwines the garlands
from the pillars of the trees, extinguishes the many twinkling lights
the sunshine hung wavering in the foliage, silences all sounds of
singing, and fills the darkened aisles and dome with a coldly-
descending mist, whose silence is extolled as above the power of
utterance,—its blinding, chill obscureness lauded as clearer than the
intelligence and warmer than the fervour of a simple and scriptural
devotion.
Atherton. You have described my experience in reading him,
though I must say he suggested nothing to me about your cathedral
of the woods, &c. His verbose and turgid style, too, is destitute of all
genuine feeling.[36] He piles epithet on epithet, throws superlative on
superlative, hyperbole on hyperbole, and it is but log upon log,—he
puts no fire under, neither does any come from elsewhere. He
quotes Scripture—as might be expected—in the worst style, both of
the schoolman and the mystic. Fragments are torn from their
connexion, and carried away to suffer the most arbitrary
interpretation, and strew his pages that they may appear to illustrate
or justify his theory.
Gower. How forlorn do those texts of Scripture look that you discern
scattered over the works of such writers, so manifestly transported
from a region of vitality and warmth to an expanse of barrenness.
They make the context look still more sterile, and while they say
there must be life somewhere, seem to affirm, no less emphatically,
that it is not in the neighbourhood about them. They remind me of
those leaves from the chestnut and the birch I once observed upon a
glacier. There they lay, foreign manifestly to the treeless world in
which they were found; the ice appeared to have shrunk from them,
and they from the ice; each isolated leaf had made itself a cup-like
cavity, a tiny open sarcophagus of crystal, in which it had lain,
perhaps for several winters. Doubtless, a tempest, which had been
vexing some pleasant valley far down beneath, and tearing at its
trees, must have whirled them up thither. Yet the very presence of
the captives reproached the poverty of the Snow-King who detained
them, testifying as they did to a genial clime elsewhere, whose
products that ice-world could no more put forth, than can such frozen
speculations as this of Dionysius, the ripening ‘fruits of the Spirit.’

You might also like