Full Download Ether and Modernity: The Recalcitrance of An Epistemic Object in The Early Twentieth Century Jaume Navarro PDF

Download as pdf or txt
Download as pdf or txt
You are on page 1of 64

Full download test bank at ebook ebookmass.

com

Ether and Modernity: The


Recalcitrance of an Epistemic
Object in the Early Twentieth
Century Jaume Navarro
CLICK LINK TO DOWLOAD

https://ebookmass.com/product/ether-and-
modernity-the-recalcitrance-of-an-epistemic-
object-in-the-early-twentieth-century-jaume-
navarro/

ebookmass.com
More products digital (pdf, epub, mobi) instant
download maybe you interests ...

Gestural Imaginaries: Dance and Cultural Theory in the


Early Twentieth Century Lucia Ruprecht

https://ebookmass.com/product/gestural-imaginaries-dance-and-
cultural-theory-in-the-early-twentieth-century-lucia-ruprecht/

The Rise and Fall of OPEC in the Twentieth Century


Giuliano Garavini

https://ebookmass.com/product/the-rise-and-fall-of-opec-in-the-
twentieth-century-giuliano-garavini/

Mixed Race Britain in The Twentieth Century Chamion


Caballero And Peter J. Aspinall (Eds.)

https://ebookmass.com/product/mixed-race-britain-in-the-
twentieth-century-chamion-caballero-and-peter-j-aspinall-eds/

Healthy Minds In The Twentieth Century: In And Beyond


The Asylum Steven J. Taylor

https://ebookmass.com/product/healthy-minds-in-the-twentieth-
century-in-and-beyond-the-asylum-steven-j-taylor/
Ancient Art and Its Commerce in Early Twentieth-century
Europe: A Collection of Essays Written by the
Participants of the John Marshall Archive Project Guido
Petruccioli (Editor)
https://ebookmass.com/product/ancient-art-and-its-commerce-in-
early-twentieth-century-europe-a-collection-of-essays-written-by-
the-participants-of-the-john-marshall-archive-project-guido-
petruccioli-editor/

Out of the Ether Matthew Leising

https://ebookmass.com/product/out-of-the-ether-matthew-leising/

Consumption and Advertising in Eastern Europe and


Russia in the Twentieth Century Magdalena Eriksroed-
Burger

https://ebookmass.com/product/consumption-and-advertising-in-
eastern-europe-and-russia-in-the-twentieth-century-magdalena-
eriksroed-burger/

The Palgrave Handbook of Twentieth and Twenty-First


Century Literature and Science The Triangle Collective

https://ebookmass.com/product/the-palgrave-handbook-of-twentieth-
and-twenty-first-century-literature-and-science-the-triangle-
collective/

Museums, Modernity and Conflict: Museums and


Collections in and of War since the Nineteenth Century
Kate Hill

https://ebookmass.com/product/museums-modernity-and-conflict-
museums-and-collections-in-and-of-war-since-the-nineteenth-
century-kate-hill/
ETHER AND MODERNITY
ETHER AND
MODERNITY
The Recalcitrance of an Epistemic Object in
the Early Twentieth Century

Edited by
JAUME NAVARRO
University of the Basque Country, Spain

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
© Oxford University Press 2018
The moral rights of the authors have been asserted
First Edition published in 2018
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: 2018935614
ISBN 978–0–19–879725–8
DOI: 10.1093/oso/9780198797258.001.0001
Printed and bound by
CPI Group (UK) Ltd, Croydon, CR0 4YY

Links to third party websites are provided by Oxford in good faith and
for information only. Oxford disclaims any responsibility for the materials
contained in any third party website referenced in this work.
ACKNOWLEDGEMENTS

This project began in 2013 when I was awarded the Marconi Fellowship by the Bodleian
Libraries at Oxford University. Alexandra Franklin, from the Department of Special
Collections and Prof. Pietro Corsi encouraged me to organise a workshop on ‘The Lure
of the Ether. Physics and Modernity’, which eventually took place in February 2014 at
the History Faculty, Oxford. Imogen Clarke, Richard Noakes, Richard Staley, Michael
Whitworth and I presented papers that were commented on by Geoffrey Cantor. In
2015, Scott Walter and I organised a special panel on the same topic at the History of
Science Society meeting in San Francisco and we were joined by Connemara Doran,
Linda Henderson and Richard Staley. It was in that meeting that we agreed to work
towards more consistent versions of our papers and to publish this edited volume. The
final papers were presented at a meeting in Donostia/San Sebastian, in March 2017,
attended by all the contributors to this volume. I would like to thank the Donostia
International Physics Center (DIPC) and the Basque Government for funding that meet-
ing, as well as project HAR2015-67831-P MINECO/FEDER, EU, for funding part of the
research for this book.
CONTENTS

Biographies of Authors ix

1. Introduction: Ether—The Multiple Lives of a Resilient Concept


Massimiliano Badino and Jaume Navarro 1
2. The Ether at the Crossroads of Classical and Modern Physics
Imogen Clarke 14
3. Transformations of Knowledge in Oliver Lodge’s Ether and Reality
Michael H. Whitworth 30
4. Poincaré’s Mathematical Creations in Search of the ‘True
Relations of Things’
Connemara Doran 45
5. Ether and Electrons in Relativity Theory (1900–11)
Scott A. Walter 67
6. Making Space for the Soul: Oliver Lodge, Maxwellian Psychics
and the Etherial Body
Richard Noakes 88
7. Lenard’s Ether and Its Vortex of Emotions: Between
Accommodating and Fighting Modern Physics with Äther and
Uräther in the German Political Context
Arne Schirrmacher 107
8. Ether and Wireless: An Old Medium into New Media
Jaume Navarro 130

9. Hunting for the Luminiferous Ether: The American Revival


of the Michelson–Morley Experiment in the 1920s
Roberto Lalli 155
10. Ether and Aesthetics in the Dialogue between Relativists and
Their Critics in the Late Nineteenth and Early Twentieth Centuries
Richard Staley 179
viii Contents

11. Umberto Boccioni’s Elasticity, Italian Futurism and the Ether of Space
Linda Dalrymple Henderson 200
12. An Ether by Any Other Name? Paul Dirac’s Æther
Aaron Sidney Wright 225

Index 245
BIOGRAPHIES OF AUTHORS

Massimiliano Badino is assistant professor of history and philosophy of science at the


University of Verona. Previously, he was the Marie Sklodowska-Curie Fellow at the
Universitat Autònoma de Barcelona and at the Massachusetts Institute of Technology.
His research focuses on history and philosophy of mathematical physics from the
Enlightenment to the twentieth century.
Imogen Clarke is an independent scholar, working in academic publishing. She com-
pleted her PhD on early twentieth-century physics at the University of Manchester in
2012, and has published articles on the history of physics with a focus on peer review,
publishing practices and the relationship between the classical and the modern.
Connemara Doran (PhD, Harvard University) is a historian of science and technology,
whose research focuses on the history and philosophy of the mathematical and physical
sciences. She is currently a Postdoctoral Fellow at Harvard University, Department of the
History of Science. Her research encompasses two areas in which the ‘creative virtue’ of
physical and mathematical reasoning has been transformative: (1) the history of modern
cosmology and (2) the history of energy resources. Her chapter in this volume draws
from her book project on the intellectual adventure to empirically determine the size
and shape of the universe—to conceptualise, measure and map the cosmos—during the
long twentieth century. She refocuses discourse onto Henri Poincaré’s odyssey to over-
come the limitations of geometric empiricism by creating an utterly new mathematics of
spatial relations.
Linda Dalrymple Henderson is the David Bruton, Jr., Centennial Professor in Art
History and Distinguished Teaching Professor at the University of Texas at Austin.
Professor Henderson’s research and teaching focus on the interdisciplinary study of
modernism, including the relation of modern art to fields such as geometry, science
and technology, and mystical and occult philosophies. In addition to numerous periodical
articles and catalogue essays, she is the author of The Fourth Dimension and Non-
Euclidean Geometry in Modern Art (Princeton University Press, 1983; new ed., MIT
Press, 2013) and Duchamp in Context: Science and Technology in the Large Glass and
Related Works (Princeton, 1998). She also co-edited with Bruce Clarke the interdisciplinary
anthology From Energy to Information: Representation in Science, and Technology, Art
and Literature (Stanford University Press, 2002). In 2008 Henderson curated and
wrote the catalogue for the exhibition Reimagining Space: The Park Place Gallery in 1960s
New York (Blanton Museum of Art, The University of Texas, 2008). She is currently
working on two book projects, ‘The Energies of Modernism: Art, Science, and
Occultism in the Early 20th Century’ and ‘The Fourth Dimension in Art and Culture
Decade-by-Decade Through the 20th Century’.
x Biographies of Authors

Roberto Lalli is a Research Scholar at the Max Planck Institute for the History of
Science and a Visiting Scholar in the Research Program on the History of the Max Planck
Society. After having received an MSc degree in Physics, he earned a PhD in International
History at the University of Milan (2011). From 2011 to 2013, he was a postdoctoral fellow
in the Program on Science, Technology, and Society at the Massachusetts Institute of
Technology. He is a historian of modern physical sciences whose work focuses on the
interconnections between social and epistemic factors in the production and circulation
of novel products in theoretical physics and in the international standardisation of scien-
tific practices. He has published extensively on the history of relativity theories, on the
transfer of quantum theory, and on the evolution of editorial practices. His monograph
analyses the attempts to build an international community of general relativity experts
during the Cold War. His current project concerns the development of new methodolo-
gies based on the concepts and tools of the network theory in order to jointly analyse the
evolution of scientific knowledge in physics, the creation of transnational scientific com-
munities, and the developments of scientific institutions.
Jaume Navarro is an Ikerbasque Research Professor at the University of the Basque
Country. Trained in physics and in philosophy he has developed a career in the history of
physics in institutions such as University of Cambridge, Imperial College and the Max
Planck Institute for the History of Science. He is author of A History of the Electron: J. J.
and G. P. Thomson (Cambridge: 2012) and co-editor of Research and Pedagogy: A History of
Quantum Theory and its Early Textbooks (Berlin: 2015).
Richard Noakes is a Senior Lecturer in History at the University of Exeter. He has
published widely on the history of physical sciences, psychical research, telecommunica-
tions and the representation of the sciences in nineteenth-century periodicals. He is the
co-editor of From Newton to Hawking: A History of Cambridge University’s Lucasian Professors
of Mathematics (2003) and Culture and Science in the Nineteenth-Century Media (2004), and
co-author of Science in the Nineteenth-Century Periodical: Reading the Magazine of Nature
(2004). He is currently finishing a monograph, Physics and Psychics: The Occult and British
Sciences, 1850–1930.
Arne Schirrmacher studied natural sciences and philosophy at the universities of
Hamburg, Oxford and Munich. After a stay at the Max Planck Institute for the History
of Science in Berlin, he became a long-term research scholar at the Deutsches Museum
in Munich with projects on the history of physics, the history of twentieth century sci-
ence communication as well as on the relations between science and war technologies.
He was also an editor of the Hilbert Edition project at the University of Göttingen for
David Hilbert’s physics lectures, and a principal investigator in a research group on the
history of quantum physics at the Max Planck Institute. Since 2010 he has been a senior
research scholar at the Department of History of Humboldt Universität. In 2015 he
won a Heisenberg Fellowship of the German Research Foundation.
Richard Staley wrote his dissertation at the University of Cambridge on the early
work of Max Born in 1992, and then co-curated two museum exhibitions in the Whipple
Biographies of Authors xi

Museum of the History of Science, ‘Empires of Physics’ and ‘1900: The New Age’,
before holding postdocs and visiting positions in Melbourne, Berlin and Chicago. From
2000 to 2013, he taught in the History of Science Department at the University of
Wisconsin–Madison, publishing Einstein’s Generation and the Origins of the Relativity
Revolution with University of Chicago Press in 2008. He is now Rausing Lecturer in the
Department of History and Philosophy of Science at the University Cambridge, con-
ducting research on the histories of physics and anthropology, the cultural history of
mechanics, and environmental physics and climate change.
Scott A. Walter teaches history and philosophy of science and technology at the
François Viète Center for Epistemology and History of Science and Technology (EA
1161), University of Nantes. He serves as editor of the ‘Henri Poincaré Papers’ website,
and has published two volumes of Henri Poincaré’s scientific correspondence, devoted
to exchanges with physicists, chemists and engineers (2007), and with astronomers and
geodesists (2016). He has published research on the history of relativity theory in Studies
in the History and Philosophy of Modern Physics, and he contributed a chapter on the his-
torical origins of spacetime to the Springer Handbook of Spacetime (2014).
Michael H. Whitworth is a lecturer in the English Faculty, University of Oxford, and
a tutorial fellow of Merton College, Oxford. He is the author of Einstein’s Wake: Relativity,
Metaphor, and Modernist Literature (OUP, 2001) and other articles and chapters on the rela-
tions of modernist literature and science. He was the co-founder of the British Society
for Literature and Science. He has also written extensively on Virginia Woolf and has
edited her novels Orlando and Night and Day.
Aaron Sidney Wright is a Postdoctoral Scholar at the Suppes Center for History and
Philosophy of Science at Stanford University. Previously, he was a Social Sciences and
Humanities Research Council Postdoctoral Fellow at the Department of History of
Science at Harvard University. He received his PhD from the Institute for the History
and Philosophy of Science and Technology at the University of Toronto. His research
is concerned with the practice of physics over time, and what this practice tells us about
science and the natural world. Particularly interested in ‘historical ontology’, he has
made detailed studies of the work of theoretical physicists who studied the vacuum, or
empty space.
1
Introduction
Ether—The Multiple Lives of a Resilient Concept

Massimiliano Badino and Jaume Navarro

1.1 ETHEREAL NARRATIVES


When did the ether disappear? Did it actually ever exist? Is it not back among some
contemporary physicists? Are you trying to chase a ghost or to resuscitate a zombie?
Questions like these were common in the academic events where the papers compiled in
this volume were presented. Indeed, the ‘Ethereal Aether’, or ‘the nothing that connects
everything’, to quote only two titles of books devoted to the history of the ether,1 has
often attracted the attention of historians of science, philosophers and physicists, and its
epistemic and ontological status been the subject of much discussion and speculation.
While the electron became the favourite probe for theories of scientific realism in times
past, the ether was a frequent case study of theories of scientific change. For quite some
time, the story of its supposed rejection via the experiments of Michelson and Morley in
the late nineteenth century reassured Popperians and was used by many physics teachers.
Later on, the ether became a major symbol of the incommensurability between classical
and modern physics, exemplifying the Kuhnian struggle between a young generation of
relativists and quantum physicists and the stubbornness of conservatives and crackpots.
More recently, social historians of science have moved away from the categories of belief
and unbelief in determining the nuances of the abandonment of the ether in the early
twentieth century, bringing in elements such as technical or pedagogical use.2
In any case, there is a major consensus in thinking that the ether was mostly erased
from any central role in physics and in culture after World War I. The best and most
often quoted work on the history of the ether, the 1981 volume Conceptions of Ether, took
for granted this timeline by limiting the study of the History of Ether Theories to the period
1740–1900.3 The recurrent question ‘what did Einstein really think of the ether’ is only
an exception to this rule, and even then the view is that the ether qua theory disappeared
in spite of (some of ) Einstein’s opinions.4

Badino, M. and Navarro, J., ‘Introduction: Ether—The Multiple Lives of a Resilient Concept’ in Ether and
Modernity, edited by J. Navarro. © Oxford University Press 2018. DOI: 10.1093/oso/9780198797258.003.0001
2 Introduction

Certainly, in the narratives of the death, rejection or demise of the ether, Einstein and
the theory of relativity play a central role. But, as the chapters in this book show, far from
killing the ether off, special and general relativity (and, to a lesser degree, quantum phys-
ics) caused an explosion of ether narratives into different directions, and only a few of
them can be traced back to a Manichean dichotomy between classical and modern. These
stories cross each other; they share contexts, aims and actors, thus making it difficult to
disentangle them into simple narrative strands. Thus, the absence of a chapter specifically
devoted to ‘Einstein and the ether’ should not be understood in terms of Edmund
Whittaker’s old strategy to diminish Einstein’s role in the development of relativity but as
a way of emphasising the complexities of the ether in the early twentieth century.5
Rather, this book is a snapshot of the multiple lives of the ether in the first decades of
the last century. From developments in pure mathematics to wireless technologies, from
modernist art to spiritualism and from popular to alternative views of physics, the chap-
ters in this book present us with an array of narratives that develop along several lines of
fracture, none of which can give us a single coherent picture. Thus, far from consolidat-
ing the traditional dichotomies between classical and modern, between British and
Continental, between material and spiritual and between ‘true’ and popular physics, this
book challenges the explanatory role of these fissures. Indeed, these dichotomies are as
many ways to tell the story of the ether: however, they fail to capture the whole story
and it is in their failure that one can catch a glimpse at the peculiar features of this con-
cept. Thus, the goal of this introduction is to follow some of these lines of fracture and
to unearth the dynamics underlying the place of the ether in the early twentieth century.

1.2 CLASSICAL AND MODERN


The most obvious way to tell the story of the ether in the twentieth century is perhaps
along the line of fracture that separates classicality and modernity. It is very tempting to
make use of the revolutionary watershed, which includes both relativity and quantum
theory, in order to safely locate the ether within the realm of ‘classical physics’ as opposed
to a newer and advanced ‘modern physics’. The narrative of the ether would thus unfold
as the battle of some diehards for hopelessly old-fashioned scientific practices and values
and against the unstoppable tide of progress. But, as some historians of science have
shown, rather than mere labels, ‘classical’ and ‘modern’ are extremely fluid, largely co-
constructed and continuously renegotiated categories which transcend the boundaries
of physics and belong to the public sphere at large.6 For this reason, the central argument
of this volume concerns the comparison of images of the ether in science and in culture
with the explicit goal of tracking the different dynamics, timescales, and meanings with
which the term ‘ether’ was appropriated by different communities. Only in this way does
it become possible to map out the complex conceptual clusters hidden under the deceiv-
ingly simple notions of classicality and modernity. What is normally used as a sharp
dichotomy reveals so many intersections, entanglements and overlaps that it was not
Classical and Modern 3

inconsistent to regard something like the ether as classical and modern at the same time,
although for different reasons in different places and by different actors.
Let us begin at the beginning, that is, with physics. In the early years of the past cen-
tury, Einstein’s relativity theory was viewed by many more as an opportunity to rethink
the ether than a reason to reject it. In Chapter 5, Scott Walter argues that the foundations
of electron theory rested on a plexus of dynamical and kinematic problems. On the one
hand, the description of electromagnetic phenomena in different states of motion called
for a redefinition of the terms ‘space’, ‘time’, ‘speed of light’ and ‘reference frame’. On
the other, the bare fact that electric charges were involved required a medium of propa-
gation for electromagnetic waves. For all its cumbersomeness, the ether allowed for a
unification of kinematics and dynamics: it was the supporting medium and the absolute
frame of reference and it was the storage device of energy. The British versions of elec-
tron theory, described by Richard Noakes in Chapter 6, made good use of these features.
In addition, the ether proved useful even when combined with the most advanced math-
ematical tools as shown by Poincaré’s combination of group transformations and
Langevin’s velocity waves to obtain the Lorentz contractions. By contrast, Einstein only
offered ‘a pair of postulates, the logical consistency of which was suspect’ and the
hypothesis of the relativity of time, a suggestion that even a flexible mind like Poincaré’s
looked upon as uninviting.
In this context, it is scarcely surprising that some physicists were reluctant to abandon
the ether completely and tried instead to rethink its role under the kinematic constraints
of Einstein’s relativity. This quest for a novel way to understand the dynamical founda-
tions of the electrodynamics of moving bodies generated what Walter calls the ‘concep-
tual drift of the concept of ether’: a series of attempts at filling up with a renewed ether
the void that Einstein’s epistemological austerity had left behind. Deprived of kinematic
properties, the ether simply became a substratum, the loftiest metaphysical concept a
physicist’s conscience could bear and deal with. It is within this framework that one has
to look at the reception of Minkowski’s work, which, slowly but steadily, became the
veritable vehicle of acceptance of special relativity. Connemara Doran’s novel step-by-
step analysis of the road from Poincaré to Minkowski’s geometrisation tools, provided in
Chapter 4, helps understand this evolution.
Efforts to make relativity and the ether happy together continued during the 1910s and
1920s in Britain. In Chapter 12, Aaron Wright describes this process as a sort of ‘domesti-
cation’ of the ether, an approach towards modern physics which marked Dirac’s physical
education and eventually fed his understanding of the role of the ether. Einstein’s 1905
famous conclusion that the ether was ‘superfluous’ contained a positive message: after
all, relativity had not pronounced the ether non-existent. This alleged compatibility
between the ether and relativity was one of the main weapons on the ether front. If the
ether could be neither observed experimentally nor undermined theoretically, the only
remaining option was to characterise it through its functions. This was the rationale
behind many of the ‘relabelling strategies’ so clearly described by Wright in the opening
sections of the chapter.
4 Introduction

This functional reinterpretation of the ether is key to understanding its vicissitudes.


Fortunately, relativity, particularly general relativity, offered plenty of conceptual space.
Its unprecedentedly complicated mathematical architecture was far from being self-
explanatory. On the contrary, several corners in its symbolic maze had to be interpreted
and many of them were open to alternative interpretations. Einstein himself admitted
this point in 1916, and in 1918 he explicitly stated that, according to general relativity, one
had to ascribe physical properties to the empty space via specific components of the
gravitational potential. As Richard Staley explains in Chapter 10, for Einstein, this math-
ematical prescription created the site for a new role of the concept of ether. This is just
one instance, albeit possibly the most illustrious, of a line of thought that became popu-
lar especially in the British scientific and public discourse: relativity had not declared the
ether dead and, in fact, it might even vindicate it eventually. Thus, for example, Ebenezer
Cunningham drew a very straightforward conclusion from special relativity: if there
are multiple inertial frames, then there should be multiple ethers. On the contrary,
as partly seen in Chapter 2, by Imogen Clarke, and Chapter 8, by Jaume Navarro,
Arthur Eddington was happy to assume only one collection of point events, and he saw
in this symbolic representation of the physical world the same connecting function for-
merly accomplished by the ether. Another interesting instance is Poincaré’s ‘structural’
reinterpretation, discussed by Doran in Chapter 4. According to Poincaré, the existence
of the ether could be reduced to ‘a natural kinship between all the optical phenomena’.
From this point of view, which could safely be defined as ontic structural realism avant la
lettre, the ether hypothesis was no longer true or false—it was simply more or less con-
venient.
These examples illustrate two deeply entwined points. First, the development of
modern physics, relativity as well as the quantum, did not happen as a gestalt jump out
of the classical world. Quite the contrary—new theories, concepts and perspectives were
conceived in constant dialogue with tradition, which never really died out. It is by explor-
ing the consistency of new ideas with regards to classical wisdom that the full extent of
their revolutionary potential can be appreciated.7 Second, and related, modern physics is
not a complete overcoming of the classical one, as the narrative mentioned at the begin-
ning of this section would have it.
The conceptual space for the ether lies at the very heart of the dialectic between clas-
sic and modern: in what modern science does not tell us—in what it does not bother to
specify, determine and interpret. Dirac, as discussed in Chapter 12, represents another
illuminating example of this entanglement between the classic and the modern. For
him, a reasonable relativistic or quantum theory of a system should be based, via analogy,
on a corresponding classical theory of the same system, a point critically reiterated in his
1951 theory of electron.8 Precisely this intuition led him from a new classical theory of
the electron to formulating an ether for quantum electrodynamics as a field of velocities
that a charge sitting in a certain point would pick up. Once again, there is a specific math-
ematical object, the electromagnetic potential, signalling the presence of the ether, but,
as Wright perceptively observes, Dirac did not confine himself to carving out a space for
Classical and Modern 5

the ether: he showed how a classical concept could still contribute to solve the riddles of
modern science.
For many experts, relativity and the ether could coexist peacefully and even fruitfully
but, for others, they were mutually exclusive. In Chapters 7 and 10, Arne Schirrmacher
and Richard Staley, respectively, explore in different ways the politically tinged support
for this claim in Germany and we shall return to this issue later in this introduction, in
Sections 1.3 and 1.4. For the moment, however, we would like to focus on the interesting
case of Dayton C. Miller, thoroughly discussed by Roberto Lalli in Chapter 9. The con-
ceptual flexibility of the ‘modern ether of relativity’, so to speak, contrasted with stricter
experimental demands, namely a very precise value—a round zero—for the ether-drift
experiments. As Miller promptly noted, no such exact result had ever been attained,
thus leaving some space for the ether. Moreover, the very possibility of performing
Michelson–Morley-like experiments presupposed the existence of the ether. These two
points made a powerful effect on the American physical community, which had built its
fortune especially on physical astronomy and recognised the ether as an indispensable
tool to make sense of their daily practices. This explains why the ether-drift experiments,
promptly if not fully satisfactorily accommodated in classical physics, were paradoxically
resuscitated by relativity.
This experimental branch of the story turned the survival of the ether into an all-or-
nothing affair. But even in this case of head-on confrontation, the actors were not reject-
ing modernity as a whole. Instead, they were denouncing relativity as a wrong track,
both scientifically and culturally. This suggests that the dialectic of modernity should not
be viewed as a clash between blocks, but rather as the result of crafty political and
rhetorical tactics aiming at formulating the most effective alignment of concepts, values
and emotions.
One can find excellent examples of this tactic in the public discourse on the ether. The
first key factor to understanding the dynamics underlying the debate on the ether in the
public sphere is the particular conjuncture of European society and culture in the early
decades on the twentieth century. New social actors and new political ideas shook the
very foundations of the international order patiently put together during the nineteenth
century. The appalling carnage of World War I and the tumultuous upheaval happening
in Russia took the entire continent aback. Behind the lights and the dances of the belle
époque, fear and uncertainty were the dominating feelings among Europeans. In this
period of emotional instability, Oliver Lodge’s plea for the ether could find more than
one interested ear.
As Imogen Clarke and Michael Whitworth duly stress in Chapters 2 and 3, respectively,
Lodge’s communication prowess was almost unmatched. In his capacity as scientific
expert and public figure, he managed to make his points heeded both at physics meetings
and in the social arena. More importantly, he knew how to get at the layman’s heart. In
an age in which the cleavage between the common man and official science was rapidly
increasing, Lodge was able to create a sense of proximity between people and the ether.
The ether was still with us—relativity had never really expunged it and it could still be
6 Introduction

seen at work as he was broadcasting his speeches on the BBC. The ether embodied all
those cognitive values of intelligibility, visualisability and intuitiveness that the math-
ematical abstruseness of modern physics wanted to obliterate. It rested on a perfectly
logical analogy: in the same way that water waves need water, so electromagnetic waves
need an electromagnetic medium to propagate. Furthermore, the ether was indispens-
able for the good face of modernity: as Chapter 8 shows, the association of the ether with
wireless technologies gave the former a clear seal of modernity and the latter a reassur-
ing sense of stability and continuity. While relativity and quantum physics were useless
for understanding technology, let alone improving it, the ether, by contrast, played a
central role. The magic of the wireless, which allowed people from the opposite ends of
the earth to speak as if together in the same room, made perfect sense once the electro-
magnetic waves were pictured as perturbation of an all-pervasive medium. As the output
of a type of physics now in jeopardy, the wireless technology literally embodied scien-
tific as well as cultural values, both modern and traditional. It is unsurprising, therefore,
that the wireless community became instrumental in sustaining the concept of the ether
well beyond its survival in official science.
In Lodge’s capable hands and with the cooperation of the wireless practitioners, the
ether became a highly symbolic notion. It obviously related to electromagnetic theory, a
chapter of science in which Britons had been unquestioned masters, from Faraday to
Maxwell, William Thomson, Joseph J. Thomson and, why not, Lodge himself. And
it also related to the wireless technology and Britain’s imperial supremacy in electrifying
and communicating the world. The ether, in sum, symbolised the cherished values of
the classical British civilisation: stability, intelligibility and common sense. And, while it
opposed all the uncomfortable novelties of the new century—moral uncertainties, social
changes, Bolshevik revolutions, Freudian abysses and cerebral mathematical theories—it
was modern in a good way: it expressed sensible progress and dynamism; it had a future
in science as well as in society. The public image of the ether was neither classical nor
modern but a cunning combination of both.

1.3 KNOWLEDGE, ETHOS AND AUTHORITY


One of the most serious flaws of the standard narrative on the ether is that it sets the entire
issue in terms of a choice between scientific theories. By making the question depend on
the intrinsic cognitive qualities of the two options, relativity theory and the ether hypoth-
esis, it regroups all the nuances of the debate into two broad categories: the former were
necessarily right, and the latter, wrong. By contrast, this book shifts the focus from the
theories to the actors of this debate, to retrieve their distinctive voices, backgrounds and
agendas. In this way, we can now appreciate elements previously neglected by historians.
One major example is the theme of authority, which permeates several articles.
In order to be an influential contributor to a debate, one needs to gain sufficient
authority; but—and this is the particularly intriguing aspect—in the ether debate, authority
Knowledge, Ethos and Authority 7

is a peculiar mixture of knowledge and ethos. Dirac, a Nobel laureate, was wittingly
provoking the scientific community in 1951 by connecting his well-respected work with
the anachronistic notion of the ‘æther’, in his battle for a new formulation of quantum
electrodynamics. Although the attempt did not meet with great acceptance, Dirac not
only used his authority and knowledge to bring back some form of ether but also used
the authority of the ether, so to speak, to create a debate on the foundations of quantum
electrodynamics.
Oliver Lodge and Arthur Eddington are two major examples of this entanglement
between ethos, knowledge and authority. In Chapter 3, Whitworth describes with com-
mendable precision Lodge’s techniques to approach his public in order to foster not only
a sense of expertise but also moral trust. By showing wisdom rather than just knowledge
and by helping the reader (or the listener) to co-create the space for the ether, Lodge
becomes the personification of those ideals of stability, reasonableness and Christian
patience which were associated with the Victorian culture and of which the ether was an
expression. And as Clarke argues in Chapter 2, Lodge and Eddington were very effective
in convincing the general public that there was a genuine scientific discussion on the ether
and that his was an active voice in this discussion. Although different on most points,
Lodge’s and Eddington’s argumentative strategy included using their scientific authority
to promote a critical attitude towards the infallibility of science. By claiming that, on the
ether problem, the jury was still out, they achieved multiple results: they presented them-
selves as humble men with no pretension to know everything, they challenged some
mainstream interpretations of relativity and they sympathetically nodded to their privileged
audiences—the wireless community in the case of Lodge, mathematical idealists in the
case of Eddington, and a broad array of liberal Christians in both.
This interplay of ethos and knowledge is key to understanding the complex dynamics
of scientific and popular culture in the early decades of the twentieth century, but it also
helps clarify the tectonic shifts within the scientific culture itself, particularly along both
the geographical and the disciplinary fracture lines. During the nineteenth century, these
differences were increasingly painted in critical terms and, often, related to specific men-
talities, spirits, cultures and ethos. The criticism levelled by William Thomson and Peter
Guthrie Tait against the Continental abuse of mathematics, for instance, and Pierre
Duhem’s famous distinction between the French and the British ‘minds’ are examples of
this train of thought. The debate on the ether was not immune from these geographical
undertones. As Noakes argues in Chapter 6, Lodge blamed Germany’s materialist, deca-
dent and ultimately dangerous morals on a philosophy that neglected, among other
things, the ether. After all, the infection of modern science had started in Germany, the
fatherland of both relativity and quantum theory. Given the peculiarities of the German
thought and ethos, this genesis was not surprising, and the result had to be properly
domesticated before any British attempt of appropriation.
But, on the other side of the Channel, they thought it differently. Notoriously, Philip
Lenard exposed relativity as contrary to the German—to wit, Aryan—spirit, using his
authority as a Nobel laureate to challenge Einstein’s views and even to accuse him of
8 Introduction

plagiarism; he also fought a bitter struggle against J. J. Thomson and George Simpson,
whom he regarded as careless and unfair experimenters. On several occasions, as
Schirrmacher describes in Chapter 7, Lenard contrasted his rigorous and systematic
approach to experimental physics with the sloppy attitude of the Britons. Much to his
dismay, the latter were also more successful in terms of publication policy and accessibil-
ity, a feature that only made Lenard more furious.
The case of Lenard shows how national differences often intersect disciplinary bound-
aries. But the issue was more complex. Relativity and its mathematical maze were not
born within the experimental domain. With the new strange player in the scientific com-
munity, the theoretical physicist, doubts could be legitimately raised that the theoreti-
cian’s mathematical dreams could simply lead too far. Unlike those of the experimenter,
the theoretician’s constraints were vague and shaky: for the latter, common sense was a
burden, a relic of a past scientific ethos. And the ether played on both sides of the divide:
for some, it was a theoretical figment of the imagination, but for others it was the most
experimentally grounded object in the fight against the excesses of modern theoretical
physics. Miller rebelled against this drift of modernity, but his story reveals an ironic
twist. His painstaking efforts to obtain precise checks of the predictions of relativity, bril-
liantly described by Lalli in Chapter 9, eventually came to a close via the experiments of
Georg Joos, who was not only a German but also a representative of a new—and mod-
ern—experimental physics. Eventually, the ultimate application of the experimenter’s
ethos made the experimenter himself a dispensable player in the scientific game.

1.4 MATTER AND SPIRIT


In nineteenth-century science, the ether served two main functions: it was the supporting
medium for electromagnetic radiation and it was the storage for the energy no longer
available for human usage. Each function required the ether to be a material entity of a
very peculiar nature. It had to be rigid enough to allow the extremely fast propagation of
transversal waves, thin enough not to slow down the planets moving through it and
capable somehow of keeping the dispersed energy until, at some point in time, it could
be released. None of these formidable properties was ever made consistent with the others:
they were all justified by conceptual requirements internal to the electromagnetic theory.
But, for many, the unresolved tension between the several properties of the ether was
more of an opportunity than a hindrance. An entity of undefinable, amphibious features,
the ether could not simply be an ordinary part of the physical domain: rather, it could be
the bridge between our world and the ultramundane. The ether became an obvious
resource to the resurgent spiritualism of the second half of the nineteenth century.
A late convert to this trend, Oliver Lodge, became a powerful advocate of this culture,
placing himself in the British tradition of natural theology. Lodge could easily inscribe
his own agenda in this unfinished project as the rhetorical strategies unfolded by Clarke,
Whitworth and Noakes show. Lodge’s favourite argument that science does not have all
Matter and Spirit 9

answers was meant in the broadest way possible. The verdict against the ether was not
carved into stone, not only because science always reassesses its results but also because
there were superior goals in intellectual inquiry, goals which science could not avoid.
Thus, the reminder that the universe was still for the better part mysterious and ‘unseen’
was not only the expression of a healthy epistemological modesty but also the evocation
of the spiritual subtext accompanying the scientific discourse. It is precisely within this
framework that Lodge could parallel electron theory and psychic research: both had
been mysterious at one time but were now ‘concrete and tractable’.
On this emotional front, World War I also played an important role. As Lodge, Lenard
and many others had lost their beloved ones in the trenches, the need to believe that the
cruel carnage was not the ultimate end grew stronger. This powerful thought, which Lodge
condensed in his book Raymond, could not find a more sympathetic audience: the stor-
age power of the ether worked for electricity and the still-mysterious soul equally well.
In Britain, the connection with the project of natural theology gave to the discourse
on the ether a characteristic psychical and almost religious tinge. By contrast, in other
areas of Europe, the tension between matter and spirit presented itself in cultural, polit-
ical and artistic varieties. Lenard’s political usage of the ether as a brand of Deutsche
Physik was part of a powerful neo-Romantic movement whose declared aim was to
recover German Naturphilosophie and a more authentic relation with the spiritual side of
nature. Aggressively opposed to the ruthless industrialisation and the blind scientificity
of the modern world, the upholders of this movement sang the praise of a mystical con-
nection between natural environment, people and spiritual tradition. This was the notori-
ous concept of the Volk, a profound combination of blood, soil and an omnipresent
‘ether’ connecting individuals in a spiritual unity.9
The connection between occultism and radical political ideas is old. In his book,
Mesmerism and the End of the Enlightenment in France, Robert Darton superbly described
how mesmerism—a practice later supported by the Society for Psychical Research—got
mixed with Rousseau’s ideas of a primeval human energy and eventually became a vec-
tor of radical politics during the Enlightenment.10 In the 1920s, many German intellectuals
moved along a similar track. Inspired and supported by the indefatigable Eugen Diederichs,
the publisher of a massive edition of Meister Eckart’s works, the self-proclaimed Free
German Youth organised mountain walks and gatherings at which Madame Blavatsky’s
theosophy was discussed in a lively manner. What the Free German Youth had in com-
mon with other spiritualistic movements with which it was in contact was its struggle
against the damaging aspects of modernity, together with a firm belief in an all-pervading
medium. One Herbert Reichstein even claimed in the 1920s that the first Aryan was
created by a shock from an electro-spiritual ether. However, the Free German Youth
was unique in that it stressed the connection between science and art. In its pretension
to explain everything, modern science progressively neglected the inward eye and
turned man into a cosmopolitan producer of material goods. The path to counter this
decline was forcefully outlined by Julius Langbehn—one of the most influential minds in
the Free German Youth—in his Rembrandt als Erzieher (also published by Diederichs):
10 Introduction

Germans must be turned into artists, and science must be turned into art. Only through
an artistic relation with the world would it become possible to recover the ‘spirit’ and
‘possession of such a spirit meant recalling that which was truly genuine, the Germanic
past, as opposed to modern and evil rationalism’.11
Science, spiritualism and art were thus intertwined in a project of cultural renewal.
However, the relations with modernity could be wavery. For the intellectuals of the Free
German Youth, art had primarily an antimodernist meaning. Fidus (the pseudonym of
Hugo Höppener), arguably its most representative artist, declared overtly that his work
originated from a connection with the ultramundane world; but, in terms of content, he
was committed to retrieving the legacy of traditional Germany. By contrast, in Chapter 11,
Linda Henderson shows very clearly that Umberto Boccioni was attempting a very dar-
ing synthesis between spiritualism and modernity via a reconfiguration of the notion of
matter. For Boccioni—not by chance an enthusiastic reader of theosophical writings
himself—the storage function of the ether, on which Lodge’s occultist arguments rested,
became the vehicle to cross the borders between ether and matter. Unlike cubism, in this
case the mixture of past, present and future was not in the memory but in the bodies
themselves: matter was ether condensed, and ether was matter vaporised. Hence, mat-
ter embodied the elan of the spirit, to use an expression of Bergson (whose works were
published in German, once again, by the energetic Diederichs).
Although the appropriation of modernity changed importantly from one current to
another, a common pattern seems recurrent, to wit, the rejection of purely rational
thought and the search for alternative ways of accessing reality, possibly a deeper reality.
In this regard, some chapters in this book demonstrate that sensory perception (at times
opportunely extended) and aesthetic intuition maintained a central role. The aesthetic
approach to science, as much as the aesthetic approach to life, aspired to attain a superior
form of knowledge, one that could not be easily conceptualised, but could serve as an
effective guide for daily tasks, scientific or otherwise. Doran, Staley and Wright insist,
with different nuances, on this theme. It is less relevant that, for Poincaré, Mach and
Dirac, the path to the unity and beauty of this world view passed through the symmetries
of mathematics or those of sensorial experience; what is relevant is that this form of
knowledge was immediate, intuitive, irreflexive and, in some sense, spiritual.

1.5 EPISTEMOLOGY AND EMOTIONS:


A PLEA TO PLURALISM
Unsurprisingly, we have circled back to knowledge. The philosophical stories we began
with insisted that the ether was the idle wheel of electromagnetic theory.12 From the
point of view of an epistemology that focuses exclusively on what leads to successful
predictions, this claim is hardly questionable: the ether as such certainly did not directly
contribute to Maxwell’s equations. This epistemology, in turn, whispered in the historians’
Epistemology and Emotions: A Plea to Pluralism 11

ear the leitmotiv we see at work in many standard narratives: once Einstein awoke the
scientific world from its ethereal slumber, the ether remained an option only for the
nostalgic, the passé, the left behind. However, the chapters in this book show that the
methodological dichotomies these narratives traditionally hinged on—classic versus
modern, progress versus conservatorism, continuity versus discontinuity—should not
be taken to be as clear-cut as they used to be, because the discourse on the ether crosses
these distinctions transversally. It is not by chance that Einstein is more evoked than dealt
with in this book, for the really interesting question is not ‘Why were people so stubborn
to stick to an outdated and blatantly false idea?’, but rather ‘What made the ether such a
resilient concept?’. This novel question forces us to jettison the hegemonic positivistic
picture of knowledge and assume a pluralistic point of view. The answer is to be found
in the complex epistemological landscape and in the structure of feeling, as Raymond
Williams used to call it,13 that allowed the ether to survive in the wake of relativity.
If we explore this landscape more carefully, we are immediately struck by the rich-
ness and variety of the discourses on scientific knowledge in which the ether can be
found. For example, as Whitworth shows in Chapter 3, Lodge’s literary ingenuity helped
him create a sense that the ether was physically present—and made good use of it in
wireless telegraphy—and no abstruse mathematics could convince us to the contrary.
The idea that we might have a sense for the ether, present among some British writers,
might be traced back to William Thomson’s presidential address at the Birmingham
Midland Institute (3 October 1883), ‘The Six Gates of Knowledge’.14 Famously, Thomson
argued that knowledge always entered through the senses which, supplemented with
the ‘sense of force’, accounted for the six gates of the title, and he did not discard the
possibility that other senses—perhaps even a sense of the ether—might at some point
be discovered.
If we expand our gaze beyond the Channel, we find that both in Der Analyse der
Empfindungen (1886) and in Erkenntnis und Irrtum (1905), Ernst Mach repeatedly used his
own body as a probe to explore the external world, although he added an important twist
to the common-sense empiricism of the British tradition. For Mach, experience was con-
stituted by ‘elements’, a notion importantly larger than just sensorial data:

Perceptions, presentations, volitions, and emotions, in short the whole inner and outer world,
are put together, in combinations of varying evanescence and permanence, out of a small
number of homogeneous elements. Usually, these elements are called sensations. But as
vestiges of a one-sided theory inhere in that term, we prefer to speak simply of elements.15

Mach was opening up a world in which the transcendental subject of the Enlightenment
progressively lost its role of legislator of nature.16 The boundaries between subject and
object became as blurry as those between matter and spirit and the ways to access
knowledge expanded to positively encompass emotions, desires and volitions. From
this perspective, mere experimental undetectability mattered very little. As Staley and
Schirrmacher show, the permanence of the ether in the Germanic culture was related to
its capability to serve epistemological functions much higher than that served by the lab.
12 Introduction

Ultimately, it is the adaptability of the ether to multiple cultural contexts, and the epis-
temological pluralism that underlies this capability, that the rich diversity of the essays in
this volume capture so effectively. Thus the obvious question is: what is it in the epis-
temological fabric of the concept of ether that allowed it to live so many lives well
beyond its early proclaimed redundancy? A preliminary answer, which is more a collec-
tion of thoughts for further research than a fully fledged response, is that the ether may
be a sort of interstitial concept. Let us close this introduction by elaborating this thought.
The reorganisation of knowledge, leading to a deep scientific revolution, always leaves
epistemic interstices. One obvious example is the famous phenomenon of the ‘Kuhn
losses’.17 Thomas Kuhn argued that, when a new paradigm takes over, it might happen
that problems previously well understood in the old paradigm suddenly become intract-
able. In other words, paradigms always leave epistemic gaps that must be filled with
resources external to them. By the same token, relativity made the ether kinematically
superfluous, but it left behind the problem of making sense of electromagnetic events
occurring in the vacuum, as Dirac would point out. It is hardly surprising, then, that many
authoritative physicists tried to fill this interstice by reconfiguring the concept of ether.
Analogous interstices can crack open at the interface between science and culture. The
chapters in this book provide numerous stories of appropriation of the ether by different
communities. The wireless practitioners, the occultists, the Italian futurists, the German
experimental physicists, and so on, had questions about which modern science only had
useless, irrelevant or even unpleasant answers, if at all. By contrast, the ether could com-
fortably live in all these spaces made available by the imperfect interconnection between
science and culture. As an interstitial concept, the ether was plastic and pliable enough to be
adapted to diverse contexts, because it was no longer a specific object but rather a multidi-
mensional concept able to serve a number of epistemic, symbolical, social, political, emo-
tional, moral and even scientific functions, some of which, in contrast, were perfectly in
tune with modernity. But such an extreme flexibility had, of course, a downside. Sitting at
the interstices between multiple discourses, the ether was not integral to any of them and
was not autonomous. Hence, it had to be sustained by the continuous effort of authorita-
tive figures energetically acting in the public sphere. This, in turn, generated a complex
dynamics of alliances, negotiations and strategies reaching out a considerable variety of
debates. It is precisely its interstitial nature that makes the ether such an effective entry
point in that thorny juncture of modernity that was the beginning of the twentieth century.

NOTES
1. Lloyd S. Swenson, Jr, The Ethereal Aether. A History of the Michelson–Morley–Miller Aether-Drift
Experiments, 1880–1930 (Austin: University of Texas Press, 1972); Joe Milutis, Ether. The Nothing
that Connects Everything (London: University of Minnesota Press, 2006).
2. Andrew Warwick, Masters of Theory. Cambridge and the Rise of Mathematical Physics (Chicago:
Chicago University Press). See also Graeme Gooday and Daniel J. Mitchell, ‘Rethinking “Classical
Notes 13

Physics”’, in Jed Z. Buchwald and Robert Fox, eds., The Oxford Handbook of the History of
Physics (Oxford: Oxford University Press, 2013), 721–64.
3. Geoffrey N. Cantor and Michael J. S. Hodge, Conceptions of Ether. Studies in the History of Ether
Theories, 1740–1900 (Cambridge: Cambridge University Press, 1981).
4. Ludwik Kostro, Einstein and the Ether (Montreal: Apeiron, 2000).
5. Edmund Whittaker, A History of the Theories of Aether and Electricity, vol. 2 (London: Longmans,
1953).
6. On this point, see Richard Staley, ‘On the Co-Creation of Classical and Modern Physics’, Isis,
96 (2005): 530–58.
7. On this process of exploration of the relation between classical and modern ideas in physics,
see Jochen Büttner, Jürgen Renn and Matthias Schemmel, ‘Exploring the Limits of Classical
Physics: Planck, Einstein, and the Structure of a Scientific Revolution’, Studies in History and
Philosophy of Modern Physics, 34 (2003): 37–59. For a pedagogical take on the same issue, see
Massimiliano Badino and Jaume Navarro, ‘Pedagogy and Research: Notes for a Historical
Epistemology of Science Education’, in Massimiliano Badino and Jaume Navarro, eds., Research
and Pedagogy: A History of Quantum Physics through its Textbooks (Berlin: Edition Open Access,
2013), 7–30.
8. On the role of analogy in the emergence of quantum physics see Olivier Darrigol, From
c-Numbers to q-Numbers: The Classical Analogy in the History of Quantum Physics (Berkeley:
University of California Press, 1992).
9. See e.g. George L. Mosse, ‘The Mystical Origins of National Socialism’, Journal of the History
of Ideas, 22 (1961): 81–96, and George L. Mosse, Nazi Culture: Intellectual, Cultural, and Social Life
in the Third Reich (Madison: University of Wisconsin Press, 1966).
10. Robert Darton, Mesmerism and the End of the Enlightenment in France (Cambridge, MA: Harvard
University Press, 1968).
11. Mosse, ‘The Mystical Origins of National Socialism’, p. 88.
12. See e.g. Philip Kitcher, The Advancement of Science (Oxford: Oxford University Press, 1993),
pp. 143–9.
13. Raymond Williams, Marxism and Literature (Oxford: Oxford University Press, 1977).
14. William Thomson, ‘The Six Gates of Knowledge’, in Popular Lectures and Addresses, vol. 1
(London: Macmillan, 1891), 260–306.
15. Ernst Mach, The Analysis of Sensation (New York: Dover, 1959), p. 22.
16. On this point, see Jürgen Habermas, Knowledge and Human Interests (Boston: Beacon Press,
1971), pp. 81–90.
17. Thomas S. Kuhn, The Structure of Scientific Revolutions (Chicago: University of Chicago Press,
3rd edn, 1996), pp. 103–10.
2
The Ether at the Crossroads of Classical
and Modern Physics
Imogen Clarke

The matter of which I have been speaking so far is the material which builds up the earth, the
sun, and the stars, the matter studied by the chemist, and which he can represent by a formula;
this matter occupies, however, but an insignificant fraction of the universe, it forms but minute
islands in the great ocean of the ether, the substance with which the whole universe is filled.
J. J. Thomson, Address of the President of the British Association for the
Advancement of Science, 1909.

2.1 INTRODUCTION
On the evening of 25 August 1909, inside the Walker Theatre in Winnipeg, Canada, the
esteemed British physicist J. J. Thomson opened the seventy-ninth meeting of the British
Association for the Advancement of Science. Thomson’s presidential address, an over-
view of current developments in physics, focused largely on the ether, which he declared
to be ‘not a fantastic creation of the speculative philosopher’ but rather ‘as essential to us
as the air we breathe’.1 Some three decades later, in his 1942 biography of Thomson,
Lord Rayleigh criticised this approach, noting that his subject had not taken the ‘fashion-
able view about ether’ but instead displayed a ‘robust confidence in its reality and neces-
sity’ in the face of mounting evidence to the contrary.2
Looking back on the physics of the first half of the twentieth century, the twenty-first
century reader is inclined to agree with Rayleigh’s assessment. Against the backdrop of
relativity theory, quantum mechanics and radioactivity, Thomson’s commitment to a
decidedly nineteenth-century theory certainly appears outdated. And yet, at the time of
the Winnipeg speech Thomson was Director of Cambridge University’s Cavendish
Laboratory, an institution renowned for its extensive role in the development of a new
style of modern physics during this period. Under Thomson’s helm, Cavendish physi-
cists moved away from a focus on precise measurement of physical constants and towards
investigations into the fundamental structure of matter, theoretical speculation and the

Clarke, I., ‘The Ether at the Crossroads of Classical and Modern Physics’ in Ether and Modernity, edited
by J. Navarro. © Oxford University Press 2018. DOI: 10.1093/oso/9780198797258.003.0002
The Ether and Discontinuity, 1909–14 15

favouring of visual assessments of experiments over metrical precision. Where previous


studies of matter had investigated it in relation to the ether, there was here a move from the
macroscopic to the microscopic, as Cavendish researchers attempted to uncover the internal
structure of rays and atoms. Thomson himself was famed for his turn-of-the-century ‘dis-
covery’ of the electron, a steadfast symbol of modern physics. Nevertheless, he maintained
throughout his life a commitment to the ether, a cornerstone of classical physics.3
How do we reconcile these two aspects of J. J. Thomson’s life, the classical and the
modern? The complexities of Thomson’s career have been discussed and picked apart,
with it now being established that Thomson himself did not conceptually connect the
‘corpuscles’, for which he found experimental evidence, with existing theories of the
electron.4 But Thomson could also be well served by a broader shift in how we frame our
histories of early twentieth-century physics, a move away from a traditional approach
that divides physics and physicists into classical and modern.5 This dichotomy obscures
much of the complexities of the period, often resulting in an incomplete picture, with
only half the story being told. Indeed, Rayleigh’s account of Thomson’s life and career is
a prime example of such failings, with the ether scarcely mentioned in the biography of
a man who saw it as fundamental to his life’s work.6 With recent scholarship challenging
the dichotomy of classical/modern, now is an opportune time to re-evaluate the ether’s
position in the twentieth century.7
In this chapter, I aim to liberate the ether from its historiographical assignment to clas-
sical physics, and consider its role in debates surrounding the future of the discipline. I
begin by looking at the discussions underway in professional spaces between 1909 and 1914,
and suggest that a physicist’s commitment to the ether does not classify them as a ‘classi-
cist’ but rather as an advocate of continuity in the discipline. I then examine the ether’s
‘popular’ life following the well-publicised 1919 eclipse expedition, and the subsequent
expository efforts by the ‘classical’ Oliver Lodge and ‘modern’ Arthur Stanley Eddington.
In doing so, I move beyond the classical–modern divide, and suggest a more substantial
role for the ether in professional and popular early twentieth-century British physics.

2.2 THE ETHER AND DISCONTINUITY, 1909–14


When thinking about the rapid changes that took place in physics in the early years of
the twentieth century, it seems as though a more helpful dichotomy than classical and
modern is that of continuity and discontinuity. The great continuous medium was under
threat by a new conception of nature as formed of discrete matter and energy. With
large changes potentially taking place in the discipline, the broader meaning of continu-
ity was also in play. Would physics progress in continuous steps, or through a discontinu-
ous shift resulting in the abandonment of long-held concepts? And within this, what
would be the fate of the ether? There was a continuum of positions on the nature and
existence of the ether, and a range of approaches to the role of the old physics in the
changing discipline.
16 The Ether at the Crossroads of Classical and Modern Physics

Perhaps the most vociferous British opponent of the ether in these early years was the
Cambridge-trained experimental physicist Norman Campbell. Campbell’s attitude
towards physics, developed while working in Thomson’s Cavendish from 1902, involved
the rejection of ideas that could not be experimentally verified.8 His 1907 textbook Modern
Electrical Theory characterised the Cambridge traditions of mathematical and experi-
mental physics as retrospectively ‘old physics’ and ‘new physics’.9 While mathematical
physicists clung to old ideas, hindering the prospect of progress, Campbell believed the
new experimentalists held no such deep commitments and were thus best placed to
move forward. Campbell became interested in relativity theory within the context of his
ongoing project to describe the electrodynamics of moving bodies without recourse to
either the ether or mathematics. In 1909 he wrote in the Proceedings of the Cambridge
Philosophical Society:

The trend of modern theory is everywhere to replace by discontinuity the continuity which
was the basis of the science of the last century. Any method which is especially applicable to
discontinuous processes is certain to be fruitful of results in every department of investiga-
tion, and any considerations which can be advanced in the elucidation of such a method are
not devoid of value.10

The idea of the physical world as fundamentally continuous, crucial from an ethereal
point of view, was further challenged in the coming years. In 1911, Ernest Rutherford
published a paper detailing his nuclear model of the atom, with the model consisting of
a central charge surrounded by a cloud of electrons. The world was now dramatically
filled with empty space, although it was entirely conceivable that the particles of matter
could still be connected by the ether. Two years later, however, Niels Bohr incorporated
the concept of discontinuous energy into Rutherford’s model by proposing that a small,
positively charged nucleus was orbited by electrons that could jump from a higher
energy orbit to a lower one, emitting a quantum of discrete energy as it did so. The con-
cept of quantum radiation had now been applied to a theory of atomic structure, com-
bining discontinuous matter and discontinuous energy.
Campbell praised Bohr’s theory in a 1914 Nature article on the structure of the atom,
supporting it on the basis of its accordance with experimental results. For Campbell,
Bohr’s theory represented a sharp divide between two groups of physicists:

There are only two alternatives open to the modern theoretical physicist: he may either sup-
pose that the principles of the older mechanics are true, and that all the brilliant results
which have followed from the application of the conceptions of Planck and Einstein to the
most diverse phenomena are illusory and devoid of evidential value; or he may suppose that
they are not true. Bohr’s theory offers him the choice in its most striking form.11

Campbell had also recently published a second version of his Modern Electrical Theory,
which again argued against the need for an ether in physical theories.12 This was reviewed
in Nature by the chemist and radioactivity researcher Frederick Soddy, who noted the
following:
The Ether and Discontinuity, 1909–14 17

Physical theories at the present moment are so shaky at the foundations that the doubt arises
sometimes whether the superstructure is not being built up too rapidly. The difficulties, now
ten years old, in reconciling the undulatory and corpuscular types of radiation in one theory,
the hopeless confusion that prevails as to the necessity for the existence of an ether, and the
modern discrete or quantum theory of energy, seem to call for a more drastic reconsideration
than we find here of many of the simplest physical conceptions and their experimental basis.13

For Soddy, Campbell’s rejection of many physical principles was still too conservative. As
a chemist, he held no deep commitment to physical theories and was instead concerned
with analysis of chemical properties and changes. He was also not one to bow down in
the face of tradition more generally, and campaigned for economic, social and institu-
tional reform in the 1920s and 1930s.14 There was perhaps here a connection between his
attitude towards rejection of long-held theories in physics, and rejection of tradition more
generally. For Soddy, where a system was no longer working, be it economical, social or
physical, the answer often lay in dramatic reform. Progress could not necessarily be
achieved through small measures but rather required profound conceptual shifts.
Of the many physicists who opposed such profound conceptual shifts regarding the
status of the ether, Lodge was perhaps the most vocal. He also benefited from a consid-
erable platform, as the president of the 1913 British Association for the Advancement of
Science. As Thomson had done in 1909, Lodge used the event as an opportunity to set
forth a defence of the ether, in front of a large audience in his home territory of Birmingham,
and an even larger readership of the subsequent published account.15 Speaking for an
hour and a half, Lodge criticised what he called ‘modern tendencies’ in science, including
the current ‘irresistible impulse to atomise everything’, which he saw as not only a prob-
lem in physics, with subatomic particles and quantum radiation, but also in biology, with
the emergence of Mendelian heredity. He also accused modern science of denying the
existence of anything which could not be readily sensed or measured. To make his point
absolutely clear, Lodge titled his talk ‘Continuity’, here referring to continuity of matter
and energy, but also continuity of thought, a link between the past and the present, a
physics that would hold onto the ether and Newton’s laws.16
The problem of continuity and discontinuity with regard to quantum developments
was further addressed two days later in a debate on radiation. James Jeans, a recent con-
vert to quantum theory, led the debate, framing the discussion in terms of continuity
versus discontinuity, and placing himself firmly on the side of discontinuity. He declared
that perhaps the ‘boldest and simplest attempt at reconciliation between the conflicting
theories lies in abandoning the ether altogether, and relying on some purely descriptive
principle, such as that of relativity’.17 This was also the view taken elsewhere by O. W.
Richardson, who concluded that, if relativity was indeed a ‘Universal Principle’, then it
followed that the ether was a ‘superfluous hypothesis’.18 Jeans was committed to a scien-
tific method that began with certain premises and from them deduced valid knowledge.19
He would not pragmatically adopt a theory simply on the basis of its current usefulness
in explaining experimental results. Thus, he was more than willing to discard the old,
providing the new had solid and viable foundations.
18 The Ether at the Crossroads of Classical and Modern Physics

Other British physicists in the room took an opposing view. The mathematician
Augustus Love refused to accept that ‘existing theories of dynamics and electrodynamics
need to be supplemented by the theory of the quanta’, instead proposing that recent
results could be interpreted within ‘ordinary theories’.20 There was a contribution from
Joseph Larmor, whose book Aether and Matter had been influential in early twentieth-
century Cambridge pedagogy.21 He commented on the new work relating to specific
heats at very low temperatures, suggesting that ‘there is nothing in it that is destructive
to the principles of physics which have led to so rich a harvest of discovery and synthesis
in the past’.22 He looked for reconciliation between the old and new ideas, continuing his
ongoing search for interactions between the ether and electrons. Furthermore, he used
the rhetoric of destruction, clearly placing the new ideas in the context of a dramatic
revolution, and emphasising the nature of discontinuity. Meanwhile Lodge stayed quiet,
but did invite the committee of Section A of the British Association to continue the
debate at his house the following Sunday.23
The discussion was written up in The Times by E. E. Fournier d’Albe, then an assistant
lecturer in physics at Birmingham.24 D’Albe was a member of the Society for Psychical
Research and had fairly radical views on what the ‘new’ physics could teach about the
nature of the soul. He believed that the discrete nature of matter, as revealed by the
electron, provided evidence for continuity of life after death.25 D’Albe was also deeply
committed to the ether, and had opposed relativity theory on the basis of a perceived
threat to it. He thus had much in common, intellectually and institutionally, with Lodge.
He did not, however, share Lodge’s opposition to quantum theory, hoping instead that
‘the investigation of this fascinating problem will teach us a great deal about the inter-
stellar aether which conveys the messages’.26
In The Times, D’Albe described the debate as a ‘pitched battle between the adherents
of the doctrines of Young and Fresnel, Maxwell and Hertz on the one hand, and the
revolutionary followers of Planck, Einstein, and Nernst on the other’. Ultimately, it was
an ‘old controversy’ between continuity and discontinuity. While the battle was still ‘rag-
ing’, d’Albe noted that opinion seemed to be in favour of the quantum theory. Detailing
the views of the opposition, he described Love as fighting ‘with conviction for the older
and more conservative view’, while Larmor was ‘somewhat pathetically seeking the way
of salvation through the falling debris of cherished views’.27 As d’Albe gleefully reported
on a battle between conservatism and revolution, he did not, despite being an adherent
to the ether, appear too threatened by the choice being made. Indeed, d’Albe, who omit-
ted a comment by James Jeans about abandoning a dynamical approach, did not neces-
sarily see a fundamental change occurring. With the ether and the quantum theory
entirely compatible in his understanding of physics, d’Albe could support both the ‘clas-
sical’ and the ‘modern’.
Not all ‘modern’ physicists were as embracing of change as d’Albe. A colleague of
both Lodge and d’Albe, Samuel McLaren was a graduate of Trinity College and worked
as a mathematics lecturer at Birmingham University from 1906 to 1913. McLaren was
keenly interested in the ‘new physics’, as he was able to grasp the complex mathematics
Eddington Versus Lodge: The Ether and Modern Physics 19

underpinning it, and also struck up a friendship with Niels Bohr during the Danish phys-
icist’s 1911 visit to England.28 McLaren was thus not unreceptive to new ideas, but he did
take issue with the more destructive consequences of some interpretations. Writing in
the Philosophical Magazine (for which Lodge was an editor), McLaren accused ‘Einstein’s
idea of the Quantum’ of being ‘destructive of the continuous medium and all that was
built upon it in the nineteenth century’. McLaren’s desire to retain the continuous
medium of the ether was more than simply a commitment to a physical principle. He
began his article by declaring that ‘the unrest of our time has invaded even the world of
Physics, where scarcely one of the principles long accepted as fundamental passes
unchallenged by all’. The problem was not simply the discontinuity of energy, but
rather the discontinuity of progress, of physics proceeding not by gradually building
upon the work of those who had gone before, but by tossing old theories aside and
replacing them with wildly different ones. And this predicament was not exclusive to
physics. Indeed, quite the opposite was true: McLaren believed that a more general
‘unrest of our time’ had infected physics, and referred to a ‘spirit of revolution’.29
Through McLaren and Lodge, we see a strong connection between discontinuous
matter and energy, and discontinuous progress. In addition, McLaren’s reference to the
‘spirit of revolution’ reveals a sense that physics was not alone in facing the challenge of
discontinuity. Indeed, the years surrounding 1913 saw equivalent developments in art and
literature, with cubist paintings and stream-of-consciousness novels breaking up time
and space. Here, as with science, long-standing authorities were being challenged (see
Chapter 11).30 And it was within this age of revolution that threats to the ether would
start to be more widely discussed. As the twentieth century advanced, conversations
about the ether, and about the nature of matter, the universe and physics more generally,
began to take place in more popular spaces. In British popular physics in the 1920s, two
key actors would appear: Arthur Stanley Eddington and Oliver Lodge. With these physi-
cists later coming to personify modern and classical physics, respectively, their shared
and divergent popular communications about the ether shed considerable light on the
nature of this apparent divide.

2.3 EDDINGTON VERSUS LODGE: THE ETHER


AND MODERN PHYSICS
In May 1919, two teams of astronomers travelled to the Brazilian municipality of Sobral
and the island of Principe, off the west coast of Africa, in order to measure the deflection
of starlight during an eclipse. As the well-known story goes, the astronomers set out to
obtain observational evidence of the general theory of relativity, proposed by Einstein in
1916. The results, in favour of Einstein’s theory, were presented at a joint meeting of the
Royal Society and Royal Astronomical Society in November 1919, widely reported in
the British press and led to increased popular interest in physics. The Cambridge astron-
omer Arthur Stanley Eddington, whose overwhelming support for Einstein’s theory has
20 The Ether at the Crossroads of Classical and Modern Physics

subsequently led to (now largely debunked) criticisms of his scientific objectivity, played
a major role in planning the expedition and carrying out the measurements, as well as
promoting the endeavour more widely.31 Eddington and other members of the Joint
Permanent Eclipse Committee ( JPEC) framed the expedition as a crucial experiment,
with a potential ‘trichotomy’ of results: a full deflection of 1.74″, confirming Einstein’s
predictions; a half-deflection, in line with Newton’s theory of gravitation; or no deflec-
tion at all, requiring an entirely new understanding altogether.32 Using connections at
The Times, the JPEC were able to lead a successful ‘publicity campaign’ during and after
the expedition, culminating in The Times’s announcement of the results under the head-
ing ‘Revolution in Science’.33
In the tale of the eclipse, Lodge plays a bit part, that of an antagonistic sceptic of
Einstein, one who storms out of the joint meeting when it becomes clear that Einstein
has emerged victorious. (Lodge subsequently maintained he had a ‘long-standing engage-
ment and a 6 o’clock train’.)34 I would argue, however, that Lodge played a significant
role in how both the expedition and physics were subsequently popularised. Take,
for example, one of the more esoteric newspaper accounts of the eclipse expedition,
published in the Aberdeen Press and Journal, a daily Scottish newspaper, on Tuesday,
11 November 1919. Headlined ‘A Great Discovery. May Reduce Physical Labour To A
Minimum’, the article announced a ‘sensational discovery’ that ‘will, it is believed, give
the clue to the nature of ether, and that, in turn, may provide the clue to the way of
utilising the tremendous energy of ether, which would give mankind control of forces
of power and energy undreamt of ’. Predicting a future with physical labour reduced to
a minimum, the author noted that Lodge, while urging caution against ‘generalising as
to the great possibilities founded upon the discovery’, had admitted that ‘a great and
splendid result has been achieved, capable of far-reaching possibilities’.35
It seems likely that the author of this report had paid less attention to the initial
announcement in The Times than to Lodge’s letter to the paper the following day. Under
the heading of ‘The Ether of Space. Sir Oliver Lodge’s Caution’, Lodge described the
eclipse result as ‘a great triumph for Einstein’, but issued ‘a caution against a strengthen-
ing of great and complicated generalisations concerning space and time on the strength
of the splendid result: I trust that it may be accounted for, with reasonable simplicity, in
terms of the ether of space’.36 Nowhere in this short letter did Lodge refer to hitherto
undreamt of stores of energy, but he had suggested this in an article in Philosophical
Magazine earlier that year.37
There was little mention of the ether in the promotional reports surrounding the
eclipse expedition. Eddington did not see general relativity as having implications for
the ether, but rather viewed the theory as an exciting development that would create
new opportunities for physical theories to be tested by astronomers and open further
avenues for research into the large-scale structure of the universe.38 But it took Lodge
little more than a single letter to The Times for the entire endeavour to be reframed (in
one Scottish newspaper anyway) as being about the ether (and about Lodge). And the
recently retired Lodge did not shy away from media attention in the subsequent years,
Eddington Versus Lodge: The Ether and Modern Physics 21

working to centre the ether in discussions about the new physics. In the magazine
Nineteenth Century and After, Lodge warned the physics world to not take the implica-
tions of the results too far and to not ‘be revolutionary to a rash and hasty extent’, argu-
ing for an interpretation of the new results in terms of a ‘generalisation’ of the old
theories, a generalisation that would hold on to the ether.39 Writing in the Fortnightly
Review in 1920, in an article unambiguously titled ‘The Ether Versus Relativity’, Lodge
insisted that his late nineteenth-century experiments needed to be repeated and that
‘extraordinary and expensive means’ were required to detect the extremely slow speed
of the ether stream. Lodge saw ‘no reason why a National Laboratory should not
undertake such an experiment’.40
Meanwhile Eddington was now in the position of trying to temper the narrative of
revolution that he had partially created, while also promoting the success of Einstein’s
theory. Writing in the Contemporary Review in late November, he mocked the hyperbolic
headlines that had appeared in newspapers: ‘REVOLUTION in Science—Newton and
Euclid Dethroned—Bending of Light—The Fourth Dimension—Warping of Space!’ He
accused such judgements of being perhaps ‘too hasty’, but admitted that the ‘fundamen-
tal nature of the change has not been exaggerated’.41 He also attempted to lay to rest any
claims of Newton’s overthrow, arguing that Newton had in fact predicted, in his Opticks,
that light could bend. Furthermore, he ended by declaring that it was ‘not necessary to
picture scientists as prostrated by the new revelations, feeling that they have got to go
back to the beginning and start again. The general course of experimental physics will
not be deflected, and only here and there will theory be touched.’42
By the time the 1920 British Association meeting came about, relativity theory had
been discussed at length. Perhaps as a result, Eddington used his presidential address to
Section A to talk about an entirely different topic, the internal constitution of stars.43
The astronomical journal, the Observatory, later lamented an absence of relativity theory
at the meeting, suggesting this was ‘because those chiefly concerned had become a little
jaded with the strenuous conflict’.44 Lodge, however, does not appear to have suffered
from this, and a large crowd gathered to hear him deliver a ‘Controversial Note on
Relativity’.45 Again admitting that Einstein’s equations were supported by experimental
observations, Lodge went on to argue that some interpretations of these equations were
‘threatening to land physicists in regions to which they had no right of entry’, into meta-
physical reasoning ‘beyond their ken’.46 He disagreed with any attempts to ‘build up on
an equation an elaborate metaphysical structure’, arguing that such equations were
open to numerous interpretations. Perhaps most damaging to those physicists who were
trying to disassociate themselves from revolution, Lodge ended his talk by suggesting
that relativists should perhaps ‘be regarded as Bolsheviks and pulled up’.47 The press duly
responded to Lodge’s successful courting of controversy, and the Bolshevik sound bite
was reported in The Guardian, the Daily Telegraph and the Daily News.48
Eddington did tackle relativity theory at the 1921 British Association meeting, in a
highly anticipated talk attended by apparently nearly 2,000 members of the association.49
The Times’s account of the event reported Eddington labelling the ether as an ‘idle
22 The Ether at the Crossroads of Classical and Modern Physics

hypothesis, unsupported by experiment and giving explanations of nothing’. The article


ended with Lodge’s contribution to the discussion:

Sir Oliver Lodge, proposing a vote of thanks to Professor Eddington for his delightful
address, said that he was not yet prepared to abandon the ether. Professor Eddington, he
added, was so immersed in the doctrine of relativity that he thought it was self-evident. He
was like the cricketer who on being asked to explain a ‘yorker’ retorted that it was just a
yorker. Einstein’s relativity theory was of great mathematical interest, but the general inter-
est taken in it throughout the world was entirely due to The Times.

As a result of Lodge’s final remark, this report was titled ‘Relativity or Ether?’ and sub-
titled ‘Sir Oliver Lodge and The Times’, allowing the ether enthusiast to hijack Eddington’s
address. With neither physicist presented as more authoritative than the other, the
framed question of ‘Relativity or Ether?’ had no clear answer in this report.
The Times was a long-standing faithful reporter of the annual British Association events,
and no doubt had sent a scientific reporter to cover this meeting, one already aware of
ongoing discussions relating to the impact of relativity theory on the ether. Local news-
papers, however, reported the event quite differently, with little mention of the continu-
ous medium. A variety of similar reports began by noting Eddington’s notoriety as the
only physicist who could intelligibly explain Einstein’s theory, and ended with a brief
mention of Lodge. However, Lodge’s explicit mention of the ether was absent in these
reports, as was Eddington’s clear dismissal of the medium. These newspapers referred to
Eddington’s suggestion that there were some physicists who believed velocity through
the ether did not exist, but this remark was not highlighted as important, nor accom-
panied by any suggestion that the ether itself was in question.50
Wherever Eddington could be found vigorously championing relativity theory, it seems
that Lodge was rarely far behind. A well-known public figure, even a minor contribution
from him was likely to be reported, injecting scepticism into any discussions of the new
physics. As a result, he was able to entirely reframe how lectures were reported, steering
the conversation onto the ether. In addition to ‘hijacking’ Eddington’s lectures, Lodge
was also starting his own conversations elsewhere, and was a prolific popular writer and
lecturer throughout the 1920s. In 1924, Ernest Benn published Lodge’s book Atoms and
Rays. Here, Lodge detailed ‘current’ knowledge about matter, discussing the structure
of the atom, quantum theory and the nature of energy. Throughout, he described the
ether as the fundamental ‘cementing substance’ that held everything together and was
responsible for the transmission of energy.51 In 1925, Lodge wrote the opening chapter
of Phases of Modern Science, a publication produced alongside the second showing of the
Royal Society’s ‘Pure Science’ exhibit at the 1924 British Empire Exhibition. In Lodge’s
chapter ‘Radiation’, he informed the reader of the ‘most usual view of radiation’ as
waves in the ether.52
That same year, Lodge delivered his sweeping account of science in the modern age,
in a series of talks on ‘Atoms and Worlds’, broadcast over the wireless in October and
November 1925. Subsequently published in Benn’s Sixpenny Library series as Modern
Eddington Versus Lodge: The Ether and Modern Physics 23

Scientific Ideas: Especially the Idea of Discontinuity, the book, resembling a small pamphlet,
came to only seventy-nine pages, a factor that, coupled with its affordable price, may
have attracted readers daunted by heftier tomes. Within the pages, Lodge put forward
his interpretation of modern science, which he saw as in flux. He made clear, in the
introduction, that ‘continuity remains the fundamental idea to which scientific philoso-
phy will in the last resort return’, but spent most of the book detailing the new and many
ways that matter and energy had been found to be discontinuous. Thus, ‘it must be
admitted that modern science is in a rather complicated, though very interesting, stage’.
Scientists had not yet ‘attained full knowledge’ and were ‘encountering a number of facts
the full explanation of which will need some generations of work on the part of our
leaders’.53 Lodge’s book was, in his mind, the presentation of a nascent and unfinished
subject, science so cutting edge that we did not yet know the answers.
Lodge’s views on the ether were not unchallenged by his peers. Atoms and Rays was
reviewed in The Observer by Edward Andrade, Professor of Physics at the Artillery
College in Woolwich. Andrade, who had written his own book on the Structure of Atom
(a treatise lacking any reference to the ether), advised his readers that Lodge was ‘rather
unorthodox . . . in his constant reference of everything back to the ether’. He remarked
that physicists had barely any knowledge at all about the ether, and knew simply that, as
Einstein had shown, it ‘has not got any mechanical properties, which rather spoils its useful-
ness’. However, overall, Andrade recommended Lodge’s book, providing the reader be
careful to differentiate ‘the certain’ from ‘the less certain’. He praised Lodge’s ‘freshness,
charm and polished simplicity of style’, and described the ‘great skill and enthusiasm’
with which Lodge discussed the quantum theory. 54
In Discovery, a popular science magazine, Atoms and Rays received a very positive
anonymous review, with the declaration that to ‘the student of Physics, as well as to every-
one who is interested in Physical Science, the appearance of a new publication by Sir Oliver
Lodge is always a memorable event’.55 Modern Scientific Ideas was also favourably reviewed,
in Nature, with reference to Andrade’s contribution to the same series, on The Atom, in which
the ether-sceptic scientist depicted nature as increasingly more atomic and discontinuous.56
The reviewer noted that the two books covered similar ground, and that it would thus be ‘of
considerable interest to note the varying manner of treatment of the same material by two
decidedly individualistic writers’.57 Notably, Andrade’s book was not depicted as being more
accurate, or even ‘better’, than Lodge’s, but simply different. In a review in Discovery, the two
books were presented as complementary.58 The author V. E. Pullin, Director of Radiological
Research at the Royal Arsenal in Woolwich (an institutional neighbour of Andrade’s Artillery
College), described Modern Scientific Ideas as ‘an excellent preamble’ to these more special-
ised books, providing an overview of modern physics. He declared that to ‘acclaim Sir Oliver
Lodge as an expounder of modern science would be to gild the lily’.59
Furthermore, the mere appearance of Lodge’s chapter in Phases of Modern Science was
a sign of tacit approval for his views, placing them at the very beginning of a reference
work accompanying an ambitious, state-funded celebration of the British Empire.60
Indeed, Lodge was very deliberately chosen for this role, appointed Vice-Chair of the
24 The Ether at the Crossroads of Classical and Modern Physics

organising committee (despite making it very clear he would be too busy to attend any
meetings) so that he could contribute to the book.61 Similarly, Lodge was invited to con-
tribute to Nature’s special relativity issue in 1921, appearing alongside such experts as
Eddington, Dyson, Crommelin, Jeans, Lorentz and even Einstein himself.62 He was
afforded an additional air of expertise by his acquaintanceship with biologist and science
journalist Peter Chalmers Mitchell, who wrote a weekly column in The Times on the
‘Progress of Science’. Mitchell, who corresponded frequently with Lodge on the subject
of spiritualism, evidently viewed the older physicist as a reliable authority on physics,
with Lodge’s name cropping up regularly in Mitchell’s column.63
Outside of The Times, smaller papers followed suit. Lodge’s 1923 Silvanus Thompson
memorial lecture, ‘The Origin or Basis of Wireless Communication’, was reported in
The Western Daily Press, with the header ‘The Ether Indispensable’. The paper quoted
Lodge’s warning not to be misled ‘by any apprehension of the theory of relativity into
supposing that that theory dispenses with the ether merely because it succeeds in ignor-
ing it’. Lodge went on to say that ‘the leaders in that theory were well aware that for
anything like a physical explanation of light or electricity or magnetism, or cohesion or
gravitation, the ether was indispensable’.64 In 1925, the Dundee Evening Telegraph reported
on Lodge’s radio talk ‘Vibrations and Waves and What They Signify’.65 Notably, the
chosen headline for this report was ‘Deep Secrets about Ether’; the existence of ether
was not questioned in the article. The Nottingham Evening Post reported Lodge’s 1928
Kelvin Lecture on ‘The Revolution in Physics’, in which Lodge noted that he personally
saw the current situation as the beginning of ‘the real theory of the ether’.66
While the lead-up to the 1920s began with concerted efforts by Eddington and others
to promote their expedition and the theory of relativity, which they believed it had decisively
proved, these efforts were somewhat derailed by a rogue Lodge. For, while Eddington was
often introduced as the primary British expert on Einstein, he did not usurp Lodge as the
expert on all physics, including (crucially) modern physics. Thus, Lodge was able to
repeatedly inject the ether into any conversation about relativity, before more widely
promoting his views on the place of the ether in modern physics. As a result, the ether
may have started the 1920s as a classical medium under threat of obsolescence at the
hands of modern physicists but, by the end of the decade, it could still be interpreted as
part of the future of the discipline. Lodge’s popularisation work was not viewed as a
desperate defence of classical physics but rather an authoritative account of the new.
The ether maintained a healthy popular existence, and a sense of relevance, up until the
end of the decade. And with Lodge’s voice among the most widely heard, the nuances
between different kinds of ethers were likely lost to a popular audience.

2.4 CONCLUSION
In 1928, Eddington’s Gifford Lecture of the previous year was published as the bestselling
Nature of the Physical World, which by 1944 had sold 80,000 copies worldwide.67 Now with
arguably a larger platform than even Lodge, Eddington was able to set forth his views on
Notes 25

the ether to a wide popular audience. His first chapter was entitled ‘The Downfall of
Classical Physics’, apparently the first attempt by a British scientist to actually define the
category of ‘classical physics’.68 Here he compared Einstein and Minkowski’s drastic
altering of our concept of time and space, and Rutherford’s nuclear model of the atom,
proposing that the latter was in fact the more dramatic. Eddington found it strange that
physicists would accuse relativists of Bolshevism (perhaps a reference to Lodge’s 1920 British
Association talk), but happily accept ‘the dissolution of all that we regard as most solid into
tiny specks floating in void’.69 Furthermore, this most revolutionary development was
entirely compatible with the concept of an ether, for, thanks to Rutherford, ‘we now realise
that the aether can slip through the atoms as easily as through the solar system’.70
While Eddington was ‘inclined to think that Rutherford, not Einstein, is the real villain
of the piece’, his definition of classical physics placed the former, not the latter, in the
category. However, Eddington tackled the impact of Einstein (and thus ‘modern phys-
ics’) on the ether in depth in a later discussion of relativity theory. He noted that with the
theory of relativity ‘evidently bound up with the impossibility of detecting absolute
velocity’, the ether was called into question, as motion with respect to the ether is
equivalent to absolute motion. Furthermore, with no ethereal frame having been found,
the notion of velocity through the ether is meaningless. However, he emphatically
stated: ‘This does not mean that the aether is abolished. We need an aether.’71
Without the ether, Eddington explained, the physical world would be analysed into
isolated particles of matter or electricity with featureless interspace. The role of the
ether was thus to provide this interspace with character, much as ‘we postulate matter or
electricity to bear the characters of the particles’. Abolishing the ether would be equiva-
lent to abolishing matter. Certainly, the ether could no longer be conceived of as a kind
of matter, but Eddington suggested that this view had ceased to be orthodox at some
point in the nineteenth century, and certainly before the advent of relativity theory. While
our conception of the ether had changed, in particular the idea of velocity through the
ether, the ether itself was ‘as much to the fore as ever it was’.72
Eddington’s ether was, of course, not the same ether as that proposed by Lodge. However,
the two men held a shared belief, that an ether played an important role in modern physics.
This they defended passionately to popular audiences, with the form of their ether differing
considerably, but the extent of their belief largely on a par. As with Lodge, Eddington’s ether
was vital in maintaining continuity in physics, providing a medium that connected together
the increasingly discrete building blocks of nature, while also forming a bridge between the
past, present and future of the discipline. Throughout the early twentieth century, adher-
ence to the ether displayed not a ‘classical’ but a continuous approach to physics.

NOTES
1. J. J. Thomson, ‘Address’, in Report of the Seventy-Ninth Meeting of the British Association for the
Advancement of Science (London: John Murray, 1910), 3–29, p. 15.
2. Lord Rayleigh, The Life of Sir J. J. Thomson (Cambridge: Cambridge University Press, 1942), p. 161.
26 The Ether at the Crossroads of Classical and Modern Physics

3. On Thomson and the Cavendish, see Jaume Navarro, A History of the Electron: J. J. and G. P.
Thomson (Cambridge: Cambridge University Press, 2012); Isobel Falconer, ‘J J Thomson and
“Cavendish” Physics’, in Frank A. J. L. James, ed., The Development of the Laboratory, (London:
Macmillan, 1989), 104–17; S. B. Sinclair, ‘J. J. Thomson and the Chemical Atom: From Ether
Vortex to Atomic Decay’, Ambix 34 (1987): 89–116.
4. George E. Smith, ‘J. J. Thomson and the Electron, 1897–1899’, in Jed Z. Buchwald and Andrew
Warwick, eds., Histories of the Electron: The Birth of Microphysics (Cambridge, MA: MIT Press,
2001), 21–76; Isobel Falconer, ‘Corpuscles, Electrons and Cathode Rays: J. J. Thomson and the
“Discovery of the Electron”’, The British Journal for the History of Science 20 (1987): 241–76.
5. See e.g. Helge Kragh, Quantum Generations (Princeton: Princeton University Press, 1999);
David M. Knight, ‘Classical Physics’, in Public Understanding of Science: A History of Communicating
Scientific Ideas (London: Taylor & Francis, 2006), 167–181; Jochen Büttner, Jürgen Renn and
Matthias Schemmel, ‘Exploring the Limits of Classical Physics: Planck, Einstein, and the
Structure of a Scientific Revolution’, Studies in the History and Philosophy of Modern Physics 34
(2003), 37–59, p. 37, in their response to Thomas S. Kuhn, Black-Body Theory and the Quantum
Discontinuity, 1894–1912 (Oxford: Oxford University Press, 1978).
6. On Thomson and the ether, see Navarro, History of the Electron.
7. Imogen Clarke, ‘Negotiating Progress: Promoting “Modern” Physics in Britain, 1900–1940’ (PhD
dissertation, University of Manchester, 2012); Richard Staley, ‘On the Co-Creation of Classical
and Modern Physics’, Isis 96 (2005): 530–58; Graeme Gooday and Daniel Mitchell, ‘Rethinking
Classical Physics’ in Jed Z. Buchwald and Robert Fox, eds., The Oxford Handbook of the History
of Physics (Oxford: Oxford University Press, 2013), 721–64.
8. AndAndrew Warwick, ‘Cambridge Mathematics and Cavendish Physics: Cunningham, Campbell
and Einstein's Relativity 1905–1911. Part II: Comparing Traditions in Cambridge Physics’, Studies in
History and Philosophy of Science 24 (1993): 1–25.
9. Norman Robert Campbell, Modern Electrical Theory (Cambridge: Cambridge University Press,
1907).
10. Robert Norman Campbell, ‘The Study of Discontinuous Phenomena’, Proceedings of the
Cambridge Philosophical Society 15 (1909): 117–36, p. 117.
11. Robert Norman Campbell, ‘The Structure of the Atom’, Nature 92 (1914): 586–7, p. 586.
12. Norman Robert Campbell, Modern Electrical Theory (Cambridge: Cambridge University Press,
1913).
13. Frederick Soddy, ‘Modern Physical Ideas and Researches’, Nature 92 (1913): 339–40, p. 339.
14. Jeff Hughes, ‘“Divine Right” or Democracy? The Royal Society “Revolt” of 1935’, Notes and
Records of the Royal Society 64, Supplement 1 (2010): S101–17. For Soddy in general, see Linda
Merricks, The World Made New: Frederick Soddy, Science, Politics, and Environment (Oxford: Oxford
University Press, 1996); George B. Kauffman, ed., Frederick Soddy (1877–1956): Early Pioneer in
Radiochemistry (Dordrecht; Lancaster: Reidel, 1986).
15. In addition to the official British Association report, Lodge’s talk was also published in book
form by J. M. Dent.
16. Oliver Lodge, ‘Continuity’, in Report of the Eighty-Third Meeting of the British Association for the
Advancement of Science (London: John Murray, 1914), 3–42.
17. ‘Friday, September 12. Discussion on Radiation’, in Report of the Eighty-Third Meeting, 376–86,
p. 380.
18. O. W. Richardson, The Electron Theory of Matter (Cambridge: Cambridge University Press,
1914), p. 325.
Notes 27

19. Matthew Stanley, ‘So Simple a Thing as a Star: The Eddington–Jeans Debate over Astrophysical
Phenomenology’, British Journal for the History of Science 40 (2007): 53–82.
20. ‘Friday, September 12’, pp. 383–4.
21. AndAndrew Warwick, Masters of Theory: Cambridge and the Rise of Mathematical Physics (Chicago:
University of Chicago Press, 2003).
22. ‘Friday, September 12’, p. 386.
23. British Association. ‘Problems of Radiation. Modern Universities and the State. Electric
Heating and Cooking’, The Times, 13 September 1913.
24. D’Albe wrote two anonymous reports of the discussion: ‘British Association. Improvement of
British Canals. Incubation of Eggs in Egypt’, The Times, 13 September 1913, and ‘British
Association. Problems of Radiation. Modern Universities and the State. Electric Heating and
Cooking’, The Times, 13 September 1913. In both cases, d’Albe wrote only the paragraphs relat-
ing to the radiation discussion (information from News International archivist).
25. Richard Noakes, ‘The “World of the Infinitely Little”: Connecting Physical and Psychical
Realities Circa 1900’, Studies in History and Philosophy of Science Part A 39 (2008): 323–34, p. 328.
26. D’Albe’s views on relativity theory and the quantum were expressed in Edmund Edward
Fournier d’Albe, ‘The Radiation Problem’, Nature 92 (1914): 689–91, p. 690.
27. ‘British Association. Improvement of British Canals. Incubation of Eggs in Egypt’.
28. McLaren is discussed in Alex Keller, ‘Continuity and Discontinuity in Early Twentieth-Century
Physics and Early Twentieth-Century Painting’, in M. Pollock, ed., Common Denominators in
Art and Science: The Proceedings of a Discussion Conference Held under the Auspices of the School of
Epistemics, University of Edinburgh, November 1981 (Aberdeen: Aberdeen University Press, 1983),
100–2.
29. Samuel B. McLaren, ‘The Theory of Radiation’, Philosophical Magazine (Series 6) 25 (1913):
43–56, p. 43.
30. Peter Gay, Modernism: The Lure of Heresy: From Baudelaire to Beckett and beyond (W. W. Norton,
New York, 2008); Robert Wohl, ‘Heart of Darkness: Modernism and its Historians’, Journal of
Modern History 74 (2002): 573–621; Christopher Butler, Early Modernism: Literature, Music, and
Painting in Europe, 1900–1916 (Clarendon Press, Oxford, 1994); Arthur I. Miller, Einstein, Picasso:
Space, Time and the Beauty that Causes Havoc (Basic Books, New York, 2002).
31. Alistair Sponsel, ‘Constructing a “Revolution in Science”: The Campaign to Promote a
Favourable Reception for the 1919 Solar Eclipse Experiments’, British Journal for the History of
Science 35 (2002): 439–67; Matthew Stanley, ‘“An Expedition To Heal the Wounds of War”: The
1919 Eclipse and Eddington as Quaker Adventurer’, Isis, 94 (2003): 57–89; John Earman and
Clark Glymour, ‘Relativity and Eclipses: The British Eclipse Expeditions of 1919 and Their
Predecessors’, Historical Studies in the Physical Sciences 11 (1980): 49–85.
32. The trichotomy is discussed most comprehensively in Earman and Glymour, ‘Relativity and
Eclipses’.
33. Sponsel, ‘Constructing a Revolution’; ‘Revolution in Science’, The Times, 7 November 1919.
While the ‘Revolution in Science’ article was not written or commissioned by the JPEC (and
was instead the work of biologist and science journalist Peter Chalmers Mitchell), it was the
indirect result of several months of promotion.
34. Oliver Lodge, ‘The Ether of Space’, The Times, 8 November 1919.
35. ‘A Great Discovery. May Reduce Physical Labor to a Minimum’, Aberdeen Press and Journal, 11
November 1919.
36. Oliver Lodge, ‘The Ether of Space. Sir Oliver Lodge’s Caution’, The Times, 8 November 1919.
Another random document with
no related content on Scribd:
“Oh, Bud! He’s such a delightful rascal. You don’t mind my calling
him that? I shouldn’t if I weren’t so fond of him. He’s absolutely
necessary to our social existence. We’d stagnate without him.”
“Bud was always a master hand at stirring things up. His methods
are a little peculiar at times, but he does get results.”
“There’s no question but that he’s a warm admirer of yours.”
“That’s because he’s forgotten about me! He hadn’t seen me for five
years.”
“I think possibly I can understand that one wouldn’t exactly forget
you, Mr. Storrs.”
She let the words fall carelessly, as though to minimize their daring
in case they were not wholly acceptable to her auditor. The point was
not lost upon him. He was not without his experience in the gentle art
of flirtation, and her technic was familiar. There was always,
however, the possibility of variations in the ancient game, and he
hoped that Mrs. Shepherd Mills was blessed with originality.
“There’s a good deal of me to forget; I’m six feet two!”
“Well, of course I wasn’t referring altogether to your size,” she said
with her murmurous little laugh. “I adore big men, and I suppose
that’s why I married a small one. Isn’t’ it deliciously funny how
contrary we are when it comes to the important affairs of our lives! I
suppose it’s just because we’re poor, weak humans. We haven’t the
courage of our prejudices.”
“I’d never thought of that,” Bruce replied. “But it is an interesting idea.
I suppose we’re none of us free agents. It’s not in the great design of
things that we shall walk a chalk line. If we all did, it would probably
be a very stupid world.”
“I’m glad you feel that way about it. For a long time half the world
tried to make conformists of the other half; nowadays not more than
a third are trying to keep the rest on the chalk line—and that third’s
skidding! People think me dreadfully heretical about everything. But
—I’m not, really! Tell me you don’t think me terribly wild and
untamed.”
“I think,” said Bruce, feeling that here was a cue he mustn’t miss, “I
think you are very charming. If it’s your ideas that make you so, I
certainly refuse to quarrel with them.”
“How beautifully you came up on that! Something tells me that I’m
not going to be disappointed in you. I have a vague sort of idea that
we’re going to understand each other.”
“You do me great honor! It will be a grief to me if we don’t.”
“It’s odd how instantly we recognize the signals when someone
really worth while swims into our ken,” she said pensively. “Dear old
Nature looks after that! Bud intimated that you’re to be one of us;
throw in your lot with those of us who struggle along in this rather
nice, comfortable town. If you enjoy grandeur in social things, you’ll
not find much here to interest you; but if just nice little companies
and a few friends are enough, you can probably keep amused.”
“If the Freemans’ friends are specimens and there’s much of this sort
of thing”—he waved his hand toward the company within—“I
certainly shall have nothing to complain of.”
“We must see you at our house. I haven’t quite Dale’s knack of
attracting people”—she paused a moment upon this note of humility
—“but I try to bring a few worth while people together. I’ve educated
a few men to drop in for tea on Thursdays with usually a few of my
pals among the young matrons and a girl or two. If you feel moved
——”
“I hope you’re not trifling with me,” said Bruce, “for I shall certainly
come.”
“Then that’s all settled. Don’t pay any attention to what Bud says
about me. To hear him talk you might think me a man-eater. My
husband’s the dearest thing! He doesn’t mind at all my having men
in for tea. He comes himself now and then when his business
doesn’t interfere. Dear Shep! He’s a slave to business, and he’s
always at work on some philanthropic scheme. I just talk about
helping the world; but he, poor dear, really tries to do something.”
Henderson appeared presently with a dark hint that Shepherd was
peeved by their long absence and that the company was breaking
up.
“Connie never plays all her cards the first time, Bruce; you must give
her another chance.”
“Oh, Mr. Storrs has promised me a thousand chances!” said Mrs.
Mills.
CHAPTER THREE
I
Sunday evening the Freemans were called unexpectedly into town
and Bruce and Henderson were left to amuse themselves.
Henderson immediately lost himself in a book and Bruce, a little
homesick for the old freedom of the road, set out for a walk. A
footpath that followed the river invited him and he lounged along, his
spirit responding to the beauty of the night, his mind intent upon the
future. The cordiality of the Freemans and their circle had impressed
him with the friendliness of the community. It would take time to
establish himself in his profession, but he had confidence in his
power to achieve; the lust for work was already strong in him. He
was satisfied that he had done wisely in obeying his mother’s
mandate; he would never have been happy if he had ignored it.
His meeting with Shepherd Mills had roused no resentment, revived
no such morbid thoughts as had troubled him on the night of his
arrival in town. Shepherd Mills was his half-brother; this, to be sure,
was rather staggering; but his reaction to the meeting was void of
bitterness. He speculated a good deal about young Mills. The
gentleness and forbearance with which he suffered the raillery of his
intimates, his anxiety to be accounted a good fellow, his serious
interest in matters of real importance—in all these things there was
something touching and appealing. It was difficult to correlate
Shepherd with his wife, but perhaps their dissimilarities were only
superficial. Bruce appraised Connie Mills as rather shallow, fond of
admiration, given to harmless poses in which her friends evidently
encouraged and indulged her. She practiced her little coquetries with
an openness that was in itself a safeguard. As they left the
Freemans, Shepherd and his wife had repeated their hope of seeing
him again. It was bewildering, but it had come about so naturally that
there seemed nothing extraordinary in the fact that he was already
acquainted with members of Franklin Mills’s family....
Bruce paused now and then where the path drew in close to the river
to look down at the moonlit water through the fringe of trees and
shrubbery. A boy and girl floated by in a canoe, the girl singing as
she thrummed a ukulele, and his eyes followed them a little wistfully.
Farther on the dull put-put-put of a motor-boat broke the silence. The
sound ceased abruptly, followed instantly by a colloquy between the
occupants.
“Damn this fool thing!” ejaculated a feminine voice. “We’re stuck!”
“I had noticed it!” said another girl’s voice good naturedly. “But such
is the life of the sailor. I wouldn’t just choose this for an all-night
camp!”
“Don’t be so sweet about it, Millicent! I’d like to sink this boat.”
“It isn’t Polly’s fault. She’s already half-buried in the sand,” laughed
the other.
Bruce scrambled down to the water’s edge and peered out upon the
river. A small power boat had grounded on a sandbar in the middle
of the stream. Its occupants were two young women in bathing suits.
But for their voices he would have taken them for boys. One was
tinkering with the engine while the other was trying to push off the
boat with an oar which sank ineffectually in the sand. In their
attempts to float their craft the young women had not seen Bruce,
who, satisfied that they were in no danger, was rather amused by
their plight. They were presumably from one of the near-by villas and
their bathing suits implied familiarity with the water. The girl at the
engine talked excitedly with an occasional profane outburst; her
companion was disposed to accept the situation philosophically.
“We can easily swim out, so don’t get so excited, Leila,” said the girl
with the oar. “And do stop swearing; voices travel a long way over
the water.”
“I don’t care who hears me,” said the other, though in a lower tone.
She gave the engine a spin, starting the motor, but the power was
unequal to the task of freeing the boat. With an exclamation of
disgust she turned off the switch and the futile threshing of the
propeller ceased.
“Let’s swim ashore and send back for Polly,” said the girl addressed
as Millicent.
“I see myself swimming out!” the other retorted. “I’m not going to
leave Polly here for some pirate to steal.”
“Nobody’s going to steal her. This isn’t the ocean, you know.”
“Well, no fool boat’s going to get the best of me! Where’s that flask?
I’m freezing!”
“You don’t need any more of that! Please give it to me!”
“I hope you are enjoying yourself,” said the other petulantly. “I don’t
see any fun in this!”
“Hello, there!” called Bruce, waving his arms to attract their attention.
“Can I be of help?”
Startled by his voice, they did not reply immediately, but he heard
them conferring as to this unlooked-for hail from the bank.
“Oh, I’m perfectly harmless!” he cried reassuringly. “I was just
passing and heard your engine. If there’s a boat near by I can pull
you off, or I’ll swim out and lift your boat off if you say so.”
“Better get a boat,” said the voice he had identified with the name of
Millicent. “There’s a boathouse just a little farther up, on your side.
You’ll find a skiff and a canoe. We’ll be awfully glad to have your
help. Thank you ever so much!”
“Don’t forget to come back,” cried Leila.
“Certainly not!” laughed Bruce and sprang up the bank.
He found the boathouse without trouble, chose the skiff as easier to
manage, and rowed back. In the moonlight he saw Millicent standing
up in the launch watching him, and as he approached she flashed an
electric torch along the side of the boat that he might see the nature
of their difficulty.
“Do you need food or medical attention?” he asked cheerfully as he
skillfully maneuvered the skiff and grounded it on the sand.
“I think we’d better get out,” she said.
“No; stay right there till I see what I can do. I think I can push you off.
All steady now!”
The launch moved a little at his first attempt to dislodge it and a
second strong shove sent it into the channel.
“Now start your engine!” he commanded.
The girl in the middle of the boat muttered something he didn’t catch.
“Leila, can you start the engine?” demanded Millicent. “I think—I
think I’ll have to row back,” she said when Leila made no response.
“My friend isn’t feeling well.”
“I’ll tow you—that’s easy,” said Bruce, noting that her companion
apparently was no longer interested in the proceedings. “Please
throw me your rope!”
He caught the rope and fastened it to the stern of the skiff and called
out that he was ready.
“Please land us where you found the boat,” said Millicent. She
settled herself in the stern of the launch and took the tiller. No word
was spoken till they reached the boathouse.
“That’s all you can do,” said Millicent, who had drawn on a long bath
wrapper and stepped out. “And thank you very, very much; I’m sorry
to have caused you so much trouble.”
This was clearly a dismissal, but he loosened the rope and tied up
the skiff. He waited, holding the launch, while Millicent tried to
persuade Leila to disembark.
“Perhaps——” began Bruce, and hesitated. It seemed unfair to leave
the girl alone with the problem of getting her friend ashore. Not to put
too fine a point on the matter, Leila was intoxicated.
“Now, Leila!” cried Millicent exasperatedly. “You’re making yourself
ridiculous, besides keeping this gentleman waiting. It’s not a bit nice
of you!”
“Jus’ restin’ lil bit,” said Leila indifferently. “I’m jus’ restin’ and I’m not
goin’ to leave Polly. I should shay not!”
And in assertion of her independence she began to whistle. She
seemed greatly amused that her attempts to whistle were
unsuccessful.
Millicent turned to Bruce. “If I could get her out of the boat I could put
her in our car and take her home.”
“Surely!” he said and bent over quickly and lifted the girl from the
launch, set her on her feet and steadied her. Millicent fumbled in the
launch, found a bath wrapper and flung it about Leila’s shoulders.
She guided her friend toward the long, low boathouse and turned a
switch.
“I can manage now,” she said, gravely surveying Bruce in the glare
of light. “I’m so sorry to have troubled you.”
She was tall and fair with markedly handsome brown eyes and a
great wealth of fine-spun golden hair that escaped from her bathing
cap and tumbled down upon her shoulders. Her dignity was in
nowise diminished by her garb. She betrayed no agitation. Bruce felt
that she was paying him the compliment of assuming that she was
dealing with a gentleman who, having performed a service, would go
his way and forget the whole affair. She drew her arm about the now
passive Leila, who was much shorter—quite small, indeed, in
comparison.
“Our car’s here and we’ll get dressed and drive back into town.
Thank you so much and—good-night!”
“I was glad to help you;—good-night!”
The door closed upon them. Bruce made the launch fast to the
landing and resumed his walk.

II
When he returned to the Freemans, Henderson flung aside his book
and complained of Bruce’s prolonged absence. “I had begun to think
you’d got yourself kidnapped. Go ahead and talk,” he said, yawning
and stretching himself.
“Well, I’ve had a mild adventure,” said Bruce, lighting a cigarette; and
he described his meeting with the two young women.
“Not so bad!” remarked Henderson placidly. “Such little adventures
never happen to me. The incident would make good first page stuff
for a newspaper; society girls shipwrecked. You ought to have taken
the flask as a souvenir. Leila is an obstreperous little kid; she really
ought to behave herself. Right the first time. Leila Mills, of course; I
think I mentioned her the other day. Her friend is Millicent Harden.
Guess I omitted Millicent in my review of our citizens. Quite a
remarkable person. She plays the rôle of big sister to Leila; they’re
neighbors on Jefferson Avenue. That’s just a boathouse on the Styx
that Mills built for Leila’s delectation. She pulls a cocktail tea there
occasionally. Millicent’s pop made a fortune out of an asthma cure—
the joy of all cut-rate druggists. Not viewed with approval by medical
societies. Socially the senior Hardens are outside the breastworks,
but Millicent is asked to very large functions, where nobody knows
who’s there. They live in that whopping big house just north of the
Mills place, and old Doc Harden gives Millicent everything she
wants. Hence a grand organ, and the girl is a regular Cecelia at the
keys. Really plays. Strong artistic bent. We can’t account for people
like the Hardens having such a daughter. There’s a Celtic streak in
the girl, I surmise—that odd sort of poetic strain that’s so beguiling in
the Irish. She models quite wonderfully, they tell me. Well, well! So
you were our little hero on the spot!”
“But Leila?” said Bruce seriously. “You don’t quite expect to find the
daughter of a prominent citizen tipsy on a river, and rather profane at
that.”
“Oh, thunder!” exclaimed Henderson easily. “Leila’s all right. You
needn’t worry about her. She’s merely passing through a phase and
will probably emerge safely. Leila’s hardly up to your standard, but
Millicent is a girl you’ll like. I ought to have told Dale to ask Millicent
here. Dale’s a broad-minded woman and doesn’t mind it at all that
old Harden’s rolled up a million by being smart enough to scamper
just a nose length ahead of the Federal grand jury carrying his rotten
dope in triumph.”
“Miss Mills, I suppose, is an acceptable member of the Freemans’
group?” Bruce inquired.
“Acceptable enough, but this is all too tame for Leila. Curious sort of
friendship—Leila and Millicent. Socially Millicent is, in a manner of
speaking, between the devil and the deep sea. She’s just a little too
superior to train with the girls of the Longview Country Club set and
the asthma cure keeps her from being chummy with the Faraway
gang. But I’ll say that Leila’s lucky to have a friend like Millicent.”
“Um—yes,” Bruce assented. “I’m beginning to see that your social
life here has a real flavor.”
“Well, it’s not all just plain vanilla,” Bud agreed with a yawn.
CHAPTER FOUR
I
Henderson made his wife’s return an excuse for giving a party at the
Faraway Country Club. Mrs. Henderson had brought home a trophy
from the golf tournament and her prowess must be celebrated. She
was a tall blonde with a hearty, off-hand manner, and given to plain,
direct speech. She treated Bud as though he were a younger
brother, to be humored to a certain point and then reminded a little
tartly of the limitations of her tolerance.
When Bruce arrived at the club he found his hostess and Mrs.
Freeman receiving the guests in the hall and directing them to a dark
end of the veranda where Bud was holding forth with a cocktail-
shaker. Obedient to their hint, he stumbled over the veranda chairs
until he came upon a group of young people gathered about Bud,
who was energetically compounding drinks as he told a story. Bruce
knew the story; it was the oldest of Bud’s yarns, and his interest
wavered to become fixed immediately upon a girl beside him who
was giving Bud her complete attention. Even in the dim light of the
veranda there was no mistaking her: she was the Millicent Harden
he had rescued from the sand bar. At the conclusion of the story she
joined in the general laugh and turned round to find Bruce regarding
her intently.
“I beg your pardon,” he said and bowed gravely.
“Oh, you needn’t!” she replied quickly.
He lifted his head to find her inspecting him with an amused smile.
“I might find someone to introduce us—Mr. Henderson, perhaps,” he
said. “My name—if the matter is important—is Bruce Storrs.”
“Possibly we might complete the introduction unassisted—my name
is Millicent Harden!”
“How delightful! Shall we dance?”
After the dance he suggested that they step out for a breath of air.
They found seats and she said immediately:
“Of course I remember you; I’d be ashamed if I didn’t. I’m glad of this
chance to thank you. I know Leila—Miss Mills—will want to thank
you, too. We must have seemed very silly that night on the river.”
“Such a thing might happen to anyone; why not forget it?”
“Let me thank you again,” she said seriously. “You were ever so
kind.”
“The incident is closed,” he remarked with finality. “Am I keeping you
from a partner? They’re dancing again. We might sit this out if I’m
not depriving you——”
“You’re not. It’s warm inside and this is a relief. We might even
wander down the lawn and look for elves and dryads and nymphs.
Those big trees and the stars set the stage for such encounters.”
“It’s rather nice to believe in fairies and such things. At times I’m a
believer; then I lose my faith.”
“We all forget our fairies sometimes,” she answered gravely.
He had failed to note at their meeting on the river the loveliness of
her voice. He found himself waiting for the recurrence of certain
tones that had a curious musical resonance. He was struck by a
certain gravity in her that was expressed for fleeting moments in both
voice and eyes. Even with the newest dance music floating out to
them and the light and laughter within, he was aware of an
indefinable quality in the girl that seemed somehow to translate her
to remote and shadowy times. Her profile—clean-cut without
sharpness—and her manner of wearing her abundant hair—carried
back loosely to a knot low on her head—strengthened his impression
of her as being a little foreign to the place and hour. She spoke with
quiet enthusiasm of the outdoor sports that interested her—riding
she enjoyed most of all. Henderson had intimated that her social life
was restricted, but she bore herself more like a young woman of the
world than any other girl he remembered.
“Maybelle Henderson will scold me for hiding you away,” she said.
“But I just can’t dance whenever the band plays. It’s got to be an
inspiration!”
“Then I thank you again for one perfect dance! I’m afraid I didn’t
appreciate what you were giving me.”
“Oh, I danced with you to hide my embarrassment!” she laughed.
Half an hour passed and they had touched and dismissed many
subjects when she rose and caught the hand of a girl who was
passing.
“Miss Mills, Mr. Storrs. It’s quite fitting that you should meet Mr.
Storrs.”
“Fitting?” asked the girl, breathless from her dance.
“We’ve all met before—on the river—most shockingly! You might just
say thank you to Mr. Storrs.”
“Oh, this is not——” Leila drew back and inspected Bruce with a
direct, candid gaze.
“Miss Harden is mistaken; this is the first time we ever met,” declared
Bruce.
“Isn’t he nice!” Leila exclaimed. “From what Millie said I knew you
would be like this.” And then: “Oh, lots of people are bragging about
you and promising to introduce me! Here comes Tommy Barnes; he
has this dance. Oh, Millie! if you get a chance you might say a kind
word to papa. He’s probably terribly bored by this time.”
“Leila’s a dear child! I’m sure you’ll like her,” said Millicent as the girl
fluttered away. “Oh, I adore this piece! Will you dance with me?”
As they finished the dance Mrs. Henderson intercepted them.
“Aren’t you the limit, you two? I’ve had Bud searching the whole
place for you and here you are! Quite as though you hadn’t been
hiding for the last hour.”
“I’m going to keep Mr. Storrs just a moment longer,” said Millicent.
“Leila said her father was perishing somewhere and I want Mr. Storrs
to meet him.”
“Yes; certainly,” said Bruce.
He walked beside her into the big lounge, where many of the older
guests were gathered.
“Poor Mr. Mills!” said Millicent after a quick survey of the room.
“There he is, listening to one of Mr. Tasker’s interminable yarns.”
She led the way toward a group of men, one of whom was evidently
nearing the end of a long story. One of his auditors, a dark man of
medium height and rather stockily built, was listening with an air of
forced attention. His grayish hair was brushed smoothly away from a
broad forehead, his neatly trimmed mustache was a trifle grayer than
his hair. Millicent and Bruce fell within the line of his vision, and his
face brightened instantly as he nodded to the girl and waved his
hand. The moment the story was ended he crossed to them, his
eyes bright with pleasure and a smile on his face.
“I call it a base desertion!” he exclaimed. “Leila brings me here and
coolly parks me. A father gets mighty little consideration these days!”
“Don’t scold! Mr. Mills—let me present Mr. Storrs.”
“I’m very glad to meet you, Mr. Storrs,” said Mills with quiet cordiality.
He swept Bruce with a quick, comprehensive scrutiny.
“Mr. Storrs has lately moved here,” Millicent explained.
“I congratulate you, Mr. Storrs, on having fallen into good hands.”
“Oh, Miss Harden is taking splendid care of me!” Bruce replied.
“She’s quite capable of doing that!” Mills returned.
Bruce was studying Franklin Mills guardedly. A man of reserves and
reticences, not a safe subject for quick judgments. His manner was
somewhat listless now that the introduction had been accomplished;
and perhaps aware of this, he addressed several remarks to Bruce,
asking whether the music was all that the jazzy age demanded;
confessed with mock chagrin that his dancing days were over.
“You only think they are! Mr. Mills really dances very well. You’d be
surprised, Mr. Storrs, considering how venerable he is!”
“That’s why I don’t dance!” Mills retorted with a rueful grin.
“‘Considering his age’ is the meanest phrase that can be applied to a
man of fifty.”
Bud Henderson here interrupted them, declaring that dozens of
people were disconsolate because Bruce had concealed himself.
“Of course you must go!” said Millicent.
“I hope to meet you again,” Mills remarked as Bruce bowed to him.
“Thank you, Mr. Mills,” said Bruce.
He was conscious once more of Mills’s intent scrutiny. It seemed to
him as he walked away that Mills’s eyes followed him.
“What’s the matter, old top?” Bud demanded. “You’re not tired?”
“No; I’m all right,” Bruce replied, though his heart was pounding hard;
and feeling a little giddy, he laid his hand on Henderson’s arm.
CHAPTER FIVE
I
Franklin Mills stood by one of the broad windows in his private office
gazing across the smoky industrial district of his native city. With his
hands thrust into his trousers’ pockets, he was a picture of negligent
ease. His face was singularly free of the markings of time. His thick,
neatly trimmed hair with its even intermixture of white added to his
look of distinction. His business suit of dark blue with an obscure
green stripe was evidently a recent creation of his tailor, and a wing
collar with a neatly tied polka-dot cravat contributed further to the
impression he gave of a man who had a care for his appearance.
The gray eyes that looked out over the city narrowed occasionally as
some object roused his attention—a freight train crawling on the
outskirts or some disturbance in the street below. Then he would
resume his reverie as though enjoying his sense of immunity from
the fret and jar of the world about him.
Bruce Storrs. The name of the young man he had met at the Country
Club lingered disturbingly in his memory. He had heard someone ask
that night where Storrs came from, and Bud Henderson, his sponsor,
had been ready with the answer, “Laconia, Ohio.” Mills had been
afraid to ask the question himself. Long-closed doors swung open
slowly along the dim corridor of memory and phantom shapes
emerged—among them a figure Franklin Mills recognized as himself.
Swiftly he computed the number of years that had passed since, in
his young manhood, he had spent a summer in the pleasant little
town, sent there by his father to act as auditor of a manufacturing
concern in which Franklin Mills III for a time owned an interest.
Marian Storrs was a lovely young being—vivacious, daring, already
indifferent to the man to whom she had been married two years....
He had been a beast to take advantage of her, to accept all that she
had yielded to him with a completeness and passion that touched
him poignantly now as she lived again in his memory.... Was this
young man, Bruce Storrs, her son? He was a splendid specimen,
distinctly handsome, with the air of breeding that Mills valued. He
turned from the window and walked idly about the room, only to
return to his contemplation of the hazy distances.
The respect of his fellow man, one could see, meant much to him.
He was Franklin Mills, the fourth of the name in succession in the
Mid-western city, enjoying an unassailable social position and able to
command more cash at a given moment than any other man in the
community. Nothing was so precious to Franklin Mills as his peace of
mind, and here was a problem that might forever menace that
peace. The hope that the young man himself knew nothing did not
abate the hateful, hideous question ... was he John Storrs’s son or
his own? Surely Marian Storrs could not have told the boy of that old
episode....
Nearly every piece of property in the city’s original mile square had
at some time belonged to a Mills. The earlier men of the name had
been prominent in public affairs, but he had never been interested in
politics and he never served on those bothersome committees that
promote noble causes and pursue the public with subscription
papers. When Franklin Mills gave he gave liberally, but he preferred
to make his contributions unsolicited. It pleased him to be
represented at the State Fair with cattle and saddle horses from
Deer Trail Farm. Like his father and grandfather, he kept in touch
with the soil, and his farm, fifteen miles from his office, was a show
place; his Jersey herd enjoyed a wide reputation. The farm was as
perfectly managed as his house and office. Its carefully tended
fields, his flocks and herds and the dignified Southern Colonial
house were but another advertisement of his substantial character
and the century-long identification of his name with the State.
His private office was so furnished as to look as little as possible like
a place for the transaction of business. There were easy lounging
chairs, a long leathern couch, a bookcase, a taboret with cigars and
cigarettes. The flat-top desk, placed between two windows,
contained nothing but an immaculate blotter and a silver desk set
that evidently enjoyed frequent burnishing. It was possible for him to
come and go without traversing the other rooms of the suite. Visitors
who passed the office boy’s inspection and satisfied a prim
stenographer that their errands were not frivolous found themselves
in communication with Arthur Carroll, Mills’s secretary, a young man
of thirty-five, trained as a lawyer, who spoke for his employer in all
matters not demanding decisions of first importance. Carroll was not
only Mills’s confidential man of business, but when necessary he
performed the duties of social secretary. He was tactful, socially in
demand as an eligible bachelor, and endowed with a genius for
collecting information that greatly assisted Mills in keeping in touch
with the affairs of the community.
Mills glanced at his watch and turned to press a button in a plate on
the corner of his desk. Carroll appeared immediately.
“You said Shep was coming?” Mills inquired.
“Yes; he was to be here at five, but said he might be a little late.”
Mills nodded, asked a question about the survey of some land
adjoining Deer Trail Farm for which he was negotiating, and listened
attentively while Carroll described a discrepancy in the boundary
lines.
“Is that all that stands in the way?” Mills asked.
“Well,” said Carroll, “Parsons shows signs of bucking. He’s thought
of reasons, sentimental ones, for not selling. He and his wife moved
there when they were first married and their children were all born on
the place.”
“Of course we have nothing to do with that,” remarked Mills, slipping
an ivory paper knife slowly through his fingers. “The old man is a
failure, and the whole place is badly run down. I really need it for
pasture.”
“Oh, he’ll sell! We just have to be a little patient,” Carroll replied.
“All right, but don’t close till the title’s cleared up. I don’t buy law
suits. Come in, Shep.”
Shepherd Mills had appeared at the door during this talk. His father
had merely glanced at him, and Shepherd waited, hat in hand, his
topcoat on his arm, till the discussion was ended.
“What’s that you’ve got there?” his father asked, seating himself in a
comfortable chair a little way from the desk.
In drawing some papers from the pocket of his overcoat, Shepherd
dropped his hat, picked it up and laid it on the desk. He was trying to
appear at ease, and replied that it was a contract calling for a large
order which the storage battery company had just made.
“We worked a good while to get that,” said the young man with a ring
of pride in his voice. “I thought you’d like to know it’s all settled.”
Mills put on his glasses, scanned the document with a practiced eye
and handed it back.
“That’s good. You’re running full capacity now?”
“Yes; we’ve got orders enough to keep us going full handed for
several months.”
The young man’s tone was eager; he was clearly anxious for his
father’s approval. He had expected a little more praise for his
success in getting the contract, but was trying to adjust himself to his
father’s calm acceptance of the matter. He drummed the edge of the
desk as he recited certain figures as to conditions at the plant. His
father disconcertingly corrected one of his statements.
“Yes; you’re right, father,” Shepherd stammered. “I got the July
figures mixed up with the June report.”
Mills smiled indulgently; took a cigarette from a silver box on the
taboret beside him and unhurriedly lighted it.
“You and Constance are coming over for dinner tonight?” he asked.
“I think Leila said she’d asked you.”
His senior’s very calmness seemed to add to Shepherd’s
nervousness. He rose and laid his overcoat on the couch, drew out
his handkerchief and wiped his forehead, remarking that it was warm
for the season.
“I hadn’t noticed it,” his father remarked in the tone of one who is
indifferent to changes of temperature.
“There’s a little matter I’ve been wanting to speak to you about,”
Shepherd began. “I thought it would be better to mention it here—
you never like talking business at the house. If it’s going to be done it
ought to be started now, before the bad weather sets in.”
He paused, a little breathless, and Mills said, the least bit impatiently:
“Do you mean that new unit at the plant? I thought we’d settled that. I
thought you were satisfied you could get along this winter with the
plant as it is.”
“Oh, no! It’s not that!” Shepherd hastily corrected. “Of course that’s
all settled. This is quite a different matter. I only want to suggest it
now so you can think it over. You see, our employees were all
mightily pleased because you let them have the use of the Milton
farm. There’s quite a settlement grown up around the plant and the
Milton land is so near they can walk to it. I’ve kept tab this summer
and about a hundred of the men go there Saturday afternoons and
Sundays; mostly married men who take their families. I could see it
made a big difference in the morale of the shop.”
He paused to watch the effect of his statements, but Mills made no
sign. He merely recrossed his legs, knocked the ash from his
cigarette and nodded for his son to go on.
“I want you to know I appreciate your letting me use the property that
way,” Shepherd resumed. “I was out there a good deal myself, and
those people certainly enjoyed themselves. Now what’s in my mind
is this, father”—he paused an instant and bent forward with boyish
eagerness—“I’ve heard you say you didn’t mean to sell any lots in
the Milton addition for several years—not until the street car line’s
extended—and I thought since the factory’s so close to the farm, we
might build some kind of a clubhouse the people could use the year
round. They can’t get any amusements without coming into town,
and we could build the house near the south gate of the property,
where our people could get to it easily. They could have dances and
motion pictures, and maybe a few lectures and some concerts,
during the winter. They’ll attend to all that themselves. Please
understand that I don’t mean this as a permanent thing. The
clubhouse needn’t cost much, so when you get ready to divide the

You might also like