Ebook A Philosopher Looks at Science 1St Edition Nancy Cartwright Online PDF All Chapter

You might also like

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

A Philosopher Looks at Science 1st

Edition Nancy Cartwright


Visit to download the full and correct content document:
https://ebookmeta.com/product/a-philosopher-looks-at-science-1st-edition-nancy-cart
wright/
More products digital (pdf, epub, mobi) instant
download maybe you interests ...

Ashoka: Portrait of a Philosopher King Olivelle

https://ebookmeta.com/product/ashoka-portrait-of-a-philosopher-
king-olivelle/

Feathertide Beth Cartwright

https://ebookmeta.com/product/feathertide-beth-cartwright/

Following the Drum Women at the Valley Forge Encampment


1st Edition Nancy K Loane

https://ebookmeta.com/product/following-the-drum-women-at-the-
valley-forge-encampment-1st-edition-nancy-k-loane/

Dance With Death 1st Edition Jack Cartwright

https://ebookmeta.com/product/dance-with-death-1st-edition-jack-
cartwright/
Lie Beside Me 1st Edition Jack Cartwright

https://ebookmeta.com/product/lie-beside-me-1st-edition-jack-
cartwright/

The Vostok Enigma 1st Edition Christopher Cartwright

https://ebookmeta.com/product/the-vostok-enigma-1st-edition-
christopher-cartwright/

Her Dying Mind 1st Edition Jack Cartwright

https://ebookmeta.com/product/her-dying-mind-1st-edition-jack-
cartwright/

Making Faithful Decisions at the End of Life 3rd


Edition Nancy J Duff

https://ebookmeta.com/product/making-faithful-decisions-at-the-
end-of-life-3rd-edition-nancy-j-duff/

At the Edge of the Rift 1st Edition Sue Gregory Paul


Jerry Nancy Tavares Jones

https://ebookmeta.com/product/at-the-edge-of-the-rift-1st-
edition-sue-gregory-paul-jerry-nancy-tavares-jones/
A PHILOSOPHER LOOKS AT SCIENCE

What is science and what can it do? Nancy Cartwright here takes
issue with three common images of science: that it amounts to the
combination of theory and experiment; that all science is basically
reducible to physics; and that science and the natural world which
it pictures are deterministic. The author’s innovative and thought-
ful book draws on examples from the physical, life and social
sciences alike, and focuses on all the products of science – not
just experiments or theories – and how they work together. She
reveals just what it is that makes science ultimately reliable and
how this reliability is nevertheless still compatible with a view of
nature as more responsive to human change than we might think.
Her book is a call for greater intellectual humility by and within
scientific institutions. It will have strong appeal to anyone who
thinks about science and how it is practised in society.

nancy cartwright is Professor of Philosophy at Durham


University and Distinguished Professor of Philosophy at the
University of California, San Diego. Her publications include
The Dappled World: A Study of the Boundaries of Science
(Cambridge, 1999), Hunting Causes and Using Them: Approaches
in Philosophy and Economics (Cambridge, 2007) and Nature, the
Artful Modeler (2019).
A Philosopher Looks at

In this series, philosophers offer a personal and philosophical


exploration of a topic of general interest.

Books in the series

Nancy Cartwright, A Philosopher Looks at Science


Raymond Geuss, A Philosopher Looks at Work
Paul Guyer, A Philosopher Looks at Architecture
Stephen Mumford, A Philosopher Looks at Sport
Onora O’Neill, A Philosopher Looks at Digital Communication
Michael Ruse, A Philosopher Looks at Human Beings
A PHILOSOPHER LOOKS AT

SCIENCE
NANCY CARTWRIGHT
University Printing House, Cambridge cb2 8bs, United Kingdom
One Liberty Plaza, 20th Floor, New York, ny 10006, USA
477 Williamstown Road, Port Melbourne, vic 3207, Australia
314–321, 3rd Floor, Plot 3, Splendor Forum, Jasola District Centre,
New Delhi – 110025, India
103 Penang Road, #05–06/07, Visioncrest Commercial, Singapore 238467

Cambridge University Press is part of the University of Cambridge.


It furthers the University’s mission by disseminating knowledge in the pursuit of
education, learning, and research at the highest international levels of excellence.

www.cambridge.org
Information on this title: www.cambridge.org/9781009201889
doi: 10.1017/9781009201896
© Nancy Cartwright 2022
This publication is in copyright. Subject to statutory exception
and to the provisions of relevant collective licensing agreements,
no reproduction of any part may take place without the written
permission of Cambridge University Press.
First published 2022
Printed in the United Kingdom by TJ Books Limited, Padstow, Cornwall
A catalogue record for this publication is available from the British Library.
Library of Congress Cataloging-in-Publication Data
names: Cartwright, Nancy, author.
title: A philosopher looks at science / Nancy Cartwright.
description: First edition. | United Kingdom ; New York, NY, USA : Cambridge University
Press, 2022. | Series: A philosopher looks at | Includes bibliographical references and index.
identifiers: lccn 2022007041 (print) | lccn 2022007042 (ebook) | isbn 9781009201889
(paperback) | isbn 9781009201896 (epub)
subjects: lcsh: Science–Philosophy. | BISAC: SCIENCE / Philosophy & Social Aspects
classification: lcc q175 .c375 2022 (print) | lcc q175 (ebook) | ddc 501–dc23/
eng20220422
LC record available at https://lccn.loc.gov/2022007041
LC ebook record available at https://lccn.loc.gov/2022007042
isbn 978-1-009-20188-9 Paperback
Cambridge University Press has no responsibility for the persistence or accuracy
of URLs for external or third-party internet websites referred to in this publication
and does not guarantee that any content on such websites is, or will remain,
accurate or appropriate.
To Tabi
contents

List of Figures page x


Acknowledgements xi

Introduction 2
Science = Theory + Experiment 3
It’s All Physics Really 5
The Laws of Physics Fix Happenings
Deterministically 7
What’s Wrong with These Three Images? 8

1 Theory + Experiment Do Not a Science Make 12


Preface 12
The Centrality of Theory and Experiment, Knowledge
and Observation 17
The Melange of Theory Ingredients 19
Concepts 20
Defining Concepts 21
Stabilising Concepts 32
Misusing Concepts 34
Models and Narratives 53
Diagrams, Illustrations and Graphs 58
Experiments and the Testing of Theory 60
Experiments: A Life of Their Own 62
Exploring 63
Creating Phenomena 66
Reconstituting Phenomena 67

vii
contents

You Can’t Build an Experiment without a Gigantic


Meccano Set 71

2 Dethroning the Queen 80


Preface 80
The Mechanical Philosophy and Its Legacy 83
Physicalism and Materialism 86
How Philosophers Talk about Reduction 89
Type-Type Reduction 91
Token-Token Reduction 93
Supervenience 97
Grounding 101
Why Chemistry Isn’t All Physics 103
Even Physics Isn’t All Physics 107
Unity at the Point of Action 120

3 A Nature More Negotiable 122


Preface 122
Whence ‘Dappled’? 126
The Disappearance of Diversity 128
A Contemporary Example Where
Difference Mattered 134
A Review of the Argument for Absolute Rule 136
Arguing from Science to the World 140
Where Is Physics Successful? 144
Tendency Laws and How to Understand Them 156
In Sum 165

Parting Thoughts 168


Some Starting Cases to Consider 169
Diagnosis 177
Take-Home Message 181
Anthony Fauci, AIDS and the Parallel Track 182

viii
contents

Barbara McClintock and the Right to Be Different 186


Conclusion 188

Notes 189
Bibliography 195
Index 202

ix
figures

Frontispiece: The Meccano set of science page 1


I.1 Typical children’s drawing of a scientist 3
I.2 An increasingly typical children’s drawing of
a scientist 4
I.3 Another typical drawing of a scientist 5
I.4 The pyramid of the sciences 6
1.1 Kites and rectangles instead of forces, masses
and accelerations 30
1.2 A cake diagram for Person A (abused as a child) 50
1.3 A cake diagram for Person B (not abused as a child) 50
1.4 A cake diagram for Person C (abused as a child:
conditions not combined) 51
1.5 Kinesin walking cargo down a microtubule trail 69
1.6 Kinesin pushing microtubule along 70
1.7 Millikan’s own model of his measurement device 72
1.8 Millikan’s actual apparatus 73
1.9 Why Millikan’s and Fletcher’s results constitute a
measurement of the charge on a droplet 75
4.1 Vajont dam 170

x
acknowledgements

Little of what I say here is new. I build heavily both on the


work of others and on my own earlier work. You can look to
that if you want to see more detailed philosophical develop-
ments of the ideas here: The Dappled World (1999), Nature,
the Artful Modeler (2019) and The Tangle of Science (with
Jeremy Hardie, Eleonora Montuschi, Matthew Soleiman and
Ann Thresher, forthcoming).
A number of people have made huge intellectual
contributions to the work: Andrew Bollhagen, Omar
El-Mawas, Elizabeth and Peter Fisher, Jeremy Hardie,
Matthew Soleiman and Richard Vagnino. Faith Bollhagen,
Lucy Charlton and Adrian Harris have contributed original
drawings. Nicola Craigs, Omar El-Mawas and Matthew
Soleiman have been invaluable in the final production.
This book was supported by funding from the
European Research Council (ERC) and the John
Templeton Foundation. The ERC funding is for Knowledge
for Use, under the European Union’s Horizon 2020 research
and innovation program (grant agreement No 667526 K4U).
The content reflects only the authors’ views, and the ERC is
not responsible for any use that may be made of the infor-
mation it contains. The Templeton funding is for The
Successes and Failures of Science through the Lens of
Intellectual Humility: Perspectives from the History and
Philosophy of Science, ID: 62247.

xi
The Meccano set of science
Drawn by Adrian Harris especially for this book. Thanks
Adrian!
Introduction

W hat does a philosopher see when she looks at


science? What do you see?
Here are three common images of science, widely
shared alike by philosophers, scientists and people in general:
1. Science theory + experiment.
2. It’s all physics really.
3. Science is deterministic: it says that what happens next
follows inexorably from what happened before.
You see indications of the first everywhere, from newspaper
reports on exciting new science results to school texts to the
deliberations of funding bodies. The second is widely held
among philosophers and is also endorsed by quite a few
scientists, though it may not seem so much a part of the
popular image of science. I think that with the exception of
worries about how quantum theory fits in, the third is
central to the popular image: it gives rise to all sorts of
familiar conundrums about the possibility of free will: Are
criminals – or even saints – really responsible for their
actions? Will the final theory of everything allow us to
predict the future with certainty? But, as I’ll explain, it’s
hard to see why you’d hold with (3) if you didn’t believe in
(2), which is generally taken as the logical foundation for (3).
That’s why I’ve included all three, putting them in this order.
Let’s look at each of these three in turn.

2
introduction

Science = Theory + Experiment


Children have been doing a lot of drawing during the Covid-
19 lockdowns. Including my young granddaughter.
‘Nan, how do you draw a scientist?’
‘I don’t know Tabi. I’m rubbish at drawing. Maybe we
can get some help online.’

We looked for simple examples under ‘cartoon sci-


entists’. The most usual image is like Figure I.1.
Happily a few are like Figure I.2 (it’s at last becom-
ing more common for children to draw scientists as
women).1
There are alternatives to pictures with test tubes and
microscopes. Almost all are images of Albert Einstein.

Figure I.1 Typical children’s drawing of a scientist


Drawn by Lucy Charlton especially for this book. Thanks Lucy!

3
a philosopher looks at science

Figure I.2 An increasingly typical children’s drawing of a


scientist
Drawn by Lucy Charlton especially for this book. Thanks Lucy!

(Oddly there are also images that look like Einstein,


renowned for his famous theories of relativity, with a test
tube – whereas Einstein didn’t do experiments at all,
let alone with a test tube.)
This suggests that the common image is: science
theory + experiment. This, for the most part, is what phil-
osophers see too.

4
introduction

Figure I.3 Another typical drawing of a scientist


Drawn by Lucy Charlton especially for this book. Thanks Lucy!

It’s All Physics Really


Tuesday, 27 April 2021, the Guardian newspaper published a
three-page journal article, ‘The clockwork universe’. The
topic was science and free will. The Guardian notes that a
‘growing chorus of scientists and philosophers argue that
free will is an illusion’. The article looks to ‘one of the most
strident of the free will sceptics, the evolutionary biologist
Jerry Coyne for a stark statement of the reason why: “free
will is ruled out, simply and decisively, by the laws
of physics”’.
This bleak view of nature and our place in it is an
exact parallel of an image of science that is deeply imbedded
in philosophical thought and that underwrites the bleak
image of nature: the pyramid of the sciences, pictured in
Figure I.4. Notice that all the sciences use the very same
bricks as physics. And each falls under physics.

5
a philosopher looks at science

Figure I.4 The pyramid of the sciences


Drawn by Rachel Hacking and originally published in
Cartwright 1999. Thanks Rachel!

Physics, they say, is the queen of the sciences. If you


understand what physics does and how, then not only do
you understand such grand stories as the motion of the
planets and space-time curvature, you also understand
everything that the other sciences have to teach, from chem-
istry and biology to psychology and medicine – at least you
would if only you were clever enough to work it out. This is
an image that has been favoured both in philosophy and in
the sciences themselves. Consider for instance the great
experimentalist at the turn of the nineteenth into the twen-
tieth century, often called ‘the father of nuclear physics’,
Ernest Rutherford. Rutherford is famously reputed to have
remarked that ‘[a]ll science is either physics or stamp col-
lecting’.2 Historian and philosopher of science Hasok Chang,
whom you will see challenging the reduction of physics to
chemistry in Chapter 2, also quotes Rutherford, adding: ‘It

6
introduction

may be considered fitting punishment that he was given the


Nobel Prize in chemistry in 1908’.3
This story that everything is built from the building
blocks of physics is often sold under the label ‘the unity of
science’, as we see from the Cambridge Elements mono-
graph titled The Unity of Science:
The notion of the unity of science is regularly connected
to the notion of reduction . . . . According to this line of
thought, unity of science just means that fundamental
physics is what everything else is ultimately based on;
the . . . [other] sciences are somehow derivative.4

As you will see, philosophers have a lot to say about the whole
idea of ‘reduction’ involved in getting from one science to
another, but the pyramid still represents the dominant view.
And after all, who could be opposed to unity? It is supposed to
be a source of special strength: united we stand, divided we fall.

The Laws of Physics Fix


Happenings Deterministically
What happens next follows inexorably from what’s happened
before – or so it is supposed. Physics rules everything and its
laws are deterministic: for a given input, one and only one
output is allowed under its laws. The inputs describe what’s
happened in the past. So exactly one future is allowed. Given
that it’s all physics really, this includes everything: all the
happenings that occur in nature. Chemical properties, the
way proteins fold, the look and feel of things, even your psy-
chological states: their future is fixed since they are ultimately
governed by the laws of physics. So, we live in a totally law-

7
a philosopher looks at science

governed world where things happen mechanically and even


predictably if only we get to know what the laws are. That’s how
the world works. This is how those philosophers and scientists
discussed in The Guardian arrive at the conclusion that, though
we may feel strongly to the contrary, free will is an illusion.
You probably already know that this story about
determinism and physics isn’t entirely right. After all, there
is radioactive decay. Whether a radioactive atom decays or
not in the next hour is open – it may or may not. That is not
fixed by the past. Still, not much is open. The probability that
it will decay is supposed to be entirely fixed by past states. Nor
can we – nor anything else – influence which it will do. What
will happen will happen, willy-nilly. That can provide the
comfort of certainty, but it’s at the cost of impotence.

What’s Wrong with These Three Images?


The idea of a law-governed world is now so entrenched that it
is hard to step back and wonder why it is so at odds with the
world as we see and experience it. Of course things are not
always as they seem. But it is a big stretch from the world as
we know it – a world where all the other sciences than physics
make great discoveries and build impressive technologies,
where we can and do make things happen and where some-
times things go as predicted but often not – to a world totally
ordered from the Big Bang onwards under the undefiable rule
of law and hence out of our control altogether.
Unease about these images increases when you look
at what scientists actually do. It is much messier and more
heterogeneous than the structured process of theory,

8
introduction

experiment and confirmation (or not) that is pictured in


these common images. That process is about discovering the
pre-existing but as yet not identified laws that govern the
universe. To see science that way requires a lot of inventive
imagination that goes far beyond what we actually see: what
is manifest to us and to scientists as they go about trying to
understand the world and how it works. If we look at what
scientists do to produce the wonderful products of science
that we so admire, like lasers, global positioning systems
(GPS) and vaccines, they do not appear to be discovering
laws and deriving results from them. It looks much more
like crafting the pieces of a Meccano set and learning how to
deploy them together, with a lot of trial and error.
The purpose of this book is to get a clearer perspec-
tive on how science produces things and why what it pro-
duces, from vaccines to spaceships, is so often so reliable.
And that is very little to do with what these images suggest.
The book is titled A Philosopher Looks at Science.
Note that it is not Philosophers Look at Science nor
Philosophy Looks at Science. That’s a good thing because
philosophers are a mixed bunch. No two of us think the
same thing. What I see when I look at science is not what
‘philosophers’ see, nor maybe not what the bulk of philoso-
phers see. But it is what many of us see who look at the
details of science as it is practised.
There’s nothing controversial about the features and
practices I shall highlight – no one would deny their role in
science. These three popular images abstract away from these
details. Of course, any image of science that is not an exact
duplicate must do that. The point of abstractions like these is

9
a philosopher looks at science

not to provide an accurate summary of the details but to


substitute a striking image that ‘gets at the heart of what’s going
on’. The trouble is that these three do the opposite in my view.
They conceal what it is that makes science so good at its job.
This book provides an alternative to these three
common images. In it I look at science, and I also look at
the world as seen through the lens of science. I focus on
what science does for us and how it does what it does for us.
The standard images are not well supported by a close
empirical look at how science produces its amazing suc-
cesses. To arrive at the standard story, it takes a good dollop
of what philosophers call ‘metaphysics’ and what I earlier
called ‘inventive imagination’ – sweeping claims far
removed from concrete details we can get our hands on.
When I look at science I see a hotch-potch of finely
crafted pieces brilliantly assembled in diverse ways to pro-
duce the myriad wonderful outputs science gives us, from
understanding to technology, and that reflects behind it a
dappled world, rich in diversity and where much is still
possible. The image of nature that I read back from my look
at science is one with space for the reality of contingency
and for our power to effect change.
This book paints, one by one, alternative pictures to
the three standard images of science that I have described
and defends why I see science that way:

1.0 Theory + experiment do not a science make


It is not just theory and experiment. All the products of
science play a crucial role in its successes: models; meas-
urement definitions, procedures and instruments;

10
introduction

concept development and validation; data collection,


analysis and curation; non-experimental studies; statis-
tical techniques; methods of approximation; case studies;
narratives; etc., etc., etc. And especially important is the
complex ways in which these get interwoven.
0
2. Dethroning the queen
We all know that it’s not all just physics in practice, far –
very far – from it. The idea that nevertheless all derives
from the building blocks of physics in principle belies
both the diversity of science and the diversity of nature,
and it can send us down the wrong research routes. As to
unity, there’s many more ways to get to it than just by
building with the same blocks.
3.0 A nature more negotiable
There’s nothing about even the most certain of the prin-
ciples we use in science to produce its marvellous successes
that implies that they represent ‘laws of nature’ that rule
with an iron hand. Often we engineer just the right circum-
stances – I shall call these ‘small worlds’ – and have the good
luck that nothing dramatically intervenes to upset them so
that what happens within them is fixed and predictable.
Sometimes these kinds of circumstances occur naturally.
Then what happens is fixed and predictable. But that‘s not
what life in general is like. And we don’t need to assume that
God has engineered the whole world to be a strange gigantic
clock in order to make sense of the pockets of exact and
approximate order we do see nor to understand why it is
sometimes possible to find or construct them.
I finish with some Parting Thoughts on the importance of
due intellectual humility in science institutions.

11
1 Theory + Experiment Do Not
a Science Make
Preface
Science has undoubtedly produced remarkable achieve-
ments from deep theories to technological devices to new
ways to measure things. These achievements, I claim, are
secured by a dense interwoven net of scientific constructions
that constrain and support each other – the concurrent,
mutually feeding back-and-forth development of ideas, con-
cepts, theories, experiments, measures, middle-level prin-
ciples, models, methods of inference, research traditions,
data and narratives that make up a scientific endeavour,
with rich interconnections with other bodies of work on
very different topics that also constrain and support it.
I think it’s significant that the images of scientists
that Tabi and I found on Google for our drawings are of
Einstein. Of course, if you are drawing a scientist you need
to make the person recognisable as a scientist, and that’s
hard when it’s theory you want to point to. Einstein has
become the canonical emblem for a theoretical scientist. But
that’s not all there is to it. Einstein represents a certain take
on theory – created by a genius, original, explaining nature’s
deepest secrets. I Googled ‘science theorists’ and the very
first entry was ‘revolutionary theories’ – grand sweeping
theories with broad explanatory stretch, like relativity, quan-
tum theory, evolution, plate tectonics and game theory lately

12
theory + experiment do not a science make

so popular in the economic and social sciences.1 Not the


tens of thousands of more local, domain-specific or lower-
level theories across the sciences that make up the bulk of
scientific theorising, like the theory of high-temperature
superconductivity, the theory of democratic peace, the
theory of thermal neutron scattering, ecological niche
theory, theories about institutional racism, protein folding,
the natural rate of unemployment, cognitive dissonance
and so on and so on.
Nor is this focus on theory and experiment and
concomitantly the explanations that theory provides – and
on getting those right – peculiar to philosophy and popular
imagination. The US National Academy of Sciences’ 2008
report, Science, Evolution, and Creationism, builds this right
into their ‘Definition of Science’: ‘The use of evidence to
construct testable explanations and predictions of natural
phenomena, as well as the knowledge generated through this
process.’ They go on to describe ‘how science works’:
Scientific knowledge and understanding accumulate
from the interplay of observation and explanation [which
sounds to me very much like ‘theory and experiment’].
Scientists gather information by observing the natural
world and conducting experiments. They then propose
how the systems being studied behave in general, basing
their explanations on the data provided through their
experiments and other observations. They test their
explanations by conducting additional observations and
experiments under different conditions. Other scientists
confirm the observations independently and carry out
additional studies that may lead to more sophisticated
explanations and predictions about future observations

13
a philosopher looks at science

and experiments. In these ways, scientists continually


arrive at more accurate and more comprehensive
explanations of particular aspects of nature.2

Similarly, the UK Science Council claims: ‘Science is the


pursuit and application of knowledge and understanding
of the natural and social world following a systematic meth-
odology based on evidence.’3 Again, this sure sounds like
‘theory and experiment’. So – it seems – science is all about
theory, knowledge and explanation and the experiments and
other observations that confirm these.
Sometimes it seems that even experiments do not
get much of a look in. Consider this lament by a particle
physicist turned historian and philosopher of science, Allan
Franklin:

One of the great anticlimaxes in all of literature occurs at


the end of Shakespeare’s Hamlet. On a stage strewn with
noble and heroic corpses – Hamlet, Laertes, Claudius,
and Gertrude – the ambassadors from England arrive
and announce that ‘Rosencrantz and Guildenstern are
dead’. No one cares. A similar reaction might be
produced among a group of physicists, or even among
historians and philosophers of science, were someone to
announce that ‘Lummer and Pringsheim are dead’. And
yet they performed some of the most important
experiments in the history of modern physics. It was
their work on the spectrum of black-body radiation,
along with that of Rubens and Kurlbaum, that showed
deviations from Wien’s Law and formed an important
part of the background to Planck’s introduction
of quantization.

14
theory + experiment do not a science make

This is symptomatic of the general neglect of experiment


and the dominance of theory in the literature on the
history and philosophy of science. In Thomas Kuhn’s
history of quantization, Black-Body Theory and the
Quantum Discontinuity, 1894–1912, Lummer, Pringsheim,
Rubens, and Kurlbaum are, at best, peripheral characters.
The title indicates what Kuhn thinks is important. We
never see what the experimental results were or find a
discussion of how they were obtained.

But, it might be said, that it is only an isolated case.


Surely everyone is aware of the famous experiments of
Galileo and the Leaning Tower of Pisa, of Thomas
Young’s double-slit interference experiment, and of the
Michelson-Morley experiment. What seems to be
generally known, particularly by scientists, about these
experiments shows the mythic treatment of experiment.
Real experiments and their roles are not often dealt with.4

I too, like the Academy of Sciences and the Science Council,


see much knowledge and explanation when I look at science,
and I agree that much of what is mentioned by them can be
conducive to achieving these. But I lament the narrowness of
focus, and my lament extends far beyond Franklin’s. When
I look at science I see far more than theory and experiment,
knowledge and explanation. As I’ve already indicated, I see a
hotch-potch: models, concepts, validation procedures, meas-
ures, classification schemes, statistical techniques, study
designs, data collection, curation and coding methods,
mathematics high and low, methods of inference, narratives
and more. These are resources ready to hand, for science to
use to build successful technological devices, make precise

15
a philosopher looks at science

predictions and create compelling accounts of the world


around us. Experiment and theory definitely must be there.
But finding them – especially those ‘high’ revolutionary
theories – among all the parts that play an essential role is
a bit like playing Where’s Wally.
For science-related institutions to endorse the
common image that science theory + experiment reflects
a lack of intellectual humility in those institutions about the
importance of theory and experiment to the conduct of
science and to what the sciences deliver. But I do not urge
attention to these just because every labourer is worthy of
their hire but also because these different products of science
are mutually supporting. Each successful endeavour in sci-
ence depends on these being up to the job they are needed
for if the endeavour is to be successful. That’s why intellec-
tual humility about theory and experiment is important in
the institutions in and around science. The different kinds of
enterprise in science are intricately interwoven and interde-
pendent. We need due attention, policing and support for all
of them if any are to perform as best they can. And failures
to attend to them can have seriously harmful consequences.
I see science and how it operates as like a giant
Meccano set, with scientists akin to a vast network of hard-
working, practised Meccano builders labouring together in
different teams on different creations. The more usual image
pictures science as uncovering nature’s deepest secrets, high-
lighting big breakthroughs, grand theories and brilliant
experiments, done by men of genius, insight and finesse.
As historian Jaume Navarro notes, ‘[t]he folk history of
science stresses crucial experiments, moments of revelation,

16
theory + experiment do not a science make

great achievements by geniuses’.5 I see science operating


differently. I see a science that sails, with effort and dedica-
tion, between the Charybdis of a hubris that assumes our
scientific successes are due to heroically wresting nature’s
secrets from her and the Scylla of diffidence that urges that
we don’t really know anything and should proceed only with
extreme caution.
Science produces remarkable and reliable outputs.
But not primarily by ingenious experiments and brilliant
theory. Rather by learning, painstakingly, on each occasion,
how to discover or create and then deploy together a pan-
oply of different kinds of highly specific scientific products
to get the job done. Every product of science – whether a
piece of technology, a theory in physics, a model of the
economy or a method for field research – depends on huge
networks of other products to make sense of it and support
it. Each takes imagination, finesse and attention to detail,
and each must be done with care, to the very highest scien-
tific standards, so that it can do the immediate job we expect
of it and because so much else in science depends on it.
There is no hierarchy of significance here. All of these
matter; each labour is indeed equally worthy of its hire.
I begin by looking at theory and experiment and all
they bring with them.

The Centrality of Theory and Experiment,


Knowledge and Observation
I begin by considering a simple principle in physics that we
probably all know: electrons are negatively charged. This is a

17
a philosopher looks at science

central part of physics theory. Later I will talk a bit about


biology, chemistry, neuroscience and the social sciences, but
I want to start with physics because it is the usual paradigm.
I want to show how misleading it is to highlight theory and
experiment even here in this most likely of places.
Here is the principle that tells us just how big the
negative charge on electrons is:
EC principle: The charge of the electron is 1.602 
10 19 Coulombs.

Already this one simple principle raises further questions.


What’s an electron? What’s charge? What are Coulombs?
Let’s suppose for the moment that we know some-
thing about what charge is and aren’t too fussed about the
units – Coulombs. Probably you know that an electron is the
smallest unit of negative charge. More questions. This sup-
poses that negative charge comes in units. It’s not continu-
ous, like the electromagnetic field or as we normally envisage
water in a beaker. How do we know that it comes in
discrete units?
All that, and much more, is answered for us in the
theory of the electron. I do not propose to say anything
much about what this theory says nor how it answers those
questions; just a short paragraph to locate it, after which
I make some general observations about it that tend to be
true of theory generally, whether in physics, biology
or economics.
The discovery of the electron is usually attributed to
J. J. Thompson in 1897 in his cathode ray experiments. But
theory about the electron both predates and postdates these

18
theory + experiment do not a science make

experiments. As to predating, from the introduction to an


impressive collection on the history of the electron:
The electromagnetic effects produced by moving charges
had been explored by Thomson and by Oliver Heaviside
in the 1880s and had been developed into a fully fledged
‘electron theory’ of matter by George Fitzgerald and,
especially, Joseph Larmor in the 1890s.6

Plus ever so much more you can read about in the collection.
As to postdating, the theory got a dramatic rewrite
with the advent of quantum mechanics and the theory of
relativity, at first by Louis de Broglie and Irwin Schrödinger
on the quantum side, which left us the image of the electron
as sometimes a small compact particle and sometimes a
smeared-out wave, and by P. A. M. Dirac, who produced a
relativistic wave equation for it. And even the latest devel-
opments in quantum electrodynamics and quantum field
theory leave many questions open. So, it is a live theory that
is still expanding and developing.
But it is not just developing in conversation with
observation and experiment as suggested in the National
Academy of Science’s discussion of how science works.
There’s far more going on.

The Melange of Theory Ingredients


You might think of theory as just a set of claims – laws or
principles or equations. But there is far more to it than that.
In particular, in this section I outline the role of a number of
other ingredients that constitute theory, making it

19
a philosopher looks at science

meaningful and useful. In particular, I look at concepts;


models and narratives; and diagrams, illustrations and
graphs. I am going to look at concepts in considerable detail
because they are after all the meat and potatoes of theory,
then turn more briefly to these other ingredients of theory.

Concepts
You can’t have theory without concepts. It is also widely
accepted that you shouldn’t admit concepts into science that
do not have exact and unambiguous meanings or that lack
clear indications of ways of connecting them with the world.
It is important for science to ensure that its concepts have
genuine empirical content – that they have a grip in the
empirical world – even if what they get a grip on is not much
as we conceive it to be. There are many other bodies of
thought, like various complex theologies, that resemble the
sciences in being highly articulated and regulated, with a
network of accepted claims and methods that tightly con-
strain what further moves can be made. What separates the
concepts of science from these is the reassurance demanded
in science that its concepts get a grip on the empirical world.
We typically do this through measurement and
experiment. When we do so the measurements and experi-
ments can play a dual role. We frequently use experiment to
test whether a theory is true or not and simultaneously to
help secure empirical content for concepts in it. In the final
section, ‘You Can’t Build an Experiment without a Gigantic
Meccano Set’, I describe in some detail Robert Millikan’s
famous oil drop experiment, which served both to measure

20
theory + experiment do not a science make

the charge on the electron, thereby providing empirical


content to the concept of ‘the electron’, and to test the
theory that negative charge comes in discrete units. The
reason I bring this up here is to point out that measurement
and experiment are not just necessary to test theory but
without them we do not have empirical concepts at all, and
without empirical concepts we do not have empirical
theory. So, just as we need concepts to have theory, so too
we need experiments or other methods of measurement to
have empirical theory – and empirical theory is what we
want in science!

Defining Concepts
The concepts we find in our theoretical sciences do not of
course spring into life fully grown. They have a prehistory:
an often long period during which they are developed,
contested and eventually, as the field of science studies puts
it, ‘stabilised’. I’ll talk about this process of stabilising con-
cepts later. For the moment let’s just think about mature
concepts in what we take to be reasonably well-established
theories. How are we to understand these theoretical con-
cepts, how do we give meaning to them? I’m going to spend
some time discussing this both because these are interesting
questions in their own right and also because the long road
we travel in thinking about these issues leads us to the
conclusion I want to underline in this chapter – how import-
ant all the other pieces beyond theory and experiment are in
the Meccano set of science, so much so that the very notion
of theory does not make sense without them.

21
a philosopher looks at science

Some theoretical concepts have esoteric names and


are unfamiliar to those who are not specialists in the field. In
particle physics we learn about leptons, in low-temperature
superconductivity about Cooper pairs. So, what is a lepton,
what is a Cooper pair? Cell biologists study the cellular cyto-
skeleton, including actin filaments, microtubules and inter-
mediate filaments. They examine the endoplasmic reticulum
and the Golgi apparatus. Cancer biologists study proto-onco-
genes and tumour suppressor genes. What are actin, microtu-
bules and intermediate filaments? What are the endoplasmic
reticulum and the Golgi apparatus? What are proto-oncogenes
and tumour suppressor genes? Social psychologists study pos-
itioning theory. What is a position? Other theoretical concepts
have familiar labels, especially in the social sciences, like dem-
ocracy, socio-economic status and aggressive behaviour. But as
these come – our everyday versions of them – these concepts
are too vague and open-ended to play a role in proper science
theory. The claims of science are meant to be clear and unam-
biguous. So, whether the concepts they employ have familiar or
unfamiliar monikers, the concepts themselves need to be made
clear and unambiguous.
The easy way to give meaning to a concept is simply to
define it. For instance, ‘leptons are considered to be fundamen-
tal particles. They have a spin 1/2 and do not partake in strong
interactions [interacting only via electromagnetic and weak
forces]. As fundamental particles, some leptons are negatively
charged.’7 That’s fine – sort of. The problem is that this defin-
ition is clear only if what it is to be a ‘fundamental particle’ is
already clearly defined, along with ‘spin 1/2’, ‘strong inter-
action’, ‘weak force’, ‘electromagnetic force’ and ‘negatively

22
theory + experiment do not a science make

charged’. Or consider ‘proto-oncogenes’. These are defined by


the US National Cancer Institute as: ‘Gene[s] involved in
normal cell growth. Mutations (changes) in a proto-oncogene
may cause it to become an oncogene, which can cause the
growth of cancer cells.’8 The clarity of this definition also
depends on the clarity of other terms such as ‘normal cell
growth’, ‘mutations’, ‘oncogenes’ and ‘cancer cells’.
What we see here is typical: one theoretical concept
defined in terms of other theoretical concepts that are them-
selves in need of clear and precise definitions. When it
comes to social science, this problem is often compounded
by the fact that there may be multiple different definitions
on offer, definitions that classify items in the world differ-
ently. Consider ‘democracy’. Definitions range from:

[G]overnment with the consent of the governed. This


formula is indeterminate with respect to institutional
forms, or the procedures by which consent is to be
expressed – questions on which consent theorists have
historically differed9

to
[A] competitive political system in which competing
leaders and organizations define the alternatives of public
policy in such a way that the public can participate in the
decision-making process10

to definitions that provide more formal criteria. For instance,


for a state to count as a democracy, its citizens must have:
1. Effective participation
2. Voting equality at the decisive stage

23
a philosopher looks at science

3. Enlightened understanding
4. Control of the agenda
5. Inclusiveness.11

Note that, just like the examples from the natural sciences,
these too end up defining one theoretical concept in terms of
other theoretical concepts. What, after all, is a ‘competitive
political system’, or ‘control of the agenda’, or ‘public par-
ticipation in political decisions’ – what even is a ‘political
decision’?
Can we ever break out of this theory circle, to define
theoretical concepts in less esoteric terms? That was the
hope in the heyday of operationalism, which was cham-
pioned by the American physicist P. W. Bridgman.
Bridgman’s method for breaking out of the theoretical circle
was to define each theoretical concept by an operation by
which it is measured: ‘We mean by any concept nothing
more than a set of operations; the concept is synonymous
with the corresponding set of operations.’12 Readers may be
familiar with this kind of doctrine from behaviourism –
sometimes called ‘rat psychology’ – which was dominant
in psychology from the 1920s through the 1950s. B. F.
Skinner is its most well-known advocate. Skinner main-
tained that psychology is the study of behaviour, not of
some invisible ‘inner mind’, and he urged that all psycho-
logical concepts, like anger, revulsion or reflection, must be
defined in behavioural terms. Note defined. Behaviour is not
a reflection of psychological states or a clue to what someone
else is experiencing. Behaving in particular ways is what it is
to be angry, revulsed or reflective.

24
theory + experiment do not a science make

Behaviourism was widely abandoned in psychology


because it just didn’t seem to work. We couldn’t articulate
distinct kinds of behaviour that fitted with each different
nuanced psychological concept that we regularly employ
and would surely not want to abandon on the grounds that
it must be a chimera since we can’t find a behavioural
definition of it. Behaviourism also completely misses out
on the ‘what-it-feels-like’ to be in these different states. So,
gradually the claim that all psychological phenomena com-
prise behaviour lost dominance to the more everyday view
that they are causally related – though, as I discuss in
Chapter 2, in the section ‘Physicalism and Materialism’,
objections to the genuine reality of the inner mind and its
states remain alive in the programmes to reduce mental
states to states of the brain.
Operationalism more widely suffers from two gen-
eral problems. First, every different measurement procedure
introduces a new concept even if, on the more usual way of
thinking about it, these procedures measure the same con-
cept but in different ways. Then it seems nature must be
littered with huge numbers of principles to bind together
the different concepts, all of which take the same value in the
same circumstances and that we normally think of as all the
same concept, just measured differently; principles like this:
‘Concept 1 and Concept 2 and Concept 3 and . . . Concept
N always have the same value.’ Second, we generally make
big efforts to defend the idea that our procedures are good
for measuring what they are supposed to. As I noted, when
I turn to discussing experiments I use the famous early
twentieth-century oil drop experiments for which Robert

25
a philosopher looks at science

Andrews Millikan won the Nobel Prize for measuring the


charge of the electron. As you will see in the final section in
this chapter, ‘You Can’t Build an Experiment without a
Gigantic Meccano Set’, Millikan did not just assert that what
he calculated from the reading on the voltmeter in his
experiment was the charge of the electrons in his experi-
ment. He argued for this claim, both with a theoretical
model – you’ll see the core of this in Figure 1.8 – and also
in a detailed description of the actual materials which his
experiment employed to show that they could play the part
required of them in securing an accurate and precise meas-
ure of the charge.
Consider another example, this time from genetics.
In his 1929 article on Heredity in the Encyclopaedia
Britannica, British biologist J. B. S. Haldane provided what
can be viewed as an operational definition of phenotype and
genotype. He wrote that ‘[a] class of organisms whose
members cannot be distinguished from one another by
observation is called a phenotype; a class which can be
distinguished from another by breeding tests is called a
genotype’.13 Such definitions were given because genes or
the genetic material had not yet been discovered. At the
time, genes were accepted as some unobservable ‘units of
heredity’ without any direct experimental handle on them.
Hence, a quasi-operational definition was beneficial. Yet,
today, with the advent of molecular genetics, which char-
acterises genotype as the ‘variant forms of a gene that are
carried by an organism’ and phenotype as the ‘observable
physical properties of an organism’ – both of which accept
a molecular characterisation of the genetic material in the

26
theory + experiment do not a science make

form of deoxyribonucleic acid (DNA) and the theoretical


models used to explain gene expression – we would con-
sider such operational ‘definitions’ superficial, serving at
best as indicators of genotype and phenotype but falling
short of proper definition. This shows that establishing the
proper definition of genotype and phenotype required a
lot of material experimental work within some broad the-
oretical framework that operational definitions fail to
appreciate.
Operationalism makes nonsense of these elaborate
efforts to defend that our measurement procedures are up to
the job of measuring the concept they are supposed to, since
after all operationalism holds that the concept just is what is
measured by those procedures.
Just looking to the procedures by which a concept is
measured may, though, be too narrow a focus in attempting
to break out of the spiral of definition of one theoretical
concept by another by another by another. In the 1930s,
1940s and 1950s, philosophers – especially logical empiri-
cists – tried valiantly to define theoretical concepts using
purely observational terms. They were called logical empiri-
cists because they argued that scientific claims should be
made entirely explicit, for instance theories should be for-
mulated formally as systems of axioms from which further
claims could be deduced as theorems. The empiricist label
was because they wanted science to be able to confirm each
claim by empirical observation. What exactly is meant by
observation was up for grabs – did it for instance mean
observable with the naked eye or was it to allow the use of
sophisticated instruments? Whichever way that is decided,

27
a philosopher looks at science

what matters for breaking out of the circle of theoretical


definition is that for each claim to be confirmable entirely by
observation, each theoretical concept in that claim needs to
have an observational correlate – some observational states
that obtain if and only if the theoretical concept obtains. So,
hurray, these observational correlates can serve as defin-
itions for these concepts and the circle is broken for each
concept.
The difficulty is that this programme failed miser-
ably. There just seems to be no way to carry it off. As Carl
Hempel, who is widely acknowledged as the leading phil-
osopher of science of the time, concluded in a famous paper
written originally in 1958: ‘[I]t is clear that theoretical for-
mulations cannot be replaced by expressions in terms of
observables only.’14 This continues to be true even if we
become for more liberal, not looking just for correlates
among features that can be observed but allowing that
theoretical concepts be defined by any terms that we already
grasp the meaning of, as Hempel suggested: ‘[W]e might
qualify a theoretical expression as intelligible or significant if
it has been adequately explained in terms which we consider
as antecedently understood.’15 The result of these repeated
failures is that philosophers settled for implicit definitions of
theoretical concepts rather than explicit ones that defined
them using concepts outside the theoretical circle. All that
means is that we give in and take the meaning of a theoret-
ical concept to be given by the axioms of the theory. As the
Stanford Encyclopedia of Philosophy notes: ‘This idea has
become almost constitutive of the very notion of a theoret-
ical term in the philosophy of science.’16

28
theory + experiment do not a science make

The problem with this is that implicit definition


leaves it open what our theories are talking about. That’s
because no matter how detailed a theory is – how many
axioms we add to its formalisation – there can never be
enough detail to pick out uniquely what the theory is about.
There will always be unintended or ‘non-standard’ interpret-
ations for it. This follows from theorems in model theory in
logic, but it is easy to get the gist of why by looking at a
caricature example.
Consider the simple theory that has one axiom, an
axiom familiar from school physics: F ma. We think of this
as telling us that the force on a material object is equal to its
mass times its acceleration. But all we really know from the
formula is that there are three quantities and the first is
equal to the product of the other two. That is just as true
of the area of a rectangle with respect to the length of its two
sides. So F could mean area of a rectangle, m the length of
the rectangle and a its width. Now let us add more detail. In
a world where only gravitation acts we can add the law of
gravity to our axiom set: FG GMm/r2. This tells us what
force a system with mass m will experience in the presence
of another of mass M a distance r away, where G is the
constant of gravity. From this we infer that in this gravity-
only world, ma F GMm/r2. But this works just as well for
rectangles if we suppose that the rectangles come overlaid
with kites, as in Figure 1.1.
Just suppose G 2, M the area of the kite and
r √d.
We can go on thickening the theory of course. As
we do so, more and more alternative interpretations will get

29
a philosopher looks at science

Figure 1.1 Kites and rectangles instead of forces, masses and


accelerations
Drawn by Lucy Charlton especially for this book. Thanks Lucy!

ruled out. Maybe it will no longer be possible to see the


equations of classical mechanics as all about rectangles. It
naturally helps in narrowing down the interpretation if some
of the concepts that figure in the theory are among that nice
collection of concepts that are antecedently understood so
that what they refer to is nailed down. For instance, we may
suppose that a in the formula F ma refers to acceleration,
which is a concept in our ordinary language vocabulary and
we are clear just what it means. (But beware of even that. We
take acceleration in the physics formula to be the rate of
change of velocity with time (dv/dt) but in some medieval
physics the acceleration of a falling body was understood
as the rate of change of velocity with distance traversed
(dv/dx).) Given this assumption, it gets harder – though
not impossible – to interpret F as the area of a rectangle.
Until each theoretical term is identified with something we
antecedently understand, we cannot be sure of a unique
interpretation. Without this, no set of axioms no matter
how long can guarantee a unique interpretation. As the

30
theory + experiment do not a science make

philosopher Hilary Putnam argued, you can’t fix what you


are talking about just by talking more and more.
So, implicit definition of theoretical concepts by
laying out the theoretical principles that are supposed to
be true of the concepts doesn’t work to fix what those
concepts refer to. Nor can we define them explicitly in terms
of other concepts we already understand outside the con-
fines of the theory.
Philosophers worried about this problem a lot. But
I don’t know of any scientists who do. I think there’s a good
reason for this. Theory isn’t just a set of claims that stands
there to inform us of what the world is like – it’s not there just
to describe the world. Theories are tools that we use to do
things in the world. We use theories to build models and we
use the models to make predictions about what will happen in
the world and to design experiments and technologies and
policies and measurement procedures that we then imple-
ment, jostling the world, picking up bits and changing them
around, to learn more about the world and to try to make it
more to our liking. These are what break us out of the spiral of
theory defining theory defining theory. When we position our
specially designed radar sensors at the ends of the court at
Wimbledon, train these on the ball that Serena Williams
serves and read out 122 miles per hour, we know we are
dealing with velocity and if we calculate its time rate of change
(so a dv/dt) and associate that with the formula F ma, then
we can be sure that F ma is not about the areas, lengths and
widths of rectangles.
We must be careful, though, about how much all
these successful interjections of theory into the world can

31
a philosopher looks at science

buy for us. They serve to rule out unintended interpretations


of our theoretical concepts, but they do not ensure that these
theoretical concepts refer to things that are really there in
the world as we conceive them in our theories. Just think
about all those theories that we used in the past to make
successful predictions about what will happen in the world
and to change things, like phlogiston theory that was used
successfully to produce breathable air, inflammable air,
shiny metal and ‘calx’ (metallic ash).17

Stabilising Concepts
Scientific concepts do not, of course, appear full blown and
fully formed out of nowhere. Rather, they are constructed
within and by science and its surrounds, contested and over
time stabilised. There are various ideas of how this happens.
To the early twentieth-century bacteriologist and philosopher
Ludwick Fleck, concepts were developed within a community
of experts and spread outward to the general public. At the
same time, the public reinforces the thinking of experts. As
Fleck urged, scientists are ‘more or less dependent, whether
consciously or subconsciously, upon “public opinion”’.18
By contrast, in the late 1970s the philosopher Bruno
Latour and the sociologist Steve Woolgar argued that a fact
is produced within the scientific community through per-
suasion – scientists must constantly convince one another
that certain statements using those concepts should be
treated as facts. The point at which a concept becomes
stable, Latour and Woolgar write, is when a statement using
the concept ‘rids itself of all determinants of place and time

32
theory + experiment do not a science make

and of all reference to its producers and the production


process’.19 The meaning of a concept becomes fixed when
it does not seem to have been constructed in the first place.
And of course, as time goes on and new things are learned
and new ideas and influences arise, concepts can also desta-
bilise and even eventually disappear.
Aside from the obvious social processes it takes to
get a concept entrenched, the stabilisation of any one con-
cept in science depends on many products of science,
including not just other concepts. Concepts become stabil-
ised in part because of how they are gradually interrelated
with other pieces of knowledge and practice, especially those
from fields of research other than the one that first
developed the concept.
Take, for example, the concept of a ‘neuron type’.
According to neuroscientists, there are not just different
types of cells in the nervous system but different types of
nerve cells in terms of their shape, electrical properties, gene
expression and, perhaps most importantly, function.
Although the concept of a neuron type is not new, since
the early 2000s neuroscientists have developed materials and
methods that support and constrain what the concept of a
neuron type can amount to. They have used genetically
modified animals and/or genetically modified viruses to
express light-sensitive proteins in specific neuron types
within the brains of mice. Upon delivering flashes of light,
they could manipulate the activity of these neuron types,
such as those that produce a particular neurotransmitter.
Consequently, they could also explore which behaviours
these types of neurons control.20 In this case, products that

33
Another random document with
no related content on Scribd:
man like as we are, and that he worshiped a God whom he could not
see, but who was always near us. Let us learn to worship
Nickelseyn’s God.” So the tribe came down from their hills to the
Christian teachers at Peshawur, and there were baptized.
XV
A. D. 1853
THE GREAT FILIBUSTER

WILLIAM WALKER, son of a Scotch banker, was born in


Tennessee, cantankerous from the time he was whelped. He never
swore or drank, or loved anybody, but was rigidly respectable and
pure, believed in negro slavery, bristled with points of etiquette and
formality, liked squabbling, had a nasty sharp tongue, and a taste for
dueling. The little dry man was by turns a doctor, editor and lawyer,
and when he wanted to do anything very outrageous, always began
by taking counsel’s opinion. He wore a black tail-coat, and a black
wisp of necktie even when in 1853 he landed an army of forty-five
men to conquer Mexico. His followers were California gold miners
dressed in blue shirts, duck trousers, long boots, bowie knives,
revolvers and rifles. After he had taken the city of La Paz by assault,
called an election and proclaimed himself president of Sonora, he
was joined by two or three hundred more of the same breed from
San Francisco. These did not think very much of a leader twenty-
eight years old, standing five feet six, and weighing only nine stone
four, so they merrily conspired to blow him up with gunpowder, and
disperse with what plunder they could grab. Mr. Walker shot two,
flogged a couple, disarmed the rest without showing any sign of
emotion. He could awe the most truculent desperado into abject
obedience with one glance of his cool gray eye, and never allowed
his men to drink, play cards, or swear. “Our government,” he wrote,
“has been formed upon a firm and sure basis.”
The Mexicans and Indians thought otherwise, for while the new
president of Sonora marched northward, they gathered in hosts and
hung like wolves in the rear of the column, cutting off stragglers, who
were slowly tortured to death. Twice they dared an actual attack, but
Walker’s grim strategies, and the awful rifles of despairing men, cut
them to pieces. So the march went on through hundreds of miles of
blazing hot desert, where the filibusters dropped with thirst, and blew
their own brains out rather than be captured. Only thirty-four men
were left when they reached the United States boundary, the
president of Sonora, in a boot and a shoe, his cabinet in rags, his
army and navy bloody, with dried wounds, gaunt, starving, but too
terrible for the Mexican forces to molest. The filibusters surrendered
to the United States garrison as prisoners of war.
Just a year later, with six of these veterans, and forty-eight other
Californians, Walker landed on the coast of Nicaragua. This happy
republic was blessed at the time with two rival presidents, and the
one who got Walker’s help very soon had possession of the country.
As hero of several brilliant engagements, Walker was made
commander-in-chief, and at the next election chosen by the people
themselves as president. He had now a thousand Americans in his
following, and when the native statesmen and generals proved
treacherous, they were promptly shot. Walker’s camp of wild
desperadoes was like a Sunday-school, his government the cleanest
ever known in Central America, and his dignity all prickles, hard to
approach. He depended for existence on the services of Vanderbilt’s
steamship lines, but seized their warehouse for cheating. He was
surrounded by four hostile republics, Costa Rica, San Salvador,
Honduras and Guatemala, and insulted them all. He suspended
diplomatic relations with the United States, demanded for his one
schooner-of-war salutes from the British navy, and had no sense of
humor whatsoever. Thousands of brave men died for this prim little
lawyer, and tens of thousands fell by pestilence and battle in his
wars, but with all his sweet unselfishness, his purity, and his valor,
poor Walker was a prig. So the malcontents of Nicaragua, and the
republics from Mexico to Peru, joined the steamship company, the
United States and Great Britain to wipe out his hapless government.
The armies of four republics were closing in on Walker’s capital,
the city of Granada. He marched out to storm the allies perched on
an impregnable volcano, and was carrying his last charge to a
victorious issue, when news reached him that Zavala with eight
hundred men had jumped on Granada. He forsook his victory and
rushed for the capital city.
There were only one hundred and fifty invalids and sick in the
Granada garrison to man the church, armory and hospital against
Zavala, but the women loaded rifles for the wounded and after
twenty-two hours of ghastly carnage, the enemy were thrown out of
the city. They fell back to lie in Walker’s path as he came to the
rescue. Walker saw the trap, carried it with a charge, drove Zavala
back into the city, broke him between two fires, then sent a
detachment to intercept his flight. In this double battle, fighting eight
times his own force, Walker killed half the allied army.
But the pressure of several invasions at once was making it
impossible for Walker to keep his communication open with the sea
while he held his capital. Granada, the most beautiful of all Central
American cities, must be abandoned, and, lest the enemy win the
place, it must be destroyed. So Walker withdrew his sick men to an
island in the big Lake Nicaragua; while Henningsen, an Englishman,
his second in command, burned and abandoned the capital.
But now, while the city burst into flames, and the smoke went up
as from a volcano, the American garrison broke loose, rifled the
liquor stores and lay drunk in the blazing streets, so the allied army
swooped down, cutting off the retreat to the lake. Henningsen,
veteran of the Carlist and Hungarian revolts, a knight errant of lost
causes, took three weeks to fight his way three miles, before Walker
could cover his embarkment on the lake. There had been four
hundred men in the garrison, but only one hundred and fifty
answered the roll-call in their refuge on the Isle of Omotepe. In the
plaza of the capital city they had planted a spear, and on the spear
hung a rawhide with this inscription:—
“Here was Granada!”
In taking that heap of blackened ruins four thousand out of six
thousand of the allies had perished; but even they were more
fortunate than a Costa Rican army of invasion, which killed fifty of
the filibusters, at a cost of ten thousand men slain by war and
pestilence. It always worked out that the killing of one filibuster cost
on the average eight of his adversaries.
Four months followed of confused fighting, in which the
Americans slowly lost ground, until at last they were besieged in the
town of Rivas, melting the church bells for cannon-balls, dying at
their posts of starvation. The neighboring town of San Jorge was
held by two thousand Costa Ricans, and these Walker attempted to
dislodge. His final charge was made with fifteen men into the heart of
the town. No valor could win against such odds, and the orderly
retreat began on Rivas. Two hundred men lay in ambush to take
Walker at a planter’s house by the wayside, and as he rode wearily
at the head of his men they opened fire from cover at a range of
fifteen yards. Walker reined in his horse, fired six revolver-shots into
the windows, then rode on quietly erect while the storm of lead raged
about him, and saddle after saddle was emptied. A week afterward
the allies assaulted Rivas, but left six hundred men dead in the field,
so terrific was the fire from the ramparts.
It was in these days that a British naval officer came under flag
of truce from the coast to treat for Walker’s surrender.
“I presume, sir,” was the filibuster’s greeting, “that you have
come to apologize for the outrage offered to my flag, and to the
commander of the Nicaraguan schooner-of-war Granada.”
“If they had another schooner,” said the Englishman afterward, “I
believe they would have declared war on Great Britain.”
Then the United States navy treated with this peppery little
lawyer, and on the first of May, 1857, he grudgingly consented to
being rescued.
During his four years’ fight for empire, Walker had enlisted three
thousand five hundred Americans—and the proportion of wounds
was one hundred and thirty-seven for every hundred men. A
thousand fell. The allied republics had twenty-one thousand soldiers
and ten thousand Indians—and lost fifteen thousand killed.
Two years later, Walker set out again with a hundred men to
conquer Central America, in defiance of the British and United States
squadrons, sent to catch him, and in the teeth of five armed
republics. He was captured by the British, shot by Spanish
Americans upon a sea beach in Honduras, and so perished, fearless
to the end.
XVI
A. D. 1857
BUFFALO BILL

THE Mormons are a sect of Christians with some queer ideas, for
they drink no liquor, hold all their property in common, stamp out any
member who dares to think or work for himself, and believe that the
more wives a man has the merrier he will be. The women, so far as I
met them are like fat cows, the men a slovenly lot, and not too
honest, but they are hard workers and first-rate pioneers.
Because they made themselves unpopular they were
persecuted, and fled from the United States into the desert beside
the Great Salt Lake. There they got water from the mountain streams
and made their land a garden. They only wanted to be left alone in
peace, but that was a poor excuse for slaughtering emigrants.
Murdering women and children is not in good taste.
The government sent an army to attend to these saints, but the
soldiers wanted food to eat, and the Mormons would not sell, so
provisions had to be sent a thousand miles across the wilderness to
save the starving troops. So we come to the herd of beef cattle
which in May, 1857, was drifting from the Missouri River, and to the
drovers’ camp beside the banks of the Platte.
A party of red Indians on the war-path found that herd and camp;
they scalped the herders on guard, stampeded the cattle and rushed
the camp, so that the white men were driven to cover under the river
bank. Keeping the Indians at bay with their rifles, the party marched
for the settlements wading, sometimes swimming, while they pushed
a raft that carried a wounded man. Always a rear guard kept the
Indians from coming too near. And so the night fell.
“I, being the youngest and smallest,” says one of them, “had
fallen behind the others.... When I happened to look up to the
moonlit sky, and saw the plumed head of an Indian peeping over the
bank.... I instantly aimed my gun at his head, and fired. The report
rang out sharp and loud in the night air, and was immediately
followed by an Indian whoop; and the next moment about six feet of
dead Indian came tumbling into the river. I was not only overcome
with astonishment, but was badly scared, as I could hardly realize
what I had done.”
Back came Frank McCarthy, the leader, with all his men. “Who
fired that shot?”
“I did.”
“Yes, and little Billy has killed an Indian stone-dead—too dead to
skin!”
At the age of nine Billy Cody had taken the war-path.
In those days the army had no luck. When the government sent
a herd of cattle the Indians got the beef, and the great big train of
seventy-five wagons might just as well have been addressed to the
Mormons, who burned the transport, stole the draft oxen and turned
the teamsters, including little Billy, loose in the mountains, where
they came nigh starving. The boy was too thin to cast a shadow
when in the spring he set out homeward across the plains with two
returning trains.
One day these trains were fifteen miles apart when Simpson, the
wagon boss, with George Woods, a teamster, and Billy Cody, set off
riding mules from the rear outfit to catch up the teams in front. They
were midway when a war party of Indians charged at full gallop,
surrounding them, but Simpson shot the three mules and used their
carcasses to make a triangular fort. The three whites, each with a
rifle and a brace of revolvers were more than a match for men with
bows and arrows, and the Indians lost so heavily that they retreated
out of range. That gave the fort time to reload, but the Indians
charged again, and this time Woods got an arrow in the shoulder.
Once more the Indians retired to consult, while Simpson drew the
arrow from Woods’ shoulder, plugging the hole with a quid of
chewing tobacco. A third time the Indians charged, trying to ride
down the stockade, but they lost a man and a horse. Four warriors
had fallen now in this battle with two men and a little boy, but the
Indians are a painstaking, persevering race, so they waited until
nightfall and set the grass on fire. But the whites had been busy with
knives scooping a hole from whence the loose earth made a
breastwork over the dead mules, so that the flames could not reach
them, and they had good cover to shoot from when the Indians
charged through the smoke. After that both sides had a sleep, and at
dawn they were fresh for a grand charge, handsomely repulsed. The
redskins sat down in a ring to starve the white men out, and great
was their disappointment when Simpson’s rear train of wagons
marched to the rescue. The red men did not stay to pick flowers.
It seems like lying to state that at the age of twelve Billy Cody
began to take rank among the world’s great horsemen, and yet he
rode on the pony express, which closed in 1861, his fourteenth year.
The trail from the Missouri over the plains, the deserts and the
mountains into California was about two thousand miles through a
country infested with gangs of professional robbers and hostile
Indian tribes. The gait of the riders averaged twelve miles an hour,
which means a gallop, to allow for the slow work in mountain passes.
There were one hundred ninety stations at which the riders changed
ponies without breaking their run, and each must be fit and able for
one hundred miles a day in time of need. Pony Bob afterward had
contracts by which he rode one hundred miles a day for a year.
Now, none of the famous riders of history, like Charles XII, of
Sweden; Dick, King of Natal, or Dick Turpin, of England, made
records to beat the men of the pony express, and in that service Billy
was counted a hero. He is outclassed by the Cossack Lieutenant
Peschkov, who rode one pony at twenty-eight miles a day the length
of the Russian empire, from Vladivostok to St. Petersburg, and by Kit
Carson who with one horse rode six hundred miles in six days.
There are branches of horsemanship, too, in which he would have
been proud to take lessons from Lord Lonsdale, or Evelyn French,
but Cody is, as far as I have seen, of all white men incomparable for
grace, for beauty of movement, among the horsemen of the modern
world.
But to turn back to the days of the boy rider.
“One day,” he writes, “when I galloped into my home station I
found that the rider who was expected to take the trip out on my
arrival had gotten into a drunken row the night before, and had been
killed.... I pushed on ... entering every relay station on time, and
accomplished the round trip of three hundred twenty-two miles back
to Red Buttes without a single mishap, and on time. This stands on
the record as being the longest pony express journey ever made.”
One of the station agents has a story to tell of this ride, made
without sleep, and with halts of only a few minutes for meals. News
had leaked out of a large sum of money to be shipped by the
express, and Cody, expecting robbers, rolled the treasure in his
saddle blanket, filling the official pouches with rubbish. At the best
place for an ambush two men stepped out on to the trail, halting him
with their muskets. As he explained, the pouches were full of
rubbish, but the road agents knew better. “Mark my words,” he said
as he unstrapped, “you’ll hang for this.”
“We’ll take chances on that, Bill.”
“If you will have them, take them!” With that he hurled the
pouches, and as robber number one turned to pick them up, robber
number two had his gun-arm shattered with the boy’s revolver-shot.
Then with a yell he rode down the stooping man, and spurring hard,
got out of range unhurt. He had saved the treasure, and afterward
both robbers were hanged by vigilantes.
Once far down a valley ahead Cody saw a dark object above a
boulder directly on his trail, and when it disappeared he knew he was
caught in an ambush. Just as he came into range he swerved wide
to the right, and at once a rifle smoked from behind the rock. Two
Indians afoot ran for their ponies while a dozen mounted warriors
broke from the timbered edge of the valley, racing to cut him off. One
of these had a war bonnet of eagle plumes, the badge of a chief, and
his horse, being the swiftest, drew ahead. All the Indians were firing,
but the chief raced Cody to head him off at a narrow pass of the
valley. The boy was slightly ahead, and when the chief saw that the
white rider would have about thirty yards to spare he fitted an arrow,
drawing for the shot. But Cody, swinging round in the saddle, lashed
out his revolver, and the chief, clutching at the air, fell, rolling over
like a ball as he struck the ground. At the chief’s death-cry a shower
of arrows from the rear whizzed round the boy, one slightly wounding
his pony who, spurred by the pain, galloped clear, leaving the
Indians astern in a ten mile race to the next relay.
After what seems to the reader a long life of adventure, Mr. Cody
had just reached the age of twenty-two when a series of wars broke
out with the Indian tribes, and he was attached to the troops as a
scout. A number of Pawnee Indians who thought nothing of this
white man, were also serving. They were better trackers, better
interpreters and thought themselves better hunters. One day a party
of twenty had been running buffalo, and made a bag of thirty-two
head when Cody got leave to attack a herd by himself. Mounted on
his famous pony Buckskin Joe he made a bag of thirty-six head on a
half-mile run, and his name was Buffalo Bill from that time onward.
That summer he led a squadron of cavalry that attacked six
hundred Sioux, and in that fight against overwhelming odds he
brought down a chief at a range of four hundred yards, in those days
a very long shot. His victim proved to be Tall Bull, one of the great
war leaders of the Sioux. The widow of Tall Bull was proud that her
husband had been killed by so famous a warrior as Prairie Chief, for
that was Cody’s name among the Indians.
There is one very nice story about the Pawnee scouts. A new
general had taken command who must have all sorts of etiquette
proper to soldiers. It was all very well for the white sentries to call at
intervals of the night from post to post: “Post Number One, nine
o’clock, all’s well!” “Post Number Two, etc.”
But when the Pawnee sentries called, “Go to hell, I don’t care!”
well, the practise had to be stopped.
Of Buffalo Bill’s adventures in these wars the plain record would
only take one large volume, but he was scouting in company with
Texas Jack, John Nelson, Belden, the White Chief, and so many
other famous frontier heroes, each needing at least one book
volume, that I must give the story up as a bad job. At the end of the
Sioux campaign Buffalo Bill was chief of scouts with the rank of
colonel.

Colonel Cody
(“Buffalo Bill”)
In 1876, General Custer, with a force of nearly four hundred
cavalry, perished in an attack on the Sioux, and the only survivor
was his pet boy scout, Billy Jackson, who got away at night
disguised as an Indian. Long afterward Billy, who was one of God’s
own gentlemen, told me that story while we sat on a grassy hillside
watching a great festival of the Blackfeet nation.
After the battle in which Custer—the Sun Child—fell, the big
Sioux army scattered, but a section of it was rounded up by a force
under the guidance of Buffalo Bill.
“One of the Indians,” he says, “who was handsomely decorated
with all the ornaments usually worn by a war chief ... sang out to me
‘I know you, Prairie Chief; if you want to fight come ahead and fight
me!’
“The chief was riding his horse back and forth in front of his men,
as if to banter me, and I accepted the challenge. I galloped toward
him for fifty yards and he advanced toward me about the same
distance, both of us riding at full speed, and then when we were only
about thirty yards apart I raised my rifle and fired. His horse fell to
the ground, having been killed by my bullet. Almost at the same
instant my horse went down, having stepped in a gopher-hole. The
fall did not hurt me much, and I instantly sprang to my feet. The
Indian had also recovered himself, and we were now both on foot,
and not more than twenty paces apart. We fired at each other
simultaneously. My usual luck did not desert me on this occasion, for
his bullet missed me, while mine struck him in the breast. He reeled
and fell, but before he had fairly touched the ground I was upon him,
knife in hand, and had driven the keen-edged weapon to its hilt in his
heart. Jerking his war-bonnet off, I scientifically scalped him in about
five seconds....
“The Indians came charging down upon me from a hill in hopes
of cutting me off. General Merritt ... ordered ... Company K to hurry
to my rescue. The order came none too soon.... As the soldiers
came up I swung the Indian chieftain’s topknot and bonnet in the air,
and shouted: ‘The first scalp for Custer!’”
Far up to the northward, Sitting Bull, with the war chief Spotted
Tail and about three thousand warriors fled from the scene of the
Custer massacre. And as they traveled on the lonely plains they
came to a little fort with the gates closed. “Open your gates and hand
out your grub,” said the Indians.
“Come and get the grub,” answered the fort.
So the gates were thrown open and the three thousand warriors
stormed in to loot the fort. They found only two white men standing
outside a door, but all round the square the log buildings were
loopholed and from every hole stuck out the muzzle of a rifle. The
Indians were caught in such a deadly trap that they ran for their lives
back to camp.
Very soon news reached the Blackfeet that their enemies the
Sioux were camped by the new fort at Wood Mountain, so the whole
nation marched to wipe them out, and Sitting Bull appealed for help
to the white men. “Be good,” said the fort, “and nobody shall hurt
you.”
So the hostile armies camped on either side, and the thirty white
men kept the peace between them. One day the Sioux complained
that the Blackfeet had stolen fifty horses. So six of the white men
were sent to the Blackfoot herd to bring the horses back. They did
not know which horses to select so they drove off one hundred fifty
for good measure straight at a gallop through the Blackfoot camp,
closely pursued by that indignant nation. Barely in time they ran the
stock within the fort, and slammed the gates home in the face of the
raging Blackfeet. They were delighted with themselves until the
officer commanding fined them a month’s pay each for insulting the
Blackfoot nation.
The winter came, the spring and then the summer, when those
thirty white men arrived at the Canada-United States boundary
where they handed over three thousand Sioux prisoners to the
American troops. From that time the redcoats of the Royal Northwest
Mounted Police of Canada have been respected on the frontier.
And now came a very wonderful adventure. Sitting Bull, the
leader of the Sioux nation who had defeated General Custer’s
division and surrendered his army to thirty Canadian soldiers, went
to Europe to take part in a circus personally conducted by the chief
of scouts of the United States Army, Buffalo Bill. Poor Sitting Bull
was afterward murdered by United States troops in the piteous
massacre of Wounded Knee. Buffalo Bill for twenty-six years
paraded Europe and America with his gorgeous Wild West show,
slowly earning the wealth which he lavished in the founding of Cody
City, Wyoming.
Toward the end of these tours I used to frequent the show camp
much like a stray dog expecting to be kicked, would spend hours
swapping lies with the cowboys in the old Deadwood Coach, or sit at
meat with the colonel and his six hundred followers. On the last tour
the old man was thrown by a bad horse at Bristol and afterward rode
with two broken bones in splints. Only the cowboys knew, who told
me, as day by day I watched him back his horse from the ring with all
the old incomparable grace.
He went back to build a million dollar irrigation ditch for his little
city on the frontier, and shortly afterward the newspapers reported
that my friends—the Buffalo Creek Gang of robbers—attacked his
bank, and shot the cashier. May civilization never shut out the free
air of the frontier while the old hero lives, in peace and honor, loved
to the end and worshiped by all real frontiersmen.
XVII
A. D. 1860
THE AUSTRALIAN DESERT

WHEN the Eternal Father was making the earth, at one time He
filled the sea with swimming dragons, the air with flying dragons, and
the land with hopping dragons big as elephants; but they were not a
success, and so He swept them all away. After that he filled the
southern continents with a small improved hopping dragon, that laid
no eggs, but carried the baby in a pouch. There were queer half-
invented fish, shadeless trees, and furry running birds like the emu
and the moa. Then He swamped that southern world under the sea,
and moved the workshop to our northern continents. But He left New
Zealand and Australia just as they were, a scrap of the half-finished
world with furry running birds, the hopping kangaroo, the shadeless
trees, and half-invented fish.
So when the English went to Australia it was not an ordinary
voyage, but a journey backward through the ages, through goodness
only knows how many millions of years to the fifth day of creation. It
was like visiting the moon or Mars. To live and travel in such a
strange land a man must be native born, bush raised, and cunning at
that, on pain of death by famine.
The first British settlers, too, were convicts. The laws were so
bad in England that a fellow might be deported merely for giving
cheek to a judge; and the convicts on the whole were very decent
people, brutally treated in the penal settlements. They used to
escape to the bush, and runaway convicts explored Australia mainly
in search of food. One of them, in Tasmania, used, whenever he
escaped, to take a party with him and eat them one by one, until he
ran short of food and had to surrender.
Later on gold was discovered, and free settlers drifted in, filling
the country, but the miners and the farmers were too busy earning a
living to do much exploration. So the exploring fell to English
gentlemen, brave men, but hopeless tenderfeet, who knew nothing
of bushcraft and generally died of hunger or thirst in districts where
the native-born colonial grows rich to-day.
Edgar John Eyre, for instance, a Yorkshireman, landed in
Sydney at the age of sixteen, and at twenty-five was a rich sheep-
farmer, appointed by government protector of the black fellows. In
1840 the colonists of South Australia wanted a trail for drifting sheep
into Western Australia, and young Eyre, from what he had learned
among the savages, said the scheme was all bosh, in which he was
perfectly right. He thought that the best line for exploring was
northward, and set out to prove his words, but got tangled up in the
salt bogs surrounding Torrens, and very nearly lost his whole party in
an attempt to wade across. After that failure he felt that he had
wasted the money subscribed in a wildcat project, so to make good
set out again to find a route for sheep along the waterless south
coast of the continent. He knew the route was impossible, but it is a
poor sort of courage that has to feed on hope, and the men worth
having are those who leave their hopes behind to march light while
they do their duty.
Eyre’s party consisted of himself and his ranch foreman Baxter, a
favorite black boy Wylie, who was his servant, and two other natives
who had been on the northward trip. They had nine horses, a pony,
six sheep, and nine weeks’ rations on the pack animals.
The first really dry stage was one hundred twenty-eight miles
without a drop of water, and it was not the black fellows, but Eyre,
the tenderfoot, who went ahead and found the well that saved them.
The animals died off one by one, so that the stores had to be left
behind, and there was no food but rotten horse-flesh which caused
dysentery, no water save dew collected with a sponge from the
bushes after the cold nights. The two black fellows deserted, but
after three days came back penitent and starving, thankful to be
reinstated.
These black fellows did not believe the trip was possible, they
wanted to go home, they thought the expedition well worth
plundering, and so one morning while Eyre was rounding up the
horses they shot Baxter, plundered the camp and bolted. Only Eyre
and his boy Wylie were left, but if they lived the deserters might be
punished. So the two black fellows, armed with Baxter’s gun, tried to
hunt down Eyre and his boy with a view to murder. They came so
near at night that Eyre once heard them shout to Wylie to desert.
Eyre and the boy stole off, marching so rapidly that the murderers
were left behind and perished.
A week later, still following the coast of the Great Bight, Wylie
discovered a French ship lying at anchor, and the English skipper fed
the explorers for a fortnight until they were well enough to go on.
Twenty-three more days of terrible suffering brought Eyre and his
boy, looking like a brace of scarecrows, to a hilltop overlooking the
town of Albany. They had reached Western Australia, the first
travelers to cross from the eastern to the western colonies.
In after years Eyre was governor of Jamaica.

II
Australia, being the harshest country on earth, breeds the
hardiest pioneers, horsemen, bushmen, trackers, hunters, scouts,
who find the worst African or American travel a sort of picnic. The
bushie is disappointing to town Australians because he has no
swank, and nothing of the brilliant picturesqueness of the American
frontiersman. He is only a tall, gaunt man, lithe as a whip, with a
tongue like a whip-lash; and it is on bad trips or in battle that one
finds what he is like inside, a most knightly gentleman with a vein of
poetry.
Anyway the Melbourne people were cracked in 1860 when they
wanted an expedition to cross Australia northward, and instead of
appointing bushmen for the job selected tenderfeet. Burke was an
Irishman, late of the Hungarian cavalry, and the Royal Irish
Constabulary, serving as an officer in the Victorian police. Wills was
a Devon man, with some frontier training on the sheep runs, but had
taken to astronomy and surveying. There were several other white
men, and three Afghans with a train of camels.
They left Melbourne with pomp and circumstance, crossed
Victoria through civilized country, and made a base camp on the
Darling River at Menindie. There Burke sacked two mutinous
followers and his doctor scuttled in a funk, so he took on Wright, an
old settler who knew the way to Cooper’s Creek four hundred miles
farther on. Two hundred miles out Wright was sent back to bring up
stores from Menindie, while the expedition went on to make an
advanced base at Cooper’s Creek. Everything was to depend on the
storage of food at that base.
While they were waiting for Wright to come up with their stores,
Wills and another man prospected ninety miles north from Cooper’s
Creek to the Stony Desert, a land of white quartz pebbles and
polished red sandstone chips. The explorer Sturt had been there,
and come back blind. No man had been beyond.
Wills, having mislaid his three camels, came back ninety miles
afoot without water, to find the whole expedition stuck at Cooper’s
Creek, waiting for stores. Mr. Wright at Menindie burned time,
wasting six weeks before he attempted to start with the stores, and
Burke at last could bear the delay no longer. There were thunder-
storms giving promise of abundant water for once in the northern
desert, so Burke marched with Wills, King and Gray, taking a horse
and six camels.
William Brahe was left in charge at the camp at Cooper’s Creek,
to remain with ample provisions until Wright turned up, but not to
leave except in dire extremity.
Burke’s party crossed the glittering Stony Desert, and watching
the birds who always know the way to water, they came to a fine
lake, where they spent Christmas day. Beyond that they came to the
Diamantina and again there was water. The country improved, there
were northward flowing streams to cheer them on their way, and at
last they came to salt water at the head of the Gulf of Carpentaria.
They had crossed the continent from south to north.
With blithe hearts they set out on their return, and if they had to
kill the camels for food, then to eat snakes, which disagreed with
them, still there would be plenty when they reached Cooper’s Creek.
Gray complained of being ill, but pilfering stores is not a proper
symptom of any disease, so Burke gave him a thrashing by way of
medicine. When he died, they delayed one day for his burial; one
day too much, for when they reached Cooper’s Creek they were just
nine hours late. Thirty-one miles they made in the last march and
reeled exhausted into an empty camp ground. Cut in the bark of a
tree were the words “Dig, 21 April 1861.” They dug a few inches into
the earth where they found a box of provisions, and a bottle
containing a letter.
“The depot party of the V. E. E. leave this camp to-day to return
to the Darling. I intend to go S. E. from camp sixty miles to get into
our old track near Bulloo. Two of my companions and myself are
quite well; the third, Patten, has been unable to walk for the last
eighteen days, as his leg has been severely hurt when thrown from
one of the horses. No person has been up here from Darling. We
have six camels and twelve horses in good working condition.
William Brahe.”
It would be hopeless with two exhausted camels to try and catch
up with that march. Down Cooper’s Creek one hundred fifty miles the
South Australian Mounted Police had an outpost, and the box of
provisions would last out that short journey.
They were too heart-sick to make an inscription on the tree, but
left a letter in the bottle, buried. A few days later Brahe returned with
the industrious Mr. Wright and his supply train. Here is the note in
Wright’s diary:—

You might also like