You are on page 1of 253

Towards Scientific Literacy

A Teachers’ Guide to the History, Philosophy


and Sociology of Science

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
Derek Hodson - 978-90-8790-507-1
Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
Towards Scientific Literacy

A Teachers’ Guide to the History, Philosophy


and Sociology of Science

Derek Hodson
Ontario Institute for Studies in Education,
University of Toronto,
Canada

SENSE PUBLISHERS
ROTTERDAM / TAIPEI

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
A C.I.P. record for this book is available from the Library of Congress.

ISBN: 978-90-8790-505-7 (paperback)


ISBN: 978-90-8790-506-4 (hardback)
ISBN: 978-90-8790-507-1 (e-book)

Published by: Sense Publishers,


P.O. Box 21858, 3001 AW
3001 AW Rotterdam
Rotterdam, The Netherlands
http://www.sensepublishers.com

Printed on acid-free paper

All Rights Reserved © 2008 Sense Publishers

No part of this work may be reproduced, stored in a retrieval system, or transmitted in any
form or by any means, electronic, mechanical, photocopying, microfilming, recording or
otherwise, without written permission from the Publisher, with the exception of any material
supplied specifically for the purpose of being entered and executed on a computer system,
for exclusive use by the purchaser of the work.

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
To Susie, for your unwavering love, inspiration and support

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
Derek Hodson - 978-90-8790-507-1
Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
CONTENTS

Dedication…………………………………………………………………………..v

Preface…………………………………………………………………………...... ix

Acknowledgements……………………………………………………………….. xi

Chapter 1 In Pursuit of Scientific Literacy: The Case for History, Philosophy


and Sociology of Science………………………………………………1

Chapter 2 Exploring Nature of Science Issues: Students’ Views


and Curriculum Images……………………………………………..... 23

Chapter 3 The Traditional View of Science: Recognizing the Myths…………... 41

Chapter 4 Exploring Alternative Views of Science: The Ideas of Popper,


Lakatos and Kuhn…………………………………………………......67

Chapter 5 Scientific Inquiry, Experiment and Theory: What Should


We Tell Our Students?.......................................................................... 85

Chapter 6 Realism or Instrumentalism: What Position for School Science?.......103

Chapter 7 Insight from the Sociology of Science: Science is


What Scientists Do…………………………………………………..123

Chapter 8 Making a case for History of Science: Going Beyond Dates


and Anecdotes …………………………………………………….…149

Chapter 9 Looking for Balance in the Curriculum: Essential Elements


in a Curriculum for Critical Scientific Literacy…………….…….....173

Chapter 10 Further Thoughts on Social Construction and Scientific


Rationality: A View for School Science…………………………….199

References………………………………………………………………………..211

Index…………………………………………………………………………….. 239

vii

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
Derek Hodson - 978-90-8790-507-1
Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
PREFACE

My interest in history of science came early. As a schoolboy in the 1950s I always


looked forward to the frequently dull science lessons being enlivened by the stories
about scientists that sprinkled the pages of science textbooks by E.J. Holmyard and
F. Sherwood Taylor. Later, as an undergraduate at the University of Manchester
Institute of Science and Technology (UMIST)), I had the great good fortune to
attend a series of lectures on the philosophy of science by David Theobald (a
natural products chemist) and a wide-ranging course in history of science, from
Thales to Dalton, by Brian Pethica (a physical chemist). These were arranged as
part of a general studies course requirement, a curriculum planning decision by
UMIST for which I will always be grateful. Subsequently, I attended a course in
history of technology at UMIST taught by D.S.L. Cardwell, pioneer of history
of technology and author of The Organization of Science in England (1957) and
Turning Points in Western Technology (1972). Later still, as a student teacher at
the University of Exeter, I was privileged to study with the eminent historian of
science, H.J.J. Winter, author of Eastern Science: An Outline of its Scope and
Contribution (published in 1952). I hereby acknowledge my gratitude to the dedi-
cated and inspired teaching of all these scholars. These rich learning experiences
have led me to a lifelong interest in the history and philosophy of science, and
more recently in the sociology of science, and to efforts over many years to instill
elements of these vast subject areas into school science curricula, courses for teachers
at both pre-service and graduate level, professional development programmes for
serving teachers, and research agenda of graduate students. This book is an out-
come of those efforts. It is intended as a journey through the literature, picking
and choosing items that I consider of particular importance in developing students’
scientific literacy, as defined in chapter 1, and teachers’ capacity to present curricula
that afford a much higher profile to HPS than has been traditional – a goal that is in
line with much recent writing in science education and a number of prominent
reports on science education published in recent years in the United States, United
Kingdom and elsewhere.
There are a number of books that cover essentially the same ground – notably
works by Chalmers (1978, 1999), O’Hear (1989), Olby et al. (1990), Riggs (1992)
and Ladyman (2002). In contrast to the approach taken by these authors, this book
is written specifically from the perspective of science education. Its principal
concern is to identify what teachers, teacher educators, curriculum developers and
science education policy makers should know about the history, philosophy and
sociology of science in order to teach more effectively about the nature of science
and scientific activity. It is important to state at the outset that this is not a ‘how to’
book. It does not focus specifically on ways to design effective teaching and
learning activities, although such a book is currently in preparation. Rather, it is a
‘what’ and ‘why’ book, intended as a guide for making an appropriate selection
from the history, philosophy and sociology of science literature for presentation to
students.

ix

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
PREFACE

Although the original sources of the figures used in chapter 3 have been
acknowledged, I wish to state here that many of them first came to my attention in
the marvellous collection of images, photographs and quotations assembled by
David Wade Chambers (1984a,b) under the titles Putting Nature in Order and Is
Seeing Believing? These books constitute an invaluable resource in the design of
courses on the nature of science.

Derek Hodson
Toronto
November, 2007

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
ACKNOWLEDGEMENTS

Writing can be a very solitary pursuit, though it was not so in this case. In
preparing this book I drew on a lifetime of reading, thinking, teaching, researching,
writing and arguing about issues in science, philosophy of science, history of
science, sociology of science and science education, and on a lifetime of thinking
about how to organize and present these ideas to students at primary, secondary,
tertiary and graduate study level. So many memories of lively and interesting
people, encounters and incidents came flooding back – some pleasant, some
stressful and emotionally challenging, some amusing and some sad, but always
enlightening. In my mind’s eye, many people, from many different schools and
universities, were my constant companions as I struggled to make coherent sense
of difficult issues. I am immeasurably indebted to all those people and events,
spread over many years, for informing and influencing my thoughts about science.
There is no doubt that my thinking has been sharpened and my views refined
through numerous discussions with colleagues and graduate students, particularly
at the Ontario Institute for Studies in Education and the University of Hong Kong. I
hereby acknowledge my indebtedness to all those people, too numerous to identify
by name.
I also extend my thanks to the British Psychological Society, Houghton Mifflin
Harcourt Publishing Company, Wiley-Liss Inc., University of California Press,
University of Illinois Press and Dr. Richard Duschl for permission to reproduce
figures and tables previously published elsewhere.
– Figure 3.1b, titled Counterchange, is attributed to Michio Kubo. Source: E.H.
Gombrich (1978) The Sense of Order: A Study in the Psychology of Decorative
Art (figure 105, page 91), London: Phaidon Press. The citation in this book
reads: “from Hide and Seek, Osaka, 1968”. The same citation is given for this
illustration (figure 1.1, page 2) in D.W. Chambers (1984) Putting Nature in
Order, Geelong: Deakin University.
– Figure 3.4 is reproduced with permission of the author and the University of
Illinois Press. Source: E.G. Boring (1930) A new ambiguous figure. American
Journal of Psychology, 42, 444. Copyright 1930 by the Board of Trustees of the
University of Illinois.
– Figure 3.7 is reproduced by permission of Houghton Mifflin Harcourt
Publishing Company. Source: James J. Gibson (1977) The Perception of the
Visual World. Boston, MA: Houghton Mifflin (p. 182). Copyright © 1971, 1950
by Houghton Mifflin Company.
– Figure 3.8a is reproduced with permission from the British Journal of
Psychology, © The British Psychological Society. Source: L.S. Penrose & R.
Penrose (1958) Impossible objects: A special type of visual illusion. British
Journal of Psychology, 49(1), 31-33 (p. 32).

xi

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
ACKNOWLEDGEMENTS

– Figure 3.9 is reproduced with permission of the author and the University of
Illinois Press. Source: P.B. Porter (1954) Another puzzle picture. American
Journal of Psychology, 67, 550-551. Copyright 1954 by the Board of Trustees
of the University of Illinois.
– Figure 3.10 is reproduced with permission of the author and the University of
Illinois Press. Source: K.M. Dallenbach (1951) A puzzle picture with a new
principle of concealment. American Journal of Psychology, 64(3), 431-433.
Copyright 1951 by the Board of Trustees of the University of Illinois.
– Table 9.2 is reprinted with permission of Wiley-Liss, Inc., a subsidiary of John
Wiley & Sons, Inc. Source: Osborne, J., Collins, S., Ratcliffe, M., Millar, R. &
Duschl, R. (2003) What “ideas-about-science” should be taught in school
science? A Delphi study of the expert community. Journal of Research in
Science Teaching, 40(7), 692-720. Copyright (2003) National Association for
Research in Science Teaching.
– Figure 10.2 is reproduced with permission from Dr. Richard Duschl, Graduate
School of Education, Rutgers University, New Brunswick, NJ, and the
University of California Press. Source: R.A. Duschl (1990) Restructuring
Science Education: The Importance of Theories and their Development. New
York: Teachers College Press (p. 87). The figure was adapted from L. Laudan
(1984) Science and Values: The Aims of Science and their Role in Scientific
Debate. Berkeley, CA: University of California Press (p. 63).
Most importantly, no words can fully express my gratitude to my wife, Sue
Hodson, for her constant and unwavering love, support and encouragement
throughout our years together. You, Susie, make everything worthwhile.

xii

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
CHAPTER 1

IN PURSUIT OF SCIENTIFIC LITERACY


The Case for History, Philosophy and Sociology of Science

Nearly forty years ago, as a young and idealistic science teacher, I enthusiastically
embraced the Nuffield Science Teaching Project’s catchphrase “Being a scientist
for the day”.1 Since those heady days, other slogans and rallying calls have come
and gone, with varying degrees of impact on classroom practice – among them:
Process, not Product, Science for All, Less is More and Children Making Sense of
the World. Since the early to mid-1990s, the most prominent has been the call
for greater scientific literacy. Indeed, the notion of scientific literacy has assumed
centre-stage in science education debate in several parts of the world and organi-
zations such as the American Association for the Advancement of Science (AAAS,
1989, 1993), the Council of Ministers of Education, Canada (CMEC, 1997) and
UNESCO (1993) have used it to frame major efforts to reform the science curri-
culum. Three questions immediately spring to mind:
– What is scientific literacy?
– Why do we need it?
– What are the curricular implications?
Given its lengthy history in the rhetoric of science education, we could perhaps
expect there to be a clear and well articulated definition of scientific literacy.
Sadly, this is not the case. While the attainment of scientific literacy has been
almost universally welcomed as a desirable goal, there is still little clarity about its
meaning (Jenkins, 1990, 1994a, 1997; Krugly-Smolska, 1990; Eisenhart et al., 1996;
Millar, 1996; Sutman, 1996; Galbraith et al., 1997; Graber & Bolte, 1997; Hurd,
1998; DeBoer, 2000; Kolsto, 2000; Laugksch, 2000; Solomon, 2001; Tippens et al.,
2000; Cajas, 2001; Ryder, 2001; Rudolph, 2005), little consensus about why we
need it, and little agreement about precisely what it means in terms of curriculum
provision. As Laugksch (2000) observes, significantly different answers to these
questions are proferred by the various stakeholders in education, or “interest groups”
as he calls them. For example, he argues that the science education community
(science teachers, teacher educators and curriculum developers, for example) regard
scientific literacy as a kind of code for the goals of science education and frame
their discussion in terms of curriculum content, pedagogy and assessment/evaluation
procedures, while those with responsibility for science policy are more concerned
with public perception of, and support for, the scientific enterprise. Yet others are
concerned with the nature of control and priority setting for science, access to
science, or keeping the wider public up-to-date on significant scientific development
via the media, zoos and museums. As will become apparent in this and succeeding
chapters, I have interests in common with each of Laugksch’s “interest groups”.
My principal goal is that all students, regardless of gender, ethnicity, religion,
sexual orientation, geographical location and current attainment levels, achieve a
1

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
CHAPTER 1

measure of critical scientific literacy. I also subscribe to the view, outlined at length
in Hodson (2003), that science education is incomplete if it does not involve students
in preparing for and taking action on matters of social and political importance – a
position that Miller (1993) characterizes as transformational education.2
My use of the term “universal critical scientific literacy” signals my rejection of
the longstanding differentiation of science education into high status, academic/
theoretical courses for those deemed (on the basis of attainment tests) to be ‘high
ability students’ and low status courses oriented towards ‘life skills’ for the rest
(Hodson & Reid, 1988). In my view, we should draw later science specialists from
a much wider pool – hopefully approaching 100% of each age cohort – of successful,
enthusiastic students who have already achieved critical scientific literacy. If the
knowledge, skills and attitudes embodied in the notion of scientific literacy are
important, as I claim, they are important for everyone. Use of the term “universal
critical scientific literacy” carries with it a commitment to a much more rigorous,
analytical, skeptical, open-minded and reflective approach to science education
than many schools provide and signals my advocacy of a much more politicized
and issues-based science education, a central goal of which is to equip students with
the capacity and commitment to take appropriate, responsible and effective action on
matters of social, economic, environmental and moral-ethical concern (Hodson,
1999, 2003).
It almost goes without saying that scientific literacy presupposes a reasonable
level of literacy in its fundamental sense (Wellington & Osborne, 2001; Fang, 2005).
Engagement in science, contribution to debate about science and access to science
education are not possible without a reasonable level of literacy. As Anderson
(1999, p. 973) states: “reading and writing are the mechanisms through which
scientists accomplish [their] task. Scientists create, share, and negotiate the meanings of
inscriptions – notes, reports, tables, graphs, drawings, diagrams”. Scientific knowledge
cannot be articulated and communicated except through text, and its associated sym-
bols, diagrams, graphs and equations. The specialized language of science makes it
possible for scientists to construct an alternative interpretation and explanation of
events and phenomena to that provided by ordinary, everyday language. Indeed, it
could be said that learning the language of science is synonymous with (or certainly
coincident with) learning science, and that doing science in any meaningful sense
requires a reasonable facility with the language. It is scientific language that shapes
our ideas, provides the means for constructing scientific understanding and expla-
nations, enables us to communicate the purposes, procedures, findings, conclusions
and implications of our inquiries, and allows us to relate our work to existing knowl-
edge and understanding.
Because of the dependence of science on text, access to science also depends
on basic literacy, and those whose ability to read and write is poorly developed
are unlikely to achieve even a rudimentary level of scientific literacy, despite the
prodigious efforts of some teachers to convey scientific understanding through
drama, dance, film and other media. Proficient reading of science text involves more
than just recognizing all the words and being able to locate specific information, it
also involves the ability to infer meaning from the text – in particular, the meaning
intended by the author – and to establish relationships between ideas, link personal

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
IN PURSUIT OF SCIENTIFIC LITERACY

experiences with text, and transfer understanding from one context to another.
Thus, it involves analysis, interpretation and evaluation. In consequence, it depends
(in part) on what the reader brings to the task in terms of text interpretation strategies.
Despite the often considerable substantive content and the highly specialized lang-
uage of science, the abilities required to extract meaning from scientific text are
largely those required to extract meaning from any text, and while content know-
ledge is vitally important, it is by no means sufficient for a proper understanding
of scientific text. Indeed, Norris and Phillips (1994) and Phillips and Norris (1999)
have shown that high school students who score highly on traditional measures of
science attainment sometimes perform very poorly when asked to interpret media
reports of scientific matters. To paraphrase the conclusions of Norris and Phillips
(2003), understanding of science text resides in the capacity to determine when
something is an inference, a hypothesis, a conclusion or an assumption, to dis-
tinguish between an explanation and the evidence for it, and to recognize when
the author is asserting a claim to ‘scientific truth’, expressing doubt or engaging in
speculation. Without this level of interpretation, the reader will fail to grasp the
essential scientific meaning. Put simply, learning to think and reason scientifically
requires a measure of facility with the forms and conventions of the language of
science. It is not solely a matter of recognizing the words, and using them appro-
priately, but also the ability to comprehend, evaluate and construct arguments that
link evidence to ideas and theories. Thus, teaching about the language of science,
and its use in scientific argumentation should be a key element in science education
at all levels and there is a clear need for much closer cooperation between science
teachers (who need to know much more about the role and function of language)
and language arts teachers (who need to know much more about the specific
characteristics of scientific language).
If it is correct that most people, including many still in school, obtain most
of their knowledge of contemporary science and technology from television,
newspapers, magazines and the Internet (National Science Board, 1998; Select
Committee, 2000), then the capacity for active critical engagement with text is not
only a crucial element of scientific literacy, it is perhaps the fundamental element.
In that sense, education for scientific literacy has striking parallels with education
in the language arts. At the very least, students need to be able to read, understand
and evaluate scientific text in a wide variety of forms and styles (textbooks, teacher
handouts, newspaper and magazine articles, press releases and news briefs, Internet
postings and product labels, as well as graphs, diagrams, tables, chemical equations
and mathematical representations), convert empirical data acquired in laboratory
and fieldwork activities into text, and articulate and communicate their thoughts,
ideas, beliefs and feelings in ways that are intelligible to the intended audience,
whether it be peers, parents, teachers or the wider public. To be fully scientifically
literate, students need to be able to distinguish among good science, bad science
and non-science, make critical judgements about what to believe, and use scientific
information and knowledge to inform decision making at the personal, employ-
ment and community level. In other words, they need to be critical consumers of
science. This entails recognizing that scientific text is a cultural artifact, and so
may carry implicit messages relating to interests, values, power, class, gender,

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
CHAPTER 1

ethnicity and sexual orientation. In the words of Yore and Treagust (2006), “No
effective science education programme would be complete if it did not support
students in acquiring the facility of oral science language and the ability to access,
produce, and comprehend the full range of science text and representations” (p. 296).
What both these authors and I have in mind is a science education very different from
the traditional uncritical and unquestioning approach that presents science as dog-
matic, fixed and certain.
Scientific literacy also presupposes some basic understanding of mathematics,
such as familiarity with simple algebraic equations, the capacity to interpret gra-
phical and statistical data, and sufficient knowledge of the mathematics of prob-
ability to understand issues of risk and cost-benefit analysis. It also presupposes
some understanding of the dependence of science on mathematics. For example,
Kepler would not have derived his Laws of Planetary Motion without Greek
knowledge of conic sections accumulated some 1800 years earlier, Hilbert’s
theory of integral equations was essential for quantum mechanics and Riemann’s
differential geometry for Einstein’s theory of relativity. We can reasonably con-
clude that the great surge in scientific knowledge since the 17th Century can be
attributed, in large part, to developments in mathematics and, in particular, the
invention of differential and integral calculus.
What else should be regarded as crucial to a claim of being scientifically
literate? Understanding the nature of science? Understanding the major theoretical
frameworks of biology, chemistry and physics? Understanding the complex relation-
ships among science, technology, society and environment? Knowing about the
historical development of the ‘big ideas’ of science and the circumstances that led
to the development of key technologies? Being aware of contemporary applications
of science? Having the ability to use science in everyday problem solving? Holding
a personal view on controversial issues that have a science and/or technology
dimension? Possessing a basic understanding of global environmental issues?

WHAT IS SCIENTIFIC LITERACY?

The term scientific literacy seems to have first appeared in the US educational
literature in papers by Paul Hurd (1958) and Richard McCurdy (1958).3 It was
enthusiastically taken up by others as a useful rallying call (see Roberts, 1983,
2007), but had little in the way of precise or agreed meaning until Pella et al.
(1966) suggested that it comprises an understanding of the basic concepts of
science, the nature of science, the ethics that control scientists in their work, the
interrelationships of science and society, the interrelationships of science and the
humanities, and the differences between science and technology – with the first
three categories being designated as the most significant. Almost a quarter century
later, Science for All Americans (AAAS 1989, p. 4) drew upon very similar cate-
gories to define a scientifically literate person as “one who is aware that science,
mathematics, and technology are interdependent human enterprises with strengths
and limitations; understands key concepts and principles of science; is familiar with
the natural world and recognizes both its diversity and unity; and uses scientific
knowledge and scientific ways of thinking for individual and social purposes.”

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
IN PURSUIT OF SCIENTIFIC LITERACY

However, to suggest that debate had been stagnant during the intervening
years would be to seriously misinterpret matters. Indeed, following the work of
Milton Pella and his co-workers, there was a period of intense debate, definition
and counter-definition, marked by notable contributions from Daugs (1970), Agin
(1974), Showalter (1974), Klopfer (1976), O’Hearn (1976) and Pella (1976), and
culminating in Gabel’s (1976) detailed analysis of the literature in terms of eight
dimensions (organization of knowledge, intellectual processes, values and ethics,
process and inquiry, human endeavour, interaction of science and technology,
interaction of science and society, interaction of science, technology and society)
and nine categories of educational objectives (Bloom’s six categories of cognitive
objectives plus three affective objectives – valuing, behaving and advocating).
As Roberts (1983) comments, “What is immediately striking about Gabel’s model is
that it includes, under the definition of scientific literacy, every category of science
education objectives… it now means virtually everything to do with science edu-
cation” (p. 22).
Of course, interest in the notion of scientific literacy has not been restricted to
those concerned with science education in school and university. As Fitzpatrick
(1960) remarked, “If the Zeitgeist is to be favorable to the scientific enterprise,
including both academic and industrial programs, the public must possess some
degree of scientific literacy, at least enough to appreciate the general nature of
scientific endeavor and its potential contributions to a better way of life… no
citizen, whether or not he is engaged in scientific endeavors, can be literate in the
modern sense until he has understanding and appreciation of science and its work”
(p. 6). He concludes: “The ultimate fate of the scientific enterprise is in no small
degree dependent upon establishing a species of scientific literacy in the general
population” (Fitzpatrick, 1960, p. 169). At about the same time, Alan Waterman
(Director of the National Science Foundation) noted that it was a matter of
urgency that “the level of scientific literacy on the part of the general public be
markedly raised… progress in science depends to a considerable extent on public
understanding and support” (Waterman, 1960, p. 1349).
In the United Kingdom there has been a tradition of concern for the public
understanding of science dating back to the early years of the 19th Century (Shen,
1975; Jenkins, 1990). In more recent times the Royal Society (1985) noted that
“Improving the public understanding of science is an investment in the future; not a
luxury to be indulged in if and when resources allow” (p. 9). The Royal Society’s
argument that scientific literacy “can be a major element in promoting national
prosperity, in raising the quality of public and private decision making and in
enriching the life of the individual” (p. 9) serves to underline the key distinction
between those who see scientific literacy as the possession of knowledge, skills
and attitudes essential to a career as a professional scientist, engineer or technician
and those who see it as the capacity to access, read and understand material with a
scientific and/or technological dimension, make a careful appraisal of it, and use
that evaluation to inform everyday decisions, including those made at the ballot box.
According to Klopfer (1969), this distinction should be reflected in a differentiated
school science curriculum: “One curricular stream… designed for students planning
to enter careers as scientists, physicians, and engineers… the other… designed for

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
CHAPTER 1

students who will become the nonscientist citizenry… housewives, service workers,
salesmen etc…. Differentiation of students (should) begin at about age fourteen
when they choose the high school they will attend” (p. 203).
Distinctions are drawn somewhat differently by Shen (1975), who sees scientific
literacy as falling into three categories: practical, civic and cultural. Practical
scientific literacy is knowledge that can be used to help solve life’s everyday
problems; civic scientific literacy is the knowledge necessary to play a full part
in key decision-making in areas such as health, use of natural resources, energy
policy and environmental protection; cultural scientific literacy involves knowing
those ideas in science that represent major cultural achievements. Essentially the
same categorization lies behind the exercise devised by Robin Millar (1993) for
teachers and others in science education to justify (or not) the inclusion of parti-
cular topics in the school science curriculum on utilitarian, democratic or cultural
grounds. Wellington (2001) makes a similar point when he argues that there are
three sets of arguments for justifying curriculum content: (i) the intrinsic value of
science education; (ii) the citizenship argument; (iii) utilitarian arguments.
(i) Intrinsic Value
– Making sense of natural phenomena; de-mystifying them.
– Understanding our own bodies, our own selves.
– Interesting, exciting, and intellectually stimulating.
– Part of our culture, our heritage.
(ii) Citizenship Arguments
– Science knowledge and knowledge of scientists’ work are needed for all citizens
to make informed decisions in a democracy.
– Key decision makers (e.g., civil servants, politicians) need knowledge of
science, scientists’ work and the limitations of scientific evidence to make key
decisions, e.g., on foods, energy resources.
(iii) Utilitarian Arguments
– Developing generic skills that are of value to all, e.g., measuring, estimating,
evaluating.
– Preparing some for careers and jobs that involve some science.
– Preparing a smaller number for careers using science or as ‘scientists’.
– Developing important attitudes/dispositions: i.e., the ‘scientific attitude’ of
curiosity, wonder, healthy skepticism, an enquiring mind, a critical/analytical
disposition.
In contrast to Klopfer’s advocacy of a differentiated curriculum, Wellington
states that each orientation is important for every student. As a variant on this
principle of common provision, the authors of Beyond 2000: Science Education
for the Future (Millar & Osborne, 1998) state that science education between the
ages of 5 and 16 (the years of compulsory schooling in the UK) should comprise
a course to enhance general scientific literacy, with more specialized science
education delayed to later years: “the structure of the science curriculum needs
to differentiate more explicitly between those elements designed to enhance
‘scientific literacy’, and those designed as the early stages of a specialist training
in science, so that the requirement for the latter does not come to distort the
former” (p. 10).4

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
IN PURSUIT OF SCIENTIFIC LITERACY

While Branscomb (1981) expanded Shen’s categorization into eight forms of


scientific literacy (methodological science literacy, professional science literacy,
universal science literacy, technological science literacy, amateur science literacy,
journalistic science literacy, science policy literacy and public science policy
literacy), each related to a particular role in society, Shamos (1995) maintained
a three-fold categorization. However, unlike other writers, Shamos sees his cate-
gories as hierarchical. For him, cultural scientific literacy is the simplest, most
basic level of literacy. It comprises the basic understanding needed to (i) make
sense of articles in newspapers and magazines, and programmes on television, (ii)
communicate with elected representatives, and (iii) follow debates on public issues
with a science and technology dimension. Functional scientific literacy builds on
cultural scientific literacy by “requiring that the individual not only have command
of a science lexicon, but also be able to converse, read, and write coherently, using
such science terms in a perhaps non-technical but nevertheless meaningful context”
(Shamos, 1995, p. 8). True scientific literacy, as Shamos calls it, involves know-
ledge and understanding of major scientific theories, including “how they were
arrived at, and why they are widely accepted, how science achieves order out of a
random universe… the role of experiment… the importance of proper questioning, of
analytical and deductive reasoning, of logical thought processes, and of reliance on
objective evidence” (p. 89). Substantially the same hierarchy is proposed by Bybee
(1997): nominal scientific literacy (knowing scientific words but not their meaning),
functional scientific literacy (reading and writing science using simple and appro-
priate vocabulary) and conceptual and procedural scientific literacy (a thorough
understanding of both the conceptual and procedural bases of science). Bybee adds
a fourth category: multidisciplinary scientific literacy – a thorough and robust
understanding of the conceptual and procedural structures of science, together with
knowledge of the history of science, an understanding of the nature of science and
appreciation of the complex interactions among science, technology and society.
As Fensham (2002) argues, the first category is no literacy at all, the second
is “functional only in a direct vocabulary sense, and not in any generalized
operational sense” (p. 17), while the fourth may comprise an unrealistic target for
many students.
A general criticism of these attempts to define scientific literacy is that they are
couched in terms of what scientists and/or science educators regard as essential
knowledge and understanding. An entirely different approach to the notion of
functional scientific literacy was taken by researchers at the University of Leeds.
Loosely structured interviews were used to ascertain the “practical knowledge in
action” assembled and deployed by individuals or groups of adults engaged in
understanding and/or acting upon some issue with a scientific dimension, such as
managing a domestic energy budget within a low, fixed income or raising a child
born with Down’s Syndrome (Layton et al., 1986, 1993; Jenkins, 1996a, 1999). It
quickly became clear that the science needed for solving the problems of everyday
life is very different in form from the science presented in the school curriculum.
For example, says Jenkins (1999, p. 705), it makes more sense in a practical
context to regard heat as “something which ‘flows’ rather than in terms of the
‘more correct’ kinetic theory of matter”. The ways in which everyday problems

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
CHAPTER 1

are confronted and solved is also very different from the way problem-solving is
approached in school science lessons.
‘Citizen thinking’, i.e., everyday thinking, turns out to be much more complex
and less well understood than scientific thinking and, as might be expected,
well adapted to decision-making in an everyday world which, unlike science
itself, is marked by uncertainty, contingency and adaptation to a range of
uncontrolled factors. (Jenkins, 1999, p. 704)
This conclusion, which has much in common with the now extensive literature in
situated cognition, has prompted Peter Fensham (2002) to state that it is “time to
change drivers for scientific literacy” and to abandon the traditional ways of iden-
tifying science content knowledge for the school curriculum. More in line with
Fensham’s recommendation than most contemporary science education programmes
would be a curriculum designed in accordance with the findings described by Law
(2002) arising from a study in which she and her co-researchers asked leading
scientists, health care professionals, local government representatives, managers
and personnel officers in manufacturing industry, and the like, about the kind of
science and the kind of personal attributes and skills that are of most value in
persons employed in their field of expertise. Similarly, Duggan and Gott (2002)
sought to identify the science used by employees in five science-based industries:
a chemical plant manufacturing cosmetics and pharmaceuticals, a biotechnology
company producing diagnostic kits for medical use, an environmental analysis
laboratory, an engineering company manufacturing pumps for the petrochemical
industry and an arable farm. What they found was that most of the necessary science
was learned on-the-job rather than in school. This finding is mirrored in work by
Chin et al. (2004) and Aikenhead (2005). The latter study concluded that school
science is “focused predominantly on declarative knowledge while workplace
knowledge is focused predominantly on procedural knowledge” (p. 129) – that is,
‘knowing that’ is emphasized in school and ‘knowing how’ in the workplace. These
studies raise important questions about the purpose of scientific literacy and the
motives of the different stakeholders in promoting it.
Before proceeding to discussion of these matters, it is worth noting that scien-
tific literacy has a significant metacognitive dimension. Students need to know
what they know, how and when that knowledge can and should be utilized, how to
recognize deficiencies in their knowledge and how to compensate for them. It is
metacognitive knowledge and skills that enable an individual to promote and monitor
her or his learning. Pintrich (2002) refers to three categories of metacognitive know-
ledge: (i) strategic knowledge – knowledge of general thinking skills to facilitate
learning and problem solving; (ii) knowledge about cognitive tasks – understanding
when and how to apply various strategies; and (iii) self knowledge – knowledge of
one’s own strengths and weaknesses. A number of authors have argued that all
meaningful construction and evaluation of knowledge, whether by means of reading,
talking or writing, depend largely on metacognition (Holliday et al., 1994; Keys,
1999; Wallace et al., 2003; Wallace, 2004). Effective readers apply a knowledge of
their own strengths and weaknesses to organize prior knowledge and relate it to new

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
IN PURSUIT OF SCIENTIFIC LITERACY

information, and to elaborate on the ideas they locate in the text; effective writers
consistently judge the match between the information and ideas they are trying to
articulate and the language they are using to represent them. To construct a scientific
argument, whether it is delivered orally or in written form, it is necessary to monitor
and evaluate the match among the various components of the argument

WHY DO WE NEED SCIENTIFIC LITERACY?

Reviewing what they describe as an extensive and diverse literature, Thomas and
Durant (1987) identify a range of arguments for promoting the public understanding
of science, including benefits to science, benefits to individuals and benefits to
society as a whole.
Benefits to science are seen largely in terms of increased numbers of recruits, greater
support for scientific research and more realistic public expectations of science.
Notwithstanding the massive research endeavours of the pharmaceutical industry, the
electronics industry, the armaments and aircraft industry, and other business enter-
prises, a great deal of the financial support for fundamental scientific research
derives from public funds.5 Thus, it is argued, the self-interests of scientists demands
that they keep the tax-payer well informed about what scientists do, and how they
validate their research findings and theoretical conclusions, because a well-informed
public is more likely to be supportive of such high levels of financial investment. In
developing this line of argument, Schwab (1962) advocated a shift of emphasis away
from the learning of scientific knowledge (the products of science) towards an
understanding of the processes of scientific inquiry because it can ensure “a public
which is aware of the conditions and character of scientific enquiry, which under-
stands the anxieties and disappointments that attend it, and which is, therefore,
prepared to give science the continuing support which it requires” (p. 38). Jenkins
(1994b) makes the related point that enhanced public understanding of science
enables scientists to be more effective in countering opposition from religious
fundamentalist groups, animal rights activists and others who might seek to constrain
or curtail scientific inquiry (and science education). A related argument advanced
by Shamos (1993) is that enhanced scientific literacy is a defence against what he
sees as the anti-science and neo-Luddite movements that are (in his words)
“threatening to undermine science”. The school science curriculum, he argues,
“should be the forum for debunking the attempts of such fringe elements to distort
the public mind, first by exposing their tactics, and then by stressing over and over
again the central role in science of objective, reproducible evidence” (p. 71).
Some years ago, Isaac Asimov (1983) suggested that “without an informed
public, scientists will not only be no longer supported financially, they will be
actively persecuted” (p. 109). More recently, Stuewer (1998) has provided a wonder-
fully colourful description of the scientist as “exactly what the medicine-man is for
the savage: namely, a mysterious ambivalent figure, who is to be worshipped as the
carrier of recondite knowledge and the agent of recondite powers; and who is at the
same time to be feared, even hated, and to be put in his place. The medicine-man
may be a power, but he is a very acceptable sacrifice to the gods” (Wiener (1950),
cited by Stuewer, p. 25). While open hostility towards science and scientists is still

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
CHAPTER 1

rare and the prospect of sacrifice remote, there is certainly much public suspicion
of science and a significant decline in public confidence in scientists, largely as
a consequence of the BSE (bovine spongiform encephalopathy) or ‘Mad Cow
Disease’ episode in the United Kingdom, anxiety about the environmental impact
of genetically engineered crops, the conspicuous failure of scientists over many
years to reach consensus about the causes and significance of global warming
and climate change,6 the appearance of scientists as ‘expert witnesses’ for both sides
in legal disputes (for example, concerning the practices of tobacco companies),
the use of ‘purpose-built’ scientific research in advertizing to promote particular
business interests, increasing domination of scientific research by industrial and
military interests, and the sometimes cavalier disregard that some scientists exhibit
towards the social and ethical issues raised by contemporary science and techno-
logy. As Barad (2000) comments, “the public senses that scientists are not owning
up to their biases, commitments, assumptions, and presuppositions, or to base
human weaknesses such as the drive for wealth, fame, tenure, or other forms of
power” (p. 229). In making his case for increased levels of scientific literacy,
Shortland (1988) states that confidence in scientists and public support for science
depend on “at least a minimum level of general knowledge about what scientists
do” (p. 307). More significantly, support depends on whether the public values
what scientists do. It is naïve to assume that enhanced scientific literacy will
inevitably translate into simple trust of scientists and unqualified support for the
work they choose to do. A scientifically literate population, with a rational view
of the world, a predisposition to think critically and the capacity to appraise
scientific evidence for themselves, is perhaps much more likely to be skeptical,
suspicious or even distrustful of scientists, and much more likely to challenge
the nature of scientific research and the direction of technological innovation. Of
course, while increased critical scrutiny of science may ensue, increased public
control of the scientific enterprise is not guaranteed by enhanced scientific literacy
alone, even if universal. What is needed is scientific literacy allied to political
literacy and a commitment to sociopolitical involvement – hence my arguments
elsewhere (Hodson, 1994, 2003) for a curriculum oriented toward citizen action.
Arguments that scientific literacy confers benefits on individuals come in a
variety of guises. For example, it is commonly argued that scientifically literate
individuals have access to a wide range of employment opportunities and are
well-positioned to respond positively and productively to the introduction of new
technologies into the workplace: “More and more jobs demand advanced skills,
requiring that people be able to learn, reason, think creatively, make decisions, and
solve problems. An understanding of science and the process of science contri-
butes in an essential way to these skills” (National Research Council, 1996, p. 2).
Moreover, those who are scientifically literate are better able to cope with the
demands of everyday life in an increasingly technology-dominated society, better
positioned to evaluate and respond appropriately to the supposed “scientific
evidence” used by advertizing agencies and politicians, and better equipped to
make important decisions that affect their health, security and economic well-
being.

10

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
IN PURSUIT OF SCIENTIFIC LITERACY

Personal decisions, for example about diet, smoking, vaccination, screening


programmes or safety in the home and at work, should all be helped by some
understanding of the underlying science. Greater familiarity with the nature
and findings of science will also help the individual to resist pseudo-scientific
information. An uninformed public is very vulnerable to misleading ideas on,
for example, diet or alternative medicine. (Royal Society, 1985, p. 10)

When people know how scientists go about their work and reach scientific
conclusions and what the limitations of such conclusions are, they are more
likely to react thoughtfully to scientific claims and less likely to reject them
out of hand or accept them uncritically. (AAAS, 1993, p. 3)
There is a cluster of arguments that focus on the intellectual, aesthetic and
moral-ethical benefits conferred on individuals by scientific literacy. In the first
two arguments there are strong echoes of C.P. Snow’s (1962) assertion that science
is “the most beautiful and wonderful collective work of the mind of man” (p. 14)
and, therefore, as crucial to contemporary culture as literature, music and fine
art. Indeed, Dawkins (1998) asserts that “the feeling of awed wonder that science
can give us is one of the highest experiences of which the human psyche is
capable. It is a deep aesthetic passion to rank with the finest that music and poetry
can deliver” (p. x). Notwithstanding the sexist language so common forty years
ago, Warren Weaver perfectly encapsulates this particular rationale for scientific
literacy:
The capacity of science progressively to reveal the order and beauty of the
universe, from the most evanescent elementary particle up through the atom,
the molecule, the cell, man, our earth with all its teeming life, the solar system,
the metagalaxy, and the vastness of the universe itself, all this constitutes the
real reason, the incontrovertible reason, why science is important, and why its
interpretation to all men is a task of such difficulty, urgency, significance and
dignity. (Weaver, 1966, p. 50)
The assertion that the ethical standards and code of responsible behaviour
acquired through scientific literacy will lead to more ethical behaviour in the wider
community is a particularly fascinating one. Paraphrasing the arguments of Jacob
Bronowski, Shortland (1988) summarizes the rationale as follows: “the internal
norms or values of science are so far above those of everyday life that their transfer
into a wider culture would signal a major advance in human civilization” (p. 310).
Harre (1986) presents a similar argument: “the scientific community exhibits a model
or ideal of rational cooperation set within a strict moral order, the whole having
no parallel in any other human activity” (p. 1). The authors of Science for All
Americans (AAAS, 1989) spell out some of these moral values as follows: “Science
is in many respects the systematic application of some highly regarded human
values – integrity, diligence, fairness, curiosity, openness to new ideas, skepticism,
and imagination” (p. 201). Studying science, scientists and scientific practice will,
they argue, help to instill these values in students.

11

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
CHAPTER 1

In other words, the pursuit of scientific truth regardless of personal interests,


ambitions and prejudice (part of the traditional image of the objective and dis-
passionate scientist) makes science a powerful carrier of ethical principles. Scientific
literacy doesn’t just result in more skilled and more knowledgeable people, it
results in wiser people, well equipped to make morally and ethically superior
decisions. Given the findings from sociology of science to be discussed in chapter 7,
there are good reasons to dismiss this claim as a ludicrous flight of fancy and to
argue, instead, that science would benefit from transfer of ethical standards in the
opposite direction.
Arguments that increased scientific literacy would confer benefits on society as
a whole include (i) the familiar and increasingly pervasive economic argument
(closely linked, of course, with military and ideological arguments7), (ii) claims
that it enriches of the cultural health of the nation and intellectual life in general,
and (iii) belief in its capacity to enhance democracy and promote responsible
citizenship In recent years, the economic argument has become the predominant
rationale for scientific literacy in North America (Garrison & Lawwill, 1992). It is a
powerful and persuasive one, as illustrated by the Government of Canada’s
(1991) attempt to establish a link between school science education and a culture of
lifelong learning as the key to the country’s prosperity.
Our future prosperity will depend on our ability to respond creatively to the
opportunities and challenges posed by rapid change in fields such as informa-
tion technologies, new materials, biotechnologies and telecommunications...
To meet the challenges of a technologically driven economy, we must not
only upgrade the skills of our work force, we must also foster a lifelong learn-
ing culture to encourage the continuous learning needed in an environment of
constant change. (Government of Canada, 1991, pp. 12 & 14)
Similarly, the authors of an Ontario Ministry of Education and Training (2000a)
document on curriculum planning and assessment state that the curriculum has been
designed to ensure that its graduates are well prepared “to compete successfully in
a global economy and a rapidly changing world” (p. 3). Thus, scientific literacy is
regarded as a form of human capital that sustains and develops the economic well-
being of a nation. Put simply, continued economic development brought about by
enhanced competitiveness in international markets (regarded as incontrovertibly a
“good thing”) depends on science-based research and development, technological
innovation and a steady supply of scientists, engineers and technicians, all of which
ultimately depend on public support for state-funded science and technology edu-
cation in school. Moreover, the argument goes, increased scientific literacy is likely
to sustain high levels of consumer demand for technologies perceived by such
scientifically literate individuals as desirable. An indication of the extent to which
this ideology has permeated the Ontario (Canada) curriculum can be gleaned from
a simple count of key words in the Technological Education Curriculum for Grades
11 and 12 (Ministry of Education & Training, 2000b): markets/marketing = 46,
consumer = 15, client/customer = 72, management = 59, industry = 217,
sustainability = 0, recycle = 0, ecosystem = 0, interdependence = 0, values = 0

12

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
IN PURSUIT OF SCIENTIFIC LITERACY

(Elshof, 2001). It is also worth noting that the People and Skills in the New Global
Economy report (Ontario Premier’s Council, 1990) was prepared by a panel
comprising five academics, seven politicians three trade union leaders and nineteen
company presidents (Davis, 2000). It is also noteworthy that there were only four
women among the 34 council members.
Before proceeding further, it is worth noting that language is not neutral. Those
who gain control of common language usage determine how we conceptualize
important issues; they set the agenda for debate and, in effect, determine the way
we think and what we value. All language carries a substantial sub-textual cargo of
meaning intended to create a particular view of the world.8 Its power is located in
the ways in which it determines how we think about society and our relations with
others, and in its impact on how we act in the world. When deployed effectively, it
creates a particular social reality. Indeed, rhetoric becomes reality and those who
think differently, and have different values, are regarded as deviant or aberrant.
Strategies include presenting one’s own position as natural or as plain common
sense, thus implying that there is a conspiracy among one’s opponents to deny the
truth or to promote what is fashionable or ‘politically correct’ (itself a term that has
acquired substantial pejorative connotations).
This is a powerful technique. First, it assumes that there are no genuine
arguments against the chosen position; any opposing views are thereby posi-
tioned as false, insincere or self-serving. Second, the technique presents the
speaker as someone brave or honest enough to speak the (previously)
unspeakable. Hence the moral high ground is assumed and opponents are
further denigrated. (Gillborn, 1997, p. 353)
Hence, when school science lessons present students, almost daily, with a lang-
uage that (i) promotes economic globalization, increasing production and unlimited
expansion, (ii) sees unfettered technological production and spiraling consumption
as “progress”, (iii) regards job satisfaction as the accumulation of wealth and
material goods, and (iv) equates excellence with competition and “winning at any
cost”, it is co-opted into the manufacture and maintenance of what Bowers (1996,
1999) calls the myths of modernity: “that the plenitude of consumer goods and
technological innovation is limited only by people’s ability to spend, that the indivi-
dual is the basic social unit… that science and technology are continually expanding
humankind’s ability to predict and control their own destiny” (1996, p. 5). At risk
from this new “reality” are the freedoms of individuals who think differently, the
spiritual well-being of those who would live differently, and the integrity of the
planet’s complex and delicate ecosystems. In Edmund O’Sullivan’s (1999) words:
Our present educational institutions which are in line with and feeding into
industrialism, nationalism, competitive transnationalism, individualism, and
patriarchy must be fundamentally put into question. All of these elements
together coalesce into a world view that exacerbates the crisis we are now
facing. (p. 27)

13

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
CHAPTER 1

Many in the economically under-developed parts of the world see globalization


as a renewed form of colonization, a form of cultural and economic imperialism
that threatens to destroy rather than foster their economic and social well-being,
that commodifies both natural resources and people, locates them on the periphery
of key decision-making (or excludes them entirely) and compounds their power-
lessness, poverty and dispossession (Muchie & Xing, 2006). As Roseneau (1992)
observes, the self-serving policies of international organizations such as the IMF,
World Bank, OECD and G-7 in many under-developed and developing nations
constitute “governance without government”.
While neo-liberalism may mean less government, it does not follow that there
is less governance. While… neo-liberalism problematises the state and is
concerned to specify its limits through the invocation of individual choice…
it involves forms of governance that encourage both institutions and indivi-
duals to conform to the norms of the market. (Larner, 2000, p. 12 – cited by
Bencze & Alsop, 2007, p. 3)

On the surface, [globalization] is instant financial trading, mobile phones,


McDonald’s, Starbuck’s, holidays booked on the net. Beneath this gloss, it is
the globalization of poverty, a world where most human beings never make
a phone call and live on less than two dollars a day, where 6,000 children
die every day from diarrhoea because most have no access to clean water.
(Pilger, 2002, p. 2)
It is clear that little of the world’s poverty, injustice, terrorism and war will be
eliminated, and few among the litany of environmental crises (ozone depletion;
global warming; land, air and water pollution; deforestation; desertification; and so
on) will be solved, without a major shift in the practices of western industrialized
society and the values that sustain them. Interestingly, one of the keys to
ameliorating the current situation may lie in increased levels of scientific literacy
among the world’s citizens – an idea explored a little in the following section and
at length in Hodson (2003).
The life-enhancing potential of science and technology cannot be realized
unless the public in general comes to understand science, mathematics and
technology and to acquire scientific habits of mind; without a scientifically
literate population, the outlook for a better world is not promising. (AAAS,
1989, p. 13)
The case for scientific literacy as a means of enhancing democracy and res-
ponsible citizenship, and resisting the consumer juggernaut, is just as strongly
made as the economic argument, though by different stakeholders and interest
groups. Thomas and Durant (1987) note that increased scientific literacy “may be
thought to promote more democratic decision-making (by encouraging people to
exercise their democratic rights), which may be regarded as good in and of itself;
but in addition, it may be thought to promote more effective decision-making

14

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
IN PURSUIT OF SCIENTIFIC LITERACY

(by encouraging people to exercise their democratic right wisely” (p. 5). Whether
wisdom is the likely outcome of enhanced scientific literacy in the wider community
depends crucially, of course, on how notions of scientific literacy are translated
into curriculum practice. In similar vein, Chen and Novick (1984) express concern
about the fragility of democracy and see enhanced scientific literacy as a means
“to avert the situation where social values, individual involvement, responsibility,
community participation and the very heart of democratic decision making will be
dominated and practiced by a small elite” (p. 425). This, of course, presupposes
that we aim for universal scientific literacy, as argued above. What is clear is that
democracy is strengthened when all citizens are equipped to evaluate Socioscientific
issues (SSI) and make informed decisions on matters of personal and public concern.
Enhancing scientific literacy might also be the most effective way to combat (i)
the naïve trust that many students have in the Internet, accepting almost anything
and everything as equally valid and reliable, and forming their views on all manner
of topics on the basis of half an hour of exploration with Google, (ii) the distrust
with which many people regard any argument that deploys statistics (because, they
say, “statistics can prove anything!”), (iii) the increasing tendency to succumb
blindly to the seductions of the now all-too-prevalent “alternative sciences” such as
iridology and reflexology, (iv) the increasing susceptibility of people to the blandish-
ments of purveyors of “miracle cures”, “revolutionary diets”, “body enhancement”
techniques and procedures, and the healing properties of crystals, and (v) the conti-
nuing fascination of so many with astrology, ESP, “ancient astronaut” theories
and spurious “mysteries” such as the Bermuda Triangle. Interestingly, and with
characteristic idiosyncrasy, Dawkins (1998) develops the argument that enhanced
scientific literacy results in better decision making into an assertion that “lawyers
would make better lawyers, judges better judges, parliamentarians better parlia-
mentarians and citizens better citizens if they knew more science, and more to the
point, if they reasoned more like scientists”. This is, in essence, the case being
argued in this chapter, although my focus is better decisions by all citizens
(whether they are journalists, business people, civil servants, teachers, police
officers, or whatever) in the context of SSI. It is also my contention that the kind of
critical scientific literacy I have in mind will serve to help scientists make better
decisions, in the sense that their judgements focus more attention on the economic,
social, cultural, political and moral-ethical dimensions of their work.

WHAT CAN WE CONCLUDE?

Where does all this propaganda for scientific literacy leave us? Can we be confi-
dent that almost half a century of debate has finally answered the three questions
posed at the beginning of the chapter? It seems that the case for greater scientific
literacy, and the kind of curriculum proposals that would follow from it, change
with social context. They are a product of their time and place: they do not easily
cross national or cultural boundaries (Tippens et al., 2000) and do not transfer
comfortably from one era to another. Ogawa (1989), for example, expresses grave
reservations about the inherent neo-colonialist aspects of ‘transplanting’ Western
conceptions of scientific literacy into the education systems of non-Western

15

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
CHAPTER 1

countries, and Clay (1996) remarks: “In its very structure our view of the world is
so deeply imbued with the dominant scientific method that to encompass the
multiplicity of other equally valid views held in societies beyond our own, we
need to encompass different scientific literacies” (p. 190). These issues, the validity
of Clay’s claim that there are such alternative literacies, and discussion of various
attempts to incorporate traditional knowledge into a curriculum that also presents
Western science are outside the scope of this book (but see Aikenhead, 2002;
Gitari, 2003, 2006; Palefau, 2005).
It almost goes without saying that as science itself changes and develops, so our
view about what counts as legitimate scientific literacy also changes. Apart from
the need to update curriculum content to keep pace with our rapidly expanding
scientific knowledge, there is an urgent need to acknowledge and respond to the
many changes in the “social and economic characteristics, ethos, practice, and
culture of science” (Hurd, 2002, p. 5). The way science is presented in many
conventional science curricula bears very little resemblance to the kind of research
carried out in the laboratories of the early 21st century, and the values that underpin
this research are very far removed from the traditional portrayal of science as the
disinterested pursuit of objective truth. With the increasing industrialization and
militarization of the scientific enterprise, for example, previous claims for the cultural
and ethical value of scientific literacy seem hopelessly misplaced, if not downright
dishonest (Jenkins, 1990), and while the claims that scientific literacy builds
economic prosperity may have become more plausible and inviting to some, they
have become less morally and ethically defensible in respect of environmental
impact and social, cultural and economic consequences for the less fortunate
members of society, both within and beyond the developed world.
So are there, one might ask, any elements of scientific literacy that are valid in
all contexts and for all time? My answer is “Yes”, if scientific literacy means
knowing what scientific resources to draw on, where to find them and how to use
them (Fourez, 1997) – including the “proper use of scientific experts” (Shamos,
1995, p. 217). My answer is “Yes”, if the real function of scientific literacy is to
confer a measure of intellectual independence, to help people learn how to think
for themselves and to reach their own conclusions about a range of issues that have
a scientific and/or technological dimension. My answer is “Yes”, if scientific lite-
racy is sought not because it improves the economy, produces more “technological
goodies” or provides more job opportunities for individuals, but because it libe-
rates the mind. As the authors of Benchmarks for Scientific Literacy (AAAS,
1993) suggest, “People who are literate in science... are able to use the habits of
mind and knowledge of science, mathematics, and technology they have acquired
to think about and make sense of many of the ideas, claims, and events that they
encounter in everyday life” (p. 322). More recently, the OECD’s Programme for
International Student Achievement (PISA) proposed that a scientifically literate
person is “able to combine science knowledge with the ability to draw evidence-
based conclusions in order to understand and help make decisions about the natural
world and the changes made to it through human activity” (OECD, 1998, p. 5) and
has “a willingness to engage in science-related issues, and with the ideas of science,
as a reflective citizen… having opinions and participating in… current and future

16

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
IN PURSUIT OF SCIENTIFIC LITERACY

science-based issues” (OECD, 2006, p. 24). There are strong echoes here of
Arons’ (1983) emphasis on the ability to “discriminate, on the one hand, between
acceptance of asserted and unverified end results, models, or conclusions, and on
the other, understand their basis and origin; that is, to recognize when questions
such as “How do we know?” “Why do we believe it?” “What is the evidence for
it?” have been addressed, answered, and understood, and when something is being
taken on faith” (p. 93). Similar capabilities have sometimes been included in the
notion of intellectual independence (Munby 1980; Aikenhead 1990; Norris 1997).
Without such capabilities, citizens are “easy prey to dogmatists, flimflam artists,
and purveyors of simple solutions to complex problems” (AAAS, 1989, p. 13) –
including, one might add, some otherwise respectable scientists, politicians and
commentators who seek to intimidate the public through their facility in a mode of
discourse unfamiliar to many citizens.
Dearden (1968) argued that personal autonomy (the achievement of which, for
him, was the prime purpose of education) has three major elements: first, an inde-
pendence from authorities; second, a disposition to test the truth of things for
oneself; third, an ability to deliberate, form intentions and choose in accordance
with a scale of values that is self-formulated. In Guy Claxton’s (1991) words,
“Nothing could be of greater value than the ability to make your own life up as you
go along: to find for yourself what is satisfying; to know your own values and
your own mind; to meet uncertainty with courage and resourcefulness; and to
appraise what others tell you with an intelligent and healthy skepticism” (p. 130).
Of course, in many aspects of modern life we are increasingly dependent on
‘experts’ and ‘authorities’ of various kinds. When dealing with socioscientific
issues and appraising new technologies, individuals will only rarely have access to all
the relevant data. In consequence, we depend on others to inform us and advise us.
For example, we are increasingly dependent on scientists, the inquiries they conduct,
and the agencies that report their studies, to tell us about the safety hazards
associated with various products and procedures, the toxic effects of pesticides,
pharmaceuticals and other materials we encounter in everyday life, the risks asso-
ciated with post-menopausal HRT and the optimal frequency of mammograms,
the threats to our health posed by the proximity of toxic waste dumps, nuclear
power plants and overhead power lines, and the large scale compromising of envi-
ronmental health through loss of biodiversity, increasing desertification, pollution
and global warming. It is crucial, therefore, that each of us understands how reliable
and valid data are collected and interpreted, and that each of us recognizes the
tentative character of scientific knowledge. It is crucial, too, that we understand
the ways in which all manner of human interests can and do shape the inquiry
and its interpretation and reporting. Without this insight, we have no alternative
but to take reports that blame or exonerate at face value, and to accept all claims
to scientific knowledge as ‘proven’. In a very real way, critical scientific literacy,
and the intellectual independence it bestows, enables us to decide which experts
we can trust and rely on. As Ungar (2000) remarks, “after a decade of clipping
articles from Science and Nature, my sense that climate change is real ultimately
boils down to picking the experts you think you can rely on” (p. 297). Ratcliffe and
Grace (2003) cite a study published by the Office of Science and Technology and

17

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
CHAPTER 1

the Wellcome Trust, in 2000, indicating that people tend to trust sources seen
as neutral and independent, such as university scientists, scientists working for
research charities or health campaigning groups, television news and documentaries,
for example. The least trusted sources are politicians and newspapers. Sources
seen as having a vested interest, such as environmental activist groups, well-
known scientists and the popular scientific press, rank somewhere in between in
terms of trustworthiness. In Elliott’s (2006) study, students were particularly skep-
tical about the relationship between science, the media and government. There is
also an ‘asymmetry of trust’: episodes that weaken or threaten trust in science tend
to receive greater exposure in the media and live longer in the public memory than
episodes that build or consolidate confidence in science and scientists. My inter-
pretation of the notion of critical scientific literacy encompasses the capacity to
read reports involving science (in all forms of communication in an informed and
critical way in order to form one’s own judgement about what to believe, what to
doubt and what to reject.
Hurd (1998) sums up this critical dimension of scientific literacy, and its roots
in learning about science, when he defines a scientifically literate person as
someone who “distinguishes and recognizes expertise, dogma, pseudoscience, the
temporal nature of knowledge, effective argumentation, and relationships among
claims, evidence, and warrants” (p. 24). What Hurd doesn’t mention is that this
kind of understanding needs to be developed in such a way that students can see
the sociopolitical embeddedness of science and technology. If science continues to
be presented as an exercise in abstract puzzle solving, devoid of social, political,
economic and cultural influences and consequences, citizens will continue to see
contemporary SSI as largely ‘technical problems’, for which experts can be relied
upon to provide the answers. What we should be seeking is political engagement
of citizens in monitoring and, to an extent, directing the course of scientific and
technological development.
It is timely, then, that the so-called Crick Report, Education for Citizenship
and the Teaching of Democracy in Schools, has prompted the establishment of
citizenship education comprising three strands – social and moral responsibility,
community involvement, political literacy – as a mandatory part of the curriculum
of all subjects in England and Wales. The declared aim of this initiative is:
… a change in the political culture of this country both nationally and locally:
for people to think of themselves as active citizens, willing, able and equipped
to have an influence in public life and with the critical capacities to weigh
evidence before speaking and acting; to build on and to extend radically to
young people the best in existing traditions of community involvement and
public service, and to make them individually confident in finding new forms
of involvement and action. (Qualifications & Curriculum Authority, 1998, p. 8)
In practice, as Davies (2004) reminds us, not all science educators who are keen
to implement science education for citizenship have a clearly articulated notion of
what responsible citizenship entails and how science education can play a part in
helping students achieve it. He quotes at length from Gamarnikow and Green’s

18

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
IN PURSUIT OF SCIENTIFIC LITERACY

(2000) argument that it so often “reproduces a version of citizenship education


unlikely to challenge the social mechanisms of inequality reproduction” (p. 1757).
I tend to agree that there is a depressing tendency to equate science education for
citizenship with the inclusion of common everyday examples as a way of motivating
students and enhancing conceptual (and possibly procedural) understanding. In other
words, the citizenship element is a mere enabling tactic; the real goal is enhanced
understanding of science content. However, I am enormously encouraged by the
authors of Science For All Americans (AAAS, 1989, p. 12), who direct attention
towards scientific literacy for a more socially compassionate and environmentally
responsible democracy when they assert that science can provide knowledge “to
develop effective solutions to… global and local problems” and can foster “the
kind of intelligent respect for nature that should inform decisions on the uses of
technology” and without which, they say, “we are in danger of recklessly destroying
our life-support system”. I am even more encouraged by the radical scholarship of
Roth and Désautels (2002, 2004) and Roth and Barton (2004) – in particular, their
vision of science education as and for sociopolitical action. Not only is responsible
social action the motive for achieving scientific literacy it is also, sometimes, the
means of achieving it.
What are the essential elements of this kind of scientific literacy? Perhaps, and
somewhat paradoxically for an overall argument so politically distant from my own,
the answer can be found in the writing of Longbottom and Butler (1999):
Science education provides ideal opportunities to engage in a wide range of
careful investigations and problem-solving activities, where mistakes and
wishful thinking are readily exposed. Science education can value creativity but
not accept personal theories as an endpoint. The ability to adjudicate between
knowledge claims in ways independent of human desires is a special feature of
science that has allowed it to build up a public body of reliable knowledge.
Science education should convey these aspects of science. (p. 488)
In common with several others, these authors seem to be saying that scientific
literacy for active citizenship, responsible environmental behaviour and social
reconstruction lies more in learning about science and in doing science than it does in
learning science.9 No science curriculum can equip citizens with thorough first-hand
knowledge of all the science underlying all important issues. Moreover, much of the
scientific knowledge learned in school, especially in the rapidly expanding fields of
the biological sciences, will be out-of-date within a few years of leaving school.
However, science education can enable students to understand the significance of
knowledge presented by others and it can enable them to evaluate the validity and
reliability of that knowledge and to understand why scientists often disagree among
themselves on such major matters as climate change (and its causes) without taking it
as evidence of bias or incompetence. Of course, they also need to know that bias and
incompetence do sometimes occur. Thus, students need to have a clear understanding
of what counts as good science (i.e., a well designed inquiry and a well argued
conclusion) and be able to detect bias and self-interest. As Geddis (1991) comments
students need to be able to “uncover how particular knowledge claims may serve the

19

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
CHAPTER 1

interests of different claimants… they need to attempt to unravel the interplay of


interests that underlie these other points of view” (p. 171).
It is certainly not my intent to argue that knowledge of the major concepts, ideas
and theories of science is unimportant.10 It would be a very curious state of affairs
indeed to claim scientific literacy and admit to knowing no science at all! Never-
theless, the science content of scientific literacy is not my concern here. Of pertinence
here are those elements of the history, philosophy and sociology of science that would
enable students to leave school with robust knowledge about the nature of scientific
inquiry and theory building, an understanding of the role and status of scientific
knowledge, an ability to understand and use the language of science appropriately
and effectively, the capacity to analyze, synthesize and evaluate knowledge claims,
some insight into the sociocultural, economic and political factors that impact the
priorities and conduct of science, a developing capacity to deal with the moral-
ethical issues that attend some scientific and technological developments, and
some experience of conducting authentic scientific investigations for themselves
and by themselves (Hodson, 2006). While we cannot provide all the science knowl-
edge that our students will need in the future (indeed, we do not know what knowl-
edge they will need) and while much of the science they will need to know has yet to
be discovered,11 we do know what knowledge, skills and attitudes will be essential to
appraising and forming a personal opinion about the science and technology
dimensions of real world issues. If students acquire good learning habits and attitudes
towards science during the school years, it will be relatively easy for them to acquire
additional scientific knowledge later on, as and when the need arises, provided that
they have also acquired the language skills to access and evaluate relevant information
from diverse sources. Of course, to be scientifically literate in the sense I am arguing
for in this chapter, students will also need the language skills to express their
knowledge, views, opinions and values in a form appropriate to their purpose and the
audience, and to participate in public debate about SSI. Learning about science is
rather different. Gaining robust familiarity with key issues in the history, philosophy
and sociology of science requires lengthy and close contact with someone already
familiar with them – that is, a teacher or scientist who can provide appropriate
guidance, support, experience and criticism.
Of course, not everyone shares the view that understanding about the nature
of science is central to science education and the notion of critical scientific literacy.
Without much in the way of justification of their view, Wilson and Cowell (1982)
assert that “anyone who believed that what (say) Popper or Kuhn were concerned with
was central to education in science would, surely, either not know the kinds of issues
these philosophers were trying to tackle or not have a firm grasp on the idea of
education” (p. 39). I will leave readers to form their own views, content in the
persuasiveness of arguments presented in this chapter. Chapters 3 to 8 identify some
of the ideas in the vast literature of history of science, philosophy of science and
sociology of science that are of value in constructing a science curriculum capable of
attaining universal scientific literacy, in the sense defined here. First, though, it is
important to consider the understanding that many students currently hold about
science and scientists, the views promoted by their teachers and textbooks, and the
ways in which this research in this area is conducted.

20

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
IN PURSUIT OF SCIENTIFIC LITERACY

ENDNOTES
1
In the Nuffield approach, learning science and doing science were regarded as equivalent activities.
Indeed, teachers were encouraged to believe that “intellectual activity is the same whether it is at the
frontier or in a third grade classroom… The schoolboy learning physics is a physicist” (Bruner,
1960, p. 14, emphasis added). Similarly, students following the ChemStudy courses in the United
States were told that “they would see the nature of science by engaging in scientific activity, thereby
‘to some extent’ becoming scientists themselves” (Pimental (1960, p. 1 and Preface) – cited by
Jenkins, 1996b, p. 137). Interestingly, some advocates of constructivist pedagogy adopt an even
more ludicrous and educationally dangerous position when they declare that science is “an activity
that is carried out by all people as part of their everyday life” (Ministry of Education, New Zealand,
1993, p. 9).
2
Miller (1993) outlines three basic positions for analysing and describing curricula: transmission,
with its focus on traditional subjects taught largely through traditional didactic methods; transaction,
in which education is seen as a dialogue between student and curriculum, and through which the
student reconstructs knowledge; transformation, which is concerned primarily with individual and
social change.
3
DeBoer (2001) also cites the Rockefeller Brothers Fund (1958) report The Pursuit of Excellence
(p. 369) as a pioneer user of the term: “Just as we must insist that every scientist be broadly
educated, so we must see to it that every educated person be literate in science” (p. 586).
4
Millar (2006) describes some initial responses to Twenty First Century Science, a major curriculum
project in England. Aimed at 15 and 16 year olds, the course comprises two equal parts: a “core
science course” focused on science education for citizenship, and a more content-oriented
“additional science” course, which is offered with either a “pure” or “applied” emphasis. According
to Millar, teachers generally perceive the citizenship emphasis of the core science course as having a
marked beneficial impact on student interest and engagement, they report favourably on the
relevance of the course, incorporation of moral-ethical issues, emphasis on ICT and its use of case
studies and debate to promote critical thinking, though the latter activities and the language demands
of the curriculum resources are reported to have created major new demands on both students and
teachers.
5
In the United States, federal funding for research carried out by the National Science Foundation, the
National Institutes of Health, the Department of Defense, the Department of Agriculture and NASA
amounts to approximately $80 billion per annum.
6
Thankfully, 2007 saw a welcome but cautious dawning of awareness of these matters, prompted in
part by Al Gore’s film documentary An Inconvenient Truth and the publication of the 4th assessment
report of the Intergovernmental Panel on Climate Change (IPCC) under the title “Climate Change
2007”. Speaking of the earlier publication of the IPCC Assessment Report for 2007, Achim Steiner
(Executive Director of the UN Environment Programme) said: “February 2nd 2007 may be
remembered as the day the question mark was removed from whether people are to blame for
climate change” (reported by Adam, 2007, p. 7)
7
Nearly 50 years ago, and writing from a US perspective, LeCorbeiller (1959) noted that economic
power is underpinned by military power (which, of course, is maintained by advances in science and
technology) and that the United States (in common with many other nations) exports scientific,
engineering and technological ‘know-how’ abroad in order to spread American influence.
8
Nowhere is this more in evidence than in A Nation at Risk (National Commission on Excellence in
Education, NCEE, 1983): “our once unchallenged preeminence in commerce, industry, science, and
technological innovation is being overtaken by competitors throughout the world… If an unfriendly
power had attempted to impose on America the mediocre educational performance that exists today,
we might well have viewed it as an act of war. As it stands, we have allowed this to happen to
ourselves. We have even squandered the gains in achievement made in the wake of the Sputnik
challenge. Moreover, we have dismantled essential support systems which helped make those gains
possible. We have, in effect, been committing an act of unthinking, unilateral, educational
disarmament.” (p. 5)
9
In a number of publications (Hodson, 1992a, 1994, 1998a), I have argued that science education is
best regarded as comprising three major elements:

21

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
CHAPTER 1

– Learning Science – acquiring and developing conceptual and theoretical knowledge.


– Learning about science – developing an understanding of the nature and methods of science;
appreciation of its history and development; awareness of the complex interactions among
science, technology, society and environment; and sensitivity to the personal, social and ethical
implications of particular technologies.
– Doing science – engaging in and developing expertise in scientific inquiry and problem-solving;
developing confidence in tackling a wide range of “real world” tasks and problems.
In recent years (Hodson, 2003), I have added a fourth component: Engaging in sociopolitical
action – acquiring (through guided participation) the capacity and commitment to take appropriate,
responsible and effective action on science/technology-related matters of social, economic,
environmental and moral-ethical concern.
10
I emphatically reject the argument made by Shamos (1995) that, for the majority of students, science
content is only of value in its exemplification of the nature of science, though I do share his view
that technology is often a more accessible subject area for scientific literacy than is science itself.
11
The BSE episode provides a graphic illustration. As Solomon and Thomas (1999) remind us, the
nature of the prion agent that most scientists consider to be a cause of the disease was unknown to
biology when most of today’s adults received their science education.

22

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
CHAPTER 2

EXPLORING NATURE OF SCIENCE ISSUES


Students’ Views and Curriculum Images

In chapter 1, I argued that a significant element of scientific literacy for the 21st
Century comes under the umbrella of what I called learning about science – that is,
developing an understanding of the nature and methods of science, an appreciation
of its history and development, and an awareness of the often complex interactions
among science, technology, society and environment. Students can only be consi-
dered scientifically literate, I argued, if they possess a robust and authentic under-
standing of what science is, how science functions, what scientists do, and how
science develops and changes over time in response to sociocultural and economic
pressures. Indeed, I proffered the idea that in the complex 21st Century world in
which we live, this aspect of scientific literacy is at least as significant, if not more
significant, than acquisition of conceptual understanding.
It is reasonable to suppose that students’ views about science are the outcome of
two interacting influences:
− Curriculum experiences – what students encounter in school science lessons;
− Informal learning experiences – what they learn via the popular media (movies, TV
and radio, newspapers, Internet sites, advertizing) and from visits to museums,
zoos, aquaria, nature reserves, field centres, and the like.
This book is concerned with the learning about science content of curriculum
experiences: the kind of experiences we provide; the kind of experiences we should
provide, but don’t; the kind of experiences we do provide, but shouldn’t. Although
I am restricting the focus of the book in this way, I remain cognizant of the need
for teachers to address the other influences on students’ views about science and,
whenever necessary, to attempt to counter them.1 I intend to argue that curriculum
experiences are of two kinds: those that we explicitly plan and those that we do
not. There are many explicit messages about science in textbooks, especially in
those early chapters that tell students what science is about and what scientists do
when they are conducting investigations; there are lots of explicit references to the
nature of science and the history of science in STS-oriented materials; on occasions,
teachers take time and trouble to emphasize particular features of science and sci-
entific inquiry during laboratory activities or in class discussions. Just as frequently,
however, ‘messages’ about the nature of science and scientific practice are not con-
sciously planned by the teacher. Rather, they are implicit messages located in the
language we use, the kind of teaching and learning activities we employ (especially
in laboratory work), the examples of science and scientists we utilize, the illustrative
and biographical material in textbooks, and so on. What is at issue here is a very
powerful hidden or implicit curriculum. In the words of Cawthron and Rowell
(1978):
23

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
CHAPTER 2

… any science curriculum and its translation into practice embody images of
the nature of man as scientist, of scientific knowledge and of the relationships
between them. These are communicated, albeit implicitly and incidentally, just
as surely as the subject matter itself. (p. 31)
Because it says science on the school timetable, students regard what they experi-
ence during that lesson as science, not as learning science. Many assume that what
they do in science lessons, particularly during hands-on activities, is what scientists
themselves do as they conduct investigations. These experiences build, over time,
into a particular set of messages about science, scientists and the scientific enter-
prise. Everything that is part of the science lesson becomes an element in this
continuous ‘story building’ about science, whether it is explicitly planned by the
teacher or not. It follows that because many of the individual messages about science
are ‘transmitted’ implicitly, simply as a consequence of teachers’ day-to-day, short-
term decisions about the conduct of lessons, a major element of the overall story
about science is the teacher’s views about the nature of science. All teachers,
therefore, need to be cognizant of the responsibility they carry for helping to form
their students’ views about science. As a first step in treating that responsibility
seriously, teachers should attempt to ascertain, examine and critique their own views.
The traditional way of ascertaining views about the nature of science is by means
of questionnaire and survey instruments using multiple choice items or Likert scales.
A large number of such instruments, mainly for use with students but perfectly appli-
cable for research on teachers’ views as well, have been developed. In fact, even
25 years ago, a literature survey by Mayer and Richmond (1982) identified at
least 32 NOS-oriented instruments, among which the best known are the Test on
Understanding Science (TOUS) (Cooley & Klopfer, 1961), the Nature of Science
Scale (NOSS) (Kimball, 1967), the Nature of Science Test (NOST) (Billeh &
Hasan, 1975) and the Nature of Scientific Knowledge Scale (NSKS) (Rubba, 1976;
Rubba & Anderson, 1978), together with a modified version (M-NSKS) developed
by Meichtry (1992). Instruments dealing with the processes of science, such as the
Science Process Inventory (SPI) (Welch, 1969a), the Wisconsin Inventory of Science
Processes (WISP) (Welch, 1969b) and the Test of Integrated Process Skills (TIPS)
(Burns et al., 1985; Dillashaw & Okey, 1980) could also be regarded as providing
information on understanding of the nature of science. In general, these instru-
ments are constructed in accordance with a particular philosophical perspective
and are predicated on the assumption that all scientists behave in the same way.
Hence teacher and/or student responses that do not correspond to the model of
science assumed in the test are adjudged to be ‘incorrect’ (see Lucas (1975),
Koulaidis & Ogborn (1995), Alters (1997a) and Lederman et al. (2002) for an
extended discussion of this issue). Moreover, as later chapters will show, many
of these instruments pre-date significant work in the philosophy and sociology
of science, and so are of severely limited value in further studies. Like research
in science itself, research in science education is a product of its time and place.
More recent reviews by Lederman (1992, 2007), Lederman et al. (1998) and
Abd-El-Khalick & Lederman (2000) describe several additional NOS instruments
that take into account the work of more recent, and even contemporary, scholars in

24

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
EXPLORING NATURE OF SCIENCE ISSUES

philosophy and sociology of science. Most notable among these newer instruments
are Conceptions of Scientific Theories Test (COST) (Cotham & Smith, 1981),
Views on Science-Technology-Society (VOSTS) (Aikenhead et al., 1989), the Nature
of Science Survey (Lederman & O’Malley, 1990), the Views of Nature of Science
Questionnaire (VNOS) (Lederman et al., 2002) and several subsequent modify-
cations (see Lederman, 2007). A particularly useful tool for use with teachers in
both pre-service and in-service programs is the Nature of Science Profile developed
by Nott and Wellington (1993).2 Its value lies in its ease of administration, wide
scope and non-judgmental nature – key factors when dealing with student teachers
and their apprehensions about impending teaching practice placements. Of particular
value when student teachers have gained some classroom experience is Nott and
Wellington’s (1996, 1998, 2000) “Critical Incidents” approach. In group settings, or
in one-on-one interviews, teachers are invited to respond to descriptions of classroom
events, many related to hands-on work in the laboratory, by answering three ques-
tions: What would you do? What could you do? What should you do? Responses,
and the discussion that ensues, may indicate something about the teachers’ views
of science and scientific inquiry and, more importantly perhaps, how this under-
standing is deployed in class-room decision making. Similar approaches using
video and multimedia materials have been used by Hewitt et al. (2003) and Wong
et al. (2008a,b). An interesting variation adopted by Murcia and Schibeci (1999)
uses a science-oriented newspaper article as a stimulus to thinking about question-
naire items.
It almost goes without saying that teachers’ views about the importance of NOS
in the curriculum will also play a major part in their decisions about the extent to
which they will emphasize learning about science in their enacted curriculum and
the nature of the learning activities they provide (Bell et al., 2000; Schwartz &
Lederman, 2002). Interestingly, Nott and Wellington (1995) comment that “teachers’
knowledge of the nature of science may be as much formed by their teaching of
science as informing their teaching of science” (p. 865, emphases added) – a con-
clusion that has major implications for both pre-service and in-service teacher
education. There is also a substantial body of research to show that teachers’ nature
of science beliefs are sometimes over-ridden in the curriculum decision making pro-
cess by more immediate concerns with classroom management, the priority affor-
ded to concept acquisition and development, apprehension about student interest in
philosophical and sociological issues and the lack of good NOS teaching resources
(Mitman et al., 1987; Carey & Smith, 1993; Hodson, 1993a; Abd-El-Khalick et al.,
1998; Lederman, 1999; Smith et al., 2000; Lederman et al., 2001) – all of which have
major implications for teacher education and teacher professional development

STUDENTS’ VIEWS ABOUT SCIENCE AND SCIENTISTS

Three decades of research into students’ alternative frameworks of understanding


in science have shown us that students come to science lessons with some prior
understanding of many of the concepts and ideas included in the curriculum, and
that it would be a mistake to assume that these views are identical to scientists’

25

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
CHAPTER 2

views. Often they are significantly different. It would be a mistake, also, to assume
that these views can be easily displaced by the ‘correct’ or preferred views embedded
in the curriculum goals. Often they are very strongly held and resistant to change.
Necessarily, the understanding students have of any phenomenon under investigation
will profoundly influence the way in which they conduct that investigation, the sig-
nificance they attach to their findings and the conclusions they reach. It was this
realization that led to the development of constructivist pedagogy – an approach
that, in many parts of the world, has become the ‘new orthodoxy’ of science teach-
ing.3 Its essential steps are:
– Identify students’ ideas and views
– Create opportunities for students to explore their ideas and test their robustness
in explaining phenomena, accounting for events and making predictions
– Provide stimuli for students to develop, modify, and where necessary change
their ideas and views
– Support their attempts to re-think and reconstruct their ideas and views.
Similar strategies might apply to learning about science. At the outset, teachers
following such a strategy would ascertain what views about science and scientists
students already hold in order to encourage students to explore them, challenge them
and perhaps develop, augment or replace them. One could, of course, try to ascertain
students’ views about science by giving them a questionnaire to complete. As dis-
cussed earlier, with respect to teachers’ views about science, there are several such
instruments in existence, some of which provide much valuable information, others
of which do not. Despite some interesting recent developments in questionnaire
design, one problem remains unsolved: students do not always interpret question-
naire items in the way the designers intended or, as Lederman and O’Malley (1990)
put it, “language is often used differently by students and researchers” (p. 237).
The designers of VOSTS attempted to circumvent this problem by using multiple
choice items derived from student writing and interviews to provide a number of
different ‘position statements’ (sometimes up to 10 positions per item), including
“I don’t understand” and “I don’t know enough about this subject to make a choice”
(Aikenhead et al., 1987; Aikenhead & Ryan, 1992). It is the avoidance of the
forced choice and the wide range of aspects covered (definitions, influence of
society on science/technology, influence of science/technology on society, influence
of school science on society, characteristics of scientists, social construction of
scientific knowledge, social construction of technology, nature of scientific know-
ledge) that give the instrument such enormous research potential. Nevertheless, as
Abd-El-Khalick and BouJaoude (1997) point out, VOSTS was conceived and written
within a North American sociocultural context and, in consequence, may have
limited validity in non-Western contexts.4 In response to concerns like these, Tsai
and Liu (2005) have developed a survey instrument that is more sensitive to socio-
cultural influences on science and students’ views of science. Rooted in similar con-
cerns about the socioculturally-determined dimensions of NOS understanding is the
Thinking about Science instrument designed by Cobern and Loving (2002) as both
a pedagogical tool (for pre-service teacher education programs) and a research tool

26

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
EXPLORING NATURE OF SCIENCE ISSUES

for assessing views of science in relation to economics, the environment, religion,


aesthetics, race and gender.
Lederman and O’Malley (1990) utilized some of the design characteristics of
VOSTS to develop an instrument comprising just seven fairly open-ended items
(e.g. “Is there a difference between a scientific theory and a scientific law? Give an
example to illustrate your answer”), to be used in conjunction with follow-up inter-
views. Not all students were prompted to focus their answers on the tentativeness
of science, as the authors intended, and although the interview stage was able to go
some way towards re-focusing attention, these findings serve to reiterate the diffi-
culty of attempting to interpret students’ understanding from their written responses
to researcher-generated questions.5 In consequence, some researchers and teachers
believe that more useful information can be obtained, especially from younger stu-
dents, by relying solely on open-ended methods such as the Draw-a-Scientist Test
(DAST) (Chambers, 1983). In his initial study, Chambers used this test with 4807
primary (elementary) school children in Australia, Canada and the United States.
He identified seven common features in their drawings, in addition to the almost
universal representation of the scientist as a man: laboratory overall; spectacles
(glasses); facial hair; ‘symbols of research’ (specialized instruments and equip-
ment); ‘symbols of knowledge’ (books, filing cabinets, etc); technological products
(rockets, medicines, machines); and captions such as ‘Eureka’ (with its attendant
lighted bulb), E = mc2 and think bubbles saying “I’ve got it” or “A-ah! So that’s
how it is”.6 In many ways, little has changed since high school students in the
United States told Margaret Mead, almost fifty years ago, that a scientist is:
… a man who wears a white coat and works in a laboratory. He is elderly or
middle aged and wears glasses. He is small, sometimes small and stout, or
tall and thin. He may be bald. He may wear a beard, may be unshaven and
unkempt. He may be stooped and tired. He is surrounded by equipment: test
tubes, Bunsen burners, flasks and bottles, a jungle gym of blown glass tubes
and weird machines with dials. The sparkling white laboratory is full of
sounds: the bubbling of liquids in test tubes and flasks, the muttering voice of
the scientist… He spends his days doing experiments. He pours chemicals
from one test tube into another. He peers raptly through microscopes. He scans
the heavens through a telescope. He experiments with plants and animals,
cutting them apart, injecting serum into animals. He writes neatly in black
books. (Mead & Metraux, 1957, pp. 386 & 387)
In the twenty or so years since Chambers’ original work, students’ drawings have
changed very little (Ward, 1986; Fort & Varney, 1989; Symington & Spurling,
1990; Jackson, 1992; Newton & Newton, 1992, 1998; Matthews, 1996; Barman,
1999; Finson, 2002),7 with research indicating that the stereotype emerges round
about grade 2 and is well-established and held by the majority of students by grade
5. Matthews (1996) claims that the stereotype is beginning to be eroded, at least in
the United Kingdom, by the increased emphasis on science in the National Curri-
culum for primary schools, though this view is not shared by Newton and Newton
(1998). Not only are these images stable across genders, they seem to be relatively

27

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
CHAPTER 2

stable across cultural differences (Chambers, 1983; Parsons, 1997; She, 1998; Song
& Kim, 1999; Finson, 2002). There are some encouraging indications that students,
and especially male students in the age range 9–12, produce drawings with fewer
stereotypical features following the implementation of gender-inclusive curriculum
experiences (Mason et al., 1991; Huber & Burton, 1995).
It may be, however, that many researchers are seriously misled by the drawings
children produce. As Newton and Newton (1998) point out, “their drawings reflect
their stage of development and some attributes may have no particular significance
for a child but may be given undue significance by an adult interpreting them”
(p. 1138). Although very young children invariably draw scientists as bald men
with smiling faces, regardless of the specific context in which the scientist is placed;
it should not be assumed that children view scientists as especially likely to be bald
and contented. Many children seem unable or disinclined to draw a distinction bet-
ween science in the outside world and science lessons in school, such that their
drawings simply reproduce caricatures of school science activities (Brandes, 1996).
Perhaps somewhat older students are consciously and purposefully presenting the
stereotypical or ‘comic book’ cartoon image of the scientist rather than giving
insight into what they really believe. As Claxton (1990) reminds us, children com-
partmentalize their knowledge and so may have at least three different versions
of the scientist at their disposal: the everyday comic book version, the ‘official’,
approved version for use in school, their personal (and perhaps private) view. It is
not always clear which version DAST is accessing, nor how seriously the ‘artists’
took the task. There is also the possibility, with students in secondary school or
university, that the response is intended to make a sociopolitical point – for exam-
ple, that there are too few women or members of ethnic minority groups engaged in
science. One way to clarify matters is to talk to students about their drawings and
the thinking behind them, ask them if they know anyone who uses science in their
work (and what this entails) or present them with writing tasks based on scientists
and scientific discovery.8 Even very young children will provide detailed explana-
tions when given the opportunity to discuss their drawings and stories with the
teacher (Sharkawy, 2006). Interestingly, discussion seems to take a different course
when this task is set in science lessons than it does in other areas of the curriculum –
perhaps because of the influence of the ‘official’ or known-to-be-approved view
referred to earlier. It is also becoming increasingly clear that young children’s
responses to open-ended writing tasks involving science, scientists and engineers are
not stable and consistent: accounts and stories produced in science lessons are very
different from those produced in the language arts (Hodson, 1993a). Interestingly, it
seems that students responding to nature of science questions sometimes provide
significantly different oral and written responses (Roth & Roychoudhury, 1994).
Despite these caveats, there is still cause for concern about the views expressed
by students. There is particular cause for concern about the kind of responses stu-
dents make when they are asked to “imagine you are a scientist… write about
your work, about the discoveries you have made and the things you have invented”.9
Just over 50% of students report accidental discovery or what one might call the
‘Alexander Fleming syndrome’: materials are left lying around in the laboratory

28

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
EXPLORING NATURE OF SCIENCE ISSUES

and suddenly, one day, the ‘answer’ appears. There is no suggestion that significant
scientific results are a consequence of a carefully planned and systematic inquiry,
no suggestion of prior theorizing; instead, scientific breakthroughs just happen,
by chance. In addition, science is widely perceived by students to be a solitary
activity: the lone scientist labours long and hard to make discoveries of heroic
stature – a view that is reinforced by questionnaire studies of students’ views about
‘doing science’ or ‘being a scientist’. Furthermore, scientists are regarded by many
students as withdrawn and remote from real life, sometimes selfish and antisocial,
and certainly less friendly than ‘normal’ people. Recently, one of my students asked
me: “How can you tell when a scientist is an extrovert?” Answer: “He looks at
your shoes when he is talking to you”. Scientists have few interests, and certainly
not in music, movies or the arts in general. Most scientists don’t have a ‘normal’
family life: they may neglect their family (many don’t have a family) because of
their preoccupation with work. Indeed, they often go into the laboratory on their
day off. Often, they are seen as careless of their appearance, generally untidy and
disorganized; sometimes they even forget to eat because of their obsession with
work. This image is remarkably similar to that given by Fournier d’Albe (1923):
To the general public the man of science is a man of mystery, a man of inhu-
man and somewhat unaccountable tastes. Not everyone goes so far as to
maintain that he is a freak because he indulges in an activity “with no money
in it”. But it seems to be generally agreed that the “scientist” is a being living
outside ordinary human spheres, not amenable to ordinary human standards,
a being who is usually harmless but may conceivably become dangerous.
(cited in Stuewer, 1998, p. 25, emphasis added)
Carl Sagan (1995) describes the common stereotypical image of the scientist in
even less complimentary terms: “Scientists are nerds, socially inept, working on
incomprehensible subjects that no normal person would find in any way interesting
– even if he were willing to invest the time required, which, again, no sensible per-
son would. ‘Get a life’, you might wish to tell them” (p. 362).
And what do students believe that scientists do when they go into the lab? They
do experiments; they find out about things; they find out new information and new
facts; they find out how things are. Scientists do lots of “finding out” and “working
things out”; they also “invent things” and “discover new stuff ”. In a study by
Coleman (1998), a number of students regard an explanation as scientific if it
includes information that not everyone knows or is immediately obvious, or infor-
mation that has to be discovered rather than looked up in a book or found from
an Internet search. For many students, it is also the case that scientists use “big
words that only they can understand”. Predictably, perhaps, many students state that
experiments are conducted to prove that such and such is the case – a view that is
regularly reinforced by the television or newspaper advertisers’ insistence that
product X is proven by experiment to do the job better than any similar product.
The notion that well conducted experiments reveal the truth is very common. Inter-
estingly, some students take the radically different view that experiments are entirely

29

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
CHAPTER 2

open-ended, a kind of “shot-in-the-dark” during which “anything can happen” – for


example, “they give stuff to rats to see what happens”.
Using interview methods and writing tasks, Joan Solomon and her colleagues
have researched students’ understanding of the nature of theory and its relationship
to experiments. Ideas about theories appear to fall into three major categories: (a) a
theory is a hunch or guess about what will happen, (b) a theory is an explanation
for why things happen (close to the view we might wish students to hold), (c) a
theory is a fact. In these studies, students sometimes referred to “proper theories”
as those theories that have been shown by tests and experiments to be true (Duveen
et al., 1993; Solomon et al., 1994, 1996). Interestingly, some students in my Toronto-
based study took the view that theories are just guesses or speculations. Comments
such as “that’s just a theory… it isn’t true” and “It’s something that you think” were
widespread, in some sense reflecting the everyday expression that “it may be OK in
theory but it doesn’t work in practice”. Intriguingly, some students in the Solomon
et al (1994) study seemed to see experiment as a ‘last resort’ when no-one knows
what to do – as in, “We’ll just have to experiment” (p. 365). Among older students
in my Toronto-based study there was sometimes a distinction drawn between
“science as it ought to be” (the kind of science that is described in text-books)
and “science as it is”. Students were not expressing the view that scientists behave
differently when self interest, commercial interests or military interests intervene
(however desirable this awareness might be); rather, they were suggesting that the
ideals of scientific inquiry are so demanding that they are beyond the reach of most
‘mere mortals’.
Interviews with students undertaking research projects as part of their final year
programme at the University of Leeds revealed that undergraduates resemble sec-
ondary school students in regarding scientific knowledge as ‘provable beyond
doubt’ on the basis of experimentally acquired data (Ryder et al., 1999). In most
cases, scientific inquiry was seen in terms of individual scientists seeking reliable
data (usually via experiment) on which to base their conclusions. Encouragingly,
their appreciation of the role of theory in directing the nature of scientific inquiry
became markedly more sophisticated during the course of the project work, though
they remained relatively unaware of the internal and external social dimensions of
scientific practice.

CURRICULUM IMAGES OF SCIENCE AND SCIENTISTS

Often, the confused and confusing views of science held by students are com-
pounded by conventional science education. There are particularly powerful mes-
sages about science embedded in laboratory activities conducted in class. As
subsequent chapters will make clear, these messages too often convey distorted or
over-simplified views of the nature of scientific investigations, especially with
respect to the role of theory. These “folk theories” of science, as Windschitl (2004)
calls them, are also held by teachers (as a consequence of their own science educa-
tion) and have substantial influence on their day-to-day curriculum decision mak-
ing, thus reinforcing similar messages embedded in school science textbooks and

30

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
EXPLORING NATURE OF SCIENCE ISSUES

other curriculum materials. Such messages are not restricted to comments on


‘scientific method’. More than a quarter century ago, Smolicz and Nunan (1975)
identified four “ideological pivots” inherent in the image of science presented
through the science curriculum. First anthropocentrism: the view of mankind as the
technologically powerful manipulator and controller of nature, with science as the
means by which we control the environment and shape it to meet our interests and
needs.10 Second, quantification: scientists are regarded not just as observers but as
measurers and quantifiers. Whatever exists in nature can (and should) be explained
in mathematical terms, best of all by means of equations. Hence, Lord Kelvin’s
dictum that we do not understand a thing until we can measure it. Third, positivistic
faith: faith in the inevitable linear progress of science towards truth about the
world, with the certainty of this knowledge being underpinned by the all-powerful
scientific method. Fourth, the analytical ideal: the assumption that phenomena and
events are best studied and explained via analysis; an entirely mechanistic view of
the world which assumes that the whole is simply, and no more than, the sum of its
parts. Two major questions spring to mind. First, are these “ideological pivots”
promoted through science education? In other words, are Smolicz and Nunan (1975)
correct in their analysis? Second, are these the underlying values of science? In
other words, is this a faithful representation of science and, therefore, an appropri-
ate set of values to promote?
A decade later, as part of a major survey of Canadian science education con-
ducted by the Science Council of Canada, Nadeau and Désautels (1984) identified
what they called five mythical values stances suffusing science education:
– Naïve realism – science gives access to truth about the universe.
– Blissful empiricism – science is the meticulous, orderly and exhaustive gathering
of data.
– Credulous experimentation – experiments can conclusively verify hypotheses.
– Excessive rationalism – science proceeds solely by logic and rational appraisal.
– Blind idealism – scientists are completely disinterested, objective beings.
The cumulative message is that science has an all-purpose, straightforward and
reliable method of ascertaining the truth about the universe, with the certainty of
scientific knowledge being located in objective observation, extensive data collec-
tion and experimental verification. Moreover, scientists are rational, logical, open-
minded and intellectually honest people who are required, by their commitment to
the scientific enterprise, to adopt a disinterested, value-free and analytical stance.
In Cawthron and Rowell’s (1978) words, the scientist is regarded by the science
curriculum as “a depersonalized and idealized seeker after truth, painstakingly
pushing back the curtains which obscure objective reality, and abstracting order
from the flux, an order which is directly revealable to him through a distinctive sci-
entific method” (p. 32). In quite startling contrast, Siegel (1991) states that:
Contemporary research… has revealed a more accurate picture of the scientist
as one who is driven by prior convictions and commitments; who is guided by
group loyalties and sometimes petty personal squabbles; who is frequently
quite unable to recognize evidence for what it is; and whose personal career

31

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
CHAPTER 2

motivations give the lie to the idea that the scientist yearns only or even mainly
for the truth. (p. 45)
A number of questions spring to mind.
1. Does the curriculum project the images identified here, and questioned by Siegel’s
remarks?
2. Are these images faithful to the nature of real science and the characteristics and
behaviour of real scientists engaged in ‘doing science’? In other words, do these
descriptions provide an authentic view of science?
3. Do students internalize these views?
4. Does it matter what image we project or what image of science students acquire?
At the time Nadeau and Désautels were writing, the answer to the first question
was undoubtedly “Yes”. While much has changed in the intervening years, thanks
largely to the STS thrust in curriculum debate, many school science curricula
and school textbooks continue to project these images (Lakin & Wellington,
1994; Cross, 1995; Knain, 2001; Clough, 2006). Loving (1997) laments that all too
often –
(a) science is taught totally ignoring what it took to get to the explanations we
are learning – often with lectures, reading text, and memorizing for a test. In
other words, it is taught free of history, free of philosophy, and in its final
form. (b) Science is taught as having one method that all scientists follow step-
by-step. (c) Science is taught as if explanations are the truth – with little
equivocation. (d) Laboratory experiences are designed as recipes with one right
answer. Finally, (e) scientists are portrayed as somehow free from human
foibles, humor, or any interests other than their work.” (p. 443)
It would be a fairly simple task to locate passages in school science textbooks that
continue to promote discredited views about nature of science issues. I have resisted
the temptation because I regard it as much more important for readers to look care-
fully and critically at the science textbooks in use in their own school or recom-
mended by the local Ministry of Education, School Board or Local Education
Authority, especially in light of the discussion of issues in the following several
chapters. Indeed, this is a task I set for all my graduate students at the outset of my
course dealing with nature of science and science education. As an aside, it is per-
haps worth mentioning that my own research shows that teachers frequently change
the model of science they project through the curriculum in accordance with the
particular topic being studied (in particular, its conceptual difficulty) and in relation
to the perceived ability level of the students (Hodson, 1993a), sometimes in disre-
gard of the views promoted by the textbook.
My answer to question 2 (with respect to the Cawthron and Rowell image) is an
unequivocal “No”, as subsequent chapters in this book will demonstrate. The ques-
tion of whether Siegel’s alternative image of science is any more authentic than
Cawthron and Rowell’s will be considered in chapter 7. Of course, expressing dis-
satisfaction with current messages about science is all very well, but teachers and
curriculum developers need access to an alternative set of messages. This book will
attempt to provide such alternatives. What will become apparent in subsequent dis-

32

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
EXPLORING NATURE OF SCIENCE ISSUES

cussion is that while there is no universally accepted view of the nature of science,
especially in regard to the sociocultural dimensions of the scientific endeavour,
there is a measure of agreement about a number of key issues relating to the con-
duct of scientific inquiry and the nature of scientific observation (McComas et al.,
1998). Chapters 3 to 8 explore some of the literature in the history, philosophy and
sociology of science that I subsequently use in chapter 9 to underpin an alternative
view of science, scientists and scientific practice suitable for the school science
curriculum.
Given the earlier description of students’ images of scientists and understanding
of experiments, the realistic answer to question 3 is likely to be “Yes”. Although not
all students will internalize all the misunderstandings of science embedded in less
enlightened science curricula, there is ample evidence that many students do leave
school with a confused, confusing, deficient or distorted view of the nature of sci-
ence and the activities of practising scientists (Ryan, 1987; Carey et al., 1989;
Larochelle & Désautels, 1991; Lederman, 1992; Duveen et al., 1993; Abell & Smith,
1994; Solomon et al., 1994, 1996; Griffiths & Barman, 1995; Driver et al., 1996;
Lubben & Millar, 1996; Barman, 1997; Leach et al., 1997; Hogan & Maglienti,
2001; Moss et al., 2001). Of course, there are encouraging signs that some teach-
ers, in some schools, are able to ensure that students develop a more authentic view of
scientific practice, sometimes as a consequence of an explicit program of study in the
history, philosophy and sociology of science, sometimes as a consequence of practi-
cally oriented experiences in laboratories, in the field or in zoos and museums.
This leaves question 4: Does it matter what image of science is presented and
assimilated? It matters insofar as it influences career choice, and so may have long
term consequences for individuals. It matters if the curriculum image of science is
such that it dissuades creative, non-conformist and politically conscious individuals
from choosing to pursue science at an advanced level. It matters if the image of
science is such that it dissuades women, members of visible minority groups and
students from lower socioeconomic status homes from entering science-related
careers or seeking access to higher education in science and engineering because they
don’t see themselves included and represented in the science curriculum. It matters
if our politicians, public servants and industrialists are so ignorant of scientific and
technological issues that their decision-making is ill-informed and uncritical. It
matters if the general population is unable to respond knowledgeably and critically
to the claims and proposals of those in society who might use scientific arguments
(and sometimes pseudoscientific or scientifically spurious arguments) to persuade,
manipulate and control. It matters simply because a significant part of humankind’s
cultural achievement is so poorly understood. To echo and adapt some of the dis-
cussion in chapter 1 for the desirability of scientific literacy itself, arguments for
equipping students with a more authentic view of science can be made on intrinsic,
utilitarian and citizenship grounds. There is ample evidence, for example, that the
unfavourable image of science and scientists to which many students are exposed
is one of the major reasons why many students turn away from science at an early
age (Holton, 1992; Wang, 1995; Gardner, 1998). Thus, it prematurely limits the pool
of talent from which future scientists are drawn, with potentially damaging effect

33

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
CHAPTER 2

on society’s economic and cultural well-being. Moreover, failing to provide every


student with an adequate understanding of the nature of science runs counter to
the demand for an educated citizenry capable of responsible and active partici-
pation in a democratic society. As I argued in chapter 1, a proper understanding of
science and the scientific enterprise is just as essential as scientific knowledge (i.e.,
conceptual understanding) in ensuring and maintaining a socially-just democratic
society.
[Scientific literacy] should help students to develop the understandings and
habits of mind they need to become compassionate human beings able to
think for themselves and to face life head on. It should equip them also to
participate thoughtfully with fellow citizens in building and protecting a soci-
ety that is open, decent, and vital. (AAAS, 1989, p. xiii)

REITERATING THE POLITICAL ARGUMENT FOR NOS UNDERSTANDING

Many of us have realized to our cost that a little knowledge can be dangerous thing.
Those with little knowledge of science, especially with little knowledge of the nature
of science, can be led to accept as dogma almost any knowledge that they don’t
fully understand, led to accept way too much on faith and on trust, led to believe
that science has all the answers to all of our problems. Central to this disturbing
situation, of course, is uncritical acceptance of the myth of the all-powerful route to
certain knowledge via the scientific method. While a lack of public understanding
of science is clearly an obstacle to proper democracy, so also is an understanding
of science rooted in myths and falsehoods about scientific method. The underlying
values and ideological pivots described by Nadeau and Désautels (1984) and
Smolicz and Nunan (1975) do not equip students with the critical skills necessary
to challenge science and scientists on matters of sociopolitical, economic and envi-
ronmental significance. Worryingly, through widespread acceptance of the myth of
scientific method, scientists can sometimes achieve a level of acclaim that leads the
public to seek their advice on matters outside the sphere of science. Nor are scien-
tists themselves always immune to the false logic that high scientific achievement
necessarily equates to wisdom on all other matters. As Bauer (1992, p. 40) remarks:
“Instances are common enough in which successful scientists succumb to the temp-
tation to see themselves as authorities not only in their own tiny field but over sci-
ence as a whole and even beyond that”. Conversely, belief in the certainty of
knowledge produced by the scientific method and the inevitability of successful
outcomes to research can lead to unrealistic expectations of science and impatience
when scientists do not immediately ‘deliver’ on society’s wants and needs.
Science in textbooks, the only science that some people know, is very different
from ‘real’ science, from science at the theoretical cutting edge or science at the
frontier of research and development, and from science that informs (or should
inform) key decision-making on matters of public interest. Quite rightly, only
knowledge that has stood the test of time as worthwhile knowledge is incorpora-
ted into textbooks for the primary and secondary school levels, and because it has
proved its value over many years, this knowledge is retained and reinforced. In

34

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
EXPLORING NATURE OF SCIENCE ISSUES

consequence, it gives the appearance of being objective and true. While scientific
knowledge in undergraduate texts, reviews and monographs is closer to ‘real’ con-
temporary scientific knowledge than the science in school textbooks, only the
research literature reveals the true nature of scientific knowledge, with its character-
istic subjectivity, inconsistency, controversy and uncertainty. This is not so much
an argument for confronting students in school with the primary research literature,
though I do see value in providing some guided access to some of that literature, as
it is an argument for assisting students to understand the status of scientific knowl-
edge, the ways in which it is generated, communicated and scrutinized by the
community of scientists, and the extent to which it can be relied upon to inform
critical decisions.
A literate citizen should be able to evaluate the quality of scientific informa-
tion on the basis of its source and the methods used to generate it. Scientific
literacy also implies the capacity to pose and evaluate arguments based
on evidence and to apply conclusions from such arguments appopriately
(National Research Council, 1996, p. 22).
These are the elements of scientific literacy to be addressed in the following chap-
ters. According to Kolsto (2001), the aspects of scientific literacy essential for
dealing adequately with controversial socioscientific issues are: understanding of
science as a social process, recognition of the limitations of science, familiarity
with the values of science and the cultivation of a critical attitude. I will argue
throughout the course of this book that the last of these elements necessarily pre-
supposes the others and constitutes the principal goal of science education.
However, there are those, such as Michael F.D. Young (1976), who claim that the
real goal of science education is scientific illiteracy and the cultivation of an
uncritical attitude. These goals are achieved, Young argues, by making science in
the early years of schooling very abstract, boring and difficult, thus ensuring that
most students will fail to learn satisfactorily and will leave school afraid of science,
unable to read scientific material in a critical way, and easily intimidated by those
who are adept at using scientific language. Science can then be used as a tool of
policy. When the language of science and technology is used in advertizing, political
addresses and official communications of any kind these scientific illiterates are
unable to ‘read between the lines’, to question or to oppose what is being presented
to them. They have little choice but to accept the word and the decisions of
‘experts’ on seal culling, forestry clear cutting, genetically modified food, BSE,
health risks from cellphone use, or any of the other 1001 issues that confront citi-
zens in their everyday lives. They have, in effect, been disenfranchised.
(They) see themselves as dependent on experts in more and more aspects
of their lives… Except in the specific context of their work, and possibly in
leisure pursuits such as car maintenance, our increasingly technologically
dominated world remains for the majority as much a mystery as the theologi-
cal mysteries of feudal times. (Young, 1976, pp. 51 & 53)

35

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
CHAPTER 2

Arguing along similar lines, Neil Postman (1992) described American society as
a “technopoly” in which citizens are socialized into accepting without question any
statement by a supposed scientific ‘expert’ if it is presented in a way that readers or
listeners perceive to be ‘scientific’ and is claimed to derive from a research study
conducted at a reputable university (no matter whether that claim is true or false).
The world we live in is very nearly incomprehensible to most of us. There is
almost no fact, whether actual or imagined, that will surprise us for very long,
since we have no comprehensive and consistent picture of the world that
would make the fact appear as an unacceptable contradiction. We believe be-
cause there is no reason not to believe. (p. 58).
More recently, Bencze (2001) has argued that contemporary science education
for the majority of students is, in effect, an apprenticeship for consumership – that
is, it seeks to create “a large mass of relatively scientifically and technologically
illiterate citizens who simultaneously serve as loyal workers and voracious, un-
questioning consumers” (p. 350). There are strong echoes here of Michael Apple’s
(1993) assertion that in the new economy-driven educational climate, students are
no longer seen as people who will participate in the struggle to build and rebuild
the social, educational, political and economic future, but as consumers: freedom is
“no longer defined as participating in building the common good, but as living in
an unfettered commercial market, with the education system… integrated into the
mechanisms of such a market” (p. 116). In similar vein, Lee and Roth (2002) assert
that in contemporary science education “both the subject matter and the method of
instruction are not geared toward generating a scientifically literate populace, but
rather function like a Fordian production line in a Foucauldian (disciplining) insti-
tution that forms employees of a certain class for a limited number of powerful
institutions” (p. 42). Marshall (1995) uses the term “busnocratic rationality” to
describe the curriculum emphasis on acquisition of skills rather than knowledge
building, and on information and information retrieval rather than knowledge, under-
standing and wisdom, allied with the notion that it should be consumers of education
(i.e., business interests and employers) rather than the clients (students) or providers
(educators) who determine curriculum, define and measure quality in education,
and set standards of attainment. Moreover, he argues, these standards are set in such
a way that they ensure a largely uneducated, uncritical and undemanding work-
force to fill society’s low-paid jobs, with any shortfall being filled by immigration
(see also Larner (2000) and Stromquist & Monkman (2000)).
Apple (2001a) sees a powerful alliance among enthusiasts for the neo-liberal
marketization of education, neo-conservatives who want “a return to higher stan-
dards and a ‘common culture’, authoritarian populist religious fundamentalists
deeply worried about secularity and the preservation of their own traditions, and
particular factions of the professionally oriented new middle class who are commit-
ted to the ideology and techniques of accountability, measurement and ‘manage-
ment’” (p. 103). Carter (2005) argues along similar lines:
Neoliberalism ‘marketizes’ everything, even notions of subjectivity, desire,
success, democracy, and citizenship, in economic terms at the same time neo-

36

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
EXPLORING NATURE OF SCIENCE ISSUES

conservatism works to preserve traditional forms of privilege and marginalize


authentic democratic and social justice agendas. More sinister still is the suc-
cess with which both ideologies have colonized the rhetoric so at the very
time reforms appear to be more just and equitable, they actually work in
opaque ways against those they purport to help… Neoliberal and neoconser-
vative forces work in tandem to marketize and reform and, as reform pro-
ceeds, to (re)distribute power back to traditional elites, effectively rejecting
recent progressive liberal moves to increase equality and social redress…
Democracy [has been] redefined] as largely synonymous with capitalism,
so that consumption becomes the new form of democratic participation, and
equity becomes isomorphic with increased choice. (p. 571)
In the so-called ‘knowledge economy’, businesses need only a relatively small
number of people who are knowledge creators and managers, whom Apple (2001b)
calls “symbolic analyzers” (those who can analyze and manipulate words, numbers,
visual representations and other symbols) and a much larger number of less skilled
and less knowledgeable workers who can and will follow instructions (Gee et al.,
1996; Lankshear, 2000; Bencze & Alsop, 2007). Thus school science education
functions to select and educate the “relatively small group of students who may
work as engineers and scientists to help companies develop and manage mecha-
nisms of production (and consumption) of goods and services… (and) large groups
of citizens who may function best as compliant workers and as enthusiastic pur-
chasers of products and services of business and industry” (Bencze, 2004,
p. 193). For the majority, emphasis is not on the development of critical thinking
but on the mastery of a given body of knowledge. Assessment and evaluation are
seen as a means of monitoring the system to determine its efficiency. The result is
an education that focuses on that which is easily and reliably assessed. As Noble
(1998) comments, business benefits from “a school system that will utilize sophis-
ticated performance measures and standards to sort students and to provide a rela-
tively reliable supply of… adaptable, flexible, loyal, mindful, expendable, ‘trainable’
workers” (p. 281). Hence the rush in many countries around the world to establish
‘standards of performance’ monitored by an imposed regime of systematic and
regular assessment via standardized tests. Education authorities insist that students,
classes, schools and whole education systems show quantifiable results, with test-
ing regimes monitoring outcomes and positioning everyone so that improvements
can be claimed by the authorities and shortfalls or deficiencies blamed on teachers
(Carter, 2005). Apple (1999, 2000, 2001a) argues that these educational standards
embody both neoliberal concerns for increased accountability, surveillance and
regulation and neoconservative desires for a return to ‘real learning’ and ‘real know-
ledge’. In this kind of technocratic approach to education, efficiency, marketability
and accountability are regarded as the ultimate virtues.
While not everyone would subscribe to the view that there is an intention to
ensure that the vast majority of the population remains scientifically illiterate as a
means to control, manipulate and oppress them, many would recognize that this is
the outcome of much contemporary science education. Scientific illiteracy is, for
many students, the consequence of what and how we teach science and technology

37

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
CHAPTER 2

in school, and makes possible the mass manipulation of scientifically illiterate peo-
ple via the popular media. Citizens who do not understand how scientific research
is done and how scientific research is scrutinized for validity and reliability, have
little option but to accept the recommendations of those they perceive to be ex-
perts, or are persuaded to accept as such. Hence my argument is that it does matter
what we teach students about science. It does matter that we teach them to under-
stand the nature of science and the scientific enterprise at a critical level. It does
matter that ordinary citizens have the capacity to read scientific text in journals
such as Scientific American, New Scientist, Discovery and Science Digest with un-
derstanding, and can make sense of scientific arguments wherever and whenever
they encounter them. It does matter that citizens have some understanding of the
ontological status of scientific knowledge and the ways in which scientific knowl-
edge is generated and validated by the community of practitioners. It does matter
that they have some sense of the historical development and cultural context of sci-
ence. Above all, it matters that future citizens are able to judge the validity of a
knowledge claim independently of other people, that they can tell the difference
between good science and bad science, and between science and non-science, and
can recognize fraudulent science and unwarranted claims when they encounter
them. It matters for social reasons; it matters for political reasons; it matters for
economic reasons; it matters, perhaps most of all, for environmental reasons. The
planet can no longer accommodate a scientifically and technologically illiterate,
uncritical, yet technologically powerful species. In the words of Carl Sagan (1995):
The consequences of scientific illiteracy are far more dangerous in our time
than in any that has come before. It’s perilous and foolhardy for the average
citizen to remain ignorant about global warming, say, or ozone depletion, air
pollution, toxic and radioactive wastes, acid rain, topsoil erosion, tropical
deforestation, exponential population growth. (p. 6)
Beyer (1998) paints a particularly bleak picture of contemporary society when
he says that we live “in a democratic-capitalist social order in which commodity
fetishism, the rule of the market, patriarchy, and White Supremacy constrain, distort,
and oppress the expression of many individuals’ humanity and their ability to act
democratically” (p. 260). Dobbin (1998) is equally dark in his vision: “Thousands of
years of human development and progress are reduced to the pursuit of ‘effi-
ciency’, our collective will is declared meaningless compared to the values of the
marketplace, and communitarian values are rejected in favour of the survival of the
fittest. A thinly disguised barbarism now passes for, is in fact promoted as, a global
human objective” (p. 1). I don’t believe that the world is quite so grim as these
authors contend, though I admit that there is no more room for complacency on the
sociopolitical front than on the environmental front. It is perhaps an opportune time,
then, lest this nightmare vision comes to pass, to restate the argument for enhanced
scientific literacy (that is both universal and critical) in terms of the politicization
of students, thus equipping them to resist technological determinism and the cul-
ture of consumerism and compliance, to fight for social justice, and to conduct
their lives in an environmentally responsible way.

38

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
EXPLORING NATURE OF SCIENCE ISSUES

To say that we are living in an era of rapid and far-reaching change, the out-
comes of which are beyond prediction, is not to say anything new or particularly
startling. But it is something to which educators, and especially science educators,
need to respond. These major social, economic and political changes, many occur-
ring on a global scale, are coincident with equally profound changes in the genera-
tion, organization and transmission of knowledge and information. Previous barriers
of time and space have been largely overcome. This instant interconnectivity has
intensified all aspects of human life, requiring that we respond to changes and pro-
posals for change within a very short period of time. Moreover, we live in an era that
generates increasing numbers of moral-ethical dilemmas but offers fewer moral
certainties. My argument thus far is that universal critical scientific literacy, inter-
preted in this book in terms of learning about science, is one of the educational im-
peratives in helping students to cope with life in this constantly changing and
uncertain world. It is a way of diverting us from the nightmare world of widespread
scientific illiteracy that Carl Sagan (1995) so gloomily speculated on in his book
The Demon-Haunted World.
I have a foreboding of… when awesome technological powers are in the hands
of the very few, and no one representing the public interest can even grasp
the issues; when the people have lost the ability to set their own agendas or
knowledgeably question those in authority; when, clutching our crystals and
nervously consulting our horoscopes, our critical faculties in decline, unable to
distinguish between what feels good and what’s true, we slide, almost without
noticing, back into superstition and darkness. (p. 25)
The curriculum I have in mind is aimed at far-reaching social change through
critical consideration of socioscientific issues (Hodson, 2003). In a number of
respects, it overlaps with futures studies (Lloyd & Wallace, 2004) – particularly
in respect of the guiding principles of futures education set out by Cornish (1977,
p. 223):
– The future is not fixed, but consists of a variety of alternatives among which we
can choose.
– Choice is necessary. Refusing to choose is itself a choice.
– Small changes through time can become major changes.
– The future world is likely to be different in many respects from the present
world.
– People are responsible for their future; the future doesn’t just happen to them.
– Methods successful in the past may not necessarily work in the future, due to
changed circumstances.
As a first step in building a future-oriented curriculum, it is necessary to con-
sider the extensive literature in the philosophy of science, the sociology of science
and the history of science, from which an appropriate selection of key issues can be
identified.

39

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
CHAPTER 2

ENDNOTES
1
To intervene effectively, of course, teachers need to be aware of these ‘popular images’ of science.
Sadly, constraints on space preclude a detailed consideration of key aspects of the ‘popular images’
of science and scientists here. Dunwoody (1993), Gough (1993), Wellington (1993), McSharry &
Jones (2002), Dimopoulos & Koulaidis (2003) and Shibley (2003) provide a useful starting point for
this exploration.
2
There may be some value in readers who are so inclined completing this profile now and then doing
so again after reading through this book. This is the kind of exercise I encourage my graduate stu-
dents to undertake. It is interesting that some students change their views substantially in response to
the history and philosophy of science issues we discuss, while others are confirmed in the views they
held at the outset.
3
This is not the appropriate place to discuss the nature of constructivist pedagogy, its strengths and
weaknesses as an approach to successful learning, or its problematic epistemology. There is already
an extensive literature dealing with these matters (see, for example, Faire & Cosgrove, 1988; Saunders,
1992; Appleton, 1993; Bell, 1993; Matthews, 1993, 1997, 1998a; Driver et al., 1994; Fensham
et al., 1994; Solomon, 1994; Phillips, 1995; Osborne, 1996; Nola, 1997; Tyson et al., 1997; Hewson
et al., 1998; Turner & Sullenger, 1999; Jenkins, 2000; Tobin, 2000; Gil-Perez et al., 2002). My own
views are elaborated in Hodson (1998a) and Hodson and Hodson (1998a,b).
4
VOSTS does appear to be robust enough to function reliably in a British context (Botton & Brown,
1998).
5
This instrument has subsequently been modified to produce the Views of Nature of Science ques-
tionnaire (Lederman et al., 2002), and further modified by Lederman and his co-workers (Lederman,
2007) and by Wong et al. (2008a,b).
6
Interestingly, Chambers found that scientists themselves also tend to draw these stereotyped pic-
tures.
7
Finson et al. (1995) have developed a checklist for identifying and quantifying the components of
students’ drawings for more efficient data analysis.
8
Miller (1992, 1993) advocates the following approach: “Please tell me, in your own words, what
does it mean to study something scientifically?”
9
The views of science and scientists presented here were collected from students aged 11-17 in a
number of schools in the greater Toronto area over the 5-year period 1999 to 2004.
10
I am consciously using mankind rather than humankind to lend increased force to Smolicz and
Nunan’s proposition. They used the term man, but at a time when authors were less sensitive to gen-
dered language. Of particular interest in this ‘nature in the service of man’ view is that we have the
right to control and manipulate the natural environment.

40

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
CHAPTER 3

THE TRADITIONAL VIEW OF SCIENCE


Recognizing the Myths

I stated in chapter 2 that school science education continues to promote and


perpetuate some grossly distorted views of science and scientists, many of which are
neatly encapsulated in Siegel’s (1991) description of the traditional image of science:
Science has traditionally been seen as the apex of rationality. The scientist, as
the traditional image has it, is the dispassionate seeker of the truth – the
person in the lab coat, untroubled by passion or emotion, unbiassed by prior
conviction, guided only by reason, patiently observing, experimenting, follow-
ing the evidence wherever it leads. In this image, the scientist believes and acts
entirely on the basis of evidence and reason. What better personification of
rationality could there be? (p. 9)
This chapter deals with three of the more persistent myths and falsehoods promoted
through the science curriculum:
– Observation provides a secure base of facts from which knowledge can be derived.
– Science starts with observation.
– Science proceeds by induction – from single observations (or a series of them)
to generalizations, laws and theories, which can be confirmed through further
observations.

MYTH 1: OBSERVATION PROVIDES DIRECT AND RELIABLE ACCESS


TO SECURE KNOWLEDGE OF THE WORLD

A significant part of the model of science projected by many school science curri-
cula is the assertion that the validity and reliability of observations are independent
of the opinions and expectations of the observer, and can be confirmed by direct
use of the senses by other observers. In other words, there is an assumption that
when we observe a phenomenon or event we all see the same thing. This assump-
tion that human observers have direct access to the properties of the external world
implies that nothing enters the mind except by way of the senses and that the mind
is a tabula rasa on which our senses inscribe a true and faithful record of the world.
This is quite simply not true, as Virginia Woolf so eloquently describes in her 1925
essay on Modern Fiction.
Examine for a moment an ordinary mind on an ordinary day. The mind receives
a myriad impressions – trivial, fantastic, evanescent, or engraved with the sharp-
ness of steel. From all sides they come, an incessant shower of innumerable
atoms; and as they fall, as they shape themselves into the life of Monday or
Tuesday, the accent falls differently from of old. (Woolf, 1925, p. 149)
41

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
CHAPTER 3

We see things differently because we are now different, as a consequence of


our experiences. Our minds are not blank slates. Rather, we interpret the sense
data that enters our consciousness in terms of previous experiences (and the sense
we made and continue to make of them), prior knowledge, beliefs and expecta-
tions. As Friedrich Nietzsche (1906/1968) said, “Everything of which we become
conscious is arranged, simplified, schematized, interpreted through and through”
(para. 477, p. 463). Immanuel Kant (1929) expresses it as follows: “Our empirical
knowledge is a compound of that which we receive through impressions, and that
which the faculty of knowledge supplies… Experience itself is a mode of know-
ledge which involves understanding” (pp. 22 & 16).
In attaching meaning to visual stimuli, or any other kind of stimulus, there are
two significant influences: first, our ability to discriminate – that is, our capacity to
attach meaning; second, our past experiences and, therefore, our expectations of
what we will see – that is, the meaning we are able to attach. Discrimination refers
to our capacity to detect differences between and among stimuli. For example, we
can perceive a shape either as a background or as a figure, though it is sometimes
difficult, in the absence of other stimuli, for our brains to ‘decide’ which is which.
In Figure 3.1a, for example, we may see a white vase on a black field or two black
faces, almost nose to nose, on a white field; in Figure 3.1b we may see black cats
or white cats, and sometimes both together. To an extent we can ‘choose’ what we
see; we can make the picture flip from one version to the other, almost at will.

Figure 3.1a. White vase or blacks faces in profile?

42

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
THE TRADITIONAL VIEW OF SCIENCE

Figure 3.1b. Black cats or white cats?

Figure 3.2 is somewhat more interesting and illustrates quite neatly how prior
knowledge (or its absence) influences perception. About eighteen years ago, when
this symbol began to appear in the popular press as advance publicity for an up-
coming movie, many adults wondered why newspapers were reproducing pictures
of someone’s tonsils. They had attached a particular meaning to the symbol. After
seeing the movie or simply becoming aware of all the surrounding publicity and

Figure 3.2. Batman logo

43

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
CHAPTER 3

merchandising most people easily recognized the Batman logo. They now had a
different meaning to attach.1
Those with previous experience of seeing 3-dimensional objects represented by
2-dimensional diagrams (as in science textbooks, for example) will readily inter-
pret Figure 3.3a as a cube, but what, one wonders, would those unfamiliar with the
convention make of this drawing? What is particularly intriguing is that we can
see the cube with either the bottom left square or the top right square as the front of
the cube. According to Gillies (1993), this effect was first noticed by the Swiss
crystallographer, L.A. Necker, as he attempted to draw a crystal he was observing
through a microscope. Exactly the same factors impact on our response to Figure
3.3b. Those unfamiliar with 2-dimensional line drawings and with flights of steps
would almost certainly interpret this diagram in other ways. Again, it is possible to
see this flight of steps from ‘on top’ or ‘from underneath’.2 Similarly, we can see
Figure 3.3c as a rabbit’s head facing left or a duck’s head facing right. For all three
figures, most people find one way of seeing the picture easier than the other, though
they can (with an effort) make the picture ‘flip’ from one to the other. In other words,
we can learn to see things differently when we know that alternatives are possible.

Figure 3.3a. Which face is the front of the cube?

44

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
THE TRADITIONAL VIEW OF SCIENCE

Figure 3.3b. Steps can be seen from ‘on tope’ or ‘from underneath’

Figure 3.3c. Rabbit or duck?

45

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
CHAPTER 3

Whether one sees an old woman or a young woman in Figure 3.4 also depends
on something other than the image falling on the retina. Significantly, it has been
found that the likelihood of a particular response (old woman or young woman)
can be increased by prior presentation of unambiguous pictures of old or young
women (Leeper, 1935). Clearly, pre-existing mental constructs cause us to see things
in a particular way. As Barlex and Carre (1985) put it: “We don’t see things as they
are, we see things as we are” (p. 4).

Figure 3.4. Old woman or young woman?

46

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
THE TRADITIONAL VIEW OF SCIENCE

Conversely, there are occasions when we do not see what we expect to see and
cannot exert conscious control on our perception. Even when we measure the lines
in the Muller-Lyer illusion (Figure 3.5) to convince ourselves that they are of equal
length, we still perceive the lower line to be longer. In this situation, what we see is
clearly not influenced by what we know. Try as we will, our knowledge, and there-
fore our expectations, cannot determine our perception.

Figure 3.5. Muller-Lyer illusion

Optical illusions created when incoming sense data are incompatible with existing
mental constructs can be a source of amusement, as in Figures 3.6a and 3.6b. They
can also be very irritating and sometimes very disconcerting because they violate
our pre-existing understanding about how things are constructed. An Ames room,
where people appear to change size as they walk across the room, is particularly
puzzling. We know that this is not happening but our expectations about the structure
of a room override our interpretation of what we see. Figure 3.7 is a variant of the
Ames room illusion that illustrates the effect quite well.
We expect Figure 3.8a and Figure 3.8b to represent real three-dimensional
objects, and when that expectation is not met we try to rationalize it, and again we
fail. Our emotional response to this continued frustration may be very illuminating
about our character and may provide interesting potential for a research project. In
some extreme cases, human brains can reject sense perceptions on the grounds of
logic and prior experience. Experiments have shown that when someone is fitted
with a pair of spectacles having lenses that invert everything, they will initially be
quite unable to make sense of what they see, and will find it almost impossible to
move around in a safe and predictable way. However, after a while, the brain adapts
and they see everything the right way up again. Subsequently, when the spectacles
are removed, the world once more appears inverted, leading to a further period of
readjustment (see Gregory, 1977).
What I am arguing here is that observation is not reliable. It is not always
indicative of the true state of affairs, as in these optical illusions. This creates a pro-
blem for scientists: how can they know when to rely on what they see? Further-
more, a change in mental construct (existing knowledge) will often bring about a
change in perception. Thus, once you have seen the faces hidden in the foliage in
the puzzle pictures often found in children’s comics, you can no longer see the
trees without seeing the faces. It is not the image falling on the retina that has
changed, but the observer. The observer now has a different perspective, a different
view of the world. For years, I was unable to see the face hidden in the snowy
landscape of Figure 3.9. For me, it remained simply a series of blotches, despite the

47

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
CHAPTER 3

insistence of a colleague that it reveals the face of Jesus of Nazareth. Then, a few
years ago, confronted by a giant reproduction of the picture at the Ontario Science
Centre and urged by my wife to squint, stare hard and think of that familiar picture
of Che Guevara, I finally saw the face. Now I cannot look at the picture without
seeing the face. My world is forever changed! What was needed in order for me to
see the face was a suitable frame of reference. For some, this frame of reference is
a Christian one; for others, including myself, it is an interest in sociopolitical revo-
lution. Perhaps there is scope for another research project here.

Figure 3.6a. Are the horizontal lines parallel?

Figure 3.6b. Which vase is wider?

48

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
THE TRADITIONAL VIEW OF SCIENCE

Figure 3.7. All three pillars are of equal height

Readers who wish to experience this kind of “perceptual revelation” for them-
selves should spend some time trying to discern what is hidden in Figure 3.10. If
success is not achieved within 10 minutes or so, and frustration levels are becoming
unbearable, the ‘answer’ can be found at the end of the chapter.3 Looking at the
picture when given this information makes all the difference.

49

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
CHAPTER 3

Figure 3.8a. Impossible figure 1

Figure 3.8b. Impossible figure 2

50

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
THE TRADITIONAL VIEW OF SCIENCE

Figure 3.9. Find the hidden man

Figure 3.10. What is hidden in this drawing?

51

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
CHAPTER 3

These simple examples are illustrative of the way in which scientific observers
have to learn how to interpret what they see through a microscope or a telescope,
observe on an oscilloscope screen or in a cloud chamber, or obtain via any of the
modern instrumental devices used by scientists. Michael Polanyi (1958) makes this
point superbly in his graphic description of a medical student struggling to make
sense of X-ray photographs.
He watches in a darkened room shadowy traces on a fluorescent screen
placed against a patient’s chest, and hears the radiologist commenting to his
assistants, in technical language, on the significant features of these shadows.
At first the student is completely puzzled. For he can see in the X-ray picture
of a chest only the shadows of the heart and the ribs, with a few spidery
blotches between them. The experts seem to be romancing about figments of
their imagination; he can see nothing that they are talking about. Then, as he
goes on listening for a few weeks, looking carefully at ever new pictures of
different cases, a tentative understanding will dawn on him; he will gradually
forget about the ribs and begin to see the lungs. And eventually, if he per-
severes intelligently, a rich panorama of significant details will be revealed to
him: of physiological variations and pathological changes, of scars, of chronic
infections and signs of acute disease. He has entered a new world. He still
sees only a fraction of what the experts can see, but the pictures are definitely
making sense now and so do most of the comments made on them. (p. 101)
Of course, the X-ray pictures have not changed at all. What has changed is the
observer (the medical student) and the observer’s ability to interpret observations
using theoretical knowledge. It is knowledge (concepts and theories) that enable us
to make meaningful, significant, scientific observations. As N.R. Hanson (1958)
puts it: “there is more to seeing than meets the eyeball”. He goes on to ask:
Would Sir Lawrence Bragg and an Eskimo baby see the same thing when
looking at an X-ray tube? Yes and no. Yes – they are visually aware of the
same object. No – the ways in which they are visually aware are profoundly
different. Seeing is not only the having of a visual experience; it is also the way
in which visual experience is had. (p. 15)
It follows that unless teachers provide extensive guidance, there can be no
guarantee that students in school science lessons will observe even the readily
observable. For example, Stevens (1978) describes how students lacking the necessary
theoretical background will fail to make even the simple observation that evaporation
of distilled water and tap water on separate microscope slides produce different
results.
This seems a simple enough observation to make, but in practice few children
succeed, and they only notice then after considerable prompting from the
teacher. The reason is not difficult to appreciate. The difference between a
clean and ‘dirty’ slide is of no interest whatever – and will therefore not be
noticed – unless the pupil already has the idea that water is a solvent, and
may contain dissolved substances. (p. 104)

52

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
THE TRADITIONAL VIEW OF SCIENCE

The key to scientific observation, as distinct from simple everyday ‘looking at


things’, is a sound theoretical frame of reference. As Peter Medawar (1967) says,
what a child sees “conveys no information until he knows beforehand the kind of
thing he is expected to see” (p. 133). Thus, it is the science teacher’s job to ensure
that students perceive the world in the appropriate way – that is, the way in which
currently accepted science (or the school version of it) deems appropriate.

MYTH 2: SCIENCE STARTS WITH OBSERVATION

The inductive process outlined in the traditional view of scientific method, and
represented diagrammatically in Figure 3.11, requires the assembly of all relevant
observations and information, from which a generalization will eventually emerge
through application of logical analysis. How, one might ask, does an innocent and
unbiased observer know what is relevant and, therefore, what to observe? The
inductive method offers no guidance on the restriction of observations to anything
less than the whole universe. As Medawar (1969) reminds us, “We cannot browse
over the field of nature like cows at pasture” (p. 29). Observation, especially scien-
tific observation, is a selective process and so requires a focus of attention and a
purpose. An observer needs an incentive to make one observation, rather than
another. Induction does not provide that incentive.

Figure 3.11. Science as induction

Making a scientific observation presupposes a view of the world that suggests


particular observations can be made and are worth making – i.e., they are of scien-
tific significance. In other words, scientific observation is neither innocent nor un-
biased. It is not ‘objective’ in the sense that school science curricula imply. Rather,
it is purposeful, theory-dependent and theory-driven. In practice, some view of the
world (some theoretical perspective) precedes observation and guides it. It is simply

53

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
CHAPTER 3

not possible to observe things for which you are conceptually unprepared – a
position admirably summed up by David Theobald (1968): “If we confront the
world with an empty head, then our experience will be deservedly meaningless.
Experience does not give concepts meaning, if anything concepts give experience
meaning” (p. 26).
Heisenberg (1971) makes the point that theory and theoretical speculation create
the possibility of new observations that lead to scientific advances: “On principle,
it is quite wrong to try founding a theory on observable magnitudes alone. In reality
the very opposite happens. It is the theory which decides what we can observe”
(p. 63). As Hanson (1958) points out, until Paul Dirac postulated the existence of
positrons, the tracks they leave in cloud chambers were not seen at all or were
dismissed as mere experimental noise. When armed with the new knowledge Dirac
gave them, physicists found clear evidence in those same cloud chamber experi-
ments of the existence of positrons. Likewise, sunspots went unrecorded in Europe
until Galileo’s work overthrew belief in the perfection of the heavens, whereas
Chinese astronomers (with no such over-riding beliefs) had been recording them
for centuries. It seems to follow that there cannot be a piece of absolutely indispu-
table observational knowledge whose meaning is not impregnated in some way by
prior belief about the world. Bauer (1992) expresses this conclusion as follows:
A mountain is, variously and for different people, an illustration of plate
tectonics, or a demonstration of God’s handiwork, or an example of the in-
comprehensibility of the world, or the abode of certain spirits, or something
else again; and each of these possible connotations determines how we see a
mountain – as old or young or eternal, as generic or idiosyncratic, and so on.
Whatever the characteristic involved, to each of us the mountain is something
different; even though we all agree that it is a fact, it is not quite the same fact
for each of us. (p. 65, emphasis in original)
Arguing along similar lines, Bohm and Peat (1987) suggest that, in respect
of making observations, it might make more sense to regard scientists as being
“somewhat like artists who produce quite different paintings of the same sitter…
Some interpretations may show creative originality while others may be mediocre.
Yet none give the final ‘truth’ about the subject” (p. 102). For the purposes of
this chapter, it is sufficient to conclude that our sensory perception of the world is
profoundly affected by the beliefs we hold and the theories to which we subscribe.
When we change our theories, we open up new ways of seeing and explaining the
world. While it would be tempting to believe that the use of scientific instruments
will simultaneously extend our capacity to observe and reduce the subjectivity of
our observations, it should be remembered that theoretical assumptions underpin
the design and construction of all scientific instruments. In short, there seems to be
no easy escape from our theoretical preconceptions.
It is the technical language of science, to which Polanyi refers in the quotation
given earlier, that conveys the meaning with which we invest our observations. All
observation statements employ theoretical language. Even an apparently simple
factual statement such as “Anhydrous copper sulphate has a solubility of 205

54

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
THE TRADITIONAL VIEW OF SCIENCE

grams per litre at 20C” only makes sense in the light of a pre-existing theoretical
framework involving concepts such as dissolving, temperature, hydration, volume
and so on. Moreover, the quality and usefulness of observation statements depend
crucially on the level of sophistication of the theoretical language available to
the observer. Without such a language, perceptions cannot be given meaning and
observations cannot be recorded and subjected to critical scrutiny.4 What is des-
cribed in scientific observations is never ‘pure phenomena’ (whatever that might
mean), but phenomena seen through particular ‘theoretical eyes’. Theoretical
knowledge opens up possibilities that otherwise would not exist and, as a science
develops and builds new theoretical knowledge, scientists are able to generate
knowledge by making different observations.5 Thus, we learn about nature and we
also learn how to learn about it, by learning (i) what constitutes significant infor-
mation, (ii) how to collect it, and (iii) how to interpret it and communicate it to
others. Smart (1968) provides the following illustrative example.
A scientifically untutored peasant could certainly make a report that a black
needle-like thing pointed to the figure ‘35’ on a round clock-like thing. He
could not report that the current through a milliammeter was thirty-five milli-
amperes, since he would not have the concept of an electric current, and still
less would he have the concept of an ampere. (p. 80)
To summarize, scientists work within a theoretical framework that guides their
actions and invests observational data with particular meaning. In a sense, theory
is both a boon and a curse: it opens up new possibilities for observation and experi-
ment but it also constrains thought and possibilities. By accepting its premises, one
accepts its particular way of looking at the world, making it difficult to make
theoretical progress. Conversely, being ‘open-minded’ sounds fine, and is extolled
as a virtue in school science because it seems to ensure that observations are un-
prejudiced by prior expectations. But if we are too ‘open-minded’ we may miss what
a particular theoretical stance would tell us. Open-minded can easily become ‘no-
minded’, leaving us with no theory with which to make sense of the world.
Because observation is theory-impregnated, it follows that observations have to
be checked for acceptability by recourse to theory. In Arthur Eddington’s words,
“It is also a good rule not to put overmuch confidence in the observational results
that are put forward until they have been confirmed by theory” (cited by Stent,
1969, p. 31). This is the reverse of what science teachers usually tell students. The
usual message is that we have to test our theories for acceptability against reliable
observations – i.e., against ‘the facts’. In practice, scientists, and students in school
science lessons, often have to reject sense data on theoretical grounds: a stick par-
tially immersed in water is not bent, nor does Mars travel backwards in its orbit
from time to time, whatever appearances might suggest. Theory tells us when
observations are mistaken and should be rejected.
The objects seen through the magnifying glass appear circled by colours of
the rainbow; is it not the theory of dispersion which teaches us to regard these
colours as created by the instrument, and to disregard them when we describe
the object observed? And how much more important this remark is when it is

55

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
CHAPTER 3

no longer a matter of a simple magnifying glass but of a powerful micro-


scope! (Duhem, 1962, p. 154)
The theory dependence of observations may, on occasions, lead scientists to
make false and misleading observations simply because the theoretical position
they have adopted is incorrect. As Beveridge (1961) remarked, more than forty
years ago, “Not only do observers frequently miss seemingly obvious things, but
what is even more important, they often invent quite false observations” (p. 98).
This ‘invention’ is not an attempt to perpetrate a deliberate falsehood, merely a
consequence of a mistaken theoretical perspective. It is a phenomenon well known
in school science lessons, where students will often see what they expect to see and
miss that which they do not anticipate. When students have a different theoretical
framework from that assumed by the teacher in the design of a laboratory activity
they may look in a different place (the ‘wrong’ place), in a different way (the ‘wrong’
way) and make different interpretations (incorrect ones), sometimes even vehe-
mently denying observational evidence that conflicts with their expectations.
Just as frequently, they may adjust or modify their observations to conform with
the expectations their existing theoretical framework gives rise to – just like
the scientists described by Beveridge. In consequence, students may go through the
entire lesson misunderstanding the purpose of the activity, the procedure and the
findings, and further compounding the misconceptions they brought to the lesson.
As Hodson (1998a) comments, “It is not too much of an exaggeration to say that,
because predictions, perceptions and explanations are all strongly influenced by
prior conceptual understanding, students who hold different frameworks of meaning
conduct different investigations, with correspondingly different learning outcomes”
(p. 28). The alternative conceptions literature contains many such examples and,
thereby, constitutes a major argument in favour of a constructivist pedagogy that
takes account of these different starting points.
Recognizing that scientific observation is theory-laden and that students may
sometimes have a different framework of reference from the teacher has another
important consequence for science teachers: that the skills of scientific observation
have to be taught and, moreover, taught and learned within particular theoretical
contexts. It is just not possible to teach someone to observe in a way that is inde-
pendent of the context in which the observation is to be made. In science lessons,
we are not teaching students to observe per se. They can already do that; they have
been making observations for many years, since long before they came to science
classes. Our responsibility is to teach them to make scientific observations, and for
that they need appropriate conceptual understanding.6

OBSERVATION AND INFERENCE

Before leaving this discussion of the relationship between observation and theory,
there is one more matter that is worthy of discussion. In common with many other
advocates of the traditional model of science, Zeitler and Barufaldi (1988) state
that “observations are perceptual, while inferences are interpretive” (p. 97) and

56

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
THE TRADITIONAL VIEW OF SCIENCE

Abruscato (1988) asserts that “nothing is more fundamental to clear thinking than
the ability to distinguish between an observation and an inference” (p. 33). Super-
ficially, this distinction sounds fine and seems to accord with what we consider to
be good practice in scientific inquiry: having respect for the evidence and not
claiming more than the data can justify. However, closer examination in the light
of earlier discussion about the theory-laden nature of scientific observation suggests
that the supposed demarcation is not always as clear as Abruscato and others might
claim. When a new theory appears, or when new scientific instruments are deve-
loped, our notion of what counts as an observation and what counts as an inference
may change. As Feyerabend (1962) points out, observation statements are merely
those statements about phenomena and events to which we can assent quickly,
relatively reliably and without calculation or inference, because we all accept, with-
out question, the theories on which they are based. Thus, where individuals draw
the line between observation and inference reflects the sophistication of their scien-
tific knowledge, their confidence in that knowledge, and their experience and
familiarity with the phenomena or events being studied. When theories are not in
dispute, when they are well understood and taken-for-granted, the theoretical langu-
age is the observation language, and we use theoretical terms in making and reporting
observations. Terms like reflection and refraction, solution and suspension, con-
duction and non-conduction, all of which are used regularly in school science as
observation terms, carry a substantial inferential component rooted in theoretical
understanding. The key point is that unless some theories are taken for granted
(and deemed to be no longer in dispute) and theory-loaded terms used for making
observations we can never make progress. We would forever be trying to retreat to
the ‘raw data’, to some position that we could regard as theory-free.
Hanson (1972) refers to terms that “carry a conceptual pattern with them”
(theoretical) and terms that are “less rich in theory, and hence less able to serve in
explanations of causes” (p. 60). As an example of the former category he cites the
word crater, which he says carries with it a cargo of astronomical theory that words
like concavity and hole do not.
[The moon] is pitted with holes and discontinuities; but to say of these that
these are craters – to say that the lunar surface is craterous – is to infuse
theoretical astronomy into one’s observations… To speak of a concavity as a
crater is to commit oneself as to its origin, to say that its creation was quick,
violent, explosive. (p. 56)
He goes on to argue that theoretical terms may be employed to organize diffuse
and seemingly unrelated aspects of a situation into a coherent, intelligible pattern,
whereas ‘phenomenal’ terms cannot be so employed. However, Hanson later weakens
this distinction by admitting that terms may be organizers in one situation and
phenomenal in another. In other words, there is no sharp distinction: “It is not that
certain words are absolutely theory-loaded, whilst others are absolutely sense-
datum words. Which are the data-words and which are the theory-words is a con-
textual question” (Hanson, 1958, p. 59).

57

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
CHAPTER 3

Perhaps no-one would quibble with Gilbert Ryle’s (1956) comment that theor-
etical terms are such that knowing their meaning and using them appropriately
necessitates a grasp of some theory.
The special terms of a science are more or less heavy with the burthen of the
theory of that science. The technical terms of genetics are theory-laden, laden,
that is, not just with theoretical luggage of some sort or other but with the
luggage of genetic theory. (p. 90)
Non-theoretical terms, presumably, carry none of this “luggage of theory”. While
this distinction sounds reasonable, it fails to account for the fact that many scien-
tific concepts appear in more than one theory. In some cases a term can only be
understood with respect to a particular theory; in other cases, the term is simply
being ‘used’ by a theory and has its meaning anchored elsewhere - in some other
theory. With respect to theory T1 a term may be ‘theory-laden’, whilst with res-
pect to another (T2) it may be ‘theory-free’, but all terms are anchored in some
theory, even if it is only the ‘theory’ of everyday commonsense. As Popper (1959)
declares:
Every description uses universal names (or symbols or ideas); every state-
ment has the character of a theory, of a hypothesis. The statement, ‘here is a
glass of water’ cannot be verified by observational experience. The reason is
that the universals which appear in it cannot be correlated with any particular
sense-experience… By the word ‘glass’, for example, we denote physical bodies
which exhibit a certain law-like behaviour, and the same holds of the word
‘water’. Universals cannot be reduced to classes of experiences. (p. 94)
When a new theory appears our notion of what is a theoretical term and what is
an observation term may change: “Introducing a new theory involves changes of
outlook both with respect to the observable and with respect to the unobservable
features of the world, and corresponding changes in the meanings of even the most
‘fundamental’ terms of the language employed” (Feyerabend, 1962, p. 29). Only when
we have good reason to doubt the theoretical grounding of a particular obser-
vational language do we retreat to relatively theory-free terms like ‘thin lines’ to
describe the features of cloud chamber photographs; when we are confident, we use
the much more powerful, theory-loaded terms such ‘track of an electron’.
We approach a problem-situation with the strongest justified description, and
only withdraw to less commital, more neutral ones when specific reason for
doubt arises – and even then, we withdraw only as far as necessary with respect
to the available reasonable alternatives (Shapere, 1982, p. 520)
There can, of course, be no scientific language sufficiently free of theoretical
assumptions that no doubt can arise, but the admission that (philosophically) doubt
can arise does not mean that we have to be doubtful and to assume that our theories
are unreliable. While we cannot claim scientific knowledge to be certain (we could
be mistaken), this does not mean that it is uncertain (i.e., unreliable and arbitrary).

58

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
THE TRADITIONAL VIEW OF SCIENCE

As Shapere (1982) assures us, “the mere possibility of doubt arising is not itself a
reason for doubt” (p. 515). While the ways in which we make sense of the world are,
in principle, open to doubt and are subject to revision and change as science pro-
gresses, the truth is that we can take action in all kinds of ways, on the basis of our
theoretical understanding, with entirely predictable outcomes. Confidence in scien-
tific knowledge is a direct consequence of science being a social process: we
collectively construct our view of the world and collectively validate it (charac-
teristics of science to be discussed in subsequent chapters). The so-called ‘facts’ of
science are simply those observation statements admitted to the corpus of scientific
knowledge because the theories and procedures on which they depend have achieved
consensus within the scientific community. As Churchland (1979) states, scientists
learn from other scientists to perceive the world as the community of scientists per-
ceives it. The exchange and criticism of ideas that leads to this consensus view is
the basis of scientific rationality and the primary reason for our confidence in
scientific observations.7
The acquisition and consolidation of new conceptual understanding carries with
it the possibility of making more sophisticated observations. For example, once
they have a theory of solubility, students see things dissolve, where previously they
saw them disappear. Moreover, once they understand that there is an important
conceptual difference between dissolving and melting, students see that it is impor-
tant for them to be careful in their use of the terms. Younger children, without this
knowledge, will continue to refer to sugar and salt ‘melting’ in water. They have
no reason to do otherwise. Students can be made aware of the ways in which their
observational skills change and develop as their theoretical understanding becomes
more sophisticated by repeating an observational exercise from earlier in the course,
or from a previous year. The new description employs observational language that
encompasses previously unknown theoretical notions. Perhaps materials in the
chemistry lab can now be observed to melt, sublime or decrepitate on heating, where
previously they had just ‘changed’. All three of these new terms include theoretical
inference. A thoughtful discussion of the theoretical assumptions underpinning the
design and construction of common laboratory instruments – beginning with some-
thing simple, like thermometers, and progressing to ammeters, voltmeters, pH meters,
and the like – can assist students to the realization that the supposed distinction bet-
ween objective observation and theoretical inference is less clear than some science
textbooks and some science teachers would assert, and is more a characteristic of
their own stage of conceptual understanding, and their confidence in that know-
ledge, than a clear demarcation between two processes of science.8
By investing their observational language with additional theoretical assumptions,
and thereby designing and conducting more sophisticated scientific investigations
and experiments, scientists can construct more elaborate theories. By repeating the
cycle at increasingly sophisticated levels, scientific understanding is propelled
forward. In a sense, science ‘hauls itself up by its own boot straps’. The same argu-
ment applies to those learning science. However, both scientist and science learner
need to be wary of assuming too much. There are occasions when inferences should
be recognized as no more than tentative. There are occasions when it is unwise to

59

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
CHAPTER 3

invest too much confidence in what observations seem to tell us. This is my some-
what oblique way of returning to the traditional view of science and the centrality
of induction in the so-called scientific method.

MYTH 3: SCIENCE PROCEEDS BY INDUCTION

In the traditional view of the scientific method, observational data can be carefully
organized and subjected to rational appraisal in order to ascertain patterns and
formulate generalizations. These generalizations, and the predictions to which they
give rise, can then be confirmed by further observation or experiment (see Figure
3.11). Hempel (1966) asks us to consider how a scientist who subscribes to an
inductivist model of science would conduct an investigation…
First, all facts would be observed and recorded, without selection or a priori
guess as to their relative importance. Secondly, the observed and recorded
facts would be analysed, compared, and classified, without hypothesis or
postulates, other than those necessarily involved in the logic of thought. Third,
from this analysis of the facts, generalizations would be inductively drawn
as to the relations, classificatory or causal, between them. Fourth, further re-
search would be deductive as well as inductive, employing inferences from
previously established generalizations. (p. 11)
Since earlier discussion has addressed the absurdity of the notion of innocent,
unbiased observation embedded in this description, attention can now be focused
on the inductive process itself – in particular, on the question of the validity of
inductive inferences from observable ‘facts’. Those who advocate this approach
argue that a generalization (an inductive inference) can be relied on if certain con-
ditions are met.
– The number of observations is large
– The observations are made under a variety of conditions
– No accepted observation statement conflicts with the derived generalization.
Suppose, then, that you have observed that a metal bar expands when heated.
First, you should repeat the test several times to ensure consistency of outcome.
Second, you should vary the conditions: iron, copper, zinc and gold bars, at high
temperature ranges and at low temperature ranges, under increased pressure and
under reduced pressure, and so on. If any metal does not expand on heating, under
any of the conditions, the generalization “all metals expand on heating” must be
discarded. However, provided that all evidence is supportive, the generalization can
be accepted. Moreover, according to the model, the generalization can subsequently
be used to make predictions, as follows:
1. All metals expand on heating
2. This railroad track is made of metal
3. This railroad track will expand on heating (and, therefore, there needs to be a
series of expansion joints to prevent buckling of the track).
This simple, straightforward and logical model of how science functions seems
fine until one starts to look carefully at the conditions for making an inductive

60

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
THE TRADITIONAL VIEW OF SCIENCE

inference. The first condition requires a large number of observations. How large is
‘large’? Will five observations suffice? Do we need to make 10, 100, 1000 or 10,000
observations? It isn’t clear. The second condition requires a wide variety of con-
ditions. Testing iron, copper and zinc bars seems reasonable enough, but what
about long bars versus short bars? How about fat bars and thin bars, or red ones
and yellow ones? On Mondays, Tuesdays and holidays? Again, it isn’t clear. The
third condition states that there should be no conflicting observation. How does one
know whether it is a genuinely conflicting observation or just a ‘bad reading’ that
should be repeated? How does one know when to accept the data, when to go back
and check it, and when to reject it entirely? It isn’t clear. In each case, the answer
to the uncertainty is located in theoretical understanding. Theory tells us how many
observations are enough, what particular variables to consider and how to dis-
tinguish good observations from bad ones. However, theory is not permitted to
intrude into the collection of data (see the Hempel quotation above). Observation is
supposed to be objective, value-free and unguided – a notion that has already been
dismissed as absurd earlier in the chapter.
So far, I have argued that inductive generalizations are not reliable because they
derive from observations, which are themselves unreliable – first, because of the
limited range of human perception9; second, because the nature of the observations
depend crucially on who makes them, as discussed at length earlier in the chapter.
Inductive inferences are also vulnerable on grounds of logic. In passing from parti-
cular observations (even if considered ‘facts’) to the generalizations that compre-
hend them, something is added. A generalization is not simply a re-statement of the
facts (or it would be valueless), but it cannot logically contain more information
than the empirical content of the statements from which it is derived, as David
Hume (1854) reminds us:
There can be no demonstrative arguments to prove that those instances of
which we have had no experience, resemble those of which we have had
experience… even after the observation of the frequent or constant conjunc-
tion of objects, we have no reason to draw any inference concerning any object
beyond those of which we have had experience. (p. 390)
In other words, an inductive inference is not logically valid. A generalization derived
from singular statements, no matter how numerous, may still turn out to be wrong.
No matter how many white swans we see, we are never justified in asserting that
“all swans are white”, because the next swan we see may well turn out to be black
– as any student studying Philosophy 101 will eagerly point out. Bertrand Russell
(1912) makes this point rather more humorously in his tale of the inductivist
turkey, who was fed each day at 9.00 am. Not wishing to jump to conclusions, the
turkey ensured that he made his observations under a wide variety of conditions
(days of the week, variations in weather, holidays, etc.). Eventually he was satis-
fied and ventured the generalization “I am always fed at 9.00 am”. Sadly, at 9.00 am
the following day, which happened to be Christmas Eve, he had his throat cut. An
inductive inference based on true singular statements, however numerous, may still

61

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
CHAPTER 3

turn out to be false. The best one can say is that, so far, the observational evidence
has not falsified the generalization.10
If inductivism is so fatally flawed, why do so many science teachers continue to
tell students that science proceeds in this way? One explanation may be that it has
the attraction of being simple and seemingly straightforward. Observations can be
made by anyone, simply by careful and diligent use of the senses. Inductive infer-
ences require nothing more than logical scrutiny of the data. My observations of
twelve science teachers in New Zealand secondary schools show that many teachers
adopt an inductivist stance with those students they consider to be less able, simply
because they believe it is “easier” for them to understand (Hodson, 1993a).11 A second
possible reason for the appeal of inductivism is that there is no personal, subjective
element involved: the validity of observation statements, when observations are care-
fully and ‘correctly’ accumulated, are independent of the opinions and expectations
of the observers. Thus, the model supports the belief that scientists are objective,
dispassionate and disinterested (the ‘white coat’ image of the scientist) – a position
that some find attractive and some find to be politically advantageous because it
invests science with a spurious authority (as discussed in chapter 2). A third argu-
ment is that it has proved successful in the past. On many occasions, inductively
produced generalizations (and the laws and theories derived from them) have proved
perfectly satisfactory in providing explanations and making predictions. There-
fore, it is argued, induction is justified by experience. Our Philosophy 101 students
should immediately disallow this argument as circular: a universal statement asser-
ting the validity of the principle of induction has been inferred from a series of
singular statements (the individual instances of previous success) – that is, by
induction. By using as justification the very argument it is trying to justify, it
assumes what it is trying to prove.
Those advocates of the inductivist method still anxious to salvage it may well
turn to probability, arguing (for example) that if a large number of metals have
been observed to expand on heating, under a wide variety of conditions, then it is
probably true that all metals expand on heating. However, this is not a solution to
the logical inadequacy of induction. Although the claim is weaker than before, it is
still subject to the same criticism; while it seems reasonable to believe that increa-
sing observational support increases the probability that a generalization is true,
probability theory tells us that it is not the case. Regardless of the number of singular
observations that a scientist makes, the ratio of ‘actual observations made’ divided
by ‘number of situations to which the generalization is claimed to apply’ is zero. In
other words, because universal generalizations claim to apply to all possible cases
the probability that the generalization is true is zero.
Faced with this reality, there are three possible courses of action. We could
accept Bertrand Russell’s (1961) conclusion that “every attempt to arrive at general
scientific laws from particular observations is fallacious… and without this principle
science is impossible” (p. 647) or, if not impossible, not able to be rationally
justified. Alternatively, we could simply take the view that it is ‘obvious’ that
induction works because ‘we use it all the time’. In other words, it is just common
sense.12 Unfortunately, ‘obvious’ things sometimes turn out to be untrue – for

62

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
THE TRADITIONAL VIEW OF SCIENCE

example, heavy objects do not fall faster than lighter ones unless there are air
resistance factors in play. Nor is common sense or majority opinion always a
reliable guide to the truth. The third option, and the one to be followed in chapter 4,
is to seek an alternative explanation for how science functions, an alternative view
of scientific method and an alternative view of the rationality of science.
It is reasonable to demand that a theory of science does two things: (i) it des-
cribes the ways in which successful episodes in science have been conducted in the
past; (ii) it provides practitioners with some advice on procedures to be adopted in
the future – i.e., some guidelines for practice. It has been argued in this chapter that
induction fares pretty poorly on the second requirement. It is unable to provide an
initial focus for study (what observations to make, and how to make them); its
alleged stance of objectivity is profoundly compromised by the theory-impregnated
nature of observational evidence; it is logically invalid. Turning to the first criterion,
we should ask how well inductivism stands up to historical scrutiny and/or
testimony from contemporary practitioners on their methods of inquiry. Is there
evidence that scientists proceed inductively? Have they done so in the past, with
any degree of success? There is clear evidence that science does use inductive
methods from time to time, despite its dubious logical status, as some of the scien-
tists interviewed by Wong and Hodson. (2008a) confirm. For example, by examining
a large number of individuals, geneticists have determined that all Down’s Syndrome
sufferers have an extra chromosome (47 instead of 46). Seemingly, no prior theo-
rizing was involved. Geneticists also make use of probability to explain inherited
characteristics such as eye colour. Isaac Newton, in his master work Philosophiae
Naturalis Principia Mathematica (1687) claimed that his law of universal gravitation
had been inferred from the existing data – in particular, from Kepler’s laws. The
problem with Newton’s claim for inductivism is that Kepler’s laws have the planets
moving in perfect ellipses around the sun, whereas Newton’s law of gravitation
predicts that planetary orbits cannot be perfect ellipses. Moreover, Kepler’s laws
are expressed in terms of distances, time intervals, velocities and so on, while
Newton’s laws use the concepts of mass and force. Clearly, there was more
involved than simple induction from the data.
Science does appear to use a form of non-deductive inference that philosophers
call “inference to the best explanation’ (IBE), sometimes known as abduction:
Given a particular set of observable events, what is the most likely explanation? For
example, one can argue for the theory of ‘continental drift’ (later developed into
plate tectonics) on the grounds that it provides a more plausible explanation than its
competitors for common geological features in widely separated regions of the
planet and the seeming overall correlations in the shapes of the major continents.
Similarly, Darwinian theory can explain a very diverse range of facts about the
living world, including anatomical similarities between species. Although each of
these facts might be explainable in other ways, the theory of evolution accounts for
all the facts in one coherent explanation. Of course, those wishing to use IBE still
need a way of deciding which of the competing hypotheses represents the best
explanation. In Darwin’s case, the criteria included elegance, simplicity and
coherence.13 To take another 19th Century example, what criteria could be used to
settle the dispute among rival explanations for Brownian motion?14 Options included

63

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
CHAPTER 3

electrical attraction between particles, convection currents and collisions between


the suspended particles and the molecules in the surrounding fluid. From our 21st
century theoretical perspective, IBE might incline us towards the third explanation.
In fact, it was not until Einstein’s precise predictions from a mathematical treat-
ment of Brownian motion, in 1905, that the kinetic theory explanation was finally
established. In other words, in the absence of other considerations IBE is insuffi-
cient to establish the validity of a theory or settle a dispute. Indeed, it may lead
scientists to choose the ‘wrong option’.

ENDNOTES
1
I have continued to use this example in my graduate classes because many students from non-
Western countries have no knowledge of Batman and so remain puzzled by the picture. Inter-
estingly, these students find it easier to accept the Batman interpretation when presented with the
colour version of yellow on black.
2
A wonderful, though perhaps fanciful, example of previous experience impacting everyday obser-
vation is provided by Turnbull (1961) in his book The Forest People (cited by Riggs, 1992, p. 17).
Turnbull reports that the pygmy inhabitants of the Ituri Forest (a region of Congo) live in an envi-
ronment where visibility is highly restricted and, in consequence, they have not developed the capacity
to gauge size and distance accurately. He describes, though one wonders with what seriousness, an
episode in which he (Turnbull) and Kenge (a tribesman) ventured into an open area outside the tree
line: “And then he saw the buffalo, still grazing several miles away, far down below. He turned to me
and said, ‘What insects are those?’ At first, I hardly understood, then I realized that in the forest vision is
so limited that there is no great need to make an automatic allowance for distance when judging size…
When I told Kenge that the insects were buffalo, he roared with laughter and told me not to tell such
stupid lies” (p. 227). Riggs comments: “the theoretical components of Kenge’s observation included the
preconception that the perceived size of an object is its true size” (Riggs, 1992, p. 17). It’s a nice story
and would illustrate my point perfectly if it were true. An alternative story for students is the Picasso
joke found in Rose (1997): “A man troubled by Picasso’s portraits with eyes facing both frontwards
and sideways asks the painter why he does not paint realist pictures. To make his point clear the man
takes out a photograph of his wife and says: ‘Like this’. The artist looks at the photo and mildly
observes: ‘Small, isn’t she?’” (p. 60).
3
The hidden figure is a white cow with black ears and a black muzzle. Once the cow is detected, you
will continue to see it. You can never return to the previous ‘innocent’ frame of mind. Try it!
4
It will be argued later that critical scrutiny by the community of scientists is the key to scientific
rationality.
5
Whether these ‘different’ observations are also ‘better’ observations will be discussed in chapter 4.
6
Knowing that children will find it easier to observe change than constancy, and differences rather
than similarities, should serve to guide choice of observation exercises in the early years.
7
The significance of consensus within the scientific community to the validity of scientific know-
ledge will be discussed in chapter 5.
8
Some discussion of whether the so-called processes of science are as discrete and theory-independent as
some science educators claim is included in chapter 4.
9
Human beings are sensitive to only a very narrow range of electromagnetic radiation and only to a
very narrow range of sound frequencies. Together with our senses of touch and smell, neither of
which is particularly sensitive, this is the ‘perceptual apparatus’ with which we confront the world.
These limited senses can be expanded through technology (microscopes, telescopes, infrared spec-
trometers, and the like) but we are still, in a very real sense, prisoners of these limited senses and our
intellectual capacity to imagine how to extend them. The real world (whatever that means) may be
suffused with entities we cannot detect and/or have not yet imagined. An added problem, as
Heisenberg pointed out, is that the observer may substantially change the object being observed simply
through the act of observation.
10
This more tentative position, and its value in enabling progress to be made, is well illustrated in the
Fable of a lost child keeping warm in the CHEM STUDY materials (1963a, pp. 3-4).

64

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
THE TRADITIONAL VIEW OF SCIENCE

11
The teachers were also fairly consistent in presenting biology as predominantly inductivist and
chemistry and physics as hypothetico-deductivist (see chapter 4 for a discussion of this latter model
of science). However, the shift in philosophic stance between subjects was more a reflection of the
learning opportunities presented by the particular topics being taught than a belief that different
sciences proceed by different methods.
12
Peter Strawson (1971) argues that induction is so fundamental to human reasoning that we don’t
need to justify it. We can just use it! Induction, he says, is one of the standards we use in deciding
whether claims to knowledge are justified – for example, whether a pharmaceutical company has
sufficient evidence to substantiate its claim for the effectiveness of a particular drug – so it is absurd
to question whether induction itself is justified. It is perfectly proper, he says, to inquire whether
there is good reason to accept a particular inductive inference, but the demand for justification of
induction itself is mistaken. He urges us to compare the situation with the question: “Is the law
legal?” It makes perfectly good sense to inquire whether a particular action is legal or not, and the
answer is provided by an appeal to the prevailing legal system and application of its set of rules. But,
he says, “it makes no sense to inquire in general whether the law of the land, the legal system as a
whole, is or is not legal. For to what legal standards are we appealing?” (p. 257)
13
The deployment of these somewhat subjective criteria in science will be discussed at length in
chapter 7.
14
Brownian motion refers to the chaotic, zig-zag motion of microscopic particles suspended in a liquid
or gas, first observed in 1827.

65

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
Derek Hodson - 978-90-8790-507-1
Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
CHAPTER 4

EXPLORING ALTERNATIVE VIEWS OF SCIENCE


The Ideas of Popper, Lakatos and Kuhn

If the traditional inductive method of science, as projected in many school science


textbooks and curriculum documents, is as seriously flawed as I argued in chapter
3, and if there are so many absurdities and difficulties associated with it, why don’t
teachers discard it? Why don’t we ‘come clean’ and tell students how science
really proceeds? We should tell them that science starts from existing knowledge
and understanding – from exploring it, thinking about it, using it, doubting it,
testing it. Then you get an idea: “I wonder if there is a relationship between x and y.”
“What would happen if I did a or b?” “If so and so is the case, it seems likely that
a, b, c”. Once you have an idea, you refine it, develop it, and then test it. This, in
essence, is what Karl Popper (1959) says in his hypothetico-deductive model of
science when he claims that science proceeds by a 4-step cycle of imagination and
criticism.
– A hypothesis is generated from the existing theoretical background – by creative
imagination or inspired guesswork.
– From the hypothesis, certain observable/testable conclusions are deduced – i.e.,
predictions are made.
– The conclusions (predictions) are tested by observation or experiment.1
– The proposition is accepted or rejected.
To use Popper’s language, if the predictions are borne out by the observational
tests, the hypothesis is corroborated, not proven. If the predictions are not borne
out by the observational tests, the hypothesis has been refuted – that is, shown to be
false – and must be discarded. It may then be modified and the cycle repeated.
According to Popper, science is a constant interplay among hypotheses, the logical
expectations to which they give rise, and observational/experimental test. It is a
constant dialogue between what might be true (the hypothesis) and what is or is
not corroborated by the experimentally determined ‘facts’. In Popper’s language,
science proceeds by means of conjectures and refutations until scientists arrive at a
proposition that satisfactorily accounts for the evidence – a proposition that is not
refuted by the observational evidence. This is not a random, hit-or-miss affair.
There is constant feedback for the modification and restructuring of hypotheses,
with each new conjecture being made in the light of the previous cycle of con-
jecture and refutation. In other words, scientists learn from their mistakes.
There are several important differences between Popper’s view of science and
the traditional inductivist view. First, it should be noted that observation comes very
late in the Popperian model of science; in the inductivist model, it comes first.
More significantly, the first step in Popper’s model is formulation of an hypo-
thesis. In other words, scientists begin their endeavours by thinking, speculating,
generating a new idea. For Popper, imagination comes first. Science is a creative,

67

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
CHAPTER 4

imaginative activity, not a dull, objective, value-free and method-driven procedure.


In philosophical terms, the most distinctive feature of Popper’s model is the
centrality of falsification, rather than verification. Every test of a scientific propo-
sition is an attempt to falsify it, not an attempt to prove it correct. Popper goes on
to argue that there is an asymmetry between confirmation and refutation. Although
a universal statement (a generalization) cannot be confirmed as true by singular
observation statements, no matter how numerous, a universal statement can be
refuted (shown to be false) by a singular observation statement, as the inductivist
turkey found out to his/her cost (see chapter 3). The observation of just one black
swan falsifies the generalization that “all swans are white”,2 while no finite number
of observations of white swans will prove the proposition true. In other words,
falsification is decisive.
By exposing hypotheses to this fierce ‘struggle for survival’ scientists ensure
that only the ‘fittest’ hypotheses are retained – where ‘fittest’ means best accords
with the observational evidence (‘best fits the facts’). By rejecting hypotheses that
fail to stand up to observational and/or experimental test, science makes progress
towards a truer description of the world. The description is ‘truer’ because some
possible explanations have been ruled out: we know that the world is ‘not like that!’
Falsificationists like myself much prefer an attempt to solve an interesting
problem by a bold conjecture, even (and especially) if it soon turns out to be
false, to any recital of a sequence of irrelevant truisms. We prefer this because
we believe that this is the way in which we can learn from our mistakes; and
that in finding that our conjecture was false we shall have learnt much about
the truth, and shall have got nearer to the truth. (Popper, 1963, p. 231)
Of course, scientists would never know if or when they had arrived at the ‘truth’.
There is no access to absolute truth because all human beings are ‘prisoners’ in
the sense that we are confined by our physical environment (which limits our pers-
pective), constrained by our senses and our limited capacity to extend them, and
restricted by the capabilities of our intellect. We can only imagine what we can
imagine! Thus, ‘scientific truth’ – insofar as the term has any meaning – is no more
than ‘our current best shot’. The theories we currently hold are simply provisional
conjectures that we have not yet managed to falsify, though we may do so tomorrow
on the basis of new evidence or a new way of looking at existing evidence. Thus,
the best corroborated theory is the one we have least reason to think is false.
We choose the theory which best holds its own in competition with other
theories; the one which, by natural selection, proves itself the fittest to sur-
vive. This will be the one which not only has hitherto stood up to the severest
tests, but the one which is also testable in the most rigorous way. A theory is
a tool which we test by applying it, and which we judge as to its fitness by
the results of its applications. (Popper, 1959, p. 108)
Although science is cautious in that sense, it is very courageous and daring
in another sense. By making predictions, new hypotheses expose themselves to
the risk of falsification. They risk being wrong! The most searching tests of a

68

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
EXPLORING ALTERNATIVE VIEWS OF SCIENCE

hypothesis centre on those predictions that are not derivable from current theory or
that contradict currently held beliefs. A hypothesis that survives such a test (i.e., is
not falsified) constitutes a significant advance in scientific knowledge. By making
daring predictions, and thereby running a greater risk of falsification, highly specu-
lative theories that survive observational/experimental testing are, in a sense, ‘better’
theories than those that claim less about the world and take fewer risks of being
wrong. Thus, boldness and risk taking are key elements in Popper’s philosophy of
science. For the inductivist, the stimulus for scientific endeavour is the patient,
systematic, meticulous and orderly collection of data, from which generalizations
will be induced. For the falsificationist, the stimulus is the leap of imagination, the
bold speculation about what might be true and the subsequent ruthless discarding
of uncorroborated speculations.

CLARITY AND PRECISION

The demand that hypotheses and theories should be falsifiable requires that they
should be clearly stated and precise. If a theory is vague, such that it is not clear
exactly what is being claimed, then attempts to falsify it can always be interpreted
in a way that is consistent with the theory. Popper (1959) talks disparagingly about
Freudian psychoanalysis as a classic example of vagueness being taken to the limit,
so that everything is compatible with it and nothing can disprove it.3 Precision is
valued because it increases the risk of falsification; lack of precision minimizes
or even avoids altogether the risk of being proved wrong. For example, the pro-
position ‘planets move around the Sun in elliptical orbits’ is more at risk of falsi-
fication than ‘planets orbit the Sun’. Similarly, to borrow the example used by
Chalmers (1999), asserting that the velocity of light is 299.8 × 10 6 m.s-1 is bolder
than claiming the velocity of light is about 300 × 10 6 m.s-1. Both are infinitely
better than saying “light travels very quickly”
For Popper, theories that claim more, in a more general sense, are preferred to
those that claim less. For example, the claim that ‘all planets move around the Sun
in elliptical orbits’ is preferred to ‘Saturn moves around the Sun in an elliptical
orbit’ because it claims much more and, therefore, is at greater risk of being wrong.
Only observations of Saturn can refute the first proposition, whereas observations
of any planet can disprove the second one. This, Popper (1959) argues, results in
science having a quasi-inductive4 trend: it proceeds from theories of high speci-
ficity to more general theories.
Theories of some level of universality are proposed, and deductively tested;
after that, theories of a higher level of universality are proposed, and, in their
turn tested with the help of those of the previous levels of universality, and so
on. The methods of testing are invariably based on deductive inferences from
the higher to the lower level; on the other hand, the levels of universality are
reached, in the order of time, by proceeding from lower to higher levels.
(Popper, 1959, p. 277)

69

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
CHAPTER 4

Thus, Newton proceeded from a theory that talked about planetary bodies in a
solar system attracting each other with a force inversely proportional to the square
of their distance apart to a law of universal gravitation – that is, all bodies attract
each other with a force inversely proportional to the square of their distance apart.
Similarly, Einstein proceeded from a theory of special relativity to a theory of
general relativity. When theories are replaced, says Popper, the new theory has to
explain all that the old one could explain and say something new – that is, it must
explain additional things: “A theory which has been well corroborated can only be
superseded by one of a higher level of universality; that is, by a theory which is
better testable and which, in addition, contains the old, well corroborated theory –
or at least a good approximation to it” (Popper, 1959, p. 276). Thus, Newton’s
physics could explain all that Aristotle’s physics could explain, it was successful
in dealing with the problematic aspects of Aristotelian physics, and it explained
phenomena and events outside the scope of Aristotle’s physics. Similarly, Einstein’s
physics was capable of explaining all that Newton’s physics explained, it accounted
for the anomalies that struck at the heart of Newtonian theory, and it made several
risky predictions, including the proposition that light waves bend when they
approach an intense gravitational field. Surviving an observational test of such a
risky prediction lends enormous support to a theory.
The theories of Kepler and Galileo were unified and superseded by Newton’s
logically stronger and better testable theory, and similarly Fresnel’s and
Faraday’s by Maxwell’s. Newton’s theory, and Maxwell’s, in their turn, were
unified and superseded by Einstein’s. In each such case the progress was
towards a more informative and therefore logically less probable theory:
towards a theory which was more severely testable because it made predic-
tions which, in a purely logical sense, were more easily refutable. (Popper,
1963, p. 220)
The demand that theories become more falsifiable (i.e., have more content) rules
out theoretical modifications designed merely to protect the theory from the threat
of falsification – what Popper calls ad hoc hypotheses.
As regards auxiliary hypotheses we decide to lay down the rule that only
those are acceptable whose introduction does not diminish the degree of
falsifiability or testability of the system in question, but, on the contrary, in-
creases it… If the degree of falsifiability is increased, then introducing the
hypothesis has actually strengthened the theory: the system now rules out
more than it did previously: it prohibits more. We can also put it like this.
The introduction of an auxiliary hypothesis should always be regarded as an
attempt to construct a new system; and this new system should then always
be judged on the issue of whether it would, if adopted, constitute a real advance
in our knowledge of the world. (Popper, 1959, p. 82)
Thus, Popper (1974) explains that in accounting for an anomaly in the motion
of Uranus, as predicted by Newton’s gravitational theory, a permitted auxiliary
hypothesis was the existence of a previously unknown planet. This modification

70

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
EXPLORING ALTERNATIVE VIEWS OF SCIENCE

was independently testable: the position of the planet was calculated and the planet
(Neptune) was discovered optically. Not permitted in these circumstances was the
ad hoc hypothesis that Uranus is unique in the solar system in not behaving as
Newtonian theory predicts. Events such as Galle’s discovery of Neptune, Hertz’s
discovery of radio waves, Eddington’s confirmation that light bends in the vicinity
of the Sun, and Powell’s observations of the first Yukawa mesons are significant
in science because, in Popper’s terminology, they corroborate a bold hypothesis
and so constitute powerful and persuasive evidence that the new theory is an
improvement on the old: “All these discoveries represent corroborations by severe
tests – by predictions which were highly improbable in the light of our previous
knowledge (previous to the theory which was tested and corroborated)” (Popper,
1963, p. 220). The historical significance of such corroborative tests contrast sharply
with the a-historical nature of confirmatory evidence in the inductivist approach,
which would seem to lend increased support for a theory regardless of when the
evidence is generated. Popper also regards the falsification of cautious hypothe-
ses as significant. Because they enable us to reject well-established and seemingly
unproblematic knowledge they are a major stimulus to creative theory building.

THE PROBLEMS OF FALSIFICATIONISM

Elegant and appealing as Popper’s arguments may be, they are not free of diffi-
culties. A major problem is his insistence that theory acceptance is tentative, while
theory rejection is decisive. As argued in chapter 3, observation statements – the
evidence on which theories are to be judged – are both unreliable (fallible) and
theory-dependent. Because observation statements are theory-dependent, a theory
may be protected from falsification until the appearance of a new theory capable
of looking at the observational evidence in a new way, or able to produce entirely
new evidence – a point to be discussed later in the chapter. Because observation
statements are fallible, and their acceptance only tentative and open to revision,
falsification is logically suspect. When theory and observation are in conflict, there
is nothing in the logic of the situation that compels us to reject the theory. If
a theory is falsified only when the observation statement that contradicts it is
accepted, theories cannot be conclusively falsified because the observation state-
ments that form the basis of the falsification may themselves prove to be false,
mistaken or misleading in the light of later developments. In short, if we cannot
rely on observations, we cannot use them in a decisive way to reject theories, just
as we cannot use them as a certain basis for accepting a theory.
While the observation of a single black swan does falsify the proposition that
all swans are white, it would be a serious mistake to regard scientific theories as
similar in scope and form to this simple generalization and, therefore, as equally
vulnerable to decisive falsification. Scientific theories are not merely simple state-
ments or generalizations; rather, they are complex explanatory systems comprising
elaborate and complex conceptual relationships. In consequence, testing and dis-
proving a theory is not quite as straightforward as Popperian falsificationism seems
to imply. We should certainly not project a view in school science that theories

71

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
CHAPTER 4

stand or fall on the basis of a single and simple “Yes or No” test. It is precisely
because theories are complex structures, supported by a complex array of other
theories (for example, theories of perception and theories underpinning scientific
instrumentation), that an apparently falsifying observation can be deflected away
from what Imre Lakatos calls the hard core of the theory. Lakatos (1974) introduces
the idea of a “protective belt” around a theory’s central principles (hard core) by
means of a story about “an imaginary case of planetary misbehaviour”. Because of
its insight and humour it is worth quoting at length.
A physicist of the pre-Einsteinian era takes Newton’s mechanics and his law
of gravitation… and calculates, with their help, the path of a newly dis-
covered small planet, p. But the planet deviates from the calculated path.
Does our Newtonian physicist consider that the deviation was forbidden by
Newton’s theory and therefore that, once established, it refutes the theory?
No. He suggests that there must be a hitherto unknown planet p’, which
perturbs the path of p. He calculates the mass, orbit, etc. of this hypothetical
planet and then asks an experimental astronomer to test his hypothesis. The
planet p’ is so small that even the biggest available telescopes cannot possibly
observe it: the experimental astronomer applies for a research grant to build
yet a bigger one. In three years time, the new telescope is ready. Were the
unknown planet p’ to be discovered, it would be hailed as a new victory of
Newtonian science. But it is not. Does our scientist abandon Newton’s theory
and his idea of a perturbing planet? No. He suggests that a cloud of cosmic
dust hides the planet from us. He calculates the location and properties of this
cloud and asks for a research grant to send up a satellite to test his calcu-
lations. Were the satellite’s instruments (possibly new ones, based on a little-
tested theory) to record the existence of the conjectural cloud, the result
would be hailed as an outstanding victory for Newtonian science. But the
cloud is not found. Does our scientist abandon Newton’s theory, together
with the idea of the perturbing planet and the idea of the cloud which hides
it? No. He suggests that there is some magnetic field in that region of the
universe which disturbed the instruments of the satellite. A new satellite is
sent up. Were the magnetic field to be found, Newtonians would celebrate a
sensational victory. But it is not. Is this regarded as a refutation of Newtonian
science? No. Either yet another ingenious auxiliary hypothesis is proposed
or… (pp. 100-101)
Embedded within this story are the key elements in the Lakatosian notion
of science as a constellation of “research programmes”, complex conceptual and
methodological structures which guide the work of scientists by indicating poten-
tially fruitful lines of development (the “positive heuristic”) and specifying what
cannot be done (the “negative heuristic”). The negative heuristic of a Lakatosian
research programme involves the stipulation that the basic assumptions underlying
the programme (the “hard core”) cannot be modified or rejected. Any mismatch
between theory and and observation is dealt with by modification of the “protective
belt” of auxiliary hypotheses surrounding the hard core, as in the story above.

72

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
EXPLORING ALTERNATIVE VIEWS OF SCIENCE

The positive heuristic of the programme comprises rough guidelines on future


development.
The positive heuristic consists of a partially articulated set of suggestions
or hints on how to change, develop the ‘refutable variants’ of the research-
programme, how to modify, sophisticate, the ‘refutable’ protective belt.
(Lakatos, 1978, p. 50)
Early work on a research programme often takes place in opposition to
apparently falsifying observations. Time has to be allowed for development of
the programme – a suitably robust and sophisticated protective belt must be con-
structed before the theory is subjected to rigorous testing. Concepts need to be
refined and conceptual relationships clearly established before a theory can be
subjected to critical experiment-based scrutiny. Once the programme is sufficiently
well developed to permit testing, it is confirmations rather than falsifications that
are considered significant.
The basic unit of appraisal must not be an isolated theory or conjunction of
theories but rather a ‘research programme’, with a conventionally accepted…
‘hard core’ and with a ‘positive heuristic’ which defines problems, outlines the
construction of a belt of auxiliary hypotheses, foresees anomalies and turns
them victoriously into examples, all according to a preconceived plan. (p. 99)
A research programme is judged to be “progressive” or “degenerating”,
depending on whether or not its theoretical growth anticipates its empirical growth
– that is, whether or not it keeps generating new facts, leads to the discovery of
new phenomena and predicts future events. Notable events of this kind were
Newton’s prediction of the return of Halley’s comet in 1758 and the discovery
of the two elements we now know as gallium and selenium on the basis of
Mendeleev’s Periodic Table.
A series of theories is theoretically progressive… if each new theory has
some excess empirical content over its predecessor, that is, if it predicts some
novel, hitherto unexpected fact. Let us say that a theoretically progressive
series of theories is also empirically progressive… if some of this excess
empirical content is also corroborated, that is, if each new theory leads to the
actual discovery of some new fact… Progress is measured by the degree… to
which the series of theories leads to the discovery of novel facts. We regard a
theory in the series ‘falsified’ when it is superseded by a theory with higher
corroborated content. (Lakatos, 1978, p. 33)
Popper’s criterion for a satisfactory theory is agreement with the observed facts;
the criterion for a satisfactory series of theories (research programme), according to
Lakatos, is that it should generate new facts and provide clear guidance for research.
A research programme is said to be ‘degenerating’ (and should be abandoned) when
it loses its coherence and/or when it no longer leads to confirmed novel pre-
dictions.

73

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
CHAPTER 4

INSURMOUNTABLE PROBLEMS

The Lakatosian modifications of Popperian theory seem to accord better with


historical evidence. If strict falsificationist methodology had been followed, some
of our best theories would have been killed off very early in their history. Imagine
the situation at the time of the Copernican revolution, for example. If the Earth
revolves around the Sun, as Copernicus suggested, there should be a variation in
the apparent size of the other planets during the year it takes to complete the
orbit. Such a variation is not observed. Therefore, Copernican theory appears to be
falsified. Galileo took a different view, arguing that the planets are much further
away than we believe and, therefore, our eyes cannot detect the change in size.
What is needed is a much better way of observing. Eventually, with telescopic
observation, Galileo was proved correct: there is an apparent variation in the size
of the other planets during the Earth’s rotation of the Sun. Galileo also had to deal
with the observation that objects dropped from the top of a tower land at the base
of the tower and not some distance away, as opponents of Copernican theory
argued they would if the Earth was rotating. He did so by introducing the idea of
relative motion, a move that challenged well-established Aristotelian views about
human perception. What Galileo did was to defend the theory against the evidence.
It was his insistence on adhering to Copernican theory in the face of apparently
falsifying evidence that led to the development of new observational methods and
new subsidiary theory (concerning the nature of human perception), and paved the
way for the eventual universal acceptance of heliocentric theory.5 As Hall (1974)
puts it:
Galileo is above all aware that the senses must be educated and assisted to
perceive realities. Thus one could know the true nature of the moon without
the telescope: that instrument simply makes reality easier to discover. (p. 72)
Similarly, early observations of the Moon’s orbit apparently falsified Newton’s
law of gravitation and it took the best part of 50 years to explain away this anomaly.
Later on, the theory survived even though it consistently failed to account fully for
the observed orbit of Mercury. Einstein’s physics finally achieved that. The reality
is that scientists don’t reject theories because of a few inconvenient facts; they
know that, given time, the facts can sometimes be changed – by reinterpreting the
observational data or obtaining better observational data. Even Popper (1974), the
principal advocate of falsificationism, says that “if we give in to criticism too
easily, we shall never find out where the real power of our theories lies” (p. 55). At
the individual level, scientists are not immune to selecting data that supports their
theoretical position and rejecting data that doesn’t, as Holton’s (1978) analysis of
Millikan’s laboratory notebooks indicates. Holton posits the notion of suspension
of disbelief to describe how Millikan withheld “final judgements concerning the
validity of apparent falsifications of a promising hypothesis” (p. 71). Far from
condemning such action, Holton regards it as an essential part of theory building:
“The chief gain was the avoidance of costly interruptions and delays that would
have been required to pin down the exact causes of discrepant observations”
(Holton, 1986, p. 12).

74

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
EXPLORING ALTERNATIVE VIEWS OF SCIENCE

One question that Lakatos does not answer is: “How long do we wait before we
label a research programme as degenerating?” It is always possible that some new
modification of the protective belt will breathe new life into an old theory. If you
are permitted to wait, asks Feyerabend (1975), why not wait a little longer?
If it is unwise to reject faulty theories the moment they are born because they
might grow and improve, then it is also unwise to reject research programmes
on a downward trend because they might recover and might attain unforeseen
splendour (the butterfly emerges when the caterpillar has reached the lowest
stage of degeneration). Hence, one cannot rationally criticise a scientist who
sticks to a degenerating programme and there is no rational way of showing
that his actions are unreasonable. (p. 185)
Feyerabend argues that Lakatos fails to provide any clear criteria for choosing
or rejecting a particular programme and, therefore, that his model of science,
whilst useful for retrospective evaluation of research programmes, offers practising
scientists no guidance on how to proceed. It is, Feyerabend asserts, no more than a
“verbal ornament… a memorial to happier times when it was still thought possible
to run a complex and often catastrophic business like science by following a
few simple and ‘rational’ rules” (Feyerabend, 1970, p. 215). Even the criterion of
successful prediction is not always sufficient for a theory to be accepted by the
community. As Brush (1990) has noted, the astrophysicist Hannes Alfven has a
remarkable record of successful predictions concerning plasma phenomena
(including magnetohydrodynamic waves, now known as Alfven waves, field aligned
currents in the Earth’s upper atmosphere and critical ionization energy phenomena,
including Uranian rings), yet his theoretical explanations of those predictions have
not been widely accepted.
As argued in the previous chapter, it is not unreasonable to insist that a satis-
factory account of science should do two things. It should account for the actual
course of science and it should provide guidance for aspiring scientists on how to
proceed. The falsificationist view of science does not meet the first of these req-
uirements: it is clearly not what scientists do, however elegant and appealing the
model may seem. There is no choice but to abandon Popperian views about the
nature of scientific method, despite its prominence in some science textbooks (those
that don’t promote inductivism generally promote hypothetico-deductivism) and
despite its exalted status as the so-called ‘scientific approach’ to educational re-
search, with its emphasis on the null hypothesis.

A REVOLUTIONARY VIEW

Despite my suggestion that Popper’s account of science should be rejected, it does


represent a major step forward from the traditional inductivist view. For me, the
most significant aspect of the Popperian model of science is that science is driven
by people and their creative imaginations, not by some cold and clinical method
to which people have to subordinate themselves. In Popper’s world of science,
people have priority over method: people determine what science to do and how

75

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
CHAPTER 4

to do it, rather than science and its method requiring that people divest them-
selves of their essential humanity at the laboratory door and assume the role of
the “depersonalized seeker after truth” (Cawthron & Rowell, 1978). Even more
interesting in these respects is Thomas Kuhn’s model of science, as outlined in his
1962/1970 book The Structure of Scientific Revolutions. According to Kuhn, science
does not proceed in an orderly, systematic and continuous way, but through a series
of revolutions and periods of consolidation (what Kuhn calls normal science).
In the beginning, in the early history of a science,6 there is no agreement on
how best to proceed. Each individual is busy defining and re-defining the field,
establishing priorities, making statements about correct procedures, deciding what
should count as evidence, and so on. The disorganized and diverse activities that
characterize pre-science, as Kuhn calls it, become structured and directed when the
community of practitioners reaches agreement on certain theoretical and methodo-
logical issues – that is, when the disciplinary matrix (the framework of theory and
method or paradigm, to use Kuhn’s term) becomes established and accepted. Kuhn
illustrates his argument by reference to the situation prior to the publication of
Isaac Newton’s Opticks, in 1704.
No period between remote antiquity and the end of the seventeenth century
exhibited a single generally accepted view about the nature of light. Instead
there were a number of competing schools and sub-schools, most of them
espousing one variant or another of Epicurean, Aristotelian, or Platonic theory…
Being able to take no common body of belief for granted, each writer on
physical opticks felt forced to build his field anew from its foundations.
(Kuhn, 1970, p. 12)
Once a particular paradigm has been established, scientists engage in normal
science: they work within the paradigm, assuming that its basic premises are valid;
they follow its rules and procedures; they use its concepts and theories to explore,
develop and extend the paradigm, widen its scope, solve its internal puzzles and
problems (both theoretical and procedural), formulate quantitative relationships
among concepts, and so on. In Kuhn’s (1977) words, normal science “aims to elu-
cidate the scientific tradition in which (the scientist) was raised rather than to
change it… The puzzles on which he concentrates are just those which he believes
can be both stated and solved within the existing scientific tradition” (p. 234).
In engaging in the paradigm articulation of normal science scientists may
encounter problems that the paradigm cannot solve or they may generate some
seemingly falsifying evidence. Initially, this is not serious. Indeed, it is to be expec-
ted: the paradigm is a human creation and cannot be expected to be perfect, to be
capable of explaining and accounting for everything within its domain (sphere of
interest and concern). It needs time to develop: concepts need to be refined, con-
ceptual relationships established and internal inconsistencies eliminated. Eventually
the problems will be satisfactorily solved. What scientists look for, says Kuhn, is
reasonable agreement between theory and observation, with ‘reasonable’ being
judged in relation to the specific circumstances. This is a significantly different

76

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
EXPLORING ALTERNATIVE VIEWS OF SCIENCE

position from the extreme falsificationist position that contrary evidence instantly
refutes a theory.
However, if a large number of unsolved problems accumulates, if the problems
are theoretically significant, in that they strike at the hard core of the theory and
consistently resist solution, or if they are socially and/or economically significant,7
a crisis develops. When a paradigm cannot cope with its internal problems, scien-
tists begin to lose confidence in it. When a paradigm is in crisis and dissatisfaction
is widespread, the time is ripe for revolution. If an individual scientist, or a group
of scientists, proffers a new set of concepts, a new theory or a new procedure that
is able to solve some of the major problems, then others will be encouraged to try
out the new ideas on their problems. If success is achieved again, still others will
be encouraged to adopt the new thinking. Eventually, if the new approach provides
a sufficiently plausible solution to the problems of the old paradigm, a revolution
will have occurred and allegiance will have switched from the old paradigm to the
new. This new paradigm now becomes the basis for a new round of normal
science, until, in due time, a new crisis develops and further revolution occurs.
While normal science is a somewhat conservative and rule-bound phase, it is the
driving force for progress because it generates the problems and anomalies that
precipitate the next revolution.
Normal science does not aim at novelties of fact or theory and, when success-
ful, finds none. New and unsuspected phenomena are, however, repeatedly
uncovered by scientific research, and radical new theories have again and
again been invented by scientists. History even suggests that the scientific
enterprise has developed a uniquely powerful technique for producing sur-
prises of this sort. (Kuhn, 1970, p. 52)
After a scientific revolution there is much work to be done. All existing data not
rendered irrelevant by the new paradigm have to be re-interpreted within the new
theoretical framework. Moreover, scientific instruments and research procedures
have to be reviewed and appropriately modified. And, of course, new textbooks
need to be written to ensure that the next generation of scientists is ‘properly’
educated – that is, in the new way of thinking and acting. Sometimes this re-writing
of science is so extensive and persuasive that all trace of preceding theoretical ideas
are swept away – a phenomenon that Kuhn calls “the invisibility of revolutions”.
[the textbooks] have to be rewritten in whole or in part whenever the langu-
age, problem-structure, or standards of normal science change. In short, they
have to be rewritten in the aftermath of each scientific revolution, and, once
rewritten, they inevitably disguise not only the role but the very existence of
the revolutions that produced them. (p. 137)

KEY FEATURES OF KUHN’S VIEW OF SCIENCE

There are four features of Kuhn’s view of science that are worthy of close atten-
tion. First, his point that scientific progress is discontinuous. For Kuhn, paradigms

77

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
CHAPTER 4

do not change in piecemeal fashion; rather, there is a wholesale shift to a new way
of thinking, bringing with it a new set of procedures and a new set of ‘facts’ about
the world. Kuhn is referring to the revolutionary nature of major theoretical
breakthroughs by scientific giants like Newton, Einstein and Darwin. Of course, as
argued in chapter 3, science is cumulative, and unless scientists assume the validity
of science already done, further progress would be impossible. However, too un-
critical an acceptance of science already done curtails further progress. Progress
emanates from doubt.
The second point of interest focuses on Kuhn’s tolerance of anomaly. Provided
they are not too serious, and strike at the very core of the paradigm, anomalies are
accepted, even expected.
During the sixty years after Newton’s original computation, the predic-
ted motion of the moon’s perigree remained only half of that observed. As
Europe’s best mathematical physicists continued to wrestle unsuccessfully
with the well-known discrepancy, there were occasional proposals for a
modification of Newton’s inverse square law. But no one took these proposals
very seriously, and in practice this patience with a major anomaly proved
justified. Clairaut in 1750 was able to show that only the mathematics of the
application had been wrong and that Newtonian theory could stand as before.
(Kuhn, 1962, p. 81)
Thus, refutation is much less significant for Kuhn than for Popper: “The
scientist who pauses to examine every anomaly he notes will never get significant
work done” (Kuhn, 1962, p. 82). Single discrepancies are not sufficient to falsify a
theory and bring about a revolution. No one seriously questioned Newtonian
physics because of discrepancies concerning the orbit of Mercury or the speed of
sound. The fact that that there are lots of observations inconsistent with Einsteinian
theory has not, so far, led to its overthrow. Paradigms are not discarded until
there is an abundance of acceptable and significant falsifying evidence and, more
importantly, an alternative theory.8
Once it has achieved the status of a paradigm, a scientific theory is
declared invalid only if an alternative candidate is available to take its
place. (Kuhn, 1970, p. 77)
A paradigm is not usually rejected on the basis of a comparison of its predicted
consequences with empirical evidence, as Popper insists. More often, it is rejected
as a consequence of a 3-way comparison of the old theory, the new theory and
observational evidence – though even this may be more difficult than it seems at
first glance. Given the arguments of chapter 2, the nature of the evidence collected
is determined by the theory, and so may be significantly different for the old and
the new paradigm – a matter to be discussed below and in chapter 5.
Kuhn’s views accord much better with historical evidence than do Popper’s.9
If a strict Popperian methodology was followed, many promising theories would
have been ‘killed off’ very early in their history, thus depriving the scientific com-
munity of many potentially interesting and productive ideas. If theories were not
remarkably resilient in the face of apparently falsifying data, scientists would be

78

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
EXPLORING ALTERNATIVE VIEWS OF SCIENCE

continually starting again and progress would be impossible. Theories are resilient
because they can accommodate ‘counter observations’ in many ways, including: (i)
challenging the reliability of the observations and/or the methods and instruments
employed; (ii) deflecting criticism to a subsidiary (and, therefore, less important)
part of the theory; (iii) postulating a ‘poorly understood’ complicating factor, to
which ‘attention will be directed in due course’. Of course, this resilience some-
times works to the detriment of scientific progress, as in Alfred Wegener’s (1915)
unsuccessful attempt to overthrow the reigning paradigm of ‘permanentism’ in
geology in favour of his notion of ‘continential drift’ – a revolution that was
delayed by more than half a century. (Frankel, 1979).
A third significant element of Kuhn’s view of science is his assertion that rival
paradigms look at the world in different ways. As argued in chapter 3, possession
of new knowledge enables us to make new observations. In other words, what you
see depends on what you know. Further, what there is to be seen depends on what
you know. Because they involve different concepts and ideas, rival paradigms
direct attention to different things, and in different ways. They have different
priorities and focus on different issues, problems and questions. Consequently,
comparison of rival paradigms is difficult, if not impossible. There is no paradigm-
independent language and there are no paradigm independent concepts, so no
paradigm-independent observations can be made or paradigm-independent experi-
ments performed. Data, in the usual meaning of the term, cannot establish the
superiority of one paradigm over another because data are perceived either through
the lens of one paradigm or the other. In short, there is no common basis on which
rival paradigms can be compared. In Kuhn’s terminology, rival paradigms are
incommensurable. In retrospect, we see Lavoisier’s isolation of oxygen in the late
18th Century as a decisive event in the establishment of his theory of combustion.
However, Priestley regarded this gas as ‘dephlogisticated air’ and its isolation as
evidence in favour of phlogiston theory. In Priestley’s world, phlogiston is released
into the air during combustion; in Lavoisier’s world, oxygen is taken out of the air
during combustion. There is nothing in the oxygen theory that corresponds directly
with phlogiston, nor is there anything in the phlogiston theory that corresponds
directly to oxygen. Yet both theories explain the observational data.10
Even when a new paradigm utilizes concepts from an old one, it does so in a
new way. Compare, for example, the concepts of mass, time, space and energy in
the Newtonian and Einsteinian paradigms. In Einstein’s special theory of relativity,
the mass of a body depends on the observer’s frame of reference and, moreover,
mass can be converted into energy; in the Newtonian paradigm, the mass of a body
is fixed and is constant for all observers, regardless of their frame of reference.
Although the term mass is used in both paradigms, the respective concepts are
incommensurable. Indeed, in the Newtonian paradigm, the concept ‘relative mass’
would be absurdly meaningless.
Since new paradigms are born from old ones, they ordinarily incorporate
much of the vocabulary and apparatus, both conceptual and manipulative,
that the traditional paradigm had previously employed. But they seldom
employ these borrowed elements in quite the traditional way. Within the new

79

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
CHAPTER 4

paradigm, old terms, concepts, and experiments fall into new relationships
one with the other. (Kuhn, 1970, p. 149)
In short, scientists before and after a paradigm change are not talking about the
same theoretical entities, even when they are using the same words.
The proponents of competing competing paradigms practice their trades in
different worlds. One contains constrained bodies that fall slowly, the other
pendulums that repeat their motions again and again. In one, solutions are
compounds, in the other mixtures. One is embedded in a flat, the other in a
curved, matrix of space. Practicing in different worlds, the two groups of
scientists see different things when they look from the same point in the
same direction… Both are looking at the world, and what they look at has not
changed. But in some areas they see different things, and they see them in
different relations one to the other. That is why a law that cannot even be
demonstrated to one group of scientists may occasionally seem intuitively
obvious to another (p. 150)
Thus, the idea of incommensurability implies that after a revolution scientists
have a new way of looking at things and new problems to work on (a new kind of
normal science). Thus, old problems are not so much solved as simply forgotten or
regarded as irrelevant. Since ancient times, theories in astronomy had struggled to
account for planetary retrogression, but in the Copernican system it ceases to be a
problem because the observed motions of the planets are a consequence of the
Earth and other planets orbiting the Sun.
If rival theories are incommensurable, there can be no crucial experiment to
decide between them. Such an experiment would require the two theories to make
mutually exclusive predictions about the same events. Because competing theories
address the world in different ways, often using different concepts, they make
different predictions about observable phenomena. Indeed, what is observable in
terms of one theory may not be observable in terms of the other. Suppose, for
example, that theory A says that light is a wave motion (electromagnetic radiation)
and theory B says it is particulate (high energy photons). As a consequence of its
behaviour as a wave motion, light will have particular observable properties, such
as interference patterns, which scientists can seek to confirm or refute through
observational evidence provided by experiment. Theory B also has particular
observable consequences for the behaviour of light, such as the photoelectric
effect, which experiments can seek to confirm or refute. Under one set of experi-
mental circumstances, light will manifest particle-like properties; under another
(experimentally incompatible) set of circumstances it will manifest wave-like pro-
perties. However, while experiments designed on the assumption that light is a
wave motion may provide support for this theory, they will not falsify the theory
that it is particulate. Nor will experiments designed on the assumption that light
comprises high energy particles falsify the theory that it is a wave motion, though
they may provide support for the particle theory. Experiments designed on the
assumption that it is neither particulate nor wave-like are not possible unless there
is a third theory that says something about the nature of light in terms other than

80

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
EXPLORING ALTERNATIVE VIEWS OF SCIENCE

particles and waves. But these experiments could not falsify either theory A or
theory B; they could only lead to judgements about the third theory. In other words,
experiments can only test theories within and against their own conceptual frame-
work. They cannot tell us anything about theories that have a different conceptual
structure. All that can be done in a situation of paradigm conflict is to evaluate a
theory ‘in terms of its own terms’, supplemented by criteria additional to the cri-
terion of simple empirical validity. Kuhn (1970) expresses this position as follows:
The choice is not and cannot be determined merely by the evaluative proce-
dures characteristic of normal science, for these depend in part upon a parti-
cular paradigm, and that paradigm is at issue. When paradigms enter, as they
must, into a debate about paradigm choice, their role is necessarily circular.
Each group uses its own paradigm to argue in that paradigm’s defence… the
status of the circular argument is only that of persuasion. It cannot be made
logically compelling for those who refuse to step into the circle. (p. 94)
This is not what we tell students in school science, nor, I suspect, in university
science courses. Chapter 5 discusses these matters at greater length.
The fourth significant element in Kuhn’s account of science is its recognition
of the importance of social factors at all stages of the scientific endeavour.
Although Popper regards the generation of hypotheses as a creative act that is not
susceptible to rational analysis, he asserts that hypotheses are tested and accepted/
rejected by orderly and logical procedures based on experiment and observation.11
In contrast, Kuhn insists that there is no purely logical argument to show the
superiority of one paradigm over another. Individual scientists may be persuaded to
adopt a particular theory for a variety of reasons, grounded in their particular
interests, personal and professional goals, values, and day-to-day concerns and
priorities. Thus, it is consensus within the community of practitioners that is the
mechanism for acceptance. Of particular significance is Kuhn’s admission that
consensus is influenced by all manner of psychological, social, political and eco-
nomic considerations. In other words, science is not entirely logical and rational, in
the sense usually employed in science textbooks. Rather, it is a value-laden and
socioculturally located enterprise – a situation to be discussed at length in chapter 7.
It is the consensus view of science embedded in Kuhn’s thesis that offends those
critics who seek objective criteria for choosing between rival theories because, they
argue, it leads to a relativist position – one theory is not superior to another, just
different.12 If Kuhn’s arguments are taken at face value, facts about the world are
paradigm-dependent and change when there is a scientific revolution. Thus, it
makes no sense to ask whether a given theory corresponds to the facts ‘as they
really are’ nor, therefore, to ask whether a theory is true. Truth itself is relative to a
particular paradigm: what is true with respect to theory T1 is not true with respect
to theory T2, and vice versa. In contrast, Popper (1972) asserts that one theory
is closer to the truth than another, whether a particular individual or group of
individuals thinks so or not. As science progresses theories approach the truth or, to
use Popper’s (1972) language, their verisimilitude increases. Furthermore, Popper
seems to regard this progress as guaranteed, provided that there is a sufficient

81

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
CHAPTER 4

number of individuals properly trained in scientific method and possessing the


right attitude of “critical rationalism”. Paul Feyerabend (1975), on the other hand,
denies that there is an objective scientific method. All attempts to characterize
it have failed, he observes, and no one model of science stands up to historical
scrutiny.
The idea of a method that contains firm, unchanging and absolutely binding
principles for conducting the business of science meets considerable diffi-
culty when confronted with the results of historical research. We find then,
that there is not a single rule, however plausible, and however firmly groun-
ded in epistemology, that is not violated at some time or another. (p. 23)
The idea of a fixed method of science rests on too naïve a view of what is a
complex and uncertain enterprise, argues Feyerabend, and the only principle that is
always applicable is the principle Anything Goes. The extent to which this view is
appropriate for school science will form part of the discussion in chapter 5 and will
be re-visited in chapter 9.

ENDNOTES
1
Popper (1972) includes two additional steps in the procedure: (i) before the conclusions are tested,
they are compared among themselves and appraised for internal consistency (there must be no
inconsistencies or mutual contradictions); (ii) the logical form of the proposition is investigated for
evidence of conceptual ambiguity, vagueness, circularity and lack of clarity.
2
Falsification is achieved if we accept that it is a genuine black swan, rather than a white swan that
appears black because it is living in a polluted environment. Disconfirming observations have to be
acceptable – a point of importance in the earlier discussion of the conditions for inductive inferences
(see chapter 3). In Popper’s (1959) words, “a few stray basic statements contradicting a theory will
hardly induce us to reject it as falsified. We shall take it as falsified only if we discover a rep-
roducible effect which refutes the theory” (p. 86, emphasis added).
3
Chalmers (1999) remarks that “there are plenty of social, psychological and religious theories that
give rise to the suspicion that in their concern to explain everything they explain nothing… Theorists
operating in this way are guilty of the fortune-teller’s evasion and are subject to the falsificationist’s
criticism. If a theory is to have informative content, it must run the risk of being falsified” (p. 64).
Popper illustrates his argument with the following example: A man pushes a child into a river with
the intention of murdering him/her, while another man jumps into the river and sacrifices his life in
order to save the child. Freudians explain the first man’s behaviour by positing that he suffered from
repression and the second man’s behaviour by saying that he achieved sublimation. Adlerians explain
the episode by saying that both men suffered from feelings of inferiority; the first man needed to
prove to himself that he could commit the crime and the second man needed to prove to himself that
he was brave enough to rescue the child at whatever personal cost. By employing concepts such as
repression, sublimation and unconscious desires, psychoanalysis is made compatible with all human
behaviour. Because they are unfalsifiable, such theories are not science (according to Popper), though
they may have value as something other than science.
4
This is not induction, which Popper goes to considerable lengths to reject, but an inductive direction
in theory building – a movement from the more particular to the more general.
5
It was a long time before the heliocentric theory of the solar system could account for all the
observational data. To do so, the theory had to adopt Keplerian elliptical orbits (rather than circular
ones), planets of different masses in mutual attraction, rotating planets, planets with satellites (moons),
and an additional planet (Neptune).

82

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
EXPLORING ALTERNATIVE VIEWS OF SCIENCE

6
Note that Kuhn discusses the development of sciences, not science as a whole. For Kuhn, there are
significant differences among the sciences.
7
The crisis that precipitated the Copernican revolution was the urgent need for an accurate calendar to
predict the holy days of the Christian church. However, Copernicus presented his ideas in instru-
mental form (see chapter 6), lest he offend against the teachings of the established church (Dreyer,
1953).
8
In science, there is rarely a period in which there is no prevailing paradigm, though there may be
occasions when no consensus has been reached about which of the competing paradigms should be
accepted.
9
Indeed, it was Kuhn’s declared intent to provide a realistic and historically accurate account of how
science develops, in contrast to previous accounts that, he says, are more like reading a tourist guide-
book about a particular country than experiencing life there.
10
It was also observed that when metals are heated to form what was then known as a calx, the resul-
ting calx weighed more than the original metal. The combustion in oxygen theory explains this per-
fectly well: the metal combines with oxygen to form a metal oxide, which necessarily weighs more
than the uncombined metal. Using phlogiston theory, the increase in weight can only be explained if
phlogiston has negative weight – an absurdity that eventually proved the final nail in the coffin for
phlogiston theory. Allchin (1997) makes a case for teaching about the ‘paradigm war’ between suppor-
ters of phlogiston theory and supporters of combustion in oxygen as a way of integrating philosophy
of science, sociology of science and history of science into the science curriculum.
11
In drawing what he considers to be a crucial distinction between the context of discovery and the
context of justification, Popper says that theory evaluation should take no account of the personal,
sociocultural and historical circumstances of its generation. Rather, all efforts should be concen-
trated on the chain of argument and the quality of the evidence: “The act of conceiving or inventing
a theory seems to me neither to call for logical analysis nor to be susceptible to it… the question
of how it happens that a new idea occurs… may be of interest to empirical psychology; but it is
irrelevant to the logical analysis of scientific knowledge” (Popper, 1959, p. 27).
12
Lakatos (1968) claimed that Kuhn’s views made theory change in science a consequence of what
he disparagingly labelled mob psychology: “According to Kuhn scientific change – from one
‘paradigm’ to another – is a mysterious conversion which is not and cannot be governed by the rules
of reason: it falls totally within the realm of (social) psychology of discovery” (p. 151).

83

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
Derek Hodson - 978-90-8790-507-1
Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
CHAPTER 5

SCIENTIFIC INQUIRY, EXPERIMENT AND THEORY


What Should We Tell Our Students?

Contemporary school science curricula are virtually unanimous in the views they
promote about the relationship between scientific knowledge and scientific method
– specifically, between theory and experiment. So, too, are school science textbooks,
most of which assert that the validity of scientific knowledge is judged solely by its
agreement with observable and experimentally acquired evidence. Indeed, the very
rationality and objectivity of science are held to be guaranteed by insisting that
theories are subjected to experimental testing by other scientists and by the assump-
tion that this testing is decisive. Thus, almost all textbooks and curriculum documents
invest enormous faith in the capacity of observation and experiment to provide
reliable data for making unequivocal decisions about the validity of theories.
There is no doubt that science is at its most powerful and most effective when it
is able to control and manipulate phenomena and events, as in laboratory experi-
ments. Interestingly, many of the events observed in experimental inquiries do not
occur in the natural world, or they are so changed by the conditions of the experiment
that they are, in essence, different events. In such circumstances, the experimental
approach is able to obtain information that is much more detailed and precise than
that arising from passive observation of uncontrived events. However, experiments
may not always be sufficient, in themselves, to provide a reliable and valid basis
for theory building about the natural world. Nor, it should be noted, are experiments
always necessary or desirable. Many fields of scientific endeavour deal with events
that are remote and inaccessible in time and space, and so make little or no use of
experiments. In these cases, theoretical conjectures have to be confirmed or refuted
by uncontrived observations. In some areas of science, experimentation may be
possible but is ruled inadmissible on ethical grounds or for reasons of safety,
difficulty or cost. In these cases, correlational studies in naturalistic settings may
play a significant role.1 Moreover, there may be field settings that are unsuitable
for experimental inquiry because they are too complex or too fragile and uncertain.
For example, an experimental approach might so distort the natural setting that it
no longer represents natural behaviour, as Bowen and Roth (2002) describe in their
discussion of observational work with lizards.
The power that results from close control is also the major weakness of the
experimental approach and a potential trap for the unwary. The arguments deve-
loped in chapters 3 and 4 concerning the theory-laden nature of scientific obser-
vation and the incommensurability of rival paradigms lead to the conclusion that
experiments can only be envisaged, designed and conducted within a particular
theoretical matrix, which governs scientists’ perceptions of the problem, deter-
mines the experimental design, influences the interpretation of results, and so on.
Theories determine which experiments are regarded as legitimate and how they are
85

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
CHAPTER 5

to be conducted. For example, in gathering data to test an hypothesis, the form of


the hypothesis and the nature and method of data collection are dictated by the very
theory that is under test. In other words, theory-independent experiments are impos-
sible. If theories are incommensurable, as Kuhn (1970) claims, there can be no
crucial experiment to decide between them (see discussion in chapter 4). In such
circumstances, the best we can do is to evaluate each theory on its own terms.
These points, made at length in Hodson (1988), serve to remind us that in
seeking to give students an understanding of the nature of scientific inquiry we
must be careful not to reinforce the many falsehoods about the role and status of
experiments perpetrated by school science textbooks. Sometimes science teachers
promote the belief that everything that falls within the province of science, and that
students learn about in school science lessons, is susceptible to direct experimental
study. They omit to mention that many theoretical advances in science did not
result from experimentation, and that many theories were developed and substan-
tiated by indirect means, such as consistency with other theoretical systems, use of
‘thought experiments’ and correlational studies, rather than by experimentally-based
observation (Hacking, 1983). Misconceptions about the nature and purpose of experi-
ments are also promoted by the popular media, especially prominent in the world of
advertising with its insistence that washing powder X is “proven by experiment to
wash whiter than any other product”.
It is not uncommon for science textbooks to assert that an hypothesis can be
rejected – and, by inference, another accepted – on the evidence provided by a
single experimental test. Many suggest that this is the only role for experiments.
It is also commonplace for teachers to insist on a clear demarcation between
hypothesis generation and hypothesis testing. This kind of naïve interpretation of
Popperian falsificationism carries with it the assumption that theory-independent
evidence is obtainable and that unambiguous testing is possible. In practice, scien-
tific experimentation is far from being a simple, straightforward matter, and science
education that portrays it as such is grossly misleading. If ‘ordinary’, day-to-day
and relatively passive observation of phenomena and events is theory dependent, as
argued in chapter 3, how much more so is the active, interrogative observation of
contrived events that constitutes experimentation? If observations are both un-
reliable (because of the frailty of our senses) and dependent on what the observer
knows and assumes about the phenomenon being studied, and if experiments are
predicated on the basis of particular assumptions about conceptual relationships,
we cannot conclusively and confidently reject a theory on the basis of observations
deriving from that experiment because the observations are impregnated with the
very theory that is under test. It is important for students to realize that every
experiment is set within a theoretical matrix (particular conceptual schemes and
theories), a procedural matrix (a community approved ‘method’ or practice under-
pinned by theories and conventions about how to conduct, record and report
experiments) and an instrumental matrix (theories underpinning the design and
construction of all scientific instruments employed in the experiment, together with
the theories of perception that underpin all observations). It is this complex of
theoretical understanding and assumptions that gives both form and purpose to
experiments. What counts as good research design, what kind of observations are

86

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
SCIENTIFIC INQUIRY, EXPERIMENT AND THEORY

sought, what measurements are regarded as legitimate, what instruments may


be utilized, and what sort of evidence is seen as crucial, are all determined by the
theoretical matrix within which the scientist is working. It is the reason why
experimental data are sometimes rejected on theoretical grounds and why there
have been occasions when theory-based calculations have proved more acceptable
than experimentally acquired data (Schaefer, 1986). Even when no error can be
found in the experimental procedure, data may be rejected because it is insuffi-
ciently plausible in terms of current theory. Of course, this tendency to back theor-
etical judgement in the face of contrary observational evidence can sometimes
work to the disadvantage of scientific development, as in the rejection of William
Bray’s elegant experimental data because current theory disallowed the notion
of oscillating chemical reactions, and continued to do so for almost 50 years
(Epstein, 1987).
Experiments form a critically important part of the scientist’s repertoire, so it
is crucial that we don’t misrepresent that role in the school science curriculum. If
Kuhn (1970) is correct, and historical evidence suggests that he is, theories are only
abandoned when there is compelling evidence (long-standing and striking at the
heart of the theory) and/or when an alternative and more promising theory becomes
available. It is misleading to present students with the idea that theories are dis-
carded because of a few negative results. In practice, all theories have to live with
anomalous data; it is a natural feature of science. We seriously mislead students
when we pretend that the kinds of experiments they perform in class constitute a
straightforward and reliable means of choosing between rival theories. Experiments
are enormously powerful for giving scientists precise information under highly con-
trived and highly controlled circumstances, but because experiments are conceived,
designed and executed with a particular complex of theoretical understanding,
considerable judgement is involved in appraising the significance of the evidence
they furnish. Whether to accept the evidence, and in consequence to accept or
reject the theory, reject the evidence, or conclude that some matters are still prob-
lematic and the experiment should be re-planned, is a decision that is not easily
made. When scientists are working at the limits of secure knowledge there are
sometimes uncertainties concerning the appropriateness of an experimental design,
the robustness and reliability of the instruments (possibly new ones, designed for
this particular inquiry) and even the ability of the technicians to achieve repeat-
able data. It is often the case that new craft expertise has to be acquired or a new
laboratory technique developed, so it may be some considerable time before the
eperiment ‘works properly’ in the sense of producing consistent data. Experimental
testing of theories is not, therefore, an infallible single step procedure; rather, it is
a multi-stage decision-making process monitored and validated by the scientific
community.
Unfortunately, as perusal of almost any school science textbook will confirm,
we present students with a very different version of the nature and purpose of
experiments. In particular, we neglect to make them aware that every experiment is
set within a theoretical matrix that determines what can be done, how it can be
done and how the data can be interpreted. The data does not speak for itself, as
some textbooks suggest when they liken scientific investigation to detective work.2

87

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
CHAPTER 5

Rather, the data means what the theory says it means. An important part of the
misrepresentation of experiment in school science is that scientific claims are
tested against reality. Not so! They are tested against our interpretation of reality,
often using evidence collected by instruments that encode in material form a great
deal of assumed theory that is subsequently implicit in the experimental con-
clusions (Latour, 1987). In addition, we seriously underestimate the complexity of
the relationships among observation, experiment and theory. What we should be
teaching students is that theory and experiment have an inter-dependent, interactive,
reflexive relationship: experiments assist theory building (by giving feedback
concerning theoretical speculation); theory, in turn, determines the kind of experi-
ments that can and should be carried out, and determines how experimentally-
acquired data are interpreted and used. Both experiment and theory, then, are tools
for thinking in the quest for satisfactory and convincing explanations. Neither has
absolute priority, though either may lead on a particular occasion. Newton et al.
(1999) express this position particularly well when they say: “Observation and
experiment are not the bedrock upon which science is built; rather, they are hand-
maidens to the rational activity of constituting knowledge claims through argument”
(p. 555). The history of science provides many examples of developments during
which theory was well ahead of experimental testing/corroboration and, equally,
lots of instances of episodes during which there was an abundance of data but no
satisfactory theory to account for it.3

LEARNING SCIENCE BY DOING SCIENCE

There is a long tradition in science education arguing that the most appropriate way
to teach science is through activities designed to mimic the activities of scientists.
This notion was particularly prominent in the penchant for discovery learning
in the 1960s. In the United Kingdom, the major impetus for the promotion of
discovery learning was the ‘progressive’ child-centred notion that inquiry-oriented
teaching is close to children’s ‘natural forms of learning’. Long-standing beliefs
that children are well-motivated by direct, inquiry-oriented experiences and learn
primarily through unstructured, play-like activities were reinforced by Piaget’s
descriptions of how the unstructured and self-directed observations and experimen-
tation of children develop via a series of stages into sophisticated formal reasoning
processes. Such was the frequency with which purely abstract reasoning seemed
to be preceded by an operational stage in which understanding was rooted in the
action itself, that there appeared to be an almost incontrovertible case for learning
science through student-driven, hands-on inquiry methods. The assertion that the
stage of independent thinking is achieved only by the “use of active methods which
give broad scope to the spontaneous research of the child or adolescent and require
that every new truth to be learnt be rediscovered” (Piaget, 1973, p. 15) was taken
by the authors of the Nuffield science projects as theoretical justification for a
revival of the heuristic approach first developed in the early 20th century by Henry
Armstrong (Jenkins, 1979). Indeed, Piaget was quoted in support of the claim that
adoption of other approaches would seriously prejudice student understanding:

88

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
SCIENTIFIC INQUIRY, EXPERIMENT AND THEORY

Each time we prematurely teach a child something he would have discovered


for himself, the child is kept from inventing it and consequently from under-
standing it completely. (Piaget, 1970, p. 715)
In the United States, the major impetus for discovery learning came from the
writings of Jerome Bruner and Joseph Schwab. In his influential essay ‘The Teaching
of Science as Enquiry’, Schwab (1962) set out an agenda for a science curriculum
emphasizing scientific inquiry as both content and method as the solution to what
he perceived as a crisis in US science education. He argued that laboratory experi-
ences should precede classroom instruction and that the laboratory manual should
“cease to be a volume which tells the student what to do and what to expect”; it
should be “replaced by permissive and open materials which point to areas in
which problems can be found” (p. 55).
According to Novak (1978) and Kirschner (1992), these assumptions were then
compounded, on both sides of the Atlantic, by thinking based on a misinter-
pretation of Ausubel’s (1968) work on ‘meaningful’ versus ‘rote’ learning. Rote
learning was falsely equated with transmission/reception methods and meaningful
learning with discovery methods. A further confusion arose from the failure to
distinguish considerations of how existing scientific knowledge is learned by
students (what I have termed learning science) from considerations of how new
scientific knowledge is generated and validated within the scientific community
(learning about science), and from experiences in which students themselves might
engage in authentic scientific inquiry (doing science) (see chapter 1, footnote 8).
Because scientists achieve their goals largely through observation and experiment,
it was assumed that the best way of learning of science is through activities based
on a model of scientific inquiry. Unfortunately, the Nuffield courses in the UK and
the BSCS, PSSC and ChemStudy projects in the US compounded these pro-
blematic assumptions by fusing progressive child-centred views emphasizing direct
experience and learning by inquiry and discovery with inductivist ideas about the
nature of scientific inquiry:
A learning-by-discovery sequence involves induction. This is the procedure
of giving exemplars of a more general case which permits the student to
induce the proposition involved… General understanding is induced from a
wealth of experience with specific cases. (Glaser, 1966, pp. 15 & 18)
In promoting the value of direct experience and an inquiry-oriented curriculum,
in stressing the motivational value of ‘finding out for oneself ’, and in their use of
terms such as observation, experiment and investigation, the curriculum developers
of the 1960s produced a model of learning which seemed to fit perfectly the tradi-
tional inductivist views of scientific method. As Cawthron and Rowell (1978) say:
It all seemed to fit; the logic of knowledge and the psychology of knowledge
had coalesced under the mesmeric umbrella term ‘discovery’ and there was
no very obvious reason for educators to look further than the traditional
inductivist-empiricist explanation of the process. (p. 38)

89

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
CHAPTER 5

What had started out as a psychological justification of learning by discovery


had slipped over into an epistemological argument. Not only was discovery learning
philosophically unsound (see chapter 3), it was pedagogically unworkable. For
example, early in the original Nuffield Physics course, students are provided with a
lever, a fulcrum and some weights (in the form of uniform metal squares) and are
invited to explore and “find out what you can”. No particular problem is stated; no
procedure is recommended: “Find out what you can about a balancing see-saw,
with different arrangements of squares (weights) on it… See if you can find out
some pattern about balancing that you could tell to other people” (Nuffield Physics,
1967, p. 186). It was assumed that the law of moments would simply emerge from
undirected, open-ended exploration. Nothing could be further from the truth. First,
because the pivot is below the centre of gravity the system doesn’t balance in the
way the students expect. If the weights are suspended below the pivot, as in a set of
scales, the beam will balance, but there is little chance that the students will dis-
cover this for themselves. Second, children tend to spread the weights irregularly
along the entire length of the beam. The complexity of this arrangement obscures
the simple relationship the teacher is seeking. Consequently, advice is proffered on
how to make the problem simpler and instructions are issued about the best way to
proceed.
Similar things happen whenever students are presented with this kind of open-
ended situation. For example, in another Nuffield Physics activity, students are
given a bar magnet, a card and a shaker of iron filings, with the aid of which they
are supposed to discover magnetic lines of force. Even teachers, who know that
these lines of force can be detected with iron filings, sometimes have great diffi-
culty revealing them. What chance is there for students, who do not know what
they looking for? They cannot discover something for which they are conceptually
unprepared. Without the relevant conceptual framework, they don’t know where to
look, how to look or how to recognize the lines as significant, even if they find
them. Additionally, there are many more entertaining things that young students
can find to do with a magnet and a batch of iron filings, as many harassed teachers
can testify.
There is also the problem that many so-called ‘experiments’ fail to give the
results that are required to meet the curriculum goals relating to content. Students
may make errors in observing, measuring or recording data, have accidents, lose
interest or just fail to finish. They may be distracted by all the clutter and ‘noise’ of
hands-on activity; they may become so immersed in the practical details of what
they are doing that they miss the conceptual significance of it (Hodson, 1993c).
Even if they ‘do everything right’, the waywardness of shoddy and sometimes
poorly maintained school apparatus may still lead them astray. As Dearden (1967)
reminds us, “a teaching method which genuinely leaves things open for discovery
also necessarily leaves open the opportunity for not discovering them” (p. 153,
emphasis added).
Many teachers responded to the pedagogical problems of discovery learning by
engaging in what came to be known as ‘directed’ or ‘guided’ discovery. It follows
this general form: the teacher guides the initial class discussion of the ‘experiment’
that is to be carried out; orchestrates the design of the ‘experiment’ – using the

90

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
SCIENTIFIC INQUIRY, EXPERIMENT AND THEORY

apparatus that has been previously laid out on the teacher’s bench; wanders around
the classroom/laboratory advising the ‘experimenters’ on matters of technique and
giving advice on recognizing and recording the significant observations (and
ignoring or re-making others); and guides the subsequent class discussion towards
the ‘discovery’ of the underlying theoretical principles to be used in the expla-
nation of the results. Throughout all these activities the teacher pretends that the
class is engaged in open-ended inquiry and deliberately suppresses (for the moment)
the very knowledge that was used in setting up the so-called inquiry. What purports
to be student-driven inquiry finishes up as a subtle but very powerful form of
teacher direction and control. To set up a situation that claims to be student-driven
and open-ended but which, in practice, demands a particular outcome is to confuse
learning science (which has a definite ‘result’ in mind) with doing science (which
hasn’t). Not only is this confusing for teachers, because they are sometimes left in
the situation where they are unable to respond properly to unexpected results, it can
be very disorienting for students. The confusion that can arise is neatly encapsu-
lated in an extract from some classroom observation work reported by Atkinson
and Delamont (1976). It concerns a series of lessons on photosynthesis in which the
class had covered growing leaves with metal foil, leaving several uncovered areas on
each leaf. After a few days the students tested for starch in various parts of the leaf.
The teacher tells the class that if starch is present only in the uncovered parts of the
leaf it can be concluded that light is necessary for starch production. Delamont reports
how one girl (Michelle) protests that such evidence constitutes proof; she suggests
that there could be other explanations. Upon which, another student (Sharon) says:
“Of course it’ll prove it. We wouldn’t be wasting our time doing it if it. didn’t”.
This episode highlights the crucial distinction between learning science and doing
science. Sharon has recognized the sham of the teacher’s supposed discovery
approach. She sees that the teacher has set up the activity in a particular way in order
to ensure a particular outcome, and thereby lead the class towards particular know-
ledge and understanding. For her, the class is learning science. Michelle, meanwhile,
believes that the class is engaged in a real scientific inquiry – a genuine experimental
study of the phenomenon of starch production in leaves. For her, the class is doing
science. The teacher’s problem is that she believes she can, and should, be doing
both. Of more significance in the context of this book, discovery learning projects
all manner of undesirable images of science. First, that it proceeds via induction;
second, that experiments are unplanned and discovery is largely accidental; third,
though it seems contradictory to the others, that scientists know what the results
of an experiment should be – hence the teacher’s impatience when activities in the
school laboratory do not turn out as the curriculum guidelines require.
While it is relatively easy to see how and why the curriculum developers of the
1960s, without the benefit of those views in the philosophy of science outlined in
chapter 44 and lacking research knowledge concerning children’s learning in
science only generated since the mid-1980s, were attracted to inductively-oriented
discovery learning, it is much more difficult to see why some science educators
continue to advocate this approach (see, for example, Sherman, 2000). One possi-
bility is that the inductivist view of science encapsulated in discovery learning is
perceived by teachers as more straightforward than other models of science and,

91

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
CHAPTER 5

therefore, easier for students to follow. Indeed, there is evidence to suggest that
teachers who adopt alternative models of science with high ability groups may still
revert to inductivist views with those classes perceived as less able (Hodson, 1993a).
A second reason is the high respect that primary school teachers have for child-
centred methods, which, because of common linguistic features (inquiry, investi-
gation, open-ended, unstructured, observation, discovery, etc.) appear to favour an
inductivist model of science (Harris & Taylor, 1983). A third reason is located
in teachers’ own inadequate views about the nature of science, which are largely
derived from their own learning experiences in school and university, reinforced
by the mythology of school science textbooks and curriculum documents. A fourth
reason is, perhaps, an emotional one: a temptation to cling to the notion that there
is a distinctive scientific method, even a precise algorithm for conducting scientific
investigations. This particular emotional need is, in part, responsible for the absur-
dities of the approach to science teaching outlined in the next section.

SCIENCE AS DISCRETE, GENERIC PROCESSES

Another major distortion of the nature of scientific inquiry is perpetrated by those


curricula that adopt the so-called ‘Process Approach’ to science education. Since
this movement in science education is addressed in some detail by Wellington
(1989), Hodson (1996) and Osborne and Simon (1996), little needs to be said here
about its history and development, beyond noting that Buchan and Jenkins (1992)
attribute its prominence in the UK during the 1980s and early 1990s to the merger
of hands-on pedagogy with the drive for rigorous and systematic assessment.
Hodson (1996) identifies an equally curious coalition underpinning process-oriented
science in the United States: “an uneasy amalgam resulting from the application of
the logical-analytical approach to curriculum planning to child-centred pedagogy”
(p. 121) – in effect, a technology of active learning. Prominent British examples of
process-oriented science include Warwick Process Science (Screen 1986, 1988),
Science in Process (Inner London Education Authority, 1987) and Active Science
(Coles et al., 1988). In the United States, the most notable example is Science –
A Process Approach (S-APA) (American Association for the Advancement of
Science, 1967), a course based largely on the work of Robert Gagne (1963, 1965).
Although S-APA is no longer in print its influence persists in an approach to
elementary science education that Cain and Evans (1990) refer to as Sciencing (see
also, Sherman (2000) and Abruscato (2004)).
Underpinning all process-oriented approaches, no matter what their origin, are
several basic assumptions, each of which deserves close critical scrutiny.
– Scientific inquiry can be described in terms of a series of discrete processes:
observing, inferring, measuring, predicting, classifying, collecting data, record-
ing data, formulating hypotheses, controlling variables, and so on.
– The processes are generic – that is, they are context-independent (i.e., theory-
free) and, therefore, transferable.
– Scientific knowledge results from engagement in these processes.
– Performance of these skills can be readily observed and accurately and reliably
measured.

92

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
SCIENTIFIC INQUIRY, EXPERIMENT AND THEORY

– By practising and developing these skills, students acquire the capability to


conduct scientific investigations.
Considerable space was devoted in chapter 3 to the argument that observation is
theory-impregnated and, therefore, that the notion of theory-free observation is
absurd. Also in chapter 3, a great deal of attention was devoted to the relationship
between observation and inference – in particular, to the spurious claim of many
science curricula that there is a simple and clear demarcation between them. Similar
arguments extend to all the other processes of science (Millar & Driver, 1987).
To be engaged in any of the processes of scientific inquiry one needs a focus of
attention: one has to classify or measure something, rather than something else,
one has to hypothesize about particular entities or events, and so on. It is not possi-
ble to engage in these activities independently of content. Moreover, the way one
classifies, measures and hypothesizes, and one’s level of sophistication in doing so,
depend crucially on one’s theoretical understanding. Science education is not about
teaching students to observe, classify, measure and hypothesize per se. They can
already do these things perfectly well, and have been doing so for many years,
since long before they came to science class, and they will carry on doing these
things after they leave the school science laboratory. What school science is
concerned with is scientific classification, scientific measurement, scientific hypo-
thesizing. What makes these processes scientific is the utilization of relevant and
appropriate science concepts in pursuit of scientific purposes. In other words, all
these processes are theory-laden and theory-driven. Put simply, doing science is a
theory-driven activity.
Scientific classification, for example, is not just a matter of noting similarities
and differences, or it would be sufficient in science lessons to classify banknotes
and postage stamps using criteria such as country of origin, colour, size and design
characteristics. Rather, it involves the utilization of scientifically significant and
appropriate categories, suited to the purpose for which the classification is being
carried out. Different purposes demand different criteria, and may involve different
theoretical understanding. It follows that success in classifying depends on appro-
priate matching of theory-based categories to purpose. It depends crucially on the
knowledge, experience, assumptions and expectations about purpose that the would-
be classifier brings to the task. A particular classification is not necessarily ‘right’ or
‘wrong’; rather, it reflects the sophistication of our knowledge and the nature of
our intent. The decision to classify an animal as mammal/reptile/fish or lives on
land/lives in water or carnivore/herbivore depends on why we wish to classify.
Is the classification intended to inform the study of evolutionary processes? Is it
part of preparation for ecological work? Or is it to help us find our way around the
local zoo? Any classroom activity involving classification or ‘looking for patterns’
(as curriculum documents sometimes describe it) is, therefore, inextricably linked
with teaching theory (appropriate concepts for classification) and identifying a
legitimate scientific purpose for the classification. As students acquire more theory
they are able to adopt increasingly abstract criteria. Thus, in chemistry, for example,
classification shifts from simple observational properties such as flammability and
solubility towards bond type and polarity.

93

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
CHAPTER 5

At the risk of beating the argument to death with a stick, the same can be said
for all the other processes of science: measuring, predicting, collecting and record-
ing data, and the like. None of them can be carried out in a theory-free way. For
example, the notion that hypotheses can be formulated or predictions made
independently of content is just too absurd to be seriously contemplated. What can
possibly constitute the basis of an hypothesis or a prediction other than good under-
standing of the phenomenon or event under consideration? Without theoretical
understanding, a hypothesis or prediction is no more than a shot-in-the-dark, a blind
guess – an activity with little or no educational value, and certainly no scientific
value. McNairy (1985), a strong advocate of the process approach, argues that
whether the student makes a ‘correct’ or ‘incorrect’ prediction does not matter. I
agree wholeheartedly, but with major and critical caveats. What does matter is that
the student has good reasons for making the prediction. What does matter is that
the student can establish a sound line of argument from her/his current under-
standing as the basis for the prediction. What does matter is that the student knows
enough about scientific inquiry to know what would constitute an appropriate and
rigorous test of the prediction. To make the argument one last time, the control of
variables (the basis of experimental testing) cannot be achieved without substantial
theoretical understanding of the phenomenon or event under study. How would a
scientist adopting a theory-free approach know what the important variables are
likely to be? In a state of ignorance, the experimenter cannot control any variables,
except fortuitously. Clearly, the planning of any experiment in which variables are
carefully and systematically varied is a theory-driven and theory-impregnated
activity.

TRANSFERABILITY

The theory-impregnated nature of scientific processes creates enormous problems


for the notion of transferability, which is, of course, a central principle of the pro-
cess approach to science education.
[Process skills] can be taught in a specific content area and subsequently used
to solve problems from other areas within science, or from subjects such as
social science or mathematics. The potential generalizability of the process
skills represents an important reason for emphasizing their development and
use. (Tobin & Capie, 1980, p. 590)
A consequence of accepting the view that experience of scientific processes in
any context is as significant as experience in any other is a commitment to the
following notions.
i. No particular conceptual understanding is significant – it doesn’t matter what
scientific content we teach.
ii. Ability to observe, classify and measure in one context can be taken as indi-
cative of a student’s capacity to do so in an entirely different context.
These arguments have been taken to an absurd degree by those who advocate
rigorous and systematic skills-based testing in science – a movement which I
have described elsewhere (Hodson, 1993b) as educationally worthless (because it

94

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
SCIENTIFIC INQUIRY, EXPERIMENT AND THEORY

trivializes learning), pedagogically dangerous (because it encourages bad teaching),


professionally debasing (because it de-skills teachers) and socially undesirable
(because of powerful hidden messages concerning control and compliance). My
concern here is that this approach to assessment and evaluation is philosophically
unsound in its claim that competence in a skill such as observing or measuring can
be learned in a particular context (any context) and subsequently transferred to an
entirely different one with no apparent loss of capability.5 One is tempted to ask in
what sense the skills learned in observing the behaviour in an ant colony can be
successfully utilized in qualitative inorganic analysis or in making astronomical
observations? In what sense does learning to dissect a dogfish help a student to use
an oscilloscope or to synthesize a complex organic molecule? If we applied this
principle of transferability between unlike contexts to the world outside school we
would happily submit to a brain operation carried out by a specialist in obstetrics.
In the real world, including the world of scientific practice, the context in which
skills are acquired is crucial to the proper performance of that skill and to our
confidence in the practitioner. In the words of Finley and Pocovi (2000), “a high
energy physicist dropped in the middle of a human genome project and asked to
collect the relevant observations would probably be clueless as to what obser-
vations were even possible, let alone which ones would be relevant to the problem
of unraveling the human genome” (p. 58).

SCIENTIFIC INQUIRY AS A SIMPLE ALGORITHMIC PROCEDURE

The final absurdity of the Process Approach to which I wish to draw attention is
the assumption that once students have acquired the separate skills of observing,
classifying, measuring, and so on, they can put them together into a procedure for
doing science. In other words, it is assumed that doing science consists in the sum
of its parts and is no more than the sum of its parts. In short, scientific inquiry
comprises an algorithm to be applied in all circumstances.
There are two principle arguments against this proposition. The first is that there
is no empirical evidence from educational research to support it. Success in
carrying out a series of decontextualized tasks focusing on observation, classifi-
cation or measurement says very little about a student’s capacity to conduct a
scientific investigation, much less about her or his ability to design such an inquiry.
It is sometimes the case that students who perform adequately on these sanitized
tasks are unable to integrate the skills into a coherent and effective strategy for
investigation. Conversely, many who perform poorly in the tests can engage in
interesting and successful scientific inquiry when they are suitably encouraged and
given the freedom and support to follow their interests.
The second counter argument relates to Paul Feyerabend’s (1975) assertion that
the idea of a fixed method of science rests on too naïve a view of what is involved
in conducting investigations and building theories. As noted at the end of chapter 4,
Feyerabend states that the only principle that is always applicable in conducting the
complex and sometimes chaotic business of science is the principle Anything Goes.
In similar vein, Percy Bridgman (1950) remarked that “scientific method, as far as

95

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
CHAPTER 5

it is a method, is nothing more than doing one’s damnedest with one’s mind, no
holds barred” (p. 278).
Much of Feyerabend’s writing – ‘epistemological anarchy’, as his critics have
labelled it – casts doubt on the view of science projected in most school science
curricula and on the public image of science and scientists. For example, he ques-
tions one of the cornerstones of school science: respect for the ‘facts’. Regardless
of whether the curriculum projects an inductivist or a falsificationist model of
science, correspondence with the facts (the observational aspects of a theory) is
taken as a measure of a theory’s ‘truth’, or verisimilitude as Popper (1972) calls it.
More particularly, lack of correspondence is taken as grounds for theory rejection.
Feyerabend (1975) takes a somewhat different view:
The suspicion arises that the absence of major difficulties is a result of the
decrease of empirical content brought about by the elimination of alter-
natives, and of facts that can be discovered with their help. (p. 43)
A theory’s success in accounting for observational evidence is a consequence of
the way in which the theory is constructed – that is, in building a theory, scientists
ensure that it explains the observational data in its domain and excludes the possi-
bility of generating the evidence that would or could refute it. Moreover, Feyerabend
argues, “empirical ‘evidence’ may be created by a procedure which quotes as its
justification the very same evidence it has produced” (p. 44). In a sense, theories
are like computer simulations: the ‘facts’ follow from the initial assumptions; the
theory is designed in such a way that it creates the factual evidence that is sub-
sequently used to justify the theory.6 The theory will explain the facts because it
was designed to do so; there are no ‘counter facts’ because the theory was designed
to exclude them, or to ensure that they cannot be revealed.
For how can we possibly test, or improve upon, the truth of a theory if it is
built in such a manner that any conceivable event can be described, and
explained, in terms of its principles? (p. 45)
The theory may well be false, but it appears valid because it creates its own con-
firmatory evidence in consequence of what it allows and disallows. Scientists who
subscribe to a particular theory may find it difficult, if not impossible, to recognize
deficiencies in that structure because their theoretical biases and expectations blind
them to the theory’s shortcomings and prevent them from obtaining, or even from
seeking, appropriate counter evidence.
How can we possibly examine something we are using all the time? How can
we analyse the terms in which we habitually express our most simple and
straightforward observations and reveal their presuppositions? (p. 32)
The only way of casting doubt on a theory is by creating a new theory, capable
of looking at the world in a different way. Often, the evidence that might refute a
theory can only be unearthed with the help of an incompatible alternative. When new
theories are employed, they often reveal new ‘facts’ of a kind very different from
those uncovered by the old theory. Thus, a new theory may be supported by an obser-

96

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
SCIENTIFIC INQUIRY, EXPERIMENT AND THEORY

vational test that was not possible within the context of the old theory; the earlier
theory may be rejected on the basis of a test that would have been quite inconcei-
vable within the conceptual framework of the old theory. Therefore, says Feyerabend
(1975), we should proceed counter inductively: introduce and elaborate hypotheses
that are inconsistent with existing theories and/or with the facts that derive from
them, even if those theories are “highly confirmed and generally accepted” (p. 105).
As the new theory is elaborated it will generate its own confirmatory evidence. If
the new theory is satisfying, for whatever reasons, allegiance will shift to it and the
old theory will simply wither away. The old theory is not falsified; it is abandoned.
The new theory does not necessarily solve the problems of the old, it dissolves
them. Old problems are removed because the new theory declares non-existent the
things that constituted the problems.
Whether one accepts Feyerabend’s principle of counter induction or not, it is
clear that correspondence with the observable facts does not afford any increased
truth status on a theory, it simply means that it may be true. However, there may be
an alternative theory that also agrees with the facts.
Theory can still vary though all possible observations be fixed. Physical
theories can be at odds with each other and yet compatible with all possible
data even in the broadest sense. In a word, they can be logically incompatible
and empirically equivalent. (Quine, 1970, p. 179)
In such circumstances, there cannot be any justification for accepting (or reject-
ing) one of the statements that is not a justification for accepting (or rejecting) the
other. Put differently, if observations provide equally good evidence for each of the
generalizations, then they provide no significant evidence for either of them.

THEORY ACCEPTANCE AND REJECTION

Posner et al. (1982) argue that a new conceptual scheme will be accepted only if
learners are dissatisfied with their current belief/understanding and have access to a
new or better idea with which to replace it. To gain acceptance the new idea should
be intelligible, plausible and fruitful. Feyerabend seems to argue that in science
itself (rather than in learning science) it may be the other way around. Ideas become
clear and reasonable only after parts of them have been used for a long time and
have been refined through use. Nor does science necessarily begin with a problem,
as Popper alleges. Often it starts with ‘idle play’, which develops into a solution to
a problem that only becomes apparent after the event. We do not think and then
act, Feyerabend says, we act and then think – much as children do. We work with
our ideas, by whatever means we can, until the theory has solved its own problems
– that is, succeeded in creating the evidence that subsequently justifies it. In
Feyerabend’s words, “Such unreasonable, nonsensical, unmethodical foreplay thus
turns out to be an unavoidable precondition of clarity and of empirical success”
(1975, p. 27). In consequence, science is much more sloppy and irrational than the
conventional image would suggest: “Without chaos, no knowledge… Without a
frequent dismissal of reason, no progress. Ideas which today form the very basis of
science exist only because there were such things as prejudice, conceit, passion;

97

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
CHAPTER 5

because these things opposed reason and because they were permitted to have
their way” (1975, p. 179). The rationality of the chosen method is only seen after-
wards, when satisfactory conclusions have emerged. Indeed, it could be argued that
rationality is created retrospectively, as part of the case scientists build to persuade
others of the validity of the findings, as Max Born (1934) remarked:
I believe there is is no philosophical highroad in science, with epistemo-
logical signposts. No, we are in a jungle and find our way by trial and error,
building our road behind us as we proceed. We do not find signposts at cross-
roads, but our own scouts erect them, to help the rest. (p. 44)
The crux of Feyerabend’s argument is that in order to make progress, scientists
should do whatever suits the particular circumstances in which they find themselves
– the nature of the problem, the range of theoretical perspectives available, the
opportunities and facilities for observation and experiment, and so on. Sometimes
one kind of move may assist progress, sometimes another. Sometimes the most
outrageous move pays off, and sometimes it proves to be just another crazy idea
that got nowhere. As Albert Einstein is reputed to have said, the scientist needs to
be “an unscrupulous opportunist” (Schilpp, 1951, p. 683), at least from the point of
view of those who advocate a strict method of science. Given the complexity of the
scientific enterprise, the myriad of different starting points for an investigation, the
major differences in knowledge, experience and personality among scientists, and
the likelihood of substantial variations in the range of facilities and resources on
which individual scientists can draw, it would be surprising if all scientists pro-
ceeded in the same way (White, 1983). Interestingly, young children consider
diversity of approach inevitable; they have no expectations of a particular method
for doing science (Hodson, 1990). Teachers and science textbooks create the
expectation of a single method through their continual reference to the scientific
method, perhaps in an effort to simplify the teaching of integrated or combined
science.
However, the assertion that ‘anything goes’ should not be interpreted as a state-
ment that science has no methods. It implies the absence of a prescribed method –
an algorithm – rather than the absence of methods. It should not be taken too literally.
Scientists wishing to make progress cannot just do anything. As Newton-Smith
(1981) observes, “Lazing in the sun reading astrology is highly unlikely to lead to
the invention of a predictively powerful theory about the constituents of the quark”
(p. 269). Nor, despite Feyerabend’s advocacy of ‘rule breaking’, can every rule be
broken. Moreover, while any particular rule may (and perhaps should) be violated
in a particular set of circumstances, it is not rational to violate all rules simul-
taneously.7 Particular rules are violated to solve particular theoretical problems, but
other rules governing the ways in which the new idea is appraised (concern for
explanatory adequacy, simplicity, clarity, accuracy, insistence on testing, and so
on) continue to exercise control over what new ideas can be generated. Science has
many methods, including widely varying techniques and procedures for conducting
experiments, analysing data and presenting results. Not to follow these procedures
carefully and systematically is, in some important sense, to be unscientific. At

98

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
SCIENTIFIC INQUIRY, EXPERIMENT AND THEORY

the very least, it is to be sloppy and unsystematic to an extent that prejudices the
scientific quality of the work being conducted. Implying that the world of the
scientist is totally anarchic does students (and the public image of science) as gross
a disservice as implying that science has a single, all-powerful algorithmic method
that is applicable in all circumstances. Indeed, in his later work, Feyerabend
substantially modified his position on methodological rules, arguing that there are
methodological rules but they are highly sensitive to their context of use and have
no applicability outside that context..
A naïve anarchist says (a) that both absolute rules and context dependent
rules have their limits and infers (b) that all rules and standards are worth-
less and should be given up… I agree with (a) but I do not agree with (b).
I argue that all rules have their limits and that there is no comprehensive
‘rationality’, I do not argue that we should proceed without rules and stan-
dards. (Feyerabend, 1978, p. 32)
It is also important to note that while the new idea may lead to the rejection of
some existing theoretical constructs, all is not discarded. Some elements of the
previous framework must remain, in the light of which the new structure makes
sense and proves fruitful: “The fact remains that no one can depart from too much
of the preselection at once and expect to make progress… What we perceive as
revolutionary innovation in a field always challenges only a little of the preselec-
tion. Only because we focus on the contrast rather than the continuity does innova-
tion seem so much of a departure” (Perkins, 1981, p. 279).
Claims to scientific knowledge have to be publicly argued and publicly justified;
the data from which conclusions are drawn and theories built have to be rep-
roducible and the chain of argument from premise to conclusion has to be clear. In
other words, science does have methods (the things that scientists routinely do8)
and it does have criteria for judging the validity of knowledge claims, but their
particular form depends on the particular circumstances: the matter under consider-
ation, the conceptual structure (‘research programme’, as Lakatos calls it) within
which the investigator frames the problem(s), the investigative techniques and
instrumentation devices available, the scientists’ familiarity with them, and so on.
Over time, new methods are introduced and old ones are refined or discarded. By
making a selection of processes and procedures from the range of those currently
available and approved by the community of practitioners, scientists choose a
‘method’ or cluster of methods that they consider contextually appropriate. There
are no universal decision criteria for what to do and how to do it. All decisions are
‘local’ – determined by the particular circumstances of individual investigations –
and, therefore, idiosyncratic.9 Oakeshott (1962) expresses this view as follows:
The coherence of scientific activity (does not) lie in a body of principles or
rules to be observed by the scientist, a ‘scientific method’; such rules and
principles no doubt exist, but they also are only abridgements of the activity
which at all points goes beyond them, in particular, in the connoisseurship of
knowing how and when to apply them. Its coherence lies nowhere but in the
way the scientist goes about his investigation, in the traditions of scientific

99

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
CHAPTER 5

inquiry. These traditions are not fixed and finished, and they are not to be
identified with merely current scientific opinion, or with an identifiable
‘method’; they are the guide in every piece of scientific investigation and at
the same time they are being extended and enlarged wherever scientists are at
work. (p. 103)
Particularly intriguing are the differences in approach adopted by practitioners
in the different sciences – something on which school science is strangely quiet.
Bauer (1992) comments as follows: “The differences among adepts of the various
sciences go beyond matters of theory, method, and vocabulary to subtler habits of
thought and even to customs of behavior, to such an extent that the differences…
can aptly be described as cultural” (p. 25). Ernst Mayr (1988) claims that many
famous controversies in biology are a direct consequence of these kinds of ‘cultural
differences’. The differences between the sciences, particularly the extent to which
the discipline tends to be theory-driven or data-driven and whether it seeks to
establish a simple linear cause and effect relationship or a complex web of inter-
relating causal factors, speaks to the issue of alternative sciences and whether, for
example, there is any meaning to expressions such as ethnoscience and feminist or
gynocentric science (see chapter 7).
When the community comes to appraise a piece of scientific research, one of its
criteria of judgement is a consideration of the methods employed. Were they well
chosen? Were they satisfactorily performed? Could/should the investigation have
been conducted differently? How was the evidence acquired? What instruments
were used, and why? What errors are likely and how, if at all, were they estimated?
Under test, too, is the craft expertise of the experimenter and the sometimes com-
plex ‘know how’ necessary to ensure that the experimental procedures ‘work
properly’. Sometimes other scientists are convinced by the quality of the data, the
elegance of the method or the persuasiveness of the argument; the particular
criteria deployed will vary with the circumstances. A claim to knowledge deriving
from experiment is likely to be accepted as valid if the various ways in which the
claim could be invalid have been thoroughly investigated and, as a result of the
investigation, have been discounted. This is what Mayo (1996) calls “severe
experimental testing”: an experiment constitutes support for a claim only when
possible sources of error have been eliminated and, in consequence, the claim
would be unlikely to “pass the series of tests” unless it were true.10 For example,
Eddington could not have observed the effect of light bending in the vicinity of the
Sun unless there is a fairly good possibility that Einstein’s theories about gravity
are correct. It is important to emphasize to students just how difficult it is to
assemble the mass of evidence necessary to install a new theory (and, thereby,
overthrow an existing one) and to present it in such a way that enables the claim to
achieve consensus within the community of practitioners.
It is the demand for consensus within the community of scientists about what
should be accepted as legitimate knowledge and the requirement that all claims to
knowledge meet community-based criteria of validity, reliability and methodo-
logical appropriateness that invests scientific knowledge with its particular kind of
authority. Unlike everyday knowledge, which needs little beyond simple consensus

100

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
SCIENTIFIC INQUIRY, EXPERIMENT AND THEORY

or the personal authority of the knower for justification, scientific knowledge has to
survive rigorous critical scrutiny by members of the scientific community, who
achieve consensus by employing well-characterized methods and clearly stated
criteria of judgement. Longino (1994) identifies four conditions that a community
of practitioners must meet if consensus is to count as knowledge rather than mere
opinion.
– There must be publicly recognized forums for criticism
– There must be publicly recognized standards for evaluation of theory and
practice
– There must be uptake of criticism – the community needs to do more than
merely tolerate dissent; it must act on it.
– There must be equality of intellectual authority – what is included or excluded
must result from critical dialogue rather than the exercise of political or eco-
nomic power.
If it survives critical scrutiny by the community, using these public methods of
evaluation and judgement, the knowledge item (model, theory, experimental pro-
cedure, instrumental technique, or whatever) becomes part of the written record
of the scientific community and is made available to others.11 Because of this
mechanism and the confidence that practitioners have in it, science is cumulative;
current researchers utilize the knowledge generated by previous scientists and,
in doing so, may develop it or discard it. In other words, science has a history, and
although a theory may be displaced by one judged to be ‘better’, it retains its
historically-located validity.

STUDENTS’ VIEWS

As noted in chapter 2, most school age students and many undergraduates hold the
view that scientific knowledge can be proven by means of careful appraisal of
experimentally acquired data, though primary (elementary) school students some-
times see experiments as unplanned or highly speculative activities that often give
rise to unexpected outcomes (Duveen et al., 1993). Understandably, as they pro-
gress through the school system, students begin to reject the ‘cartoon image’ of
scientific inquiry in favour of the ‘school curriculum image’ (Solomon, et al.,
1996). Whether this results in an authentic view of experiment and the relationship
between evidence and theory depends, of course, on the nature of the curriculum
and the priorities of the teacher.
In recent years there has been widespread adoption of the so-called ‘fair testing’
approach to investigative work in school science, especially in the United Kingdom
(Watson et al., 1999; Watson, 2000). It is an approach that reinforces the view that
scientific inquiry consists in the isolation and systematic manipulation of variables.
This ‘standardization’ of scientific inquiry contributes little to students’ understan-
ding of the methods employed in ecology, geology, astronomy and meteorology –
what Mayer and Kumano (1999) call the “systems sciences”, in which it is field
work (rather than lab work) and the monitoring of changes over time (rather than
regarding time as a variable to be controlled) that are important, and where empha-
sis shifts from analysis to wholism and quantitative relationships are replaced by

101

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
CHAPTER 5

‘thick description’ in qualitative terms. Discussion in chapters 7, 8 and 9 will raise


questions about whether these elements of the ‘systems sciences’ should be
regarded as impregnating all sciences in the 21st Century. Perhaps this would be
the way in which scientific knowledge could be transformed into scientific wisdom
and attention focused on the kinds of social and environmental problems discussed
in chapters 1 and 2.
The ubiquity of the ‘fair testing’ approach may also be partly responsible for the
major findings of studies by Millar et al. (1994), Leach (1999), Robertson (1999)
and Ryder and Leach (2000) that students at all academic levels have little aware-
ness of the significance of theory in the interpretation of data and, conversely, often
fail to recognize the role of evidence in the appraisal of theory. Hogan (2000) has
pointed out that many students have difficulty relating their experiences in school
science lessons during hands-on, investigative work to their ‘formal’ knowledge of
the protocols, practices and procedures of the scientific community. One way of
attending to this shortcoming is to provide more metatalk about purposes and
procedures during laboratory work and field work. Another way is to provide more
explicit instruction about the relationships between models and theories in science
– an issue that is addressed in chapter 6.

ENDNOTES
1
Bencze (1996) provides a detailed argument for the promotion of correlational studies in school
science and a critical discussion of their distinctive features, including systematic inquiry via statis-
tical control.
2
Gallagher and Ingram (1984) are typical of those textbook authors who adopt this position when
they say that “science is about asking questions… you ask scientific questions when you are reason-
ably sure that the answers you get can be trusted” (p. 6). In other words, there is a true and certain
explanation located in the facts revealed by observation and experiment.
3
A detailed case for including the history of science in the school curriculum is provided in chapter 8.
4
The deliberations that led to the discovery learning movement preceded the publication of key works
by Kuhn, Lakatos and Feyerabend. Also, given the lengthy time lag between the publication of
major works in philosophy and psychology and their appearance in the consciousness of teachers, it
is safe to say that even Popper’s work was not well known among science educators until well into
the 1960s.
5
A much more detailed analysis of the philosophical problems associated with this approach can be
found in Hodson (1992b).
6
This situation is evident in school science when ‘experiments’ are conducted to ‘prove Ohm’s law’
using equipment calibrated on the assumption that Ohm’s law is valid.
7
It should be emphasized that it is not irrational, as some of Feyerabend’s critics assert, to violate a
rule, any more than it is necessarily rational to follow a set of rules. The nature of the rules and the
context of action determine the rationality of the conduct.
8
The notion that science is simply ‘what scientists do’ will be discussed in chapter 7.
9
The implications of this position for teaching students about science and providing opportunities for
them to conduct their own scientific investigations are addressed in Hodson (1992b; 1993b,c).
10
Questions about the ‘truth’ of scientific knowledge are examined in chapter 6.
11
The extent to which the community-approved procedures for appraising and validating scientific
knowledge are able to guarantee the integrity of the scientific enterprise will be discussed further in
chapters 6 and 7.

102

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
CHAPTER 6

REALISM OR INSTRUMENTALISM
What Position for School Science?

Closely linked with arguments about how scientists generate new scientific
knowledge (see chapter 5) is consideration of the role and status of that knowledge.
Some years ago, in the early days of the Nuffield science teaching projects in the
UK, Ernest Coulson (then Director of the Nuffield Chemistry project) said that
it was important for students to develop “a proper view of theory”. At the time, I
was teaching in a Nuffield ‘pilot school’, where curriculum materials and new
approaches to teaching and learning science were being trialled, and my problem
was that I didn’t know what “a proper view of theory” is supposed to be. For
example, it was fairly common for science textbooks, teachers and curriculum
documents to draw an analogy between science and detective work, thus implying
that there is a correct version of events and phenomena, and that the application
of good scientific procedures and the adoption of appropriate attitudes will enable
scientists to arrive at the truth about the world. Philosophers of science refer to
this position as naïve realism. It is a view that is still promoted in some science
textbooks. At the other extreme of the ‘philosophical spectrum’ is instrumentalism
(one of several variants of ‘anti-realism’1) – the notion that theories are simply
convenient devices (‘fictions’) that enable us to predict, manipulate and control.
Created entities like atoms, electrons, genes, gravitational fields and black holes
are built into theoretical structures designed to account for ‘the world beneath
surface appearances’ and give scientists a means of interacting with the real world
in predictable ways.2 Provided that we solve our problems quickly and accurately
(e.g., predicting the time of the next solar eclipse), it really doesn’t matter whether
the theory employed to solve the problem or make the calculation is true or not,
or whether the theoretical constructs really exist. Moreover, because it has no other
value, the ‘predictive device’ can be discarded once it has ‘done the job’. Truth is
irrelevant. Utility is the significant criterion. Thus, the kinetic theory of gases enables
us to state precise relationships among volume, temperature and pressure and to
predict the behaviour of gases on heating, but whether the molecules postulated in
the theory actually exist, or not, is irrelevant. The position is neatly encapsulated in
this quotation form the work of Ernst Mach (1911)
In the investigation of nature, we have to deal only with knowledge of the
connexion of appearances with one another. What we represent to ourselves
behind the appearances exists only in our understanding, and has for us the
value of a memoria technica or formula, whose form, because it is arbitrary
and irrelevant, varies very easily with the standpoint of our culture. (p. 49)

103

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
CHAPTER 6

Toulmin (1963) adopts a less extreme form of instrumentalism, arguing that we


cannot claim that electrons and protons exist but we can act as though they exist –
that is, we can rely on them, and their postulated entities, as useful guides for
dealing with whatever practical and theoretical problems may arise. Theories are
not assertions about the world; rather, they are “techniques of explanation” – maps
to enable us to find our way around the phenomena we wish to study, just as
ordnance survey maps enable us to get around the British countryside and the
famous map of the London Underground helps us to negotiate the city. Hence,
theories are regarded as ‘holding’ rather than true, as ‘not holding’ rather than
‘false’. They are adopted or neglected, rather than believed or denied.
Both realism and instrumentalism are beset with enormous difficulties. Naïve
realism sets standards for the acceptability of knowledge claims that simply cannot
be met. We cannot have direct access to truth about the world via our senses
(which are fallible) and our created conceptual systems (which may be wrong for
any or all of the reasons discussed in chapters 3 and 4). We may wish to compare
our theory of the world with the real world, but we have no means of doing so.
We can investigate nature and develop theoretical understanding of the world,
but we cannot compare what we think we know with the truth to see how
well we are doing. (Ellis, 1985, p. 69)
As I argued earlier, we are ‘prisoners’ of our woefully inadequate senses (our
physical frailty) and our capacity to theorize (our intellectual frailty). Moreover,
whatever we interact with in order to build our conception of it could be changed in
some possibly unknowable way by the interaction. In short we risk contamination
and distortion through interference, so that all we can find out is information/
knowledge about “interacted-with-things”. Again, there is no guarantee of direct
access to the real world. We have no way of stepping outside our minds and
bodies to see the world in a theory-free way, assisted by perfect senses. Our
success in building a model or theory about the world cannot be checked against
the reality we are theorizing about. Even the instruments we build to give us
greater access to information about the world are ‘contaminated’ by the theoretical
assumptions make in their design and construction. In consequence, agreement
with the ‘facts’ (observable evidence) does not mean that a theory is true. Firstly,
the facts may be wrong – after all, they are interpreted responses to sense data
collected with severely limited ‘tools’ (our senses or our instrumental extensions of
them). Secondly, there may be other theories that also agree with the facts. The
observational evidence may be susceptible to interpretation in more than one way,
to suit more than one theory. Duhem (1962) argues as follows: Suppose that the
hypotheses of a particular theory are able to explain all known appearances. What
can be concluded is that they may be true, not that they are necessarily true. In
order to make legitimate this last conclusion it would have to be proved that no
other system of hypotheses could possibly be imagined which could explain the
appearances just as well.
If two theories can provide contradictory accounts for the same observable
phenomena, yet make similar predictions about observable events, they are equally

104

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
REALISM OR INSTRUMENTALISM

well corroborated by the available evidence. However, they cannot both be true in
the sense that the theoretical entities they utilize in explanation and the interactions
they postulate about those entities exist independently of our theorizing about them.
We may wish to choose between these two theories on grounds of simplicity,
elegance or similarity with other theories, as Kuhn (1970) describes (see also
chapter 7), but these criteria say nothing about the truth of the chosen theory. We
cannot assume truth from explanatory power.
These problems lead some theorists and some science teachers and textbooks
to embrace instrumentalism. Interestingly, this tendency seems to be greater in
textbooks for the physical sciences than the biological sciences. Perhaps the exotic
language of contemporary physics, with its leptons, quarks (and their traits of colour,
flavour, strangeness, taste and charm), gluons and black holes lends strength to the
notion that physics is an elaborate fiction. Perhaps the complexities of general
relativity theory and quantum mechanics are such that the task of trying to imagine
the real world described by the equations is just too daunting. Much easier to regard
them simply as organizing tools. In school science, talk of ‘frictionless planes’,
‘ideal gases’ and ‘point masses’ seems to point clearly to an instrumentalist position.
Nothing in the real world matches these conceptualizations and descriptions.
Another argument for anti-realism is that the methods by which scientists
investigate knowledge claims are so extensively theory-dependent and theory-
driven (see chapters 4 and 5) that they are not so much a means of discovery as a
means of constructing and re-constructing knowledge, and so cannot tell us any-
thing about the world outside those constructions.3 In other words, if the ideas and
procedures that scientists use to investigate the world are social constructs then
what they tell us about the world is also a social construct, unless they ‘hit on the
truth’ by accident. Therefore, it makes sense to admit that theories are simply
fabrications of the human mind, invented to give us a measure of control and
predictive capability. Indeed, Laudan (1984) argues that because the realist goal is
unattainable, and because there is no means to measure how close a theory is to
the truth, then scientists who pursue realist goals are acting irrationally. Thus, his
position is that kinetic theory uses created entities such as atoms and molecules as
devices to account for the behaviour of real gases such as oxygen and nitrogen, but
makes no claim that such entities really exist. The behaviour of real objects like bar
magnets and iron filings is explained by means of theoretical devices such as
magnetic fields without any suggestion that magnetic fields are real. Some would
respond that moral perfection is also beyond our reach but that doesn’t mean we
shouldn’t strive for it or that it is irrational to do so. Indeed, it is generally held to
make one a better person. So with science, perhaps: striving for truth makes you
a better scientist and serves as a ‘corrective thrust’ by demanding explanatory
breadth rather than mere computational efficiency.

THE LANGUAGE OF OBSERVATION AND THE LANGUAGE OF THEORY

Many of those inclined to favour anti-realist views of scientific knowledge draw


a sharp distinction between terms like dog, tree, yellow and hot, which refer to

105

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
CHAPTER 6

observable objects, properties, relations and events, and terms such as electron,
gene and magnetic field, which refer to unobservable (i.e., theoretical) entities and
can only be understood within the context of the theoretical system in which they
are located. The position is clearly stated by Carnap (1956).
It is customary and useful to divide the language of science into two parts, the
observation language and the theoretical language. The observation language
uses terms designating observable properties and relations for the description
of observable things or events. The theoretical language, on the other hand,
contains terms which may refer to unobservable events, unobservable aspects
or features of events. (p. 38)
The position seems to be that what we can observe, and describe via the obser-
vational language is real, while what we cannot observe – and have to infer, as the
explanation of the observation – is theoretical and can be regarded as ‘fiction’. In
other words, what is plainly observable to persons of normal perceptual ability is
real and what is unobservable is not real. It is worth noting that both the Ontario
Ministry of Education (1987) and the Alberta Ministry of Education (1993) urge
teachers to draw a distinction between the descriptive language of science and the
theoretical language of science, perhaps indicating a strongly instrumentalist view
of scientific knowledge. My position, as outlined in chapters 3, 4 and 5, is that
all scientific language is impregnated with theory. All terms used in designing,
conducting and reporting scientific investigations are theory-laden, though in many
cases we don’t recognize that we are using a theory-loaded term because the theory
that underpins it is no longer problematic. The theory has become part of everyday
language, and is now accepted without question. The point at which an individual
begins to regard scientific terms as ‘theoretical’ is determined by her or his parti-
cular level of theoretical sophistication. Children, science teachers and research
scientists will draw that line in different places. One of the reasons why many
people find scientists difficult to understand, and accuse them of using esoteric
jargon, is that they don’t have the conceptual understanding that has been built into
that particular everyday language of scientific reporting. Problems can be consi-
derable when faced with statements such as “Our eyes are sensitive to electro-
magnetic waves in the frequency range 4 × 1014 to 7.5 × 1014 Hz (wavelength
between 4 × 10–7 and 7 × 10–7)” and “Once transcription has been successfully
initiated, the RNA polymerase continues along the DNA molecule until it en-
counters terminator sequences on the non-transcribed DNA strand”. Even simple
observation statements can be problematic: “the pH was found to be 4.73” or “the
resistance of this coil is 2.5 ohms”.
Given the arguments concerning the theory-laden nature of observation outlined
in chapter 3, we could conclude that ‘we see what we choose to see’ simply by
opting to use a particular theory, rather than some other theory. Does it follow,
therefore, as the anti-realist position on the distinction between observational and
theoretical language would seem to imply, that we can decide for ourselves what is
real and what is fictional? Like Maxwell (1962), I take this as an absurd position
to hold.

106

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
REALISM OR INSTRUMENTALISM

Our drawing of the observational-theoretical line at any given point is an


accident and a function of our psychological make-up, our current state of
knowledge, and the instruments we happen to have available and, therefore, it
has no ontological significance whatever. (p. 14)
Hempel (1958) draws a distinction between directly observable and indirectly
observable: “In regard to an observable term it is possible, under suitable circum-
stances, to decide by means of direct observation whether the term does or does not
apply to a given situation… Theoretical terms, on the other hand, usually purport to
refer to not directly observable entities and their characteristics. They function… in
scientific theories intended to explain empirical generalizations” (p. 42). This
elaboration of the language doesn’t seem to help at all. It isn’t entirely clear what
directly observable and indirectly observable mean. And so it is not entirely clear
what is supposed to be real and what is simply created/fictional. It is easy to
observe a dog or a fish in a direct way, so dogs and fish must be real. So far, so
good. What about an amoeba? It can’t be observed with the unaided senses, but
it can be observed through a microscope (which can be regarded as a simple
extension of our senses). Is an amoeba directly observable, and therefore to be
classed as real, or is it indirectly unobservable, and therefore fictional? Does it
mean that the amoeba was at one time a mere fiction (because it could not be
observed by direct use of the senses) but became real once the microscope had
been invented? Jupiter’s moons are not visible with the naked eye. Did they only
become real at the time telescopic observations became available? Or did they only
become real when, in 2006, a NASA spacecraft landed on Titan? Viruses and
protein molecules are not visible through a microscope but can be ‘observed’ using
an electron microscope. Does the use of an electron microscope, which utilizes
senses that humans do not possess, count as direct observation, and therefore
qualify a protein molecule for real rather than fictional status? What can we claim
to ‘see’ in the ultrasound imaging of a developing foetus? It is indirect observation,
of course, but we know that the foetus is real enough! As an aside, viruses rep-
resent an excellent example of those numerous occasions in the history of science
when scientists, having postulated unobservable entities and incorporated them into
theoretical structures which they have subsequently confirmed, via procedures
deemed satisfactory to the scientific community, use those theories to devise and
build instruments that detect and measure the very entities they had earlier
postulated.4
To return to the main question, what can we say about electrons? They can’t be
‘seen’ at all, though we can observe their effects – in a cloud chamber, for
example. Does that make them real or fictional? Neutrons do not produce tracks in
a cloud chamber, so their presence has to be inferred. Is this the point at which we
draw the line between direct and indirect observation, and by implication between
real and fictional? Or is the line to be drawn at photons, which are in principle
unobservable? If the demarcation between directly observable and indirectly obser-
vable (and unobservable, for that matter), is not clear, are instrumentalists also
unclear about what is real and what is fictional? In practice, there is a continuum
between observable and unobservable, via various degrees of indirect observation,

107

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
CHAPTER 6

as increasingly sophisticated instrumentation intervenes between the observer and


the observed. Does this mean that there is a gradual transition from real to fictional
status? Is a protein molecule or a virus ‘seen’ with an electron microscope less
real than a synthetic polymer molecule or a bacterium ‘seen’ through an optical
microscope?
One more point is worth making. If human faculties had been different,
the boundary between what is unobservable and observable, or unperceivable/
perceivable using other senses, would have been different. Would this entail a
different set of realities? For a realist, it is unreasonable to draw the line between
real and fictional at a point that is an accidental consequence of our particular
physiological make-up. It also raises questions about species with radically different
sense organs from our own. What for us is a fiction may be real for them. The
endpoint of this kind of speculation is that our inability to observe is no guarantee
that an entity doesn’t exist.
Even if we do reduce theoretical entities to mere devices for prediction the fact
still remains that something causes things to behave as they do. There is ‘some-
thing out there’. Some things are real: you and I, for instance; this book; that table.
It is also true that some things are heavier than others, or yellower, hotter, further
away and moving faster than others. It is not all a matter of how we choose to
describe the world, how we elect to see it. The real world plays some part in that
decision about how to describe it. So we may as well try to figure out what it is.
I am arguing that science is one of the ways in which we attempt to find out what
that ‘something’ is. Scientists are not concerned with mere prediction, they are
concerned with what is. In a sense, instrumentalists have the cart before the horse:
the capacity to make successful predictions is not the driving force for doing
science, it is the consequence of successful attempts to build explanatory systems.
Theories work as well as they do, in terms of predictions, not because we act as if
theoretical entities exist, but because they really do exist – or close approximations
to them exist. It is hard to believe that regularities that appear on the macroscopic
level (what we see) should be just as if they were due to things on the theoretical
level. Smart (1968, p. 152) asks why sense data such as “a cat’s tail on the left of a
sofa” should be followed by sense data of “a cat’s head and whiskers on the right
of a sofa” unless there is a cat walking behind the sofa. It isn’t just as if there is a
cat walking behind the sofa, there really is a cat walking behind it. This is, in
essence, the same inference to best explanation IBE) discussed in chapter 3. As
Newton-Smith (1990) argues:
I may never see a mouse in my house. But I may infer the existence of the
mouse on the grounds that if there were a mouse that would provide the best
explanation I can think of for the disappearance of bits of food in the night,
the occurrence of scrabbling noises and the appearance of droppings. In the
scientific context, a theory, such as Thomson’s theory of the electron, is
argued to be more likely to be true than any rival theory on the grounds that it
provides the best available explanation. (p. 184)

108

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
REALISM OR INSTRUMENTALISM

A more interesting and compelling argument against instrumentalism and other


forms of anti-realism, and perhaps an easier one for school age students to follow,
is that it offers no incentive for the scientific endeavour. Once in possession of
a satisfactory explanation that adequately accounts for the data, and enables accu-
rate predictions to be made, there is no reason for an instrumentalist to seek
alternatives. Why bother to do more? Only a realist, someone whose drive is to
find out what the universe is really like, will continue the effort. In common with
Newton (1997), I would argue that most (if not all) scientists are realists5 and, not
with standing the technological benefits that sometimes accrue from advances in
science, I would argue that realism is the major driving force for science. Indeed,
it was precisely because Galileo took a realist stance towards Copernican theory,
and set about solving its problems, that progress was made. He would not have
persisted with his work unless he believed that by doing so he was getting closer to
the truth. Science is driven by a realist position. We need a realist interpretation of
phenomena if we are to make sense of the world, and making sense of the world is
the main reason why we engage in theory building.
If we have nothing but instrumental laws, they may explain in the sense of
enabling us to predict, but they don’t explain in the sense of reducing the
brutishness of brute facts. (Smart, 1968, p. 152 )
The realist goal also underlies the drive for theory unification. Two theories may
be very successful in their own domains but in conflict from a realist perspective –
for example, quantum mechanics and general relativity. The strenuous effort to
produce a single unified theory signals very clearly that the scientific community
(or, at least, the sub-group of theoretical physicists) has a realist outlook. If the goal
of science was merely predictive there would be no reason to proceed beyond
existing theories that work perfectly well in an instrumental capacity.

WHY BELIEVE IN MIRACLES?

Instrumentalism cannot easily explain why theories are successful and, in parti-
cular, why they have predictive capability, nor how an idea introduced expressly to
solve a particular problem can lead to the solution of quite different problems. If
scientific theories are not (approximately) true, how is it possible for them to solve
diverse problems and to yield such accurate (and sometimes very surprising)
observational predictions? Surely, the argument goes, the continuing success of a
particular theory in solving both theoretical and practical problems, its ability to
provide common explanations for diverse phenomena, and its capacity to grow,
develop and extend its domain, are good grounds for adopting a realist position. In
other words, its success can be taken as evidence that its ‘hard core’ is approxi-
mately true. How can a theory be no more than a fictional calculating device if it
can predict real and surprising phenomena such as the bending of light in the
vicinity of an intense gravitational field, and with surprising regularity? How can
instrumentalist theories, which are supposed to be mere calculating devices, lead to
the discovery of new phenomena, using concepts that are supposedly no more than

109

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
CHAPTER 6

theoretical fictions? In Putnam’s (1975) words, “the positive argument for realism
is that it is the only philosophy that doesn’t make the success of science a miracle”
(p. 73), a view echoed by Karl Popper (1974): “it would be a highly improbable
coincidence if a theory like Einstein’s could correctly predict very precise measure-
ments not predicted by its predecessors unless there is ‘some truth’ in it” (p. 1193).
Van Fraassen’s counter-argument is that predictive success is simply a consequence
of the theory having been designed to account for observable phenomena and
events. It doesn’t mean that the theory is true. As the history of science shows, we
should be very careful about assuming our theories are true just because they fit the
data; many people have assumed that in the past, only to be proved wrong.
The success of current scientific theories is no miracle. It is not even sur-
prising to the scientific (Darwinist) mind. For any scientific theory is born
into a life of fierce competition, a jungle red in tooth and claw. Only success-
ful theories survive – the ones which in fact latched on to actual regularities
in nature. (van Fraassen, 1980, p. 40)
Van Fraassen proceeds to argue that the success of a theory in providing
explanations, making predictions and enhancing our ability to manipulate objects
in the real world is a good reason for accepting and using a theory, but not a good
reason for believing it (accepting it as true), on the grounds that “credibility varies
inversely with informativeness” (p. 280). The more informative the theory, the more
opportunities there are for it to be false. In other words, the more a theory claims,
the less likely it is to be true. He concludes, “I assume that no one can coherently
call one hypothesis less likely to be true than another while professing greater
credence in it” (p. 294).
It is much more difficult for instrumentalism to account for the embarrassing
occasions when, as a consequence of improved instrumentation, a once purely theor-
etical entity (a fiction) becomes an observable one (and, therefore, real). It cannot
readily account for those occasions when science creates phenomena and events
that do not exist in the natural world, especially when they go on to form the basis
of significant technological artifacts that give us the ability to interact with the real
world in powerful new ways. Arguing this line, Hacking (1983, 1991) regards
engineering success as the strongest case for realism. Perhaps electrons were once
merely theoretical fictions, but now that we can manipulate them in a controlled
way and use them to bring about effects in something else we are justified in
believing that they are real.
The best kinds of evidence for the reality of a postulated or inferred entity is
that we can begin to measure it or otherwise understand its causal powers.
The best evidence, in turn, that we have this kind of understanding is that we
can set out, from scratch, to build machines that will work fairly reliably, taking
advantage of this or that causal nexus. Hence, engineering, not theorizing, is
the best proof of scientific realism about entities. (Hacking, 1991, p. 258)
However, Fine (1986) argues that this instrumental success is hard-earned and
less convincing of the realist status of theories than supporters would claim.

110

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
REALISM OR INSTRUMENTALISM

Even when success appears at the end of the road, it generally crowns a long
history of frustration and failure… an enormous amount of plain old trial and
error… I think a reasonable historical picture would be to draw each success
as sitting on top of a great mountain of failures. In inviting us to explain the
instrumental success of science, the realist performs a sort of conjuring trick,
and directs our line of sight only along the successful tips of the mountains of
failures. (p. 152)
Nevertheless, if scientists act as if theoretical entities exist, and that supposition
proves robust (i.e., things behave as scientists had supposed), there must be
something real about that entity. While knowledge of their behaviour is some
evidence of their existence, it is the power to use unobservable entities that justifies
our belief in the reality of what we manipulate. To paraphrase Ogborn (1995), elec-
tromagnetic waves were once a theoretical construction of Maxwell’s, now they are
used to broadcast television programmes; genes were once a speculation of Mendel’s,
now they are a tool for producing genetically engineered organisms. Hacking (1983)
sums up this realist argument with the simple, down-to-earth statement that if there
are “standard emitters with which we can spray positrons and electrons… then they
are real” (p. 24). For Giere (1988), engineering is the driving force for science: “The
development of science depends at least as much on new machines as it does on
new ideas” (p. 138). He argues that entities like protons and electrons, once regar-
ded as highly theoretical and problematic, have been ‘tamed’ and ‘harnessed’ by
technology and are now research tools for investigating entities that are still pro-
blematic, such as gluons.
George Thomson (1965) claims that the discovery of new effects and new pheno-
mena is one of the aims (perhaps the major aim) of theory-building. It is difficult to
see how an instrumentalist view of science can meet this aim. A ‘theoretical fiction’
is acceptable if it accounts for the phenomenon it is designed to explain, it need not
go further. Indeed, it cannot go further unless it was designed to do so. As Popper
(1963) says, “If theories are instruments for prediction, then we must assume that
their purpose must be determined in advance, as with other instruments” (p. 118).

CHOOSING BETWEEN REALISM AND INSTRUMENTALISM

It is necessary to ask whether there a major problem for realists in accounting for
theory change. Not only are theories sometimes refuted, but also the very entities
that comprise them (phlogiston and the notion of the aether, for example) are
discarded. Anti-realists argue that empirical success and predictive capability is no
guarantee that the particular entities comprising a theory actually exist. Indeed,
Laudan (1981) points to a long list of now abandoned theories that once had pre-
dictive power, including the crystalline spheres of ancient and medieval astronomy,
the phlogiston theory of burning, the caloric theory of heat and various theories of
spontaneous generation. Clearly, we no longer believe that the entities in these
theoretical structures exist. Therefore, he argues, we have good reason to believe
that the entities we currently accept as real will “go the same way” as phlogiston
and caloric – that is, into the dustbin of bad or discarded scientific ideas. However,

111

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
CHAPTER 6

theory replacement is not a problem for realists if these events are regarded as
occasions when our theoretical understanding is brought closer into conformity
with the real world, as occasions when science is learning from its mistakes and
getting closer to the truth by using both observable and unobservable (theoretical)
entities as tools for thinking. The fact that theories have anomalies, and that from
time to time we abandon entire theories (and even the entities that comprise
them), tells us that there is ‘something out there’ that is different from out current
conception of it. Our view of the precise nature of that ‘something’ is always subject
to change and revision. This is the impetus for further study, further investigation,
further theory building. Insofar as they are prepared to conjecture that their theor-
etical entities exist in the physical world, realists are speculative and bold, whilst
instrumentalists are cautious and defensive.
While it is fair to say that knowledge of the mechanisms of real events is pro-
duced by scientific practice, it is not true to say that the actual mechanisms are
so produced. The world behaves in particular ways, independently of us and our
thoughts about it. While instrumentalism is content to allow scientific concepts and
theories to be regarded as artificial constructs that enable us to impose order on the
physical world, realism asserts that it is because the physical world is ordered
that science becomes possible, and order can be perceived. Science attempts to
ascertain the mechanisms underlying events in the real world. The nature of the
physical world can only be known from a study of science, but its nature is not
determined by science (as social constructivists would have us believe). There is a
reality, which is largely unknown, but this reality is, at least in part, knowable. This
view of science can accommodate the changing nature of scientific theory with
the ‘unchanging’ nature of the physical world it attempts to explain.6 It is well
described by Bhaskar (1975).
The causal structures and generative mechanisms of nature must exist and act
independently of the conditions that allow men access to them, so that they
must be assumed to be structured and intransitive, i.e., relatively independent
of the patterns of events and the actions of men alike… Structure and
mechanisms then are real and distinct from the patterns of events that they
generate; just as events are real and distinct from the experiences in which
they are apprehended… The ultimate objects of scientific understanding are
neither patterns of events nor models but the things that produce and the
mechanisms that generate the flux of the phenomena of the world. Scientists
attempt to discover the way things act, a knowledge typically expressed in
laws; and what things are, a knowledge… typically expressed in real defini-
tions. (pp. 56 & 66)
This scientific knowledge is not constructed ‘from scratch’ each time; rather,
new knowledge is built from and by means of existing knowledge. The more
science speculates and investigates, the more it knows and the more it is able to
know. But science has to start somewhere, even though we cannot be certain of the
‘reality’ of the starting point (i.e., the validity of our initial knowledge). Nash
(1963) coins an interesting phrase when he talks about the principle of corrigible
fallibility.

112

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
REALISM OR INSTRUMENTALISM

However we conceive science, we cannot begin our work without some facts
and some ideas provisionally accepted as unchallengeable. But the ideas may
be quite wrong; the facts ill-observed or laden with ‘credulity’. Obsessed by
such fears we don’t do science! From such obsession scientists are freed by
what I call the principle of corrigible fallibility. It is a principle of action.
Beginning with the best available facts and ideas, we proceed vigorously in
the faith that any errors in them will be revealed by the interaction of facts
and ideas – by the interaction of rational and empirical elements, in neither of
which we have, or can have, absolute confidence. We amend out hypotheses
in the light of our experiments; but we also reject (as errors), ‘correct’, and
explain away some of our data when they conflict with ‘indubitable principles’.
In such unquestioning acceptance of principles to which all experience is
made to conform… science may seem at one with divination, or magic. Yet
science progresses, as magic does not, and not simply because science had
the good fortune to hit on the ‘right’ principles at the outset. It did not! But it
learned better principles. (p. 83)
It could be argued that the objections to a realist interpretation of Copernican
theory were twofold. First, the theory lacked sufficient factual support of its own,
while Ptolemaic theory could account for all that Copernican theory claimed to
explain. Second, it was inconsistent with certain observations and with well-
confirmed physical theory (Aristotle’s). The conservative response would have
been to retain Ptolemy’s epicycles; they were superior in the instrumentalist sense
of making better predictions. However, the history of science shows us that
inconsistency with other existing theory is not sufficient grounds for a realist to
dismiss a new theory. Indeed, it was precisely because a realist interpretation of
Copernican theory was adopted that a new and better dynamics was developed.
Without a realist drive there would have been no point in seeking to improve on the
predictive capability of Ptolemaic epicycles. Realism provides the incentive for the
development of better articulated theories, new theories and better instruments for
observation. Before Copernican theory could fully account for the observational
data a number of refinements were necessary: Keplerian elliptical orbits (rather
than circular ones), planets of significantly different masses in mutual attraction,
planetary satellites, spinning planets. Eventually, the theory was capable of explain-
ing the ‘facts’, but the imperfect theory was retained throughout - despite the poor
supporting evidence. Why? Because scientists really believed they were in pursuit
of the truth and retained the theory despite its shortcomings. In Feyerabend’s (1964)
words, “the realistic position encourages research and stimulates progress, whereas
instrumentalism is more conservative and therefore liable to lead to dogmatic
petrifaction” (p. 302).
Of course, because scientific knowledge is produced in a social context, the goal
of ascertaining ‘how the world is’ (the realist goal) is bound to be impacted to
some extent, and in varying degrees from theory to theory, by the personal and
professional goals of scientists, the interests and priorities of funding agencies, and
the cluster of economic, political and moral-ethical influences that impregnate the

113

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
CHAPTER 6

sociocultural context in which scientific practice is located. These matters will be


addressed in chapters 7 and 8.

MODELS, THEORIES AND CRITICAL REALISM

Speculating further on why we accumulate scientific knowledge may throw up a


number of responses.
– To overcome fear – the pursuit of rational explanations is preferable to super-
stition.
– To reduce uncertainty – making the world (or our perception of it) more stable,
more predictable and less disconcerting.
– To understand – to satisfy our curiosity by gaining a sense of ‘what lies behind
things’.
– To solve problems – especially problems relating to food, shelter, health, trans-
port, and the like.
– To create artifacts – tools to make life easier, toys to make life more pleasurable.
– To create ‘a better life’ – a somewhat vague but still powerful motivator for
many people.
Perhaps the status of scientific knowledge is related to our purpose in seeking it,
to the role it plays once it has been acquired. Science is, of course, highly valued
for its practical achievements and for its material benefits. In other words, it has
instrumental value. But, according to Popper (1963), it is valued more for its ability
“to free our minds from old beliefs, old prejudices, and old certainties, and to offer
us in their stead new conjectures and daring hypotheses. Science is valued for its
liberalizing influence – as one of the greatest of the forces that make for human
freedom” (p. 102). Science liberates because scientists dare to create theories in
contradiction to everyday experience, dare to go beyond the world of the senses.
Indeed, much of scientific theory is counter-intuitive and counter to everyday
commonsense. Eddington (1928) provides a wonderful example in his description
of two tables.
One of them has been familiar to me from earliest years. It is a commonplace
object of that environment which I call the world. How shall I describe it?
It has extension; it is comparatively permanent; it is coloured; above all it
is substantial… Table No. 2 is my scientific table. It is a more recent
acquaintance and I do not feel so familiar with it. It does not belong to the
world previously mentioned – that world which spontaneously appears
around me when I open my eyes, though how much of it is objective and how
much subjective I do not here consider. It is part of a world which in more
devious ways has forced itself on my attention. My scientific table is mostly
emptiness. Sparsely scattered in that emptiness are numerous electric charges
rushing about with great speed; but their combined bulk amounts to less than
a billionth of the table itself. (pp. ix & x)
Science liberates because scientists dare to challenge common sense appearances,
received knowledge and the dictates of the state and organized religion. By trying

114

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
REALISM OR INSTRUMENTALISM

to explain the regularities deduced from their theoretical speculations, scientists


explain the known by the unknown. Sometimes scientists aim at a true description
of the world and a true explanation of observable facts (a description of these facts
must be deducible from the theory).7 On other occasions they are not seeking a
‘true’ description of the world; rather, a convenient predictive instrument is all that
is required. When they wish to explain ‘how things are’ in the universe, when they
are trying to make sense of it, scientists develop theories – our current ‘best shot’
at the truth. Each theoretical development takes us a little nearer to the truth about
the universe, though we recognize that we are unlikely ever to achieve that goal
and wouldn’t know if we had. When we simply want ‘to get the job done’ (make a
prediction, achieve a measure of control, etc.) we invent theoretical models.
Sidney Morgenbesser (1969) asserts that “a scientist is justified in using a theory
T to accomplish a given end, if he has good reasons for believing that his theory-
guided act will accomplish his end” (p. 213). I take this to mean that when scientists
wish to predict an event they may employ an instrumental model, but when they
wish to explain they must use a realist theory, and that the choice depends solely
on their particular purpose at the time. Thus, the view I wish to promote for school
science is that scientists play both a realist game and an instrumentalist game, as
determined by their immediate purpose: they develop theories when they aim to
explain and describe the real world; they use convenient models when they wish
for no more than a quick and accurate calculation or prediction. This position,
which some have termed critical realism (Jacoby & Spargo, 1992), enables scientists
to be realist about some theories (those they consider to be genuine attempts to
uncover the truth) and instrumentalist about others (those they find useful but do
not accept as true descriptions). I believe there is enormous value in referring to the
former as theories and the latter as models. In contrast, instrumentalists are always
instrumentalist and blur the distinction between theory and model.
It is interesting that a model originally designed simply to achieve a short-term
instrumentalist goal may sometimes be developed, over time, into a realist theory.
Sometimes entities that we ‘invent’ to gain a measure of control turn out to have a
real existence. Smart (1968) argues that while many contemporary theories are
instrumental, they are moving towards realism.
Modern physics is instrumentalist, but may nevertheless constitute an approxi-
mation to, or a hint at, a realist theory. That is, a realist theory, even though it is
unknown to us as yet, is ‘in the offing’. For if this were not so, how could
the success of an instrumentalist theory be explained? Otherwise the success
of the instrumentalist theory would have to depend on too many mere
coincidences at the macroscopic level. (p. 163)
It seems that models often function as the means of building progressively more
complex (and realist) theories. They function as ‘scaffolding devices’, by means of
which scientists make interim sense of observational data as they develop and
extend their thinking. They play a key role in thought experiments and in the design
and conduct of real experiments. Like scaffolding, models may be discarded once
they have served their function. Indeed, it may be that some models used early in a

115

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
CHAPTER 6

theory’s developmental history prove incompatible with the premises of the final,
refined theory despite their pivotal role in its construction. As observational
support for the once tentative theoretical model increases, and it resists continued
criticism and efforts to falsify it, it becomes accepted as the new theory (the new
paradigm) and enters what Kuhn (1970) calls the phase of normal science, during
which it is developed and refined (see chapter 4). Within normal science activities
scientists may use a range of models in problem solving, prediction and further
theory articulation – e.g., the so-called ‘billiard ball model’ to represent collisions
of gas molecules. Indeed, within all significant theoretical frameworks there is
a “family of models” (Giere, 1988, 1999) that scientists may deploy for solving
short-term, practical problems, making calculations and predictions. The full
theoretical force of the theory does not have to be engaged if instrumental models
within it can suffice. Eventually, when scientists encounter insurmountable prob-
lems with the theory, find severe limitations to it, or construct a theory that is more
satisfying, the old theory is rejected and reverts to the status of a model. It may still
be useful for making predictions, but we no longer believe that it explains matters.
From the critical realist position, it is not illogical to retain a falsified or superseded
theory in an instrumental capacity, provided that its status is recognized and
acknowledged. It may be that within a restricted domain of application, a falsified
theory (in its new status as model) is more useful than the currently accepted theory
because it is easier to use. For example, in many instances calculations using New-
tonian physics give the same numerical results as would relativity or quantum
mechanics if applied to the same problem, but the calculations are much easier! In
a sense, school science does this all the time, and certainly in physics, where most
of what is taught is pre-20th Century physics and a long way from the ‘cutting
edge’ of contemporary research.
Within the critical realist position it is not illogical to utilize alternative instru-
mental models, even incompatible or contradictory ones, to deal with different
aspects of the same phenomenon. As Morgenbesser (1969) comments, “a scientist
may be justified in using T1 for predictive or calculating purposes and using T2 for
related ones though their conjunction is self-contradictory” (p. 213). For example,
we can use a wave model of light to solve problems relating to interference patt-
erns and a particle model of light to solve a problems relating to photoelectric
effects. The use of diverse and sometimes mutually contradictory models is fairly
common practice in engineering, along with the use of models that function only
within a narrow range of conditions. According to Cartwright (1983), we design
models according to the specific features of the situation we are interested in, with-
out worrying too much about whether the model is compatible with another model
we might use for other purposes. In exemplification, she cites a text on quantum
optics in which various and mutually incompatible models are deployed to explain
the various properties of lasers. She also notes that the need for idealized (fictional)
models is greatest when scientists wish to use sophisticated mathematics.
In general, nature does not prepare situations to fit the kinds of mathematical
theories we hanker for. We construct both the theories and the objects to
which they apply, then match them piecemeal onto real situations, deriving –

116

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
REALISM OR INSTRUMENTALISM

sometimes with great precision – a bit of what happens, but generally not
getting all the facts straight at once. The fundamental laws do not govern
reality. What they govern has only the appearance of reality and the appearance
is far tidier and more readily regimented than reality itself. (p. 162)
My concern is that we often confuse students by not distinguishing sufficiently
between different kinds of theoretical structures. Indeed, most curricula are silent
on this matter. We leave students to form their own views about whether science
has a realist or an instrumentalist thrust. Little wonder that many students are con-
fused about the status of scientific knowledge. Some may come to believe that
some of the strange calculating and predicting devices, that I would urge teachers to
call models, are intended to explain - i.e., that they represent what scientists believe
about real events and phenomena. Others may simply conclude that all scientific
explanations are as fanciful as these models, and for them, science will lose its
credibility. It is worth remembering the constructivist mantra: ‘new ideas will only
be accepted by students if they perceive them to be intelligible, plausible and fruitful’
(Posner et al., 1982). Theories needs to be intelligible and plausible, and fruitful in the
sense of being able to explain; models need to be intelligible and fruitful, though not
necessarily plausible. In other words, models may fail the plausibility test even
though they are excellent in terms of fruitfulness. There is, of course, an important
link between the use of models in science for theory building and problem solving
and the use of models by students and teachers in learning scientific concepts and
acquiring new theoretical knowledge. Although these matters fall outside the scope
of this book, it is important to note that the nature of mental models has long been
an area of research in cognitive psychology, dating back to the seminal work of
Johnson-Laird (1983) and Gentner and Stevens (1983). In recent years, the topic of
models and modelling has generated considerable interest among science educators
(Gilbert et al., 1998a,b; Franco et al., 1999; van Driel & Verloop, 1999; Gilbert &
Boulter, 2000; Greca & Moreira, 2000, 2002; Coll & Treagust, 2002, 2003a,b;
Justi & Gilbert, 2002a,b,c, 2003; Treagust et al., 2002; Davies & Gilbert, 2003;
Taber, 2003; Gilbert, 2004; Kowasaki et al., 2004; Coll & Taylor, 2005; Coll et al.,
2005; Justi & van Driel, 2005), an interest that can be categorized into three prin-
cipal areas of concern:
– The particular models and theories produced by scientists as explanatory
systems, including the history of their development.
– The ways in which scientists utilize models as cognitive tools in their day-to-
day problem solving, theory articulation and theory revision.
– The role of models and modelling in science pedagogy.
Of particular interest to the substance of this book is the diversity among
scientists in terms of how they utilize models as tools for thinking (Hesse, 1970;
Gilbert & Mulkay, 1984; Giere, 1988; Stavy, 1991; Nersessian, 1995; Ogborn &
Martins, 1996; Penner et al., 1997; Ault, 1998).

117

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
CHAPTER 6

WHAT IS REAL?

While I am content that the critical realist perspective is a productive one for
school science, and a considerable improvement on simply leaving students to
work it out for themselves, there remains a problem. Which theoretical entities
should be accepted as real and which are to be regarded as fictional? Sometimes it
is easy to decide, but in the ‘cutting edge’ physics to which Smart refers in the
passage quoted earlier it may be more problematic. As discussed earlier in the
chapter, Hacking (1983, 1991) adopts the position that theoretical entities are real
if we can use them to bring about significant changes in a predictable way. He also
states that the confidence we have in the existence of particular entities (because
of our ability to manipulate them) does not necessarily extend to the complex
theoretical structures that use these entities in explanation of the processes under-
pinning observable phenomena – a position also taken by Cartwright (1983) and
Giere (1988). Thus, we can be realist about some of the entities in a theory even if
the theory is not yet regarded as a true explanation of events. For example, we can
believe in the existence of atoms but regard the accounts of crystallographers about
how the atoms are arranged within particular structures as uncertain. Moreover,
we can continue to believe in the existence of entities even when a theory that
employs them is known to be false or has been superseded. Furthermore, as argued
previously, scientists may simplify and idealize a realist theory into a range of
instrumental models in order to utilize it more effectively in solving a range of
problems or disposing of an anomaly. In other words, a theory may be regarded
as comprising a family of models. In Kuhn’s (1970) account of science, models
are constantly evaluated for their effectiveness in dealing with the paradigm’s
problems (and those which prove ineffective are discarded), but theories are only
evaluated in times of crisis, and only against a competing theory.
One way of dealing with the dilemma about precisely what should be regarded
as real is to draw a distinction between the statements ‘theoretical entities are
real’ and ‘theoretical entities exist’. There is a distinction between our conceptual
systems, which are human constructs and are susceptible to change, and the actual
world, to which our conceptual systems bear some relation. The physical world
is real and scientific theories are real, but they are not identical (there is an
ontological distinction between them). Dragons and phlogiston are both real – they
have properties and particular relationships with other concepts that are also well
understood (such as fair maidens and burning) – but they do not exist. Other non-
existing but real entities include unicorns, hobbits, leprechauns and taniwha.8
Similarly, electrons and magnetic fields are real concepts, having an existence
independent of individuals and having relationships with other concepts whether
or not anyone is aware of that relationship. But in these cases we do not know
whether they exist or not. We believe that they do, and act as if they do, but a
change in theory may lead us to change our view in the same way that a change in
theory led us to change of views regarding the existence of dragons and phlogiston.
In a sense, all knowledge is real and has an existence independent of the opini-
ons and feelings of individuals, though individuals are, of course, necessary for the
initial generation and development of knowledge. Particular relationships among

118

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
REALISM OR INSTRUMENTALISM

the ideas that comprise a complex matrix of scientific knowledge exist, and the
entities within the theory relate to the available evidence in particular ways, whether
or not any individual accepts the theory as a true account of events or believes the
entities contained in the theory actually exist in the real physical world. In other
words, they have an objective existence. ‘Objective’ is used here in contrast to
‘subjective’ and does not necessarily mean ‘true’. A theory can be objective and
false – for example, phlogiston theory is not a true account of combustion, but its
concepts and principles have an objective existence and can be critically addressed.
The most complete description of this view of scientific knowledge is Karl Popper’s
Objective Knowledge, published in 1972, in which he postulates the existence of
three distinct ‘worlds’: the actual physical world (world 1), the world of human
thought processes (world 2) and the world of objective knowledge (world 3).
In this pluralistic philosophy the world consists of at least three ontologi-
cally distinct sub-worlds; or, as I shall say, there are three worlds: the first is
the physical world or the world of physical states; the second is the mental
world or the world of mental states; and the third is the world of intelligibles,
or of ideas in the objective sense; it is the world of possible objects of
thought; the world of theories in themselves, and their logical relations; of
arguments in themselves; and of problem situations in themselves. (p. 154)
The third world (the product of human activity) is separate from but related to
the second world (the subjective act of thinking) which produced it in just the same
way as a spider’s web is separate from but related to the spider’s act of spinning it.
Scientific knowledge is created by people, but once it has been created it is
independent of its creator and has properties which can be studied without regard
to its origin. It may even have consequences unforeseen by those who created it;
there may be both internal relationships and external relationships (to knowledge in
other theories) that remain undiscovered for some time. Interaction between the
physical world and the theoretical world is via individual consciousness, which is
firmly anchored to and influenced by relationships in the world of theories. The
world of objective knowledge (World 3) is not an exact account of the actual
physical world (World 1); rather, it is our current ‘best shot’ at describing it. These
views will change as science improves the match between World 1 and World 3
through a combination of theorizing, speculating, experimenting and arguing. This
position is well captured by Bohm (1957):
The world as a whole is objectively real, and that as far as we know, it has a
precisely describable and analyzable structure of unlimited complexity. This
structure must be understood with the aid of a series of progressively more
fundamental, more extensive, and more accurate concepts, which series will
furnish, so to speak, a better and better set of views of the infinite structure of
objective reality. We should, however, never expect to obtain a complete
theory of this structure, because there are almost certainly more elements in
it than we can possibly be aware of at any particular stage of scientific
development. (p. 100)

119

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
CHAPTER 6

Figure 6.1. Popper’s ‘three worlds’ view

Popper’s position stands diametrically opposed to the views of radical


constructivists such as Ernst von Glasersfeld (1987, 1989, 2007), who asserts that
knowledge exists only in the minds of cognizing beings (it cannot reside in books,
for example) and so will vary substantially from individual to individual, even when
addressing the same phenomenon. Scientific knowledge, like any other knowledge,
is simply that which individual ‘epistemic agents’, given their particular experiences
and traditions of thought and language, consider ‘viable’ for them.9
Knowledge does not reflect an ‘objective’ ontological reality, but exclusively
an ordering and organization of a world constituted by our experience (von
Glasersfeld, 1987, p. 199)

The basic elements out of which an individual’s conceptual structures are


composed and the relations by means of which they are held together cannot
be transferred from one language user to another… they must be abstracted
from individual experience (von Glasersfeld, 1989, p. 132)
Returning to Popper’s ‘three worlds view’, it follows that the terms in which
individual scientists think (the way they ‘see’ the world) depend on the particular
selection they make from the objective third world of theories. They make this
selection in response to a complex interaction of personal beliefs and values,
impregnated with social norms, pressures and influences. And, in turn, the parti-
cular selection made by an scientist exerts a powerful influence on the kind of
science that is done and the way in which it is done, and strongly influences the
kind of knowledge that is generated. It is the cluster of social influences on science
– internally, with respect to the day-to-day practice of scientists; externally, in
terms of the ways in which science impacts (and is impacted by) the sociocultural
context in which it is located - that is the principal concern of chapter 7.

120

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
REALISM OR INSTRUMENTALISM

ENDNOTES
1
The term ‘antirealist’ may be used to refer to advocates of a number of philosophical positions
opposing realism, including instrumentalists, idealists, phenomenalists, empiricists, conventionalists,
constructivists and pragmatists (Fine, 1991). Instrumentalism and constructivism are the two
antirealist positions most relevant to the discussions in this chapter.
2
While instrumentalists may regard these entities as fictions, they are fictions that are purposefully
created; they are not mere flights of fancy. Hence, utilizing an invented concept like gene and
building it into a complex explanatory structure, for example, is not to be regarded as merely an
intellectually satisfying scientific fantasy.
3
This position, adopted by most contemporary sociologists of science, will be elaborated in chapter 7.
4
Of course, if the theoretical assumptions underpinning the design and construction of particular
scientific instruments are wrong or merely incomplete, then the data are spurious and the existence
of the entities that they appear to confirm is in serious doubt.
5
Wong and Hodson (2008a,b) provide some compelling evidence for this claim from interviews with a
number of leading scientists.
6
‘Unchanging’ is used here in a relative sense. Of course, the universe is constantly changing, although
within the normal lifetime of human beings there is no substantial change at the gross level.
7
‘True’ is being used in the somewhat restricted sense of scientifically true: not yet falsified, even by
the most rigorous tests.
8
Taniwha: mythical water-living beast in Maori folklore. Taniwha stories are used to ‘persuade’
young children to avoid dangerous places such as river estuaries and tidal caves.
9
A detailed critique of von Glasersfeld’s views can be found in Suchting (1992), Matthews (1993,
1998a) and Kelly (1997), with a response in von Glasersfeld (1992). A special issue of Science &
Education (1997, 6(1-2)) includes ten articles and a lengthy bibliography dealing with philosophical
aspects of constructivism.

121

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
Derek Hodson - 978-90-8790-507-1
Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
CHAPTER 7

INSIGHT FROM THE SOCIOLOGY OF SCIENCE


Science is What Scientists Do

It has been argued a number of times in previous chapters that empirical adequacy
is insufficient, in itself, to establish the validity of a theory: consistency with the
observable ‘facts’ does not mean that a theory is true,1 only that it might be true,
along with other theories that may also correspond with the observational data.
Moreover, empirical inadequacy (theories unable to account for all the ‘facts’ in
their domain) is frequently ignored by individual scientists in their fight to establish
a new theory or retain an existing one. It has also been argued that because experi-
ments are conceived and conducted within a particular theoretical, procedural
and instrumental framework, they cannot furnish the theory-free data needed
to make empirically-based judgements about the superiority of one theory over
another. What counts as relevant evidence is, in part, determined by the theoretical
framework the evidence is intended to test. It follows that the rationality of science
is rather different from the account we usually provide for students in school.
Experiment and observation are not as decisive as we claim. Additional factors that
may play a part in theory acceptance include the following: intuition, aesthetic
considerations, similarity and consistency among theories, intellectual fashion,
social and economic influences, status of the proposer(s), personal motives and
opportunism.
Although the evidence may be inconclusive, scientists’ intuitive feelings about
the plausibility or aptness of particular ideas will make it appear convincing. The
history of science includes many accounts of scientists ‘sticking to their guns’
concerning a well-loved theory in the teeth of evidence to the contrary, and some-
times in the absence of any evidence at all. Marton et al. (1994) surveyed eighty-
three Nobel laureates in physics, chemistry and medicine about the role of intuition
in their research. Seventy-two were in no doubt about its importance. Michael
Brown, joint winner with Joseph Goldstein of the 1985 Nobel Prize in medicine
for their work on cholesterol metabolism, commented “As we did our work… we
would go from one step to the next, and somehow we would know which was the
right way to go. And I can’t really tell how we knew that” (Marton et al., 1994,
p. 461). Rita Levi-Montalcini (1986 winner, with Stanley Cohen, for their dis-
covery of growth factors) said: “Intuition… is something unconscious, which, all
of a sudden, comes out of a clear sky to you and is absolutely a necessity, more
than logic… You’ve been thinking about something… for a long time… then all
of a sudden, the problem is opened to you in a flash, and you suddenly see the
answer” (pp. 462 & 465) and Konrad Lorenz (joint winner in 1973, with Karl von
Frisch and Nikolaas Tinbergen, for work on animal behaviour) remarked: “[You
keep] all known facts afloat, waiting for them to fall into place, like a jigsaw
puzzle… If you try to permutate your knowledge, nothing comes out of it. You

123

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
CHAPTER 7

must give a sort of mysterious pressure, and then rest, and suddenly BING… the
solution comes” (p. 467). Perhaps, as so often, Albert Einstein (1918/1954) says
it best:
The supreme task of the physicist is to arrive at those universal elementary
laws from which the cosmos can be built up… There is no logical path to
these laws; only intuition, resting on sympathetic understanding of experience,
can reach them (p. 221)
Elegance, simplicity and parsimony can be significant factors in gaining support
for a theory. As Martin Amis (1995) says, in his novel The Information, “In the
mathematics of the universe beauty helps tell us whether things are false or true”
(p. 8), while Richard Feynman (1965) remarked that “You can recognize truth by
its beauty and simplicity… When you get it right, it is obvious that it is right. The
truth always turns out to be simpler than you thought” (p. 171). Similarly, in a
conversation with Einstein, Werner Heisenberg (1971) argued that the simplicity
of good ideas is a strong indication of their truth: “I believe, just like you, that
the simplicity of natural laws has an objective character, that is not just the result
of thought economy. If nature leads us to mathematic forms of great simplicity
and beauty… we cannot help thinking that they are ‘true’, that they reveal a
genuine feature of nature” (p. 68). Max Born (1924) argues in similar vein when he
says that relativity theory was accepted long before supporting experimental/
observational evidence became available because it made science “more beautiful
and grander”. Also commenting on the elegance of Einstein’s work, Paul Dirac
(1980) states –
Anyone who appreciates the fundamental harmony connecting the way Nature
runs and general mathematical principles must feel that a theory with the
beauty and elegance of Einstein’s theory has to be substantially correct…
One has a great confidence in the theory arising from its great beauty, quite
independent of its detailed successes… One has an overpowering belief that
its foundations must be correct quite independent of its agreement with
observation”. (p. 44)
In another essay, Dirac (1963) says, “It is more important to have beauty in
equations than to have them fit experiments” (p. 47). Miller (2006) argues that it
was Dirac’s insistence on beauty at the expense of ‘facts’ that led to the discovery
of antiparticles. Referring to his and Francis Crick’s elucidation of the structure
of DNA, Jim Watson (1980) reports that Rosalind Franklin accepted the fact that
the structure was “too pretty not to be true” (p. 124). More extensive discussion of
the role of aesthetic criteria in science can be found in McAllister (1996).
A new theory is more likely to be accepted when it is consistent with other
well-established theories and is less likely to be accepted when it is in conflict with
them (Laudan, 1977). Thus, Copernican theory had some initial problems because
it was inconsistent with Aristotelian physics- a problem that was solved by Galileo.
Perhaps scientists have expectations of a grand unifying theory, so they look for
common explanations or common kinds of explanations.2 Holton (1981, 1986,

124

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
INSIGHT FROM THE SOCIOLOGY OF SCIENCE

1988) refers to a number of cross-disciplinary “themata”, usually in the forms of


dyads or triads, that guide the work of scientists by presenting choices between,
for example, theories based on constancy/equilbrium versus evolution versus
catastrophic change, and between driving forces such as hierarchy and unity or
reductionism and holism. Holton (1988) also notes the conscious or unconscious
preoccupation with symmetry shown by many scientists.
Trends in other disciplines (philosophy, psychology, sociology and politics, for
example) can be influential. Some theories are clearly products of their time and
the prevailing intellectual climate. For example, phrenology (Combe, 1825, 1828)
was perfectly suited to the socio-political climate of Victorian England and Wilhelm
Reich’s theorizing about orgonne energy and its impact on sexual behaviour was
similarly well-suited to the late 1940s/early 1950s (Boadella, 1985). Matters of
contemporary social and economic significance will, of course, influence research
priorities; they may also have impact on what findings are published and, in turn,
play a part in forming scientists’ views about acceptability and validity. Despite
some severe theological problems, acceptance of Copernican theory was hastened
because it solved a persistent problem associated with fixing the precise dates of
Holy Days in the Christian calendar. Dirac (1980) comments that Einstein’s theory
of relativity was largely unknown outside a small circle of scientists until the end
of the First World War, when it had enormous impact.
It came at a time when everyone was sick of the war…People wanted some-
thing new. Relativity provided just what was wanted and was seized upon by
the general public and became the central topic of conversation. It allowed
people to forget for a time the horrors of the war they had come through.
Innumerable articles about relativity were written in newspapers, magazines
and everywhere. Never before or since has a scientific idea aroused so much
and such wide-spread interest. (p. 42)
In the contemporary world, the interests of those who fund the research are seen
to impact significantly on the findings that are given prominence, though those
who are charitable may cite unconscious motivation at work, rather than vested
interest.
To be admitted to the corpus of approved scientific knowledge, theories have
to be socially, culturally and politically acceptable. In the Stalinist Soviet Union,
Darwinian ideas underpinning evolutionary biology were rejected in favour of
Lysenko’s theorizing about Lamarckism because transmission of acquired charac-
teristics was much more compatible with Marxist views about human nature, with
disastrous impact on agriculture and widespread famine (Joravsky, 1970; Lewontin
& Levins, 1976; Lecourt, 1977). Archaeologists in Nazi Germany were strongly
encouraged by the state to publish ‘findings’ indicating that the great achievements
of Classical Greece were actually the work of Germanic peoples who had migrated
south to escape a series of natural disasters (Arnold, 1990, 1992). In explanation of
these aberrations, Bloor (1974) asks us to consider “the sorts of things that people
think about… They think about objects like computers or oracles, about substances
like oxygen or phlogiston, and about states like being infected by a virus or being

125

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
CHAPTER 7

ritually unclean. Clearly peoples’ habitual categories of thought vary from culture
to culture. To these categories they apply socially-varying standards of efficiency,
cogency and satisfactoriness. The ideas that are in people’s minds are in the currency
of their time and place… The terms in which they think do not emanate from their
subjective psyches. They come from the public domain into their heads during
socialization” (p. 71).
Issues of power, influence and prestige play a major part in determining the
acceptability of theories. Who is proposing this view? Who else supports it or
accepts it? Are they prominent and important members of the community of
scientists? According to Latour and Woolgar (1979), the credibility of the proposal
and the credibility of the proposer are virtually identical, and whether a knowledge
claim is judged plausible or implausible is strongly influenced by who proposed it,
where the work was done, and how it was accomplished. In other words, the
context of the research is just as important as the content of the knowledge claim.
In Knorr-Cetina’s (1981) words, “Scientists speak about the motives and interests
which presumably gave rise to the ‘finding’, about the material resources available
to those who did the research, and about ‘who stands behind’ the results. They
virtually identify the results… with the circumstances of their generation” (p. 7).3
Evaluating the skills of researchers (their ‘track record’) proves to be a
good strategy for assessing the quality of their work (whether in accepting it
uncritically or scrutinizing it more thoroughly). No one can afford the time
or resources to assess every detail or repeat every experiment, however
ideal that may seem… Evidence may be more fundamental, but credibility
is generally primary, in the sense of being applied first. (Allchin, 2004,
pp. 942 & 943)
With regard to personal motives, an individual scientist is likely to ask: “Is this
theory likely to play a useful role in my own work?” “Will it help to solve my
particular problems?” In the group environment of a research institution, parti-
cularly if it has commercial interests, junior scientists may be strongly influenced
in their decisions by what colleagues believe (especially if there is a powerful
pressure group) or by what the Head of Department advises. Pickering (1982)
describes a situation, which he argues is not uncommon, in which the chosen
alternative (the ‘charm’ model of the properties of hadrons in the 1970s) was the
one most compatible with the research group’s existing practice and offering
opportunity for its extension. In other words, groups have a vested interest in the
continued use of approaches that have served them well in the past.
Like other professionals, scientists have to earn a living and so will seek out
research fields with potential for growth. Personal ambition and the desire for
recognition and fame will prompt some scientists to ask: “Is there a career to be
built in this area?” “Is this the kind of research that academic journals are
publishing at the moment?” “Is there likely to be a book in it?” “Is there good
conference potential here?” “Is espousing these views likely to be of value in an
interview situation or a seminar?” “Is this the kind of research for which funding
agencies are awarding substantial research grants?”

126

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
INSIGHT FROM THE SOCIOLOGY OF SCIENCE

If one takes the view that scientists are members of a community of practice,
and that the ideas of particular scientists will only become accepted as scientific
knowledge when they achieve consensus within the community, it follows that all
of the influences that impact on people in their day-to-day lives have to be consi-
dered as possible or likely influences on the conduct of the scientific community –
that is, as influences on the science that is carried out and the scientific ideas that
gain acceptance. Of course, the empirical under-determination of theories does not
prove that social and affective factors play a part in theory acceptance and theory
rejection, though it makes it much more likely.
Acknowledging this likelihood constitutes a major challenge to the much-
vaunted objectivity of science promulgated by school textbooks. By failing to
address these influences, the simple-minded accounts of theory acceptance/rejection
that we promote through the school science curriculum are insulting to students,
and often flatly contradict what they read about real scientists like Galileo Galilei,
Albert Einstein, Barbara McClintock, Rachel Carlson, Jane Goodall and Jim Watson.
What textbooks often seem to omit from their accounts of theoretical development
is the personal dimension – the ways in which the decisions and actions of scientists
are influenced by their worldviews, feelings, attitudes and prejudices. Because of
the theory-laden nature of observation (and theory-impregnated nature of all the
other processes of science) and the empirical under-determination of theories,
what one ‘sees’ as an observer, the proper conduct of experiments, the adequacy
of a theoretical explanation, and so on, are all open to dispute, contestation and
modification. Rather than attempting to reduce science to a cold, clinical, de-
personalized method, and rather than presenting science as independent of the
society in which it is located, we should be emphasizing the ways in which
knowledge is negotiated within the scientific community by a complex interaction
of imagination, experiment, theoretical argument and personal opinion. And we
should be promoting the view that science is a theory-driven and creative endea-
vour, influenced throughout by social, economic, political and moral-ethical fac-
tors as they impact on the decision makers – the ‘gatekeepers’ or guardians of the
community’s store of knowledge. As Robert Young (1987) says:
Science is not something in the sky, not a set of eternal truths waiting for
discovery. Science is practice. There is no other science than the science that
gets done. The science that exists is the record of the questions that it has
occurred to scientists to ask, the proposals that get funded, the paths that get
pursued… Nature ‘answers’ only the questions that get asked and pursued long
enough to lead to results that enter the public domain. Whether or not they
get answered, how far they get pursued, are matters for a given society, its
educational system, its patronage system and its funding bodies. (pp. 18 & 19).
Not only is science a social activity in the sense that society (writ large) deter-
mines “the science that gets done”, but also in the sense that the rules of scientific
procedure and the legitimacy of the ‘product’ are determined by a community of
practitioners. As Bereiter et al. (1997) remark, “If there is anything distinctive
about science, it is not to be found in the workings of individual minds, but in the

127

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
CHAPTER 7

way scientists conduct themselves as a community” (p. 333). Again, there are
two levels of community here. First, it is rare for scientists to work entirely alone.
Most experimental scientists in universities, research institutes and industry, are
members of a laboratory team working on closely related problems over a number
of years. Beyond the specific laboratory team an individual is also a member
of the institution’s team of scientists and of the ‘invisible team’ of scientists in
all other institutions who conduct similar and related research. Even without this
collaboration, scientists are still dependent on one another for the ‘conditions’
under which they work (including availability of ‘starting knowledge’, established
techniques, scientific apparatus, etc.). Second, the wider community of scientists
determines what counts as acceptable scientific practice and exercises strict control
of what is admitted to the corpus of accepted knowledge through its system of peer
review. This community also exercises strict control over the education of future
scientists and initiation of newcomers into the community of practice.
Consideration of the larger social context in which science is practiced,
including the ways in which the sociopolitical, economic and military significance
of science and technology lead to major concerns about control and accountability
(Dickson, 1988; Callon, 1995; Cozzens & Gieryn, 1995; Gieryn, 1995), are outside
the scope of this book, save to note that different societies, because they are located
in different physical and social environments and have different values and aspi-
rations, are likely to have different priorities for science and technology. It is
interesting to speculate on the extent to which science is culturally determined –
that is, to what extent are the questions we ask, the kind of problems we perceive
and try to solve, the ways in which we conceptualize the world, the ways in which
we think and proceed towards solution of our problems, the criteria we generate for
judging ‘success’, and the nature of the knowledge produced, a reflection of the
needs, interests, values and aspirations of society? These questions raise important
issues concerning the demarcation criteria of science. What can be changed (aims,
values, concepts, methods, criteria of validity, and so on) and the activity still be
characterized as science? What distinguishes science from pseudoscience? Do terms
such as Feminist or Gynocentric science and First Nations science or African
science have any meaning? Is the adoption of a particular worldview a sine qua
non of science or is it possible to operate within more than one worldview? These
and related matters will be addressed, albeit briefly, later in the chapter.
With regard to sociological considerations internal to the practice of science,
Ziman (1968) comments as follows: “Far from being the sum of independent, indi-
vidual researches, the continuous compilation of innumerable disconnected facts,
observations and theories, scientific knowledge is the joint social product of the
members of… ‘Invisible Colleges’” (p. 61). In other words, scientific knowledge is
the product of a complex social activity in which the creative acts of individual
scientists are embedded: knowledge claims have to be argued according to the
‘rules of the game’ laid down by the community of scientists and expressed in a
language and form determined by the community. An individual scientist’s confi-
dence in the significance of her/his work is insufficient to establish it as part of
the body of knowledge; it must withstand critical scrutiny by the community by
whatever means the community decides is appropriate. Experiments are repeated

128

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
INSIGHT FROM THE SOCIOLOGY OF SCIENCE

(sometimes with variations) by other scientists; chains of evidence are audited;


conclusions and explanations are critically examined; and sometimes the claim is
amended, modified or reformulated. Kuhn (1970) encapsulates the social dimension
of scientific practice in his description of normal science: problem solving and
theory articulation within a ‘received tradition’. Individual scientists are deeply
committed to the tradition and to the research practices it encompasses; they use
the resources of conceptual and procedural knowledge already established (what
Kuhn calls the current paradigm) to address some of the problems defined by the
paradigm as legitimate; when successful, the solutions are incorporated into the
tradition and made available as resources for use by others. According to Ziman
(1978), it is the community-driven nature of the enterprise that makes science
unique. Scientific knowledge, he argues, can be distinguished from other forms
of knowledge because it is “consensible” (expressed in mutually intelligible and
unambiguous language) and “consensual” (the ultimate criterion of validity is
consensus). To gain acceptance, knowledge must meet the standards, expectations
and needs of the community, which are both formidable and strictly enforced through
peer review. What counts as authentic knowledge and proper procedure is jealously
guarded, as witness the rapid closing of ranks whenever there is a claim to know-
ledge by an outsider. These standards are maintained by means of rigorous ‘rites
of passage’ (education and research training), which individuals must complete
successfully before they can gain admission to the community. Once admitted,
scientists must conform to the community’s rules and conventions, including the
adoption of a strict linguistic code. Otherwise, they risk expulsion. While this is a
way for the community to exclude charlatans, it is also an essentially conservative,
restrictive and elitist stance that may serve to exclude members of minority groups.
Charges that, in consequence of these constraints, science is both sexist and racist
will be addressed later in the chapter.

THE NORMS OF SCIENTIFIC PRACTICE

Some sixty years ago, Robert Merton identified four “functional norms” or
“institutional imperatives”, transmitted by precept and example, which govern
the practice of science and the behaviour of individual scientists, whether or not
they are aware of it (Merton, 1973).4 Not only do these norms constitute the most
effective and efficient way of generating new scientific knowledge, they also
provide a set of ‘moral imperatives’ that serves to ensure good and proper conduct.
– Universalism – science is universal (i.e., its validity is independent of the
context in which it is generated) because evaluation of knowledge claims in
science uses objective, rational and impersonal criteria rather than criteria based
on personal, national or political interests, and is independent of the reputation
of the particular scientist or scientists involved.
– Communality – science is a cooperative endeavour and the knowledge it gene-
rates is publicly owned. Scientists are required to act ‘in the common good’,
avoid secrecy and publish their findings and conclusions so that all scientists
may use and build upon the work of others.

129

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
CHAPTER 7

– Disinterestedness – science is a search for truth simply for its own sake, free
from political or economic motivation or strictures, and with no vested interest
in the outcome. Because attempts to exploit the ignorance or credulity of non-
scientists or to fabricate results in pursuit of commercial or personal gain are
strictly outside the code of approved scientific conduct, scientists have tradi-
tionally enjoyed a good reputation for ethical behaviour.
– Organized scepticism – all scientific knowledge, together with the methods
by which it is produced, is subject to rigorous scrutiny by the community in
conformity with clearly established procedures for judging such matters as
methodological appropriateness, chain of argument from data to conclusions,
and testability. The ‘emotional neutrality’ of these procedures ensures that all
knowledge claims are treated similarly, regardless of their origin.
Two additional norms have been proposed by Barber (1962).
– Rationality – science uses rational methods to generate and validate its claims to
knowledge.
– Emotional neutrality – scientists are not so committed to an existing theory
or procedure that they will decline to reject it or adopt an alternative when
empirical evidence points to it.
Similar ideas underpin Garfinkel’s (1967) “rules of interpretive procedure”,
which he contrasts with everyday reasoning: the rule of unlimited doubt asserts that
scientists will not limit their scepticism by the kind of ‘practical considerations’
that govern everyday life; the rule of knowing nothing allows scientists to suspend
their own knowledge in order to see where it leads, while testing in everday life
proceeds on the basis of what can be taken for granted; the rule of universalised
others enables scientists to trust the findings of others; the rule of publicisability
requires that all findings are made public, regardless of personal motives and
interests.
Many contemporary sociologists of science argue that the so-called ‘Mertonian
norms’ of scientific conduct do not guide practice; rather, they are used retro-
spectively by scientists to dignify what they have done, and to impress non-scientists.
Mitroff (1974), for example, suggests that the ‘emotional neutrality’ of organized
scepticism is frequently over-ruled by the ‘emotional commitment’ of scientists
struggling to overcome difficulties and setbacks. Indeed, he postulates a counter-
norm for each of the norms listed above.5
– Particularism – the personal or professional attributes of the researcher, and the
status of the institution, are frequently taken into account in the evaluation of
scientific contributions.
– Solitariness – ownership and control of distribution of scientific knowledge reside
with the individual scientist (or group) who produced it. On occasions results are
withheld until a patent has been secured or delayed until their announcement
will have greater impact.
– Interestedness – many scientists have personal agendas for engaging in parti-
cular research and may have a vested interest in the outcomes – even more so
when research is funded by commercial organizations.
– Exercise of judgement – the expert opinion of experienced scientists plays a
prominent role in the evaluation of knowledge claims. Moreover, the research

130

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
INSIGHT FROM THE SOCIOLOGY OF SCIENCE

of newcomers is subject to much more rigorous checks than the work of esta-
blished scientists.
– Non-rationality – scientists do not always act in a fully functional manner and
scientific advances can result from non-rational as well as rational actions.
– Emotional commitment – commitment to a theory is essential for its advancement;
disinterest leads to stagnation. On occasions, however, commitment in spite of
substantial contrary evidence becomes unreasonable.
In practice, it seems that scientists simply act as they see fit and attempt to
rationalize it afterwards. Hence, rather than regarding science as a distinctive way
of proceeding, to which all scientists have to conform, it makes more sense to
regard science as (no more than) what scientists actually do. Conventions (such
as Mertonian norms) do not direct the actions of scientists, they are simply what
the collective actions of scientists amount to – at least, in their retrospective
rationalizations. In Mulkay’s (1979) words, “it seems more appropriate to portray
the ‘norms of science’, not as defining clear social obligations to which scientists
conform, but as flexible vocabularies employed by participants in their attempts
to negotiate suitable meanings for their own and others’ acts in various social
contexts… What is clear is that it is highly misleading to regard the diffuse reper-
toire of standardized verbal formulations as the normative structure of science or
to maintain that it contributes in any direct way to the advance of scientific
knowledge” (p. 72).
If convention is not a determinant of action, but its product, then the beliefs,
practices and values of scientists are reduced to a set of phenomena to be observed
(directly or indirectly), analyzed and rationalized. In approaching this description
of the scientific endeavour, there are two possible approaches. One is to ask
scientists about aspects of their practice, using questionnaires, surveys and inter-
views; the other is to observe them as they engage in their day-to-day practice in
the laboratory, making detailed field notes of events and audiotaping conversations
between scientists for subsequent discourse analysis.
The former approach, well exemplified by the work of Hagstrom (1965), tends
to focus on the large-scale characteristics of science, in particular its growth, organi-
zation and established mechanisms for admitting and enculturating newcomers. It
also aims at generalizability: what scientists say about their practice is regarded as
applying to each and every situation. In contrast, observational studies tend to be
smaller scale studies in which researchers use an ethnographic approach to describe
and interpret day-to-day events and interactions between scientists in particular
situations. Those researchers who adopt this case study approach sometimes make
little or no attempt to generalize, regarding it as the reader’s responsibility to deter-
mine what, if anything, can be transferred and used to inform and interpret other
situations. What these studies reveal is that science is much less linear, much less
certain and much more disordered than the conventional image suggests. They also
reveal a noticeable mismatch between the rhetoric and the practice of scientists, thus
giving good grounds for subscribing to Albert Einstein’s (1933) remark that “if you
want to find out anything from [scientists] about the methodology they use… don’t
listen to their words, fix your attention on their deeds” (p. 270).

131

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
CHAPTER 7

An important element of this mismatch centres on the cluster of personal


characteristics and attributes that have long been regarded in the rhetoric of science
education as essential for the successful pursuit of science and, so the science
curriculum rhetoric goes, are clearly exhibited in the day-to-day practice of success-
ful scientists: superior intelligence, objectivity, rationality, emotional neutrality,
open-mindedness, willingness to suspend judgement, intellectual integrity and
communality. Like Mertonian norms, these so-called ‘scientific attitudes’ are said
to guarantee proper scientific practice by ensuring that (i) all knowledge claims
are treated sceptically until their validity can be judged according to the weight of
evidence, (ii) all evidence is carefully considered before decisions about validity
are made, and (iii) the idiosyncratic prejudices of individual scientists do not in-
trude into the decision making. Of course, in traditional science curricula ‘evidence’
is always taken to mean empirical evidence – that is, agreement with the observed
‘facts’. Thus, proper scientific practice is seen to comprise a dispassionate appraisal
of that empirical evidence (the ‘facts’) before decisions are taken. It is now more
than forty years since Roe (1961) suggested that scientists rarely possess these
‘scientific attitudes’, although (she says) they think that they do. They, too, sub-
scribe to the myth of the emotionally-detached, disinterested and impartial scientist.
Or they continue to promote this false image because they perceive it to be in their
interests to imply a connection between the disinterested approach and the ‘truth’
of the findings in the battle to maintain high levels of public funding for scientific
research.6 Roe concludes: “The creative scientist, whatever his field, is very deeply
involved emotionally and personally in his work” (p. 456). Mahoney’s (1979) con-
clusions about the attitudes and characteristics of scientists, derived from writings
by sociologists, historians and scientists, make particularly interesting reading.
– Superior intelligence is neither a prerequisite nor a correlate of high scientific
achievement.
– Scientists are often illogical in their work, particularly when defending a pre-
ferred view or attacking a rival one.
– Scientists’ perceptions of reality are dramatically influenced by their theoretical
expectations.
– Scientists are often selective, expedient and not immune to perceptual bias and
distortion of the data.
– Scientists are among the most passionate of professionals. Their theoretical and
personal biases often colour their alleged openness to the data.
– Scientists are often dogmatically tenacious and inflexible in their opinions, even
when contradictory evidence is overwhelming.
– Scientists are skilled in ‘expedient reasoning’ – that is, bending their arguments
to fit their purposes.
– Scientists are not paragons of humility and disinterest. Rather, they are often
selfish, ambitious and petulant defenders of personal recognition and territoriality.
Vitriolic episodes and bitter disputes over personal credit and priority are common.
– Scientists often behave in ways that are diametrically opposite to communal
sharing of knowledge. They are frequently secretive and suspicious of others,
and will frequently suppress data until they have established priority of discovery.

132

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
INSIGHT FROM THE SOCIOLOGY OF SCIENCE

– Far from being a ‘suspender of judgement’, the scientist is often an impetuous


truth-spinner who rushes to hypotheses and theories long before the data
warrants.
Following their study of scientists and engineers involved in NASA’s Apollo
Project, Mitroff and Mason (1974) concluded that scientists are arranged along a
continuum from extreme speculative scientists, who “wouldn’t hesitate to build a
whole theory of the solar system based on no data at all” (p. 1508), to data bound
scientists, who “wouldn’t be able to save their own hide if a fire was burning next
to them because they’d never have enough data to prove the fire was really there”
(p. 1508). These conclusions echo Mahoney’s earlier description of a continuum
of scientists: “At one extreme there are ‘speculophobics’ – scientists who devoutly
avoid any ventures beyond the data. Their opposites are ‘hypothophiliacs’ – scientists
who need less than a hint of evidence to draw sweeping generalizations and ambi-
tious models” (Mahoney, 1979, p. 357). Contrary to the school textbook stereotype,
the scientists who produce the most significant work are those who disregard the
so-called ‘scientific attitudes’; those who conform may be excellent technicians,
but they don’t make the theoretical breakthroughs and procedural innovations. This
conclusion is a reminder that we need to be sure which ‘game’ we are playing in
school science education, and that we should make it clear to students, too. Careful
attention to detail and painstaking accumulation of data is crucial to good science,
to the effective conduct of laboratory testing, to the maintenance of safety and
to the sound operation of science-based industries, but theoretical breakthroughs
are not made this way. As discussed below, creative, theory-building scientists are
those who ‘break the rules’ (see discussion of Feyerabend’s principle “Anything
Goes” in chapter 5). We need to provide our students with experiences capable of
developing both sets of attributes.
In contrast to the caricature of scientific inquiry portrayed in many school text-
books, real scientific inquiry is holistic, fluid, flexible, reflexive, context-dependent
and idiosyncratic (see chapter 5). It is characterized by frequent false starts, blind
alleys and improvised modifications; it can be, and often is, re-directed by un-
expected events and by unanticipated technical problems; it can be profoundly
influenced by the availability (or not) or particular laboratory equipment or its
unexpected malfunctioning, by the publication of a research paper in the same field
or chance conversation with another researcher. Among others, Collins (1985)
and Latour and Woolgar (1986) describe scientific practice as a long and tedious
process for creating order out of disorder, information out of ‘noise’ and organi-
zation out of chance events. First, there is the difficult business of extracting data
from ‘background noise’ – that is, deciding what is significant and what is not, and
imposing meaning frameworks to create a coherent data set. Traditional ‘scientific
attitudes’ are of little value here; much more important are intuition and tenacity in
the face of conflicting evidence and counter argument from other scientists. Of
course, arguing for a particular view will usually involve arguing against another –
that is, negating alternative interpretations or rendering them less plausible through
argument and skilful presentation of data, with each side in the dispute seeking to
‘explain away’ inconvenient data. Critics will cite researchers’ negative results as

133

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
CHAPTER 7

grounds to dismiss any controversial idea with which they disagree, with any
residual positive results being explained away as incompetence, delusion or even
fraud. The proponents of the new idea, on the other hand, account for competitors’
negative results from attempted replication in terms of failure to reproduce exactly
the same conditions used by them to obtain positive results (Collins & Pinch, 1993).
Replication of experimental results, long held to be a cornerstone of scientific
method, is much less straightforward than the school science curriculum implies, and
what is eventually admitted as satisfactory replication may be a consequence of
careful negotiation that relies heavily on intuition and value judgements and takes
account of the prestige of the research group (Collins & Pinch, 1993; Knorr-Cetina,
1995). There are a number of reasons why replication is not a simple matter: while
scientists engaged in cutting edge research may agree on basic conceptual issues,
they may profoundly disagree on proper design of experiments, specifications for
instruments and interpretation of data. Complex scientific investigations cannot be
reduced to a simple algorithm that can be checked at each stage (see discussion in
chapter 5); a great deal of the ‘know how’ that goes into the design of incisive
experiments is tacit and intuitive, while much of the expertise that goes into their
effective conduct is located at the craft level and is acquired through lengthy on-
the-job experience. Given these complexities, it isn’t always clear whether a second
experiment really is a replication or just another (different) experiment. Moreover,
as Collins (1975) points out, most research groups would not consider exact copies
of experiments desirable. First, it would reflect no particular credit on the group,
whereas a better experiment would do so, and would give them a competitive edge.
Second, a new effect is more powerfully demonstrated or a new explanation more
effectively made when more than one approach is used – what social scientists call
triangulation. In consequence, attention shifts from replication to what counts as a
‘good experiment’. Of course, in one sense a ‘good experiment’ is one that provides
evidence for one’s theoretical claim, a ‘bad experiment’ is one that doesn’t.
In contrast to the traditional notion that scientists complete their investigations
and theorizing prior to disseminating their findings and conclusions, laboratory
studies reveal a much more fluid and interactive relationship in which scientists are
engaged in a continuing struggle to persuade themselves and others that their data
are important and their interpretations are valid. For some, the laboratory is no
longer a site for the discovery and validation of knowledge but a site for the
manufacture of knowledge (Knorr-Cetina, 1981, 1983).
The study of scientific knowledge is primarily seen to involve an investigation
of how scientific objects are produced in the laboratory rather than a study of
how facts are preserved in scientific statements about nature. (Knorr-Cetina,
1983, p. 119)
Latour and Woolgar (1979, 1986) regard scientific work as essentially a form of
writing and the laboratory as an instrument of persuasion through which conjec-
tural statements are transformed into statements of ‘fact’. In other words, meaning
is selectively constructed and re-constructed until it is capable of convincing the
group with power of decision on acceptability. Recognizing that scientific knowledge

134

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
INSIGHT FROM THE SOCIOLOGY OF SCIENCE

is negotiated within the community of practitioners opens up the somewhat alarming


prospect that the outcome of scientific disputes may ultimately depend on the argu-
mentation skills, prestige and material resources that participants in a dispute can
mobilize in making their case.
If there exists an unavoidable indeterminacy in principle in relation to scien-
tific decisions, then it may be that the rhetorical brilliance of those advoca-
ting a particular outcome, the political saliency of the findings, or the support
proponents can draw on, etc., may tip the balance in favour of a specific
choice. (Knorr-Cetina & Mulkay, 1983, p. 12)
It almost goes without saying that if the skills of persuasion and dissuasion
are an important element of scientific expertise, they are an important element of
scientific literacy and should be fostered in the science curriculum (Osborne, 2001;
Duschl & Osborne, 2002) – an idea that will be elaborated a little in chapter 9.

SEARCHING FOR AUTHENTICITY

Laboratory studies go beyond concern with the nature and conduct of experiments
to shed light on the laboratory as a ‘cultural space’ within which knowledge is
constructed by the collective efforts of scientists. They give us some understanding
of “the bricolage, tinkering, discourse, tacit knowledge and situated actions that
build local understandings and agreements” (Fujimura, 1992, p. 170) and the sub-
sequent debating, persuading and political manoeuvring involved in gaining the
interest and support of scientists outside the immediate group – support that is
essential if the research is to become part of accepted scientific knowledge.
Scientific objects are not only ‘technically’ manufactured in laboratories but
are also inextricably symbolically or politically construed, for example, through
literary techniques of persuasion such as one finds embodied in scientific
papers, through the political stratagems of scientists in forming alliances and
mobilizing resources, or through the selections and decision translations which
‘build’ scientific findings from within. (Knorr-Cetina, 1992, p. 115)
Scientists rarely study objects and phenomena as they occur in nature. Indeed,
some of the power of experimentation lies in the capacity to control and contrive
situations so that scientists do not need to deal with all the complexities of studying
an object or phenomenon as it is (it can be simplified, idealized and stylized) or
where it is in the real world (by re-locating it to the laboratory, all kinds of techno-
logical resources can be brought to bear). Nor do scientists need to accommodate to
an event when it happens to occur (natural cycles can be replaced by more con-
venient ones – more frequent, more predictable). Thus, experiments are represen-
tations of reality – processed, manipulated and partial versions of it, rather than the
‘real thing’. Everything in the laboratory is, in some sense, a human construct: the
knowledge from which scientists begin, the materials, the measuring devices and
other technical tools are all ‘created entities’; laboratory animals are selectively
bred; plants are specially grown; pure chemicals are manufactured; and even the

135

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
CHAPTER 7

water supply is sterilized and deionized. Because of the complexity, fluidity and
context-specific character of experimentation, scientists themselves could be regar-
ded as a kind of ‘technical device’: repositories of experiential knowledge, tacit
understanding, scientific intuition and capacity for critical judgement. Technicians,
through their capacity to optimize experimental conditions, to build, maintain and
use apparatus specially suited to the circumstances, and to ‘feel’ what is reasonable
and acceptable as experimental results, can also be regarded as ‘instruments’ for
experimentation (Knorr-Cetina, 1992).
Within this socially constructed working environment, the business of doing
science proceeds through a series of decisions and negotiations about problems,
priorities, procedures, measurements, interpretations, arguments and conclusions.
As argued in earlier chapters, there is no one way to proceed on all occasions;
scientific method is not a pre-determined algorithmic procedure. Decisions about
the most suitable way to proceed are local and context-specific, determined by
what is already known about the objects, phenomena or event under investigation
and strongly influenced by: (i) access to information (including library provision),
(ii) who can be consulted, (iii) what materials, facilities, equipment and other
resources are available, (iv) extent of technician support, (v) availability of service
laboratories, and the characteristics of their practices, (vi) funding and conditions/
restrictions governing its specific deployment (equipment versus technician support
versus research assistants, for example), and even (vii) seemingly trivial matters such
as office and laboratory regulations and business hours. Moreover, because of the
complexity and uncertainty of scientific investigation there is considerably more
at issue than the execution of a pre-determined plan. Fluidity is the key: every deci-
sion about procedure is made in the light of a previous decision (and its outcome)
and has consequences for future decisions – both within a particular inquiry and
within a research programme as a whole. In other words, the products of a
particular scientific investigation are not only decision-impregnated, they are also
decision-impregnating in the sense that they point to new problems and (to an
extent) direct their solution. In effect, the logic of scientific research is an oppor-
tunistic one: scientists do what they can in the circumstances and adjust their research
goals and intentions to the changing circumstances, modifying their approach in the
light of successes and difficulties encountered, exploiting unanticipated opportunities,
and constantly monitoring the progress of competing research groups. Success
depends on the ability of the scientist to think on several different levels simul-
taneously, maintaining an overview while analysing the individual elements, piecing
together seemingly disparate bits of information, recognizing clues and sensing
what moves are appropriate, and at every stage trying to anticipate and counter
likely criticisms of the research. Given the messy, fluid and contingent nature of
the process, it is not easy to predict the course of an investigation or the likelihood
of eventual success. In consequence, it is not possible to devise a set of decision-
making criteria for would-be researchers to study and apply. This is not to say that
scientists are unsystematic or irrational, but it is to say that the openness of
scientific inquiry, the context embedded and intuitive nature of the decision making,
the craft element of bench expertise, and the tacit knowledge and understanding

136

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
INSIGHT FROM THE SOCIOLOGY OF SCIENCE

involved in using laboratory instruments skilfully and effectively cannot be easily


captured in words.
Of course, there is a major mismatch between the way scientific inquiry is
conducted and the way it is reported, and a similar mismatch between the pri-
vate language of argument and negotiation within the laboratory (and embedded
in laboratory notebooks) and the public language of scientific argument used in
academic journals. The need to communicate one’s findings as precisely and effi-
ciently as possible, and to make them convincing to others, determines the lin-
guistic form of the research report and the scientific paper. Gone are the references
to crises, compromises and intuition; replaced by an account in which references
to human agency are reduced to a minimum, a text in which the physical world is
made to ‘speak for itself’. Thus, the emergence of the ‘correct view’ is portrayed as
arising unproblematically from the data (Mulkay et al., 1983). More than 40 years
ago, Peter Medawar (1963) asked the provocative question: “Is the scientific paper
a fraud?” In the sense that it is constructed to persuade readers of a particular point
of view rather than to describe the day-by-day events of the investigation, it is
a fraud. It frequently conceals the situationally contingent and opportunistic logic
of the inquiry, renders the choice of method straightforward and unproblematic
and misrepresents the motives for the work in an effort to provide a clear and
logically compelling argument for the validity of the findings and the author’s
particular interpretation and explanation. As Knorr-Cetina (1981) so disarmingly
puts it: “The scientific paper hides more than it tells on its tame and civilised
surface” (p. 94).
Latour and Woolgar (1979, 1986) see the creation and deployment of text as the
dominant feature of laboratory life and regard enculturation into the scientific
community as being, in significant part, the learning of literary techniques for per-
suading, interesting and influencing others and the mastery of ‘inscription devices’
for enhancing the persuasive power of texts.
It seems that whenever technicians are not actually handling complicated pieces
of apparatus, they are filling in blank sheets with long lists of figures; when
they are not writing on pieces of paper, they spend considerable time writing
numbers on the sides of hundreds of tubes, or pencilling large numbers on the
fur of rats… When the observer moves from the bench space to the office
space, he is greeted with yet more writing. Xeroxed copies of articles, with words
underlined and exclamation marks in the margins, are everywhere. Drafts of
articles in preparation intermingle with diagrams scribbled on scrap paper,
letters from colleagues and reams of paper spewed out by the computer in the
next room; pages cut from articles are glued to other pages; excerpts from draft
paragraphs change hands between colleagues while more advanced drafts pass
from office to office being constantly altered, retyped, recorrected, and even-
tually crushed into the format of this or that journal. (1979, pp. 48 & 49)
Ideas become established, both within the laboratory and beyond, not simply
on the strength of the data but through repeated use of data and interpretive
commentary in linguistic forms that take away doubt about the preferred view and

137

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
CHAPTER 7

cast doubt on the rival. According to Latour (1987), avoiding ‘isolation’ is a key
strategy in winning the argument, hence the heavy use of footnotes, quotations and
references to supporting literature – all of which have to be separately addressed
and countered by those intent on opposing the argument. Opposing views should
be anticipated and dealt with in the text through careful and selective referencing of
other research. Latour’s (1987) advice is to “do whatever you need to the former
literature to render it as helpful as possible for the claims you are going to make”
(p. 37). A convincing text is arranged in layers – in Latour’s words, “a folded array
of successive defence lines” (p. 48) to deal with opponents- with each claim being
supported by references to other work, cross referenced to other parts of the paper
and bolstered by technical detail, diagrams, graphs, equations and the like. Within
the laboratory, and the scientific community at large, particular significance is
attached to the operation of apparatus capable of providing some kind of ‘written
output’ in the form of charts, tables, graphs and figures. Judicious deployment
of such ‘inscriptions’, as Latour and Woolgar (1979, 1986) and Latour (1987) call
them, gives some increased persuasive power to scientific papers, while reference
to expensive ‘inscription devices’ (the technologies that produce them) sends an
important message to readers about the resources available to the research team,
thereby enhancing the group’s credibility and the likelihood of its ‘products’ being
regarded as important. Indeed, these authors ultimately made sense of the ‘laboratory
life’ they were studying (their particular sense of it, that is) in terms of the
relationships they perceived among the inscription capabilities of scientific apparatus,
what they call the “manic passion for marking, coding and filing” and the skills of
persuading others through construction of scientific text.
Thus, the observer could even make sense of such obscure activities as a
technician grinding the brains of rats, by realising that the eventual end pro-
duct of such activity might be a highly valued diagram. (Latour & Woolgar,
1979, p. 52)
Latour (1987) talks about what he calls “black boxes”, a means of encoding
certain aspects of a perceived reality within an instrument, procedure or theoretical
structure. Using particular black boxes is a useful tactic in mounting a scientific
argument because their use signals the legitimacy of the embedded view of the
world. With its previously problematic aspects effectively hidden from sight, a
black box becomes a powerful tool of persuasion because the intellectual effort that
would be involved in mounting an objection to the validity and reliability of the
data it produces. As Latour remarks, controversies end when black boxes are
successfully deployed.
Black box is used by cyberneticians whenever a piece of machinery or a set of
commands is too complex. In its place they draw a little box about which
they had to know nothing but its input and output… no matter how con-
troversial their history, how complex their inner workings, how large the
commercial or academic networks that hold them in place, only their input
and output count. (Latour, 1987, p. 2)

138

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
INSIGHT FROM THE SOCIOLOGY OF SCIENCE

Of course, to be an influential contribution to science, a paper has to be


sufficiently convincing and sufficiently important to ensure that ‘significant others’
make use of it, preferably without modification or qualification of its key points.
Thus, while scientists extend knowledge in ways that serve their particular interests,
they also have to serve the needs and interests of (some) others – for example, by
solving a problem that others have encountered, creating a problem that others
cannot ignore, developing a powerful new technique for generating data, striking at
the fundamental assumptions of the rival’s research programme, or some such
provocative move.

NORMS, INTERESTS AND VALUES

In the scientific world described by Merton (1973) the academic community is


simultaneously a communication system and a reward system. The contributions
of individuals are assessed by an audience of peers and if judged to be original
and significant they are allowed to be published, and are made available for all
to use. Hagstrom (1965) describes it as a ‘barter system’ in which the scientist
gets recognition (which may subsequently translate into tangible rewards such as
tenure, promotion, research grants, or even a Nobel Prize) in ‘payment’ for provi-
ding new intellectual resources (information, ideas, techniques, etc.) and, in turn,
such recognition stimulates the kind of approved behaviour that leads to further
recognition. There is implicit agreement among scientists to adhere to the insti-
tutional norms described by Merton and Barber because they provide the optimal
way of producing certified knowledge which, in turn, can be exchanged for recog-
nition and reward. If everyone ‘abides by the rules’ then everyone ‘gets a fair go!’
In Storer’s (1966) words:
Scientists subscribe to the norms of science first of all because of their
importance for the continued, adequate circulation of the commodity in
which they are mutually interested… It is the occasional reinforcement given
these norms by the scientist’s awareness of their relevance to his own interest
in obtaining competent response to his work rather than the general goal of
science, which I feel accounts for their continued moral potency… these
norms are natural concomitants of the desire to be creative. (p. 84)
It is noteworthy that Storer includes successful completion of creative acts in his
description of rewards. As Wong and Hodson (2008a,b) report, many scientists cite
this as a major driving force for their research, alongside curiosity, “finding out how
things work”, “making a contribution”, improving the quality of human life and the
sheer enjoyment of solving complex problems.
A decade after Hagstrom’s influential work, Bourdieu (1975) introduced the
notion of a capitalist market economy in which scientists are engaged in a competi-
tive struggle for a monopoly on scientific credit – symbolic capital acquired by
scientists through research and other technical achievements and capable of con-
version into resources for further scientific production. Thus, the cooperative world
of science envisaged by Merton and Hagstrom was replaced by an essentially

139

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
CHAPTER 7

antagonistic world, in which scientists seek to dominate the field and to discredit
competitors.
The scientific field is the locus of a competitive struggle, in which the specific
issue at stake is the monopoly of scientific authority, defined inseparably as
technical capacity and social power. (Bourdieu, 1975, p. 19)
Latour and Woolgar (1979, 1986) modify Bourdieu’s account quite radically by
suggesting that scientists do not simply seek credits or rewards; rather, they are
investors in credibility – a form of capital that is capable of ready conversion into
research grants, equipment, and in due course, further knowledge production.
Scientists ‘invest’ in fields and topics that promise the greatest return for their
efforts, and the credibility they gain from producing new knowledge is converted
into resources for ‘reinvestment’ in further research. Therefore, the goal is not the
pursuit of truth, nor the gaining of recognition per se. Successful research groups
are those capable of accelerating and expanding the credibility-investment cycle. In
a sense, production (publication) for the sake of production is the hallmark of
scientific capitalism.
In the traditional forms of ‘basic’ or ‘fundamental’ research envisaged by Merton
(1973), usually located in universities and/or government research institutes,
‘pure scientists’ constitute their own audience: they determine the research goals,
recognize competence, reward originality and achievement, legitimate their own
conduct and discourage attempts at outside interference. In the contemporary
world, research is often dependent on expensive technology and so must meet the
needs and serve the interests of those sponsors whose funds provide the resources.
In consequence, scientists have lost a substantial measure of autonomy and are
now seen, in some situations, as advocates for a particular point of view rather
than disinterested arbiters of the ‘truth’. The vested interests of the military and
commercial sponsors of research, particularly tobacco companies, the petroleum
industry, the food processing industry, pharmaceutical companies and the nuclear
power industry, can often be detected in research design – especially in terms of
what and how data are collected, manipulated and presented. More subtly, in what
data are not collected, what findings are omitted from reports and whose voices are
silenced. Commercial interests may influence the way research findings are made
public (e.g., press conferences rather than publication in academic journals) and the
way the potential impact of adverse data is minimized – for example, in publicizing
the value of oral contraceptives and hormone replacement therapy, the increased
risks of cervical cancer, breast cancer and thromboembolism were given little
attention. Science is no longer the disinterested search for truth and the free and
open exchange of information portrayed in the school science textbook stereotype.
Rather, it is a highly competitive enterprise in which scientists may be driven by
self-interest and career building, financial inducements provided by business and
commerce, or by the ‘political imperatives’ of military interests. The cynical inter-
pretation of this state of affairs is that the science that ‘gets done’ is the science that
is in the interests of the rich and powerful, the self-interest of particular scientists,
and the companies that employ them. Chambers (1984c) deals with these matters

140

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
INSIGHT FROM THE SOCIOLOGY OF SCIENCE

admirably, and appropriately for both school science and science teacher education,
in his compelling account of the manufacture and use of pesticides, the DDT con-
troversy and the work of Rachel Carson.7
One consequence of this struggle for recognition, credit and credibility is the
creation of boundaries between different sub-disciplines of science and between
‘insiders’ and ‘outsiders’. Science is increasingly fragmented, diversified and
specialized. The days when any one individual could command knowledge of all
science are long gone. While subject disciplines retain some political significance
at the university level in terms of teaching, most of the effective work at the
research level goes on in sub-groups of scientists with common interests (but
little in common with other sub-groups). Across institutions, these sub-groups
form ‘invisible’ research networks (Mulkay, 1977), whose members differ from the
members of other groups in terms of research interests (area of concern, phenomena
studied, etc.), theoretical frameworks employed, procedures, instruments and techno-
logy deployed, to such an extent that Hacking (1992) regards their practices as
incommensurable. The greatest difference between sub-disciplines is in terms of
the scientists themselves – in particular, their experiences, skills and network of
informants. Research groups seek to retain this expertise within the group because
that is where their investment capital is primarily located. Because a substantial
component of this expertise is tacit and not amenable to transfer in written form,
even if it were desirable, it has to be learned on-the-job and through personal
contact with experts. As Ravetz (1971) says: “In every one of its aspects, scientific
inquiry is a craft activity depending on a body of knowledge which is informal and
partly tacit” (p. 103).
Of course, proliferation of research groups and sub-disciplines leads to even more
intense competition for limited resources, with possible deleterious impact on the
ways in which scientists conduct their affairs. Callon et al. (1986) see the efforts
of scientists to rally resources, and their methods of doing so, as similar in kind
to those of politicians and entrepreneurs: “Controlling resources, controlling the
environment, and controlling the world that is being built, all of these are aspects
of the entrepreneurial activity of scientists. In a sense then, they are not only prac-
ticing science – they are also practicing politics, economics, and sociology” (p. 10).
While the Mertonian norm of ‘communality’ demands openness of communication,
competition fosters secrecy and concealment of information. Both commercial and
military interests lead to increasing secrecy and to privatization of knowledge.
Removal of science from public scrutiny and all its attendant safeguards, and the
tendency to ‘cut corners’ brought on by impatience to get quick results, strike at
our ability to distinguish sound from unsound knowledge, and plausible from
implausible chains of arguments, and the public is left with no real means of
judging the ‘warrant for belief’. When secrecy replaces openness, and the pursuit
of personal reward replaces the search for explanation, science is profoundly
devalued and the public can no longer trust what scientists tell them. When Latour
and Woolgar (1986) assert that “each text, laboratory, author and discipline strives
to establish a world in which its own interpretation is made more likely by virtue of
the increasing number of people from whom it extracts compliance” (p. 285) they
seem to suggest that any way of playing the game is acceptable if it brings success.

141

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
CHAPTER 7

Indeed, Pinch and Collins (1984) go so far as to endorse deception when they
suggest that one of the research groups they studied would have been more
successful if the members had suppressed ‘inconvenient’ evidence – a position
which Slezak (1994) characterizes as “truth is what you can get away with”. Sadly,
there are scientists, most notably the psychometrician Cyril Burt, who take this
dictum at face value and fraudulently manufacture the ‘evidence’ on which they
build their theories. Indeed, it seems that fraud, the most blatent form of self-
interest serving, is much more widespread than we might expect or that scientists
will generally admit to, or that we tell students in school (Broad & Wade, 1982;
Kohn, 1986; Bell, 1992; Wible, 1992; Judson, 2004).8 While they are a long way
from fraud, Wong and Hodson (2008a,b) report several examples of what might be
termed ‘sharp practice’: using institutional power to stifle dissent, publishing false
data to mislead rivals, and using the peer review system to delay competitors.

SCIENTIFIC KNOWLEDGE AS A SOCIAL CONSTRUCT

It has been asserted a number of times in this book that we do not and cannot know
the world as it ‘really is’. Rather, we describe and explain the world in terms of the
conceptual frameworks that scientists have developed. While material reality will
closely limit or constrain the number of possible ways of knowing, reality alone is
insufficient to determine the truth or what we believe about it. We have to supply
much of it, through our imagination. In this chapter, it has been argued that science
(including the processes of theory-building) is a socially-embedded activity. In
other words, scientific knowledge is socially constructed through the practices of
the scientific community, in response to the demands, needs and interests of the
wider community that surrounds and supports it. Furthermore, the procedures of
science (experiment, observation, etc.) are based on conventions that are, themselves,
human constructs, subject to variation and the outcome of negotiation among
scientists. Some sociologists of science (and, of course, some philosophers of
science, historians of science and scientists) go further, claiming that the know-
ledge produced by scientists is no more than a social construct: “Scientific activity
is not about ‘nature’, it is a fierce fight to construct reality” (Latour & Woolgar,
1979, p. 2, emphasis added). In short, science creates its own reality.
There is a sense in which this is self-evidently true: the science and technology
of the modern laboratory is sometimes capable of creating phenomena and objects
that previously did not exist.9 The laser is a prime example of such a technology.
So, too, is Dolly the cloned sheep. For the past 50 years or so the research of most
chemists and materials scientists has been directed towards the synthesis of mole-
cules and materials that do not exist in nature. However, the case made by social
constructivists is that all scientific knowledge is simply a social construct, and no
more than a social construct. Crucially, it could have been constructed differently.
In other words, scientific knowledge is not just socially constructed, it is a social
construct, a mere social convention. Moreover, if scientific knowledge and its
means of production are social products, and so could be otherwise, it follows that
the criteria of judgement in science must also be a social construct. Indeed, rationality
itself is simply a human convention, and could be otherwise.

142

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
INSIGHT FROM THE SOCIOLOGY OF SCIENCE

From Shapin’s (1979) classic study of phrenology it is fairly easy to see why
and how a theory linking intellectual ability with social class was developed in
Victorian England, but the author goes on to make the point that phrenology cannot
be understood apart from the historical contexts of use and sociopolitical interests.
This assertion has been taken by some scholars as a warrant to shift from the
position that knowledge is influenced by the social context in which it is produced
to the position that all knowledge is solely determined by the social context. The
only reality is the social reality, and all knowledge is relative to the social context
of its production. Truth becomes simply the beliefs and values that happen to
prevail among members of influential communities of practice.
Knowledge for the sociologist is whatever men take to be knowledge. It
consists of those beliefs… which are taken for granted or instituionalised, or
invested with authority by groups of men. (Bloor, 1976, p. 2)
I argued in chapter 4 that all the so-called processes of science are theory
impregnated and theory driven. The way in which entities are classified, for example,
depends on one’s purpose and theoretical knowledge. Borges (1964, p. 103) reports
that the ancient Chinese encyclopaedia, The Celestial Emporium of Benevolent
Knowledge, divides the animal kingdom into the following categories: (a) those
that belong to the Emperor; (b) embalmed ones; (c) those that are tamed or trained;
(d) suckling pigs; (e) mermaids (or sirens); (f) fabulous ones; (g) stray dogs; (h) those
that are included in this classification; (i) those that tremble as if they were frenzied
or mad; (j) innumerable ones; (k) those drawn with a very fine camel’s hair brush;
(l) others; (m) those that have just broken a flower vase; (n) those that from a
distance look like flies.10 Why does it sound so odd, even ridiculous, like some-
thing from a Monty Python sketch? In part, it is because our own familiar classi-
fication system has pre-determined the way we think and we see no need to include
imagined animals, such as “fabulous ones’ and mermaids, in our classification.
In part, it is because the (a), (b), (c) style of presentation sets up an expectation
that the categories are mutually exclusive, logical, purposeful and inclusive of all
animals. We are then presented with the category “included in this classification”.
The system clearly doesn’t fit any purpose we might have. Nor can we ascertain
its underlying logic. But, as argued in chapter 5, the categories we employ in classi-
fication are simply those that we have decided are significant and logical. They
don’t emerge from our observations; rather, we impose them on our observations.
We choose to observe and classify in one way rather than in some other way
because it suits us to do so, because we are comfortable with that classification,
because it serves our needs, purposes and interests.11 In other words, the categories
we employ are convenient, socially constructed ones. Learning to classify ‘properly’
is a matter of learning to employ the classification system sanctioned by the parti-
cular community we inhabit. History tells us that the ways in which we classify
things have changed over time. They also vary quite substantially from culture to
culture. But it takes an example like the one presented by Borges to remind us of
the extent of social construction in familiar everyday language and in scientific
terminology. In Foucault’s (1970) words, “the thing we apprehend in one great

143

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
CHAPTER 7

leap, the thing that, by means of the fable, is demonstrated as the exotic charm of
another system of thought, is the limitation of our own, the stark impossibility of
thinking that” (p. xv)
Chapter 6 included a brief discussion of the ways in which scientists use a range
of analogies and metaphors in building theories to account for natural phenomena.
In other words, scientists attempt to make sense of the unknown in terms of what is
familiar – both to themselves and to others. Kepler used metaphors from music and
geometry to formulate views about planetary orbits; Rutherford and Bohr used
the analogy of planetary motion to think about atomic structure. The everyday
language of science is rich in embedded metaphor, simile and analogy: electrical
resistance, the flow of electrons, genetic code, messenger RNA, natural selection,
computer virus, and even hormonal permissiveness and antagonism. Because the
effectiveness of such imagery is rooted in its familiarity, metaphors and analogies
do not easily cross cultural boundaries (Nashon, 2003). It is tempting to speculate on
whether the various ‘thematic features’ of scientific thought identified by Holton
(1986) are also culturally specific, particularly the tendency to use dichotomies
and hierarchies. Similar questions arise with respect to what Jurgen Habermas
(1971) calls the universal human interest in the ‘technical control of nature’ and the
anthropocentric motivation identified by Smolicz and Nunan (1975) as one of four
“ideological pivots” underpinning both science and science education (see chapter 2).
One further point should serve to complete the case for science as a social
construct. According to some sociologists of science, academic journals are not
devices for disseminating truth; rather, they are the means by which the scientific
community imposes meaning – that is, enforces the current paradigm. Firstly, the
distinctive language is a social construct and could be different. It is an imposed
convention that serves a particular purpose – ensuring conformity to imposed ways
of seeing the world. Forcing people to use a particular language is an important
first step in controlling the way they think. Throughout history language has been
used as an instrument of control, especially in colonial conquest. Secondly, the whole
enterprise of refereeing for publication via peer review is a tradition established by
the community. It, too, is a social construct and part of the mechanism by which
scientific knowledge is ‘manufactured’, and is intimately concerned with issues of
power, prestige and reward (in the form of funded research contracts, for example).
If the ways in which we classify and observe are ‘conventional’ (that is, social
conventions that we choose, for whatever reasons we regard as important),
if scientific theories are often metaphoric in origin and, therefore, socially and
culturally dependent, if science is conceived, conducted and reported within a
language that is constructed to serve the purposes of a particular group, and if
knowledge is intimately bound up with human interests of various kinds, then it
seems to follow that scientific knowledge is determined by the prevailing socio-
cultural climate (Barnes & Edge, 1982). In other words, it is a social construct –
and possibly no more than a social construct. Scientific knowledge is ‘what scientists
say it is’, for the reasons they choose. It could be otherwise. It has no special status.
Science is ‘no better’ than any other knowledge about natural phenomena and
events; it is just ‘different’. This position is used to underpin arguments for the
legitimacy of ethnosciences and argue for the inclusion of traditional environmental

144

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
INSIGHT FROM THE SOCIOLOGY OF SCIENCE

and technological knowledge in the school curriculum, and to substantiate the claim
that science is inherently sexist and racist – in that it reflects the interests, ways of
thinking and values of Western males.12
Although a reasoned consideration of these arguments and claims is outside the
scope of this book,13 it is worthwhile rehearsing (very briefly) the basic ‘common
sense’ argument, on the grounds that it informs the position we might take with
respect to science education. Since most scientists, for whatever social reasons,
have been male there is good reason to expect that the science produced will be
androcentric (male centred and male biased) simply because it reflects the outlook,
interests, priorities and experiences of those who produced it. If it is legitimate
to refer to masculine ways of thinking, then science could be expected to reflect
them, and some would argue that prominent features of science such as analysis,
quantification and the use of hierarchies and dichotomies are, indeed, illustrations
of such masculine bias. So, too, the underlying values of science relating to control
and manipulation of the natural environment. Furthermore, the argument goes, because
science plays such a powerful ideological role in our society it has functioned to
legitimate social inequality between the sexes – in other words, to discriminate
against women. For many years, women were denied access to science – presumably
in order to retain power and influence in the hands of men. In more recent years,
with the removal of formal barriers to access, the masculine face of science has
functioned to dissuade, limit or restrict access by making women feel uncom-
fortable or ‘out of place’. It is a truism that a society14 in which one gender (or
race) is dominant is likely to distribute resources disproportionately, with the
greater share going to the dominant group and the inequity justified on the basis of
presumed inherent differences between dominant and subordinate groups. Thus,
allegations of gender bias come in a variety of forms:
– Claims about the unequal composition of the profession and the unfair treatment
of women in it/by it.
– Claims about the failure of the profession to investigate matters of interest and
importance to women.
– Charges that scientific research victimizes women.
– Charges that scientists have projected their culture’s gender stereotypes onto the
natural world.
– Assertions that the language of science is masculine.
– Assertions that scientific thinking itself is essentially a masculine way of
theorizing and arguing.
– Assertions that the underlying values of science are masculine.
To all these charges can be added all the evidence and argument that science
education is biased in terms of gender and serves to disadvantage young women –
another vast literature that is outside the scope of this book.15 In addition, these
arguments and claims about gender bias are fairly easily extended to encompass
ethnicity. The case is essentially the same: because most scientists are and have
been Western scientific practice and the knowledge it generates has a Western bias.
Furthermore, science has often been used to oppress and to discriminate against
particular ethnic groups.

145

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
CHAPTER 7

Acknowledging that science is a socially embedded practice, admitting to the


possibility that some of our scientific knowledge may be no more than a social
convention, and recognizing the potential for bias and distortion in the construction
of science, raises questions about the demarcation criteria of science – in particular,
whether the distinctiveness of science is located in its area of concern, concepts,
methods, traditions, linguistic conventions, criteria of validity or underlying values.
It prompts further questions about which of these elements could be changed and
the activity still be characterized as science? If certain elements could be changed,
should they be changed? Would science be improved (in some significant way) by
such changes? If so, in what ways? If change is considered desirable, how can it be
brought about?

ENDNOTES
1
‘True’ is being used here in the sense of ‘scientifically true’ (the theory has achieved consensus
within the community of scientists) - as discussed in previous chapters.
2
This may be a Western characteristic. The traditional environmental knowledge of Aboriginal
Australians, for example, puts much more emphasis on usefulness in a specific context than on
generalizability across contexts as a criterion of validity. Hence, there is no need to seek similar or
consistent explanations (Christie, 1991).
3
One particularly startling study of the peer review process suggests that papers are sometimes
accepted more on the basis of institutional affiliation of the authors than on the academic merits of
the paper (Peters & Ceci, 1982, 1985) – a view endorsed by several of the scientists interviewed by
Wong and Hodson (2008a,b).
4
Merton first outlined this theoretical framework in a 1942 essay titled “Science and technology in a
democratic order”, published in the Journal of Legal and Political Sociology. The citations in this
chapter refer to a subsequent work published in 1973 (see references).
5
Only six of Mitroff’s list of eleven counter-norms are directly relevant to the discussion in this
chapter.
6
Science teachers may also feel some vested interest in sustaining this image of science as a means of
enhancing their status in school (Gaskell, 1992).
7
See Dillon (2005) for a commentary on the continuing significance of Silent Spring (Carson, 1962),
written to commemorate the 40th anniversary of its original publication.
8
O’Rafferty (1995) makes a clear and strong case for a study of fraud and scientific misconduct to be
part of the science curriculum.
9
Cartwright (1983) suggests that one of the primary purposes of experiments is to create phenomena
and events as evidence for the validity of theoretical speculation.
10
Similarly incomprehensible to contemporary science is the assertion by Parcelsus that good doctors
should not have red beards and his belief that an infusion made from a plant with a leaf pattern
resembling a snake will provide protection against poisons.
11
Although it is an urban myth that Inuktitut and the five Yup’ik languages have more than a hundred
words for classifying snow, these languages do have more than English because there is a need for
them. There are probably between 15 and 18, depending on interpretation, compared with about 6 to
9 in English (including snowflakes, falling snow, snow ‘on the ground’, snowdrift, slush, blizzard,
sleet, frost and ice).
12
Not surprisingly, many scholars (especially scientists and philosophers) have little patience with this
proposition. Mario Bunge’s (1992) response is characteristically forthright: “There are no such
things as proletarian science or Aryan science, black mathematics, or feminine philosophy. These
are just political or academic rackets. To be sure, learning prospers more in some groups or societies
than others, but so does superstition” (p. 51).
13
Useful starting points for teachers wishing to acquire some awareness of this literature are Shepherd
(1993), Keller and Longino (1996), Etkowitz et al. (2000), Lederman and Bartsch (2001).
14
‘Society’ is used here in the several senses that ‘community’ has been utilized throughout the
chapter – the laboratory group in which a scientist works, the group of scientists engaged in similar

146

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
INSIGHT FROM THE SOCIOLOGY OF SCIENCE

and related research, the various official governing bodies (including publishing mechanisms), and
society as a whole.
15
Part of the drive for universal critical scientific literacy necessarily includes establishment of a much
more inclusive form of science education – with respect to gender, ethnicity, class and sexual
orientation (Gaskell & Willinsky, 1995; Roychoudhury et al., 1995; Barton, 1998; Barton &
Osborne, 1998). It is also important for students to be made aware of gender bias in science and
science education.

147

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
Derek Hodson - 978-90-8790-507-1
Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
CHAPTER 8

MAKING A CASE FOR HISTORY OF SCIENCE


Going Beyond Dates and Anecdotes

Developing an argument for including history of science in the school science


curriculum should entail more than strenuous efforts by those with a keen interest
in history seeking a suitable role for it. The curriculum is already overloaded with
content and little is to be served by further addition unless the additional (histori-
cal) perspectives can enhance the achievement of key learning goals or assist stu-
dents in reaching understanding that is difficult or impossible to achieve in other
ways. Of course, one could argue for history of science as an alternative to hands-
on activities in situations where experiments are difficult to perform, expensive,
dangerous, unacceptably time-consuming or ethically inadmissible, but that is no
more an argument for deploying history of science than for using computer simula-
tions or watching videos. What is needed is an argument that focuses specifically
on what history of science offers in the way of a distinctive learning experience.
Such arguments have a long history in school science education (Sherratt, 1982,
1983; Brock, 1989), though the nature of the argument, and the success of its advo-
cates in convincing curriculum developers, has shifted substantially over the years.
Not surprisingly somewhat different arguments are advanced by different ‘stake-
holders’ – that is, by historians of science, philosophers of science, sociologists of
science, scientists and science educators.
Among the most prominent, persistent and successful arguments is that history of
science promotes student learning of science – that is, it assists concept acquisition
and concept development. In his splendid overview of these arguments, Matthews
(1994) gives prominence to the views of Ernst Mayr (1982):
I feel that the study of the history of a field is the best way of acquiring an
understanding of its concepts. Only by going over the hard way by which
these concepts were worked out – by learning all the earlier wrong assump-
tions that had to be refuted one by one, in other words by learning all past
mistakes – can one hope to acquire a really thorough and sound understand-
ing. In science one learns not only by one’s own mistakes but by the history
of the mistakes of others. (p. 20).
There are at least three important points here. First, one learns from recognizing,
analyzing and correcting one’s own mistakes; second, one learns from descriptions
of the mistakes of others; third, a thorough understanding of conceptual niceties
can only be acquired by putting oneself in the position of the ‘pioneers’ of the dis-
cipline and appreciating the specific problems with which they grappled. A fourth,
closely related argument, deriving from Piaget’s (1970b, 1972) writing on genetic
epistemology, is that individual learning histories in science tend to mirror the his-
torical development of scientific concepts (Driver & Easley, 1978; McCloskey,
149

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
CHAPTER 8

1983; Nussbaum, 1983, 1989; McDermott, 1984; Hashweh, 1986; Vosniadou &
Brewer, 1987; Nersessian, 1989; Mintzes et al., 1991; Gil & Solbes, 1993; Tsaparlis,
1997; van Driel et al., 1998). If this is the case, it follows that knowledge of the
historical development of a discipline can sometimes help teachers to anticipate,
understand and deal with the conceptual difficulties that students are likely to en-
counter and the kind of misconceptions they are likely to hold (Wandersee, 1985;
Steinberg et al., 1990; Barker, 1995). Studying episodes in the history of science
will indicate the kind of questions teachers should ask, the observational evidence
they should seek and the experiments they should conduct in order to enhance stu-
dent learning. In that sense, history of science is just as much for teachers as it is
for students.
As Sequeira and Leite (1991) point out, history of science is particularly helpful
in identifying parallels between the problems that accompany conceptual change
strategies with students and the historical resistance to new ideas and the overthrow
of existing ideas. An explicit comparison of historical misconceptions with current
scientific explanations helps both teachers and students to understand the ways in
which common sense everyday understanding frequently reinforces misconcep-
tions and, thereby, lays a good foundation for challenging and changing students’
own misconceptions in a non-threatening way. It is both intriguing and comforting
for students to discover that eminent scientists of the past had views similar to their
own. They realize that it was a perfectly good and sensible view to hold, even
though it later turned out to be incorrect, and so they are better placed, both cogni-
tively and emotionally, to engage in critical confrontation of other ideas they hold.
Campanario (2002) extends Sequeira and Leite’s (1991) argument to embrace a
thorough study of the scientific community’s resistance to new ideas and episodes
of ‘delayed recognition’ in order to shed light on the problems of conceptual change
in students. Many ‘good ideas’ in science were initially rejected through simple
errors of judgement and because of ‘conceptual conservatism’, metaphysical and
religious objections, concerns about methods of data generation, intuition, per-
sistent simplistic assumptions (e.g., that proteins, rather than DNA, must be the
carrier of genetic information on grounds of their structural diversity) and pro-
fessional rivalries. Studying such episodes can help students to reflect on, and per-
haps gain greater control over, their own learning. Such metacognitive skills will,
of course, enhance future learning, not least because students will have more con-
fidence when facing the emotional demands of changing their own conceptions.
Rudolph and Stewart (1998) advocate a similar approach: studying opposition to
Charles Darwin’s The Origin of Species as a way of shedding light on the difficulties
that some students (and some adults) experience in accommodating the notion of
evolution by natural selection. Similarly, Gauld (1998) urges teachers to present
original arguments by Newton and his contemporaries as a means of rendering
Newton’s Third Law more plausible to students.
A somewhat different version of a parallelism argument between the history of
science and individual learning histories was made in a report, Science Teaching in
Secondary Schools, by the British Association for the Advancement of Science
(BAAS, 1917), cited by Sherratt (1982) and Jenkins (1989). Drawing on the ideas

150

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
MAKING A CASE FOR HISTORY OF SCIENCE

of Percy Nunn, the report claimed that both proceed through three phases of interest
in understanding the natural world: wonder, utility and systematizing knowledge for
the purposes of intellectual satisfaction. At about 11 years old, children respond to
“the direct appeal of striking and beautiful phenomena”; then, after discovering and
appreciating the utility of scientific knowledge, they are ready (at age 17, or there-
abouts) to accept and study the systematization of knowledge. Sherratt (1982) also
cites a passage from the Spens Report (Board of Education, 1938) that clearly illus-
trates the application of Nunn’s views to the history of electricity.
(It began) with a period of wonderment and delight in marvellous and bizarre
phenomena for the first time brought to light… it passed to the exploitation of
electricity in the service of man… and was completed by the contemporary
phase – initiated by the great work of Clerk Maxwell – in which the physicist
seeks to construct a picture of the whole material world in terms of electrical
entities. (Spens Report, 1938, cited in Sherratt, 1982, p. 228)
However, it is important not to make too much of these kinds of parallel. There
are some very important differences between children’s naïve thinking in the 21st
Century, as they struggle to learn existing science, and the earlier speculations of
scientists as they struggled to create that science, largely due to very significant
differences in social, cultural and intellectual contexts. As noted in chapter 7, how
one thinks and the worldview that underpins beliefs and values are just as much a
product of one’s time and place as what one thinks and what one thinks about (i.e.,
the theories to which one subscribes).1 Added to which there are, of course, major
differences in experience and maturity of intellectual judgement that result in sub-
stantial differences in the way conclusions are reached and in the extent to which
metacognitive awareness impacts the thinking processes (Vosniadou & Brewer,
1987; McCloskey & Kargon, 1988; Gauld, 1991). Moreover, to reiterate a point
made earlier in this book, doing science and learning science are significantly dif-
ferent activities: scientists struggling to create new concepts and conceptual struc-
tures to solve the theoretical and empirical problems confronting them are engaged
in a very different exercise from students struggling to use existing concepts, or
variations of them, to solve problems set by teachers. While scientists’ past views
were likely to have been embedded in a complex, wide ranging and coherent expla-
natory system, children’s views are just as likely to be part of an assembly of
‘minitheories’ (Claxton, 1991), each of which is generated to engage satisfactorily
with particular phenomena and events. In pointing to other traps for the unwary,
Duschl et al. (1992) remind us that teachers attempting to use an historical approach
on the basis of parallelism are faced with a crucial decision about whether to
present a theory in the ahistorical context of justification (without regard to pre-
decessor theories) or in the historical context of discovery and development, argu-
ing that the former is more appropriate to episodes of normal science and the latter to
scientific revolutions. They also suggest using “the context of justification whenever
new theory requires only weak restructuring on the part of the student and the con-
text of development when it requires radical restructuring” (p. 28).

151

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
CHAPTER 8

Nevertheless, despite all these caveats and difficulties much is to be gained by


leading students through part of that historical journey by means of carefully selected
episodes and topics (‘historical vignettes’, as Wandersee (1990) calls them), using
curriculum materials that pitch the level of cognitive demand at an appropriate
level for the students.
The legitimate, sure, and fruitful method of preparing a student to receive a
physical hypothesis is the historical method. To retrace the transformations
through which the empirical matter accrued while the theoretical form was
first sketched; to describe the long collaboration by means of which common
sense and deductive logic analyzed this matter and modelled that form until
one was exactly adapted to the other: that is the best way, surely even the
only way, to give to those studying physics a correct and clear view of the
very complex and living organization of this science. (Duhem, 1962, p. 268)
In developing their notion of Large Context Problems (LCPs), Stinner and
Williams (1998) add a motivational element to claims for the educational benefits of
including historical studies in the curriculum – the excitement of sharing “the frus-
trations and rewards of the intellectual struggles of those who have made important
scientific discoveries” (p. 1029). LCPs are ‘science stories’ built around powerful
unifying ideas chosen for their curriculum saliency. Stinner and Williams (1993)
argue the value of LCPs on two grounds: first, well-designed LCPs make very
effective curriculum resources; second, their construction can play an effective role
in sensitizing student teachers to the value of historical studies and awakening
awareness of students’ alternative frameworks of understanding. In the context of
teacher education, they say, the approach works best when the stories are created
collaboratively by instructor and students in conformity with the following guide-
lines:
– Map out a context with one unifying central idea both important to science and
likely to capture student interest.
– Provide experiences that can be related to the students’ everyday world at a
level and in a way that makes sense to them.
– Create and develop a storyline that will dramatize and highlight the main idea,
even though the story elements may not all be historically accurate.
– Ensure that the main ideas, concepts and problems to be addressed arise ‘natu-
rally’ from consideration of the context.2
– Ensure that the concepts are ‘diversely connected’ (i.e., within the storyline and
with everyday experience) and that there is scope for individual extension and
elaboration.
These features of good LCP design reflect Kenealy’s (1989) assertion that tell-
ing a coherent story, with a beginning, a middle and an end, and in which the char-
acters’ actions are seen to make sense in relation to the overall theme, may be the
best way of conveying complex ideas: the key, he says, is “a careful construction,
an understandable rationale, and minor and major closures, all told while paying at-
tention to making this technical construction interesting and exciting” (p. 214).

152

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
MAKING A CASE FOR HISTORY OF SCIENCE

Similar pedagogic motives and design principles underlie the detailed case study of
the controversy between creationism and evolution developed by Michael Ruse
(1989), the teaching unit on atomic theory (Thales to Dalton) designed by Jutta
Luhl (1992), the history of sickle cell anaemia used to teach population genetics
(Howe, 2007), and the use of historically significant experiments by Seroglou et al.
(1998) to clarify student thinking about electromagnetism. An interesting variation
on this theme is Lochhead and Dufresne’s (1989) idea of using ‘dialogues’ be-
tween important scientists – for example, between Aristotle and Galileo – as a way
of elucidating conceptual difficulties and illustrating the chain of argument con-
necting theory and observation. More recently, Hoadley and Linn (2000) have de-
scribed an approach simulating ‘historical debate’ (between Kepler and Newton)
using online asynchronous discussion.

HUMANIZING SCIENCE AND BRIDGING THE GAP

A second well-used argument for the inclusion of history of science in the school
science curriculum is that it humanizes science and science education. Jenkins
(1989, p. 25) identifies two principal reasons why this claim was “pressed with par-
ticular vigour in the years immediately following the First World War and, more
generally, during the inter-war period”, particularly in the United Kingdom. First,
the rapid growth in the history of science as an academic discipline; second, the re-
action to what many saw as the ‘prostitution’ of science, particularly German sci-
ence, in pursuit of military goals. He cites the work of Richard Livingstone (1916)
criticizing contemporary science education because it “tells us hardly anything
about man. The man who is our friend, enemy, kinsman, partner, colleague, with
whom we live and have our business, who governs or is governed by us, never
comes within our view” (cited in Jenkins, 1989, p. 25). Close to a century later, this
is still the case for the science education of many students; science is still presented
in a way that is austere, depersonalized, authoritarian and remote from personal
experience History of science can go some way towards changing this unattractive
face of science education by showing students that scientists are people, too, with
all the hopes and fears that drive the rest of us, and by showing them that science is
conducted for and in the interests of people.3 As Matthews (1994) remarks, “the
lives and times of the great and not-so-great scientists are usually full of interesting
and appealing incidents and issues that students can read about, debate and reen-
act… History is a way of putting a face on Boyle’s Law, Ohm’s Law, Curie’s dis-
coveries, Mach bands, Planck’s Constant and so on” (p. 52). History of science
shows students that the concepts and theories of science were developed by par-
ticular people, at particular times, in response to particular problems, needs and
interests. The ‘big ideas’ of science did not just appear, fully developed, sometime
in the recent past.
Shedding light on the motives, feelings, thoughts, commitments, apprehensions,
triumphs, failures, mistakes, changes of plan and struggles of scientists is a way

153

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
CHAPTER 8

of making science more inviting and more accessible. For example, by exploiting
their natural interest in people, students can be subtly led to an interest in what
scientists do, and why and how they do it, thus countering some of the current dis-
inclination to study science. As I intimated in the Preface to this book, my own
interest in science was triggered by the historical vignettes and anecdotes included
in school science textbooks written by E.J. Holmyard and F. Sherwood Taylor,
though more systematic theoretical arguments and empirical research data on the
ways in which history of science can impact favourably on attitudes to science,
especially for girls, are provided by Russell (1981), Nielsen and Thomsen (1990),
King (1991) and Thomsen (1998).4 By providing more authentic information about
the realities of doing science and by helping them to recognize that scientists are
not ‘super human geniuses’ but ordinary people, just like them, some students
may come to realize that they can become scientists, too, provided they have suffi-
cient determination and persistence. In fact, history of science as a kind of careers
education was listed by two respondents in Galili and Hazan’s (2001a) study of
Israeli teachers’ motives for including history and philosophy of science in the
curriculum.
A related argument is that history of science, especially historical accounts of
20th and 21st Century science, can be an effective way to bridge the ‘gap’ between
the two cultures of arts and sciences so lamented by C.P. Snow (1962) – thus
simultaneously broadening the perspectives of future scientists/engineers and
ensuring that future politicians and business leaders have some basic understanding
of science, scientists and scientific development.5 As Conant (1951) asserted, more
than a half-century ago, all citizens should have a robust understanding of the rela-
tionships among government, industry and commerce, education, and issues of
scientific research and development. In a sense, this is the argument advanced in
chapter 1 that some basic knowledge of the history of science (and knowledge of
the philosophy and sociology of science) is an essential component of scientific
literacy for citizenship and assists the development of the necessary skills for
addressing current policy and potential future developments in a more critical way
– that is, serious and critical consideration of questions about environment, health,
natural resources and the like, informed confrontation of ethical issues and realistic
preparation for change brought about by technological innovation. Of course, pro-
viding common ground for Arts and Science specialists to study is not without its
problems: a proper appreciation of the historical development of major ideas in
science requires both substantial scientific understanding and substantial historical
sensitivity, and questions must be asked about whether either party would be suffi-
ciently well-equipped to engage in serious study. Added to which is the problem
that many science specialists are depressingly narrow in their interests, and may be
reluctant to engage seriously with material that does not contribute directly to their
mastery of science content, while arts specialists may exhibit an initial aversion to
studying anything relating to science because of previous unhappy and unsuccess-
ful experiences with scientific content. In crude terms, science specialists may have
little interest in history and non-science specialists may have no more interest in
history of science than in science itself. Schwartz (1995) draws on thirty years of

154

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
MAKING A CASE FOR HISTORY OF SCIENCE

experience with such courses to proffer some sage advice on course structure and
content, and to remind us that developing and teaching successful inter-disciplinary
courses requires “imagination, self-confidence, a willingness to learn and to take
risks, a fairly thick skin, and some resiliency with which to respond to the inevi-
table failures” (p. 1040). For Kauffman (1989), the value of such an attempt is in
achieving some balance between the scientist’s emphasis on abstractions and uni-
versals and the historian’s and sociologists’s emphasis on localized contextual
details and particularistic meanings. “Seen in the unifying light of the Tai Chi”, he
says, “apparent opposites disappear, and, paradoxically, disadvantages can be seen
as advantages” (p. 87).

THE NATURE OF SCIENCE

A fourth argument is that history of science provides rich insight into the nature of
science and scientific inquiry, an argument that lay behind James Conant’s famous
Harvard Case Histories in Experimental Science (1957), designed to give students
a feel for “the tactics and strategies of science”, and the much lamented Harvard
Project Physics course (HPPC) for US secondary schools (Rutherford et al., 1970).
Essentially the same point was made by John Ziman (1980) in his argument for
history as a means of dispelling myths about scientific inquiry and as an antidote to
naïve scientism because it shows the scientist as “a real person, the child of his
times, vainly seeking glory and all the rest of it. It can illustrate by example all that
one might wish to say about the personal dimension of research” (p. 121). Of
course, history of science and philosophy of science are mutually enriching: just as
history of science informs, illuminates and exemplifies the philosophy of science,
so the philosophy of science assists us to make better sense of the history of
science. As Lakatos (1978) says: “Philosophy of science without history of science
is empty; history of science without philosophy of science is blind” (p. 102).
Case studies in the history of science can illustrate many of the points made in
chapters 3 to 6, such as the relationship between observation and theory, the the-
ory-dependence of experimentation, and the role and status of theories and models
(Arons, 1983; Kauffman, 1989; Betts, 1992; Lin & Chen, 2002; Wang & Marsh,
2002). They show that scientific knowledge is, in principle, refutable and therefore
constantly under development and subject to change. They show that science is not
conducted in an intellectual, social and cultural vacuum; rather, it occurs in a cli-
mate of constant criticism and debate, it is subject to economic and political forces,
and it is often dependent on (as well as productive of) technological innovation.
Historical case studies reveal the part played in scientific inquiry by intuition, luck,
competition, technological ingenuity and sheer hard work, and illustrate how scien-
tists are often dogmatically tenacious in pursuit of their views, even in the face of
contradictory evidence. Above all, they reinforce Feyerabend’s (1975) point that
there is no one universally applicable, simple and straightforward method of con-
ducting scientific investigations. By showing that at different times in history there
have been significant differences in what counts as an appropriate research ques-
tion, what constitutes a satisfactory investigation or experiment, what makes a

155

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
CHAPTER 8

particular concept acceptable and what counts as a satisfactory explanation, histori-


cal studies contribute to a deeper understanding of the nature of science. By draw-
ing attention to significant changes in the way observational evidence is collected,
processed and analyzed, and in the way scientific work is reported and dissemi-
nated,6 they assist students in developing their understanding of and ability to use
the discourse of science – arguably the most significant aspect of scientific literacy
for citizenship (see chapter 1).
By studying episodes in the history of science, especially through original
papers, and by reading biographical and autobiographical material, students can
gain insight into the ways in which particular scientists thought about conceptual
and procedural issues, designed investigations and went about solving problems.
They are given graphic illustration of the interplay of theory and experiment: scien-
tists design new experiments to counter an opponent’s theoretical position and the
opponent counters with a modified theory, a new experiment, or a technological
innovation capable of making more accurate readings or gathering an entirely new
kind of observational evidence. There may be value, once in a while, in students
replicating experiments that were highly significant in those respects (Kreuzman,
1995; Heering, 2000; Hottecke, 2000) or engaging in role play (Allchin et al.,
1999). History also shows us that there is a difference between the way science is
done and the way science is reported. It shows us that the creative phase is rule-
breaking (just as Feyerabend claims), while the validation phase has to conform to
the strict standards of the community of practitioners.7
I must compare myself to a mountain climber, who without knowing the way
climbs up slowly and laboriously, must often turn around because he can
go no further, discovers new trails, sometimes through reflection, sometimes
through accident, which again lead him forward a little, and finally, if he
reaches his goal, finds to his shame a Royal Road on which he could have
traveled up, if he would have been clever enough to find the right beginning.
(Helmholtz (1892), cited in Stuewer (1998), p. 16)
While these insights into the nature of scientific inquiry are valuable in their own
right as a contribution to learning about science, they may also have transfer value
and lead to improved capacity for doing science and for constructing meaning. This
is what Seroglou and Koumaras (2001) mean when they argue that history of sci-
ence has a major role in the science curriculum because it contributes to improved
problem-solving. What they have in mind is an expanded notion of problem solving
that includes generating new conceptual models or improving, revising and devel-
oping existing ones in order to ensure a better fit between theory and observational
data, especially experimentally acquired data (see also, Stewart & Hafner, 1991).
In studying how particular conceptual models and theories were developed and
established, students quickly realize that the supposedly unambiguous facts recorded
in textbooks have not always been so unambiguous, nor have our taken-for-granted
explanations of phenomena and events always been so well-established. Facts are
not individual pieces of data but interpretations of observations linked in meaningful
ways into explanatory systems, and underlying and underpinning these complex

156

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
MAKING A CASE FOR HISTORY OF SCIENCE

conceptual frameworks is a host of assumptions and a wealth of interpreted data,


many of which are not apparent on first encountering them. Some models and theo-
ries are so familiar that their fundamental assumptions remain hidden, their inade-
quacies are glossed over, their layers of embedded meaning remain unexplored and
their misleading aspects ignored. Historical studies enable theories to be ‘unpacked’
to reveal their components and the influences that led to their construction, thereby
leading students to a deeper understanding of how theoretical systems are con-
structed and what criteria must be satisfied in order for them to be validated (Hagen
& Kugler, 1990; Castro & Carvalho, 1995; Weck, 1995; Justi & Gilbert, 1999).
Many important questions are raised: What constitutes a well-built theory? How
are convincing arguments made? What factors are likely to influence acceptance
and rejection by the community? It is evident that the simpler and more elegant the
model, the more likely it is to be believed, and the more intuitive the model, the
harder it will be to displace it or modify it in the light of new data, especially if it
is reinforced through the shared understanding consequent upon membership of other
sub-cultural groups. Kipnis (1996) argues that this approach becomes considerably
more effective when reinforced by what she calls “investigative experimentation”.
In a sense, she is promoting the view that students learn science by doing science or
by imitating scientists, as she describes it.8 In fact, the scientific inquiry in which
students participate is carefully reconstructed to reflect historical development and to
highlight specific conceptual issues, so perhaps learning science (and learning about
science) by imitating particular scientists, but using modern apparatus and simplified
experimental design, is a more apt description of the approach. In some ways, it
resembles teacher demonstrations and computer simulations rather more than the
open-ended scientific inquiry suggested by the term doing science, despite claims
by Kipnis that she is trying to get away from teacher demonstrations.
Lawrenz and Kipnis (1990) make some bold claims for the success of the
‘historical-investigative’ approach, including enhanced conceptual understanding and
thinking ability, increased enjoyment of learning and greater interest in experimenting.
While these claims should be viewed cautiously, there are good grounds for regarding
this cluster of approaches as well placed to generate interest by showing students that
science is often controversial, thereby stimulating curiosity in contemporary scien-
tific debates and encouraging students to scrutinize their own beliefs and the beliefs
of others by challenging underlying assumptions and critiquing evidence for them.
Allchin et al. (1999) have shown that attitudes to science are positively impacted
by courses for non-science majors that include historically-contextualized and rela-
tively open-ended investigations of science problems. It might also be expected that
students will become more confident in their reasoning ability and better prepared
for discussion of controversial issues in science – a crucial part of scientific literacy
(see chapter 1). Furthermore, one might expect enhanced metacognitive skills
and self-regulatory cognitive strategies that enable individuals to reflect on, con-
struct meaning from, monitor and regulate their own learning (Duschl & Erduran,
1996). Ultimately, success will depend on the way in which teachers implement
the approach and, in particular, on the sensitivity with which they deal with student
debate of ideas. Monk and Osborne (1997) provide a wealth of invaluable advice

157

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
CHAPTER 8

in their 6-step model combining historical material, detailed consideration of issues


in philosophy of science and constructivist pedagogy.
– Presentation of phenomena or events – carefully chosen for their historical sig-
nificance, inherent interest and capacity to generate questions, problems and
predictions.
– Eliciting students’ views, ideas and explanations – using the usual range of con-
structivist strategies: skilful teacher questioning, use of concept maps, word asso-
ciation, art work, group discussion, writing activities of various kinds, etc.
– Historical studies – presented orally by the teacher or via text, multimedia, stu-
dent research, museum visits, and the like.
– Devising suitable tests – designing experiments and other kinds of investigations
to evaluate and judge the range of ideas collected.
– Presenting the current scientific view – if it has not already arisen, the current
view (or the school science curriculum version of it) should be introduced using
whatever means is appropriate, with particular emphasis on the reasons why this
view has been accepted.
– Review and evaluation – reviewing learning progress (an essential element of
constructivist pedagogy) and reiterating the basic thrust of the approach – that
is, how to distinguish between justified and unjustified beliefs.
The value of this approach is that it shifts emphasis from the mere acquisition of
theoretical knowledge towards a concern with the ‘warrant for belief’: How do we
know? What are the reasons for holding this belief? What is the evidence for this
view? How was that evidence acquired? Interpreting theoretical disputes from the
history of science requires students to adopt rival perspectives on the interpretation
of data, and to enter discussion about what counts as relevant data and appropriate
ways of collecting them. Thus, it is excellent preparation for appraising new theo-
retical propositions in relation to existing ideas. By studying and understanding
the ways in which scientists construct knowledge, students are assisted in building
metacognitive insight into their own knowledge construction. By utilizing histori-
cal material and emphasizing epistemological concerns, the approach meets several
of the learning about science goals embedded in the notion of scientific literacy
discussed in chapter 1, and simultaneously addresses one of the affective issues
surrounding constructivist pedagogy – that is, the reluctance of some students to
contribute their own ideas. Perhaps it is comforting to know that others, including
some eminent scientists, have held erroneous beliefs; perhaps contributing ideas to
class discussion can be seen not so much as a potential threat to self-esteem as an
opportunity to enrich discussion. Perhaps it is not too extravagantly optimistic to
speculate that successful experiences of this kind might lead to increased tolerance
of and appreciation for the views of others – a matter central to the role of history
of science in multicultural and antiracist education (see below). Support for these
speculations can be found in Crawford’s (1993) description of how she has used
Michael Faraday’s experimental diaries to create a classroom climate that encour-
ages students to (i) articulate their anxieties and uncertainties, (ii) develop the emo-
tional commitment to struggle with theoretical and practical problems and (iii)
learn to think independently.

158

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
MAKING A CASE FOR HISTORY OF SCIENCE

STS, CURRICULUM INTEGRATION AND CULTURAL ENLIGHTENMENT

Historical case studies are ideal vehicles for teaching about the complex interactions
among science, technology, society and environment (STSE education). Clearly,
such things as the role of market forces and military interests, patents and the need
for secrecy, problems of scaling up from the research scale to the production scale,
issues of safety and social responsibility, environmental impact, and the ever-present
moral-ethical concerns, are all easier for students to comprehend when they are set
in a real life context. Moreover, studying episodes from the past enables students to
reflect on such issues with a sense of perspective that is not always possible with
current events and controversies. By judicious choice of examples, teachers can de-
monstrate occasions when (i) science precedes technological application, (ii) tech-
nology precedes the science that eventually explains it, (iii) scientific development
and technological development are entirely independent, and (iv) scientific and
technological development are mutually dependent and interactive. Layton (1988)
makes the point that even when technology is ‘just’ applied science (item (i) above),
there is still a lot of hard work to be done. He gives the example of Perkin’s devel-
opment of the synthetic dye Mauve, and says that its translation into a commercial
product required “knowledge, skills and personal qualities very different from those
needed for the test tube oxidation of aniline” (p. 371). At the outset, Mauve would
not take evenly on large batches of cloth, there was no suitable mordant for cotton,
raw materials were not readily available, handling concentrated nitric and sulphuric
acids on a ‘factory scale’ presented all manner of engineering problems, there were
problems of marketing associated with consumer reluctance, and so on. Notions
such as optimization, feedback modelling, systems analysis, critical path planning
and risk assessment have to be included whenever science is applied to ‘real world’
situations.
Of course, real world problems are rarely the simple matters of cause and effect
portrayed in traditional science curricula; rather, attempts at solution often reveal
layers of increasing complexity and uncertainty that cannot be contained within a
particular disciplinary framework. Problems in science and technology become in-
extricably linked with considerations in economics, politics, aesthetics and moral
philosophy. In general, a scientific solution is valid/acceptable if it conforms to the
rationality of science – i.e., if it has observational or experimental support, if it is
internally consistent, if it is consistent with other accepted theory, and so on. It
helps if it is elegant and parsimonious, though these criteria may not be considered
essential (see earlier chapters and chapter 9). In technology a solution has to work,
of course, but it also has to be efficient, cost-effective, durable, and perhaps aes-
thetically pleasing as well. There may also be critical considerations relating to social
and environmental impact. Technologies are rarely ‘good’ in an absolute sense.
Rather, they are good from some perspectives, less good (or even undesirable)
from others. In that case, whose perspective is to count, whose interests are to be
served, whose values are to be upheld? One person’s acceptable risk or cost is an-
other person’s intolerable hazard, social disruption or cultural insensitivity. Tech-
nology is inescapably a social activity, determined by the prevailing distribution

159

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
CHAPTER 8

patterns of wealth and power. History of science and technology provides the
means for the curriculum to acknowledge these realities.
Of course, the extent to which the curriculum depiction of scientific practice
should be politicized, are matters of continuing debate and considerable contro-
versy (Solomon & Aikenhead, 1994; Yager, 1996). Somewhat surprisingly, given
the caution with which American educators often approach sociopolitical matters,
Heilbron (1987) has suggested that students should study, in historical context, the
“institutions and apparatus of science and the relations among the sciences, govern-
ment, universities, and the military” (p. 554).9 While only a few science teachers will
be willing to follow this advice, many more may be willing to use historical case
studies to show how and why scientists have often disagreed on fundamental issues,
thus laying important groundwork for (i) making the point that controversy is a
perfectly normal part of science, (ii) illustrating how the course of science has often
been profoundly influenced by the sociocultural context in which it is located, and
(iii) identifying and confronting some of the complex ethical issues generated by
scientific developments (Dunn, 1993). Clearly, history of science has a key role to
play in dispelling the all-too-prevalent myth that science is conducted in a sociopoli-
tical and moral-ethical vacuum. By illustrating how science is produced within par-
ticular social, economic, historical and technological contexts, history of science
makes a major contribution to critical scientific literacy, as defined in chapter 1
(Lemke, 2001; Allchin, 2004). This particular role for the history of science is encap-
sulated in a 19-year old statement from the Department of Education and Science
(1989):
Pupils should be given opportunities to develop their knowledge and under-
standing of how scientific ideas change through time and how their nature
and the use to which they are put are affected by the social, moral, spiritual
and cultural contexts in which they are developed. In doing so they should
begin to recognize that while science is an important way of thinking about
experience, it is not the only way. (p. 32)
It is the flavour of this passage that underpins my sixth argument for the history
of science: its role in multicultural and antiracist education (Hodson, 1993d, 1999;
Hodson & Dennick, 1994). Studies in the history of medicine, astronomy, optics
and technology are particularly rich in examples of the achievements of Chinese,
Indian, Arab and African scientists. Sadly, much of that history is ignored or mis-
represented in school science textbooks. In lamenting this distorted history, Dennick
(1992a) comments that “throughout European history since the renaissance there
has been a tendency to disparage and down grade the discoveries and achievements
of other cultures and historians have been very prone to give credit where it is not
due” (p. 81). In evidence, he cites the assertion by Joseph Needham (1954, 1969)
that the three greatest inventions of the previous millennium (paper-making and
printing, gunpowder, and the navigational compass) were each used in China several
centuries before their supposed discovery by Westerners. Oddly, Chinese science is
often dismissed as ‘mere technology’, thereby creating the impression that it was

160

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
MAKING A CASE FOR HISTORY OF SCIENCE

little more than haphazard trial and error (‘suck-it-and-see’, as my fellow Lancas-
trians would say) or, at best, a series of chance discoveries.
Throughout ancient Chinese, Egyptian, and other cultures of the time, there
was, in effect, no real attempt to understand or to generalize. If devices or
explanations of the universe worked, they were accepted as best or true.
(Loving, 1997, p. 440)
Needham (1981) tells us that there is considerable evidence to the contrary:
Chinese science was theoretically driven, though its philosophical basis was radi-
cally different from that of Western science; it involved systematic observation and
careful experimentation.
Indian, Arab and African scientific achievements have been similarly trivialized
or falsely attributed to Westerners. For example, when the Arab contribution to the
growth of science is mentioned at all (and that is only very rarely), it is not seen
as extending beyond the role of custodian of ancient Greek knowledge, which
was later passed back to European scholars. In reality, what was passed on was
a substantial corpus of new scientific understanding and practice (Winter, 1952;
Hill, 1993; Hobson, 2004; Saliba, 2007) and “a sophisticated culture uniting art,
religion and science in a profound world view which is still very much alive today”
(Dennick, 1992b, p. 6). Ignoring or falsely attributing the work of Muslim scientists
to Europeans is a practice that, according to Sardar (1989), has been going on for
centuries: “Piracy was so common that as early as the twelfth century a decree was
issued in Seville forbidding the sale of scientific writings to Christians because
the latter translated the writings and published them under another name” (p. 10).
A similar fate befell much of Indian science (Machwe, 1979; Kumar & Kenealy,
1992). In addition, the agricultural theory and practice of the pre-Columbian peo-
ples of the Americas were subsumed within European science without any acknow-
ledgement (Weatherford, 1988). Even more serious in the context of arguments
for multicultural and antiracist education is the systematic trivialization, distortion
and suppression of African cultural history and its scientific and technological
achievements – a key element, of course, in the racist ideology that was formerly
used to legitimize slavery and colonial exploitation, and still serves to deny a strong
sense of cultural identity to many people of African descent living outside Africa.
Despite the spectacular achievements of the ancient civilizations of Ethiopia, Benin
and Zimbabwe, for example, the myth is still propagated that significant African
history only began with the Imperialist invasions. Moreover, the great civilization
of Egypt (and its achievements in science and technology) is presented without any
acknowledgement of its African roots. It almost goes without saying that the goals
of multicultural/antiracist science education require that we develop more authentic
historical studies for use in primary and secondary schools.
A seventh argument is that history of science assists curriculum integration,
always a powerful argument with primary school (elementary school) teachers.
Science developed alongside and in tandem with mathematics, technology, theo-
logy, philosophy, psychology and commerce; it was (and continues to be) affected

161

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
CHAPTER 8

by and influenced by each of the fields, and by politics, religion, ethics, literature,
art, music and all forms of mass culture. Thus, history of science can provide stu-
dents with access to a deep understanding of cultural and intellectual interaction, the
kind of interaction highlighted in James Burke’s superb and highly idiosyncratic
television series Connections and Jacob Bronowski’s classic series The Ascent
of Man.10 Matthews (1994) makes this particular argument at some length, and
very persuasively, with respect to Galileo’s physics. He concludes: “A historical
approach to science allows students to connect the learning of specific scientific
topics with their learning of mathematics, literature, political history, theology,
geography, philosophy and so on. When the richness of science’s history is appre-
ciated, then collaboration between science, drama, mathematics, history and religious
studies teachers in schools can fruitfully be encouraged” (p. 53). While cross-
disciplinary themes have been relatively common in primary/elementary schools in
the past, they are increasingly under threat from the ‘accountability movement’,
with its emphasis on increasingly analytical forms of assessment.11 In secondary
and tertiary education, cross-disciplinary approaches have always been rare, and
whenever implemented have tended to be the province of particular enthusiastic,
energetic and charismatic individuals willing to work tirelessly in pursuit of a
‘dream’ in order to assemble a suitable team of teachers. Particularly striking illus-
trations of what can be achieved can be found in Samson and Weininger’s (1995)
description of their integration of history of science and art history, together with
relevant social history, philosophy, literature and psychology, into a course for en-
gineering students titled “Light, Vision and Understanding” and Schwartz’s (1995)
account of a multidisciplinary course on the history of alchemy.
The eighth and final argument to justify the inclusion of history of science in the
school science curriculum rests on the simple assertion that knowing science has a
history, that it didn’t just appear in sophisticated and developed form at some time
in the 20th Century, is an essential component of scientific literacy and of a sound
liberal education.
Such puzzling concepts as force, energy, etc., are man-made and were evolved
in an understandable sequence in response to acutely felt and very real problems.
They were not handed down by some celestial textbook writer to whom they
were immediately self-evident. (Cardwell, 1963/1964, p. 120, cited in Brush,
1974, p. 1163)
Sometimes students are genuinely surprised that people ‘back then’ were conduct-
ing experiments and building theories. Perhaps the creation of a sense of scientific
history is the only justification that we need – in a sense, history of science for its
own sake. Science is an important part of our cultural heritage and knowing some-
thing of its history and development is an important part of what it means to be
a cultured and educated citizen (Arons, 1983; Jenkins, 1989; Bybee et al., 1991;
Millar, 1993, 1996; Wellington, 2001).12 As Monk and Osborne (1997) argue, teach-
ing science and teaching about science in an ahistorical way makes as little sense
as studying Shakespeare without discussing the sociocultural context of Elizabethan

162

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
MAKING A CASE FOR HISTORY OF SCIENCE

England. In short, science education is seriously deficient if it fails to provide satis-


factory answers to the following series of questions in ways that are interesting and
accessible to all students:
– Where did science come from?
– How did it get that way?
– Who were the significant people in its development?
– Why did they choose to study these issues and problems?
In Ernst Mach’s (1960) view, those who know the history of science are much
better positioned to evaluate current developments: “They that know the entire
course of the development of science, will, as a matter of course, judge more freely
and more correctly of the significance of any present scientific movement than
they, who, limited in their views to the age in which their own lives have been
spent, contemplate merely the momentary trend that the course of intellectual events
takes at the present moment” (p. 8).

WHAT KIND OF HISTORY OF SCIENCE?

Needless to say, there has been some strong opposition to the inclusion of history
of science in the school science curriculum. First, from those historians who see
history in science lessons as poor history or even as a fabrication of history in sup-
port of current scientific views. Second, from those scientists who see it as divert-
ing attention (and time) from what they see as ‘proper science’. There is also
concern that history of science can have an adverse impact on students by under-
mining the ‘certainties’ of the scientific world (Brush, 1974). There are strong ech-
oes here of Thomas Kuhn’s (1977) point that education in science is primarily
‘indoctrination’ into the prevailing paradigm and his warning that the proper per-
formance of ‘normal science’ and the subsequent capacity to challenge the prevail-
ing paradigm is likely to be compromised by too early a questioning of its
fundamentals. Thus, whether to include history of science in the curriculum, or not,
is not the only question; it is also important to ask what kind of history of science
we should include (if any). A basic distinction is that between internalist accounts
and externalist accounts. The former concentrate solely on the development of
scientific concepts and their role in theoretical explanations, excluding all but
the most cursory consideration of the sociocultural context in which the ideas
were developed and the socio-economic factors that might have motivated their
development. Externalist accounts, on the other hand, seek to describe and explain
the growth of science in terms of the personal circumstances of individual scien-
tists and/or the social and cultural climate of the wider society in which the work
was conducted.
Clearly, the motives for including history of science will be a powerful influ-
ence on the kind of history included. Those concerned to assist conceptual under-
standing may be inclined to interpret scientific history from a 21st Century
perspective, frequently ignoring superseded ideas or regarding them as seriously
misguided – an approach that has been termed “Whiggish” history.13 Such accounts
may seriously distort history by criticizing scientists of the past for failing to

163

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
CHAPTER 8

meet modern standards of data collection and experimental design, and may ridi-
cule those scientists for being unaware of some of our taken-for-granted modern
knowledge. They portray scientific knowledge as emerging in simple and predict-
able fashion from scientists’ struggles to solve theoretical rather than practical
problems, with one scientific development leading directly and inexorably to the
next until the current position was reached. When current theories are taken as the
yardstick, those who initially opposed the ideas that eventually led to those theories
are regarded as incompetent or perverse, while those who accepted and developed
them are credited with exceptional foresight – a kind of ‘villains and heroes’
approach to scientific history. Because it is assumed that there has been one
universally applicable method in use since the outset of the scientific endeavour,
theories that were once accepted but were subsequently falsified are regarded as
the product of scientists’ errors. By evaluating early scientific investigations using
modern criteria and standards, Whiggish historians of science ignore altogether any
appraisals made at the time about whether an experiment was appropriate and reli-
able, whether a theory was intelligible and whether an argument was convincing.
The fact that a belief doesn’t stand up to critical scrutiny now, in the 21st Century,
does not mean that it was irrational to hold such a view at the time it was proposed.
Unencumbered by modern notions of rationality, scientists of the past had
to make decisions about the acceptability of contemporary theories by their
criteria rather than by ours. We may have the hubris to imagine that our theo-
ries of rationality are better than theirs (and they may well be), but how does
it help historical understanding to evaluate the cogency of past theories utiliz-
ing evaluative measures which we know were not operative (not even in
approximate form) in the case at hand? (Laudan, 1977, p. 129)
We can only understand the past in its own terms; the intellectual standards of
the present sometimes have little relevance to a proper understanding of events in
the distant past.14 Accounts that are more faithful to historical circumstances and
sociocultural influences entail consideration of the various by-ways, diversions,
false paths and dead ends of science, recognition that science is frequently complex
and uncertain, and acknowledgement that not all inquiries are fruitful. A ‘proper’
history of science attends to the theoretical and practical problems that motivated
new ideas and new procedures, and takes cognizance of the metaphysics and
worldview prevailing at the time. In these respects, ‘time slices’ rather than ‘verti-
cal history’ may be more appropriate for the curriculum – that is, consideration of
the range of ideas current at any one time, how these ideas were generated and how
they were received, interpreted, modified and utilized in further work. However,
simply noting that an idea was widely accepted at a particular time does not neces-
sarily imply that scientists were sound in their judgement. Taking theory appraisal
and acceptance at face value and invoking social influences as the major causative
factor leads to the kind of relativism discussed in chapter 7. It leaves no way of
judging whether one theory was better than another, no way of judging whether, by
the standards in force at the time, a judgement was well founded. Thus, there is a
problem: How can we judge episodes in the past without imposing and invoking

164

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
MAKING A CASE FOR HISTORY OF SCIENCE

present-day criteria and standards? Laudan’s (1977) response, which will be dis-
cussed at some length in chapter 9, is that scientists behave ‘correctly’ (i.e., ration-
ally) when they choose in favour of those ideas effective in solving the most urgent
practical problems and the most significant theoretical ones. This simple criterion,
he argues, is both trans-temporal and trans-cultural, and to do other than invoke it
is to act irrationally.
The notion of progress is both a distinctive feature of science and a major driv-
ing force for scientific endeavour. Thus, a history of science is necessarily evalua-
tive. “It cannot be merely descriptive but will be required to pass judgement, to
evaluate some ideas negatively as epistemological obstacles which need to be
overcome and rejected, and others positively as epistemological acts of scientific
genius. What is negatively evaluated is… excluded from contemporary scientific
thought… what is positively evaluated continues to play a role in science” (Tiles,
1984, p. 13). It follows that a history of science cannot be written once and for all;
a change in current scientific thinking will precipitate a re-evaluation of past
thought. Thus, there is what Gaston Bachelard (1951) referred to as a “recurrent
history” of science, one that is continually re-told in the light of the present, one
that tells how scientific thinking reached its present state. Recurrent history of sci-
ence is the attempt to show “not merely how we came to the present views but also
why… the reasons for rejecting previous theories, for modifying previous concepts,
and thus the reason behind the acceptance of currently accepted views” (Tiles,
1984, p. 15). The reasons under scrutiny are not primarily the psychological, socio-
logical, economic and political factors that influence science, but the justifications
used by scientists in addressing and making decisions about conceptual and proce-
dural issues. In that sense, this reasoning is also part of the present justification of
our theoretical positions and must be understood and subjected to critical scrutiny
if further progress is to be made.
In order to evaluate the past properly the historian of science must know the
present; to the best of his ability he must learn the science whose history he
plans to write. And it is through this that, whether one likes it or not, the his-
tory of the sciences has a strong connection with the science of the moment.
It is when the historian of science is initiated into the modernity of science
that he is also able to uncover more, and more subtle, nuances in the historic-
ity of science. (Bachelard, 1951, p. 9, cited in Kragh, 1987, p. 92)
Chamizo (2007) describes recurrent history as a focus on “distinguishing the
‘sanctioned’ from the ‘lapsed’. The latter is the history of false paths, of errors and
illusions, of prejudice and mystification, whereas the first… is the history of the
thoughts that continue (in) the science of the present time… Recurrent history
helps us to understand, first the context in which ‘wrong’ ideas were once consid-
ered ‘right’, and second how (and why) such context changed” (p. 207). Attention
is focused on the kind of problems that a theory was designed to solve (and why
those problems were considered important), the extent to which a particular theory
was successful in solving them, and the reasons why previous attempts were not
successful and had to be altered or abandoned. Because recurrent history addresses

165

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
CHAPTER 8

scientific error and its correction it focuses on both the nature and the limitations of
scientific justification.
Perhaps what suits the purposes of science education best is a careful delineation
of the key conceptual issues set within a rich and lively sociocultural context. But,
as Klein (1972) comments, what the science teacher wants of historical examples
is precisely what the history teacher would reject as poor history: the one delights
in a “sharply defined single insight’, the other in “the rich complexity of fact’”
(pp. 12 & 18). History of science in the curriculum is necessarily selective in terms
of extent, scope and style. Sometimes consciously and sometimes unconsciously,
selected history of science can easily become distorted history or distorted science.
Great scientists are often portrayed as ‘super heroes’, possessed of almost super-
human intellectual capability, extraordinary determination and startling resilience
in the face adversity, while scientists who made mistakes are sometimes seen as
fools. More worryingly, history of science can be used (and has been used) to pro-
mote the view that science is the exclusive province of Western males and the be-
lief that the science they produce is infallible, authoritative and not to be
questioned by mere students (or even by citizens in general). In contrast, the focus
on people, social circumstances, false trails and fierce competition between rivals
can easily lead to a distorted and over-simplified version of the science, with little
acknowledgement or recognition of the complexities of key conceptual, philoso-
phical and mathematical issues, thus producing what we might call “historical
noise”.15
A number of researchers report that despite the long tradition advocating history
of science in the school science curriculum (at least in some countries), the histori-
cal development of key ideas in science is generally very poorly presented in sci-
ence textbooks and other curriculum materials, rarely extending beyond a few
names and dates, some pictures, and the odd colourful anecdote (Chiappetta et al.,
1991; Niaz, 1998, 2000; Leite, 2002; Rodriguez & Niaz, 2002; Williams, 2002).
Such accounts not only trivialize the sociocultural dimensions, they also misrepre-
sent the nature of scientific inquiry and theory building (Bauer, 1992; Hodson,
1998b). In Schwab’s (1962) words, they present science as an “unmitigated rheto-
ric of conclusions in which current and temporary constructions of scientific know-
ledge are conveyed as empirical, literal, and irrevocable truths” (p. 24). Allchin
(2003) argues that the use of narrative form, chosen specifically for its appeal to
students, leads inevitably to distortion. All narratives, he says, risk drifting into
myth because of common narrative elements.
– Monumentality – not only are scientists larger-than-life heroic figures, often
working single-handedly, but the situations they face and the problem they solve
are of immense significance.
– Idealization – all nuances are laid aside in pursuit of a straightforward, ‘black-
and-white’ storyline.
– Affective drama – stories are made more entertaining, persuasive and memora-
ble by enhancing emotional and aesthetic elements.
– Explanatory and justificatory narrative – the storyline is always structured to
justify the authority and certainty of the scientific conclusion.

166

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
MAKING A CASE FOR HISTORY OF SCIENCE

Leite (2002) provides a checklist (comprising eight major dimensions) for ana-
lyzing the historical content of science textbooks. Its value extends well beyond the
simple comparison of rival textbooks (useful as that can be for teachers contem-
plating the first time introduction of history and philosophy of science into their
courses) to opportunities for reflection on the diversity and relative priorities of
issues to be included in a history-oriented course. However, despite the valuable
guidance that such a checklist can provide, the development of new curriculum
materials is a complex and difficult task. Achieving the appropriate balance is not
easy. As Stuewer (1998) remarks: “A scholar cannot be constrained by logic and
strive for simplicity and at the same time give full weight to illogicality and strive
to portray complexity” (p. 17).
The problem of selection for the curriculum is not just a matter of what counts
as the ‘best history of science’, it is impacted by consideration of what is in stu-
dents’ best interests. Is it necessarily in students’ best interests for them to be given
an authentic version of the history of science? Introducing history into the science
curriculum opens a Pandora’s box by telling students that scientists do not always
behave as rational, open-minded investigators who pursue knowledge through
carefully controlled experiments and dispassionate consideration of the evidence.
Is a ‘warts and all’ account of science likely to make students more or less enthusi-
astic about science? Is there some risk associated with an approach that reveals
how scientists sometimes steal ideas from others, distort data and cut corners in
their determination to establish the primacy of their work? Is the picture of the sci-
entist emerging from the history and sociology of science even more harmful than
the stereotyped image described in chapter 2? When Stephen Brush (1974) asked
this question, more than a quarter century ago, his answer was that the way scien-
tists behave (according to historians) may not constitute a good role model for
impressionable students.16
These writings do violence to the professional ideal and public image of
scientists as rational, open-minded investigators, proceeding methodically,
grounded incontrovertibly in the outcome of controlled experiments, and
seeking objectively for the truth, let the chips fall where they may. (p. 1164)
De-mythologizing science through historical and sociological studies may lower
the esteem in which scientists are held, so such studies are unlikely to be in scien-
tists’ interests. As Holton (1975) remarks, “the apparent contradiction between the
often ‘illogical’ nature of actual discovery and the logical nature of well-developed
physical concepts is perceived by some as a threat to the very foundations of sci-
ence and rationality itself” (p. 329). Are there situations in which it makes sense to
project a series of stylized and mythical stories about great scientists because they
make science more memorable and more interesting? Are there other occasions
when it is more sensible to ‘re-write’ historical accounts somewhat in order to re-
duce conceptual complexity and focus on sociocultural issues, or to focus attention
more clearly on conceptual aspects without too much distracting social, political
and economic detail (de Berg, 1997). Much depends, of course, on the age and
stage of intellectual and emotional development of the students, and on the ways

167

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
CHAPTER 8

in which different versions of ‘history’ contribute to the notion of critical scien-


tific literacy discussed in chapter 1 and to the drive to re-orient nature of science
teaching.
It may be that the notion of ‘authentic history of science’ is illusory. Scientists
rarely leave detailed accounts of how they came to make their discoveries. As noted
earlier, published accounts are rationally reconstructed after the event, while labo-
ratory notebooks are usually little more than pages of data augmented from time to
time by cursory indications of procedures adopted and critical reflective comments.
Histories created after the events (by historians and biographers) are necessarily
conjectures about the influences, circumstances and chains of argument that led to
particular decisions. Those conjectures are influenced by the author’s understand-
ing of the science and by her/his beliefs about the nature of science, as well as by
knowledge of the sociocultural and political context. In other words, all histories
comprise interpreted evidence. In organizing and interpreting data, historians have
to make decisions about which actions and episodes they regard as significant;
different assumptions about significance lead to different histories. In common
with scientific accounts, historical accounts are empirically under-determined; for
any set of recorded historical data, there may be more than one plausible explana-
tion. Furthermore, it is not always clear what would count as confirming or falsify-
ing evidence, and it is not always clear what counts as reliable data. Just as the
assumptions we make about other times and places influence the significance we
attach to particular historical data and determine the way we interpret events, so the
assumptions that our predecessors made influenced the data they thought it impor-
tant to record. As discussed earlier, science educators are urged to avoid a Whig-
gish interpretation of history, though it is tempting to speculate on the extent to
which it is possible to write an account of anything from a standpoint other than the
present. Also, one wonders whether an account, if it could be written, would have
any relevance for us unless it is based on the concerns of the present. It might serve
merely to encourage students to worry about theoretical and social problems that
have already been satisfactorily solved or rendered obsolete.
Despite expressing these concerns, I incline to the view that criticizing science
teachers for projecting bad history is to miss the point. The history need only be as
complex and sophisticated as necessary to make particular points in pursuit of learn-
ing goals associated with enhancing conceptual understanding, humanizing science,
illustrating sociocultural context, addressing methodological issues, or whatever.
It is not necessary to tell everything known or conjectured about historical
episodes to use them to humanize science and scientists, to rationalize past
problem choices, to explode the myth of objectivity, to underline the aesthe-
tics and pigheadedness of creativity, and to teach that the validation and
refinement of new science is a social, not an individual, affair. (Heilbron,
1987, p. 558)
The science curriculum is designed for students to learn science, learn about
science and engage in science; its purpose is not to teach history, teach about his-
tory, or give experience of doing history. It is well known that practical work is

168

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
MAKING A CASE FOR HISTORY OF SCIENCE

more successful in bringing about learning when it is designed with specific goals
in mind and that activities successful in bringing about concept acquisition and
development are not necessarily well suited to teaching about scientific inquiry or
developing bench skills (Hodson, 1993c). By extension, history of science should
be deployed differently for different purposes. Differences in purpose, leading to
different questions and different interpretive frameworks, are evident in the different
ways that Karl Popper, Imre Lakatos, Thomas Kuhn, and Gerald Holton describe
the final overthrow of Newtonian physics by Albert Einstein. It is here that Allchin’s
(1995) distinction between rational reconstructions (using modern scientific pers-
pectives) and historical simulations is useful: the former being useful for dealing
with conceptual and procedural issues, the latter being well suited to dealing with
the humanitarian aspects and for addressing social, economic and moral-ethical
issues.17 These differences in presentation reflect Latour’s (1987) distinction bet-
ween “ready made science” and “science in the making” and speak to the distinc-
tion between studies that deal with the context of justification and those dealing
with the context of discovery. Kragh (1992) suggests that teachers should ‘come
clean’ and make students aware of those situations in which history has been dis-
torted to enhance the conceptual or philosophical aspects, and when the science has
been simplified in order to focus on the sociocultural dimensions of an episode in
the history of science. As an example of what she has in mind, she cites the follow-
ing passage, in which Wichmann (1971, p. 19) introduces an account of the devel-
opment of quantum theory.
The reader should realize that our discussion is extremely deficient as a his-
torical account: we could not possibly hope to do justice to the very interest-
ing development of quantum physics in a few pages. We are looking at the
situation at the beginning of this century in retrospect, and it is then easy to
see that these three problems [Planck’s law of blackbody radiation, the stability
of atoms and the photoelectric effect] were key problems. However, if we ex-
amine the publications for the year 1900… we find that the majority of physi-
cists were concerned with very different things. (Kragh, 1992, p. 360)
Moreover, as Justi and Gilbert (1999) remind us, the conceptual models used by
teachers are not always coincident with models deployed in the history of science.
Often they are modified or hybridized in order to meet particular pedagogical pur-
poses.

PERSISTENT PROBLEMS

Although arguments for including the history of science in the school science
curriculum were made as early as 1855 (Jenkins, 1989) and have continued to be
made at regular intervals ever since, there has not been widespread uptake of the
advice, and surprisingly little research into the impact of history of science on student
learning has been published. As mentioned previously, there is strenuous opposi-
tion to history of science on the grounds that it ‘dilutes’ science courses, takes too
long, and deflects attention away from content acquisition. Consequently, there is

169

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
CHAPTER 8

widespread apprehension that science departments in universities will not afford


such courses the same value as conventional science courses, and may not accept
them as entry qualifications. There is concern about who would teach such courses.
Do science teachers know enough history? Do history teachers know enough sci-
ence? Is either group sufficiently confident to take on the responsibility?
Although the discussion in this chapter places the responsibility firmly in the
science teachers’ camp, many teachers already express grave concerns about time
constraints – arguing that with courses already seriously overloaded it would be
difficult to find the time and create the opportunity to adopt an historical perspec-
tive. It is inevitable that some teachers will be concerned about the additional
demands on them, feeling that they lack expertise in the field and do not have
access to good teaching resources. Many teachers lack confidence in dealing with
historical matters, just as they lack confidence in dealing with philosophical and
sociological matters.
It is clear that curricula designed along the lines being argued in this book will
not be successfully implemented unless serious attention is given to teachers’ con-
cerns and steps taken to develop appropriate curriculum materials. Particularly
promising in this respect are the continuing efforts of the Project 2061 team to
develop an integrated collection of computer-based, Internet-accessible collection
of HPS resources – to be known as System 61 (Rutherford, 2001). Above all, there
is urgent need for research and professional development in the area of pedagogical
content knowledge – how to translate history of science into teachable form. The
same urgency clearly attends other aspects of the nature of science, particularly the
philosophy and sociology of science. Looming over all debate about curriculum is
the ever-present spectre of assessment. Given the increasing concern of education
authorities with assessment, evaluation and accountability, and given the wide-
spread belief that students will not regard anything seriously unless it is rigorously
assessed, there is an urgent need for the development of good and robust assess-
ment procedures and instruments that address the historical, philosophical and
sociological dimensions of science and technology.
One final point: there are teachers who believe that encouraging students to
study ideas from the past, most of which are now regarded as false, serves merely
to confuse them; there are teachers who are unconvinced that history will motivate
students, believing that some (perhaps most) students will regard history as even
more tedious to study than science; there are teachers who believe that school age
students are insufficiently mature, both intellectually and emotionally, to deal with
the complexity and subtlety of issues arising from historical studies. Short of Min-
istry of Education directives (or their local equivalent), nothing will convince these
teachers to introduce history of science into their courses (or philosophy of science
and sociology of science, either) except clear research evidence indicating its edu-
cational value. One of the paradoxes facing the curriculum developer is that teach-
ers will not try a radical curriculum alternative unless there is good research
evidence in its favour, but that evidence will not be forthcoming unless someone
implements the programme. While there is some evidence that Harvard Project
Physics attracted a much wider clientele than regular science courses (Brush, 1989)

170

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
MAKING A CASE FOR HISTORY OF SCIENCE

and may have improved attitudes to science (Welch & Walberg, 1972), there is
little or no evidence to suggest that it improved content knowledge (Ahlgren &
Walberg, 1973). Unfortunately, both this course and that devised by Connelly et al.
(1977) to focus on historically significant examples of scientific inquiry have long
since been abandoned. However, some encouragement can be taken from the fol-
lowing research studies, selected for inclusion in this chapter because they show
that the benefits of history of science seem to cross cultural boundaries, even when
the science on which they focus is Western science. First, a study by Galili and
Hazan (2001b), in Israel, shows that a history-oriented course on optics can have a
positive effect on both conceptual and procedural understanding. Second, research
by Lin et al. (2002) shows that the performance of Taiwanese students in solving
conceptual problems on the properties of gases and on atomic theory improved
when historical studies involving debates, role play and hands-on activities were
used. Third, California-based research by Kafai and Gilliland-Swetland (2001)
showed that even primary school age children can reach an appreciation that data
collection methods can and should be different at different times in history and in
different circumstances. Fourth, and particularly striking, Hazari (2006) demon-
strates that the inclusion of history of science in secondary school science courses
(in the United States) impacts significantly on the attainment of women in univer-
sity physics courses. Interestingly, for men, history of science is not a predictor of
success.

ENDNOTES
1
Just how different the intellectual contexts can be is graphically illustrated by Chalmers (1998) in his
comparison of the ways in which Demokritos and Dalton regarded atoms.
2
Of course, attention must also be paid to the requirements of the curriculum.
3
In my view, the curriculum should also address the sociopolitical and economic forces that result in
science and technology sometimes serving the interests of the rich and powerful at the expense of
the poor and powerless (Hodson, 1994, 2003).
4
The history of technology may be even more effective in achieving these goals (Nielsen, 1993).
5
At a more mundane level, history of science may be one of the few ways of generating interest
among those students who have no other interest in science but are required by matriculation regula-
tions to take a science course (Galili & Hazan, 2001b).
6
Even the conventions of scientific reporting and methods of establishing claims to originality have
changed over time and are, themselves, products of history.
7
Jenkins (1989) makes the intriguing argument that introducing history of science into the curriculum
is an effective way of shifting curriculum emphases, and so can be regarded as a tool for curriculum
reform – for example, using historical evidence about the use of experiments in theory building to
change the way teachers design and implement practical work.
8
In my opinion this is a seriously flawed view, as I argue at length in Hodson (1985, 1988, 1993c,
1996). It has led to the absurdities of discovery learning, the excessive emphasis on the so-called
processes of scientific inquiry, a focus on sterile laboratory exercises and neglect of the sociocultural
aspects of science.
9
My own view is that we should orient the science curriculum towards preparing students to take
sociopolitical action. The curriculum proposals outlined in Hodson (1994, 2003) are concerned to
produce activists: people who will fight for what is right, good and just; people who will work to
re-fashion society along more socially-just lines; people who will work vigorously in the best inter-
ests of the biosphere.

171

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
CHAPTER 8

10
Although they have less impact than the television versions, both these series are available in book
form: Bronowski (1977) and Burke (1978).
11
This is certainly the case in Ontario and in New Zealand - the educational systems with which I am
most familiar. Teachers now find it virtually impossible to meet the demands of the plethora of
learning outcomes specified for each individual subject through inter-disciplinary projects.
12
Brockman (1995) argues that the great intellectual achievements of science constitute a ‘third cul-
ture’, of equal value to literacy and numeracy.
13
The term “Whig history” was coined by Herbert Butterfeld (1931). It may be defined as an attempt
to explain the past in terms of the present, to investigate the past in order to support conclusions in
the present and to fit the past into an explanatory scheme applicable to the present, as distinct from
the attempt to explain the past on its own merits. It is not too much of an exaggeration to say that
Whiggish history sees the beliefs, practices and institutions of the present as the goals of previous
efforts. Kragh (1987) uses the term anachronic to describe this approach to history and the term
diachronic for attempts to interpret historical events in the light of views prevalent at the time:
“In the diachronical perspective one imagines oneself to be an observer in the past, not just of the
past” (p. 90).
14
As Foucault (1970) remarks: “Historians want to write histories of biology in the eighteenth century;
but they do not realize that biology did not exist then, and that the pattern of knowledge that has
been familiar to us for a hundred and fifty years, is not valid for a previous period” (p. 15).
15
The term noise has been used to describe the features of practical work that frequently distract stu-
dents from attending to the key conceptual and procedural aspects (Hodson, 1993c, 1996). Similarly,
mathematical noise refers to unnecessarily complex arithmetical and graphing tasks that may divert
students’ attention away from the underlying science. Historical noise distracts students from the
science embedded in historical case studies.
16
In a more recent paper, Brush (2002) says that he now regards some scientists, notably James Clerk
Maxwell, Max Planck and Edwin Hubble, as good role models.
17
Historical simulations may still modify the history slightly, but solely in order to emphasize rather
than eliminate the sociocultural and intellectual climate of the time.

172

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
CHAPTER 9

LOOKING FOR BALANCE IN THE CURRICULUM


Essential Elements in a Curriculum for Critical Scientific Literacy

While there are still some contentious issues concerning precisely what scientific
literacy entails and why we need it (see discussion in chapter 1), there is a measure
of consensus on some of its basic components.1
− A general understanding of some of the fundamental ideas, principles and theories
of science, and the ability to use them appropriately and effectively.
− Some knowledge of the ways in which scientific knowledge is generated, vali-
dated and disseminated.
− Familiarity with the form, structure and purpose of scientific language.
− The capacity to read and interpret scientific data and, at a general level, to
evaluate their validity and reliability.
− The ability to evaluate a scientific argument or claim to knowledge.
− Some awareness of the sociocultural and cognitive circumstances surrounding
the history and development of some of the ‘big ideas’ of science, and the origin
and development of important technologies.
− An appreciation of the complexity of inter-relationships among science, tech-
nology, society and environment
− A commitment to critical understanding of contemporary socioscientific issues at
the local, regional, national and global levels, including their historical roots and
underlying values, together with a willingness to take appropriate and responsible
action, and encourage others to do so.
− The capacity and willingness to address moral-ethical issues associated with
scientific research and the deployment of scientific knowledge and techno-
logical innovations.
− General interest in science, together with a willingness and capacity to update
and acquire new scientific and technological knowledge in the future.
It is immediately apparent that issues in the history, philosophy and sociology
of science (hereafter, HPS) impact on all ten categories. This is true even of the
first item if one takes the view that understanding and using an item of scientific
knowledge entails knowledge of its role and status as well as its strengths, weak-
nesses and relationships to other knowledge items.
My intention in the earlier chapters of this book was to cast doubt on the image
of science and scientists that school science curricula and school science textbooks
have traditionally promoted. Indeed, elsewhere (Hodson, 1998b) I have identified
ten common myths and falsehoods promoted, sometimes explicitly and sometimes
implicitly, by the science curriculum.2 They are reproduced in Table 9.1 alongside
a broadly similar list of falsehoods generated by McComas (1998) from his critical
reading of science textbooks.

173

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
CHAPTER 9

Table 9.1. Myths and falsehoods about science

Hodson (1998) McComas (1998)


Observation provides direct and Hypotheses become theories that in turn
reliable access to secure become laws.
knowledge. Scientific laws and other such ideas are
Science always starts with absolute.
observation. A hypothesis is an educated guess.
Science always proceeds via A general and universal scientific method
induction. exists.
Science comprises discrete, Evidence accumulated carefully will result
generic processes. in sure knowledge.
Experiments are decisive. Science and its methods provide absolute
Scientific inquiry is a simple proof.
algorithmic procedure. Science is procedural more than creative.
Science is a value-free activity. Science and its methods can answer all
Science is an exclusively questions.
Western, post-Renaissance Scientists are particularly objective.
activity. Experiments are the principal route to
The so-called ‘scientific scientific knowledge.
attitudes’ are essential to the Scientific conclusions are reviewed for
effective practice of science. accuracy.
All scientists possess these Acceptance of new scientific knowledge is
attitudes. straightforward.
Science models represent reality.
Science and technology are identical.
Science is a solitary pursuit.

Sweeping away the old is one thing; finding an acceptable set of alternatives is
somewhat different. Many science educators will share Israel Scheffler’s alarm at
some of the alternatives that have been advanced.
The extreme alternative that threatens is the view that theory is not controlled
by data, but that data are manufactured by theory; that rival hypotheses
cannot be rationally evaluated, there being no neutral court of observational
appeal nor any shared stock of meanings; that scientific change is a product
not of evidential appraisal and logical judgment, but of intuition, persuasion
and conversion; that reality does not constrain the thought of the scientist but
is rather itself a projection of that thought. (Scheffler, 1967, p. v)
Longbottom and Butler (1999) express similar concerns when they state that “if
we go along with those who deny that modern science provides a privileged view
of the world… we fall into an abyss where skeptical postmodernists, who have lost
faith in reason, dismiss all knowledge claims as equally arbitrary and assume the
universe to be unreliable in its behavior and incapable of being understood” (p. 482).

174

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
LOOKING FOR BALANCE IN THE CURRICULUM

Stanley and Brickhouse (1995) regard such remarks as examples of what Bernstein
(1983) called “Cartesian anxiety”: the fear that if we do not retain our belief in the
traditional objective foundations of scientific method we have no rational basis for
making any knowledge claims. In short, fear that belief in scientific progress will
be replaced by scientific change consequent upon power struggles among competing
groups, with ‘victory’ always going to the better resourced. Fear that scientific
knowledge is no longer to be regarded as the product of a rigorous method or set of
methods; instead, it is merely the way a particular influential group of scientists
happens to think, and can persuade, cajole or coerce others into accepting.
The purpose of this chapter is to identify aspects of HPS that are suitable for
inclusion in the school science curriculum. In making our selection, do we face
a simple but stark choice between the traditional and the post-modern? Are we
required to choose between the image of the scientist as a cool, detached seeker-
after-truth patiently collecting data from which conclusions will eventually be
drawn, when all the evidence is in hand, and that of “an agile opportunist who will
switch research tactics, and perhaps even her entire agenda, as the situation requires”
(Fuller, 1992, p. 401)? Which view is the more authentic? Equally important, what
should we tell students? What is in their interests? Some years ago, Stephen Brush
(1974) asked: “should the history of science should be rated X?” The question is
just as pertinent to the philosophy of science and the sociology of science. Should
we expose students to the anarchistic epistemology of Paul Feyerabend? Should we
lift the lid off the Pandora’s box that is the sociology of science? Would students
be harmed by too early an exposure to these views? When we seek to question (and
possibly reject) the certainties of the traditional view of science, are we left with
no firm guidance, no standards and no shared meaning? Does recognition of the
sociocultural baggage of science entail regarding science as just one cultural arti-
fact among many others, with no particular claim on our allegiance? Is any kind
of compromise possible between these extremes and among this diversity? Can we
retain what is still ‘good’ and ‘useful’ about the old view of science (such as con-
ceptual clarity and stringent testing) while embracing what is ‘good’ and ‘useful’ in
the new (such as sensitivity to sociocultural dynamics and awareness of the possi-
bility of error, bias, fraud and the misuse of science)? Can the curriculum achieve
a balance that is acceptable to most stakeholders? In short, what particular items
from all the argument and counter argument reviewed in earlier chapters would
constitute an educationally appropriate and teachable selection? What nature of
science (NOS) understanding can be taken-for-granted and regarded as no longer in
dispute? Is there any consensus among scholars about an acceptable alternative to
the traditional view that will allay the fears expressed by Scheffler and others?
Responses to a 20-item Likert-type questionnaire on “15 tenets of NOS” led
Alters (1997a) to conclude that there is no consensus – at least, not among the 210
philosophers of science he surveyed.3 In the words of Larry Laudan et al. (1986):
The fact of the matter is that we have no well-confirmed general picture of
how science works, no theory of science worthy of general assent. We did once
have a well developed and historically influential philosophical position, that of

175

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
CHAPTER 9

positivism or logical empiricism, which has by now been effectively refuted.


We have a number of recent theories of science which, while stimulating
much interest, have hardly been tested at all. And we have specific hypotheses
about various cognitive aspects of science, which are widely discussed but
wholly undecided. If any extant position does provide a viable understanding
of how science operates, we are far from being able to identify which it is.
(p. 142)
Interestingly, despite this categorical denial of consensus, it seems that the authors
of several recent science curriculum reform documents (AAAS (1989, 1993) and
NRC (1996), among others), with their increasing emphasis on NOS issues, are
in fairly substantial agreement on the elements of NOS that should be included
in the school science curriculum (McComas & Olson, 1998) – a view endorsed
by Cleminson (1990), Abd-el-Khalick et al. (1998), McComas et al. (1998) and
Cobern and Loving (2001).4 Indeed, Abd-El-Khalick and BouJaoude (1997), Abd-
El-Khalick et al. (1998) and Lederman et al. (2002) state that while philosophers
and sociologists might disagree on some aspects of NOS, these disagreements are
irrelevant to K-12 students and their teachers. In an effort to shed further light on
this matter, Osborne et al. (2003) conducted a Delphi study to ascertain the extent
of agreement among 23 participants drawn from the ‘expert community’ on what
ideas-about-science should be taught in school science. The participants included
five scientists, five persons categorized as histo-rians, philosophers and/or
sociologists of science, five science educators, four science teachers and four
science communicators. Although there was some variation among individuals,
there was broad agreement on nine major themes: scientific method and critical
testing; scientific creativity; historical development of scientific knowledge;
science and questioning; diversity of scientific thinking; analysis and interpretation
of data; science and certainty; hypothesis and prediction; cooperation and collabo-
ration. A comparison of these themes with those distilled from the science
education standards documents in McComas and Olson’s (1998) study reveals
many simila-rities (Table 9.2).
Essentially the same list is found in Lederman et al (2002): science is tentative,
empirically based, subjective (in the sense of being theory dependent and impacted
by the scientists’ experiences and values), socioculturally embedded and, in part,
the product of human imagination and creativity. These authors are also concerned
that students draw a distinction between observation and inference (see my com-
ments in Chapter 3) and understand the functions of and relationships between
theories and laws.5 A number of questions immediately spring to mind. First, is this
apparent consensus deliberately pitched at such a trivial level that nobody could
possibly quibble with it? For example, statements such as “science is an attempt to
explain natural phenomena’, “scientists are creative” and “science and technology
impact each other” – all items in the consensus list developed by McComas et al.
(1998) – don’t claim anything particularly insightful. Of course, some would argue
that a list of relatively trivial items is better than no list at all. Second, we should
ask whether the list includes the ‘big issues’ with which philosophers of science

176

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
LOOKING FOR BALANCE IN THE CURRICULUM

Table 9.2. Comparison of themes in Delphi study and McComas & Olson study
(Osborne et al., 2003)

have traditionally grappled? And third, we should ask whether it is a consensus


arrived at in retrospective justification of a previously agreed set of wider curri-
culum proposals? We should also ask about the people involved in reaching con-
sensus on NOS items. Given their diverse concerns, we might expect substantial
disagreements among philosophers of science, sociologists of science, scientists,
historians of science and science educators (Pomeroy, 1993; Abd-El-Khalick et al.,
1998).6 So how is consensus achieved within such a diverse group and can that
information be helpful in achieving the curriculum balance being sought in this
chapter? When disagreements arise, whose views of NOS are to count? Should we
defer to the philosophers of science, on the grounds that they spend the whole of
their professional lives grappling with the important questions? “No” say Smith et al.
(1997) because their concerns are “esoteric, inaccessible, and probably inappro-
priate for most K-12 instruction” (p. 1102). Should we put our trust in scientists, on
the grounds that those engaged on a day-to-day basis with the enterprise of science
must know what it entails? Some readers will recall that Peter Medawar (1969)
once famously remarked that if you ask a scientist about scientific method “he will
adopt an expression that is at once solemn and shifty-eyed: solemn, because he
feels he ought to declare an opinion; shifty-eyed, because he is wondering how
to conceal the fact that he has no opinion to declare” (p. 11). Pitt (1990) has also
observed that most scientists have poor understanding of NOS issues and “generally
don’t even know the history of their own discipline” (p. 16). What is interesting
is that it clearly doesn’t matter. While NOS knowledge is crucial to the deep and

177

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
CHAPTER 9

reflective understanding of science implied in the term critical scientific literacy,


it seems to be irrelevant to the practice of science.
Science, broadly considered, is incomparably the most successful enter-
prise human beings have ever engaged upon; yet the methodology that has
presumably made it so, when propounded by learned laymen, is not attended
to by scientists, and when propounded by scientists is a misrepresentation
of what they do. Only a minority of scientists have received instruction in
scientific methodology, and those that have done so seem no better off.
(Medawar, 1982, p. 80)
In Weinberg’s (1987) words, “philosophy of science is just about as useful to
scientists as ornithology is to birds” (p. 433). Interestingly, many of the scientists
interviewed by Schwartz and Lederman (2006) and Wong and Hodson (2008a,b) did
have fairly sophisticated views about NOS, though they don’t realize it until they
discuss issues in depth. Inevitably, their day-to-day concerns as researchers and theory-
builders leave little time for reflection on matter philosophical. Nonetheless, some
of the scientists in these studies also expressed very narrow and naïve views of
NOS, which reminds us that NOS-oriented teaching in school and university is just
as important in courses aimed at the pre-professional education of future scientists
as it is in more general courses. Scientists are citizens, too! They also need the kind
of critical awareness that will equip them to address complex socioscientific issues.
In returning to the question of whose views of NOS are to count, should we
follow Albert Einstein’s (1933) advice and put our faith in sociologists and ethno-
graphers?
If you want to find out anything from the theoretical physicists about the
methodology they use, I advise you to stick closely to one principle: don’t
listen to their words, fix your attention on their deeds. (p. 270)
Or should we rely on historians of science, on the grounds that a robust theory
of science must stand up to close historical scrutiny? After all, what value is there
in an account of science for which there is little or no historical support? Of course,
historians of the philosophy of science (Losee (1993) and Heidelberger and Stadler
(2002), for example) tell us that NOS knowledge has changed as science itself has
changed, and so we can’t simply look to the past for guidance on the present con-
cerns of NOS knowledge.
No more than a casual browse in the library is necessary to show that philo-
sophy of science is a dynamic, changing and increasingly specialized field of study.
Within the discipline, philosophers of quantum mechanics may have few concerns
in common with philosophers of biology. Some philosophers of science hold that
there is no universal nature of science because the sciences themselves have no
unity. The best that can be said is that there is a ‘family resemblance’, with common
interests and some areas of methodological and conceptual agreement7 – what
Loving (1997) calls a “loose configuration of critical processes and conceptual
frameworks, including various methods, aims, and theories all designed to shed
light on nature” (p. 437). Thus, Mayr (1988, 1997) has criticized NOS views on

178

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
LOOKING FOR BALANCE IN THE CURRICULUM

the grounds that they are nearly always derived from physics. Biology, he argues,
is markedly different in many respects, not the least significant being that many
biological ideas are not subject to the kind of falsificationist scrutiny advocated by
Popper and given such prominence in school science textbooks: “It is particularly
ill-suited for the testing of probabilistic theories, which include most theories in
biology… And in fields such as evolutionary biology… it is often very difficult, if
not impossible, to decisively falsify an individual theory” (Mayr, 1997, p. 49). It is
more than 70 years since Bachelard (1934/1985) expressed broadly similar views
about diversity among the sciences.
When one looks at science what is immediately striking is that its oft-alleged
unity has never been a stable condition, so that it is quite dangerous to
assume a unitary epistemology. Not only does the history of science reveal a
regular alternation between atomism and energetics, realism and positivism,
continuity and discontinuity, rationalism and empiricism; and not only is the
psychology of the scientist engaged in active research dominated one day by
the unity of scientific laws and the next by the diversity of things; but even
more, science is divided, in actuality as well as in principle, in all of its
aspects. (p. 14)
As a particular science progresses, and new theories and procedures are deve-
loped, the nature of scientific reasoning changes. Thus, any consideration of NOS
must be “contextual, conditional, with an eye to open horizons: ‘closed’ answers
must, for that reason alone, be suspect, indeed rejected” (Suchting, 1995, p. 20). In
short, we should seriously question whether views in the philosophy of science that
were arrived at some years ago can any longer reflect the nature of 21st Century
science, especially in rapidly developing fields in the biological sciences.
Similar arguments were developed in chapter 5 to address and subsequently
reject the fundamental assertion of the so-called “Process Approach” to science
education: that science comprises a set of discrete, generic processes, and that
these processes can be acquired in any context and subsequently used with compe-
tence in any other context. I argued that scientific inquiry is perhaps better under-
stood as situated practice, in which the particular conceptual schemes deployed
and the specific procedures and instrumental techniques utilized play a crucial role.
It is not too much of an exaggeration to state that the different sub-disciplines of
science employ distinctive styles of argumentation, rules of evidence and criteria of
judgement. In other words, the specifics of scientific rationality change between
sub-disciplines, with each sub-discipline playing the game of science according
to its own rules. It is not surprising, therefore, that students often experience great
difficulty in generalizing even the simplest process skills from one context to
another – an observation that has been given theoretical legitimacy by the situated
cognition movement (Lave, 1988; Brown et al., 1989; Hennessy, 1993; McLellan,
1996). This argument has been eloquently and forcefully made by Rudolph (2000).
Understanding of the operations of science has changed dramatically over the
past 40 years. Educators need to begin to exploit the vast literature of the
science studies community, not to develop some universalist picture of science,

179

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
CHAPTER 9

the value of which is questionable, but to begin to understand what the various
practices of science look like in all their myriad forms, in order to provide some
reasonably authentic context in which to situate the scientific knowledge claims
of the curriculum. (p. 409)
Elby and Hammer (2001) also argue that the current consensual list of NOS items
is too general and broad, and that it is neither valid nor productive of good learning
of science: “a sophisticated epistemology does not consist of blanket generalizations
that apply to all knowledge in all disciplines and contexts; it incorporates contextual
dependencies and judgments” (p. 565). Essentially the same point is made by Clough
(2006) when he says that “while some characteristics [of NOS] are, to an acceptable
degree, uncontroversial… most are contextual, with important and complex excep-
tions” (p. 463). Thus, instead of trying to find and promote broad generalizations
about the nature of science, scientific inquiry and scientific knowledge, teachers
should be building an understanding of NOS from examples of the daily practice of
diverse groups of scientists engaged in diverse practices, and should be creating
opportunities for students to experience, explore and discuss the differences in
knowledge and its generation across multiple contexts. It is for this reason that
NOS-oriented research needs to study the work of scientists active at the frontier of
knowledge generation..
Lederman (1995) questions whether school age students could cope with diverse
and sometimes conflicting views of how science is conducted, citing Perry’s (1970)
conclusion that even adults often experience difficulty with relativistic thinking.
If we were to attempt to promote student understanding of the multiplicity of
conceptions of science an enormous dilemma would quickly arise. We would
be expecting students to internalize a particular world view or paradigm so
that they can understand subject matter derived from that particular paradigm.
We would then be asking these same individuals to internalize yet another
paradigm so that they can understand the subject matter perspectives derived
from the alternative paradigm. (p. 374)
I am not so pessimistic as Lederman. In my experience, when presented with
appropriate examples, students have no difficulty in recognizing that different
situations require different strategies, depending on the nature of the problems, the
expertise of the research group, and the resources available. Indeed, students find
this perfectly natural; it is teachers who create the expectation of one fixed approach
to scientific inquiry through their constant references to “the scientific method”
(Hodson, 1990). I do not accept the extreme interpretation of Kuhn’s incom-
mensurability thesis on which Lederman’s anxieties are based (see below); nor do I
equate recognition of a plurality of approaches to scientific inquiry with relativism.
While scientists are free to choose whatever approach they consider best suited to
the immediate situation, their choices will be subject to close critical scrutiny by
the community of practitioners. Among other considerations, appraisal will focus
on the methods employed: Were they appropriate? Were they satisfactorily per-
formed? Could/should the investigation have been conducted differently? In short,
some scientific investigations are better than others, and other scientists will make

180

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
LOOKING FOR BALANCE IN THE CURRICULUM

judgements to that effect, judgements that are not arbitrary or subject to personal
whim.
While I am arguing for a much more fluid and context-dependent view of what
constitutes scientific practice, I also recognize that telling students, too early in their
science education, that scientific inquiry is context-dependent and idiosyncratic could
be puzzling, frustrating and even off-putting. This is a similar point to Brush’s (1974)
concern that teaching history of science can have an adverse effect on students by
undermining their confidence in science and scientists. Claxton (1997) expresses
similar anxiety: “Perhaps the ‘truth’ about science would be too complex, or too
destabilizing for most young people to handle. After all, we protect them, as a
matter of principle, from many aspects of the adult world for which we judge them
to be unready” (p. 76). One approach is to take our cue from secondary school
chemistry curricula, where we often begin with some very simple representations,
such as “elements are either metals or non-metals” or “bonding is either covalent or
electrovalent”. We then proceed to qualify these assertions in all manner of ways:
“there are varying degrees of metallic/non-metallic character, depending on atomic
size and electron configuration”; “there is a range of intermediate bond types,
including polarized covalent bonds and lattices involving highly distorted ions,
as well as hydrogen bonding, van der Waal’s forces, and so on”. Similarly, in the
early years, we may find it useful to characterize scientific inquiry as a fairly standard
set of steps. Within this simple representation we can emphasize the importance of
making careful observations (using whatever conceptual frameworks are available
and appropriate to the students’ current stage of understanding), taking accurate
measurements, systematically controlling variables, and so on. As students become
more experienced they can be introduced to variations in approach that are necessary
as contexts change – for example, the startlingly different approaches adopted
by experimental particle physicists, synthetic organic chemists and evolutionary
biologists. Eventually, students can experience for themselves the joys (and frust-
rations) of having to work out the procedures for themselves. With these experiences
comes the realization that doing science successfully involves learning to ‘think on
one’s feet’. In a sense, this progression is similar to those cases in science where
scientists begin with a conceptual model (which they know is not ‘true’) and proceed
through debate and experimentation to refine and develop it into a more complex
theory (see chapter 6). The model is merely a devise to help them think more clearly
and to gain a measure of predictive control, while the theory is believed to explain
the reality they are studying. By analogy, the early childhood version of science
can be regarded as a model of science and the more robust and sophisticated later
version as a theory of science and scientific practice.

WHAT DO STUDENTS REALLY NEED TO KNOW?

While I wholeheartedly endorse Rudolph’s (2000) view that we need to address the
diversity of scientific practice, we still have to confront the question of what specific
ideas about science to include in the school science curriculum. Science teachers
and science curriculum developers are still faced with the task of making a selection

181

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
CHAPTER 9

capable of capturing (at least some of) the complexity of science. Answering the
‘what NOS items do we include?’ question presupposes a satisfactory answer to the
‘why learn about science?’ question posed in chapter 1. Like all aspects of curri-
culum decision-making, decisions about inclusion of particular NOS items will be
informed, explicitly or implicitly, by the wider sociopolitical context in which the
decisions are located. In other words, any representation of science and scientists
will carry with it a powerful sub-text of sociopolitical meaning and will reveal the
interests and biases of the decision makers.
Rudolph (2002) argues that previous representations of science and its episte-
mology in US science education during the second half of the 20th Century, deriving
from the work of Dewey and Schwab, were formulated to advance particular social
and political interests. Both men emphasized process over content, but Dewey vested
authority in the individual and the capacity of the individual to use a universally
applicable scientific method in the solution of wider social problems, while Schwab
emphasized the specialized expertise of the community of scientists and the capacity
of scientists to use complex, context-specific methods guided by prior concep-
tual frameworks in ways that are not readily accessible to the lay public – thus
engendering respect for scientists and support for the scientific enterprise. Whether
or not one agrees with Rudolph’s analysis, the simple point remains that there are
no socially neutral images of science and scientists. Each and every representation
will carry implicit or explicit messages. It is incumbent on us, therefore, to make
our wider intentions clear. We need to be clear about the wider goals of science
education, and why we choose to include NOS issues in the curriculum, and we
also need to give some thought to the consequence of how the implicit goals of
science education are likely to be perceived by students. My own motives for
advocating NOS teaching were discussed at length in chapters 1 and 2 (see also
Hodson, 2006).
It is unrealistic as well as inappropriate to expect students to become highly
skilled philosophers and sociologists of science, just as it is fruitless to pursue
sophisticated expertise in history of science. The science teacher’s responsibilities
encompass education in science, education about science and opportunities for
students to do science. In chapter 8, I argued that we should subordinate history of
science to the major goals of science education. Similarly, we should select NOS
items for the curriculum in relation to other educational goals: the need to motivate
students and assist them in developing positive attitudes towards science; the need
to pay close attention to the cognitive goals and emotional demand of specific
learning contexts; the creation of opportunities for students to experience doing
science for themselves; the capacity to address complex socioscientific issues with
critical understanding; concern for multiculturalism and antiracism; and so on. The
degree of sophistication of the NOS items we include should be appropriate to the
stage of cognitive and emotional development of the students and compatible with
other long and short-term educational goals. There are numerous goals for science
education, and education in general, that can, will and should impact on decisions
about the NOS content of lessons. Our concern is not just good philosophy of
science, good sociology of science or good history of science, not just authenticity

182

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
LOOKING FOR BALANCE IN THE CURRICULUM

and preparation for sociopolitical action (my ultimate goal for science education),
but the educational needs and interests of the students – all students. Selection of
NOS items should consider the changing needs and interests of students at different
stages of their science education, as well as taking cognizance of the views of
‘experts’ and promoting the wider goals of (i) authentic representation of science
and (ii) politicization of students. In short, we need to ask what is in students’ best
interests at any given time – a decision that is best made by the teacher, of course.
It is considerations like these that prompted Matthews (1998b) to advocate
the pursuit of “modest goals” concerning HPS in the school science curriculum.
In his words, “there is no need to overwhelm students with cutting edge questions”
(p. 169). I agree. But agreement only serves to raise the question of “What should
these modest goals comprise?” At the very least, we should include the following:
consideration of the relationship between observation and theory; the role and
status of scientific explanations (including the processes of theory building and
modelling); the nature of scientific inquiry (including experiments, correlational
studies, blind and double blind trials, etc); the history and development of major
ideas in science; the sociocultural embeddedness of science and the interactions
among science, technology, society and environment; the distinctive language of
science; the ways in which scientific knowledge is validated through criticism,
argument and peer review; moral-ethical issues surrounding science and technology;
error, bias, vested interest, fraud and the misuse of science for sociopolitical ends;
the relationship between western science and indigenous knowledge. A number
of these elements are present in some science curricula, but more often than not
they are implicit, part of the “hidden curriculum”, embedded in language, text-
book examples, laboratory activities and the like, and so dependent, ultimately, on
teachers’ nature of science views – hence the need for teacher education to pay
close attention to the nature of science.
More controversially, Jenkins (1994c) argues that it is incumbent on science
teachers to introduce students to some of the more radical views concerning nature
of science: the ideas of Paul Feyerabend, for example, the autobiographical accounts
provided by scientists such as James Watson and Richard Feynman, and the socio-
logical writing of Bruno Latour and Steve Woolgar. While I am strongly sym-
pathetic to Jenkins’ argument, I also recognize that we need to be circumspect,
modest in our goals and wary of promoting extreme views. In chapter 5, for example,
I warned that too literal an interpretation of Feyerabend’s (1975) dictum that
“anything goes” would constitute a gross disservice to students. While his free-
wheeling approach certainly applies to the creative phase of scientific inquiry
(what some have called the context of discovery) it is less applicable to the context
of justification, for which there are strict procedures relating to judgements about
reliability, validity and appropriateness. This more circumspect approach might
also incline us to consider that Feyerabend’s anarchistic view is fine (maybe even
necessary) for what Kuhn calls revolutionary science, but is entirely inappropriate
for the effective conduct of normal science, where a more dispassionate and sys-
tematic approach is essential. Careful attention to detail and painstaking accu-
mulation of data are crucial to many aspects of good science, to the effective conduct

183

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
CHAPTER 9

of laboratory testing, to the maintenance of safety and to the sound operation of


science-based industries. But theoretical breakthroughs are not made this way.
Science needs both technicians and creative theory-builders, and we need to
provide students with knowledge, understanding and experiences that acknowledge
this dual need and lay the groundwork for all manner of future careers. Kuhn
himself wonders whether “flexibility and open-mindedness have not been too
exclusively emphasized as the characteristic requisites for basic research” (Kuhn,
1963, p. 342). He continues, “almost none of the research undertaken by even the
greatest scientists is designed to be revolutionary, and very little of it has such
effect” (p. 343).
Interestingly, one of the scientists interviewed by Wong and Hodson (2008a) reported
that funding bodies often have clear expectations about the conduct of scientific
investigations and are suspicious of open-ended approaches, which they regard as
little more than “fishing expeditions”. Hence, he argued, it is sometimes necessary
to pretend to be engaged in more systematic inquiries. Further support for adopting
a more fluid view of scientific investigation is provided by scientists in fields such
as molecular biology and materials science, where data collection that previously
took months to complete is now achieved by high-speed computers in a matter of
minutes, rendering prior hypothesizing and theorizing unnecessary (or certainly
less important than in the past). Hence scientific investigations in which data are
obtained first and then interesting problems are identified by “data mining” have
become much more common in recent years. Despite the theoretical and procedural
difficulties associated with inductive approaches outlined in chapter 3, such methods
do have a long history; they have been successfully used in science and have led to
significant advances in many fields (Kipnis, 1996). It behoves us to tell students
this ‘truth’ about the history of science.
We should also ‘tell the truth’ about anomalous data. It is certainly not the case
that scientists immediately abandon a theory when conflicting data arise. As Chinn
and Brewer (1993, 1998) report, scientists’ responses may include any of the
following, depending on the particular circumstances: (i) ignore the data, (ii) reject
the data, often claiming methodological error, (iii) express uncertainty about the
validity of the data, (iv) exclude anomalous data because it is outside the sphere of
concern of the theory, (v) hold data in abeyance for later attention, (vi) reinterpret
the data, (vii) change some peripheral aspect of the theory, (viii) change the theory.
From a study of undergraduates in science labs, Lin (2007) has added a ninth
category: (ix) express uncertainty about interpretation of data. Thus, in moves
(i) and (ii) the data are declared invalid, in moves (iv) to (viii) the data are accepted
as valid, and in moves (iii) and (ix) the scientist is undecided on the validity of
the data.
Similarly, too literal an interpretation of statements about the tentative nature of
science can be counter-productive, leading students to regard all science as no
more than temporary (Harding & Hare, 2000).8 Scientific knowledge is tentative
because it is based, ultimately, on empirical evidence that may be incomplete and
because it is collected and interpreted in terms of current theory - theory that may
eventually be changed as a consequence of the very evidence that is collected. In

184

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
LOOKING FOR BALANCE IN THE CURRICULUM

all these endeavours the creative imagination of individual scientists is impacted by


all manner of personal experiences and values. Moreover, the ‘collective wisdom’
of the scientific community that supports the practice, scrutinizes the procedures
and evaluates the products, is also subject to complex sociopolitical, economic and
moral-ethical forces. In consequence, there can be no certainty about the knowl-
edge produced. However, to admit that absolute truth is an impossible goal is not to
admit that we are uncertain about everything. We know many things about the
universe even though we recognize that many of our theoretical systems are still
subject to revision, or even rejection. There are several closely related issues to
consider. First, very specific claims about phenomena and events may be regarded
as true even though the theories that account for the events are regarded as tentative.
Because the whole necessarily extends beyond the parts of which it is comprised,
the whole may be seen as tentative while the parts (or some of them) are regarded
as certain. Most theories are tentative when first developed, but are accepted as
‘true’ (in a scientific sense) when they have been elaborated, refined and successfully
used, when they are consistent with other theories and strongly supported by
evidence. Teachers make a grave mistake when they encourage students to regard
all science as tentative. Indeed, if scientists did not accept some knowledge as well
established we would be unable to make progress. Cromer (1993) describes the
situation particularly well:
The experimental and theoretical basis of some of our fundamental knowl-
edge is so extensive that there is little likelihood of its being changed to any
significant degree. This is an astonishing assertion, given the breathtaking
pace of discovery today. But the pace of discovery is possible precisely
because our fundamental knowledge is so complete. (p. 6)
Nevertheless, there is always the possibility that some new theory will challenge
and possibly displace these taken-for-granted ideas. Open-mindedness is the key
virtue of the scientist. Hence the seeming paradox that although scientists have
good reason for accepting particular theories as true they do so with the proviso
that they can subsequently change their minds if new evidence or new ways of
looking at existing evidence (brought about by a new theoretical perspective)
become available.
When compelling evidence has accumulated, scientists will come to regard
the theory as settled and established. The claim in question, however, conti-
nues to rest on available evidence, and one’s open-mindedness consists not in
suspending judgment where there would seem to be compelling evidence, but
in being willing to reconsider present conclusions in the light of whatever
evidence or rival interpretation appears. (Harding & Hare, 2000, p. 230)
An important element of science education concerns understanding and being
able to make effective and appropriate use of the ideas that the community of
scientists takes as well-established (what I have called learning science), provided
of course that the warrant for such belief is well-argued. There is little value in
encouraging students to doubt every scientific proposition they encounter (Norris,

185

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
CHAPTER 9

1995). As Elby and Hammer (2001) say, “it would be naïve to think that quantum
theory is the long-sought Theory of Everything. At the same time, it would be
unsophisticated to consider it tentative that the earth is round or that the heart
circulates blood through the body” (p. 557). As Ravetz (1997) has written: “Science,
as it is taught, even up to degree level, admits of uncertainty, doubt and debate only
rarely if at all. It is facts all the way… In some ways this assumption of certainty is
justified. So much of science is unproblematic, that the uncontested facts can more
than fill any course; and young minds do not need to be perturbed by uncertainties
while they are still mastering difficult concepts and techniques” (p. 7). Thus, the
statement “scientific knowledge is tentative.. it has only temporary status” (see
Cleminson, 1990, p. 437) is grossly misleading. Much of the scientific knowledge
that students encounter in class is no longer tentative. Rather, it is well established,
taken-for-granted and used in the production of further knowledge. It is category-
cally not the case that it could be invalidated by a simple experiment in a science
class, and we do students a gross disservice by suggesting that it could. This is not
to say that students shouldn’t be given the opportunity to ‘disprove’ hypotheses or
to challenge some existing ideas. What is sought is a judicious balance: not a
blanket belief in the certainty of scientific knowledge or its uncertainty, but appro-
priate levels of confidence, skepticism and open-mindedness, depending on the
particular knowledge item and the context of its consideration. Indeed, there are
good grounds for believing that students who appreciate that, from time to time,
scientific knowledge is discarded in favour of better alternatives are able to use
their own scientific knowledge and understanding in more flexible and creative
ways than those who are taught to regard all scientific knowledge as absolute truth.
According to Langer et al. (1989), they seem also to be more emotionally and
intellectually resilient, simply because they are better prepared for negative or
unexpected outcomes.
Perhaps we should also be a little circumspect in the way that we present the
empirical underdetermination of scientific theory. It is certainly the case that
observational evidence alone is insufficient to guarantee the truth of a theory, as
discussed previously. It is also the case that, in principle, there can be more than
one plausible explanation for any given set of data. In practice, however, it is only
rarely that this situation arises, and in many situations scientists may struggle to
find even one satisfactory explanation for the data. Nor should it be assumed that
when alternative explanations of the same set of data are generated they are all
equally likely, equally acceptable, equally believable. Underdetermination has, in
general, not been a problem for science. Scientists are not in perpetual disagree-
ment. As discussed earlier, for all kinds of reasons scientists do reach consensus,
and often surprisingly quickly.
Perhaps the aspects of HPS where the greatest degree of caution is required are
those relating to Thomas Kuhn’s incommensurability thesis and to questions sur-
rounding the extent to which we consider scientific knowledge, and the practices
that produce it, to be social constructs.

186

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
LOOKING FOR BALANCE IN THE CURRICULUM

INCOMMENSURABILITY, REVOLUTIONS, NORMAL SCIENCE


AND THEORY CHOICE

Kuhn (1970) argues that because of the theory-dependence of observations


rival theories about a phenomenon give rise to observation statements that, in
the extreme, are mutually unintelligible. Therefore, he claims, such theories are
“incommensurable” and there is no rational method for choosing between them in
a way that is acceptable or meaningful to the advocates of the competing theories.9
Not only is this situation logically possible, says Kuhn, it has been evident on
more than one occasion in the history of science. He cites the transition from
Ptolemaic to Copernican astronomy and the shift from Newtonian mechanics to
special relativity. In consequence, Kuhn argues, judgements between theories are
based on a clutch of non-empirical and non-objective criteria. There are at least
two issues to consider. First, is incommensurability as extensive and fundamental
as Kuhn argues? Second, does the empirical under-determination of theories
necessarily mean that theory change is reduced to a purely social process? Are
there any rational criteria to judge the value of competing theories?
According to Kuhn’s account of scientific revolutions the key concepts of an
existing theory are discarded in favour of entirely new ones. For example, in the
late 18th Century the notion of phlogiston as the defining property of inflammable
substances was discarded in favour of an explanation based on combustion in
oxygen – an explanation involving concepts that were not part of previous expla-
nations. Moreover, even when concepts are retained (or, at least, the terms are
retained) their meaning is sometimes subtly or radically altered as they move from
theory to theory – for example, the meaning of mass, time and space as they move
from classical mechanics to relativity theory or acid and base as they move from
the Lowry-Bronsted theory of acids as hydrogen ion donors to the Lewis theory of
acid-base behaviour involving electron pair donation for covalent bond formation.
But does this really mean that scientists advocating these different theories cannot
understand each other? It seems unlikely. Is there no transfer of meaning across
‘paradigmic frontiers’? Again, it seems unlikely. Even when scientific knowledge
changes radically and rapidly there is often a chain of developments connecting
them, a chain through which a rational evolution can be traced, and although invested
with different meanings in the rival theories, concepts like mass, space and acid
constitute a “family of uses” in which “ an ancestry-descent relationship” can be
discerned (Shapere, 1984).
Theories are not simple statements about the world; rather, they are complex
structures of inter-related ideas nested within a much wider array of supporting
theories. Within this complex, different theories will have areas of intersection,
areas of common interest and aspects of experience for which they may proffer
exactly the same explanation. All of this is transferred across paradigms regardless
of the key theoretical difference at issue. In a so-called Kuhnian scientific revo-
lution, there are, of course, some radical conceptual innovations, but much of our
science is retained unchanged and some is retained in only slightly modified form.
Perhaps, then, scientific revolutions are not quite so revolutionary as Kuhn suggests.
Nor are they so rare. The posing and solving of conceptual problems is a regular

187

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
CHAPTER 9

feature of scientific practice; it is not confined to the short-lived periods of


paradigm crisis envisaged by Kuhn. There is little historical evidence to identify
periods in science when there was no questioning of basic theoretical assumptions
and no interest in identifying the conceptual problems that a paradigm generates.10
This is not to suggest that all scientists are constantly questioning the fundamentals
of their discipline and can never agree on anything; at any one time, many scien-
tists will be engaged in normal science – i.e., ‘taking the paradigm for granted’
and working within it to extend its applicability (‘puzzle solving” as Kuhn calls it).
It is to suggest, however, that conceptual change and replacement is a relatively
common feature of (modern) science and it is to suggest that normal science
(with no questioning of basic principles ) is not quite so widespread nor so prosaic
an activity as Kuhn implies. It is, perhaps, more realistic to regard science as
occurring at many different and intermediate points on a continuum from major
revolution to standard (algorithmic) puzzle solving, and to regard the goals of
science, the methods of inquiry and the theoretical knowledge of science as being
constantly modified (Laudan, 1984a). Kuhn (1963) states that “much of the work
undertaken within a scientific tradition is an attempt to adjust existing theory or
existing observation in order to bring the two into closer and closer agreement…
(it) aims to elucidate the scientific tradition in which (the scientist) was raised
rather than to change it” (pp. 348 & 349). Nevertheless, he says, this paradigm-
based research has often produced a revolution.
History also tells us that major scientific revolutions are usually the outcome of
lengthy periods of concentrated work on persistent problems by groups of con-
tributors (see the DNA double helix story, for example), and only rarely the out-
come of the heroic efforts of a unique visionary like James Clerk Maxwell. Nor is
the impact of a scientific revolution always immediate. Although we quite rightly
credit Charles Lyell with bringing about a revolution in geology, most of his
contemporaries did not accept his views, nor would Darwin’s theory of evolution
have proved the ‘fittest theory’ without T.H. Huxley’s vigorous campaigning.11
Even the Copernican revolution was not a sudden and universal shift of opinion;
despite Galileo’s heroic efforts, it was many years before physics was sufficiently
advanced for the dispute to be finally settled beyond doubt.
As discussed in chapter 7, Kuhn’s notion of incommensurability, the Duhem-
Quine thesis of the empirical under-determination of theories, and recognition of
the theory-dependence of experiments, seem to require adoption of the position
that theories are accepted for reasons other than empirical adequacy. For some, this
is taken to mean that theory acceptance has as much to do with the social context
as it has with the intrinsic merits of the idea under scrutiny. It almost goes without
saying that a key issue in the debate about suitable NOS views for inclusion in the
school science curriculum concerns the mechanism for theory appraisal. Why is a
particular scientific paradigm or theory accepted? What are the principal criteria
of scientific credibility? The curriculum should no more promote the view that
science has infallible methods for adjudicating the claims of rival theories than
it should promote the view that scientific knowledge is produced by arbitrary

188

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
LOOKING FOR BALANCE IN THE CURRICULUM

methods and has only subjective and socially contingent criteria for judging the
worth of knowledge claims and for accepting or rejecting theories.
While the theory-ladenness of observation is generally taken to rule out the
possibility of using observations to judge the merits of rival theories, this is only
the case if the theory with which the observation is laden is the theory that is under
test. In other circumstances (i.e., when the observation is impregnated with a
theory that is not under test), observational tests are perfectly legitimate. Further-
more, even though rival theories may employ different concepts and conceptual
relationships, they address a common area of interest, otherwise they wouldn’t be
rival theories. While there will be two unique sets of empirical problems, these sets
may be relatively small. Moreover, there will also be some common problems that
the theories must attempt to solve, so the rivals can be appraised in terms of how
well they succeed in doing so. Even without common ground, there are many
objective/rational ways of evaluating and appraising theories. Principal among
them is the capacity to make accurate and interesting predictions, though it should
be noted that there are many theories now known to be false that were good
predictors. Laudan (1977) regards problem solving capability as the major criterion
of a good theory: solved problems count in favour of a theory; anomalous problems
(unsolved by the theory but solved by one or more of its competitors) count against
it; unsolved problems (not yet solved by any theory) constitute the priorities for
future research. Of course, the ability to solve problems is no more a guarantee of a
theory’s truth than the ability to make predictions: Ptolemy’s epicycles solved the
problem of the retrograde motion of the planets perfectly well.
Consideration of observational evidence and the capacity to solve empirical
problems does not exhaust the scope for rational appraisal. Laudan (1977) identifies
“methodological well-foundedness” (p. 59) as a significant criterion. Certainly, a
sloppy or careless approach, ill-considered methods of data collection, vague and
inconsistent arguments, unwarranted conclusions and structural incoherence are
good grounds for rejecting a theory. Even Thomas Kuhn (1977), regarded by some
as an irrationalist (and certainly the inspiration for those who are irrationalists
or social constructivists), identifies four objective factors additional to empirical
accuracy: scope (a theory should be wide in scope and should extend beyond the
facts it was designed to explain; it should also unite phenomena that would other-
wise require separate explanation), fruitfulness (it should provide a framework for
on-going research, perhaps revealing previously unrecognized relationships or
leading to the discovery of new phenomena), consistency (it should be consistent,
both internally and with respect to other accepted theories) and simplicity (when
theories are empirically equivalent, the simplest is preferred). Following van
Fraassen (1980), we might add precision, practical utility and the capacity to unify
otherwise disparate fields to Kuhn’s list. For Kuhn, these criteria do not constitute
a set of rules to dictate or determine theory choice, but a set of values to influence
and inform it. Of course, they may point in different directions. For example, while
simplicity favoured Copernican theory, consistency favoured Ptolemaic theory.
Because competing but empirically equivalent theories will satisfy these four
criteria to different degrees, and because the criteria will be weighted differently

189

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
CHAPTER 9

by different people and may be differently weighted at different times in the


appraisal, “choice between competing theories depends on a mixture of objective
and subjective factors, or of shared and individual criteria” (Kuhn, 1977, p. 325).
So it seems that Kuhn is willing to admit that theories can be compared, although
science does not have an infallible means of doing so. In Kuhn’s (1970) words, we
have “no neutral algorithm for theory choice, no systematic decision procedure
which, properly applied, must lead each individual… to the same decision” (p. 200).
An element of subjective judgement, or what we might call “scientific common
sense”, will often be needed to decide between competing theories. Seen in this
light, Kuhn’s suggestion that the adoption of a new paradigm involves a certain act
of faith does not seem quite so radical, nor so destructive of scientific rationality.
In practice, the criteria are not so easy to apply. Even the criterion of simplicity is
somewhat ambiguous. For some, simplicity might mean fewer variables; for others,
it might mean less complex mathematics. Nor is there any compelling reason
to assume that nature itself is simple. Attractive as simplicity is as a criterion of
theoretical acceptability, it is a human preference and its appeal is no guarantee of
its appropriateness (Bunge, 1963).

SCIENTIFIC KNOWLEDGE AS A SOCIAL CONSTRUCT

It is important that the school science curriculum achieves a sensible balance


between the view that science is absolute truth, ascertained by value-free dis-
interested individuals using entirely objective and reliable methods of inquiry, and
the dangerously relativist view that ‘scientific truth’ is any view that happens to
suit the prevailing cultural climate or reflects the interests of those in power. To
paraphrase Helen Longino (1990), the first account is a logical analysis that is
historically unsatisfactory and the second is an historical analysis that is logically
unsatisfactory. There is no doubt that scientific practice is profoundly influenced
by the social context in which it is located. The key point at issue is the extent of
this social influence. As Giere (1988) puts it: “The real issue is the extent to which,
and by what means, nature constrains scientific theorizing” (p. 55). For those at
one extreme, social factors are acknowledged to influence research priorities, but
little else; for those located at the other extreme, scientific knowledge is regarded
as social in content as well as in origin. In other words, the second position is that
scientific knowledge is no more than a social construct and, therefore, has no more
legitimate a claim to describe reality than any other form of knowledge. One res-
ponse to this extraordinary proposition (extraordinary to scientists and science
educators, that is) is to point out that such theorists are hoist by their own petard: if
all knowledge is merely a social construct (and has no particular claim to truth)
then so is the proposition that all knowledge is merely a social construct, and has no
particular claim to truth. Another response is to ask for a clear illustration of the
ways in which social conditions determine scientific knowledge. It is unremarkable
to state that science is produced in a social context, but where is the evidence that
specific elements of the social context lead unerringly and inevitably to particular
scientific knowledge? In pursuing this line of questioning, Slezak (1994) asks what

190

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
LOOKING FOR BALANCE IN THE CURRICULUM

elements in the social circumstances surrounding Isaac Newton led directly to the
inverse square law? In different social circumstances, Slezak asks, would Newton
have formulated the inverse cube law of universal gravitation? And, in turn, was
Newtonian physics replaced by Einsteinian physics simply because the social
climate changed or because the new theory explained more and solved some
outstanding problems? Developing an interest in gravity and choosing to express
that understanding in mathematical terms was, no doubt, a consequence of the
immediate social context in which Newton found himself; so, too, his underlying
assumption that the universe can be understood in mechanistic terms; but the
precise relationship at which Newton arrived (the inverse square law) was deter-
mined by features, relationships and phenomena in the real world.
Recognizing that ideas emerge and are accepted and utilized under particular
intellectual circumstances, and that these circumstances are impacted by social and
political factors, is not to admit that the knowledge produced is simply a product of
the wider social circumstances. And certainly not without more convincing evi-
dence than the sociology of science has so far been able to produce. How can
social constructivists account for those situations in which very different theories
emerged in the same social milieu? How can they explain why we continue to
believe in theories that were produced in social circumstances very different from
those of the present day? Boyle’s Law was a product of 17th Century England,
Darwinian evolutionary theory a product of 19th Century England, and Heisenberg’s
Uncertainty Principle a product of early 20th Century Germany – each very different
from my social circumstances in early 21st century New Zealand. Yet I, like many
others, continue to accept these propositions. The crucial point is that these ideas
were adopted principally because of their capacity to address matters of urgency in
science, not because of their sociopolitical appropriateness.12 Indeed, it is fair to
say that, far from being socially determined, the profoundly innovative theoretical
advances brought about by scientists like Copernicus, Galileo, Newton and Darwin
flew in the face of social convention – particularly, religious convictions. In
addition, there are notable examples of theories being neglected despite their con-
formity with the prevailing zeitgeist. The ultimate requirement is that theories
have empirical support, regardless of their appropriateness to the prevailing socio-
political climate. As O’Hear (1989) comments, “ Despite all his prestige and the
full backing of an oppressive and dictatorial ideological state apparatus, Lysenko was
not able to make his wheat grow” (p. 214).
Slezak (1989) has argued that scientific knowledge can now be produced
outside any social context by means of computers. He describes how the BACON
software has enabled computers to use raw data to generate some of the accepted
laws of chemistry and physics – laws that were originally developed in very dif-
ferent sociocultural circumstances.
The success of BACON and AM programs provide empirical confirmation of
the view that the specific historical circumstances which undoubtedly attend
scientific reasoning do not play a decisive role… On the contrary, the pro-
grams demonstrate that the widely varying social factors attending the origin-
nal discoveries played no part in determining their specific contents. (p. 571)

191

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
CHAPTER 9

I would argue that the interests, expectations and understanding of those who
designed the BACON and AM programs are necessarily embedded in the software
and in decisions about what data to input, and in what form. Stated baldly, the
computer will only search for solutions in ways it is instructed to search. In that
sense, there is a social context.
Some sociologists go on to argue that it is not just facts and explanations that are
socially determined, but criteria of truth and the nature of scientific rationality
itself. In other words, decisions about the validity of arguments, the truth of know-
ledge claims, the correct application of principles, and so on, are all made in terms
of criteria determined by the socially dominant group – in this case, by influential
people within the community of scientists. The so-called “strong programme” in the
sociology of knowledge begins the chain of argument with the perfectly reasonable
proposition that we do not know the world as it really is, but only as mediated
through our conceptual frameworks. This is no different from the arguments I have
made in earlier chapters of this book. The argument continues as follows: this
framework is a human construction – it could have been otherwise – and is relative
to the social structure (i.e., it was agreed within the community of practitioners). Is
it still a reasonable position? It is only a short step from the assertion that all
knowledge is socially constructed to the claim that the criteria by which we decide
on truth and falsehood are also socially constructed and could, therefore, be
otherwise – that is, not only knowledge, but rationality itself is merely a convention
among scientists. Eventually, the chain of argument arrives at the position that
truth and falsity are simply ‘institutionalized conventions’. In other words, ‘truth is
as you choose to see it’ or as important members of the community of scientists
choose to see it; the real world imposes no constraints. Bhaskar (1975) responds as
follows:
Taken literally, it would imply that a chromosome count is irrelevant in
determining the biological sex of an individual, that the class of the living
is only conventionally divided from the class of the dead, that the chemical
elements reveal a continuous gradation in their properties, that tulips merge
into rhododendron bushes and solid objects fade gaseously away into empty
space. (p. 213)
Of course, some would say that this is how the world is constituted. There are
sexually ambiguous persons, just as there are transsexuals and hermaphrodites, and
many would argue that gender is a more socially appropriate category descriptor
than sex simply because it is less categorical. Many people believe that the dead
are accessible at certain times and in certain places, and that they constitute part of
the ‘spiritual fabric’ of society. Despite spectacular advances in molecular biology
it is still not always clear what ‘life’ consists in, or whether entities such as viruses
should be classified as living or non-living. My understanding of chemistry is
exactly what Bhaskar jokes about: categories such as acid and base, covalent and
electrovalent, metal and non-metal impose distinctions that do not adequately
describe the intermediate character of many substances. Moreover, solid objects
have a vapour pressure and many can be detected by smell at quite a distance, as

192

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
LOOKING FOR BALANCE IN THE CURRICULUM

observation of canine behaviour quickly reveals. The notion of social determination


is not so easily shrugged off when we remember that witches, dragons, mermaids,
phlogiston and transmutation of the elements have all had their ‘day in the sun’.
Arguments like these lead to the proposition that different groups could legiti-
mate different standards of validity, truth and correctness, according to their own
interests and their level of sociopolitical power. The reference to ‘power’ is key:
reality is never likely to be determined by all of us; it is determined by some of us.
Given the relationship between knowledge production and commercial, political
and military power, “truth is what the powerful say it is” and those in positions of
power and influence determine reality for the rest of us. It is these relationships
between power and knowledge and between power and scientific practice that we
should address through the politicization of the science curriculum (Hodson, 1994,
2003). For example, we should ask: Whose interests are being served by particular
kinds of scientific research? Whose interests are being promoted by particular
kinds of technological innovations? Whose interests support and determine the
treatment of scientific and technological issues in the media? How is public
opinion on scientific and technological matters formed and consent to particular
policies manufactured by media manipulation? It is here that the sociology of
science makes its unique and powerful contribution to the attainment of scientific
literacy (see arguments in chapter 1). But just how influential is the dominant
social group in determining scientific knowledge and, therefore, what counts as
reality? To what extent is scientific knowledge determined by the nature of things,
the nature of the method by which it is generated and the need to provide a
coherent argument linking evidence and conclusions? To what extent is it inde-
pendent of those who use the method? Clearly there is a sense in which reality is
determined by those in power, just as much in science as in any other area of
human concern.13 There is no doubt that particular aspects of reality are only
available to the initiated. Only the privileged members of the ‘scientific club’ have
the conceptual understanding and the necessary language resources to access
and understand the data and its significance as support for a particular belief. Data
have no meaning outside a particular interpretative framework; explanations have
no meaning outside a particular linguistic framework, as the following example
illustrates:.
NH4NO3 dissolves in water endothermically because the lattice energy of
anhydrous NH4NO3 is greater than the combined enthalpies of solvation of
the ammonium ion and nitrate ion.
By controlling access to the community of practitioners and to the discourse
of science, we can determine reality differently for different people.14 Also, we
can control what knowledge is admitted into the knowledge store of science by
controlling the organs of publication (academic journals, the media, the Internet).
Does this mean that the ‘powerful’ groups can authorize any knowledge they
choose? Are there any constraints on this power? It would be grossly misleading to
believe that there is only one conceivable representation of the world, or that we
can know what the world is like independently of our conceptual structures. But

193

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
CHAPTER 9

this is not to say that the world is merely a construct of the human mind, that our
knowledge is purely arbitrary or that individuals are free to make unique inter-
pretations of sense data or to fabricate any world that happens to suit them. The
admission that observation is theory-dependent and that theories are created by
individuals does not mean that science loses all objectivity. The admission that
theoretical explanations could be different does not reduce science to mere fashion,
prejudice or social convention. Science cannot be anything that important scientists
choose to say it is. The world does limit us in some ways. Others put it more starkly:
whatever social constructivists say about knowledge being a human creation, wasps
still sting and dogs still bite, sound takes time to travel across a pond, wood burns
and soap bubbles cannot be nailed to the wall. Limits are set to the ways in which
scientists’ interests determine knowledge by (i) the interests of others and (ii) the
nature of the real world. As Longino (1990) argues, what constitutes the world of
science is not a given, but nor is it entirely fabricated. It is a product of the inter-
action between the external real world and our intellectual needs and capabilities.
Of course, what we (as scientists) contribute may change over time in response
to sociocultural change, technological innovation and new theorizing. Moreover,
there are always features of the real world that we ignore or that remain unappre-
hended. Social forces do not determine how the natural world is constituted and
how it behaves, but they do ‘open scientists’ eyes’ in particular ways, direct their
attention to particular phenomena and events, and impact on the ways in which
they make sense of them using procedures devised and sanctioned by scientists and
using technological artifacts designed in accordance with existing theory.

REALISM AND INSTRUMENTALISM REVISITED

As the foregoing discussion makes clear, questions about the extent to which we
should regard scientific knowledge as a social construct speak directly to the debate
about the relative merits of realism and instrumentalism as appropriate positions
for the school science curriculum (see chapter 6). Realists, whether they are philo-
sophers, historians or sociologists of science, are concerned with what scientists do
in order to elucidate the nature of reality – that is, how they go about the process of
discovery and theory building. In contrast, social constructivists and instrumenta-
lists argue that science is not involved in finding out about things that already exist
(the realist position) but in creating entities can be used to construct reality – thus
asserting that entities such as electrons, magnetic fields and genes do not exist in
any meaningful way outside this social construction. If we accept Bloor’s (1976)
argument that science should be regarded as creating the phenomena it describes,
we should presumably treat viruses, volcanoes and nuclear fission as mere social
constructs/conventions. Only a fool would do so! Richard Dawkins (1995) puts the
absurdity of this position in a nutshell when he says: “Show me a relativist at thirty
thousand feet and I will show you a hypocrite” (p. 32).
As discussed in chapter 6, the standard case against antirealists of all kinds
hinges on the fruitfulness of theories, as Putnam (1978) argues with respect to both
Newtonian physics and relativity theory.

194

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
LOOKING FOR BALANCE IN THE CURRICULUM

If these objects don’t really exist at all, then it is a miracle that a theory which
speaks of gravitational action at a distance successfully predicts phenomena;
it is a miracle that a theory which speaks of curved space-time successfully
predicts phenomena. (p. 19)
When we use scientific theories to make technological interventions, the devices
we make usually work, and often work very successfully: spacecraft with their in-
board telescopes are capable of sending back close-up pictures of Mars, Jupiter or
Saturn; laser surgery does enable some sight defects to be repaired; nuclear fission
can be harnessed to produce electricity. At a theoretical level, it could be argued
that disputes in science reach closure precisely because the ‘successful’ theory more
closely corresponds to reality – that is, it has more realist features than its com-
petitors. Of course, we do not always need sophisticated scientific theories in order
to build successful technological artifacts. An approximate or simplified theory is
often good enough for practical purposes. Spectacles, telescopes and microscopes,
for example, can all be successfully manufactured using optical theories dating
from the mid-17th Century. Rival theories are sometimes equivalent for techno-
logical purposes and even falsified theories can be used successfully in some
practical contexts (Bunge, 1967).
Sociologists have done a great service in illuminating aspects of the day-to-day
lives of practising scientists and in ‘putting a human face’ on science. However, it
seems that in their eagerness to give priority to the social dimensions of science
they may be guilty of ignoring the most important facet of the scientific enterprise:
the theories generated by scientists and the reasons why these theories are so
successful. In drawing attention to this misplaced focus, Laudan (1990) asks us to
consider an analogy.
No one would entertain for a moment the idea that the central problem in the
history of music was how symphonic performers organised themselves. They
would reject out of hand the idea that the central problem in the history of
painting should be the question of how painters’ patronage was arranged. Yet
it is today seriously being proposed… that the central questions about science
have to do, not with the master works which science has produced, but rather
with the mundane minutiae of the life of a ‘normal’ scientist or with the ways
in which scientists garner political support for their activities. (p. 52)
So what position should we adopt in the school science curriculum regarding the
realist-instrumentalist debate? Solomon (1999) asks: “Should we teach that electrons,
the energy concept, and the colliding molecules in a gas, which we have such
difficulty making real and believable to our students, are not real at all? And which
should we teach first – their relativity or their unreality?” (p. 4). I share her con-
clusion that “science simply cannot be taught at all if the students are to be told that
there is no point in believing what is being said” (p. 5). Millar (1997) makes the
same argument: “Is it wise, from the perspective of effective teaching and learning,
to portray demanding and difficult ideas as tentative conjectures? Is this likely to
increase students’ motivation to undertake the challenging mental work of coming
to terms with these ideas, so that they can use them as tools for their own thinking

195

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
CHAPTER 9

about new problems and situations?” (p. 91). Millar goes on to state that there is
moral issue here. For example, he says, “the germ theory of disease is intended to
provide students with understanding which can be practically useful in taking
decisions about their own health, is it responsible to portray generally accepted
core knowledge as tentative conjecture?” (p. 91). At the other extreme, Eflin et al.
(1999) advocate naïve realism (belief in the absolute truth of scientific knowledge)
on the grounds that anything else is too difficult for students. I profoundly disagree
and, as argued in chapter 6, I am inclined to adopt a critical realist position on the
grounds of authenticity, flexibility and pedagogical utility.
I argue for the authenticity of critical realism on the grounds that most scientists
are realists of a kind, for whom the principal goal of science is to obtain knowledge
of maximum clarity, validity, reliability and predictive capability. Indeed, all the
scientists in the study conducted by Wong and Hodson (2008a,b) subscribed to the view
that the predictive power and technological applicability of current scientific theories
can be taken as strong evidence that in many respects we are getting closer to
knowing some of the truths about the universe. Although we cannot describe the
world except through the conceptual frameworks we have devised, we can test the
adequacy of those frameworks by interacting with the world. Science assumes that
there is regularity in nature (regularity is not solely a creation of human imagi-
nation) and that whatever underpins this regularity can be ascertained. Thus, the
prime thrust of science is to generate explanations of nature and to use them to
control, manipulate and generate more knowledge. The authors of Science for All
Americans (AAAS, 1993) express it as follows: “Scientists assume that even if
there is no way to secure complete and absolute truth, increasingly accurate approxi-
mations can be made to account for the world and how it works” (p. 4).
My claim for flexibility rests on the point made in chapter 6 that critical realists
can be realist about those conceptual structures they regard as genuine attempts to
uncover the truth and can be instrumentalist about those that have utilitarian value
(for prediction, control, etc) but no claim to truth. I suggested that the former should
be termed theories and the latter models. Matthews (1998b, p. 166) argues that this
position also enables students to recognize that “science is a human creation, that it is
bound by historical circumstances, that it changes over time, that its theories are
underdetermined by empirical evidence, that its knowledge claims are not absolute,
that its methods and methodology change over time, that it necessarily deals in
abstractions and idealizations, that it involves certain metaphysical positions, that its
research agendas are affected by social interests and ideology” and still claim that in
many respects it describes the real world.
My criterion of pedagogical utility is based on the assumption that realism is
more likely to motivate students, just as it motivated scientists like Galileo and
Einstein. Students who are critical realists are more likely to think that they have
the ability to build understanding and explanations for themselves through careful
and critical reflection on their experiences (especially those we present in science
lessons). But if scientific knowledge is absolute truth, this knowledge has to be
learned; and if scientific knowledge is nothing more than a human invention (i.e., a
mere social convention), then learning science also reduces to learning what others

196

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
LOOKING FOR BALANCE IN THE CURRICULUM

prescribe. A critical realist position enables students to construct their own models
to solve short-term problems, where initially the only criterion of acceptability is
‘what works’. By comparing and contrasting their model with models constructed
by others, students see the importance of robustness, breadth and scope, and pre-
dictive capability. Later, they come to appreciate how further rounds of inquiry and
reflection can result in these models being developed into realist theories.

ENDNOTES
1
Although a discussion of technological literacy is outside the scope of this book, a combined notion
of scientific and technological literacy would also include: (i) understanding of basic technological
concepts, principles of design and criteria of evaluation, and (ii) possession of a range of basic
hands-on skills, the capacity and confidence to tackle technological problems, and the willingness to
develop new knowledge and skills.
2
I am not claiming that all ten myths are promulgated by all science curricula. Rather, that most
curricula promote one or more of them and, across the range of curricular provision, all ten are
evident.
3
Smith et al. (1997) accuse Alters of bias in designing the questionnaire and interpreting the responses
to make the differences appear greater than they really are – charges that Alters (1997b) con-
temptuously dismisses. There is also some dispute over what the term nature of science (NOS)
encompasses. In recent publications, Lederman et al. (2002) and Lederman (2006, 2007) seek to
restrict its use to the characteristics of scientific knowledge (i.e., to epistemological considerations)
and to exclude all activities related to the collection and interpretation of data and the derivation and
publication of conclusions. The definition adopted in this book is considerably broader, and
includes: the characteristics of scientific inquiry; the role and status of scientific knowledge; how
scientists work as a social group; how science impacts, and is impacted by, the social context in
which it is located; the language of science; its underlying values; and so on. Thus, my usage is
closer to the notion of “ideas about science” used by Bartholomew et al (2004): nature of science,
plus “the social influences on science and technology, the nature of causal links, risk and risk
assessment, and the impact of science and technology on society” (p. 656).
4
Good and Shymansky (2001), however, question whether there really is a consensus, arguing that
there are major inconsistencies between and within the lists of NOS items in Benchmarks for
Scientific Literacy (AAAS, 1993) and National Science Education Standards (NRC, 1996), depend-
ing on whether the reader adopts a philosophy of science perspective or a history of science
perspective.
5
Lederman (2007) is insistent that students and teachers recognize that “observations are descriptive
statements about natural phenomena that are ‘directly’ accessible to the senses (or extensions of the
senses) and about which several observers can reach consensus with relative ease… Inferences, on
the other hand, go beyond the senses” (p. 835). He also states that “laws are statements or descriptions
of the relationships among observable phenomena… Theories, by contrast, are inferred explana-
tions for observable phenomena” (p. 835).
6
Osborne et al. (2003) detected some differences between the views of scientists and the views of
philosophers and sociologists, and some substantial differences among science teachers.
7
Questions about the unity/disunity of the sciences raise important questions concerning our capacity
to formulate robust criteria to demarcate science and non-science – a crucial element of scientific
literacy.
8
There is a certain irony here. Teachers often have to work very hard in the early years of science
education to get students to appreciate that scientific knowledge is tentative. Youngsters are naturally
inclined to take all knowledge (especially knowledge located in books or heard on television) as
secure and certain. Of course, by making such an assumption the world can be rendered stable and
predictable, and so much easier to cope with.
9
There is a sense in which the situation envisaged by Kuhn is nonsensical. It theories really are
incommensurable there is no way of deciding whether they address different phenomena or address
the same phenomenon in ways that are incompatible and, therefore, requiring of resolution. There
are no ways of judging compatibility and incompatibility, difference or sameness (Longino, 1990).

197

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
CHAPTER 9

10
Kuhn (1970) and Lakatos (1974) both argue that the central tenets of an established theory (paradigm
or research programme) are immune from criticism – an assertion that does not correspond to the
historical evidence.
11
Although a distinguished scientist in his own right, Thomas Huxley (1825-1895) is often remem-
bered as “Darwin’s Bulldog” for his vigorous advocacy of evolutionary theory.
12
Of course, the observation that some scientific knowledge is unintelligible to those outside the
particular social context of its production is clear evidence of a measure of social construction,
though it shouldn’t be taken as evidence of a causal link between social circumstances and scientific
knowledge.
13
We should also turn the critical spotlight on ourselves, and ask to what extent are we, as science
teachers, concerned to determine reality for those we teach? And, in doing so, whose interests are we
serving?
14
Essentially the same situation exists with any highly context-dependent and culturally specific
knowledge – only those who gain admission to the sub-culture and its specialized language will have
access to this knowledge.

198

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
CHAPTER 10

FURTHER THOUGHTS ON SOCIAL CONSTRUCTION


AND SCIENTIFIC RATIONALITY
A View for School Science

Thirty years after Stephen Brush (1974) famously asked whether the history of
science should be rated X, Douglas Allchin (2004) posed a similar question
regarding the sociology of science. For Mario Bunge (1991), the answer is clear. In
a savage attack that is worth quoting at length he says that the sociology of science
is guilty of adopting seven misleading tenets that, taken together, constitute “a
grotesque cartoon of scientific research” (p. 525).
These are externalism, or the thesis that conceptual content is determined by
social context; constructivism or subjectivism, the idea that the inquiring
subject constructs not only his accounts of facts but also the facts themselves,
and possibly the entire world; relativism, or the thesis that there are no objective
and universal truths; pragmatism, or emphasis on action and interaction at the
expense of ideas, and the equation of science with technology; ordinarism, or
the thesis that scientific research is pure perspiration and no inspiration, and
the refusal to accord science a special status and to distinguish it from ideology,
pseudoscience, or even antiscience; adoption of obsolete psychological
doctrines, such as behaviorism and psychoanlayis, and the substitution of a
number of unscientific or even antiscientific philosophies – such as linguistic
philosophy, phenomenology, existentialism, hermeneutics, ‘critical theory’,
poststructuralism, deconstructivism, or the French school of semiotics, as the
case may be – for positivism, rationalism, and other classical philosophies.
(pp. 524-525)

Equally deplorably, but at the opposite extreme, the accounts of scientific ration-
ality presented in many science curricula fail to acknowledge that sociopolitical
forces play any role at all in determining the direction of science and establishing
research priorities. By neglecting to consider the day-to-day practices of scientists,
they fail to distinguish between ‘science-in-the-making’ and ‘ready made science’
(the edited, sanitized, ahistorical and systematized account of science common to
so many textbooks). They take no account of the daily struggle to extract meaning-
ful data, develop robust techniques, establish credible interpretations and create
convincing arguments.
The question at issue here and in the previous chapter is: what would constitute
a sensible balance between the view that scientific knowledge is entirely inde-
pendent of the social context in which it is generated and the position that says all
knowledge (including science) is no more than a social construct? While it would
be absurd to claim that the products of scientific inquiry and theory-building cannot
199

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
CHAPTER 10

be understood outside the sociocultural context in which they were produced, as


some sociologists argue, appreciation of the sociocultural milieu within which
particular scientists work (or worked) provides a context for a better understanding
of their priorities, working styles and criteria of judgement. This applies just as
much to a full understanding of the elegant rationalist work of Isaac Newton as it
does to understanding the principles of alchemy, the theory of phrenology or the
practice of acupuncture.
Perhaps there is value in drawing a distinction between epistemic relativism and
judgemental relativism (Knorr-Cetina & Mulkay, 1983). Epistemic relativism asserts
that knowledge is rooted in a particular time and place. Knowledge doesn’t just
mimic nature, it also reflects the society that produces it (at least in some respects).
Judgemental relativism makes the additional claim that all forms of knowledge are
‘equally valid’. In other words, any one human construction is as good as any other.
We can acknowledge the first without accepting the second. Socially produced
knowledge still has to meet its purpose, and for science that purpose is to explain
the nature and behaviour of the natural world in reasonably comprehensible terms,
and with predictive capability. And it has to gain consensus among other prac-
titioners who have this same commitment to theory building. The appropriate
question is not “Is science objective and rational?” but rather “What is the nature of
scientific rationality and objectivity?”
In much recent sociological writing scientific practice is portrayed as an exer-
cise in literary persuasion, in which theory acceptance depends less on a rational
appraisal of evidential adequacy than on shrewd manoeuvres, tactics and rhetoric.
As Latour and Woolgar (1986) state, the worth of a theory is established and mea-
sured by the “increasing number of people from whom it extracts compliance”
(p. 285) and “the degree of accuracy (or fiction) of an account depends on what is
subsequently made of the story, not on the story itself” (p. 284). Although it is no
more than common sense to state that a knowledge claim only becomes established
when it is used (as it stands or in modified form) by other scientists, Latour and
Woolgar seem to imply that whether this occurs or not is simply a consequence of
the scientist’s skill with literary techniques and her/his capacity to use inscription
devices to lend rhetorical power to the argument (see chapter 7). Drawing heavily
on the work of Foucault (1970, 1972/1982) on the emergence of professional dis-
course within a network of social, political and economic relations, Latour and
Woolgar conclude that science is simply a form of discourse that has acquired a
particular form of organization and a set of historically and socially contingent rules
to establish what is admissible and what is true/false. These characteristics could
have been developed differently. In consequence, the knowledge that is accepted as
science simply reflects the interests and ideology of those currently in positions of
power. Using that rationale, Blondlot’s theories about N-rays were not so much
discredited on intellectual grounds as out-manoeuvred on sociopolitical grounds
(Klotz, 1980; Nye, 1980; Ashmore, 1993).
Fuller (1992) argues that the opportunistic and frequently self-serving activities
of scientists are rational in some broad understanding of the term because they
often succeed in what they are trying to do (publish articles, persuade others, win
research grants, and so on), or they fail “in ways that permit them to continue and

200

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
FURTHER THOUGHTS ON SOCIAL CONSTRUCTION

improve upon their efforts” (p. 401). Acknowledging that published accounts of
science are sanitized and depersonalized by omitting all reference to dead ends and
failed experiments, and by minimizing the role of intuition and maximizing the role
of rational, systematic planning is not to say that scientific writing is a ‘confidence
trick’ or an exercise in fiction building. The scientific paper is not autobiography or
history, it is not a narrative account of what happened. Rather, it is a rational
reconstruction of events into a chain of logical argument and evidence designed to
persuade readers of the legitimacy of a particular point of view. There is more
involved here than literary techniques. No matter how persuasive the rhetoric,
argument is won or lost on grounds of rational criteria that extend beyond the
immediate sociopolitical context of the various players. For most scientists there is
a key role for observational and experimental support, successful predictions,
logical argument, internal and external consistency, absence of decisive counter-
examples, and so on. Isaac Newton’s views did not prevail because he was a better
politician, and adept at out-manoevring his rivals; they prevailed because they were
scientifically significant, rationally justified and cogently argued.
I am inclined to follow Laudan’s (1977) advice that there is no need to invoke
social causes, either internal to science or external to it, unless there are no rational
reasons at all that can be invoked. It may well be, as Forman (1971) argues, that
Heisenberg’s Uncertainty Principle was readily accepted by German physicists in
the late 1920s because it enabled them to repudiate, to an extent, the charge that
science was overly rationalistic, mechanical and deterministic and left no scope for
human values. But arguing that the social context created the overall intellectual
climate in which a particular idea was regarded as acceptable is not to establish
social conditions as the primary cause of its acceptance.1 The Uncertainty Principle
was accepted primarily because it was an exceptionally good idea, because it
solved a number of long-standing problems, and because it enabled the discipline
of physics to make progress. In Laudan’s words, “the chief way of being scientifi-
cally reasonable or rational is to do whatever we can to maximize the progress of
scientific research traditions” (1977, p. 124). This can involve solving empirical
problems or making conceptual changes and adjustments that increase the theory’s
problem-solving effectiveness. It is this second rational move that is being invoked
here. The consequence of regarding factors in the social-cultural context as the sole
or principal reason for accepting an idea is the reduction of science to whim, caprice,
social conditioning or powerful propaganda. If we cannot regard the knowledge
base of science as reasonably secure and rationally justified we cannot regard its
claims any more seriously than those of astrologers, religious prophets or crystal
ball gazers. This is an unacceptable position for the science curriculum.2

SAVING SCIENTIFIC RATIONALITY

It is important to acknowledge that science is practised by people and that their


knowledge, experience and interests will influence science in all kinds of ways –
deciding priorities, influencing procedures, controlling access to the community,
controlling publication, and so on. Clearly, scientific knowledge is created, sustained,
transmitted and modified through social processes. Whether they are employed in

201

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
CHAPTER 10

universities, research institutes or industry, most scientists work as members of a


team. At the day-to-day level, an individual scientist is likely to be a member of a
group engaged in a specific investigation or series of investigations. At the wider
level, the scientist is a member of the institution’s team of scientists, with the
collective responsibilities that follow from it, and a member of the ‘invisible
team’ of scientists extending across all institutions who conduct similar and related
research. Science is a social practice in the sense that scientists are dependent on
one another for the intellectual and technical resources with which they work.
Existing theoretical and procedural knowledge, investigative techniques, labo-
ratory apparatus and instruments, and so on, collectively constitute the research
context within and upon which further progress is based. Science is a social
activity in the sense that the rules of scientific procedure and the legitimacy of the
‘product’ are determined by the community of practitioners. The wider community of
scientists determines what counts as acceptable scientific practice and exercises
strict control over what is admitted to the corpus of accepted knowledge through its
system of peer review. This community also exercises control over the education of
future scientists and the initiation of newcomers into the community of practice.
Furthermore, scientific practice is sociohistorically situated. Scientific theories are
human constructions; they depend for their existence on strongly motivated human
agents building and developing them, exploring their adequacy and rationality,
using them to address both theoretical and practical problems, and so on. Immersed
in the prevailing sociocultural milieu, scientists cannot remain immune from other
cultural and ideological influences. History tells us that scientific ideas come from
many sources, many of them outside the sphere of science. The cultural context is
likely to shape the questions that are asked, the topics that get pursued, the obser-
vations that are made and attended to, and what overall theoretical perspectives
are likely to find favour within the community. Thus, Darwin’s theorizing on evo-
lution in terms of ‘survival of the fittest’ was a consequence of his experiences in
the ruthlessly competitive society of Victorian England and the 19th Century
science of phrenology (or craniology) was promoted as a rational justification for a
society that already believed that women and non-Europeans are intellectually
inferior to Caucasian men (Fee, 1979; Gould, 1981; Hodson & Prophet, 1986). It
almost goes without saying that because scientists draw ideas from their cultural
location there is the ever-present danger of bias and distortion. Furthermore, socio-
cultural pressures can also function to resist or even exclude particular lines of
research and explanation, while encouraging others.
Although they may be reluctant to accept the findings of sociological and
ethnographic studies of science laboratories, most practicing scientists would readily
acknowledge the (often significant) role played by intuition, hunch, luck, greed,
personal needs, publishing pressures, and the like. They might admit Knorr-Cetina’s
(1995) assertion that scientists can, on occasions, be guilty of practices that are not
“open and above board”, such as hoarding of information, implementing perso-
nal and group biases, engaging in plagiarism, showing blind trust in their own data
or theory while dismissing those of rivals without sufficient consideration. Many
would also acknowledge that competition plays a key role and, for some, may even
be the major driving force. As David Hull (1988) observes, as long as there are

202

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
FURTHER THOUGHTS ON SOCIAL CONSTRUCTION

rewards for publishing papers, formulating new theoretical propositions and


developing new techniques and instruments, there will be scientists ready and
willing to produce them, and many who will strive to be the ‘first’ or the ‘best’. Of
course, competition can also lead to wasteful duplication of research and, in the
rush to publish, to the generation of substandard work and the publication of
incomplete studies. Even worse, it can lead to attempts to mislead competitors,
steal other scientists’ results, discredit other researchers by spreading false infor-
mation about them, and the regrettably all-too-common tactic of releasing part of
the research findings in support of an alternative explanation/theory in a deliberate
attempt to mislead or distract competitors (Monhardt et al., 1999; Wong and Hodson,
2008b).
The question for us, as curriculum builders and teachers, is whether these social,
cultural and economic factors constitute the main driving force for science or
whether they are influences that simply make science a human (and therefore imper-
fect) endeavour. Recognizing that science is a social activity, and that its methods
and procedures were established by people and are sustained by authority and
custom, is not to say that the scientific knowledge produced is empirically inade-
quate, socially expedient, irrationally believed or likely to be false. Rationality can
be retained in our account of science as a guarantee that the methods of appraisal
we choose to employ produce knowledge that is robust enough to solve empirical
and conceptual problems, and has some direct relationship to the actual world. We
choose particular methods because they have some objective value in helping us to
reach our principal goal of ‘getting a handle on the nature of reality’. The ratio-
nality of science is located in (i) careful and critical experimentation, observation
and argument and (ii) critical scrutiny of the procedures and products of the enter-
prise by other practitioners. It is a community-regulated and community-monitored
rationality. Science is socially constructed through critical debate. And those
involved in it have a commitment to maintain certain rigorous debating standards.
Longino (1990) makes the point that “it is the social character of scientific know-
ledge that both protects it from and renders it vulnerable to social and political
interests and values “ (p. 12). First, she says, there are public forums for the presen-
tation and criticism of evidence, methods, assumptions and reasoning – in particular,
conferences, academic journals and the system of peer review. Second, there are
shared and publicly available standards that critics must invoke in appraising work,
including but not restricted to empirical adequacy (as discussed earlier in this book).
Third, the scientific community makes changes and adjustments in response to
critical debate and is clearly seen by practitioners and members of the public to do
so. Fourth, the right to submit work for peer appraisal and criticism is open to all
practitioners; so, too, the right to publicly criticize the work of others. Of course,
the way in which the community of scientists exercises this public scrutiny at any
one time is subject to a whole range of social, cultural, political and economic
factors and the inclination of individual scientists to be persuaded by a particular
argument or to be swayed by particular evidence depends, in part, on their back-
ground knowledge, assumptions and values. In these respects, as Longino states,
the procedures of scientific appraisal are vulnerable to charges of social relativism.
The resilience of science in the face of such charges is a consequence of its

203

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
CHAPTER 10

openness to criticism by individuals of diverse backgrounds, experiences, interests


and underlying values. Questions will be asked about the appropriateness, extent
and accuracy of the data, how it was collected and interpreted, and whether the
conclusions follow directly from the data, and so on. The explanation will be scruti-
nized for internal consistency and for consistency with other accepted theories.
Particular attention will be directed to the background theory and assumptions under-
pinning the research design, and to the deployment of auxiliary theories and choice
of instrumentation and measurement methods. The possibility that these questions
may be answered differently by different appraisers, and that different perspectives
will be brought to bear in the appraisal process, is the reason for upholding the
principle of academic equality and is one of the guarantees of scientific objectivity.3
In short, Longino (1990, 2002) argues that the critical scrutiny exerted on scientific
ideas by peer review and public critique via conferences and journals is the cen-
terpiece of scientific rationality and a guarantee of the objectivity and robustness of
the knowledge developed.
The formal requirement of demonstrable evidential relevance constitutes a
standard of rationality and acceptability independent of and external to any
particular research program or scientific theory. The satisfaction of this standard
by any program or theory, secured, as has been argued, by inter-subjective
criticism, is what constitutes its objectivity. (Longino, 1990, p. 75)
Of course, one of the main concerns surrounding the increasing commerciali-
zation and militarization of science is that the demands of patent secrecy and
military secrecy remove science from public scrutiny, and thereby compromise
its objectivity. The privatization of science strikes directly and damagingly at our
collective ability to distinguish sound from unsound work, plausible from implau-
sible arguments, and even truth from falsehood. And, of course, science can be
rational and still get it wrong, in the sense of accepting ideas that subsequently turn
out to be false. The key point is that we improve our science because of constant
critical scrutiny by people with wide and varied experience and reflecting a range
of value positions. Over time, we identify wrong science and bad science (such as
science with clear evidence of bias), we detect and expose fraud, and we discriminate
science from non-science. The point at issue for the science curriculum is that all
accepted scientific knowledge has a well-argued and well-supported warrant for
belief. While acknowledging that verification of theories is not possible, this
position does recognize that science makes progress. Progress arises from continual
criticism and efforts to meet criticisms through modification and/or replacement of
theoretical structures. As Rorty (1991) argues, what makes science ‘special’ is that
scientists have done a better job than most other groups in implementing certain
values – in particular, reliance on persuasion rather than coercion and willingness
to consider alternative ideas.
With the goal of recreating this community-based and criticism-based model of
science in the classroom, Carl Bereiter (1994) identifies four basic requirements for
the establishment of progressive discourse within the community of scientists/
students. First, a commitment to work towards a common understanding satisfactory

204

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
FURTHER THOUGHTS ON SOCIAL CONSTRUCTION

to all; second, a commitment to frame questions and propositions in ways that


allow evidence to be brought to bear on them; third, a commitment to expand the
body of collectively valid propositions (what Bereiter has in mind here is similar to
the Lakatosian notion of a ‘hard core’, the propositions that participants will not
deny, even though they may not actively endorse them); fourth, a commitment to
allow any belief to be subjected to criticism if it will advance the discourse.
Bereiter notes that item four is the most problematic commitment. To paraphrase
Wittgenstein (1969), although you can doubt everything, you cannot doubt every-
thing at once; and to paraphrase Shapere (1982), the possibility of doubt is not
sufficient reason to doubt. We need to ask whether certain doubts should be taken
seriously. We should not doubt unless there is good reason to do so, no more
than we should believe unless there is good reason to do so. There comes a point
when doubt must be put aside. Without an established and taken-for-granted set
of theoretical propositions, significant further progress is impossible; but without
doubt, the possibility of scientific revolution is ruled out.

A MODEL FOR THE SCHOOL CURRICULUM

Popper’s ‘three worlds’ view of scientific knowledge, introduced towards the end
of chapter 6, might usefully be modified to take account of some of the socio-
logical perspectives developed in chapter 7. Replacing the second world of sub-
jective thought processes (the ‘mental world’ in Figure 6.1) by the complex social
activity of scientific practice (see Figure 10.1) provides a simple model for enable-
ing teachers to raise important questions about authentic scientific practice and address
a cluster of seemingly unconnected issues: equity and exclusion; values in science
and science education; ethnosciences, pseudoscience and scientific fraud; the issue of
realism versus instrumentalism or social constructivism.

Figure 10.1. Popper’s ‘three worlds’, adapted to include scientific practice

The model implies that scientific practice has the same kind of status as scientific
knowledge: it exists independently of its practitioners, though it is created, moni-
tored, sustained and validated by its practitioners and may, from time to time, be

205

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
CHAPTER 10

changed or modified as practitioners see fit or the situation demands. Importantly, the
world of scientific practice (World 2) is no longer a ‘secret garden’. It is open to
public scrutiny, and thereby subject to criticism and change. Scientific knowledge
(World 3) can be regarded as a ‘conceptual net’: concepts at differing levels of
abstraction are linked by generalizations of varying degrees of certainty and linked
to the natural world through ‘instances’ expressed in observable terms (Hesse,
1980). Scientific practice is the creative extension of this conceptual net to fit new
circumstances. In principle, inquiry is open-ended; it could lead anywhere. Nothing
in the ‘net’ fixes its future form, though clearly it influences the way in which
research is designed and conducted and constitutes the background against which
problems are defined and addressed. According to Popper (1972), scientific
problems exist because scientific knowledge has an autonomous existence outside
the minds of individuals or groups of scientists and, in consequence, some prob-
lems may remain unsolved or even unrecognized for some time. He compares this
situation with a nesting box in a garden. Eventually a bird will utilize an oppor-
tunity for nesting that has been there, undetected by other birds, ever since it was
installed. In the same way, problems concerning the adequacy of theories in
accounting for observations, problems concerning conceptual relationships within
or between theories, or whatever, are sensed by individuals who then use the approved
procedures of existing scientific practice to attack them and solve them. In this
undertaking scientists are driven by their individual perceptions, ideas and intuition,
and are guided by fellow practitioners. Science has achieved its remarkable success
not because the problems it tackles are simple or because nature is particularly easy
to study, but because scientists have refined and regulated their activities into a
particularly effective scientific practice: “a social endeavour, which over the cen-
turies has developed an approach appropriate to its limited goals, and where the
work of each individual is informed and controlled by that of his colleagues in this
endeavour, of the past, the present and the future” (Ravetz, 1971, p. 180).
Ravetz (1971) has likened the scientific enterprise to cathedral building, with
many individuals working over a long period of time on different aspects of a com-
plex structure. Scientific knowledge is achieved, he says, “by a complex social
endeavour, and derives from the work of many craftsmen in their very special
interaction with the world of nature” (p. 81). The production of scientific know-
ledge, like cathedral building, requires a community of practitioners with shared
knowledge, standards and goals, and it requires individuals with diverse and wide-
ranging skills and capabilities working under a measure of overall direction and
supervision (the “positive heuristics” of various research programmes and the con-
tinuing critical scrutiny by the community’s senior practitioners). The scientific
knowledge generated by individuals or teams of individuals, and subsequently
validated and accepted by the community, has an objective existence – just as a
cathedral built by individual artisans working to a community plan has an objective
existence. Thus, the intensely personal activity of creative research (within a team
or as an individual) produces objective, impersonal knowledge by means of pro-
cedures developed, validated and supervised by the community of practitioners, or
the community’s nominated agents. Basing their speculation on existing knowledge,
individual scientists sense a problem, formulate a strategy for investigating it,

206

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
FURTHER THOUGHTS ON SOCIAL CONSTRUCTION

collect evidence, and so on, using procedures, techniques and instrumental methods
developed and authenticated by predecessors, and subsequently present an account
of the work for critical evaluation by peers in the formal ‘public language’ of
science. Students can be sensitized to the existence of these different stages, and to
the crucial distinction between the free and open ‘private language’ of creative
thought and the formal, objective and de-personalized ‘public language’ of scientific
publication by incorporating what we might call ‘conference opportunities’ into
open-ended project work.
Laudan (1977) reminds us that contemporary scientific practice is largely
an exercise in problem solving. He divides problems into two broad categories:
empirical problems (any feature of the natural world that is need of explanation)
and conceptual problems. Conceptual problems can be internal (logical incon-
sistencies, ambiguities, circularities, etc) or external (conflicts with other theories),
while empirical problems can be categorized as solved problems (those satis-
factorily explained by the theory), unsolved problems (those that fall outside the
scope of the theory) and anomalous problems (those not solved by the theory, but
solved by a rival theory). Solved problems are important because they constitute a
substantial part of the warrant for belief in the theory’s central premises (what
Lakatos would call the “hard core”) and are especially important if rival theories
are unable to solve them. However, as history reminds us, solved problems are no
guarantee of truth: false theories can work just as well as true ones, and there are
many examples of discarded theories that were previously accepted as true because
they ‘worked well’. The goal of the researcher is to maximize the number and scope
of solved problems and minimize the number and scope of anomalies and con-
ceptual problems (compare the Lakatosian notion of a ‘positive heuristic’). Of
course, anomalies are potentially the most serious threat to a theory’s viability
because they can encourage scientists to abandon a theory in favour of a rival.
Closely related to empirical problems, though not mentioned by Laudan, are the
technical problems associated with the collection of experimental data. In many
fields of science progress is increasingly dependent on the design of new instru-
mentation techniques – not simply to speed up data collection but to provide the
corroborative evidence for theory building that may not yet exist. The theories that
best solve problems will contribute most to meeting the aims of science: making
accurate and reliable predictions, establishing manipulative control, increasing the
scope and precision of explanatory systems and, whenever possible, integrating
and simplifying the various components of those explanatory systems.4 Laudan
(1984a,b) makes the point that there is a constant and complex process of mutual
adjustment and mutual justification among aims, methods and theories. Aims
justify methods and must harmonize with theories; methods justify theories and are
chosen to meet aims; theories constrain methods and must harmonize with aims
(Figure 10.2). This view sits well with what we should be trying to teach students
about the contextual contingency of scientific inquiry.
As Abrams and Wandersee (1995) point out, contemporary science is highly
dependent on research funding, and regardless of whether funds are provided by
industry, the military, universities or government, there is increasing pressure for

207

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
CHAPTER 10

Figure 10.2. Interactions among aims, methods & theories, after Laudan
(see Duschl, 1990)

research to be directed towards the solving of practical problems, creating com-


ercial opportunities, establishing military advantage and meeting public interest needs.
These “societal aims”, as Abrams and Wandersee (1995) call them, prioritize
research funding and, in turn, drive the development of new research methods. In
many cases, concern with fundamental theoretical issues (sometimes known as
“Blue Sky” research) is pushed to the sidelines, a concern expressed by several of
the scientists interviewed by Wong and Hodson (2008b). In the pursuit of critical
scientific literacy two key questions for students to consider focus on who determines
the extent and direction of research funding and how research priorities are
established. Other questions follow: Should our priorities be different? How can we
intervene to re-prioritize? Asking such questions and subjecting the scientific
enterprise to close critical scrutiny is an essential element in the politicization of
students, which I take to be the ultimate purpose of achieving universal critical
scientific literacy.

TEACHING AND LEARNING METHODS

Research tells us that teaching students and student teachers about the nature of
science is much more successful when it is both explicit and supportive of critical
reflection (Abd-El-Khalick & Lederman, 2000). Teaching about science, like
any other cognitive learning outcome, should embody a specific tangible content
around which teachers can plan a systematic and coherent sequence of learning
experiences and assessment/evaluation activities. Implicit approaches, in which
students are expected to acquire understanding by ‘reading between the lines’ as
they engage in classroom activities, particularly practical work, have been shown

208

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
FURTHER THOUGHTS ON SOCIAL CONSTRUCTION

to be, at best, only moderately successful in achieving their goals. Also, as with
any other complex learning goal, adequate curriculum time is essential. Nature of
science learning cannot be rushed.
Making matters explicit does not entail a didactic approach. Rather, it suggests a
carefully focused approach using a variety of teaching and learning methods,
including hands-on activities, contemporary and historical case studies, reading and
writing activities and, most importantly of all, opportunities for guided reflection
and critical discussion. Figures 10.1 and 10.2 are useful in focusing teacher atten-
tion on key issues. For example, Figure 10.2 is useful for concentrating attention
on the complex relationship between theory and scientific inquiry, how scientists
utilize models as cognitive tools in their day-to-day problem solving, how theories
are articulated and revised, and how scientific arguments are constructed. The
notion of scientific practice in Figure 10.1 focuses attention on social dimensions
within the community of science and on the relationships of the community of
scientists with the wider society and its institutions. It reminds us that this com-
munity, like all other communities of practice, has its own distinctive language,
beliefs, codes of behaviour and values, each of which is a focus for curriculum
attention. Demarcation criteria loom large in such considerations: What is science?
Is it characterized principally by its methods, explanatory systems, criteria of
validity, underlying values, or some other criterion? How does science differ from
other forms of knowledge? How much room is there for manoevre and change
without the activity losing its designation as science? Demarcation is a crucial
aspect of critical scientific literacy – in particular, being able to distinguish good
science from bad science, detecting bias and distortion and differentiating among
science, pseudoscience and non-science The emphasis on science as a practice
reminds us that because science is conducted by people it is vulnerable, on
occasions, to bias, influence of vested interest, distortion and misuse for socio-
political, business and military ends. It is also subject to errors and may sometimes
follow false trails. Pressures on scientists can be such that they engage in mis-
conduct and even fraud. Teachers have important decisions to make about the
extent to which they will raise these kinds of issues in the curriculum. This book is
a plea to those teachers to afford learning about science a much higher profile than
has been usual in the past.

ENDNOTES
1
It is noteworthy that Forman neglects to mention that several major players in these events were not
German. Bunge (1991) reminds us that Schrodinger was Austrian, de Broglie was French, Dirac was
English, Bohr was Danish and Einstein is best described as a ‘citizen of the world’. Moreover, “the
place to which all the quantum physicists at the time flocked in pilgrimage and called ‘the Mecca of
the quantum theory’ was Copenhagen, not Gottingen, Berlin, Leipzig or Munich” (p. 541).
2
Of course, it is important to ask whether the desire to promote science as more rationally based than
these other knowledge claims, and therefore more valid and reliable, has any sound foundation.
Or whether it is, itself, merely the outcome of social conditioning.
3
Among several others, Harding (1998) has developed this into a generalized argument for the
urgency of involving more women, more members of racial/ethnic minority groups and more people
from non-industrialized cultures in the appraisal of scientific practice.

209

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
CHAPTER 10

4
Some would argue that creating artifacts to meet human wants and needs and the creation of a better
life also constitute legitimate aims for science (see chapter 6), while others would argue that these
aims are technological rather than scientific. In the complex world of 21st Century technoscience this
distinction is increasingly irrelevant.

210

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
REFERENCES
Abd-El-Khalick, F., Bell, R. L., & Lederman, N. G. (1998). The nature of science and instructional
practice: Making the unnatural natural. Science Education, 82(4), 417–437.
Abd-El-Khalick, F., & BouJaoude, S. (1997). An exploratory study of the knowledge base for science
teaching. Journal of Research in Science Teaching, 34, 673–699.
Abd-El-Khalick, F., & Lederman, N. G. (2000). Improving science teachers’ conceptions of the nature
of science: A critical review of the literature. International Journal of Science Education, 22(7),
665–701.
Abell, S. K., & Smith, D. C. (1994). What is science? Preservice elementary teachers’ conceptions of
the nature of science. International Journal of Science Education, 16, 475–487.
Abrams, E., & Wandersee, J. (1995). How does biological knowledge grow? A study of life scientists’
research practices. Journal of Research in Science Teaching, 32(6), 649–664.
Abruscato, J. (1988). Teaching children science (2nd ed.). Englewood Cliffs, NJ: Prentice Hall.
Abruscato, J. (2004). Teaching children science: Discovery methods for the elementary and middle
grades. Boston: Pearson Allyn & Bacon.
Adam, D. (2007, November 5). Solutions v. sophistry. The Guardian, “The Green List” supplement, pp. 6–7.
Agin, M. L. (1974). Education for scientific literacy: A conceptual frame of reference and some
applications. Science Education, 58(3), 403–415.
Ahlgren, A., & Walberg, H. (1973). Changing attitudes towards science among adolescents. Nature,
245, 187–190.
Aikenhead, G. S. (1990). Scientific/technological literacy, critical reasoning, and classroom practice. In S. P.
Norris & L. M. Phillips (Eds.), Foundations of literacy policy in Canada (pp. 127–145). Calgary: Detselig.
Aikenhead, G. S. (2002). Cross-cultural science teaching: Rekindling traditions for aboriginal students.
Canadian Journal of Science, Mathematics and Technology Education, 2(3), 287–304.
Aikenhead, G. S. (2005). Science-based occupations and the science curriculum: Concepts of evidence.
Science Education, 89(2), 242–275.
Aikenhead, G. S., Fleming, R. W., & Ryan, A. G. (1987). High school graduates’ beliefs about science-
technology-society. I. Methods and issues in monitoring student views. Science Education, 71(2),
145–161.
Aikenhead, G. S., & Ryan, A. G. (1992). The development of a new instrument: “Views on Science-
Technology-Society” (VOSTS). Science Education, 76, 477–491.
Aikenhead, G. S., Ryan, A. G., & Fleming, R. W. (1989). Views on science-technology-society. Saskatoon:
Department of Curriculum Studies (Faculty of Education), University of Saskatchewan.
Alberta Education. (1993). Achieving the vision. Edmonton: Alberta Education, Policy and Planning Board.
Allchin, D. (1995). How not to teach history of science. In F. Finley, D. Allchin, D. Rhees, & S. Fifield
(Eds.), Proceedings of the third international history, philosophy and science teaching conference (Vol. 1,
pp. 13–22). Minneapolis, MN: University of Minnesota Press.
Allchin, D. (1997). Rekindling phlogiston: From classroom case study to interdisciplinary relationships.
Science & Education, 6(5), 473–509.
Allchin, D. (2003). Scientific myth-conceptions. Science Education, 87(3), 329–351.
Allchin, D. (2004). Should the sociology of science be rated X? Science Education, 88(6), 934–946.
Allchin, D., Anthony, E., Bristol, J., Dean, A., Hall, D., & Lieb, C. (1999). History of science – with labs.
Science & Education, 8, 619–632.
Alters, B. J. (1997a). Whose nature of science? Journal of Research in Science Teaching, 34(1), 39–55.
Alters, B. J. (1997b). Nature of science: A diversity or uniformity of ideas? Journal of Research in Science
Teaching, 34(10), 1105–1108.
American Association for the Advancement of Science (AAAS). (1967). Science- A process approach.
Washington, DC: Ginn.
American Association for the Advancement of Science (AAAS). (1989). Science for all Americans.
A Project 2061 report on literacy goals in science, mathematics, and technology. Washington, DC: AAAS.
American Association for the Advancement of Science (AAAS). (1993). Benchmarks for scientific literacy.
Oxford: Oxford University Press.
Amis, M. (1995). The information. London: Flamingo.
Anderson, C. W. (1999). Inscriptions and science learning. Journal of Research in Science Teaching, 36,
973–974.
Apple, M. W. (1993). Official knowledge: Democratic education in a conservative age. New York:
Routledge.
Apple, M. W. (1999). Power, meaning and identity: Essays in critical educational studies. New York: Peter
Lang.

211

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
REFERENCES

Apple, M. W. (2000). The hidden costs of reform. Educational Policy, 14, 429–435.
Apple, M. W. (2001a). Educating the ‘Right’ way: Markets, standards, god, and inequality. New York:
RoutledgeFalmer.
Apple, M. W. (2001b). Creating profits by creating failures: Standards, markets, and inequality in education.
International Journal of Inclusive Education, 5(2 & 3), 103–118.
Appleton, K. (1993). Using theory to guide practice: Teaching science from a constructivist perspective.
School Science & Mathematics, 93, 269–274.
Arnold, B. (1990). The past as propaganda: Totalitarian archaeology in Nazi Germany. Antiquity, 64,
464–478.
Arnold, B. (1992). The past as propaganda. Archaeology, 45, 30–37.
Arons, A. B. (1983). Achieving wider scientific literacy. Daedalus, 112(2), 91–122.
Ashmore, M. (1993). The theatre of the blind: Starring a Promethean prankster, a phoney phenomenon, a
prism, a pocket and a piece of wood. Social Studies of Science, 23, 67–106.
Asimov, I. (1983). Popularizing science. Nature, 306(5939), 119.
Atkinson, P., & Delamont, S. (1976). Mock-ups and cock-ups: The stage management of guided discovery
instruction. In M. Hammersley & P. Woods (Eds.), The process of schooling: A sociological reader (pp.
133–142). London: Routledge & Kegan Paul.
Ault, C. R. (1998). Criteria of excellence for geological inquiry: The necessity of ambiguity. Journal of
Research in Science Teaching, 35(2), 189–212.
Ausubel, D. P. (1968). Educational psychology: A cognitive view. New York: Holt, Rinehart & Winston.
Bachelard, G. (1951). L’Activite rationaliste de la physique contemporaine. Paris: Presses Universitaires de
France.
Bachelard, G. (1934/1985). The new scientific spirit (A. Goldhammer, Trans.). Boston: Beacon. (Original
work published 1934)
Barad, K. (2000). Reconceiving scientific literacy as agential literacy. In R. Reid & S. Traweek (Eds.), Doing
science + culture (pp. 221–258). New York: Routledge.
Barber, B. (1962). Science and the social order. New York: Collier.
Barker, M. (1995). Esteeming prior knowledge: Historical views and students’ intuitive ideas about plant
nutrition. In F. Finley, D. Allchin, D. Rhees, & S. Fifield (Eds.), Proceedings of the third international
history, philosophy and science teaching conference (Vol. 1, pp. 97–102). Minneapolis, MN: University
of Minnesota Press.
Barlex, D., & Carre, C. (1985). Visual communication in science. Cambridge: Cambridge University Press.
Barman, C. R. (1997). Students’ views about scientists and science: Results from a national study. Science
and Children, 35(9), 18–24.
Barman, C. R. (1999). Students’ views about scientists and school science: Engaging K-8 teachers in a
national study. Journal of Science Teacher Education, 10(1), 43–54.
Barnes, B., & Edge, D. (1982). General introduction. In B. Barnes & D. Edge (Eds.), Science in context:
Readings in the sociology of science (pp. 1–12). Milton Keynes: Open University Press.
Bartholomew, H., Osborne, J., & Ratcliffe, M. (2004). Teaching students “ideas-about-science”: Five
dimensions of effective practice. Science Education, 88(5), 655–682.
Barton, A. C. (1998). Feminist science education. New York: Teachers College Press.
Barton, A. C., & Osborne, M. D. (1998). Marginalized discourses and pedagogies: Constructively con-
fronting science for all. Journal of Research in Science Teaching, 35(4), 339–340.
Bauer, H. H. (1992). Scientific literacy and the myth of the scientific method. Chicago: University of Illinois
Press.
Bell, B. (1993). Children’s science, constructivism and learning in science. Geelong: Deakin University
Press.
Bell, R. (1992). Impure science. New York: John Wiley.
Bell, R. L., Lederman, N. G., & Abd-El-Khalick, F. (2000). Developing and acting upon one’s conception of
the nature of science: A follow-up study. Journal of Research in Science Teaching, 37(6), 563–581.
Bencze, J. L. (1996). Correlational studies in school science: Breaking the science-experiment-certainty
connection. School Science Review, 78(282), 95–101.
Bencze, J. L. (2001). Subverting corporatism in school science. Canadian Journal of Science, Mathematics
and Technology Education, 1(3), 349–355.
Bencze, L. (2004). School science for/against social justice. In S. Alsop, L. Bencze, & E. Pedretti (Eds.),
Analysing exemplary science teaching: Theoretical lenses and a spectrum of possibilities for practice
(pp. 193–202). London: Routledge-Falmer.

212

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
REFERENCES

Bencze, L., & Alsop, S. (2007). School science for the people and the planet: Enabling education
in a milieu of global economization. Paper presented at the 3rd biennial ‘Provoking Curriculum’
conference, Banff, February.
Bereiter, C. (1994). Implications of postmodernism for science, or, science as progressive discourse.
Educational Psychologist, 29(1), 3–12.
Bereiter, C., Scaradamalia, M., Cassells, C., & Hewitt, J. (1997). Postmodernism, knowledge building, and
elementary science. The Elementary School Journal, 97(4), 329–340.
Bernstein, R. J. (1983). Beyond objectivism and relativism: Science, hermeneutics, and praxis. Philadelphia:
University of Philadelphia Press.
Betts, D. M. (1992). Developing an appreciation of the nature of science using the history of the periodic law
in a secondary school chemistry course. In G. L. C. Hills (Ed.), Proceedings of the second international
conference on history and philosophy of science and science teaching (Vol. 1, pp. 105–113). Kingston,
ON: Queen’s University.
Beveridge, W. I. B. (1961). The art of scientific investigation. London: Heinemann.
Beyer, L. E. (1998). Schooling for democracy: What kind? In L. E. Beyer & M. W. Apple (Eds.), The
curriculum: Problems, politics and possibilities (pp. 245–263). Albany, NY: State University of New
York Press.
Bhaskar, R. (1975). A realist theory of science. Hemel Hempstead: Harvester Wheatsheaf.
Billeh, V. Y., & Hasan, O. (1975). Factors affecting teachers’ gain in understanding the nature of science.
Journal of Research in Science Teaching, 12(3), 209–219.
Bloor, D. (1974). Popper’s mystification of objective knowledge. Science Studies, 4, 65–76.
Bloor, D. (1976). Knowledge and social imagery. London: Routledge & Kegan Paul.
Boadella, D. (1985). Wilhelm Reich: The evolution of his work. London: Arkana.
Board of Education. (1938). Secondary education with special reference to grammar schools and technical
high schools (The Spens report). London: H.M. Stationery Office.
Bohm, D. (1957). Causality and chance in modern physics. Philadelphia: University of Pennsylvania Press.
Bohm, D., & Peat, F. D. (1987). Science, order, and creativity. New York: Bantam Books.
Borges, J. L. (1964). The analytical language of John Wilkins. In J. L. Borges (Ed.), Other inquisitions 1937–
1952 (pp. 101–105). Austin, TX: University of Texas Press.
Born, M. (1924). Einstein’s theory of relativity (H. L. Brose, Trans.). London: Methuen.
Born, M. (1934). Experiment and theory in physics. Cambridge: Cambridge University Press.
Botton, C., & Brown, C. (1998). The reliability of some VOSTS items when used with preservice secondary
science teachers in England. Journal of Research in Science Teaching, 35(1), 53–71.
Bourdieu, P. (1975). The specificity of the scientific field and the social conditions of the progress of reason.
Social Science Information, 14(6), 19–47.
Bowen, G. M., & Roth, W.-M. (2002). Of lizards, outdoors and indoors: Translating worlds in ecological
fieldwork. In D. Hodson (Ed.), OISE Papers in STSE education (Vol. 3, pp. 167–189). Toronto: Imperial
Oil centre for Studies in Science, Mathematics and Technology Education, OISE-University of Toronto.
Bowers, C. A. (1996). The cultural dimensions of ecological literacy. Journal of Environmental Education,
27(2), 5–11.
Bowers, C. A. (1999). Changing the dominant cultural perspective in education. In G. A. Smith & D. R.
Williams (Eds.), Ecological education in action: On weaving education, culture and the environment (pp.
161–178). Albany, NY: State University of New York Press.
Brandes, A. A. (1996). Elementary school children’s images of science. In Y. Kafai & M. Resnick (Eds.),
Constructionism in practice: Designing, thinking, and learning in a digital world (pp. 37–69). Mahweh,
NJ: Lawrence Erlbaum.
Branscombe, A. W. (1981). Knowing how to know. Science, Technology & Human Values, 6(36), 5–9.
Bridgman, P. W. (1950). The reflections of a physicist. New York: Philosophical Library.
British Association for the Advancement of Science (BAAS). (1917). Science teaching in secondary schools.
London: BAAS.
Broad, W., & Wade, N. (1982). Betrayers of the truth. New York: Simon & Schuster.
Brock, B. (1989). Past, present, and future. In M. Shortland & A. Warwick (Eds.), Teaching the history of
science (pp. 30–41). Oxford: Basil Blackwell.
Brockman, J. (1995). The third culture: Beyond the scientific revolution. London: Charles Scribner.
Bronowski, J. (1977). The ascent of man. London: Book Club Associates/BBC.
Brown, J. S., Collins, A., & Duguid, P. (1989). Situated cognition and the culture of learning. Educational
Researcher, 18, 32–42.

213

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
REFERENCES

Bruner, J. S. (1960). The process of education. Cambridge, MA: Harvard University Press.
Brush, S. (1974). Should the history of science be rated X? Science, 183(4130), 1164–1172.
Brush, S. G. (1989). History of science and science education. Interchange, 20(2), 60–70.
Brush, S. G. (1990). Prediction and theory evaluation: Alfven on space plasma phenomena. Eos
(Transactions, American Geophysical Union), 71(2), 17–28.
Brush, S. (2002). Cautious revolutionaries: Maxwell, Planck, Hubble. American Journal of Physics, 70(2),
119–127.
Buchan, A. S., & Jenkins, E. W. (1992). The internal assessment of practical skills in science in England and
Wales, 1960–1991: Some issues in historical perspective. International Journal of Science Education,
14(4), 367–380.
Bunge, M. (1963). The myth of simplicity: Problems of scientific philosophy. Upper Saddle River, NJ:
Prentice Hall.
Bunge, M. (1967). Technology as applied science. Technology and Culture, 8, 329–347.
Bunge, M. (1991). A critical examination of the new sociology of science. Part 1. Philosophy of the Social
Sciences, 21(4), 524–560.
Bunge, M. (1992). A critical examination of the new sociology of science. Part 2. Philosophy of the Social
Sciences, 22(1), 46–76.
Burke, J. (1978). Connections. Boston: Little, Brown & Co.
Burns, J. C., Okey, J. R., & Wise, K. C. (1985). Development of an integrated process skills test: TIPS II.
Journal of Research in Science Teaching, 22(2), 169–177.
Butterfield, H. (1931). The whig interpretation of history. London: G. Bell & Sons.
Bybee, R. W. (1997). Towards an understanding of scientific literacy. In W. Graber & C. Bolte (Eds.),
Scientific literacy: An international symposium (pp. 37–68). Kiel: IPN, University of Kiel.
Bybee, R. W., Powell, J. C., Ellis, J. D., Giese, J. R., Parisi, L., & Singleton, L. (1991). Integrating the history
and nature of science and technology in science and social studies curriculum. Science Education, 75(1),
143–155.
Cain, S. E., & Evans, J. M. (1990). Sciencing: An involvement approach to elementary science methods.
Columbus, OH: Merrill.
Cajas, F. (2001). The science/technology interaction: Implications for science literacy. Journal of Research
in Science Teaching, 38(7), 715–729.
Callon, M. (1995). Four models for the dynamics of science. In S. Jasanoff, G. Markle, J. Peterson, & T.
Pinch (Eds.), Handbook of science and technology studies (pp. 29–63). Thousand Oaks, CA: Sage.
Callon, M., Law, J., & Rip, A. (Eds.). (1986). Mapping the dynamics of science and technology. London:
Macmillan.
Campanario, J. M. (2002). The parallelism between scientists’ and students’ resistance to new scientific
ideas. International Journal of Science Education, 24(10), 1095–1110.
Capra, F. (1983). The turning point: Science, society and the rising culture. New York: Bantam.
Cardwell, D. S. L. (1957). The organization of science in England. London: Heinemann.
Cardwell, D. S. L. (1972). Turning points in western technology: A study of technology, science and history.
New York: Science History Publishers.
Carey, S., Evans, E., Honda, M., Jay, E., & Unger, C. (1989). “An experiment is when you try it and see if it
works”: A study of grade 7 students’ understanding of the construction of scientific knowledge.
International Journal of Science Education, 11, 514–529.
Carey, S., & Smith, C. (1993). On understanding the nature of scientific knowledge. Educational
Psychologist, 28(3), 235–251.
Carnap, R. (1956). The methodological character of theoretical concepts. In H. Feigl & M. Scriven (Eds.),
Minnesota studies in philosophy of science (Vol. I, pp. 38–76). Minneapolis, MN: University of
Minnesota Press.
Carson, R. (1962). Silent spring. New York: Houghton Mifflin.
Carter, L. (2005). Globalisation and science education: Rethinking science education reforms. Journal of
Research in Science Teaching, 42(5), 561–580.
Cartwright, N. (1983). How the laws of physics lie. Oxford: Oxford University Press.
Castro, R. S., & Carvalho, M. P. (1995). The historic approach in teaching: Analysis of an experience.
Science & Education, 4(1), 65–85.
Cawthron, E. R., & Rowell, J. A. (1978). Epistemology and science education. Studies in Science Education,
5, 31–59.
Chalmers, A. F. (1978). What is this thing called science? Milton Keynes: Open University Press.

214

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
REFERENCES

Chalmers, A. F. (1999). What is this thing called science? (3rd ed.). Buckingham: Open University Press.
Chalmers, A. (1998). Retracing the ancient steps to atomic theory. Science & Education, 7, 69–84.
Chambers, D. W. (1983). Stereotypic images of the scientist: The draw-a-scientist test. Science Education,
67(2), 255–265.
Chambers, D. W. (1984a). Is seeing believing? Geelong: Deakin University Press.
Chambers, D. W. (1984b). Putting nature in order. Geelong: Deakin University Press.
Chambers, D. W. (1984c). Worm in the bud: Case study of the pesticide controversy. Geelong: Deakin
University Press.
Chamizo, J. A. (2007). Teaching modern chemistry through ‘recurrent historical teaching methods’. Science
& Education, 16, 197–216.
CHEM STUDY. (1963a). Chemistry – An experimental science. San Francisco: W.H. Freeman.
CHEM STUDY. (1963b). Teachers’ guide for chemistry – An experimental approach. San Francisco: W.H.
Freeman.
Chen, D., & Novick, R. (1984). Scientific and technological education in an information society. Science
Education, 68(4), 421–426.
Chiappetta, E. L. M., Sethna, G. H., & Dillman, D. A. (1991). A qualitative analysis of high school chemistry
textbooks for scientific literacy themes and expository learning aids. Journal of Research in Science
Teaching, 28, 939–951.
Chin, P., Munby, H., Hutchinson, N. L., Taylor, J., & Clark, F. (2004). Where’s the science? Understanding
the form and function of workplace science. In E. Scanlon, P. Murphy, J. Thomas, & E. Whitelegg
(Eds.), Reconsideraing science learning (pp. 118–134). London: RoutledgeFalmer.
Chinn, C. A., & Brewer, W. F. (1993). The role of anomalous data in knowledge acquisition: A theoretical
framework and implications for science instruction. Review of Educational Research, 63, 1–49.
Chinn, C. A., & Brewer, W. F. (1998). An empirical test of a taxonomy of responses to anomalous data in
science. Journal of Research in Science Teaching, 35, 623–654.
Christie, M. (1991). Aboriginal science for the ecologically sustainable future. Australian Science Teachers
Journal, 37, 26–31.
Churchland, P. (1979). Scientific realism and the plasticity of mind. Cambridge: Cambridge University Press.
Claxton, G. (1990). Teaching to learn: A direction for education. London: Cassell.
Claxton, G. (1991). Educating the inquiring mind: The challenge for school science. New York: Harvester
Wheatsheaf.
Claxton, G. (1997). Science of the times: A 2020 vision of education. In R. Levinson & J. Thomas (Eds.),
Science today: Problem or crisis? (pp. 71–86). London: Routledge.
Clay, J. (1996). Scientific literacy: Whose science? Whose literacy? In D. Baker, J. Clay, & C. Fox (Eds.),
Challenging ways of knowing: In english, mathematics and science (pp. 184–193). Lewes: Falmer Press.
Cleminson, A. (1990). Establishing an epistemological base for science teaching in the light of contemporary
notions of the nature of science and of how children learn science. Journal of Research in Science
Teaching, 27, 429–445.
Clough, M. W. (2006). Learners’ responses to the demands of conceptual change: Considerations for
effective nature of science instruction. Science Education, 15, 463–494.
Cobern, W. W., & Loving, C. C. (2001). Defining ‘science’ in a multicultural world: Implications for science
education. Science Education, 85, 50–67.
Cobern, W. W., & Loving, C. C. (2002). Investigation of preservice elementary teachers’ thinking about
science. Journal of Research in Science Teaching, 39(10), 1016–1031.
Coleman, E. B. (1998). Using explanatory knowledge during collaborative problem solving in science.
Journal of the Learning Sciences, 7(3 & 4), 387–428.
Coles, M., Gott, R., & Thornley, T. (1988). Active science. London: Collins.
Coll, R. K., & Taylor, I. (2005). The role of models and analogies in science education: Implications from
research. International Journal of Science Education, 27, 183–198.
Coll, R. K., & Treagust, D. F. (2002). Learners’ mental models of covalent bonding. Research in Science &
Technological Education, 20(2), 241–268.
Coll, R. K., & Treagust, D. F. (2003a). Learners’ mental models of metallic bonding: A cross-age study.
Science Education, 87(5), 685–707.
Coll, R. K., & Treagust, D. F. (2003b). Investigation of secondary school, undergraduate and graduate
learners’ mental models of ionic bonding. Journal of Research in Science Teaching, 40(5), 464–486.
Coll, R. K., France, B., & Taylor, I. (2005). The role of models/and analogies in science education:
Implications from research. International Journal of Science Education, 27(2), 183–198.

215

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
REFERENCES

Collins, H. (1975). ‘The seven sexes’: A study in the sociology of a phenomenon, or the replication of
experiments in physics. Sociology, 9, 206–224.
Collins, H. M. (1985). Changing order: Replication and induction in scientific practice. London: Sage.
Collins, H. M., & Pinch, T. J. (1993). The golem: What everyone should know about science. Cambridge:
Cambridge University Press.
Combe, G. (1825). A system of phrenology (2nd ed.). Edinburgh: John Anderson.
Combe, G. (1828). The constitution of man in relation to external objects. Edinburgh: MacLachlan &
Stewart.
Conant, J. B. (1951). On understanding science: An historical approach. New York: New American Library.
Conant, J. B. (Ed.). (1957). Harvard case histories in experimental science (2 Vols.). Cambridge, MA:
Harvard University Press.
Connelly, F. M., Finegold, M., Clipsham, J., & Wahlstrom, M. (1977). Science inquiry and the teaching of
science: Patterns of inquiry project. Toronto: OISE Press.
Cooley, W. W., & Klopfer, L. E. (1961). Manual for the test on understanding science. Princeton, NJ:
Educational Testing Service.
Cornish, E. (1977). The study of the future: An introduction to the art and science of understanding and
shaping tomorrow’s world. Bethesda, MD: World Futures Society.
Cotham, J., & Smith, E. (1981). Development and validation of the conceptions of scientific theories test.
Journal of Research in Science Teaching, 18, 387–396.
Council of Ministers of Education, Canada. (1997). Common framework of science learning outcomes.
Toronto: CMEC Secretariat.
Cozzens, S. E., & Gieryn, T. F. (Eds.). (1990). Theories of science in society. Bloomington, IN: Indiana
University Press.
Crawford, E. (1993). A critique of curriculum reform: Using history to develop thinking. Physics Education,
28, 204–208.
Cromer, A. (1993). Uncommon sense: The heretical nature of science. Oxford: Oxford University Press.
Cross, R. (1995). Conceptions of scientific literacy: Reactionaries in ascendancy in the state of Victoria.
Research in Science Education, 25(2), 151–162.
Daugs, D. R. (1970). Scientific literacy – Re-examined. The Science Teacher, 37(8), 10–11.
Davies, I. (2004). Science and citizenship education. International Journal of Science Education, 26(14),
1751–1763.
Davies, T., & Gilbert, J. (2003). Modelling: Promoting creativity while forging links between science
education and design and technology education. Canadian Journal of Science, Mathematics and
Technology Education, 3(1), 67–82.
Davis, B. (2000). Skills mania: Snake oil in our schools? Toronto: Between the Lines Press.
Dawkins, R. (1995). River out of eden. New York: Basic Books.
Dawkins, R. (1998). Unweaving the rainbow: Science, delusion, and the appetite for wonder. New York:
Houghton Mifflin.
Dearden, R. F. (1967). Instruction and learning by discovery. In R. S. Peters (Ed.), The concept of education
(pp. 135–155). London: Routledge & Kegan Paul.
Dearden, R. F. (1968). The philosophy of primary education. London: Routledge & Kegan Paul.
de Berg, K. C. (1997). The development of the concept of work: A case where history can inform pedagogy.
Science & Education, 6, 511–527.
De Boer, G. E. (2000). Scientific literacy: Another look at its historical and contemporary meanings and its
relationship to science education reform. Journal of Research in Science Teaching, 37(6), 582–601.
DeBoer, G. (2001). Scientific literacy: Another look at its historical and contemporary meanings and its
relationship to science education reform. Journal of Research in Science Teaching, 37(6), 582–601.
Dennick, R. (1992a). Analysing multicultural and antiracist science education. School Science Review,
73(264), 79–88.
Dennick, R. (1992b). Multicultural and antiracist science education: Theory and practice. Nottingham:
School of Education, University of Nottingham.
Department of Education & Science/Welsh Office. (1989). Science in the national curriculum. London:
HMSO.
Dickson, D. (1988). The new politics of science. Chicago, IL: University of Chicago Press.
Dillashaw, F. G., & Okey, J. R. (1980). Test of integrated process skills for secondary school science
students. Science Education, 64(5), 601–608.

216

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
REFERENCES

Dillon, J. (2005). Silent spring: Science, the environment and society. School Science Review, 86(316),
113–118.
Dimopoulos, K., & Koulaides, V. (2003). Science and technology education for citizenship: The potential
role of the press. Science Education, 87(2), 241–256.
Dirac, P. A. M. (1963). The evolution of the physicist’s picture of nature. Scientific American, 208(5), 45–63.
Dirac, P. A. M. (1980). The excellence of Einstein’s theory of gravitation. In M. Goldsmith, A. Mackay, &
J. Woudhuysen (Eds.), Einstein: The first hundred years (pp. 41–46). Oxford: Pergamon Press.
Dobbin, M. (1998). The myth of the good corporate citizen: Democracy under the rule of big business.
Toronto: Stoddart.
Dreyer, J. L. E. (1953). A history of astronomy from Thales to Kepler. New York: Dover.
Driver, R., & Easley, J. (1978). Pupils and paradigms: A review of literature related to concept development
in adolescent science students. Studies in Science Education, 5, 61–84.
Driver, R., Asoko, H., Leach, J., & Scott, P. (1994). Constructing scientific knowledge in the classroom.
Educational Researcher, 23, 5–12.
Driver, R., Leach, J., Millar, R., & Scott, P. (1996). Young people’s images of science. Buckingham: Open
University Press.
Duggan, S., & Gott, R. (2002). What sort of science education do we really need? International Journal of
Science Education, 24, 661–679.
Duhem, P. M. M. (1962). The aim and structure of physical theory (P. P. Wiener, Trans.). London:
Atheneum Press.
Dunn, R. (1993). Empires of physics – A new initiative in science education. School Science Review, 75,
135–137.
Dunwoody, S. (1993). Reconstructing science for public consumption: Journalism as science education.
Geelong: Deakin University Press.
Duschl, R. A. (1990). Restructuring Science education: The importance of theories and their development.
New York: Teachers College Press.
Duschl, R. A., & Erduran, S. (1996). Modelling the growth of scientific knowledge. In G. Welford, J.
Osborne, & P. Scott (Eds.), Research in science education in Europe – Current issues and themes
(pp. 153–165). London: Falmer Press.
Duschl, R. A., & Osborne, J. (2002). Supporting and promoting argumentation discourse in science
education. Studies in Science Education, 38, 39–72.
Duschl, R. A., Hamilton, R. J., & Grandy, R. E. (1992). Psychology and epistemology: Match or mismatch
when applied to science education? In R. A. Duschl & R. J. Hamilton (Eds.), Philosophy of science,
cognitive psychology and educational theory and practice (pp. 19–47). Albany, NY: State University of
New York Press.
Duveen, J., Scott, L., & Solomon, J. (1993). Pupils’ understanding of science: Description of experiments or
“a passion to explain”? School Science Review, 75(271), 19–27.
Eddington, A. (1928). The nature of the physical world. Cambridge: Cambridge University Press.
Eflin, J. T., Glennan, S., & Reisch, G. (1999). The nature of science: A perspective from the philosophy of
science. Journal of Research in Science Teaching, 36(1), 107–116.
Einstein, A. (1918). Principles of research (address in celebration of Max Plank’s 60th birthday). In A.
Einstein (Ed.), (1954) Ideas and opinions (pp. 220–226). New York: Dell Publishing.
Einstein, A. (1933). On the method of theoretical physics. Herbert Spencer lecture, Oxford, June 10.
Reprinted in C. Seelig (Ed.), Ideas and opinions of Albert Einstein (S. Bargmann, Trans., pp. 263–270).
New York: Wings Books.
Eisenhart, M., Finkel, E., & Marion, S. F. (1996). Creating the conditions for scientific literacy: A re-
examination. American Educational Research Journal, 33, 261–295.
Elby, A., & Hammer, D. (2001). On the substance of a sophisticated epistemology. Science Education, 85(5),
554–567.
Elliott, P. (2006). Reviewing newspaper articles as a technique for enhancing the scientific literacy of
student-teachers. International Journal of Science Education, 28(11), 1245–1265.
Ellis, B. (1985). What science aims to do. In P. M. Churchland & C. Hooker (Eds.), Images of science
(pp. 48–74). Chicago: University of Chicago Press.
Elshof, L. (2001). Worldview research with technology teachers. Toronto: Unpublished PhD thesis,
University of Toronto.
Epstein, I. R. (1987, March 30). Patterns in time and space: Created by chemistry. Chemical & Engineering
News, 24–36.

217

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
REFERENCES

Etzkowitz, H., Kemelgor, C., & Uzzi, B. (Eds.). (2000). Athena unbound: The advancement of women in
science and technology. Cambridge: Cambridge University Press.
Faire, J., & Cosgrove, M. (1988). Teaching primary science. Hamilton, New Zealand: University of Waikato
Education Centre.
Fang, Z. (2005). Scientific literacy: A systematic functional linguistics perspective. Science Education, 89(2),
335–347.
Fee, E. (1979). Nineteenth century craniology: The study of the female skull. Bulletin of the History of
Medicine, 53, 415–433.
Fensham, P. J. (2002). Time to change drivers for scientific literacy. Canadian Journal of Science,
Mathematics and Technology Education, 2(1), 9–24.
Fensham, P. J., Gunstone, R. F., & White, R. T. (Eds.). (1994). The content of science: A constructivist
approach to its teaching and learning. London: Falmer Press.
Feyerabend, P. K. (1962). Explanation, reduction and empiricism. Minnesota Studies in the Philosophy of
Science, 3, 28–97.
Feyerabend, P. K. (1964). Realism and instrumentalism: Comments on the logic of factual support. In M.
Bunge (Ed.), The critical approach to science and philosophy (pp. 281–308). New York: Free Press.
Feyerabend, P. K. (1970). Consolations for the specialist. In I. Lakatos & A. Musgrave (Eds.), Criticism and
the growth of knowledge (pp. 195–230). Cambridge: Cambridge University Press.
Feyerabend, P. K. (1975). Against method: Outline of an anarchistic theory of knowledge. London: New Left
Books.
Feyerabend, P. K. (1978). Science in a free society. London: New Left Books.
Feynman, R. (1965). The character of physical law. Cambridge, MA: MIT Press.
Fine, A. (1986). Unnatural attitudes: Realist and instrumentalist attachments to science. Mind, 95(378), 149–
179.
Fine, A. (1991). The natural ontological attitude. In R. Boyd, P. Gasper, & J. D. Trout (Eds.), The philsophy
of science (pp. 261–277). Cambridge, MA: MIT Press.
Finley, F. N., & Pocovi, M. C. (2000). Considering the scientific method of inquiry. In J. Minstrell & E. H.
van Zee (Eds.), Inquiry into inquiry learning and teaching in science (pp. 47–62). Washington,
DC: AAAS.
Finson, K. D. (2002). Drawing a scientist: What we do and do not know after fifty years of drawing. School
Science & Mathematics, 102(7), 335–345.
Finson, K. D., Beaver, J. B., & Cramond, B. L. (1995). Development and field test of a checklist for the
draw-a-scientist test. School Science & Mathematics, 95(4), 195–205.
Fitzpatrick, F. L. (Ed.). (1960). Policies for science education. New York: Teachers College, Columbia
University.
Fort, D. C., & Varney, H. L. (1989). How students see scientists: Mostly male, mostly white, and mostly
benevolent. Science and Children, 26, 8–13.
Forman, P. (1971). Weimar culture, causality, and quantum theory, 1918–1927: Adaptation by German
physicists and mathematicians to a hostile intellectual environment. In R. McCormmach (Ed.), Historical
studies in the physical sciences (Vol. 3, pp. 1–115). Philadelphia: University of Pennsylvania Press.
Foucault, M. (1970). The order of things: An archaeology of the human sciences (A. Sheridan, Trans.). New
York: Pantheon.
Foucault, M. (1982). The archaeology of knowledge (A. Sheridan, Trans.). New York: Pantheon. (First
published by Tavistock: London)
Fourez, G. (1997). Scientific and technological literacy as a social practice. Social Studies of Science, 27(6),
903–936.
Franco, C., Barros, H. L., Colinvaux, D., Krapas, S., Queiroz, G., & Alves, F. (1999). From scientists’ and
inventors’ minds to some scientific and technological products: Relationships between theories, models,
mental models and conceptions. International Journal of Science Education, 21(3), 277–291.
Frankel, H. (1979). The career of continental drift theory. Studies in History and Philosophy of Science,
10, 21–66.
Fujimura, J. H. (1992). Crafting science: Standardized packages, boundary objects, and ‘translation’. In A.
Pickering (Ed.), Science as practice and culture (pp. 168–211). Chicago: University of Chicago Press.
Fuller, S. (1992). Social epistemology and the research agenda of science studies. In A. Pickering (Ed.),
Science as practice and culture (pp. 390–428). Chicago: University of Chicago Press.
Gabel, L. L. (1976). The development of a model to determine perceptions of scientific literacy. Unpublished
PhD thesis, Columbus, OH: Ohio State University.

218

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
REFERENCES

Gagne, R. M. (1963). The learning requirements for inquiry. Journal of Research in Science Teaching,
1(2), 144–153.
Gagne, R. M. (1965). The psychological bases of science – A process approach. Washington, DC: American
Association for the Advancement of Science.
Galbraith, P. L., Carss, M. C., Grice, R. D., Endean, L., & Warry, M. (1997). Towards scientific literacy for
the third millennium: A view from Australia. International Journal of Science Education, 19, 447–467.
Galili, I., & Hazan, A. (2001a). Experts’ views on using history and philosophy of science in the practice of
physics instruction. Science & Education, 10, 345–367.
Galili, I., & Hazan, A. (2001b). The effect of a history-based course in optics on students’ views about
science. Science & Education, 10, 7–32.
Gallagher, R., & Ingham, P. (1984). Chemistry made clear. Oxford: Oxford University Press.
Gamarnikow, E., & Green, A. (2000). Citizenship education and social capital. In D. Lawton, J. Cairns, & R.
Gardner (Eds.), Education for citizenship. London: Continuum.
Gardner, P. L. (1998). The development of males’ and females’ interest in science and technology. In L.
Hoffmann, A. Krapp, K. A. Renninger, & J. Baumert (Eds.), Interest and learning (pp. 41–57). Kiel:
IPN, University of Kiel.
Garfinkel, H. (1967). Studies in ethnomethodology. Englewood Cliffs, NJ: Prentice Hall.
Garrison, J. (2002). Questioning the cultural function of science education: An endorsement and response to
Rudolph. Science Education, 87(1), 80–89.
Garrison, J. W., & Lawwill, K. S. (1992). Scientific literacy: For whose benefit? In S. Hills (Ed.),
Proceedings of the second international conference on the history and philosophy of science and science
education (Vol. 1, pp. 337–349). Kingston: Queen’s University.
Gaskell, J., & Willinsky, J. (Eds.). (1995). Gender in/forms curriculum. New York: Teachers College Press.
Gaskell, J. P. (1992). Authentic science and school science. International Journal of Science Education,
14, 265–272.
Gauld, C. (1991). History of science, individual development and science teaching. Research in Science
Education, 21, 133–140.
Gauld, C. (1998). Making more plausible what is hard to believe: Historical justifications and llustrations of
Newton’s third law. Science & Education, 7, 159–172.
Geddis, A. (1991). Improving the quality of science classroom discourse on controversial issues. Science
Education, 75, 169–183.
Gee, J. P., Hull, G., & Lankshear, C. (1996). The new work order: Behind the language of the new
capitalism. Boulder, CO: Allen & Unwin/Westview Press.
Gentner, D., & Stevens, A. (1983). Mental models. Hillsdale, NJ: Lawrence Erlbaum.
Giere, R. N. (1988). Explaining science: A cognitive approach. Chicago: University of Chicago Press.
Giere, R. N. (1999). Science without laws. Chicago: University of Chicago Press.
Gieryn, T. F. (1995). Boundaries of science. In S. Jasanoff, G. Markle, J. Peterson, & T. Pinch (Eds.),
Handbook of science and technology studies (pp. 393–443). Thousand Oaks, CA: Sage.
Gil, D., & Solbes, J. (1993). The introduction of modern physics: Overcoming a deformed vision of science.
International Journal of Science Education, 15(3), 255–260.
Gilbert, J. K. (2004). Models and modelling: Routes to more authentic science education. International
Journal of Science and Mathematics Education, 2(1), 115–130.
Gilbert, J. K., Boulter, C., & Rutherford, M. (1998a). Models in explanations, Part 1: Horses for courses?
International Journal of Science Education, 20(1), 83–97.
Gilbert, J. K., Boulter, C., & Rutherford, M. (1998b). Models in explanations, Part 2: Whose voice? Whose
ears? International Journal of Science Education, 20(2), 187–203.
Gilbert, J., & Boulter, C. (2000). Developing models in science education. Dordrecht: Kluwer.
Gilbert, N., & Mulkay, N. (1984). Opening Pandora’s box. Cambridge: Cambridge University Press.
Gillborn, D. (1997). Racism and reform. British Educational Research Journal, 23, 345–360.
Gil-Perez, D., Guiasola, J., Moreno, A., Cachapuz, A., Pessoa de Carvalho, A. M., Martinez Torregrosa, J.,
et al. (2002). Defending constructivism in science education. Science & Education, 11(6), 557–571.
Gillies, D. (1993). Philosophy of science in the twentieth century: Four central themes. Oxford: Blackwell.
Gitari, W. (2003). An inquiry into the integration of indigenous knowledges and skills in the Kenyan
secondary science curriculum: A case study of human health knowledge. Canadian Journal of Science,
Mathematics and Technology Education, 3(2), 195–212.
Gitari, W. (2006). Everyday objects of learning about health and healing and implications for science
education. Journal of Research in Science Teaching, 43(2), 172–193.
Glaser, R. (1966). Variables in discovery learning. In L. S. Shulman & E. R. Keislar (Eds.), Learning by
discovery: A critical appraisal (pp. 13–26). Chicago: Rand McNally.
219

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
REFERENCES

Good, R., & Shymansky, J. (2001). Nature-of-science literacy in Benchmarks and Standards: Post-
modern/relativist or modern/realist? Science & Education, 10, 173–185.
Gough, N. (1993). Laboratories in fiction: Science education and popular media. Geelong: Deakin
University Press.
Gould, S. J. (1981). The mismeasure of man. Harmondsworth: Penguin.
Government of Canada. (1991). Prosperity through competitiveness. Ottawa: Ministry of Supply and
Services Canada.
Graber, W., & Bolte, C. (Eds.). (1997). Scientific literacy: An international symposium. Kiel: Institut fur die
Padagogik der Naturwiseenschaften (IPN) an der Universitat Kiel.
Greca, I. M., & Moreira, M. A. (2000). Mental models, conceptual models and modelling. International
Journal of Science Education, 22(1), 1–11.
Greca, I. M., & Moreira, M. A. (2002). Mental, physical and mathematical models in the teaching and
learning of physics. Science Education, 86(1), 106–121.
Gregory, R. L. (1977). Eye and brain: The psychology of seeing. London: Weidenfeld & Nicolson.
Griffiths, A. K., & Barman, C. R. (1995). High school students’ views about the nature of science: Results
from three countries. School Science & Mathematics, 95(2), 248–255.
Habermas, J. (1971). Knowledge and human interests. Boston: Beacon Press.
Hacking, I. (1983). Representing and intervening. Cambridge: Cambridge University Press.
Hacking, I. (1991). Experimentation and realism. In R. Boyd, P. Gasper, & J. D. Trout (Eds.), The
philosophy of science (pp. 247–260). Cambridge, MA: MIT Press.
Hacking, I. (1992). The self-vindication of the laboratory sciences. In A. Pickering (Ed.), Science as practice
and culture (pp. 29–64). Chicago: University of Chicago Press.
Hagen, J. B., & Kugler, C. (1992). Using history to teach principles of biology to college students: The case
of cell theory. In G. L. C. Hills (Ed.), Proceedings of the second international conference on history and
philosophy of science and science teaching (Vol. 1, pp. 471–481). Kingston, ON: Queen’s University.
Hagstrom, W. O. (1965). The scientific community. London: Basic Books.
Hall, A. R. (1967). The significance of Galileo’s thought for the history of science. In E. McMullin (Ed.),
Galileo: Man of science (pp. 67–81). New York: Basic Books.
Hanson, N. R. (1958). Patterns of discovery. Cambridge: Cambridge University Press.
Hanson, N. R. (1972). Observation and explanation. London: Allen & Unwin.
Harding, P., & Hare, W. (2000). Portraying science accurately in classrooms: Emphasizing open-mindedness
rather than relativism. Journal of Research in Science Teaching, 37(3), 225–236.
Harding, S. (1998). Is science multicultural? Bloomington, IN: Indiana University Press.
Harre, R. (1986). Varieties of reason. Oxford: Blackwell.
Harris, D., & Taylor, M. (1983). Discovery learning in school science: The myth and the reality. Journal of
Curriculum Studies, 15(3), 277–289.
Hashweh, M. Z. (1986). Towards an explanation of conceptual change. European Journal of Science
Education, 8(3), 229–249.
Hazari, Z. (2006). Gender differences in introductory university physics: The influence of high school physics
preparation and affect. Unpublished PhD thesis, University of Toronto.
Heering, P. (2000). Getting shocks: Teaching secondary school physics through history. Science &
Education, 9, 363–373.
Heidelberger, M., & Stadler, F. (Eds.). (2002). History of philosophy of science: New trends and
perspectives. Dordrecht: Kluwer.
Heilbron, J. L. (1987). Applied history of science. Isis, 78(4), 552–563.
Heisenberg, W. (1971). Physics and beyond: Encounters and conversations. New York: Harper & Row.
Hempel, C. G. (1958). The theoretician’s dilemma. In H. Feigl, M. Scriven, & G. Maxwell (Eds.), Minnesota
studies in philosophy of science (Vol. II, pp. 37–98). Minneapolis, MN: University of Minnesota Press.
Hempel, C. G. (1966). Philosophy of natural science. Englewood Cliffs, NJ: Prentice Hall.
Hennessy, S. (1993). Situated cognition and cognitive apprenticeship: Implications for classroom learning.
Studies in Science Education, 22, 1–41.
Hesse, M. (1980). Revolutions and reconstructions. Brighton: Harvester Press.
Hewitt, J., Pedretti, E., Bencze, L., Vaillancourt, B. D., & Yoon, S. (2003). New applications for multimedia
cases: Promoting reflective practice in preservice teacher education. Journal of Technology and Teacher
Education, 11(4), 483–500.
Hewson, P., Beeth, M. E., & Thorley, N. R. (1998). Teaching for conceptual change. In B. Fraser & K.
Tobin (Eds.), International handbook of science education (pp. 199–218). Dordrecht: Kluwer.

220

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
REFERENCES

Hill, D. R. (1993). Islamic science and engineering. Edinburgh: Edinburgh University Press.
Hoadley, C. M., & Linn, M. C. (2000). Teaching science through online, peer discussions: SpeakEasy in the
knowledge integration environment. International Journal of Science Education, 22(8), 839–857.
Hobson, J. M. (2004). The eastern origin of western civilization. Cambridge: Cambridge University Press.
Hodson, D. (1985). Philosophy of science, science and science education. Studies in Science Education, 12,
25–57.
Hodson, D. (1988). Experiments in science and science teaching. Educational Philosophy & Theory,
20, 53–66.
Hodson, D. (1990). Making the implicit explicit: A curriculum planning model for enhancing children’s
understanding of science. In D. E. Herget (Ed.), More history and philosophy of science in science
teaching (pp. 292–310). Tallahassee, FL: Florida State University Press.
Hodson, D. (1992a). In search of a meaningful relationship: An exploration of some issues relating to
integration in science and science education. International Journal of Science Education, 14, 541–562.
Hodson, D. (1992b). Assessment of practical work: Some considerations in philosophy of science. Science &
Education, 1, 115–144.
Hodson, D. (1993a). Philosophic stance of secondary school science teachers, curriculum experiences, and
children’s understanding of science: Some preliminary findings. Interchange, 24(1&2), 41–52.
Hodson, D. (1993b). Against skills-based testing in science. Curriculum Studies, 1(1), 127–148.
Hodson, D. (1993c). Re-thinking old ways: Towards a more critical approach to practical work in school
science. Studies in Science Education, 22, 85–142.
Hodson, D. (1993d). In search of a rationale for multicultural science education. Science Education, 77(6),
685–711.
Hodson, D. (1994). Seeking directions for change: The personalisation and politicisation of science
education. Curriculum Studies, 2, 71–98.
Hodson, D. (1996). Laboratory work as scientific method: Three decades of confusion and distortion.
Journal of Curriculum Studies, 28(2), 115–135.
Hodson, D. (1998a). Teaching and learning science: Towards a personalized approach. Buckingham: Open
University Press.
Hodson, D. (1998b). Science fiction: The continuing misrepresentation of science in the school curriculum.
Curriculum Studies, 6(2), 191–216.
Hodson, D. (1999). Going beyond cultural pluralism: Science education for sociopolitical action. Science
Education, 83(6), 775–796.
Hodson, D. (2003). Time for action: Science education for an alternative future. International Journal of
Science Education, 25(6), 645–670.
Hodson, D. (2006). Why we should prioritize learning about science. Canadian Journal of Science,
Mathematics and Technology Education, 6(3), 293–311.
Hodson, D., & Dennick, R. (1994). Antiracist education: A special role for the history of science and
technology. School Science & Mathematics, 94(5), 255–263.
Hodson, D., & Hodson, J. (1998a). From constructivisn to social constructivism: A Vygotskian perspective
on teaching and learning science. School Science Review, 78(289), 33–41.
Hodson, D., & Hodson, J. (1998b). Science education as enculturation: Some implications for practice.
School Science Review, 80(290), 17–24.
Hodson, D., & Prophet, R. B. (1986). A bumpy start to science education. New Scientist, 1521, 25–28.
Hodson, D., & Reid, D. J. (1988). Science for all: Motives, meanings and implications. School Science
Review, 69, 653–661.
Hogan, K. (2000). Exploring a process view of students’ knowledge about the nature of science. Science
Education, 84(1), 51–70.
Hogan, K., & Maglienti, M. (2001). Comparing the epistemological underpinnings of students’ and
scientists’ reasoning about conclusions. Journal of Research in Science Teaching, 38(6), 663–687.
Holliday, W., Yore, L., & Alvermann, D. (1994). The reading-science learning-writing connection:
Breakthroughs, barriers and promises. Journal of Research in Science Teaching, 31, 877–893.
Holton, G. (1975). On the role of themata in scientific thought. Science, 188(4186), 328–338.
Holton, G. (1978). The scientific imagination: Case studies. Cambridge: Cambridge University Press.
Holton, G. (1981). Thematic presuppositions and the direction of science advance. In A. F. Heath (Ed.),
Scientific explanation (pp. 1–27). Oxford: Oxford University Press.
Holton, G. (1986). The advancement of science and its burdens: The Jefferson lectures and other essays.
Cambridge: Cambridge University Press.

221

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
REFERENCES

Holton, G. (1988). Thematic origins of scientific thought: Kepler to Einstein. Cambridge, MA: Harvard
University Press.
Holton, G. (1992). How to think about the ‘anti-science’ phenomenon. Public Understanding of Science,
1, 103–128.
Hottecke, D. (2000). How and what can we learn from replicating historical experiments? A case study.
Science & Education, 9, 343–362.
Howe, E. M. (2007). Untangling sickle-cell anemia and the teaching of heterozygote protection. Science &
Education, 16(1), 1–19.
Huber, R. A., & Burton, G. M. (1995). What do students think scientists look like? School Science &
Mathematics, 95(7), 371–376.
Hull, D. L. (1988). Science as a process: An evolutionary account of the social and conceptual development
of science. Chicago: University of Chicago Press.
Hume, D. (1854). Of the understanding. In D. Hume (Ed.), Philosophical works (Vol. 1). Edinburgh: Adam
and Charles Black.
Hurd, P. D. (1958). Science literacy: Its meaning for American schools. Educational Leadership,
16(1), 13–16.
Hurd, P. D. (1998). Scientific literacy: New minds for a changing world. Science Education, 82(3), 407–416.
Hurd, P. D. (2002). Modernizing science education. Journal of Research in Science Teaching, 39(1), 3–9.
Inner London Education Authority (ILEA). (1987). Science in process. London: Heinemann.
Jackson, T. (1992). Perceptions of scientists among elementary school children. Australian Science Teachers
Journal, 38, 57–61.
Jacoby, B., & Spargo, P. (1992). An appropriate science education for Africa: Deweyan instrumentalism or
critical realism? In G. L. C. Hills (Ed.), History and philosophy of science in science education (Vol. 1,
pp. 545–557). Kingston, ON: Queen’s University Press.
Jenkins, E. W. (1979). From Armstrong to Nuffield: Studies in twentieth-century science education in
England and Wales. London: John Murray.
Jenkins, E. (1989). Why the history of science? In M. Shortland & A. Warwick (Eds.), Teaching the history
of science (pp. 19–29). Oxford: Basil Blackwell.
Jenkins, E. (1990). Scientific literacy and school science education. School Science Review, 71(256), 43–51.
Jenkins, E. W. (1994). Scientific literacy. In T. Husen & T. N. Postlethwaite (Eds.), The international
encyclopaedia of education (2nd ed., Vol. 9, pp. 5345–5350). Oxford: Pergamon Press.
Jenkins, E. W. (1994b). HPS and school science education: Remediation or reconstruction? International
Journal of Science Education, 16(6), 613–623.
Jenkins, E. W. (1994c). Public understanding and science education for action. Journal of Curriculum
Studies, 26, 601–611.
Jenkins, E. W. (1996a). The ‘nature of science’ as a curriculum component. Journal of Curriculum Studies,
28(2), 137–150.
Jenkins, E. W. (1996b). Scientific literacy: A functional construct. In D. Baker, J. Clay, & C. Fox (Eds.),
Challenging ways of knowing: In english, mathematics and science (pp. 43–51). Lewes: Falmer Press.
Jenkins, E. W. (1997). Scientific and technological literacy: Meanings and rationales. In E. W. Jenkins (Ed.),
Innovations in science and technology education. Volume VI: Rationality, meanings and measures of
technological literacy (pp. 11–40). Paris: UNESCO.
Jenkins, E. W. (1999). School science, citizenship and the public understanding of science. International
Journal of Science Education, 21(7), 703–710.
Jenkins, E. W. (2000). Constructivism in school science education: Powerful model or the most dangerous
intellectual tendency? Science & Education, 9, 599–610.
Johnson-Laird, P. N. (1983). Mental models. Cambridge: Cambridge University Press.
Joravsky, D. (1970). The Lysenko affair. Chicago: University of Chicago Press.
Judson, H. F. (2004). The great betrayal: Fraud in science. Orlando, FL: Harcourt.
Justi, R., & Gilbert, J. K. (1999). History and philosophy of science through models: The case of kinetics.
Science & Education, 8(3), 287–307.
Justi, R. S., & Gilbert, J. K. (2002a). Modelling, teachers’ views on the nature of modelling and implications
for the education of modellers. International Journal of Science Education, 24(4), 369–388.
Justi, R. S., & Gilbert, J. K. (2002b). Science teachers’ knowledge about and attitudes towards the use of
models in learning science. International Journal of Science Education, 24(12), 1273–1292.

222

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
REFERENCES

Justi, R. S., & Gilbert, J. K. (2002c). Models and modelling in chemical education. In J. K. Gilbert, O. de
Jong, R. Justi, D. F. Treagust, & J. H. van Driel (Eds.), Chemical education: Towards research-based
practice (pp. 47–68). Dordrecht: Kluwer.
Justi, R. S., & Gilbert, J. K. (2003). Teachers’ views on the nature of models. International Journal of
Science Education, 25(11), 1369–1386.
Justi, R., & van Driel, J. (2005). The development of science teachers’ knowledge on models and modelling:
Promoting, charaterizing, and understanding the process. International Journal of Science Education,
27(5), 549–573.
Kafai, Y. B., & Gilliland-Swetland, A. J. (2001). The use of historical materials in elementary science
classrooms. Science Education, 85(4), 349–367.
Kant, I. (1929). The critique of pure reason (N. K. Smith, Trans.). London: Macmillan.
Kauffman, G. B. (1989). History on the chemistry curriculum. Interchange, 20(2), 81–94.
Kawasaki, K., Herrenkohl, L. P., & Yeary, S. A. (2004). Theory building and modelling in a sinking and
floating unit: A case study of third and fourth grade students’ developing epistemologies of science.
International Journal of Science Education, 26(11), 1299–1324.
Keller, E. F., & Longino, H. E. (Eds.). (1996). Feminism and science. Oxford: Oxford University Press.
Kelly, G. J. (1997). Research traditions in comparative context: A philosophical challenge to radical
constructivism. Science Education, 81, 355–375.
Kenealey, P. (1989). Telling a coherent ‘story’: A role for the history and philosophy of science in a physical
science course. In D. E. Herget (Ed.), History and philosophy of science in science teaching
(pp. 209–220). Tallahasssee, FL: Florida State University.
Keys, C. W. (1999). Revitalizing instruction in scientific genres: Connecting knowledge production with
writing to learn in science. Science Education, 83, 115–130.
Kimball, M. E. (1967). Understanding the nature of science: A comparison of scientists and teachers. Journal
of Research in Science Teaching, 5(2), 110–120.
King, B. B. (1991). Beginning teachers’ knowledge of and attitudes towards history and philosophy of
science. Science Education, 75(1), 135–141.
Kipnis, N. (1996). The ‘historical-investigative’ approach to teaching science. Science & Education,
5, 277–292.
Kirschner, P. A. (1992). Epistemology, practical work and academic skills in science education. Science &
Education, 1(3), 273–299.
Klein, M. J. (1972). The use and abuse of historical teaching in physics. In S. G. Brush & A. L. King (Eds.),
History in the teaching of physics (pp. 12–18). Hanover, NH: University of New England Press.
Klopfer, L. E. (1969). Science education in 1991. School Review, 77(3–4), 199–217.
Klopfer, L. E. (1976). Scientific literacy reexamined. Science Education, 60(1), 95.
Klotz, I. M. (1980). The N-ray affair. Scientific American, 242(5), 122–131.
Knain, E. (2001). Ideologies in school science textbooks. International Journal of Science Education, 23(3),
319–329.
Knorr-Cetina, K. D. (1981). The manufacture of knowledge: An essay on the constructivist and contextual
nature of science. Oxford: Pergamon Press.
Knorr-Cetina, K. D. (1983). The ethnographic study of work: Towards a constructivist interpretation of
science. In K. D. Knorr-Cetina & M. Mulkay (Eds.), Science observed: Perspectives on the social study
of science (pp. 115–140). London: Sage.
Knorr-Cetina, K. (1992). The couch, the cathedral, and the laboratory: On the relationship between experi-
ment and laboratory in science. In A. Pickering (Ed.), Science as practice and culture (pp. 113–137).
Chicago: University of Chicago Press.
Knorr-Cetina, K. (1995). Laboratory studies: The cultural approach to the study of science. In S. Jasanoff, G.
Markle, J. Peterson, & T. Pinch (Eds.), Handbook of science and technology studies (pp. 140–166).
Thousand Oaks, CA: Sage.
Knorr-Cetina, K. D., & Mulkay, M. (1983). Introduction: Emerging principles in social studies of science. In
K. D. Knorr-Cetina & M. Mulkay (Eds.), Science observed: Perspectives on the social study of science
(pp. 1–17). London: Sage.
Kohn, A. (1986). False prophets. Oxford: Bail Blackwell.
Kolsto, S. D. (2000). Consensus projects: Teaching science for citizenship. International Journal of Science
Education, 22(6), 645–664.
Kolsto, S. D. (2001). Scientific literacy for citizenship: Tools for dealing with the science dimension of
controversial socioscientific issues. Science Education, 85(3), 291–310.

223

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
REFERENCES

Koulaidis, V., & Ogborn, J. (1995). Science teachers’ philosophical assumptions: How well do we
understand them? International Journal of Science Education, 17(3), 273–283.
Kragh, H. (1987). An introduction to the historiography of science. Cambridge: Cambridge University Press.
Kragh, H. (1992). A sense of history: History of science and the teaching of introductory quantum theory.
Science & Education, 1, 349–363.
Kreuzman, H. (1995). Stepping into science: Historical experiments and scientific method. In F. Finley, D.
Allchin, D. Rhees, & S. Fifield (Eds.), Third international history, philosophy and science teaching
conference. Proceedings Volume 1 (pp. 636–647). Minneapolis, MN: University of Minnesota.
Krugly-Smolska, E. T. (1990). Scientific literacy in developed and developing countries. International
Journal of Science Education, 12(5), 473–480.
Kuhn, T. S. (1962). The structure of scientific revolutions. Chicago: University of Chicago Press.
Kuhn, T. S. (1963). The essential tension: Tradition and innovation in scientific research. In C. W. Taylor &
F. Barron (Eds.), Scientific creativity: Its recognition and development (pp. 341–354). New York: John
Wiley.
Kuhn, T. S. (1970). The structure of scientific revolutions (2nd ed.). Chicago: University of Chicago Press.
Kuhn, T. S. (1977). The essential tension: Selected studies in scientific traditions and change. Chicago:
University of Chicago Press.
Kumar, A., & Kenealy, P. (1992). An issue of cultural diversity in science. In S. Hills (Ed.), The history and
philosophy of science in science education (Vol. II). Kingston: Queen’s University.
Ladyman, J. (2002). Understanding philosophy of science. London: Routledge.
Lakatos, I. (1968). Criticism and the methodology of scientific research programmes. Proceedings of the
Aristotelian Society, 69, 149–186.
Lakatos, I. (1974). Falsification and the methodology of scientific research programmes. In I. Lakatos &
A. Musgrave (Eds.), Criticism and the growth of knowledge (pp. 191–196). Cambridge: Cambridge
University Press.
Lakatos, I. (1978). The methodology of scientific research programmes (J. Worrall & G. Currie, Eds.).
Cambridge: Cambridge University Press.
Lakin, S., & Wellington, J. (1994). Who will teach the ‘nature of science’? Teachers’ views of science and
their implications for science education. International Journal of Science Education, 16(2), 175–190.
Langer, E., Hatem, M., Joss, J., & Howell, M. (1989). Conditional teaching and mindful learning: The role of
uncertainty in education. Creativity Research Journal, 2, 139–150.
Lankshear, C. (2000). Literacy. In D. A. Gabbard (Ed.), Knowledge and power in the global economy:
Politics and the rhetoric of school reform (pp. 87–94). Mahwah, NJ: Lawrence Erlbaum.
Larner, W. (2000). Neo-liberalism: Policy, ideology, governmentality. Studies in Political Economy,
63, 5–26.
Larochelle, M., & Désautels, J. (1991). “Of course, it’s just obvious”: Adolescents’ ideas of scientific
knowledge. International Journal of Science Education, 13, 373–389.
Latour, B. (1987). Science in action: How to follow scientists and engineers through society. Cambridge,
MA: Harvard University Press.
Latour, B., & Woolgar, S. (1979). Laboratory life: The social construction of scientific facts. Beverley Hills,
CA: Sage.
Latour, B., & Woolgar, S. (1986). Laboratory life: The construction of scientific facts. Princeton, NJ:
Princeton University Press.
Laudan, L. (1977). Progress and its problems: Toward a theory of scientific growth. Berkeley, CA:
University of California Press.
Laudan, L. (1981). A confutation of convergent realism. Philosophy of Science, 48(1), 19–49.
Laudan, L. (1984a). Science and values. Berkeley, CA: University of California Press.
Laudan, L. (1984b). Explaining the success of science: Beyond epistemic realism and relativism. In J. T.
Cushing, C. F. Delaney, & G. M. Gutting (Eds.), Science and reality: Recent work in the philosophy
of science (pp. 83–105). Notre Dame, IN: Notre Dame University Press.
Laudan, L. (1990). The history of science and the philosophy of science. In R. C. Olby, G. N. Cantor, J. R. R.
Christie, & M. J. S. Hodge (Eds.), Companion to the history of modern science (pp. 47–59). London:
Routledge.
Laudan, L., Donovan, A., Laudan, R., Barker, P., Brown, H., Leplin, J., et al. (1986). Scientific change:
Philosophical models and historical research. Synthese, 69, 141–223.
Laugksch, R. C. (2000). Scientific literacy: A conceptual overview. Science Education, 84(1), 71–94.

224

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
REFERENCES

Lave, J. (1988). Cognition in practice: Mind, mathematics and culture in everyday life. New York:
Cambridge University Press.
Law, N. (2002). Scientific literacy: Charting the terrains of a multifaceted enterprise. Canadian Journal of
Science, Mathematics and Technology Education, 2(2), 151–176.
Lawrenz, F., & Kipnis, N. (1990). Hands-on history of physics. Journal of Science Teacher Education,
1(3), 54–59.
Layton, D. (1988). Revaluing the T in STS. International Journal of Science Education, 10, 367–378.
Layton, D., Davey, A., & Jenkins, E. W. (1986). Science for specific social purposes (SSSP): Perspectives on
adult scientific literacy. Studies in Science Education, 13, 27–52.
Layton, D., Jenkins, E., MacGill, S., & Davey, A. (1993). Inarticulate science? Driffield: Studies in
Education.
Leach, J. (1999). Students’ understanding of the co-ordination of theory and evidence in science.
International Journal of Science Education, 21(8), 789–806.
Leach, J., Driver, R., Millar, R., & Scott, P. (1997). A study of progression in learning about ‘the nature of
science’: Issues of conceptualisation and methodology. International Journal of Science Education, 19,
147–166.
LeCorbeiller, P. (1959). Education in science: Prerequisite for national survival. Daedalus, 88(1), 170–174.
Lecourt, D. (1977). Proletarian science? The case of Lysenko. Manchester: Manchester University Press.
Lederman, M., & Bartsch, I. (Eds.). (2001). The gender and science reader. New York: Routledge.
Lederman, N. G. (1992). Students’ and teachers’ conceptions of the nature of science: A review of the
research. Journal of Research in Science Teaching, 29, 331–359.
Lederman, N. G. (1995). Suchting on the nature of scientific thought: Are we anchoring curricula in
quicksand. Science & Education, 4, 371–377.
Lederman, N. G. (1999). Teachers’ understanding of the nature of science and classroom practice: Factors
that facilitate or impede the relationship. Journal of Research in Science Teaching, 36(8), 916–929.
Lederman, N. G. (2006). Research on nature of science: Reflections on the past, anticipations of the future.
Asia-Pacific Forum on Science Learning and Teaching. Retrieved from http://www.ied.edu.hk/apfslt/
v7_issue1/foreword/index.html
Lederman, N. G. (2007). Nature of science: Past, present, and future. In S. K. Abell & N. G. Lederman
(Eds.), Handbook of research on science education (pp. 831–879). Mahwah, NJ: Lawrence Erlbaum
Associates.
Lederman, N. G., Abd-El-Khalick, F., Bell, R. L., & Schwartz, R. S. (2002). Views of nature of science
questionnaire: Toward valid and meaningful assessment of learners’ conceptions of nature of science.
Journal of Research in Science Teaching, 39(6), 497–521.
Lederman, N. G., & O’Malley, M. (1990). Students’ perceptions of tentativeness in science: Development,
use and sources of change. Science Education, 74(2), 225–239.
Lederman, N. G., Schwartz, R. S., Abd-El-Khalick, F., & Bell, R. L. (2001). Pre-service teachers’
understanding and teaching of nature of science: An intervention study. Canadian Journal of Science,
Mathematics and Technology Education, 25(8), 923–948.
Lederman, N., Wade, P., & Bell, R. L. (1998). Assessing the nature of science: What is the nature of our
assessments? Science & Education, 7(6), 595–615.
Lee, S., & Roth, W.-M. (2002). Learning science in the community. In W.-M. Roth & J. Désautels (Eds.),
Science education as/for sociopolitical action (pp. 37–66). New York: Peter Lang.
Leeper, R. (1935). A study of a neglected portion of the field of learning: The development of sensory
organization. Journal of Genetic Psychology, 46, 41–75.
Leite, L. (2002). History of science in Science education: Development and validation of a checklist for
analysing the historical content of science textbooks. Science & Education, 11, 333–359.
Lemke, J. L. (2001). Articulating communities: Sociocultural perspectives on science education. Journal of
Research in Science Teaching, 38(3), 296–316.
Lewontin, R., & Levins, R. (1976). The problem of Lysenkoism. In H. Rose & S. Rose (Eds.), The
radicalisation of science (pp. 32–64). London: Macmillan.
Lin, H. S., & Chen, C. C. (2002). Promoting preservice chemistry teachers’ understanding about the nature
of science through history. Journal of Research in Science Teaching, 39, 773–792.
Lin, H. S., Hung, J. Y., & Hung, S. C. (2002). Using the history of science to promote students’ problem-
solving ability. International Journal of Science Education, 24(5), 453–464.
Lin, J.-Y. (2007). Responses to anomalous data obtained from repeatable experiments in the laboratory.
Journal of Research in Science Teaching, 44(3), 506–528.

225

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
REFERENCES

Livingstone, R. (1916). A defence of classical education. London: Macmillan.


Lloyd, D., & Wallace, J. (2004). Imaging the future of science education: The case for making futures studies
explicit in student learning. Studies in Science Education, 39, 139–177.
Lochhead, J., & Dufresne, R. (1989). Helping students understand difficult science concepts through the use
of dialogues with history. In D. E. Herget (Ed.), History and philosophy of science in science teaching
(pp. 221–229). Tallahassee, FL: Florida State University.
Longbottom, J. E., & Butler, P. H. (1999). Why teach science? Setting rational goals for science education.
Science Education, 83(4), 473–492.
Longino, H. E. (1990). Science as social knowledge: Values and objectivity in scientific inquiry. Princeton,
NJ: Princeton University Press.
Longino, H. (1994). The fate of knowledge in social theories of science. In F. T. Schmitt (Ed.), Socializing
epistemology: The social dimensions of knowledge (pp. 135–157). Lanham, MD: Rowman & Littlefield.
Longino, H. (2002). The fate of knowledge. Princeton, NJ: Princeton University Press.
Losee, J. (1993). A historical introduction to the philosophy of science (3rd ed.). Oxford: Oxford University
Press.
Loving, C. C. (1997). From the summit of truth to its slippery slopes: Science education’s journey through
positivist-postmodern territory. American Educational Research Journal, 34(3), 421–452.
Lubben, F., & Millar, R. (1996). Children’s ideas about the reliability of experimental data. International
Journal of Science Education, 18, 955–968.
Lucas, A. (1975). Hidden assumptions in measures of ‘knowledge about science and scientists’. Science
Education, 59(4), 481–485.
Luhl, J. (1992). Teaching of social and philosophical background to atomic theory. Science & Education,
1, 193–204.
Mach, E. (1911). History and root of the principle of the conservation of energy (P. E. B. Jourdain, Trans.).
London: Open Court.
Mach, E. (1960). The science of mechanics: A critical and historical account of its development (T. J.
McCormack, Trans.). La Salle, IL: Open Court. (Originally work published 1893)
Machwe, P. (1979). Hinduism: Its contribution to science and civilisation. New Delhi: Vikas Publishing.
Mahoney. M. J. (1979). Psychology of the scientist: An evaluative review. Social Studies of Science, 9(3),
349–375.
Marshall, J. D. (1995). Foucault and neo-liberalism: Biopower and busno-power. In Philosophy of education.
Proceedings of the Philosophy of Education Society. Retrieved from http://www.ed.uiuc.edu/EPS/PES-
Yearbook/95_docs/marshall.html
Marton, F., Fensham, P. J., & Chaiklin, S. (1994). A Nobel’s eye view of scientific intuition: Discussions
with the Nobel prize winners in physics, chemistry and medicine (1970–1986). International Journal of
Science Education, 16(4), 457–473.
Mason, C. L., Kahle, J. B., & Gardner, A. L. (1991). Draw-a-scientist test: Future implications. School
Science & Mathematics, 91(5), 193–198.
Matthews, B. (1996). Drawing scientists. Gender and Education, 8(2), 231–243.
Matthews, M. R. (1992). History, philosophy and science teaching: The present rapprochement. Science &
Education, 1(1), 11–48.
Matthews, M. R. (1993). Constructivism and science education: Some epistemological problems. Journal of
Science Education and Technology, 2, 359–370.
Matthews, M. R. (1994). Science teaching: The role of history and philosophy of science. New York:
Routledge.
Matthews, M. R. (1997). Introductory comments of philosophy and constructivism in science education.
Science & Education, 6, 5–14.
Matthews, M. R. (Ed.). (1998a). Constructivism in science education: A philosophical analysis. Dordrecht:
Kluwer.
Matthews, M. R. (1998b). In defence of modest goals when teaching about the nature of science. Journal of
Research in Science Teaching, 35(2), 161–174.
Maxwell, G. (1962). The ontological status of theoretical entities. In H. Feigl & G. Maxwell (Eds.),
Minnesota studies in philosophy of science (Vol. III, pp. 3–27). Minneapolis, MN: University of
Minnesota Press.
Mayer, V., & Kumano, Y. (1999). The role of system science in future school curricula. Studies in Science
Education, 34, 71–91.

226

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
REFERENCES

Mayer, V. J., & Richmond, J. M. (1982). An overview of assessment instruments in science. Science
Education, 66(1), 49–66.
Mayo, D. (1996). Error and the growth of experimental knowledge. Chicago: University of Chicago Press.
Mayr, E. (1982). The growth of biological though: Diversity, evolution and inheritance. Cambridge, MA:
Harvard University Press.
Mayr, E. (1988). Towards a new philosophy of biology: Observations of an evolutionist. Cambridge, MA:
The Belknap Press of Harvard University Press.
Mayr, E. (1997). This is biology: The science of the living world. Cambridge, MA: Harvard University Press.
McAllister, J. W. (1996). Beauty and revolution in science. Ithaca, NY: Cornell University Press.
McCloskey, M. (1983). Intuitive physics. Scientific American, 248(4), 122–130.
McCloskey, M., & Kargon, R. (1988). The meaning and use of historical models in the study of intuitive
physics. In S. Straudd (Ed.), Ontogeny, phylogeny and historical development (pp. 49–67). Norwood, NJ:
Ablex.
McComas, W. F. (1998). The principal elements of the nature of science: Dispelling the myths. In W. F.
McComas (Ed.), The nature of science in science education: Rationales and strategies (pp. 41–52).
Dordrecht: Kluwer.
McComas, W. F., Clough, M. P., & Almazroa, H. (1998). The role and character of the nature of science in
science education. In W. F. McComas (Ed.), The nature of science in science education: Rationales and
strategies (pp. 3–39). Dordrecht: Kluwer.
McComas, W. F., & Olson, J. K. (1998). The nature of science in international education standards
documents. In W. F. McComas (Ed.), The nature of science in science education: Rationales and
strategies (pp. 53–70). Dordrecht: Kluwer.
McCurdy, R. C. (1958). Towards a population literate in science. The Science Teacher, 25(7), 366–369, 408.
McDermott, L. C. (1984). Research in conceptual understanding of mechanics. Physics Today, 37, 23–32.
McLellan, H. (Ed.). (1996). Situated learning perspectives. Englewood Cliffs, NJ: Educational Technology
Publications.
McNairy, M. R. (1985). Sciencing: Science education for early childhood. School Science & Mathematics,
85, 383–393.
McSharry, G., & Jones, S. (2002). Television programming and advertisements: Help or hindrance to
effective science education? International Journal of Science Education, 24(5), 487–497.
Mead, M., & Metraux, R. (1957). Image of the scientist among high school students. Science, 126, 384–390.
Medawar, P. B. (1963, September 12). Is the scientific paper a fraud? The Listener, 377–378.
Medawar, P. B. (1967). The art of the soluble. London: Methuen.
Medawar, P. B. (1969). Induction and intuition in scientific thought. London: Methuen.
Medawar, P. B. (1982). Plato’s republic. Oxford: Oxford University Press.
Meichtry, Y. J. (1992). Influencing student understanding of the nature of science: Data from a case of
curriculum development. Journal of Research in Science Teaching, 29, 389–407.
Merton, R. K. (1973). The sociology of science: Theoretical and empirical investigations. Chicago:
University of Chicago Press.
Millar, R. (1993). Science education and public understanding of science. In R. Hill (Ed.), ASE secondary
science teachers’ handbook (pp. 357–374). Hemel Hempstead: Simon & Schuster.
Millar, R. (1996). Towards a science curriculum for public understanding. School Science Review, 77(280),
7–18.
Millar, R. (1997). Science education for democracy: What can the school curriculum achieve? In R.
Levinson & J. Thomas (Eds.), Science today: Problem or crisis? (pp. 87–101). London: Routledge.
Millar, R. (2006). Twenty first century science: Insights from the design and implementation of a scientific
literacy approach in school science. International Journal of Science Education, 28(13), 1499–1521.
Millar, R., & Driver, R. (1987). Beyond processes. Studies in Science Education, 14, 33–62.
Millar, R., & Osborne, J. (Eds.). (1998). Beyond 2000: Science education for the future. London: King’s
College London School of Education.
Millar, R., Lubben, F., Gott, R., & Duggan, S. (1994). Investigating in the school science laboratory:
Conceptual and procedural knowledge and their influence on performance. Research Papers in
Education, 9, 207–248.
Miller, A. I. (2006). A thing of beauty. New Scientist, 189(2557), 50–52.
Miller, J. D. (1992). Toward a scientific understanding of the public understanding of science and
technology. Public Understanding of Science, 1, 23–26.

227

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
REFERENCES

Miller, J. D. (1993). Theory and measurement in the public understanding of science: A rejoinder to Bauer
and Schoon. Public Understanding of Science, 2, 235–243.
Miller, J. P. (1993). The holistic teacher. Toronto: OISE Press.
Ministry of Education, Ontario. (1987). Curriculum guidelines, science: Intermediate and senior divions,
part 1, program outline and policy. Toronto: Queen’s Printer for Ontario.
Ministry of Education and Training. (2000a). The Ontario curriculum grades 9 to 12: Program planning and
assessment. Toronto: Queen’s Printer for Ontario.
Ministry of Education and Training. (2000b). The Ontario curriculum grades 11 & 12: Technological
education. Toronto: Queen’s Printer for Ontario.
Ministry of Education, New Zealand. (1993). Science in the New Zealand curriculum. Wellington: Learning
Media.
Mintzes, J. J., Trowbridge, J. E., Arnaudin, M. W., & Wandersee, J. H. (1991). Children’s biology: Studies
on conceptual development in the life sciences. In S. M. Glynn, Yeaney, & Britton (Eds.), The
psychology of learning science (pp. 179–201). Hillsdale, NJ: Lawrence Erlbaum.
Mitman, A. L., Mergendoller, J. R., Marchman, V. A., & Packer, M. J. (1987). Instruction addressing the
components of scientific literacy and its relation to student outcomes. American Educational Research
Journal, 24(4), 611–633.
Mitroff, I. I. (1974). The subjective side of science: A philosophical inquiry into the psychology of the Apollo
moon scientists. Amsterdam: Elsevier.
Mitroff, I. I., & Mason, R. O. (1974). On evaluating the scientific contribution of the Apollo missions via
information theory: A study of the scientist-scientist relationship. Management Science: Applications, 20,
1501–1513.
Monk, M., & Osborne, J. (1997). Placing the history and philosophy of science on the curriculum: A model
for the development of pedagogy. Science Education, 81, 405–424.
Monhardt, R. M., Tillotson, J. W., & Veronesi, P. D. (1999). Same destination, different journeys: A
comparison of male and female views on becoming and being a scientist. International Journal of
Science Education, 21(5), 533–551.
Morgenbesser, S. (1969). The realist-instrumentalist controversy. In S. Morgenbesser, P. Suppes, & M.
White (Eds.), Philosophy, science and method (pp. 200–218). London: St Martin’s Press.
Moss, D. M., Abrams, E. D., & Robb, J. (2001). Examining student conceptions of the nature of
science. International Journal of Science Education, 23(8), 771–790.
Muchie, M., & Xing, L. (Eds.). (2006). Globalisation, inequality and the commodification of life and
well-being. London: Adonis & Abbey.
Mulkay, M. J. (1977). Sociology of the scientific research community. In I. Spiegel-Rosing & D. Price
(Eds.), Science, technology and society: A cross disciplinary perspective (pp. 93–148). London: Sage.
Mulkay, M. (1979). Science and the sociology of knowledge. London: Allen & Unwin.
Mulkay, M., Potter, J., & Yearley, S. (1983). Why an analysis of scientific discourse is needed. In K. D.
Knorr-Cetina & M. Mulkay (Eds.), Science observed: Perspectives on the social study of science
(pp. 171–203). London: Sage.
Munby, H. (1980). Analyzing teaching for intellectual independence. In H. Munby, G. Orpwood, & T.
Russell (Eds.), Seeing curriculum in a new light: Essays from science education (pp. 11–33). Toronto:
OISE Press.
Murcia, K., & Schibeci, R. (1999). Primary student teachers’ conceptions of the nature of science.
International Journal of Science Education, 21(11), 1123–1140.
Nadeau, R., & Déasutels, J. (1984). Epistemology and the teaching of science. Ottawa: Science Council of
Canada.
Nash, L. K. (1963). The nature of the natural sciences. Boston: Little Brown.
Nashon, S. M. (2003). Teaching and learning high school physics in Kenyan classrooms using analogies.
Canadian Journal of Science, Mathematics and Technology Education, 3(3), 333–345.
National Commission on Excellence in Education (NCEE). (1983). A nation at risk. Washington, DC:
Government Printing Office.
National Research Council. (1996). National science education standards. Washington, DC: National
Academy Press.
National Science Board. (1998). Science and engineering indicators – 1998. Arlington, VA: National
Science Foundation.
Needham, J. (1954). Science and civilization in China (7 Vols.). Cambridge: Cambridge University Press.
Needham, J. (1969). The grand titration: Science and society in east and west. London: Allen & Unwin.

228

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
REFERENCES

Needham, J. (1981). Science in traditional China: A comparative perspective. Hong Kong: Chinese
University Press.
Nersessian, N. (1989). Conceptual change in science and science education. Synthese, 80, 163–183.
Nersessian, N. J. (1995). Should physicists preach what they practice? Science & Education, 4, 203–226.
Newton, D. P., & Newton, L. D. (1992). Young children’s perceptions of science and the scientist.
International Journal of Science Education, 14, 331–348.
Newton, L. D., & Newton, D. P. (1998). Primary children’s conceptions of science and the scientist: Is the
impact of a national curriculum breaking down the stereotype? International Journal of Science
Education, 20(9), 1137–1149.
Newton, P., Driver, R., & Osborne, J. (1999). The place of argumentation in the pedagogy of school science.
International Journal of Science Education, 21(5), 553–576.
Newton, R. (1997). The truth of science: Physical theories and reality. Cambridge, MA: Harvard University
Press.
Newton-Smith, W. H. (1981). The rationality of science. London: Routledge & Kegan Paul.
Newton-Smith, W. H. (1990). Realism. In R. C. Olby, G. N. Cantor, J. R. R. Christie, & M. J. S. Hodge
(Eds.), Companion to the history of modern science (pp. 181–195). London: Routledge.
Niaz, M. (1998). From cathode rays to alpha particles to quantum of action: A rational reconstruction of the
atom and its implications for chemistry textbooks. Science Education, 82, 527–552.
Niaz, M. (2000). A rational reconstruction of the kinetic molecular theory of gases based on history and
philosophy of science and its implications for chemistry textbooks. Instructional Science, 28, 23–50.
Nielsen, H. (1993). The endless spiral. Science & Education, 2, 169–181.
Nielsen, H., & Thomsen, P. V. (1990). History and philosophy of science in physics education. International
Journal of Science Education, 12(3), 308–316.
Nietschze, F. (1906/1968). The will to power (revised text of original 1901 publication). Book 3: Principles
of a new evaluation (W. Kaufmann & R. J. Hollingdale, Trans.). New York: Vintage Books.
Noble, D. D. (1998). The regime of technology. In L. E. Beyer & M. W. Apple (Eds.), The curriculum:
Problems, politics and possibilities (pp. 267–283). Albany, NY: State University of New York Press.
Nola, R. (1997). Constructivism in science and science education: A philosophical critique. Science &
Education, 6, 55–83.
Norris, S. P. (1995). Learning to live with scientific expertise: Toward a theory of intellectual communalism
for guiding science teaching. Science Education, 79(2), 201–217.
Norris, S. P. (1997). Intellectual independence for nonscientists and other content-transcendent goals of
science education. Science Education, 81, 239–258.
Norris, S. P., & Phillips, L. M. (1994). Interpreting pragmatic meaning when reading popular reports of
science. Journal of Research in Science Teaching, 31, 947–967.
Norris, S. P., & Phillips, L. M. (2003). How literacy in its fundamental sense is central to scientific literacy.
Science Education, 87(2), 224–240.
Nott, M., & Wellington, J. (1993). Your nature of science profile: An activity for science teachers. School
Science Review, 75(270), 109–112.
Nott, M., & Wellington, J. (1995). Probing teachers’ views of the nature of science: How should we do it and
where should we be looking? In F. Finley, D. Allchin, D. Rhees, & S. Fifield (Eds.), Proceedings of
the third international history, philosophy and science teaching conference (Vol. 2, pp. 864–871).
Minneapolis, MN: University of Minnesota Press.
Nott, M., & Wellington, J. (1996). Probing teachers’ views of the nature of science: How should we do it and
where should we be looking? In G. Welford, J. Osborne, & P. Scott (Eds.), Science education research in
Europe (pp. 283–294). London: Falmer Press.
Nott, M., & Wellington, J. (1998). Eliciting, interpreting and developing teachers’ understandings of the
nature of science. Science & Education, 7, 579–594.
Nott, M., & Wellington, J. (2000). A programme for developing understanding of the nature of science
in teacher education. In W. F. McComas (Ed.), The nature of science in science education: Rationales
and strategies (pp. 293–313). Dordrecht: Kluwer.
Novak, J. D. (1978). An alternative to Piagetian psychology for science and mathematics education. Studies
in Science Education, 5, 1–30.
Nuffield Physics. (1967). Teachers’ guide 1. London: Longman/Penguin.
Nussbaum, J. (1983). Classroom conceptual change: The lesson to be learned from the history of science. In
H. Helm & J. D. Novak (Eds.), Proceedings of the international seminar on misconceptions in science
and mathematics (pp. 272–281). Ithaca, NY: Cornell University.

229

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
REFERENCES

Nussbaum, J. (1989). Classroom conceptual change: Philosophical perspectives. International Journal


of Science Education, 11(5), 530–540.
Nye, M. J. (1980). N-rays: An episode in the history and psychology of science. Historical Studies in the
Physical Sciences, II(1), 125–156.
Oakeshott, M. (1962). Rational conduct. In M. Oakeshott (Ed.), Rationalism in politics and other essays (pp.
80–110). London: Methuen.
Office of Science and Technology/The Wellcome Trust. (2000). Science and the public: A review of science
communication and public attitudes to science. London: The Wellcome Trust.
Ogawa, M. (1989). Beyond the tacit framework of ‘science’ and ‘science education’ among science
educators. In E. Whitelegg, J. Thomas, & S. Tresman (Eds.), Challenges and opportunities for science
education (pp. 17–21). London: Paul Chapman.
Ogborn, J. (1995). Recovering reality. Studies in Science Education, 25, 3–38.
Ogborn, J., & Martins, I. (1996). Metaphorical understandings and scientific ideas. International Journal of
Science Education, 18(6), 631–652.
O’Hear, A. (1989). Introduction to the philosophy of science. Oxford: Clarendon Press.
O’Hearn, G. T. (1976). Science literacy and alternative futures. Science Education, 60(1), 103–114.
Olby, R. C., Cantor, G. N., Christie, J. R. R., & Hodge, M. J. S. (1990). Companion to the history of modern
science. London: Routledge.
Ontario Premier’s Council. (1990). People and skills in the new global economy. Toronto: Queen’s Printer
for Ontario.
O’Rafferty, M. H. (1995). Developing sociological insights on scientific norms using case studies on
misconduct in science. In F. Finley, D. Allchin, D. Rhees, & S. Fifield (Eds.), Proceedings of the third
international history, philosophy and science teaching conference (Vol. 2, pp. 905–912). Minneapolis,
MN: University of Minnesota Press.
Organization for Economic Cooperation and Development (OECD). (1998). Instrument design: A
framework for assessing scientific literacy. Report of Project Managers Meeting. Arnhem: Programme
for International Student Assessment.
Organization for Economic Cooperation and Development (OECD). (2006). Assessing scientific, reading
and mathematical literacy: A framework for PISA 2006. Paris: OECD.
Osborne, J. F. (1996). Beyond constructivism. Science Education, 80(1), 53–82.
Osborne, J. (2001). Promoting argument in the science classroom: A rhetorical perspective. Canadian
Journal of Science, Mathematics and Technology Education, 1(3), 271–290.
Osborne, J., Collins, S., Ratcliffe, M., Millar, R., & Duschl, R. (2003). What “ideas-about-science”
should be taught in school science? A Delphi study of the expert community. Journal of Research in
Science Teaching, 40(7), 692–720.
Osborne, J., & Simon, S. (1996). Primary science: Past and future directions. Studies in Science Education,
26, 99–147.
O’Sullivan, E. (1999). Transformative learning: Educational vision for the 21st century. London: Zed
Books.
Palefau, T. H. (2005). Perspectives on scientific and technological literacy in Tonga: Moving forward in the
21st century. Unpublished PhD thesis, University of Toronto, Toronto.
Parsons, E. C. (1997). Black high school females’ images of the scientist: Expression of culture. Journal of
Research in Science Teaching, 34(7), 745–768.
Pella, M. O. (1976). The place or function of science for a literate citizenry. Science Education,
60(1), 97–101.
Pella, M. O., O’Hearn, G. T., & Gale, C. W. (1966). Referents to scientific literacy. Journal of Research in
Science Teaching, 4, 199–208.
Penner, D. E., Giles, N. D., Lehrer, R., & Schauble, L. (1997). Building functional models: designing an
elbow. Journal of Research in Science Teaching, 34(2), 125–143.
Perkins, D. P. (1981). The mind’s best work. Cambridge, MA: Harvard University Press.
Perry, W. (1970). Forms of intellectual and ethical development in the college years: A scheme. New
York: Holt, Rinehart & Winston.
Peters, D., & Ceci, S. (1982). Peer review practices of psychological journals: The fate of published articles
submitted again. Behavioral and Brain Sciences, 5, 187–195.
Peters, D., & Ceci, S. (1985). Beauty is in the eye of the beholder. Behavioral and Brain Sciences,
8(4), 747–749.

230

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
REFERENCES

Phillips, D. C. (1995). The good, the bad, and the ugly: The many faces of constructivism. Educational
Researcher, 24(7), 5–12.
Phillips, L. M., & Norris, S. P. (1999). Interpreting popular reports of science: What happens when
the reader’s world meets the world on paper? International Journal of Science Education, 21(3),
317–327.
Piaget, J. (1970). Science education and the psychology of the child. New York: Orion Press.
Piaget, J. (1973). To understand is to invent: The future of education. New York: Grossman. (First published
by UNESCO, 1948)
Pickering, A. (1982). Interests and analogies. In B. Barnes & D. Edge (Eds.), Science in context: Readings in
the sociology of science (pp. 125–146). Milton Keynes: Open University Press.
Pilger, J. (2002). The new rulers of the world. London: Verso.
Pimental, G. C. (Ed.). (1960). Chemistry: An experimental approach. San Francisco: Freeman.
Pinch. T. J., & Collins, H. M. (1984). Private science and public knowledge: The committee for the
scientific investigation of the paranormal and its use of the literature. Social Studies of Science, 14,
521–546.
Pintrich, P. R. (2002). The role of metacognitive knowledge in learning, teaching, and assessing. Theory
into Practice, 41(4), 220–227.
Pitt, J. C. (1990). The myth of science education. Studies in Philosophy and Education, 10, 7–17.
Polanyi, M. (1958). Personal knowledge: Towards a post-critical philosophy. London: Routledge &
Kegan Paul.
Pomeroy, D. (1993). Implications of teachers’ beliefs about the nature of science: Comparison of the
beliefs of scientists, secondary science teachers, and elementary teachers. Science Education, 77,
261–278.
Popper, K. R. (1959). The logic of scientific discovery. London: Hutchinson.
Popper, K. R. (1963). Conjecture and refutations. London: Routledge & Kegan Paul.
Popper, K. R. (1972). Objective knowledge. Oxford: Oxford University Press.
Popper, K. R. (1974). Replies to my critics. In P. A. Schilpp (Ed.), The philosophy of Karl Popper (Vol.
2, pp. 961–1197). La Salle, IL: Open Court.
Posner, G. J., Strike, K. A., Hewson, P. J., & Gertzog, W. A. (1982). Accommodation of a scientific
conception: Toward a theory of conceptual change. Science Education, 66, 211–227.
Postman, N. (1992). Technopoly: The surrender of culture to technology. New York: Alfred A. Knopf.
Putnam, H. (1975). Mathematics, matter and method: Philosophical papers (Vol. 1). Cambridge:
Cambridge University Press.
Putnam, H. (1978). Meaning and the moral sciences. London: Routledge & Kegan Paul.
Qualifications and Curriculum Authority (QCA). (1998). Education for citizenship and the teaching of
democracy in schools. London: QCA.
Quine, W. V. O. (1970). On the reasons for the indeterminacy of translation. Journal of Philosophy,
67, 178–183.
Ratcliffe, M., & Grace, M. (2003). Science education for citizenship: Teaching socio-scientific issues.
Maidenhead: Open University Press.
Ravetz, J. R. (1971). Scientific knowledge and its social problems. Oxford: Clarendon Press.
Ravetz, J. R. (1997). Simple scientific truths and uncertain policy realities: Implications for science
education. Studies in Science Education, 30, 5–17.
Riggs, P. J. (1992). Whys and ways of science: Introducing philosophical and sociological theories of
science. Melbourne: Melbourne University Press.
Roberts, D. A. (1983). Scientific literacy: Towards balance in setting goals for school science
programs. Ottawa: Science Council of Canada.
Roberts, D. A. (2007). Scientific literacy/science literacy. In S. K. Abell & N. G. Lederman (Eds.),
Handbook of research on science education (pp. 729–780). Mahwah, NJ: Lawrence Erlbaum
Associates.
Robertson, I. (1999). Key evidence in testing hypotheses. In M. Bandiera, S. Caravita, E. Torracca, &
M. Vicenti (Eds.), Research in science education in Europe (pp. 193–200). Dordrecht: Kluwer.
Rockefeller Brothers Fund. (1958). The pursuit of excellence: Education and the future of America.
Garden City, NY: Doubleday.
Rodriguez, M. A., & Niaz, M. (2002). How in spite of the rhetoric, history of chemistry has been
ignored in presenting atomic structure in textbooks. Science & Education, 11, 423–441.
Roe, A. (1961). The psychology of the scientist. Science, 134(3477), 456–459.

231

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
REFERENCES

Rorty, R. (1991). Objectivity, relativism and truth: Philosophical papers (Vol. 1). Cambridge: Cambridge
University Press.
Rose, H. (1997). Science wars: My enemy’s enemy is – only perhaps – my friend. In R. Levinson &
J. Thomas (Eds.), Science today: Problem or crisis? (pp. 51–64). London: Routledge.
Rosenau, J. (Ed.). (1992). Governance without government: Order and change in world politics.
Cambridge: Cambridge University Press.
Roth, W.-M., & Barton, A. C. (2004). Rethinking scienctific literacy. New York: RoutledgeFalmer.
Roth, W.-M., & Désautels, J. (2002). Science education as/for sociopolitical action. New York:
Peter Lang.
Roth, W.-M., & Désautels, J. (2004). Educating for citizenship: Reappraising the role of science
education. Canadian Journal of Science, Mathematics and Technology Education, 4(2), 149–168.
Roth, W.-M. & Roychoudhury, A. (1994). Physics students’ epistemologies and views about knowing
and learning. Journal of Research in Science Teaching, 31, 5–30.
The Royal Society. (1985). The public understanding of science. London: Royal Society.
Roychoudhury, A., Tippins, D. J., & Nichols, S. E. (1995). Gender-inclusive science teaching: A
feminist-constructivist approach. Journal of Research in Science Teaching, 32(9), 897–924.
Rubba, P. (1976). Nature of scientific knowledge scale. Bloomington, IN: Indiana University School of
Education.
Rubba, P. A., & Anderson, H. O. (1978). Development of an instrument to assess secondary school
students’ understanding of the nature of scientific knowledge. Science Education, 62(4), 449–458.
Rudolph, J. L. (2000). Reconsidering the ‘nature of science’ as a curriculum component. Journal of
Curriculum Studies, 32(3), 403–419.
Rudolph, J. L. (2002). Portraying epistemology: School science in historical context. Science Education,
87, 64–79.
Rudolph, J. L. (2005). Inquiry, instrumentalism, and the public understanding of science. Science
Education, 89(5), 803–821.
Rudolph, J. L., & Stewart, J. (1998). Evolution and the nature of science: On the historical discord and
its implications for education. Journal of Research in Science Teaching, 35(10), 1069–1089.
Ruse, M. (1989). Making use of creationism: A case study for the philosophy of science classroom.
Studies in Philosophy and Education, 10(1), 81–92.
Russell, B. (1912). The problems of philosophy. Oxford: Oxford University Press.
Russell, B. (1961). History of western philosophy. London: George Allen & Unwin.
Russell, T. L. (1981). What history of science, how much and why? Science Education, 65(1), 51–64.
Rutherford, F. J. (2001). Fostering the history of science in American science education. Science &
Education, 10, 569–580.
Rutherford, F. J., Holton, G., & Watson, F. G. (1970). Harvard project physics: Text. New York: Holt,
Rinehart & Winston.
Ryan, A. G. (1987). High school graduates’ beliefs about science-technology-society. IV: The
characteristics of scientists. Science Education, 71, 489–510.
Ryder, J. (2001). Identifying science understanding for functional scientific literacy. Studies in Science
Education, 36, 1–44.
Ryder, J., & Leach, J. (2000). Interpreting experimental data: The views of upper secondary school and
university science students. International Journal of Science Education, 22(10), 1069–1084.
Ryder, J., Leach, J., & Driver, R. (1999). Undergraduate science students’ images of science. Journal of
Research in Science Teaching, 36(2), 201–219.
Ryle, G. (1956). Dilemmas. Cambridge: Cambridge University Press.
Sagan, C. (1995). The demon-haunted world: Science as a candle in the dark. New York: Random
House.
Saliba, G. (2007). Islamic science and the making of the European renaissance. Cambridge, MA: MIT
Press.
Samson, M. D., & Weininger, S. J. (1995). Light, vision and understanding: Using the history of science
and art history in an engineering college curriculum. In F. Finley, D. Allchin, D. Rhees, & S. Fifield
(Eds.), Proceedings of the third international history, philosophy and science teaching conference (Vol.
2, pp. 988–996). Minneapolis, MN: University of Minnesota Press.
Sardar, Z. (1989). Explorations in Islamic science. London: Mansell.
Saunders, W. L. (1992). The constructivist perspective: Implications and teaching strategies for science.
School Science & Mathematics, 92, 136–141.

232

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
REFERENCES

Schaefer, H. F. (1986). Methylene: A paradigm for computational quantum chemistry. Science, 231,
1100–1107.
Scheffler, I. (1967). Science and subjectivity. Indianapolis, IN: Bobbs-Merrill.
Schilpp, P. A. (1951). Albert Einstein: Philosopher-scientist. New York: Tudor.
Schwab, J. J. (1962). The teaching of science as enquiry. In J. J. Schwab & P. F. Brandwein (Eds.), The
teaching of science (pp. 3–103). Cambridge, MA: Harvard University Press.
Schwartz, A. T. (1995). Three decades of humanizing the scientists and simonizing the humanist: Some
retrospective reflections. In F. Finley, D. Allchin, D. Rhees, & S.Fifield (Eds.), Proceedings of the third
international history, philosophy and science teaching conference (Vol. 2, pp. 1031–1041). Minneapolis,
MN: University of Minnesota Press.
Schwartz, R. S., & Lederman, N. G. (2002). ‘It’s the nature of the beast’: The influence of knowledge
and intentions on learning and teaching the nature of science. Journal of Research in Science
Teaching, 39(3), 205–236.
Schwartz, R. S., & Lederman, N. G. (2006). Exploring contextually-based views of NOS and scientific
inquiry: What scientists say [Tentativeness, creativity, scientific method, and justification]. Paper
presented at annual meeting of the National Association for Research in Science Teaching, San
Francisco, CA, April.
Screen, P. (1986). Warwick process science. Southampton: Ashford.
Screen, P. (1988). A case for a process approach: The Warwick experience. Physics Education, 23(3),
146–149.
Select Committee on Science and Technology, House of Lords. (2000). Science and Society. 3rd Report,
Session 1999–2000. London: HMSO.
Sequeira, M., & Leite, L. (1991). Alternative conceptions and history of science in physics teacher
education. Science Education, 75(1), 45–56.
Seroglou, F., Koumaras, P., & Tselfes, V. (1998). History of science and instructional design: The case
of electromagnetism. Science & Education, 7, 261–280.
Seroglou, F., & Koumaras, P. (2001). The contribution of the history of physics in physics education.
Science & Education, 10, 153–172.
Shamos, M. H. (1993). STS: A time for caution. In R. E. Yager (Ed.), The science, technology, society
movement (pp. 65–72). Washington, DC: National Science Teachers Association.
Shamos, M. (1995). The myth of scientific literacy. New Brunswick, NJ: Rutgers University Press.
Shapere, D. (1982). The concept of observation in science and philosophy. Philosophy of Science, 49,
485–525.
Shapere, D. (1984). Reason and the search for knowledge: Investigations in the philosophy of science.
Dordrecht: Reidel.
Shapin, S. (1979). Homo phrenologicus: Anthropological perspectives on an historical problem. In B.
Barnes & S. Shapin (Eds.), Natural order: Historical studies of scientific culture (pp. 41–71).
London: Sage.
Sharkawy, A. (2006). An inquiry into the use of stories about scientists from diverse sociocultural back-
grounds in broadening grade one students’ images of science and scientists. Toronto: Unpublished
PhD thesis, University of Toronto.
She, H.-C. (1998). Gender and grade level differences in Taiwan students’ stereotypes of science and
scientists. Reearch in Science & Technological Education, 16, 125–135.
Shen, B. S. P. (1975). Scientific literacy and the public understanding of science. In S. B. Day (Ed.), The
communication of scientific information (pp. 44–52). Basel: Karger.
Shepherd, L. (1993). Lifting the veil: The feminine face of science. Boston: Shambala Publications.
Sherman, S. J. (2000). Science and science teaching: Science is something you can do. Boston: Houghton
Mifflin.
Sherratt, W. J. (1982). History of science in the science curriculum: An historical perspective. Part I: Early
interest and roles advocated. School Science Review, 64(227), 225–236.
Sherratt, W. J. (1983). History of science in the science curriculum: An historical perspective. Part II: Interest
shown by teachers. School Science Review, 64(228), 418–424.
Shibley, I. A. (2003). Using newpapers to examine nature of science. Science & Education, 12(7), 691–702.
Shortland, M. (1988). Advocating science: Literacy and public understanding. Impact of Science on Society,
38(4), 305–316.
Siegel, H. (1991). The rationality of science, critical thinking, and science education. In M. R. Matthews
(Ed.), History, philosophy and science teaching: Selected readings (pp. 45–62). Toronto: OISE Press.

233

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
REFERENCES

Slezak, P. (1989). Scientific discovery by computer as emopirical refutation of the strong programme. Social
Studies of Science, 19, 563–600.
Slezak, P. (1994). Sociology of scientific knowledge and scientific education: Part 1. Science & Education,
3, 265–294.
Smart, J. J. C. (1968). Between science and philosophy: An introduction to philosophy of science. New York:
Random House.
Smith, C. L., Maclin, D., Houghton, C., & Hennessey, M. G. (2000). Sixth grade students’ epistemologies
of science: The impact of school science experiences on epistemological development. Cognition &
Instruction, 18(3), 349–422.
Smith, M. U., Lederman, N. G., Bell, R. L., McComas, W. F., & Clough, M. P. (1997). How great is the
disagreement about the nature of science: A response to Alters. Journal of Research in Science Teaching,
34(10), 1101–1103.
Smolicz, J. J., & Nunan, E. E. (1975). The philosophical and sociological foundations of science education:
The demythologizing of school science. Studies in Science Education, 2, 101–143.
Snow, C. P. (1962). The two cultures and the scientific revolution. Cambridge: Cambridge University Press.
Solomon, J. (1994). The rise and fall of constructivism. Studies in Science Education, 23, 1–19.
Solomon, J. (1999). Meta-scientific criticisms, curriculum innovation and the propagation of scientific
culture. Journal of Curriculum Studies, 31(1), 1–13.
Solomon, J. (2001). Teaching for scientific literacy: What could it mean? School Science Review, 82(300),
93–96.
Solomon, J., & Aikenhead, G. (Eds.). (1994). STS education: International perspective on reform. New
York: Teachers College Press.
Solomon, J., & Thomas, J. (1999). Science education for public understanding of science. Studies in Science
Education, 33, 61–90.
Solomon, J., Duveen, J., & Scott, L. (1994). Pupils’ images of scientific epistemology. International Journal
of Science Education, 16, 361–373.
Solomon, J., Scott, L., & Duveen, J. (1996). Large-scale exploration of pupils’ understanding of the
nature of science. Science Education, 80(5), 493–508.
Song, J., & Kim, K.-S. (1999). How Korean students see scientists: The images of the scientist.
International Journal of Science Education, 21(9), 957–977.
Stanley, W. B., & Brickhouse, N. W. (1995). Science education without foundations: A response to
loving. Science Education, 79(3), 349–354.
Stavy, R. (1991). Using analogy to overcome misconceptions about conservation of matter. Journal of
Research in Science Teaching, 28(4), 305–313.
Steinberg, M. S., Brown, D. E., & Clement, J. (1990). Genius is not immune to persistent mis-
conceptions: Conceptual difficulties impeding Isaac Newton and contemporary physics students.
International Journal of Science Education, 12(3), 265–273.
Stent, G. S. (1969). The coming of the golden age. New York: Natural History Press.
Stevens, P. (1978). On the Nuffield philosophy of science. Journal of Philosophy of Education, 12,
99–111.
Stewart, J., & Hafner, R. (1991). Extending the conception of ‘problem’ in problem-solving research.
Science Education, 75(1), 105–120.
Stinner, A., & Williams, H. (1993). Conceptual change, history and science stories. Interchange,
24(1&2), 87–103.
Stinner, A., & Williams, H. (1998). History and philosophy of science in the science curriculum. In B. J.
Fraser & K. G. Tobin (Eds.), The international handbook of science education (pp. 1027–1045).
Dordrecht: Kluwer.
Storer, N. (1966). The social system of science. New York: Holt Rinehart & Winston.
Strawson, P. F. (1971). Introduction to logical theory. London: Methuen.
Stromquist, N. P., & Monkman, K. (Eds.). (2000). Globalization and education: Integration and
contestation across cultures. London: Rowman & Littlefield.
Stuewer, R. E. (1998). History and physics. Science & Education, 7, 13–30.
Suchting, W. A. (1992). Constructivism deconstructed. Science & Education, 1, 223–254.
Suchting, W. A. (1995). The nature of scientific thought. Science & Education, 4, 1–22.
Sutman, F. X. (1996). Science literacy: A functional definition. Journal of Research in Science
Teaching, 33(5), 459–460.

234

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
REFERENCES

Symington, D., & Spurling, H. (1990). The ‘draw a scientist test’: Interpreting the data. Research in
Science & Technological Education, 8(1), 75–77.
Taber, K. S. (2003). Mediating mental models of metals: Acknowledging the priority of the learner’s
prior learning. Science Education, 87(5), 732–756.
Theobald, D. W. (1968). An introduction to the philosophy of science. London: Methuen.
Thomas, G., & Durant, J. (1987). Why should we promote the public understanding of science? In M.
Shortland (Ed.), Scientific literacy papers (pp. 1–14). Oxford: Oxford University Department for
External Studies.
Thomsen, P. V. (1998). The historical-philosophical dimension in physics teaching: The Danish
experiences. Science & Education, 7, 493–503.
Thomson, G. P. (1965). Some thoughts on the scientific method. In R. S. Cohen & M. W. Wartofsky
(Eds.), Boston studies in the philosophy of science (Vol. II, pp. 81–92).
Tiles, M. (1984). Bachelard: Science and objectivity. Cambridge: Cambridge University Press.
Tippens, D. J., Nichols, S. E., & Bryan, L. A. (2000). International science educators’ perceptions of
scientific literacy. In S. K. Abell (Ed.), Science teacher education: An international perspective.
Dordrecht: Kluwer.
Tobin, K. (2000). Constructivism in science education: Moving on. In D. C. Phillips (Ed.), Construc-
tivism in education (pp. 227–253). Chicago: National Society for the Study of Education.
Tobin, K. G., & Capie, W. (1980). Teaching process skills in the middle school. School Science &
Mathematics, 80, 590–597.
Toulmin, S. E. (1963). The philosophy of science. London: Hutchinson.
Treagust, D. F., Chittleborough, G., & Mamiala, T. L. (2002). Students’ understanding of the role of
scientific models in learning science. International Journal of Science Education, 24(4), 357–368.
Tsai, C.-C., & Liu, S.-Y. (2005). Developing a multi-dimensional instrument for assessing students’
epistemological views toward science. International Journal of Science Education, 27(13), 1621–1638.
Tsaparlis, G. (1997). Atomic and molecular structure in chemical education. Journal of Chemical
Education, 74(8), 922–925.
Turnbull, C. M. (1961). The forest people. London: Jonathan Cape.
Turner, S., & Sullenger, K. (1999). Kuhn in the classroom, Lakatos in the lab: Science educators confront the
nature-of-science debate. Science, Technology & Human Values, 24(1), 5–30.
Tyson, L. M., Venville, G. J., Harrision, A. G., & Treagust, D. F. (1997). A multidimensional framework for
interpreting conceptual change events in the classroom. Science Education, 80, 121–140.
UNESCO. (1993). International forum on scientific and technological literacy for all. Final Report. Paris:
UNESCO.
Ungar, S. (2000). Knowledge, ignorance and the popular culture: Climate change versus the ozone hole.
Public Understanding of Science, 9, 297–312.
van Driel, J., de Vos, W., & Verloop, N. (1998). Relating students’ reasoning to the history of science:
The case of chemical equilibrium. Research in Science Education, 28(2), 187–198.
van Driel, J. H., & Verloop, N. (1999). Teachers’ knowledge of models and modelling in science.
International Journal of Science Education, 21(11), 1141–1153.
van Fraassen, B. C. (1980). The scientific image. Oxford: Oxford University Press.
von Glasersfeld, E. (1987). Construction of knowledge. Salinas, CA: Intersystems Publications.
von Glasersfeld, E. (1989). Cognition, construction of knowledge, and teaching. Synthese, 80, 121–140.
von Glasersfeld, E. (1992). Constructivism reconstructed: A reply to Suchting. Science & Education, 1,
379–384.
Von Glasersfeld, E. (2007). Key works in radical constructivism (M. Larochelle, Ed.). Rotterdam: Sense
Publishers.
Vosniadou, S., & Brewer, W. F. (1987). Theories of knowledge restructuring in development. Review of
Educational Research, 51(1), 51–67.
Wallace, C. S. (2004). Framing new research in science literacy and language use: Authenticity, multiple
discourses, and the ‘third space’. Science Education, 88(6), 901–914.
Wallace, C. S., Tsoi, M. Y., Calkin, J., & Darley, W. M. (2003). Learning from inquiry-based laboratories in
nonmajor biology: An interpretive study of the relationships among inquiry experience, epistemologies,
and conceptual growth. Journal of Research in Science Teaching, 40, 986–1024.
Wandersee, J. H. (1985). Can the history of science help science educators anticipate students’
misconceptions? Journal of Research in Science Teaching, 23(7), 581–597.

235

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
REFERENCES

Wandersee, J. H. (1990). On the value and use of the history of science in teaching today’s science:
Constructingg historical vignettes. In D. E. Herget (Ed.), More history and philosophy of science in
science teaching (pp. 278–283). Tallahassee, FA: Florida State University.
Wang, M. C. (1995). The effect of a role model project upon the attitudes on ninth-grade science students.
Journal of Research in Science Teaching, 32(2), 195–204.
Wang, H. A., & Marsh, D. D. (2002). Science instruction with a humanistic twist: Teachers’ perception and
practice in using the history of science in their classrooms. Science & Education, 11, 169–189.
Ward, A. (1986). Magician in a white coat. School Science Review, 68(243), 348–350.
Waterman, A. T. (1960). National science foundation: A ten year resume. Science, 131(3410), 1341–1354.
Watson, J. D. (1980). The double helix: A personal account of the discovery of the structure of DNA (G. S.
Stent, Ed.). New York: Norton.
Watson, R. (2000). The role of practical work. In M. Monk & J. Osborne (Eds.), Good practice in science
teaching: What research has to say (pp. 57–71). Buckingham: Open University Press.
Watson, R., Goldsworthy, A., & Wood-Robinson, V. (1999). What is not fair with investigations? School
Science Review, 80(292), 101–106.
Weatherford, J. (1988). Indian givers: What the native Americans gave to the world. New York: Crown.
Weaver, W. (1966). Why is it so important that science be understood? Impact of Science on Society,
XVI(1), 41–50.
Weck, M. A. (1995). Are today’s models tomorrow’s misconceptions? In F. Finley, D. Allchin, D. Rhees, &
S.Fifield (Eds.), Proceedings of the third international history, philosophy and science teaching
conference (Vol. 2, pp. 1286–1294). Minneapolis, MN: University of Minnesota Press.
Wegener, A. (1915). The origin of continents and oceans (J. Biram, Trans.). New York: Dover. (Republished
in 1966 by Dover)
Weinberg, S. (1987). Newtonianism, reductionism and the art of congressional testimony. Nature, 330,
433–437.
Welch, W. W. (1969a). Science process inventory. Minneapolis, MN: University of Minnesota.
Welch, W. W. (1969b). Wisconsin inventory of science processes. Madison, WI: University of Wisconsin
Scientific Literacy Research Center.
Welch, W., & Walberg, H. (1972). A national experiment in curriculum evaluation. American
Educational Research Journal, 9, 373–383.
Wellington, J. J. (Ed.). (1989). Skills and processes in science education. London: Routledge.
Wellington, J. (1993). Using newspapers in science education. School Science Review, 74(268), 47–52.
Wellington, J. (2001). What is science education for? Canadian Journal of Science, Mathematics and
Technology Education, 1(1), 23–38.
Wellington, J., & Osborne, J. (2001). Language and literacy in science education. Buckingham: Open
University Press.
White, F. C. (1983). Knowledge and relativism III: The sciences. Educational Philosophy & Theory, 15,
1–29.
Wible, J. R. (1992). Fraud in science: An economic approach. Philosophy of the Social Sciences, 22(1),
5–27.
Wichman, E. H. (1971). Berkeley physics course. Vol. 4: Quantum physics. New York: McGraw Hill.
Wiener, N. (1950). The human use of human beings: Cybernetics and society. Boston: Houghton
Mifflin.
Williams, J. D. (2002). Ideas and evidence in science: The portrayal of scientists in GCSE textbooks.
School Science Review, 84(307), 89–101.
Wilson, J., & Cowell, B. (1982). Methods of subject-teaching: A challenge to current thinking.
Westminster Studies in Education, 5, 37–41.
Windschitl, M. (2004). Caught in the cycle of reproducing folk theories of ‘inquiry’: How pre-service
teachers reproduce the discourse and practices of an atheoretical scientific method. Journal of
Research in Science Teaching, 41(5), 481–512.
Winter, H. J. J. (1952). Eastern science: An outline of its scope and contribution. London: John Murray.
Wittgenstein, L. (1969). On certainty (D. Paul & G. E. M. Anscombe, Trans.). Oxford: Blackwell.
Wong, S. L. & Hodson, D. (2008a). From the horse’s mouth: What scientists say about scientific investi-
gation and scientific knowledge. Science Education (in press).
Wong, S. L., & Hodson, D. (2008b). More from the horse’s mouth: What scientists say about science as
a social practice. Science Education (under review).

236

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
REFERENCES

Wong, S. L., Kwan, J., Yung, B. H. W., & Hodson, D. (2008a). Turning crisis into opportunity: Nature of
science and scientific inquiry as illustrated in the scientific research on severe acute respiratory syndrome
(SARS). Science & Education (in press).
Wong, S. L., Kwan, J., Yung, B. H. W., & Hodson, D. (2008b). Turning crisis into opportunity: Enhancing
student teachers’ understanding of the nature of science through a case study of scientific research in
severe acute respiratory syndrome (SARS). International Journal of Science and Education (in press).
Woolf, V. (1925). Modern fiction. In V. Woolf (Ed.), The common reader (pp. 146–154). London:
Hogarth Press.
Yager, R. E. (Ed.). (1996). Science/technology?Society as reform in science education. Albany, NY:
State University of New York Press.
Yore, L. D., & Treagust, D. F. (2006). Current realities and future possibilities: Language and science
literacy – Empowering research and informing instruction. International Journal of Science
Education, 28(2–3), 291–314.
Young, M. F. D. (1976). The schooling of science. In G. Whitty & M. F. D. Young (Eds.), Explorations
in the politics of school knowledge (pp. 47–61). Driffield: Nafferton Books.
Young, R. M. (1987). Racist society, racist science. In D. Gill & L. Levidow (Eds.), Anti-racist science
teaching (pp. 16–42). London: Free Association Books.
Zeitler, W. R., & Barufaldi, J. P. (1988). Elementary school science: A perspective for teachers. New
York: Longman.
Ziman, J. M. (1968). Public knowledge. Cambridge: Cambridge University Press.
Ziman, J. M. (1978). Reliable knowledge. Cambridge: Cambridge University Press.
Ziman, J. M. (1980). Teaching and learning about science and society. Cambridge: Cambridge
University Press.

237

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
Derek Hodson - 978-90-8790-507-1
Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
INDEX

Abd-El-Khalick, F. 25, 25, 26, 176, Doing science 2, 29, 32, 88, 91, 93,
177, 208 95, 136, 151, 154, 156, 157, 182
Aesthetics 11, 27, 123, 159 Economy, neoliberalism &
Agreement with facts 74, 76–79, 81, globalization 12–14, 16, 27, 34,
85, 87, 96, 97, 104, 110, 113, 36–38
119, 123, 124, 132, 168, 184 Elegance & simplicity 105, 124,
Aikenhead, G. 8, 16, 17, 25, 26, 160 157, 189, 190
Allchin, D. 83, 126, 157, 160, 166, Ethics 11, 12, 16, 20, 39, 85, 129,
199 130, 142, 154, 159, 160, 173,
American Association for the 183, 196
Advancement of Science (AAAS) Experiment 29, 30, 59, 67–69, 73,
1, 4, 11, 14, 16, 17, 19, 34, 92, 79, 80, 81, 85–91, 94, 100, 101,
196, 197n 119, 127, 134–136, 155, 164,
Anything goes 82, 95, 98, 133, 183 183, 201, 203
Argumentation 3, 9, 33, 35, 38, 99, Falsification & refutation 67–71,
119, 130, 133–135, 137, 138, 73–75, 77, 78, 86, 96
141, 145, 153, 157, 173, 179, Feyerabend, P.K. 57, 58, 75, 82,
183–185, 189, 200, 201, 203, 209 95–99, 102n, 113, 155, 183
Bencze, J.L., 14, 36, 37, 102n Fluidity, holism & idiosyncrasy
Bias 19, 145, 175, 182, 183, 202, 97–101, 131, 133, 135–137, 156,
204, 208 164, 179–181, 184
Brush, S. 75, 163, 167, 170, 172n, Giere, R.N. 111, 116–118, 190
175, 181, 199 Hacking, I. 86, 110, 118, 141
Competition & collaboration 139, Hodson, D. 2, 10, 14, 20–22n, 25,
140, 141, 202, 203 28, 32, 39, 40n, 56, 62, 63, 86,
Consensus in science 59, 81, 90, 92, 94, 98, 102n, 121n, 139,
99–101, 127–130, 135, 200, 202 142, 146, 160, 171n–174, 178,
Consensus in science education 33, 180, 182, 184, 193, 196, 202,
173–177, 186 203, 208
Constructivism 26, 56, 97, 117, 150, Holton, G. 33, 74, 124, 125, 144,
152, 158 167, 169
Corroboration 67–71, 105 Humanizing science 153–155, 168
Creativity 67, 68, 75, 81, 97, 127, Hypotheses & conjectures 67–69,
139, 142, 156, 183, 185 71, 81, 86, 93, 94, 97, 110
Critical realism 114–118, 196 Implicit & explicit 3, 13, 23, 24, 35,
Critical scientific literacy 2, 10, 15, 95, 138, 157, 173, 182, 183, 208
17, 18, 20, 35, 38, 39, 160, 168, Incommensurability 79–81, 85, 86,
178, 208 141, 180, 186–188, 203
Désautels, J. 19, 31–34 Induction 53, 60–64, 67, 69, 71, 75,
Direct & indirect observation 107, 89, 91, 92, 96, 184
108 Inference 56–60, 93, 107, 108, 176
Discovery learning 88–91 Intellectual independence 16, 17,
Diversity in science 33, 204 158

239

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
INDEX

Intuition 123, 124, 133, 134, 136, Observation 41–60, 67–71, 73, 74,
141, 155, 157, 201, 202 76, 78–81, 85, 86, 88, 89, 91,
Jenkins, E.W. 1, 5, 7–9, 16, 88, 92, 93–95, 97, 104, 106–108, 113,
150, 153, 162, 169, 171n, 183 127, 155, 156, 176, 183, 187,
Knorr-Cetina, K.D. 126, 134, 135, 189, 194, 201, 203
136, 200, 202 Osborne, J. 2, 6, 40n, 135, 157, 176,
Kuhn, T.S. 20, 76, 77, 79–81, 83n, 177, 197n
86, 87, 102n, 105, 116, 118, 129, Paradigm 76–81, 85–87, 116, 118,
163, 169, 184, 186–190, 197n, 123, 129, 180, 187, 188, 190
198n Political issues 2, 10, 12, 14, 15,
Lakatos, I. 72, 73, 83n, 102n, 155, 18–20, 34, 35, 37, 38, 154, 159,
169, 198n, 205, 207 160, 173, 183, 193, 208
Language of science 2–4, 20, 35, 36, Popper, K.R. 20, 58, 67–71, 74,
79, 105, 106, 129, 137, 138, 144, 81–83n, 110, 111, 114, 119, 120,
145, 156, 173, 183, 193, 207, 208 169, 205, 206
Latour, B. 88, 126, 133, 134, 137, Prediction 68–71, 73, 75, 78, 80, 94,
138–142, 169, 183, 200 104, 105, 108–111, 113, 189,
Laudan, L. 105, 111, 164, 175, 188, 200, 201, 207
189, 195, 201, 207, 208 Proof & certainty 29, 30, 101, 103
Lederman, N.G. 24–27, 33, 40n, Quasi-induction 69, 70
176, 178, 180, 197n, 208 Recurrent history 165, 166
Literacy 2–4, 7, 20, 137, 138 Rudolph, J.L. 1, 150, 179, 181,
Longino, H. 101, 146n, 190, 194, 182
197n, 203, 204 Scientific method (algorithmic)
Matthews, M.R. 121n, 149, 162, 92–95, 98, 99, 101, 134, 136,
183, 196 155, 179
Medawar, P. 53, 137, 177, 178 Scientific method/investigation 30,
Merton, R.K. 129, 130, 132, 139, 140 31, 32, 34, 60, 82, 92, 95, 97–
Metacognition 8, 150, 151, 157, 158 102, 105, 106, 112, 127, 133,
Millar, R. 1, 6, 21n, 33, 93, 102, 162, 134, 136, 137, 140, 154–156,
195 164, 180, 183
Miracles 109, 110, 115, 195 Scientific revolutions 76–78, 80, 81,
Models 104, 114–118, 155, 157, 97, 99, 116, 151, 163, 183, 184,
169, 181, 183, 196 187, 205
Non-Western science 16, 100, 128, Shapere, D. 58, 59, 187, 205
144, 160, 161, 183 Social construction 81, 105, 112,
Normal science 76, 77, 80, 81, 116, 118, 119, 125, 127, 128,
129, 151, 163, 183, 188 134–136, 142–146, 153, 164,
Norms, interests & values 139–145, 188, 190–194, 199–209
159, 182, 183, 193, 204, 208 Sociocultural & economic influences
Norms of scientific practice 81, 102, 113, 120, 123, 125–128,
129–135, 139 135, 139–141, 144, 151, 153,
Nuffield 1, 88, 89, 103 155, 159, 160, 162, 163, 173,
Objectivity 9, 31, 62, 63, 81, 85, 183, 185, 191, 202
119, 127, 129, 130, 175, 190, Solomon, J. 1, 22n, 30, 33, 40n, 160,
194, 200, 203, 204, 206 195

240

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access
INDEX

Technological success 110, 111, Vested interest 125, 126, 130, 140,
118, 195 183
Tentative nature of science 17, 27, Wellington, J. 2, 6, 25, 31, 92, 162
35, 71, 78, 155, 184 Whiggish history 163, 164, 168, 169
Theories as complex structures Wong, S.L. 25, 40n, 63, 121n, 139,
71–75, 79, 118, 119, 151, 156, 142, 146n, 178, 184, 196, 203,
157, 187, 206 208
Values 10, 11, 14, 17, 31, 35, 81,
127, 128, 141–145, 185, 189,
204, 208

241

Derek Hodson - 978-90-8790-507-1


Downloaded from Brill.com11/09/2020 11:12:23AM
via free access

You might also like